<<

metals

Article The Effect of on the Microstructure and Properties of Ultralow-Temperature Rolled Mg–2Y–0.6Nd–0.6Zr

Hongmin Yu 1,2,3, Wei Li 1,2,3,* , Yun Tan 1,2,3 and Yuanbiao Tan 1,2,3

1 College of Material and Metallurgy, Guizhou University, Guiyang 550025, China; [email protected] (H.Y.); [email protected] (Y.T.); [email protected] (Y.T.) 2 Guizhou Key Laboratory for Mechanical Behavior and Microstructure of Materials, Guizhou University, Guiyang 550025, China 3 National and Local Joint Engineering Laboratory for High-Performance Metal Structure Material and Advanced Manufacturing Technology, Guiyang 550025, China * Correspondence: [email protected]; Tel.: +86-0851-8362-7068

Abstract: Mg–2Y–0.6Nd–0.6Zr alloy was first deformed by equal channel angular pressing (ECAP), then rolled and deformed under ultralow temperature conditions (liquid nitrogen immersion), and finally annealed. Optical microscopy (OM), electron backscatter diffraction technology (EBSD), and transmission electron microscopy (TEM) were used to analyze the evolution of the multiscale microstructure and changes in the mechanical properties of the alloy under ultralow temperatures and various annealing conditions. The results showed that the alloy treated with ECAP obtained fine grains, and a large number of fine twins were formed during the ultralow-temperature rolling process, which promoted the improvement of its hardness and strength and provided numerous preferential nucleation sites. The annealing made it easier to induce recrystallization and improve the   recrystallization nucleation rate. The twin boundary produced by the alloy after ultralow-temperature rolling and the uniform fine grains formed by annealing resulted in excellent strength and plasticity Citation: Yu, H.; Li, W.; Tan, Y.; Tan, − − Y. The Effect of Annealing on the of the alloy. The twins formed after rolling under ultralow temperatures were mainly {1012} <1011> Microstructure and Properties of tensile twins. The alloy had comprehensive mechanical properties with a tensile strength of 186.15 ◦ Ultralow-Temperature Rolled MPa and an elongation of 29% after annealing at 350 C for 10 min. Mg–2Y–0.6Nd–0.6Zr Alloy. Metals 2021, 11, 315. https://doi.org/ Keywords: alloy; ultralow-temperature rolled; annealing; recrystallization; twin 10.3390/met11020315

Academic Editor: Hamid Jahed 1. Introduction Received: 26 January 2021 Magnesium alloys offer the advantages of low , high specific strength, high Accepted: 8 February 2021 Published: 11 February 2021 specific rigidity, excellent electromagnetic shielding effectiveness, and recyclability. These materials have broad prospects for industrial applications, such as in aerospace and elec-

Publisher’s Note: MDPI stays neutral tronics, and are known as the “green structural engineering materials of the 21st cen- with regard to jurisdictional claims in tury” [1]. Magnesium and magnesium alloys have a hexagonal close-packed (HCP) crystal published maps and institutional affil- structure; thus, under low-temperature deformation, it is difficult to activate slip, facilitat- iations. ing the formation of a strong texture and resulting in poor room-temperature formability, which limits the wide application of magnesium alloys in engineering materials [2–5]. With the development of modern industrial technology, more and more structural parts require materials that have both high strength and good plasticity. In recent decades, many severe plastic deformation (SPD) technologies for obtaining ultrafine grain (UFG) Copyright: © 2021 by the authors. Licensee MDPI, Basel, Switzerland. materials have been developed, including high-pressure torsion (HPT) [6,7], multiaxial This article is an open access article (MAF) [8,9], accumulative roll-bonding (ARB) [10,11], and equal channel angular distributed under the terms and pressing (ECAP) [12–14]. These deformation methods can obtain ultrafine grains, but the conditions of the Creative Commons plasticity and thermal stability of the structure obtained by these deformation methods Attribution (CC BY) license (https:// are poor. Jahadi et al. [15] studied the effect of ECAP process on the microstructure and creativecommons.org/licenses/by/ mechanical properties of forged AM30 magnesium alloy. They found that the alloy grain ◦ 4.0/). size was refined after four passes ECAP at 275 C, and the mechanical properties were

