Next Article in Journal
Orientation of Conjugated Polymers in Single Crystals: Is It Really Unusual for the Polydiacetylene Backbone to Be Aligned Almost Perpendicular to the Hydrogen Bond Network?
Previous Article in Journal
Taguchi Robust Design of Phase Transfer Catalytic Hydrolysis of Polyethylene Terephthalate (PET) Waste in Mild Conditions: Application for the Preparation of Metal–Organic Frameworks
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Research Progress on Texture Regulation of Rare-Earth Magnesium Alloys

1
School of Materials Science and Chemical Engineering, Harbin University of Science and Technology, Harbin 150040, China
2
Key Laboratory for Light-Weight Materials, Nanjing Tech University, Nanjing 211816, China
*
Author to whom correspondence should be addressed.
Solids 2025, 6(1), 11; https://doi.org/10.3390/solids6010011
Submission received: 28 December 2024 / Revised: 27 February 2025 / Accepted: 4 March 2025 / Published: 7 March 2025

Abstract

:
Magnesium and its rare-earth alloys are extensively studied for their lightweight properties and high specific strength, making them attractive for aerospace, automotive, and biomedical applications. However, their hexagonal close-packed structure leads to a strong basal texture, limiting plasticity and formability at room temperature. Considerable research has been devoted to texture control strategies, including alloying, thermomechanical processing, and recrystallization mechanisms, yet a comprehensive understanding of their effects remains an ongoing research focus. This review summarizes recent advances in texture regulation of rare-earth magnesium alloys, focusing on the role of RE elements (Gd, Y, Nd, Ce) and non-RE elements (Zn, Ca) in modifying basal texture and enhancing mechanical properties. The influence of key processing techniques, such as extrusion, rolling, equal channel angular pressing, and rotary shear extrusion, is discussed in relation to their effects on recrystallization behavior. Additionally, the mechanisms governing texture evolution, including continuous dynamic recrystallization, discontinuous dynamic recrystallization (DDRX), and particle-stimulated nucleation, are critically examined. By integrating recent findings, this review provides a systematic perspective on alloying strategies, processing conditions, and recrystallization pathways, offering valuable insights for the development of high-performance magnesium alloys with improved formability and mechanical properties.

1. Introduction

Magnesium (Mg) and its alloys are characterized by low density and high specific strength, making them widely used in automotive, aerospace, medical, and other industries [1,2,3,4,5]. Due to their hexagonal close-packed (HCP) crystal structure, these alloys typically develop a distinct basal texture during conventional plastic deformation. As a result, their plasticity at room temperature is limited, restricting their applications. To enhance the mechanical properties and corrosion resistance of magnesium alloys, rare-earth (RE) elements, such as gadolinium (Gd), yttrium (Y), and neodymium (Nd), are incorporated [6,7,8].
In recent years, various methods have been explored by researchers to alter the texture of alloys. These approaches aim to reduce or modify the basal texture, thereby decreasing anisotropy and ultimately enhancing the plasticity of magnesium alloys. As a result, this can lead to improved forming characteristics for magnesium alloy materials [9,10,11,12]. Commonly adopted texture control approaches for rare-earth magnesium alloys include adding alloying elements, adjusting process parameters, modifying processing methods, and regulating recrystallization processes. The addition of elements in rare-earth magnesium alloys directly affects the texture strength, orientation distribution, and recrystallization behavior. Its mechanism mainly involves solute strengthening, stacking fault energy regulation, recrystallization texture evolution, and the influence of precipitates on texture control [13,14,15]. By modifying processing methods or adjusting deformation process parameters during hot working processes, it is feasible to exert control over recrystallization processes and thereby manipulate the type and strength of deformation texture in magnesium alloys. A special deformation process can also be used to introduce shear deformation in magnesium alloys, which helps weaken the basal texture and optimize the overall alloy texture [16,17].
The typical deformation textures of rare-earth magnesium alloys are primarily determined by grain rotation due to slip and orientation shear induced by twinning. The typical deformation textures in rare-earth magnesium alloys are extrusion texture (Figure 1a) and rolling texture (Figure 1b). The texture of the extruded sheet composed of rare-earth magnesium alloy, as shown in Figure 1, closely resembles that of the rolled counterpart, with both predominantly exhibiting a (0001) basal texture. Under high-temperature extrusion, some grains experience low-angle deflection of the basal surface, forming a distinct, discrete texture [18,19]. The (0002) pole figure in Figure 1a highlights this characteristic rare-earth texture [20]. The symmetrical pattern of this texture is characterized by peak intensity occurring 30° to 40° from the normal direction (ND) towards the extrusion direction (ED). Notably, this texture shows a significantly lower pole density compared to the pronounced basal texture commonly observed in conventional AZ/AM series magnesium alloys, which are typically used in commercial applications [21,22,23]. Similarly, during the rolling process of rare-earth magnesium alloys, both Ce and Gd alloys exhibit a noticeable bimodal texture (Figure 1b) [24]. The basal poles are oriented at an angle of approximately ±20° from the normal direction of the sheet, leaning towards the rolling direction (RD). Neither of the two rolled textures demonstrates significant basal pole diffusion along the transverse direction (TD). The deflection of magnesium alloy (0001) basal poles along the rolling direction under hot rolling conditions is primarily attributed to the initiation of pyramidal slip and the formation of secondary twinning at temperatures typically above 400 °C [25,26].
Recent studies have shown that rare-earth magnesium alloys form weaker basal textures or even transition to non-basal textures after recrystallization. Rare-earth elements, as typical alloying elements, regulate the formation of recrystallization textures in magnesium alloys. The type and content of rare-earth elements significantly influence the formation of these textures [27,28,29,30]. In addition to rare-earth elements, the addition of calcium, aluminum, and zinc also has a significant impact on the recrystallization texture characteristics of rare-earth magnesium alloys [13,31,32,33,34]. Therefore, more research is focused on controlling the texture of magnesium alloys, particularly on understanding the mechanisms behind texture evolution and controlling recrystallization behavior [35,36,37,38]. From the perspective of recrystallization texture mechanisms, nucleation orientation and the competition of orientations during grain growth are key factors in forming weak or non-basal textures.
Regarding these issues, the latest research offers some solutions. Researchers have co-alloyed rare-earth elements with cost-effective elements, like Ca or Zn, utilizing the synergistic effects of multiple elements to reduce stacking fault energy, promote non-basal slip, and stimulate nucleation. Advanced processing techniques, such as asymmetric rolling and shear-assisted extrusion, introduce controlled shear strains, causing further tilting of the basal pole and achieving an isotropic Mg-Gd sheet with a tensile yield strength of up to 258 MPa [39]. Meanwhile, studies on recrystallization dynamics reveal how the segregation of rare-earth solutes at grain boundaries inhibits basal-oriented grain growth, while rare-earth-rich precipitates lead to dynamic recrystallization that produces a random texture. These findings not only clarify the relationship among elements, deformation, and recrystallization mechanisms but also establish predictive models for texture design
This paper provides a comprehensive review of the current methods for regulating the texture of rare-earth magnesium alloys, focusing on how alloying elements and processing techniques influence texture evolution. The discussion also covers the challenges faced in controlling the recrystallization behavior and the mechanisms behind texture formation. The insights gained from this review are expected to guide the development of high-performance, heat-resistant magnesium alloys.

2. Effect of Alloying Elements on Texture

The texture of rare-earth magnesium alloys is significantly influenced by the type and content of alloying elements. Different rare-earth elements exhibit distinct effects on recrystallization behavior and texture evolution. For instance, Gd and Y strongly weaken basal texture and can even introduce a tilted texture, thereby enhancing isotropy. In contrast, Nd and Ce primarily contribute to grain refinement, improving ductility. Moreover, the content of alloying elements dictates the extent of texture modification—low content (<3 wt.%) facilitates texture optimization and promotes non-basal slip, whereas high content (>5 wt.%) may lead to the precipitation of secondary phases, suppressing dynamic recrystallization and consequently affecting formability. Therefore, the judicious selection of rare-earth elements and their content is crucial for optimizing the texture and mechanical properties of magnesium alloys.

2.1. Effect of Element Types on Texture

The addition of alloying elements to magnesium alloys is an effective method for modifying texture characteristics and reducing the intensity of the basal texture. Among these, the introduction of rare-earth elements significantly alters the material’s intrinsic properties, influencing dislocation motion mechanisms and enhancing the activation of non-basal slips [36,40,41]. Commonly used rare-earth elements in magnesium alloys include Gd, Y, Nd, and Ce. These elements each have distinct effects on recrystallization behavior, leading to the formation of different textures depending on the specific rare-earth element used.
The effect of Gd element on the texture of extruded magnesium alloy can be observed in Figure 2a. The basal texture exhibits an ED-split texture. As the Gd content increases, the texture intensity gradually decreases [42]. Correspondingly, in the inverse pole figure (IPF), all sheets show strong texture components between [2 1 1 4] and [2 1 1 2]. Alloys based on Mg-RE-Zn, which are strengthened by the long-period stacking ordered (LPSO) phase, are regarded as promising materials for engineering applications due to their outstanding mechanical characteristics [43,44,45,46]. As shown in Figure 2b, a narrow-LPSO phase was produced in the Mg-13Gd-4Y-2Zn-0.5Zr magnesium alloy rod formed by forging and extrusion, and particle-stimulated nucleation (PSN) was introduced [47]. The particle promotes the dynamic recrystallization behavior, and a small quantity of bulk LPSO phase is between the crystals, thereby achieving the purpose of controlling the texture [48]. The initial coarse grains exhibiting random lamellar LPSO orientations in rare-earth magnesium alloys play a role in creating a bimodal structure, as illustrated in Figure 2c. The process involves the accumulation of dislocations from grains with varying orientations during the compression phase, the location and growth development of DRX grains, the alteration of the lamellar LPSO phase’s direction, and the evolution of the final bimodal structure. It is noted that the addition of the Gd element affects the texture and improves the mechanical characteristics of the alloy.
Among rare-earth magnesium alloys dominated by Y elements, the WE series alloys [49,50] are relatively common, with WE43 being one of the most studied. As shown in Figure 3a, the texture of the rolled WE43 rare-earth magnesium alloy does not appear at the peak position in the center but is dispersed along the TD direction at the 20–30° positions. This texture promotes the formation of {10 1 2} extension twinning, and after compression, a more dispersed basal texture appears [51]. The addition of Y not only weakens the basal texture but also inhibits basal slip, while promoting non-basal slip, which facilitates deformation at room temperature. Therefore, non-basal slip is an important deformation mechanism in WE43. As shown in Figure 3b, multiaxial forging of the WE43 rare-earth magnesium alloy induces the formation of a new <54 1 3 > rare-earth texture component after the third deformation stage, promoting a more random texture [52]. This texture evolution is attributed to either particle-stimulated nucleation mechanisms or continuous dynamic recrystallization. In the cast WE43 magnesium alloy, intermetallic compounds are distributed around the matrix, forming a rod-like eutectic structure. Upon aging, the β phase precipitates (Figure 3c), influencing the texture and enhancing the alloy’s strength and hardness [53]. Similar results have been reported in studies of the WE71 alloy. During rolling, numerous twins are formed; however, rare-earth elements inhibit dynamic recrystallization [54]. This suggests that high concentrations of Y can impede recrystallization, delay grain boundary migration, and suppress the recrystallization process. As illustrated in Figure 3d, both nucleation and grain growth inhibition contribute to the weakening of the basal texture.
In addition to elements, like Gd and Y, which are recognized for their effectiveness in optimizing texture, Ce, Yb, and Nd are often investigated for their potential to alter the texture of magnesium alloys [55,56,57,58]. In the drawing process of Mg-Nd-Zn-Zr alloy, as shown in Figure 4a, both the grain size and the second-phase particle size initially increase with the drawing temperature and then decrease [59]. The texture evolves from the rare-earth texture characteristic of the extruded sample to the basal texture, leading to a significant reduction in texture density. This change results in alloy rods with improved plasticity.
Compared to other rare-earth elements, Ce has the lowest solubility limit in Mg, but its effect on texture modification is pronounced. As shown in Figure 4b, the addition of Ce significantly improves the microstructure of the alloy, resulting in a finer grain size. Furthermore, due to the activation of non-basal slip and the occurrence of recrystallization, the alloy exhibits a more randomized texture distribution [60]. This alteration promoted the formation of alloy recrystallization, activated prismatic and pyramidal slip mechanisms, and reduced the {0001} pole density. In contrast to the mechanism by which the Ce element influences the texture of rare-earth magnesium alloys, the strength and elongation of rare-earth magnesium alloy rods containing the Yb element are primarily influenced by thermally stable nano-precipitates in the extrusion direction [61]. In comparison to magnesium (Mg), Yb possesses a larger atomic size and shares a similar atomic position with Gd in magnesium alloys. This similarity can influence the orientation and preferred growth direction of new grains during the hot extrusion process. As shown in Figure 4c, the inclined texture’s pole density is reduced, and ductility is improved.
In rare-earth magnesium alloys, non-rare-earth elements play a significant role in weakening the basal texture and can even alter the recrystallized texture [62,63]. The primary non-rare-earth elements added to these alloys are Zn and Ca. In alloys composed of Zn and various rare-earth elements (including Gd, Nd, Ce and La), as shown in Figure 5, all studied alloys exhibit increased ductility and reduced anisotropy [64]. Zn and rare-earth elements typically form precipitated phases, with Mg-Zn-Ce precipitates primarily located at grain boundaries, where they are continuously distributed. In contrast, Mg-Zn-Gd particles are smaller in size and have a lower volume fraction. In alloys containing Nd, the particles are particularly fine, with sizes smaller than 1 μm. The La alloy, on the other hand, exhibits a more complex precipitation microstructure. The addition of Zn to RE-rich Mg alloys promotes the formation of larger precipitates, such as LPSO structures. During the solution treatment, RE atoms diffuse from high-concentration regions (such as 18R-LPSO phases or the grain boundaries of eutectic phases) to low-concentration regions (the matrix) to achieve a certain equilibrium, as shown in Figure 6a. Differences in the cooling processes of the Mg-11Gd-4Y-2Zn-0.5Zr alloy result in varying LPSO structures, including both low-density and high-density 18R and 14H-LPSO phases [27]. The presence of high-density stacking faults significantly inhibits dynamic recrystallization (DRX) behavior. As a result, deformation textures with higher pole density tend to display stronger basal textures, while random dynamic recrystallization textures play a significant role in weakening the deformation texture.
We investigated the effects of adding Ca and Zn to Mg-0.6Gd, as shown in Figure 6b. Both plates displayed a texture characterized by an ED split, where the minor additions of Ca and Zn slightly reduced the peak pole density of the Mg-Gd plate, concurrently enhancing its mechanical properties [65]. Similarly, the addition of Ca will weaken the extrusion texture of magnesium alloys. The primary reason for this effect is the recrystallization induced by particles, which is triggered by second-phase particles containing calcium, leading to the creation of grains that are recrystallized with a random orientation [66,67]. In the study of the Mg-9Al-3Zn-1Mn-6Ca-2Nd alloy, two types of LPSO phase textures were identified: the LPSO phase and the C15 phase. The synergistic effect of these two phases led to grain refinement in the magnesium alloy and the development of a random texture [68]. From these studies, it can be concluded that the addition of Ca has a similar effect to that of rare-earth elements. Ca has an atomic size similar to that of rare-earth elements and exhibits high solubility in magnesium. During thermal deformation, it can produce a texture resembling that of rare-earth elements, weakening the texture. Therefore, Ca can replace some rare-earth elements, thus reducing the cost of the alloy.
In rare-earth magnesium alloys, the incorporation of rare-earth elements induces lattice distortion, thereby reducing the basal texture [69,70]. This phenomenon is tightly linked to both the c/a ratio and the stacking fault energy of the alloy. More specifically, the introduction of rare-earth elements not only lowers the c/a ratio but also elevates the stacking fault energy, thereby facilitating the activation of non-basal slip systems. Consequently, non-basal texture components are markedly enhanced. Ultimately, this change may lead to the basal pole of the texture being widely distributed in the TD.

2.2. Effect of Element Content on Texture

The composition of the alloy significantly influences the formation of RE textures. As the content of RE elements increases, the texture optimization effect becomes more pronounced, ultimately altering the original deformation texture and resulting in a typical RE texture. Wang [71] examined how varying Gd concentrations influence the mechanical characteristics of alloys. Figure 7a demonstrates that Gd merges into the magnesium matrix, which boosts the ultimate tensile strength (UTS = 344 MPa) through solid solution strengthening and activates the prismatic slip system, consequently leading to a notable enhancement in elongation (EL = 15.7%). When the Gd content reaches 13 wt.%, the grain orientation becomes increasingly random with further addition, leading to a significant reduction in texture intensity and a decrease in maximum pole density to 6.92. Changes in Gd concentration not only reduce the texture but also affect the dimensions and morphology of the precipitates found in the alloy. Kim et al. [72] investigated the microstructure and mechanical characteristics of extruded Mg-Gd alloys, with x ranging from 1 to 15 wt.%. According to Figure 7b, the authors observed marked alterations in the grain size, the precipitation phase, texture, and tensile properties of the extruded alloys as the Gd content changed. At a Gd content of 10 wt.%, the recrystallized grain size is observed to increase; conversely, in the Mg-15Gd alloy, grain size reduces due to the presence of numerous fine Mg5Gd precipitates that restrict the movement of grain boundaries. With an increase in Gd content, the orientation of the texture peak experiences a notable shift, transitioning the recrystallized region’s texture direction from <2 1 1 1> to <0001>. Furthermore, the tensile strength of the extruded alloys enhances as the Gd content rises.
Furthermore, other studies have focused on Mg alloys with high Gd contents (>5 wt.%), particularly investigating the strengthening effects of combining Gd and Y to form LPSO phases. Examples include Mg-9.8Gd-3.6Y-2.2Zn-0.4Zr [73], Mg-8.4Gd-5.3Y-1.65Zn-0.59Mn [74], and Mg-9.5Gd-4Y-2Zn-0.3Zr [75].
The Mg-1Y and Mg-5Y alloys studied by Yang et al. [76] both exhibited a weakened basal texture before ECAP. The Mg-5Y alloy consistently has a smaller grain size and exhibits a weaker basal structure. Prior to ECAP, the maximum basal splits along the TD and is tilted approximately 56° in the PF, which is a typical rare-earth texture feature. As shown in Figure 8a, after ECAP, a large number of sub-peaks are observed in the PF, with the maximum intensity either remaining unchanged or slightly reduced. The main peak of the IPF is at <11 2 0>, and the IPF intensity decreases to below 2.0 mrd (with the lowest intensity being 1.2 mrd). It is suggested that there is no significant preferential orientation. Additionally, the polarity distribution of the PF after ECAP tends to cluster around 45° between ED and TD. This leads to a softer basal slip orientation, thereby activating more basal slipping. Zhao et al. [77] studied the microstructure and tensile properties of extruded Mg-1Gd-0.5Zn-xCe sheets with varying Ce contents (0~1.2 wt.%) at room temperature. The alloy was solution-treated at 500 °C for 15 h and then extruded at 430 °C with an extrusion ratio of 32:1. As shown in Figure 8b, the addition of Ce refines the grain size and leads to the formation of a new Mg12Ce phase, hindering the growth of dynamically recrystallized grains during the extrusion process, thereby weakening the texture of the Mg-1Gd-0.5Zn-based alloy. Grain refinement and the large bulk Mg12Ce phase provide the highest strength in Mg-1Gd-0.5Zn sheets with 1.2 wt.% Ce addition.
Based on the above research, increasing RE content in magnesium alloys enhances strength but reduces elongation. Therefore, when determining the amount of rare-earth elements to add, the effects on strength and toughness should be carefully balanced. A thorough evaluation is necessary. Different amounts of non-rare-earth elements mainly influence the precipitation phase, which then affects the alloy’s texture.
The addition of a low concentration of Ca to the Mg-1Y alloy, as illustrated in Figure 9a, results in the formation of fine, dispersed second phases within the alloy. This observation suggests that the incorporation of a small amount of Ca does not significantly increase the quantity of second phases precipitating in the Mg-1Y-xCa alloy [78]. Additionally, the texture strength of the alloys shows a weakening trend with increasing Ca content. Zn is incorporated into the Mg-Gd-Y-Zn alloy, leading to the formation of the W phase as its concentration rises to 1~2 at.% [79]. During the extrusion process, the W phase not only increases in quantity but also disintegrates into particles, resulting in the loss of its characteristic herringbone shape. Figure 9b illustrates that the W phase particles and LPSO phase facilitate dynamic recrystallization through the PSN effect, leading to a weakened basal texture. The mechanical properties of alloys are significantly affected by grain size, texture, and the presence of secondary phases. The effect of non-rare-earth elements on rare-earth magnesium alloys is primarily reflected in their influence on the precipitation phase texture. Higher additions of these elements lead to the formation of coarser precipitation phases, which hinders the weakening of the alloy texture. The added amount of rare-earth elements, such as Nd, Y, and Ce, is determined by their solubility in magnesium alloys. These elements weaken the texture by forming deformation bands containing twins and restricting grain growth [80]. Different elements exhibit varying solid solubility in rare-earth magnesium alloys. As the content of these elements increases, the strength of the alloy decreases, a trend that remains unaffected by heat treatment. The critical addition number of elements in the alloy is not constant and is influenced by the deformation method employed. Consequently, selecting an appropriate processing technology is crucial for controlling the alloy’s texture.