Metals 2021, 11, 315. https://doi.org/10.3390/met11020315 https://www.mdpi.com/journal/metals Metals 2021, 11, 315 2 of 15

improved. Garces et al. [16] also found that ECAP can achieve the effect of fine grain strengthening. Researching how to obtain high-strength materials with good plasticity and thermal stability has important theoretical significance and application value. Sarker et al. [17] studied the change of twin width caused by detwinning and the subsequent change of texture composition strength by changing the loading direction of predeformed AM30 magnesium alloy and found that the twin width decreased linearly with the increase of true strain. Sarker et al. [18] studied the change of twin width in extruded magnesium alloy during compression and found that the twin had three growth stages with the increase of strain due to the interaction between twin and dislocation. Sarker et al. [19] studied the plastic deformation of AM30 magnesium alloy in direction, which could be divided into three different stages. With the decrease of strain rate, twinning occurred in stage A. In stage B, there was a strong twin dislocation interaction, which showed that the strain hardening rate increased with the extension of strain range. In stage C, the strain hardening rate decreased due to the decrease of twin resistance to dislocation slip. Sarker et al. [20] studied the effect of twinning induced by prestrain on the subsequent texture and hardening behavior of extruded AM30 alloy. It was observed that in two-step extrusion direction (ED) and transverse direction (TD) compression, with the increase of prestrain, the YS strength and bending strength increased and the hardening ability decreased. In the second compression process, new twins were formed along the TD direction, which led to grain refinement, thus increasing the strength (YS) effect based on the Hall–Petch relationship. Sarker et al. [21] explored the relationship between hardening behavior, texture evolution, and detwinning of extruded AM30 alloy when the strain path was changed. It was found that detwinning occurred with the increase of prestrain, and the texture was weakened by precompression in the ED and recompression in the normal direction (ND). Sarker et al. [22] studied the texture composition of AM30 magnesium alloy and its evolution with increasing compression strain and found that the c-axis always rotated in the direction of anticompression due to the tensile twins. Although twins can be obtained by compression at room temperature, a large number of twin boundaries will be obtained by reducing the rolling temperature, and better strength and hardness will be obtained. A large number of twin boundaries can be obtained through ultralow-temperature deformation, which can be equivalent to grain boundaries, strengthening metal materials by effectively hindering the movement of dislocations [23]. After ultralow-temperature rolling, the strength of the alloy is greatly improved, but its plasticity is greatly reduced. To make an alloy with high strength and good plasticity is therefore a challenge. The room temperature formability of the alloy can be further improved by refining the grain structure and improving the texture [24–26]. Two methods are generally used for magnesium alloy matrixes; one method is to control the activity of deformation twins and to promote the activation of nonbasal surface slip, and the second method is to refine the grain structure and improve the texture by cold deformation and subsequent annealing to form static recrystallization. Ultralow-temperature rolling is more conducive to twinning deformation of metals by reducing the deformation temperature because the dislocation movement requires greater stress at lower temperatures, making it difficult for the dislocation to slip and climb. Therefore, it can accumulate a large amount of high-energy positions to provide more preferential nucleation positions for subsequent annealing, which is more conducive to the formation of finer and uniform recrystallized grains during the annealing process, meaning the alloy has excellent plasticity. Sarker et al. [27] studied the recrystallization behavior of precompressed AM30 magnesium alloy and the microstructure and texture changes of annealed samples during compression. It was found that the twins in precompressed AM30 magnesium alloy gradually disap- peared, the volume fraction of twins decreased, and the texture weakened with the increase of annealing time, while the compress yield strength (CYS) decreased and the fracture strain increased with the increase of annealing temperature. Kim et al. [28] studied the changes in twinning and annealing behavior caused by the difference in grain size within the material Metals 2021, 11, 315 3 of 15

and found that the microstructure difference between the coarse-grained region and the fine-grained region caused by the precompression in the subsequent annealing process almost disappeared and got a uniform microstructure. Song et al. [29] used a process route combining equal channel angular extrusion and annealing to simultaneously improve the resistance and mechanical properties of Mg–2Zn-Mn–Ca–Ce alloy and achieved double yield/ultimate strength and elongation, while the mechanical integrity of the alloy was better maintained after corrosion. Chen et al. [30] used cumulative cold rolling and intermediate annealing processes to prepare ultrafine-grained ZK61 magnesium alloy sheets. Multiple static recrystallizations clearly achieved refinement and homogenization of the structure during this process. Guo et al. [31] studied the effects of recrystallization and grain growth during annealing in optimizing the texture of calcium-added magne- sium alloy sheets and found that the recrystallized nucleus retained the orientation of its deformed matrix in AZ31, which provided limited potential for optimizing texture through annealing treatment. Therefore, it can be seen that the alloy can obtain good plasticity during the annealing process. The purpose of this work was to study the static recrystallization behavior of the subsequent annealed alloy as well as the change rule of twinning in the process of static recrystallization and its influence on the microstructure and mechanical properties by im- proving the traditional deformation process and introducing a large number of dislocations and twins after grain refinement.