3. Effect of Processing on Texture of Rare-Earth Magnesium Alloys

This chapter explores the relationship between processing parameters, deformation methods, and texture evolution, analyzing the effects of extrusion ratio, extrusion temperature, rolling temperature, and strain rate on texture. It also introduces advanced plastic forming techniques, such as rotary shear extrusion and cross-rolling, explaining how they regulate texture through specific stress states and deformation mechanisms. By systematically examining the influence of different processing methods, this chapter provides theoretical support and technical guidance for optimizing the plasticity and strength of rare-earth magnesium alloys.

3.1. Processing Parameter

Various processing methods significantly affect the texture characteristics and mechanical properties of magnesium alloys by altering grain structure, recrystallization behavior, and slip system activation. Typically, the dynamic recrystallization of magnesium alloys can be regulated by modifying the process parameters, which in turn reduces the strength of the basal texture [81,82]. A higher extrusion ratio (e.g., increasing from 10 to 25) during the extrusion process helps activate various slip systems in magnesium alloys, leading to a weaker basal texture or the formation of non-basal textures. As shown in Figure 10a, the Mg-Gd-Mn alloys exhibit a bimodal texture characterized by a combination of equiaxed recrystallized grains and elongated deformed grains, which varies with different extrusion ratios [83]. In the VM20ER25 alloy, the texture components exhibit a relatively dispersed distribution, presenting a typical rare-earth texture, where most grain c-axes are inclined at 45° along the ED. This specific orientation facilitates the activation of {10 1 2} tensile twins. When the extrusion ratio is reduced to 10, the texture characteristics slowly disappear. This is primarily due to the increased proportion of non-recrystallization materials. Increasing the extrusion ratio effectively enhances the elongation of the alloy but reduces its tensile strength. Similarly, as shown in Figure 10b, the Mg-Gd-Y-Sm-Zr alloy undergoes the extrusion–shear (ES) forming process, and a comparison of the tensile properties at extrusion ratios (λ) of 9, 16, 25, and 36 reveals that λ = 25 exhibits the highest ultimate tensile strength (303 MPa), yield strength (216 MPa), and elongation (25.47%). The superior tensile performance is attributed to the synergistic effect of grain refinement and texture weakening [84].
Different from the extrusion deformation, the volume of the second phase changes with the deformation process during the rolling process. In the rolled Mg-7Y-5Zn-0.1V alloy after forging, as shown in Figure 11, the 18R-LPSO phase increases with deformation, while the 14H-LPSO phase dissolves during hot deformation [85]. The rolling process is carried out at 450 °C, with total deformation amounts of 30%, 40%, 50%, and 60%. Figure 11b shows that in alloys subjected to significant rolling deformation, no typical texture is present, and the texture intensity is reduced. Figure 11c illustrates the structural evolution from forging to rolling. The second-phase particles influence the recrystallization process, which in turn affects texture formation. The direct-chill-cast WE43 magnesium alloy was hot-rolled at 480 °C [86]. As the reduction increased, the strength of the basal texture significantly decreased, but typical basal and rare-earth textures did not form. The incomplete recrystallization of the microstructure is considered to be responsible for the formation of this non-basal texture [87,88,89].
During alloy processing, temperature is a crucial parameter for regulating the texture. Higher temperatures promote the initiation of non-basal surface slips. As shown in Figure 12, Ling et al. [90] studied the Mg-9.5Gd-4Y-2Zn-0.5Zr alloy, which was extruded at 480–500 °C with an extrusion rate of 0.2–0.4 mm/s and an extrusion ratio of 23:1 to serve as the billet. The deformation behavior was further investigated at strain rates of 0.001 s−1, 0.01 s−1, 0.1 s−1, and 1 s−1, as well as deformation temperatures of 350 °C, 400 °C, 450 °C, and 500 °C. The texture features multiple high-pole-density {0001} peaks aligned along the normal axis, centered around the CD, which can be attributed to non-basal slip promoted by rare-earth elements. As the deformation temperature increases, the maximum pole density of {0001} first increases, then decreases, and then increases again. This phenomenon is explained by two effects: (1) the replacement of randomly oriented DRX grains produced by thermal deformation weakens the alloy texture; (2) at low strain rates, the rotation and growth of DRX grains reduce grain orientation randomness. Chen et al. [91] found that in Mg-Zn-Y alloy, higher temperatures reduce the torsion of LPSO, increase the recrystallized area, and significantly weaken the texture’s pole density. The effects of different strain rates on texture were also analyzed, with results similar to those at varying temperatures. As the strain rate increases, the maximum pole density increases, indicating that higher strain rates weaken dynamic recrystallization. The resulting uneven microstructure forms more sub-grains and deformed grains, with a stronger {0001} texture distribution at an angle to the CD. Therefore, when selecting processing methods, the setting of processing parameters (deformation temperature, deformation rate, and deformation extent) is the main factor influencing the texture of rare-earth magnesium alloys and plays a decisive role in the recrystallization ratio and texture morphology after processing.

3.2. Special Deformation Mode

In recent years, significant advancements have been made in the plastic forming methods of magnesium alloys, including equal channel angular pressing (ECAP) [92], rotary shear extrusion, accumulative roll bonding [93], cross-rolling, electro-plastic forming [94], differential speed rolling, and liner protection rolling [95]. While the deformation mechanisms in each process may differ, all of them contribute to either the weakening of the basal texture or the promotion of non-basal texture formation during deformation.
Rotary shear extrusion (RSE) technology simultaneously introduces torsion and shear deformation, offering advantages, such as low forming load, high strain, and more complete deformation. The Mg-9Gd-4Y-2Zn-0.5Zr alloy was processed at 420 °C using the RSE method (Figure 13a), which effectively eliminated the lamellar and bulk LPSO phases, resulting in a more uniform distribution of the particle phase within the grain [39]. The RSE-treated alloy exhibited a single texture with a strong [2 1 1 0] orientation, significantly weakening the texture intensity. Notably, weak texture components were observed at the [2 1 1 4] and [2 1 1 0] orientations (Figure 13), which are typically associated with rare-earth components. The RSE process also induced DRX, resulting in a smaller overall grain size [96]. Similarly, the microstructure of Mg-3.8Y-2.6 RE-0.45Zr alloy processed by ECAP shows that the grain nucleation volume fraction increases, but the grain growth is inhibited, and the final grain size decreases. The findings indicate that the substantial deformation procedure is capable of efficiently refining the grains and decreasing the texture density.
We conducted a comparison of the microstructure and mechanical properties of Mg-8Li-6Zn-1Y plates produced by unidirectional rolling (UR) and cross-rolling (CR) technologies. The process involves four rolling passes at 493 K, with a total thickness reduction of 8 mm and a 20% reduction per pass [97]. The schematic diagrams of the two rolling methods are shown in Figure 14. In this study, after UR, the fractured I-phase particles tend to aggregate at the α-Mg/β-Li phase interfaces. However, the CR process not only severely fragments the bulk I-phase but also results in a more uniform distribution of I-phase particles within the matrix phases. The range of the fine-grained areas in CR samples is larger, and the extreme densities of the textures of the two rolling methods are not very different at this stage. The Mg-10Gd-2Y-0.4Ag-0.4Zr alloy exhibits enhanced dynamic recovery in CR plates, which diminishes the driving force for dynamic recrystallization, thereby leading to an increase in grain size [98]. Rare-earth magnesium alloys typically exhibit a TD split texture following hot rolling, characterized by a deviation from the ND to the TD. The texture of the UR plate primarily consists of the base texture, whereas the base texture of the CR plate displays a distinct texture distribution, with its c-axis deviating between the TD and the RD. This variation is attributed to the influence of the rolling path. Some studies have linked this texture change to the PSN mechanism [87,99] or the SBN mechanism [100,101,102].

4. Effect of Recrystallization on Texture

The optimization of rare-earth magnesium alloys through recrystallization primarily aims to weaken the basal texture and promote the formation of non-basal textures. Two common mechanisms of recrystallization in magnesium alloys are continuous dynamic recrystallization (CDRX) and discontinuous dynamic recrystallization (DDRX) [18,103].
The dynamic recovery of dislocations encourages CDRX, as local stress prompts the development of sub-grain boundaries. Various dislocation types contribute to the rotation of these sub-grains. At the grain boundaries, the presence of rare-earth elements hampers the movement of original grain boundaries, thereby preventing DDRX. Moreover, incorporating rare-earth elements increases the cross-slip capability of dislocations, which notably enhances the strain capacity of magnesium alloys along the c-axis. As shown in Figure 15, the presence of rare-earth elements at the boundaries of grains obstructs the movement of these boundaries and restricts DDRX. [104]. The newly formed grains via dynamic recrystallization typically maintain a comparable orientation to the deformed grains, leading to some degree of texture weakening; however, this influence is not substantial enough to markedly change the type of texture. [105,106]. The formation mechanism of DDRX involves the bulging of original high-angle grain boundaries (HAGBs), which are then subdivided by low-angle grain boundaries (LAGBs), forming dislocation walls at the bulging sites [107]. As the strain increases, dislocations accumulate at these protruding walls, eventually transforming into new HAGBs and recrystallized grains. In rare-earth magnesium alloys, DDRX is often hindered by the precipitation of rare-earth phases, making it difficult to form, and the newly developed grains typically preserve an orientation similar to that of the matrix, restricting their capacity to diminish the texture.
Recrystallization nucleation is a critical mechanism for generating newly oriented grains and plays a key role in the texture optimization of magnesium alloys. During the recrystallization process that results in the development of rare-earth textures, the main mechanisms involved include twin-mediated nucleation, shear band nucleation, and particle-stimulated nucleation.
Twin-mediated recrystallization refers to a mechanism in which plastically deformed twins in magnesium alloys serve as nucleation sites. The orientation of certain twins differs significantly from that of the matrix, and they store large amounts of deformation energy. Consequently, these twins become key sites for subsequent recrystallization nucleation. As shown in Figure 16, larger strains and stresses accumulate at grain boundaries, twin boundaries, and dislocation entanglements [108]. The main twin type is {10 1 2} extension twins, and most LAGBs are concentrated near grain boundaries and their intersections, where they can act as nucleation points for recrystallization during annealing. Twins can alter the orientation of grains, converting hard orientations to softer basal slip orientations, thus facilitating more basal slip to accommodate deformation. The rotation of the slip surface in the rolling plane is induced by the basal slip, which results in the creation of extra dislocations that engage with grain boundaries, twins, and various other dislocations. This interaction increases the accumulated stress, which provides the driving force for static recrystallization during the subsequent annealing process. While undergoing annealing, dislocations that are densely packed move and intertwine, ultimately leading to the creation of LAGBs, sub-grains, and HAGBs, which subsequently develop into recrystallized grains. The process of recrystallization modifies the original grain orientations, resulting in a non-basal texture that somewhat diminishes the strength of the basal texture.
Shear bands are microstructures that form in alloys after undergoing severe deformation. Due to the localized plastic flow within these bands, they store significant strain energy, making them potential nucleation sites for recrystallization [109]. During the deformation process, several shear bands containing finer dynamic recrystallized grains are generated, as shown in Figure 17. In the shear band, the grains resulting from dynamic recrystallization display distinct characteristics of basal texture, with the c-axis orientation inclined at an angle of about ±20° relative to the normal direction of the rolling process. Shear bands play a crucial role in developing rare-earth textures, which is linked to the nucleation of recrystallized grains occurring within these bands [110]. They play a pivotal role in the deformation mechanisms during the finishing rolling process. Nevertheless, the processes controlling the initiation and spread of shear bands are not yet fully understood and require additional study.
Particle-stimulated nucleation describes a phenomenon where the presence of stable second-phase particles causes high strain accumulation around them due to dislocation pinning and other effects during deformation [111,112,113,114]. This accumulation promotes the preferential recrystallization and nucleation of these particles during thermal deformation or subsequent annealing. Research indicates that the PSN mechanism activates solely when second-phase particles exceed a certain size (>1 μm). Zhang et al. [115] investigated the extruded Mg-6Zn-0.5Zr-xNd (wt.%) alloy and found that the addition of Nd enhanced the precipitation process, leading to an increase in the size of the second-phase particles. As shown in Figure 18a, some second-phase particles exceeded 1 μm, which clearly facilitated the PSN mechanism. For the alloy with 4 wt.% Nd, the volume fraction of recrystallized grains increased, and the {0001} pole density slightly decreased. As observed by Salandari-Rabori et al. [52], particles larger than 1 μm effectively hindered dislocation movement, accelerated recrystallization, and promoted the formation of randomly oriented recrystallized grains. Similarly, some research noted that two large LPSO phases in the alloy stimulated substantial recrystallization, resulting in a maximum recrystallization fraction of 42.6% [26]. The recrystallization process of three different LPSO phase configurations is shown in Figure 18c. High-rare-earth magnesium alloys typically contain multiple second phases, making particle-stimulated nucleation a significant recrystallization nucleation mechanism. Although recrystallization occurs extensively, the overall texture density is not significantly weakened. Therefore, further research is needed to understand the role of PSN in texture weakening in magnesium alloys.

5. Conclusions

This review systematically summarizes the influence of alloying elements, processing methods, and recrystallization mechanisms on the texture evolution of rare-earth magnesium alloys. The findings highlight that rare-earth (RE) elements, such as Gd, Y, Nd, and Ce, play a critical role in weakening the basal texture by promoting non-basal slip and affecting recrystallization behavior. Among them, Gd significantly reduces the basal pole density from 12.3 to 6.92 when its content increases to 13 wt.% in Mg alloys. Additionally, the incorporation of Y in WE-series alloys leads to texture randomization, particularly at concentrations above 5 wt.%, where a new <5 4 1 3 > texture component forms, reducing anisotropy.
Processing conditions also exert a substantial influence on texture modification. Increasing the extrusion ratio from 10 to 25 in Mg-Gd-Mn alloys leads to a notable decrease in texture intensity and an improvement in elongation, reaching 25.47%. Similarly, rolling at higher temperatures above 400 °C activates pyramidal slip and secondary twinning, reducing basal texture alignment. In extrusion–shear (ES) processing, an extrusion ratio of 25 results in an optimal balance of yield strength (216 MPa) and ultimate tensile strength (303 MPa), along with significant texture weakening.
Recrystallization mechanisms, such as particle-stimulated nucleation and twin-mediated recrystallization, further influence texture evolution. Second-phase particles exceeding 1 μm effectively trigger PSN, increasing recrystallized grain volume fraction and weakening basal texture. In Mg-6Zn-0.5Zr-xNd alloys, adding 4 wt.% Nd increases the recrystallization fraction while reducing the {0001} pole density. Moreover, in alloys containing LPSO phases, the recrystallization fraction reaches 42.6%, further contributing to texture modification.
In summary, controlling the texture of rare-earth magnesium alloys requires a comprehensive approach combining alloy composition, optimized processing parameters, and recrystallization mechanisms. Future research should focus on refining processing techniques, such as rotary shear extrusion and accumulative roll bonding, to further optimize texture and enhance mechanical properties, ensuring the development of high-performance magnesium alloys with improved ductility and reduced anisotropy.