2. Material and Methods The experimental material was as-cast Mg–2Y–0.6Nd–0.6Zr alloy. The central portion of an ingot of the material was cut, and it was processed into Φ12 mm × 80 mm, followed by homogenization at 450 ◦C for 6 h. The ECAP die structure used in the experiment is shown in Figure1, where ϕ = 120◦, Ψ = 30◦, the extrusion path is the Bc (that is, rotated 90◦ around the longitudinal axis of the sample in a counter-clockwise direction after each extrusion pass), ND and ED (RD) represent the normal and extrusion direction (rolling direction) in the sample, and TD is the transverse direction. A total of four ECAP extrusion passes were made. The deformation temperature was 400 ◦C for the first two passes and 350 ◦C for the last two passes; the sample was kept warm for 10 min before each pass at an extrusion speed of 0.4 mm/s. A three-phase asynchronous rolling mill was used to carry out rolling tests on the extruded alloy at a speed of 970 r·min−1 under ultralow temperature conditions for two rolling passes. The bar was soaked in liquid nitrogen for approximately 15 min before each rolling pass, and the sample was rotated 90◦ along the axis before the second rolling. After the rolling was completed, the sample was soaked in liquid nitrogen for approximately 10 min before removal from the mill. Finally, the samples were annealed under different conditions at annealing temperatures of 250, 300, and 350 ◦C with a 1 h holding time; the optimal high temperature was selected as 350 ◦C. A short annealing time was used to observe and analyze the effect of changes in the twins and the dislocation content on the properties of Metals 2021, 11, x FOR PEER REVIEW 4 of 16 magnesium alloys, that is, annealing times of 3, 6, 10, and 30 min were investigated, with air cooling after annealing.

FigureFigure 1.1. EqualEqual channelchannel angularangular pressingpressing (ECAP)(ECAP)mold mold structure.structure.

All the observation surfaces used for testing are ED(RD) − ND. For simplicity, the ultralow-temperature rolled Mg–2Y–0.6Nd–0.6Zr alloy (ECAP + rolling) is referred to as the ultralow-temperature rolled alloy. The sample was ground with 400 #~10000 # sand- paper, then polished to a “mirror surface” effect, and lastly corroded with picric acid etch- ing solution. The microstructure of the ultralow-temperature rolled samples was observed before and after annealing with an OLMPUS-BH2 optical microscope (Olympus, Tokyo, Japan). Image-Pro Plus image analysis software (Image-Pro Plus 6.0, Media Cybernetics, Inc., Rockville, MD, USA) was used to measure the grain size by the average linear inter- cept method. The samples were tested with a HV-1000 digital microhardness tester (Su- zhou Haichuan Science and Education Equipment Co. Ltd, Suzhou, China) under a 100 g load with a loading time of 15 s and a holding time of 10 s. Because the uneven defor- mation at both ends of the sample needed to be removed, it could not reach the standard size length. Therefore, the nonstandard size with extensometer was adopted. Figure 2 shows that the tensile sample adopted a sheet-like nonstandard size because of sample size limitations. An Instron 5565 electronic universal testing machine (Instrang Shanghai Test Equipment Trading Co., Ltd. Instron, Shanghai, China) was used to measure the ten- sile properties of the sample. At least three samples for each state were tested at a tensile rate of 0.51 mm/min. The microtexture and grain orientation of the overall alloy structure were tested by electron backscatter diffraction technology (EBSD) with a S-3400 N scan- ning electron microscope (German Zeiss, Jena, German). Thin-slice samples with 5 mm length, 5 mm width, and 1 mm thickness were cut from the middle of the sample. Firstly, the samples were mechanically polished with 400 #~3000 #–SiC water-abrasive paper. Then, the observation surface of the sample was mechanically polished with alumina slurry to obtain the ideal state of bright surface and no scratch. Leica EMRES 102 argon ion polishing instrument (Leica Microsystems Shanghai Trading Co., Ltd, Shanghai, China) was used to polish the EBSD samples, and the specimens were polished according to the following sequence of polishing parameters: (1) voltage of 5 KV, ion gun angle of 15°, and polishing time of 40 min; (2) voltage of 4.5 KV, ion gun angle of 15°, and polishing time of 50 min; (3) voltage of 3.5 KV, ion gun angle unchanged, and low-voltage polishing time of 60 min. The sample was inclined at 70°, the acceleration voltage was 20 KV, the beam density was 18 A/m2, the step size EBSD scan was 0.6 µm, and the working distance was 15–16 mm. Channel 5 software (Oxford Instruments, Abingdon, UK) was used for statistics and analysis of EBSD data. To prepare transmission electron microscopy (TEM) sample, the sample was first cut into a disc with a thickness of 1 mm and then mechani- cally ground to a thickness of less than 50 mm. Then, a sample with a diameter of 3 mm was punched out from the cutting disc. The thin area was obtained by ion milling, and Gatan 691 argon ion polishing instrument (Gatan Inc., Pleasanton, CA, USA) was used to polish the TEM samples. The specimens were polished according to the following se- quence of polishing parameters: (1) thinning: voltage of 5.3 KV, ion gun angle of 8°, and polishing time of 50 min; (2) expansion: voltage of 4.5 KV, ion gun angle of 5°, and polish- ing time of 30 min; (3) thinning: voltage of 3.6 KV, ion gun angle of 3°, and low-voltage polishing time of 15 min. The dislocations and twins were observed and analyzed by a FEI Talos F200X TEM (FEI Company, Hillsboro, OR, USA).