Author Contributions

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

Funding

This research was funded by the Natural Science Foundation of Jiangsu Province (BK20232011).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Li, K.; Chen, W.; Yin, B.; Ji, C.; Bai, S.; Liao, R.; Yang, T.; Wen, P.; Jiang, B.; Pan, F. A comparative study on WE43 magnesium alloy fabricated by laser powder bed fusion coupled with deep cryogenic treatment: Evolution in microstructure and mechanical properties. Addit. Manuf. 2023, 77, 103814. [Google Scholar] [CrossRef]
  2. Li, M.; Yang, Q.; Zhao, Z.; Zhang, J.; Li, J. Full liquid phase sintering of binder jetting printed magnesium alloy. J. Manuf. Process. 2023, 108, 194–203. [Google Scholar] [CrossRef]
  3. Pang, S.; Zhao, W.; Qiu, T.; Liu, W.; Jiao, L.; Wang, X. Study on surface quality and mechanical properties of micro-milling WE43 magnesium alloy cardiovascular stent. J. Manuf. Process. 2023, 101, 1080–1090. [Google Scholar] [CrossRef]
  4. Weiler, J.P. The role of the Mg17Al12-phase in the high-pressure die-cast magnesium-aluminum alloy system. J. Magnes. Alloys 2023, 11, 4235–4246. [Google Scholar] [CrossRef]
  5. Li, R.; Liu, Q.; Zhang, R.; Li, Y.; Ma, Y.; Huo, H.; Gao, Y.; Zuo, P.; Wang, J.; Yin, G. Achieving high-energy-density magnesium/sulfur battery via a passivation-free Mg-Li alloy anode. Energy Storage Mater. 2022, 50, 380–386. [Google Scholar] [CrossRef]
  6. Mei, W.; Li, Q.; Chen, X. Atomic simulation study on the effect of Y atom and grain boundary on tensile deformation in polycrystalline magnesium alloy. J. Mater. Res. Technol. 2023, 23, 931–942. [Google Scholar] [CrossRef]
  7. Arrabal, R.; Mingo, B.; Pardo, A.; Matykina, E.; Mohedano, M.; Merino, M.C.; Rivas, A.; Maroto, A. Role of alloyed Nd in the microstructure and atmospheric corrosion of as-cast magnesium alloy AZ91. Corros. Sci. 2015, 97, 38–48. [Google Scholar] [CrossRef]
  8. Wang, R.; Xue, S.; Zhou, W.; Wang, X.; Li, Z.; Meng, Z.; Ma, L. Microstructure evolution and springback behavior in single-pass roll bending of magnesium alloy plates. J. Alloys Compd. 2024, 1009, 176935. [Google Scholar] [CrossRef]
  9. Du, Y.-z.; Liu, D.-j.; Ge, Y.-f.; Jiang, B.-l. Effects of deformation parameters on microstructure and texture of Mg–Zn–Ce alloy. Trans. Nonferrous Met. Soc. China 2020, 30, 2658–2668. [Google Scholar] [CrossRef]
  10. Shokri, M.; Zarei-Hanzaki, A.; Abedi, H.R.; Cho, J.H. On the microstructure and RE-texture evolution during hot tensile deformation of Mg-Gd-Y-Zn-Zr alloy. J. Mater. Res. Technol. 2021, 15, 6974–6989. [Google Scholar] [CrossRef]
  11. Wang, C.; Ning, H.; Liu, S.; You, J.; Wang, T.; Jia, H.-J.; Zha, M.; Wang, H.-Y. Enhanced ductility and strength of Mg-1Zn-1Sn-0.3Y-0.2Ca alloy achieved by novel micro-texture design. Scr. Mater. 2021, 204, 114119. [Google Scholar] [CrossRef]
  12. Zhao, L.-Q.; Wang, C.; Chen, J.-C.; Ning, H.; Yang, Z.-Z.; Xu, J.; Wang, H.-Y. Development of weak-textured and high-performance Mg–Zn–Ca alloy sheets based on Zn content optimization. J. Alloys Compd. 2020, 849, 156640. [Google Scholar] [CrossRef]
  13. Malik, A.; Nazeer, F.; Abrar, S.; Cheng, C.; Chen, Z.; Fouda, A.M. Interfacial bonding, texture evolution, and mechanical properties of different structured extruded Mg-7Gd-4Y-1Zn-0.2Zr and Mg-3Zn-1Al-1Ca-0.5Mn composites. J. Magnes. Alloys 2024. [Google Scholar] [CrossRef]
  14. Wu, Y.P.; Zhang, X.M.; Deng, Y.L.; Tang, C.P.; Zhong, Y.Y. Effect of compression conditions on the microstructure and texture of a Mg–RE alloy. Mater. Sci. Eng. A 2015, 644, 152–158. [Google Scholar] [CrossRef]
  15. Zhang, G.; Meng, Y.; Yan, F.; Gao, Z.; Yan, Z.; Zhang, Z. Microstructure and texture evolution of Mg-RE-Zn alloy prepared by repetitive upsetting-extrusion under different decreasing temperature degrees. J. Alloys Compd. 2020, 815, 152452. [Google Scholar] [CrossRef]
  16. Wu, Y.-p.; Xiong, H.-q.; Jia, Y.-z.; Xie, S.-h.; Li, G.-f. Microstructure, texture and mechanical properties of Mg–8Gd–4Y–1Nd–0.5Zr alloy prepared by pre-deformation annealing, hot compression and ageing. Trans. Nonferrous Met. Soc. China 2019, 29, 976–983. [Google Scholar] [CrossRef]
  17. Zhao, L.; Ma, G.; Jin, P.; Yu, Z. Role of Y on the microstructure, texture and mechanical properties of Mg–Zn–Zr alloys by powder metallurgy. J. Alloys Compd. 2019, 810, 151843. [Google Scholar] [CrossRef]
  18. Jamili, A.M.; Zarei-Hanzaki, A.; Abedi, H.R.; Minárik, P. The grain boundary character distribution in thermomechanically processed rare earth bearing magnesium alloy. J. Alloys Compd. 2019, 798, 158–166. [Google Scholar] [CrossRef]
  19. Deng, J.-F.; Tian, J.; Zhou, Y.; Chang, Y.; Liang, W.; Ma, J. Plastic deformation mechanism and hardening mechanism of rolled Rare-Earth magnesium alloy thin sheet. Mater. Des. 2022, 218, 110678. [Google Scholar] [CrossRef]
  20. Li, T.; Rao, J.; Zheng, J.; Yin, D.; Shou, H.; Zhang, Y.; Shi, R.; Jing, W.; Xia, L. Anisotropic cyclic deformation behavior of an extruded Mg-3Y alloy sheet with rare earth texture. J. Magnes. Alloys 2022, 10, 1581–1597. [Google Scholar] [CrossRef]
  21. Akyuz, B. Influence of Al content on machinability of AZ series Mg alloys. Trans. Nonferrous Met. Soc. China 2013, 23, 2243–2249. [Google Scholar] [CrossRef]
  22. Liu, F.-J.; Ji, Y.; Bai, Y.-X. Influence of multipass high rotating speed friction stir processing on microstructure evolution, corrosion behavior and mechanical properties of stirred zone on AZ31 alloy. Trans. Nonferrous Met. Soc. China 2020, 30, 3263–3273. [Google Scholar] [CrossRef]
  23. Gronostajski, Z.; Kaczyński, P.; Polak, S.; Bartczak, B. Energy absorption of thin-walled profiles made of AZ31 magnesium alloy. Thin-Walled Struct. 2018, 122, 491–500. [Google Scholar] [CrossRef]
  24. Basu, I.; Al-Samman, T. Triggering rare earth texture modification in magnesium alloys by addition of zinc and zirconium. Acta Mater. 2014, 67, 116–133. [Google Scholar] [CrossRef]
  25. Zhou, M.; Xu, Y.; Liu, Y.; Duan, M.; Xia, Z.; Huang, L.; Zhu, R.; Ye, H.; Peng, L.; Wu, Y.; et al. Microstructures and mechanical properties of Mg-15Gd-1Zn-0.4Zr alloys treated by ultrasonic surface rolling process. Mater. Sci. Eng. A 2021, 828, 141881. [Google Scholar] [CrossRef]
  26. Han, F.; Luo, X.; Liu, Q.; Hou, Z.; Marthinsen, K.; Wu, G.L.; Hatzoglou, C.; Kontis, P.; Huang, X. The effect of grain size and rolling reduction on microstructure evolution and annealing hardening response of a Mg-3Gd alloy. J. Magnes. Alloys 2023. [Google Scholar] [CrossRef]
  27. Zheng, C.; Chen, S.F.; Cheng, M.; Zhan, S.; Li, Y.; Yang, Y.S. Controlling dynamic recrystallization via modified LPSO phase morphology and distribution in Mg-Gd-Y-Zn-Zr alloy. J. Magnes. Alloys. 2023, 11, 4218–4234. [Google Scholar] [CrossRef]
  28. Prasad, A.; Lee, S.Y.; Singh, S.S.; Gosvami, N.N.; Jain, J. Texture weakening and enhanced recrystallization kinetics in a La containing Mg–Dy alloy. Mater. Chem. Phys. 2022, 277, 125537. [Google Scholar] [CrossRef]
  29. Molodov, K.D.; Al-Samman, T.; Molodov, D.A. Effect of gadolinium on the deformation and recrystallization behavior of magnesium crystals. Acta Mater. 2022, 240, 118312. [Google Scholar] [CrossRef]
  30. Abbasi, M.; Bagheri, B.; Sharifi, F. Simulation and experimental study of dynamic recrystallization process during friction stir vibration welding of magnesium alloys. Trans. Nonferrous Met. Soc. China 2021, 31, 2626–2650. [Google Scholar] [CrossRef]
  31. Tan, M.; Wang, J.; Wu, Y.; Jin, C.; Sun, Y.; Song, L.; Zhang, Y.; Wang, J.; Huo, J.; Gao, M. Accelerated crystallization kinetics and grain refinement in Mg-Ni-Y metallic glass via multiple rare earth elements doping. J. Alloys Compd. 2024, 999, 175080. [Google Scholar] [CrossRef]
  32. Wei, Q.; Yuan, L.; Ma, X.; Zheng, M.; Shan, D.; Guo, B. Strengthening of low-cost rare earth magnesium alloy Mg-7Gd-2Y–1Zn-0.5Zr through multi-directional forging. Mater. Sci. Eng. A 2022, 831, 142144. [Google Scholar] [CrossRef]
  33. Kan, Z.; Yang, G.; Xiao, L.; Wang, C.; Qin, H.; Wang, B.; Jie, W. Asymmetric tension-compression mechanical behaviors of extruded Mg-4.58Zn-2.6Gd-0.18Zr alloy with double-peak fiber texture. J. Alloys Compd. 2024, 989, 174291. [Google Scholar] [CrossRef]
  34. Yang, M.; Jia, H.-L.; Jiang, R.; Zhou, X.-L.; Ma, P.-K.; Li, Z.-G.; Zha, M.; Wang, H.-Y. Excellent synergy of formability and strength of a Mg-Zn-Y-Ca-Zr alloy by tailoring segregation-assisted weak elliptical ring texture. J. Magnes. Alloys 2024. [Google Scholar] [CrossRef]
  35. Li, J.R.; Xie, D.S.; Zeng, Z.R.; Song, B.; Xie, H.B.; Pei, R.S.; Pan, H.C.; Ren, Y.P.; Qin, G.W. Mechanistic investigation on Ce addition in tuning recrystallization behavior and mechanical property of Mg alloy. J. Mater. Sci. Technol. 2023, 132, 1–17. [Google Scholar] [CrossRef]
  36. Li, M.; Wang, X.; Xiao, Z.; Liu, Y.; Huang, Y. Effect of dynamic recrystallization, LPSO phase, and grain boundary segregation on heterogeneous bimodal microstructure in Mg-9.8Gd-3.5Y-2.0Zn-0.4Zr alloy. J. Mater. Res. Technol. 2023, 26, 3863–3880. [Google Scholar] [CrossRef]
  37. Zhang, D.; Liu, C.; Jiang, S.; Gao, Y.; Wan, Y.; Chen, Z. Effects of dynamic recrystallization mechanisms on texture evolution in Mg-Gd-Y-Zr-Ag alloy during hot compression. J. Alloys Compd. 2023, 944, 169190. [Google Scholar] [CrossRef]
  38. Zhang, N.; Liu, M.; Li, Q.; Chen, X.; Zheng, Z. Effect of long-period stacking ordered phase on dynamic recrystallization in Mg-Y-Zn-Zr alloy processed by backward extrusion. J. Mater. Res. Technol. 2023, 26, 2384–2393. [Google Scholar] [CrossRef]
  39. Yan, L.; Zhang, Z.; Xue, Y.; Xu, J.; Dong, B.; Li, X. Effect of rotating shear extrusion on the microstructure, texture evolution and mechanical properties of Mg-Gd-Y-Zn-Zr alloy. J. Alloys Compd. 2022, 906, 164406. [Google Scholar] [CrossRef]
  40. Zheng, Y.-q.; Zhang, Y.; Tian, Y.-q.; Chen, L.-s. Microstructure characteristics and degradation behaviors of medical rare-earth Mg-10.6Gd-0.3Ag-x(Y/Zn) alloys in vitro. J. Alloys Compd. 2024, 997, 174967. [Google Scholar] [CrossRef]
  41. Zheng, Z.; Dong, Z.; Jiang, B.; Cheng, Y.; Zhang, A.; Song, J.; Li, T.; Pan, F. Evolution of strength with rare-earth content in highly-alloyed Mg-Gd-Y alloys. Scr. Mater. 2024, 238, 115772. [Google Scholar] [CrossRef]
  42. Zhao, J.; Jiang, B.; Xu, J.; He, W.; Huang, G.; Pan, F. The influence of Gd on the recrystallisation, texture and mechanical properties of Mg alloy. Mater. Sci. Eng. A 2022, 839, 142867. [Google Scholar] [CrossRef]
  43. Deng, Y.; Yan, H.; Li, Q.; Chen, J.; Xia, W.; Su, B.; Wu, M.; Yu, Y.; Song, M. Enhancing strength and ductility of low RE content Mg-Gd-Y-Zr alloy via a novel thermomechanical treatment based on multi-directional forging. J. Alloys Compd. 2023, 958, 170535. [Google Scholar] [CrossRef]
  44. Fei, H.-j.; Xu, G.-l.; Liu, L.-b.; Bo, H.; Zeng, L.-j.; Chen, C.-p. Phase equilibria in Mg-rich corner of Mg–Ca–RE (RE=Gd, Nd) systems at 400 °C. Trans. Nonferrous Met. Soc. China 2013, 23, 881–888. [Google Scholar] [CrossRef]
  45. Luo, K.; Zhang, L.; Wu, G.; Liu, W.; Ding, W. Effect of Y and Gd content on the microstructure and mechanical properties of Mg–Y–RE alloys. J. Magnes. Alloys 2019, 7, 345–354. [Google Scholar] [CrossRef]
  46. Luo, X.; Yang, H.; Zhou, J.; Jiang, B.; Feng, Q.; Zeng, Y.; Li, W.; Dong, Z.; Song, J.; Xu, J.; et al. Achieving outstanding heat-resistant Mg-Gd-Y-Zn-Mn alloy via introducing RE/Zn segregation on α-Mn nanoparticles. Scr. Mater. 2023, 236, 115672. [Google Scholar] [CrossRef]
  47. He, Y.; Guo, R.; Zhao, X.; Zhang, Z. High strength and high ductility Mg-Gd-Y-Zn-Zr alloys obtained by controlling texture and dynamic precipitation through LPSO phase structure of different initial morphologies. J. Rare Earths 2024, 42, 2217–2230. [Google Scholar] [CrossRef]
  48. Meng, D.; Xu, Y.; Jia, J.; Luo, J.; Xu, B. Inhomogeneous deformation behavior and recrystallization mechanism of Mg-Gd-Y-Zn-Zr alloy containing only intragranular lamellar LPSO phase. J. Alloys Compd. 2024, 1003, 175713. [Google Scholar] [CrossRef]
  49. Chen, W.Q.; Deng, S.; Zhao, X.J.; Luo, S.N. Active {11 2 1} and {10 1 2} extension twinning in a WE43 Mg alloy. J. Mater. Sci. Technol. 2023, 144, 93–101. [Google Scholar] [CrossRef]
  50. Yu, Z.; Xu, X.; Shi, K.; Du, B.; Han, X.; Xiao, T.; Li, S.; Liu, K.; Du, W. Development and characteristics of a low rare-earth containing magnesium alloy with high strength-ductility synergy. J. Magnes. Alloys 2023, 11, 1629–1642. [Google Scholar] [CrossRef]
  51. Xiang, C.; Xu, Y.; Yang, Y.; Ding, H.; Wang, Z. Plastic deformation mechanisms and constitutive modeling of WE43 magnesium alloy at various strain rates and temperatures. J. Mater. Res. Technol. 2024, 29, 4110–4128. [Google Scholar] [CrossRef]
  52. Salandari-Rabori, A.; Zarei-Hanzaki, A.; Abedi, H.R.; Lecomte, J.S.; Khatami-Hamedani, H. Micro and macro texture evolution during multiaxial forging of a WE43 magnesium alloy. J. Alloys Compd. 2018, 739, 249–259. [Google Scholar] [CrossRef]
  53. Jiang, H.S.; Zheng, M.Y.; Qiao, X.G.; Wu, K.; Peng, Q.Y.; Yang, S.H.; Yuan, Y.H.; Luo, J.H. Microstructure and mechanical properties of WE43 magnesium alloy fabricated by direct-chill casting. Mater. Sci. Eng. A 2017, 684, 158–164. [Google Scholar] [CrossRef]
  54. Liu, J.B.; Zhang, K.; Han, J.T.; Li, X.G.; Li, Y.J.; Ma, M.L.; Yuan, J.W.; Shi, G.L.; Li, M.; Lu, C.F. Study on the microstructure and mechanical Properties of WE71 magnesium alloy. Mater. Sci. Eng. A 2015, 625, 107–113. [Google Scholar] [CrossRef]
  55. Gao, S.; Wei, X.; Le, Q.; Ren, L.; Hu, J.; Wu, J.; Wang, T.; Liao, Q.; Bao, L. Magnesium alloys strengthened with a combination of light rare earth element Ce and heavy rare earth element Y. Mater. Today Commun. 2025, 43, 111682. [Google Scholar] [CrossRef]
  56. Woo, S.K.; Pei, R.; Al-Samman, T.; Letzig, D.; Yi, S. Plastic instability and texture modification in extruded Mg-Mn-Nd alloy. J. Magnes. Alloys 2022, 10, 146–159. [Google Scholar] [CrossRef]
  57. Zhang, Z.; Kim, J.; Pu, H.; Zhou, S.; Hu, Y.; Huang, G.; Jiang, B.; Pan, F. Improving mechanical properties of Mg–Zn-Nd-Zr alloy by low alloying with Yb and corresponding strengthening mechanisms. J. Mater. Res. Technol. 2024, 33, 2023–2034. [Google Scholar] [CrossRef]
  58. Li, L.; Zhang, C.; Lv, H.; Liu, C.; Wen, Z.; Jiang, J. Texture development and tensile properties of Mg–Yb binary alloys during hot extrusion and subsequent annealing. J. Magnes. Alloys 2022, 10, 249–265. [Google Scholar] [CrossRef]
  59. Zhou, B.; Wang, X.; Jiang, J.; Chen, C.; Zhang, X.; Niu, J.; Pei, J.; Huang, H.; Wang, D.; Yuan, G. Superior strength-ductility synergy of Mg-Nd-Zn-Zr alloy rod achieved by drawing at elevated temperatures. Int. J. Plast. 2024, 175, 103944. [Google Scholar] [CrossRef]
  60. Zhang, S.; Fan, Y.; Zhao, H.; Dong, X.; Guo, C.; Zhang, S.; Weng, K.; Xiao, Y. A comparative study on the effect of 0.5 wt% Ca or Ce on the microstructure, texture, and mechanical properties of the hot-rolled Mg-4Li alloy. Mater. Today Commun. 2024, 39, 108540. [Google Scholar] [CrossRef]
  61. Li, L.; Wang, Y.; Zhang, C.; Wang, T.; Lv, H.; Yu, W. Ultrafine-grained Mg–Zn–Yb–Zr alloy with simultaneously improved strength and ductility processed by axisymmetric hot extrusion. Vacuum 2020, 173, 109157. [Google Scholar] [CrossRef]
  62. Xie, H.; Wang, J.; Li, Y.; Song, J.; Hu, H.; Qin, L.; Zhao, H.; Li, C.; Cui, Y.; Tan, Z.; et al. Fast shot speed induced microstructure and mechanical property evolution of high pressure die casting Mg-Al-Zn-RE alloys. J. Mater. Process. Technol. 2024, 331, 118523. [Google Scholar] [CrossRef]
  63. Mouhib, F.; Pei, R.; Erol, B.; Sheng, F.; Korte-Kerzel, S.; Al-Samman, T. Synergistic effects of solutes on active deformation modes, grain boundary segregation and texture evolution in Mg-Gd-Zn alloys. Mater. Sci. Eng. A 2022, 847, 143348. [Google Scholar] [CrossRef]
  64. Al-Samman, T.; Li, X. Sheet texture modification in magnesium-based alloys by selective rare earth alloying. Mater. Sci. Eng. A 2011, 528, 3809–3822. [Google Scholar] [CrossRef]
  65. Jun, Z.; Bin, J.; Yuan, Y.; Wang, Q.; Yuan, M.; Tang, A.; Huang, G.; Zhang, D.; Fusheng, P. Understanding the enhanced ductility of Mg-Gd with Ca and Zn microalloying by slip trace analysis. J. Mater. Sci. Technol. 2021, 95, 20–28. [Google Scholar] [CrossRef]
  66. Wang, L.; Hu, S.; Xi, C.; Zhou, W.; Sakai, T.; Yang, X.; Huo, Q. Tensile creep anisotropy and its weakening mechanism in a dilute Mg-Ca alloy. J. Magnes. Alloys 2024. [Google Scholar] [CrossRef]
  67. Li, Z.H.; Cheng, D.; Wang, K.; Hoglund, E.R.; Howe, J.M.; Zhou, B.C.; Sasaki, T.T.; Ohkubo, T.; Hono, K. Revisited precipitation process in dilute Mg-Ca-Zn alloys. Acta Mater. 2023, 257, 119072. [Google Scholar] [CrossRef]
  68. Li, Z.; Liu, F.; Yuan, A.; Duan, B.; Li, Y.; Li, X. Effect of rolling deformation on microstructure and texture of spray-deposited magnesium alloy containing Mg-Nd-Zn typed LPSO. J. Mater. Sci. Technol. 2017, 33, 630–636. [Google Scholar] [CrossRef]
  69. He, Z.; Zhang, Y.; Liang, J.; Li, Y. Quantitative investigation on the individual contribution of twinning and dislocation slip to creep in dilute Mg-Y alloys with various grain sizes. J. Alloys Compd. 2024, 990, 174473. [Google Scholar] [CrossRef]
  70. Sarebanzadeh, M.; Orozco-Caballero, A.; Llorca, J. Accurate determination of active slip systems for improved geometrical criteria of basal-to-basal slip transfer at grain boundaries in pure Mg. Acta Mater. 2023, 243, 118536. [Google Scholar] [CrossRef]
  71. Wang, S.; Niu, E.; Zhang, Q.; Ouyang, S.; Shen, X.; She, J.; Tang, A.; Chen, X.; Pan, F. Advancing biodegradable Mg-Gd-Y alloy for cardiovascular stent applications. Mater. Today Commun. 2024, 41, 110893. [Google Scholar] [CrossRef]
  72. Kim, S.-H.; Jung, J.-G.; You, B.S.; Park, S.H. Microstructure and texture variation with Gd addition in extruded magnesium. J. Alloys Compd. 2017, 695, 344–350. [Google Scholar] [CrossRef]
  73. Wang, X.; Zhao, Y.; Li, M.; Tang, S.; Huang, Y.; Liu, Y.; Huang, C. Achieving enhanced mechanical properties of extruded Mg-Gd-Y-Zn-Zr alloy by regulating the initial LPSO phases. Mater. Sci. Eng. A 2024, 917, 147411. [Google Scholar] [CrossRef]
  74. Wang, Y.; Rong, W.; Wu, Y.J.; Peng, L.M.; Chen, J.; Ding, W.J. Effects of Mn addition on the microstructures and mechanical properties of the Mg-15Gd-1Zn alloy. J. Alloys Compd. 2017, 698, 1066–1076. [Google Scholar] [CrossRef]
  75. Chen, M.; Tao, X.; Wang, L.; Ma, K.; Zhang, D.; Li, Y.; Wei, S.; Zhang, Z.; Zhao, Z.; Bai, P. Phase transformation during solid-solution and aging treatment of a Mg-Gd-Y-Zn-Zr alloy containing LPSO and W phases. J. Alloys Compd. 2024, 1005, 176176. [Google Scholar] [CrossRef]
  76. Yang, W.; Quan, G.F.; Ji, B.; Wan, Y.F.; Zhou, H.; Zheng, J.; Yin, D.D. Effect of Y content and equal channel angular pressing on the microstructure, texture and mechanical property of extruded Mg-Y alloys. J. Magnes. Alloys 2022, 10, 195–208. [Google Scholar] [CrossRef]