Metals 2021, 11, 315 4 of 15

All the observation surfaces used for testing are ED(RD) − ND. For simplicity, the ultralow-temperature rolled Mg–2Y–0.6Nd–0.6Zr alloy (ECAP + rolling) is referred to as the ultralow-temperature rolled alloy. The sample was ground with 400 #~10000 # sandpaper, then polished to a “mirror surface” effect, and lastly corroded with picric acid etching solution. The microstructure of the ultralow-temperature rolled samples was observed before and after annealing with an OLMPUS-BH2 optical microscope (Olympus, Tokyo, Japan). Image-Pro Plus image analysis software (Image-Pro Plus 6.0, Media Cy- bernetics, Inc., Rockville, MD, USA) was used to measure the grain size by the average linear intercept method. The samples were tested with a HV-1000 digital microhardness tester (Suzhou Haichuan Science and Education Equipment Co. Ltd, Suzhou, China) under a 100 g load with a loading time of 15 s and a holding time of 10 s. Because the uneven deformation at both ends of the sample needed to be removed, it could not reach the standard size length. Therefore, the nonstandard size with extensometer was adopted. Figure2 shows that the tensile sample adopted a sheet-like nonstandard size because of sample size limitations. An Instron 5565 electronic universal testing machine (Instrang Shanghai Test Equipment Trading Co., Ltd. Instron, Shanghai, China) was used to measure the tensile properties of the sample. At least three samples for each state were tested at a tensile rate of 0.51 mm/min. The microtexture and grain orientation of the overall alloy structure were tested by electron backscatter diffraction technology (EBSD) with a S-3400 N scanning electron microscope (German Zeiss, Jena, German). Thin-slice samples with 5 mm length, 5 mm width, and 1 mm thickness were cut from the middle of the sample. Firstly, the samples were mechanically polished with 400 #~3000 #–SiC water-abrasive paper. Then, the observation surface of the sample was mechanically polished with alumina slurry to obtain the ideal state of bright surface and no scratch. Leica EMRES 102 argon ion polishing instrument (Leica Microsystems Shanghai Trading Co., Ltd., Shanghai, China) was used to polish the EBSD samples, and the specimens were polished according to the following sequence of polishing parameters: (1) voltage of 5 KV, ion gun angle of 15◦, and polishing time of 40 min; (2) voltage of 4.5 KV, ion gun angle of 15◦, and polishing time of 50 min; (3) voltage of 3.5 KV, ion gun angle unchanged, and low-voltage polishing time of 60 min. The sample was inclined at 70◦, the acceleration voltage was 20 KV, the beam density was 18 A/m2, the step size EBSD scan was 0.6 µm, and the working distance was 15–16 mm. Channel 5 software (Oxford Instruments, Abingdon, UK) was used for statistics and analysis of EBSD data. To prepare transmission electron microscopy (TEM) sample, the sample was first cut into a disc with a thickness of 1 mm and then mechanically ground to a thickness of less than 50 mm. Then, a sample with a diameter of 3 mm was punched out from the cutting disc. The thin area was obtained by ion milling, and Gatan 691 argon ion polishing instrument (Gatan Inc., Pleasanton, CA, USA) was used to polish the TEM samples. The specimens were polished according to the following sequence of polishing parameters: (1) thinning: voltage of 5.3 KV, ion gun angle of 8◦, and polishing time of 50 min; (2) expansion: voltage of 4.5 KV, ion gun angle of 5◦, and polishing time of 30 min; (3) thinning: voltage of 3.6 KV, ion gun angle of 3◦, and low-voltage polishing time of Metals 2021, 11, x FOR PEER REVIEW 5 of 16 15 min. The dislocations and twins were observed and analyzed by a FEI Talos F200X TEM (FEI Company, Hillsboro, OR, USA).

FigureFigure 2. 2. DimensionDimension diagram diagram of of tensile tensile specimen specimen (units: (units: mm). mm).

3. Results and Discussion 3.1. Microstructure Analysis Figure 3 shows the microstructure of an ultralow-temperature rolled alloy annealed at different temperatures for 1 h. As shown in Figure 3a, the large degree of deformation of the alloy produced a large number of twins and elongated deformed grains in the struc- ture. Figure 3b demonstrates that after annealing the alloy at 250 ℃ for 1 h, there were more twins and deformed grains within the structure as well as more fine recrystallized grains, which were mainly distributed in the crystal grains. The distortion energy was higher in some regions of the grains, such as the boundary and the twin junction. After annealing the alloy at 300 °C for 1 h, gradual growth of the recrystallized grains resulted in few twins, as shown in Figure 3c. Increasing the temperature to 350 °C caused the twin to disappear, and more uniform equiaxed grains were formed.

Figure 3. Microstructure of ultralow-temperature rolled alloy (a) before annealing and (b–d) after annealing at various temperatures: (b) 250 °C, (c) 300 °C, and (d) 350 °C.