  77. Zhao, J.; Jiang, B.; Wang, Q.; Yang, H.; Yuan, M.; Huang, G.; Zhang, D.; Pan, F. Influence of Ce content on the microstructures and tensile properties of Mg-1Gd-0.5Zn alloys. Mater. Sci. Eng. A 2021, 823, 141675. [Google Scholar] [CrossRef]
  78. Ci, W.; Deng, L.; Chen, X.; Liu, C.; Pan, F. Effect of minor Ca addition on microstructure and corrosion behavior of Mg–Y–Ca alloys. J. Mater. Res. Technol. 2023, 26, 7502–7515. [Google Scholar] [CrossRef]
  79. Liao, H.; Kim, J.; Lv, J.; Jiang, B.; Chen, X.; Pan, F. Microstructure and mechanical properties with various pre-treatment and Zn content in Mg-Gd-Y-Zn alloys. J. Alloys Compd. 2020, 831, 154873. [Google Scholar] [CrossRef]
  80. Hantzsche, K.; Bohlen, J.; Wendt, J.; Kainer, K.U.; Yi, S.B.; Letzig, D. Effect of rare earth additions on microstructure and texture development of magnesium alloy sheets. Scr. Mater. 2010, 63, 725–730. [Google Scholar] [CrossRef]
  81. Peng, J.; Zhang, Z.; Li, Y.; Zhou, W.; Wu, Y. Twinning-induced dynamic recrystallization and micro-plastic mechanism during hot-rolling process of a magnesium alloy. Mater. Sci. Eng. A 2017, 699, 99–105. [Google Scholar] [CrossRef]
  82. Zhang, L.; Li, Y. Dynamic recrystallization mechanism, texture evolution development and mechanical characteristics of a Mg–8.7Gd–4.18Y–0.42Zr magnesium alloy by ECAP. Prog. Nat. Sci. Mater. Int. 2024, 34, 376–388. [Google Scholar] [CrossRef]
  83. Zhao, T.; Dai, Y.; Zheng, T.; Hu, Y.; Jiang, B.; Pan, F. The extrusion ratios modification of Mg-2Gd-xMn alloys to exploration of the strengthening mechanisms. J. Mater. Res. Technol. 2024, 30, 2147–2157. [Google Scholar] [CrossRef]
  84. Pei, Y.; Yuan, M.; Wei, E.; Zhang, Z.; Zhao, M.; Ma, H.; Teng, B. Cooperative achievement in grain refinement and texture weakening of Mg-Gd-Y-Sm-Zr alloy fabricated by asymmetric extrusion-shear process. Mater. Sci. Eng. A 2024, 895, 146262. [Google Scholar] [CrossRef]
  85. Liu, W.; Zhang, C.; Guo, P.; Shi, Q.; Feng, H.; Zhang, S.; Han, F. Achieving enhanced mechanical properties in the Mg–Y–Zn–V alloy by refining grains and improving the morphology and distribution of LPSO via pre-forging and subsequent rolling. J. Mater. Res. Technol. 2024, 29, 2493–2504. [Google Scholar] [CrossRef]
  86. Li, Q.-k.; Yan, H.; Chen, R.-s. Effect of rolling reduction on deformation mechanism and twinning behavior of WE43 magnesium alloy. Trans. Nonferrous Met. Soc. China 2022, 32, 3901–3913. [Google Scholar] [CrossRef]
  87. Liang, S.; Ren, Q.; Li, X.; Luan, S.; Li, X.; Jin, P.; Wang, J. The dynamic recrystallization induced by Al2Nd and promoted by Al2RE in a hot compressed Mg–1Al-1.5Nd-1.5Y-0.5Gd alloy. J. Mater. Res. Technol. 2024, 33, 8232–8245. [Google Scholar] [CrossRef]
  88. Nienaber, M.; Bohlen, J.; Yi, S.; Kurz, G.; Kainer, K.U.; Letzig, D. Influence of Ca addition on the dynamic and static recrystallization behavior of direct extruded flat profiles of Mg-Y-Zn alloy. J. Magnes. Alloys 2023, 11, 3736–3748. [Google Scholar] [CrossRef]
  89. Wang, X.; Gao, M.; Miao, L.; Tan, L.; Yang, K. Study of rotational dynamic recrystallization during warm compression of an Mg-Zn-Nd-Zr alloy. Mater. Today Commun. 2024, 38, 107928. [Google Scholar] [CrossRef]
  90. Ling, Z.; Yang, K.; Wang, L.; Ma, K.; Zhang, D.; Li, Y.; Zhang, Z.; Zhao, Z.; Bai, P. Hot deformation behavior and microstructural evolutions of a Mg-Gd-Y-Zn-Zr alloy prepared by hot extrusion. J. Alloys Compd. 2024, 1009, 176916. [Google Scholar] [CrossRef]
  91. Chen, T.; Chen, Z.; Shao, J.; Wang, R.; Mao, L.; Liu, C. The role of long period stacking ordered phase in dynamic recrystallization of a Mg–Zn–Y alloy during hot compression. J. Alloys Compd. 2020, 818, 152814. [Google Scholar] [CrossRef]
  92. Hao, J.; Zhao, L.; Zhang, J.; Wang, H.; Cheng, W.; Li, B.; Bai, Y. Study on equal channel angular pressing process of Mg–Zn–Y–Mn–Ti magnesium alloy enhanced by lamellar LPSO phase and spherical W phase. J. Mater. Res. Technol. 2023, 25, 7078–7089. [Google Scholar] [CrossRef]
  93. Luo, X.; Feng, Z.; Yu, T.; Huang, T.; Li, R.; Wu, G.; Hansen, N.; Huang, X. Microstructural evolution in Mg-3Gd during accumulative roll-bonding. Mater. Sci. Eng. A 2020, 772, 138763. [Google Scholar] [CrossRef]
  94. Li, X.; Li, X.; Ye, Y.; Zhang, R.; Kure-Chu, S.-Z.; Tang, G. Deformation mechanisms and recrystallization behavior of Mg-3Al-1Zn and Mg-1Gd alloys deformed by electroplastic-asymmetric rolling. Mater. Sci. Eng. A 2019, 742, 722–733. [Google Scholar] [CrossRef]
  95. Kwak, T.Y.; Lim, H.K.; Han, S.H.; Kim, W.J. Refinement of the icosahedral quasicrystalline phase and the grain size of Mg–9.25Zn–1.66Y alloy by high-ratio differential speed rolling. Scr. Mater. 2015, 103, 49–52. [Google Scholar] [CrossRef]
  96. Minárik, P.; Zemková, M.; Veselý, J.; Bohlen, J.; Knapek, M.; Král, R. The effect of Zr on dynamic recrystallization during ECAP processing of Mg-Y-RE alloys. Mater. Charact. 2021, 174, 111033. [Google Scholar] [CrossRef]
  97. Wang, D.; Xu, D.; Wang, B.; Yan, C.; Wang, S.; Xu, X.; Zhang, L.; Lu, C. Effect of cross rolling on the microstructure and mechanical performance of a dual-phase structured Mg-8Li-6Zn-1Y (in wt.%) alloy. J. Mater. Sci. Technol. 2024, 176, 132–144. [Google Scholar] [CrossRef]
  98. Chen, H.; Han, Y.; Liu, C.; Chen, Z. Effect of unidirectional rolling and cross-rolling on microstructure and mechanical anisotropy of Mg-Gd-Y-Ag-Zr plates. J. Alloys Compd. 2023, 960, 170680. [Google Scholar] [CrossRef]
  99. An, D.; Qian, B.; Wu, R.; Wang, X.; Hou, L.; Ma, X.; Zhang, J. Influence of power dissipation value and deformation activation energy on recrystallization in compression deformation behavior of Mg-Li-Zn-Y alloy. J. Rare Earths 2024, 42, 2341–2349. [Google Scholar] [CrossRef]
  100. Guan, D.; Rainforth, W.M.; Gao, J.; Ma, L.; Wynne, B. Individual effect of recrystallisation nucleation sites on texture weakening in a magnesium alloy: Part 2- shear bands. Acta Mater. 2018, 145, 399–412. [Google Scholar] [CrossRef]
  101. Sagapuram, D.; Viswanathan, K. Evidence for Bingham plastic boundary layers in shear banding of metals. Extrem. Mech. Lett. 2018, 25, 27–36. [Google Scholar] [CrossRef]
  102. Shi, B.Q.; Wang, Y.Z.; Shang, X.L.; Zhao, L.Y.; Li, C.Q.; Chen, D.C.; Nie, B.H.; Chen, R.S.; Ke, W. Microstructure evolution of twinning-induced shear bands and correlation with ‘RD-split’ texture during hot rolling in a Mg-1.1Zn-0.76Y-0.56Zr alloy. Mater. Charact. 2022, 187, 111853. [Google Scholar] [CrossRef]
  103. Ding, N.; Du, W.; Li, S.; Liu, K.; Du, X.; Yu, Z. Strain rate dependence of dynamic recrystallization and texture evolution in hot compressed Mg-Gd-Er-Zr alloy. J. Magnes. Alloys 2023, 13, 161–171. [Google Scholar] [CrossRef]
  104. Han, L.; Yu, Y.; Wei, D.; Wang, X.; Zhao, G.; Wang, G. The synergistic and interactive effects of slip systems and dynamic recrystallization on the weakening basal texture of Mg-Y-Nd-Zr-Gd magnesium alloy. Mater. Des. 2024, 237, 112583. [Google Scholar] [CrossRef]
  105. Li, Q.; Yan, H.; Liu, H.; Chen, R. Dynamic recrystallization mechanism and near-isotropic mechanical properties of WE43 magnesium alloy sheets rolled at different temperatures. Mater. Charact. 2022, 193, 112259. [Google Scholar] [CrossRef]
  106. Sabat, R.K.; Samal, P.K.; Ahamed, S.M. Effect of strain path on the evolution of microstructure, texture and tensile properties of WE43 alloy. Mater. Sci. Eng. A 2018, 715, 348–358. [Google Scholar] [CrossRef]
  107. Pan, F.; Zeng, B.; Jiang, B.; Atrens, A.; Dong, H. Deformation mechanism and microstructure evolution during on-line heating rolling of AZ31B Mg thin sheets. Mater. Charact. 2017, 124, 266–275. [Google Scholar] [CrossRef]
  108. Li, J.; Li, X.; Xu, T.; Yan, H.; Mao, Y.; Chen, R. Regulating the grain refinement and rolling properties of coarse-crystalline Mg-Zn-Gd-Ca-Mn alloy through multi-pass cold rolling and annealing. Mater. Sci. Eng. A 2024, 911, 146940. [Google Scholar] [CrossRef]
  109. Yan, H.; Xu, S.W.; Chen, R.S.; Kamado, S.; Honma, T.; Han, E.H. Twins, shear bands and recrystallization of a Mg–2.0%Zn–0.8%Gd alloy during rolling. Scr. Mater. 2011, 64, 141–144. [Google Scholar] [CrossRef]
  110. Wang, J.; Zhu, G.; Wang, L.; Zhu, Q.; Vasilev, E.; Zeng, X.; Knezevic, M. Dislocation-induced plastic instability in a rare earth containing magnesium alloy. Materialia 2021, 15, 101038. [Google Scholar] [CrossRef]
  111. Wang, X.; Li, M.; Huang, Y.; Liu, Y.; Huang, C. Deformation behavior of LPSO phases with regulated morphology and distribution and their role on dynamic recrystallization in hot-rolled Mg–Gd–Y–Zn–Zr alloy. J. Mater. Res. Technol. 2023, 26, 6121–6134. [Google Scholar] [CrossRef]
  112. Chen, W.; Ma, J.; Cui, C.; Zhang, W.; Wang, W.; Liu, X.; Yang, J.; Cui, G. Texture role in the mechanical property improvement contributed by grain refinement for Mg-2.6Nd-0.55Zn-0.5Zr alloy subjected to extrusion process. Mater. Sci. Eng. A 2022, 831, 142185. [Google Scholar] [CrossRef]
  113. Dong, B.; Chai, J.; Wu, G.; Zhang, H.; Zhang, Z.; Yu, J.; Xue, Y.; Meng, M. Dynamic recrystallization behavior and strengthening mechanism of Mg–Gd–Y–Zn–Zr alloy during isothermal MDF process. J. Mater. Res. Technol. 2022, 21, 1419–1433. [Google Scholar] [CrossRef]
  114. Gao, W.; Dong, X.; Fan, Y.; Zhao, H.; Guo, C.; Weng, K.; Xiao, Y. A comparative study on the microstructure, texture, and mechanical properties of extruded Mg–9Li and Mg–9Li-0.3Ca alloys. Mater. Sci. Eng. A 2022, 854, 143844. [Google Scholar] [CrossRef]
  115. Zhang, Z.; Kim, J.; Li, M.; Gao, Y.; Hu, Y.; Jiang, B.; Pan, F. Effects of Nd content on the microstructures and mechanical properties of ZK60 Mg alloy and corresponding strengthening mechanisms. Mater. Sci. Eng. A 2024, 901, 146504. [Google Scholar] [CrossRef]
Figure 1. (a) {0002} pole figure of extruded Mg-3Y alloy [20]; (b) {0001} pole figures of rolled Mg-Ce and Mg-Gd alloy [24].
Figure 1. (a) {0002} pole figure of extruded Mg-3Y alloy [20]; (b) {0001} pole figures of rolled Mg-Ce and Mg-Gd alloy [24].
Solids 06 00011 g001
Figure 2. (a) {0002} pole figures, inverse pole figures, and stress–strain curves of the extruded Mg-Gd alloys [42]; (b) optical microstructure (OM), inverse pole figure, TEM, and LPSO phase SAED spots of Mg-13Gd-4Y-2Zn-0.5Zr alloy [47]; (c) schematic of bimodal texture formation under compression [48].
Figure 2. (a) {0002} pole figures, inverse pole figures, and stress–strain curves of the extruded Mg-Gd alloys [42]; (b) optical microstructure (OM), inverse pole figure, TEM, and LPSO phase SAED spots of Mg-13Gd-4Y-2Zn-0.5Zr alloy [47]; (c) schematic of bimodal texture formation under compression [48].
Solids 06 00011 g002
Figure 3. (a) IPF maps and pole figures of WE43 in different states [51]; (b) basal pole figures of 1–3 MAF passes of the WE43 alloys [52]; (c) microstructures of casting WE43 alloy and heat-treated WE43 alloys [53]; (d) IPF maps and pole figures of the as-cast and as-extruded WE71 [54].
Figure 3. (a) IPF maps and pole figures of WE43 in different states [51]; (b) basal pole figures of 1–3 MAF passes of the WE43 alloys [52]; (c) microstructures of casting WE43 alloy and heat-treated WE43 alloys [53]; (d) IPF maps and pole figures of the as-cast and as-extruded WE71 [54].
Solids 06 00011 g003
Figure 4. (a) {0001} pole figures of the different samples [59]; (b) {0002} pole figures and IPF-ND maps of rolled Mg-Li alloys [60]; (c) EBSD images, pole figures, and stress–strain curves of representative microstructures in different position [61].
Figure 4. (a) {0001} pole figures of the different samples [59]; (b) {0002} pole figures and IPF-ND maps of rolled Mg-Li alloys [60]; (c) EBSD images, pole figures, and stress–strain curves of representative microstructures in different position [61].
Solids 06 00011 g004
Figure 5. Optical micrographs and SEM micrographs of cast structure of ZEK100 magnesium alloy with rare-earth elements and the evolution of texture of rare-earth elements after hot rolling [64].
Figure 5. Optical micrographs and SEM micrographs of cast structure of ZEK100 magnesium alloy with rare-earth elements and the evolution of texture of rare-earth elements after hot rolling [64].
Solids 06 00011 g005
Figure 6. (a) Schematic of solid solution process of GWZK114, precipitation sequence diagram at different cooling rates, and pole figures of different grains [27]; (b) IPF maps, (0001) pole figures. and tensile property curve of the Mg-Gd alloys [65].
Figure 6. (a) Schematic of solid solution process of GWZK114, precipitation sequence diagram at different cooling rates, and pole figures of different grains [27]; (b) IPF maps, (0001) pole figures. and tensile property curve of the Mg-Gd alloys [65].
Solids 06 00011 g006
Figure 7. (a) IPF maps, pole figures, and mechanical properties with the direction paralleling to ED of GW alloy [71]; (b) inverse pole figures of Mg-Gd and extrusion direction deflection and grain area fraction curves, pole figure of the recrystallized region, and tensile curves of as-extruded Mg-Gd alloys [72].
Figure 7. (a) IPF maps, pole figures, and mechanical properties with the direction paralleling to ED of GW alloy [71]; (b) inverse pole figures of Mg-Gd and extrusion direction deflection and grain area fraction curves, pole figure of the recrystallized region, and tensile curves of as-extruded Mg-Gd alloys [72].
Solids 06 00011 g007
Figure 8. (a) {0001} pole figures, inverse pole figures, and mechanical properties of the Mg-Y alloys for different ECAP processes [76]; (b) IPF maps and stress–strain curves of the extruded sheets [77].
Figure 8. (a) {0001} pole figures, inverse pole figures, and mechanical properties of the Mg-Y alloys for different ECAP processes [76]; (b) IPF maps and stress–strain curves of the extruded sheets [77].
Solids 06 00011 g008
Figure 9. (a) {0001} pole figures and microstructure of Mg-1Y-xCa (x = 0~0.2 wt.%) alloys [78]; (b) SEM images, pole figures, and the stress–strain curve of Mg-Gd-Y-Zn alloys [79].
Figure 9. (a) {0001} pole figures and microstructure of Mg-1Y-xCa (x = 0~0.2 wt.%) alloys [78]; (b) SEM images, pole figures, and the stress–strain curve of Mg-Gd-Y-Zn alloys [79].
Solids 06 00011 g009
Figure 10. (a) Pole figure, inverse pole figure, and mechanical properties of extruded alloys [83]; (b) IPF maps and {0001} pole figures of the Mg-Gd-Y-Sm-Zr alloy with different extrusion ratio [84].
Figure 10. (a) Pole figure, inverse pole figure, and mechanical properties of extruded alloys [83]; (b) IPF maps and {0001} pole figures of the Mg-Gd-Y-Sm-Zr alloy with different extrusion ratio [84].
Solids 06 00011 g010
Figure 11. (a) Statistical chart of proportions of different precipitated phases.; (b) pole figure and ODF maps of D80 and R60 alloys; (c) schematic diagram of the Mg-Y-Zn-V alloy processing process [85].
Figure 11. (a) Statistical chart of proportions of different precipitated phases.; (b) pole figure and ODF maps of D80 and R60 alloys; (c) schematic diagram of the Mg-Y-Zn-V alloy processing process [85].
Solids 06 00011 g011
Figure 12. (a) Pole figures of the Mg-9.5Gd-4Y-2Zn-0.5Zr alloy at different temperatures [90]; (b) IPF maps and pole figures of Mg-1.99Zn-8.02Y at different strain rates [91]; (c) pole figures of the Mg-9.5Gd-4Y-2Zn-0.5Zr alloy at different strain rates [90]; (d) IPF maps and pole figures of Mg-1.99Zn-8.02Y at different temperatures [91].
Figure 12. (a) Pole figures of the Mg-9.5Gd-4Y-2Zn-0.5Zr alloy at different temperatures [90]; (b) IPF maps and pole figures of Mg-1.99Zn-8.02Y at different strain rates [91]; (c) pole figures of the Mg-9.5Gd-4Y-2Zn-0.5Zr alloy at different strain rates [90]; (d) IPF maps and pole figures of Mg-1.99Zn-8.02Y at different temperatures [91].
Solids 06 00011 g012
Figure 13. (a) Schematic of the BE and RSE; (b) FEM consequences of effective strain for the BE and RSE methods; (c) {0001} pole figures and inverse pole figures of the BE and RSE [39].
Figure 13. (a) Schematic of the BE and RSE; (b) FEM consequences of effective strain for the BE and RSE methods; (c) {0001} pole figures and inverse pole figures of the BE and RSE [39].
Solids 06 00011 g013
Figure 14. (a) UR processing diagram; (b) CR processing diagram [97].
Figure 14. (a) UR processing diagram; (b) CR processing diagram [97].
Solids 06 00011 g014
Figure 15. Schematic diagram of dynamic recrystallization [104].
Figure 15. Schematic diagram of dynamic recrystallization [104].
Solids 06 00011 g015
Figure 16. Microstructure results of Mg-1.9Zn-2.0Gd-0.2Ca-0.2Mn alloy: (ac) OM image; (d) IPF map; (e) KAM of area in (d); (f) IGMA of area in (d); (g) pole figure; (h) type of twin; (i) orientation difference distribution image; (j) schematic diagram of deformation and recrystallization [108].
Figure 16. Microstructure results of Mg-1.9Zn-2.0Gd-0.2Ca-0.2Mn alloy: (ac) OM image; (d) IPF map; (e) KAM of area in (d); (f) IGMA of area in (d); (g) pole figure; (h) type of twin; (i) orientation difference distribution image; (j) schematic diagram of deformation and recrystallization [108].
Solids 06 00011 g016
Figure 17. (a) OM image of Mg2.0Zn0.8Gd sheet; (b) EBSD image of part of the area in OM image; (c) pole figures of a different grain in (b) [109].
Figure 17. (a) OM image of Mg2.0Zn0.8Gd sheet; (b) EBSD image of part of the area in OM image; (c) pole figures of a different grain in (b) [109].
Solids 06 00011 g017
Figure 18. (a) SEM image, GOS image, engineering strain–stress curves of ZKNd604, and volume fractions of different grains [115]; (b) inverse pole figure maps of the PSN mechanism (the red arrow shows the Mg24Y5 eutectic phase) [52]; (c) schematic diagram of different deformation processes [26].
Figure 18. (a) SEM image, GOS image, engineering strain–stress curves of ZKNd604, and volume fractions of different grains [115]; (b) inverse pole figure maps of the PSN mechanism (the red arrow shows the Mg24Y5 eutectic phase) [52]; (c) schematic diagram of different deformation processes [26].
Solids 06 00011 g018
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

Liu, W.; Wei, B.; Li, R.; Wang, X.; Wu, H.; Fang, W. Research Progress on Texture Regulation of Rare-Earth Magnesium Alloys. Solids 2025, 6, 11. https://doi.org/10.3390/solids6010011

AMA Style

Liu W, Wei B, Li R, Wang X, Wu H, Fang W. Research Progress on Texture Regulation of Rare-Earth Magnesium Alloys. Solids. 2025; 6(1):11. https://doi.org/10.3390/solids6010011

Chicago/Turabian Style

Liu, Weiyan, Boxin Wei, Rengeng Li, Xin Wang, Hao Wu, and Wenbin Fang. 2025. "Research Progress on Texture Regulation of Rare-Earth Magnesium Alloys" Solids 6, no. 1: 11. https://doi.org/10.3390/solids6010011

APA Style

Liu, W., Wei, B., Li, R., Wang, X., Wu, H., & Fang, W. (2025). Research Progress on Texture Regulation of Rare-Earth Magnesium Alloys. Solids, 6(1), 11. https://doi.org/10.3390/solids6010011

Article Metrics

Back to TopTop