Recrystallized grains first nucleate and grow in the high-energy regions of a grain and then expand to spread over the entire grain. High-energy regions, such as high-den- sity dislocation areas, junction of twin boundaries, and grain boundaries, provide prefer- ential recrystallization sites during heat treatment. The deformation twins formed by ul- tralow-temperature rolling can increase the grain boundary energy inside grains and ef- fectively prevent the excessive growth of recrystallized grains. Figure 4 indicates that the microstructure was gradually transformed from deformed grains formed by ultralow-

Metals 2021, 11, x FOR PEER REVIEW 5 of 16

Metals 2021, 11, 315 5 of 15 Figure 2. Dimension diagram of tensile specimen (units: mm).

3. Results and Discussion 3. Results and Discussion 3.1.3.1. Microstructure Microstructure A Analysisnalysis FigureFigure 3 shows thethe microstructuremicrostructure of of an an ultralow-temperature ultralow-temperature rolled rolled alloy alloy annealed annealed at atdifferent different temperatures temperatures for for 1 h.1 h. As As shown shown in in Figure Figure3a, 3a the, the large large degree degree of deformationof deformation of ofthe the alloy alloy produced produce ad largea large number number of of twins twins and and elongated elongated deformed deformed grains grains in thein the structure. struc- ture.Figure Figure3b demonstrates 3b demonstrates that afterthat after annealing annealing the alloy the al atloy 250 at °C250for ℃ 1 for h, there 1 h, there were were more moretwins twins and deformedand deformed grains grains within within the structure the structure as well as as well more as finemore recrystallized fine recrystallized grains, grains,which werewhich mainly were mainly distributed distributed in the crystal in the grains.crystal Thegrains. distortion The distortion energy was energy higher was in highersome regionsin some of regions the grains, of the such grains, as the such boundary as the boundary and the twin and junction. the twin Afterjunction. annealing After annealingthe alloy atthe 300 alloy◦C forat 300 1 h, °C gradual for 1 h, growth gradual of thegrowth recrystallized of the recrystallized grains resulted grains in few result twins,ed inas few shown twins, in Figureas shown3c. Increasingin Figure 3c. the Increasing temperature the totemperature 350 ◦C caused to 350 the °C twin cause tod disappear, the twin toand disappear, more uniform and more equiaxed uniform grains equ wereiaxedformed. grains were formed.

FigureFigure 3. 3. MicrostructureMicrostructure of of ultralow ultralow-temperature-temperature rolled rolled alloy alloy (a ()a before) before annealing annealing and and (b (–bd–)d after) after annealingannealing at at various various temperatures: temperatures: (b (b) )250 250 °C,◦C, (c (c) )300 300 °C,◦C, and and (d (d) )350 350 °C.◦C.

RecrystallizedRecrystallized grains grains first first nucleate nucleate and and grow grow in in the the high high-energy-energy regions regions of of a a grain grain andand then then expand expand to to spread spread over over the the entire entire grain. grain. High High-energy-energy regions, regions, such suchas high as-den- high- sitydensity dislocation dislocation areas, areas, junction junction of twin of boundaries twin boundaries,, and grain and boundaries, grain boundaries, provide provideprefer- entialpreferential recrystallization recrystallization sites during sites duringheat treatme heat treatment.nt. The deformation The deformation twins formed twins formed by ul- tralowby ultralow-temperature-temperature rolling rolling can increase can increase the grain the grain boundary boundary energy energy inside inside grains grains and andef- fectivelyeffectively prevent prevent the the excessive excessive growth growth of recrystallized of recrystallized grains. grains. Figure Figure 4 indicates4 indicates that that the microstructurethe microstructure was wasgradually gradually transforme transformedd from from deformed deformed grains grains formed formed by by ultralow ultralow-- temperature rolling to undistorted equiaxed grains of annealing for different times at 350 ◦C. It is noticeable from Figure4b that after annealing the sample at 350 ◦C for 3 min, many twins remained in the sample interior and many fine static recrystallized grains appeared in the unevenly deformed structure, mainly in the interior of metastable grains and grain boundaries. As shown in Figure4c, a few twins remained in the structure, and the grains had not yet completed the static recrystallization process after annealing the sample for 6 min. This result was obtained because the short annealing period did not provide sufficient time and energy for small grains to nucleate and grow. As can be seen from Figure4d, the deformed grains gradually became fine recrystallized grains after annealing the sample for 10 min. As shown in Figure4e, the entire metallographic structure was composed of uniformly sized equiaxed grains after annealing the alloy for 30 min. The Metals 2021, 11, x FOR PEER REVIEW 6 of 16

temperature rolling to undistorted equiaxed grains of annealing for different times at 350 °C. It is noticeable from Figure 4b that after annealing the sample at 350 °C for 3 min, many twins remained in the sample interior and many fine static recrystallized grains appeared in the unevenly deformed structure, mainly in the interior of metastable grains and grain boundaries. As shown in Figure 4c, a few twins remained in the structure, and the grains had not yet completed the static recrystallization process after annealing the sample for 6 min. This result was obtained because the short annealing period did not

Metals 2021, 11, 315 provide sufficient time and energy for small grains to nucleate and grow. As can be 6seen of 15 from Figure 4d, the deformed grains gradually became fine recrystallized grains after an- nealing the sample for 10 min. As shown in Figure 4e, the entire metallographic structure was composed of uniformly sized equiaxed grains after annealing the alloy for 30 min. Thestatic static recrystallization recrystallization process process was completed,was completed, and the and resulting the resulting grains grains were approximatelywere approx- imately15.5 µm 15.5 in size µm on in average.size on average.

FigureFigure 4. 4. MicrostructureMicrostructure of of ultralow ultralow-temperature-temperature rolled rolled alloy alloy (a ()a before) before annealing annealing and and (b (–be–) eafter) after annealing for different times: (b) 3 min, (c) 6 min, (d) 10 min and (e) 30 min. annealing for different times: (b) 3 min, (c) 6 min, (d) 10 min and (e) 30 min.

3.2.3.2. TEM TEM A Analysisnalysis InIn order order to to observe observe the the dislocation dislocation changes changes and and the the formation formation of of recrystallized recrystallized grains grains ofof the the alloy alloy before before and and after after annealing, annealing, the the bright bright field field image image of of the the sample sample before before and and afterafter annealing annealing at at 350 350 °C◦C for for 10 10 min min (Figure (Figure 5)5 )was was used used for for observation observation and and analysis. analysis. It canIt can be beseen seen from from Figure Figure 5a,c5a,c that that a large a large accumulated accumulated strain strain resulted resulted in ina high a high density density of dislocationof dislocation after after the thealloys alloys under underwentwent ultralow ultralow-temperature-temperature rolling. rolling. It is It interesting is interesting to noteto note that that these these high high-density-density dislocations dislocations entangle entangledd and and accumulate accumulatedd near near the the grain boundaries into dislocation cells, and the grain boundaries were rough. It can be seen from Figure5b,d that for the sample annealed at 350 ◦C for 10 min, part of the dislocation cell shape occurred during the annealing process, and the cell walls were flattened to form

subgarins. The original rough grain boundaries also became smooth. In addition, the subgrain boundaries with larger dislocation density migrated toward the surrounding subgrains with larger dislocation difference and swallowed adjacent deformations. The de- formed grains and subgrains were gradually transformed into high-angle grain boundaries, becoming the center of recrystallization nucleation and growing to form recrystallized grains. The original dislocation entanglement position distribution became dispersed, and the grain boundaries were clear between recrystallized grains. Metals 2021, 11, x FOR PEER REVIEW 7 of 16

boundaries into dislocation cells, and the grain boundaries were rough. It can be seen from Figure 5b,d that for the sample annealed at 350 °C for 10 min, part of the dislocation cell shape occurred during the annealing process, and the cell walls were flattened to form subgarins. The original rough grain boundaries also became smooth. In addition, the sub- grain boundaries with larger dislocation density migrated toward the surrounding sub- grains with larger dislocation difference and swallowed adjacent deformations. The de- formed grains and subgrains were gradually transformed into high-angle grain bounda- ries, becoming the center of recrystallization nucleation and growing to form recrystal- Metals 2021, 11, 315 lized grains. The original dislocation entanglement position distribution 7 bec of 15ame dis- persed, and the grain boundaries were clear between recrystallized grains.

Figure 5. Bright field images of ultralow-temperature rolled sample before and after annealing at Figure350 ◦C 5. for Bright 10 min. field (a, cimages) 0 min, of (b ultralow,d) 10 min.-temperature (Note: there rolled could sample be the possible before and formation after annealing of an at 350artificial °C for generation 10 min. ( ofa,c few) 0 smallmin, ( dislocationsb,d) 10 min during. (Note: ion there milling.) could be the possible formation of an arti- ficial generation of few small dislocations during ion milling.) Figure6 shows TEM and the corresponding selected area electron diffraction images of theFigure alloy. As 6 shows can be seenTEM from and Figure the corresponding6a, the dislocation selected density area of the elect alloyron was diffraction relatively images high after ultralow-temperature rolling, and a high-density dislocation zone was formed of the alloy. As can be seen from Figure 6a, the dislocation density of the alloy was rela- around the twins with staggered dislocations. It is clear from Figure6c that the deformation tivelytwin boundary high after was ultralow not straight,-temperature and the nearbyrolling, lattice and stripesa high- ofdensity the matrix dislocation and twins zone was formedoverlapped. around The the arrow twins in the with figure staggered shows that dislocations. stacking faults It is onclear the from basal planeFigure were 6c that the deformationobserved in both twin the boundary matrix and was the twins not straight at the atomic, and scale, the nearby which is lattice similar stripes to the resultsof the matrix andobserved twins by overlap Matsudaped et. al.The [32 arrow] in their in studythe figure on the shows interaction that betweenstacking a faults long-period on the basal planestacked were ordered observed phase in and both deformation the matrix twins and in the the twins rapid at solidification the atomic ofscale, a Mg which97Zn1Y is2 similar toalloy. the Diffractionresults observed calibration by ofMatsuda the twins et in al. the [32] selected in their area study showed on that the these interaction spots were between a − longthe diffraction-period stacked spots of ordered {1012} tensile phase twins. and deformation This result was twins obtained in the because rapid thesolidification tension of a − − Mg97Zn1Y2 alloy. Diffraction calibration of the twins in the selected area showed that these force on the c-axis mainly produces {1012} ̄ <1011> tensile twins during alloy deformation. spots were the diffraction spots of {1012} −tensile twins. This result was obtained because The critical shear stress of 1–2 MPa [33] of {1012} tensile twins ̄  ̄ is very small. The external the tension force on the c-axis mainly produces {1012} <1011> tensile twins during alloy force exerted by the deformation of the alloy under ultralow temperature ̄ conditions can deformation.easily reach the The critical critical shear shear stress. stress of 1–2 MPa [33] of {1012} tensile twins is very small.

Metals 2021, 11, x FOR PEER REVIEW 8 of 16

Metals 2021, 11, 315 8 of 15 The external force exerted by the deformation of the alloy under ultralow temperature conditions can easily reach the critical shear stress.

FigureFigure 6. 6. TransmissionTransmission electron electron microscopy microscopy (TEM (TEM)) and and selected selected area area electron electron diffraction diffraction (SAED) (SAED) images of ultralow-temperature rolled alloy: (a,b) bright field images, (c) b-pattern twin and two- images of ultralow-temperature rolled alloy: (a,b) bright field images, (c) b-pattern twin and two- dimensional lattice diagram near a boundary, and (d) diffraction pattern corresponding to twin dimensional lattice diagram near a boundary, and (d) diffraction pattern corresponding to twin crystal crystal area. (Note: there could be the possible formation of an artificial generation of few disloca- tionsarea. during (Note: ion there milling could.) be the possible formation of an artificial generation of few dislocations during ion milling.)

3.3.3.3. EBSD EBSD A Analysisnalysis UUltralow-temperatureltralow-temperature rolled rolled alloy alloy has has a a large large storage storage capacity, capacity, and and its its driving driving force force forfor recrystallization recrystallization during during the the annea annealingling process process is also is also large, large, meaning meaning the thedegree degree of re- of crystallizationrecrystallization is ishigh. high. Figure Figure 7 7presents presents a adistribution distribution map map of of recrystallized recrystallized grains of the ultralowultralow-temperature-temperature rolled rolled alloy alloy at at 350 350 °C◦C before before and and after after annealing annealing for for different different times. times. AsAs the annealing annealing time time increased, increased the, the proportion proportion of deformed of deformed grains grains (red) decreased (red) decrease sharply,d sharply,the number the number of subcrystalline of subcrystalline grains (yellow)grains (yellow) increased, increase and thed, and fraction the fraction of recrystallized of recrys- tallizedgrains (blue)grainsincreased (blue) increase fromd 11.62% from 11.62% to 39.55%. to 39.55%. The occurrence The occurrence of static of recrystallizationstatic recrystal- lizationconsiderably considerably weakened weaken theed strong the strong texture texture formed formed during during the deformation the deformation process process and andsoftened soften theed the structure. structure. It isIt is apparent apparent from from Figure Figure7d 7d that that for for the the ultralow-temperatureultralow-temperature rolledrolled alloy alloy annealedannealed atat 350350◦ C°C for for 10 10 min, min a, largea large number number of fineof fine grains grains appeared appear compareded com- paredto annealing to annealing for 3 and for 5 3 min and (Figure 5 min7 (Figureb,c). Because 7b,c). theBecause recrystallization the recrystallization process is controlledprocess is controlledby thermal by activation, thermal activation, the recrystallization the recrystallization temperature temperature will reduce will with reduce the increase with the of increaseannealing of time,annealing the recrystallization time, the recrystallization nucleation is nucleation faster, and is the faster amount, and of the nucleation amount perof nucleationunit volume per is more, unit volume meaning is the more, final meaning recrystallization the final grain recrystallization size is smaller. grain size is smaller.Figure 8 manifests images of the grain orientation of the alloy before and after an- nealing at 350 ◦C for different times. There are a host of twins and elongated deformed grains in ultralow-temperature rolled alloy, including many metastable structures inside the grains, generous dislocation entanglements, and dislocation cells, which strengthen the basal plane orientation of the alloy grains. The rolled alloy has a relatively large internal strain and stored distortion energy such that the high-energy regions act as nucleation sites during subsequent annealing, resulting in the formation of many fine recrystallized grains. Because the growth of the recrystallized core is realized by migration of the large-angle interface, the recrystallization will eliminate or change the original deformation texture. As the annealing time increased, the deformation texture was greatly reduced, and the

Metals 2021, 11, 315 9 of 15

recrystallization texture was significantly increased. The basal-oriented crystal grains were Metals 2021, 11, x FOR PEER REVIEW 9 of 16 significantly reduced, and the column-oriented crystal grains increased, that is, the blue crystal grains increased.

Metals 2021, 11, x FOR PEER REVIEWFigureFigure 7. 7. DistributionDistribution map map of of recrystallized recrystallized grains grains of of ultralow ultralow-temperature-temperature ro rolledlled alloy alloy (a) ( abefore)10 before of 16 annealingannealing and and ( (bb––dd)) after after annealing annealing for for different different times: times: ( (bb)) 3 3 min, min, ( (cc)) 6 6 min, min, and and ( (dd)) 10 10 min. min.

Figure 8 manifests images of the grain orientation of the alloy before and after an- nealing at 350 °C for different times. There are a host of twins and elongated deformed grains in ultralow-temperature rolled alloy, including many metastable structures inside the grains, generous dislocation entanglements, and dislocation cells, which strengthen the basal plane orientation of the alloy grains. The rolled alloy has a relatively large inter- nal strain and stored distortion energy such that the high-energy regions act as nucleation sites during subsequent annealing, resulting in the formation of many fine recrystallized grains. Because the growth of the recrystallized core is realized by migration of the large- angle interface, the recrystallization will eliminate or change the original deformation tex- ture. As the annealing time increased, the deformation texture was greatly reduced, and the recrystallization texture was significantly increased. The basal-oriented crystal grains were significantly reduced, and the column-oriented crystal grains increased, that is, the blue crystal grains increased.

Figure 8. ImageImage showing showing grain grain orientation orientation im imageage of of ultralow ultralow-temperature-temperature rolled alloy ( a) before before annealing and ( b–d) after annealing at different times: ( b) 3 min, (c) 6 min, and (d) 10 min.

The phase difference (θ) between adjacent grains is used to divide the grain bound- aries into low-angle grain boundaries (LAGB, 2° < θ < 10°) and high-angle grain bounda- ries (HAGB, θ > 10°). Figure 9 is a misorientation distribution diagram of the alloy before and after annealing at 350 °C for different times, showing the peak distribution of the misorientation angle at approximately less than 10° and a significant increase in the peak orientation angle at approximately 86°. As shown in Figure 9a, the peak at 86° was higher than at the other angles, except for dislocation slip and grain boundary sliding, where {10  ̄  ̄ 12}(86° <1120>) tensile twinning also contributed to plastic deformation under this condi- tion. This result shows that due to the low critical shear stress, 86° tensile twins were pref- erentially generated at the initial stage of plastic deformation. Grain boundaries and twin boundaries hindered the movement of dislocations during plastic deformation, resulting in the accumulation of dislocations; furthermore, the twins generated at the initial stage of plastic deformation changed the grain orientation, thereby activating the nonbasal sur- face slip system and promoting interactions among the grain boundaries, twin bounda- ries, and dislocations in the subsequent deformation process, which enhanced the room- temperature tensile plastic deformation ability of the alloy. The ultralow-temperature rolled alloy promoted the formation of a wide range of low-angle grain boundaries due to the interaction of high-density twin boundaries and adjacent dislocations. Statistics of the high-angle and low-angle grain boundaries under different conditions showed that the proportion of low-angle grain boundaries in the ultralow-temperature rolled alloy was 54.9% after annealing at different times. The proportion of high-angle grain boundaries increased, in contrast to the decrease in the proportion of low-angle grain boundaries. In addition, after annealing at 350 °C for 3, 6, and 10 min, the proportions of the low-angle grain boundaries were 43.8%, 34.8%, and 33.4%respectively. Combined with Figure 6a, it can be seen that a large number of dislocation entanglements and dislocation cells were generated inside the grains of the ultralow-temperature rolled alloy. The walls of the dis- location cells were entangled with a large number of dislocations, forming a low-angle grain boundary by dislocation plugging. The accumulation of the dislocation density re-

Metals 2021, 11, 315 10 of 15

The phase difference (θ) between adjacent grains is used to divide the grain boundaries into low-angle grain boundaries (LAGB, 2◦ < θ < 10◦) and high-angle grain boundaries (HAGB, θ > 10◦). Figure9 is a misorientation distribution diagram of the alloy before and after annealing at 350 ◦C for different times, showing the peak distribution of the misorientation angle at approximately less than 10◦ and a significant increase in the peak orientation angle at approximately 86◦. As shown in Figure9a, the peak at 86 ◦ was higher than at the other angles, except for dislocation slip and grain boundary sliding, where − − {1012}(86◦ <1120>) tensile twinning also contributed to plastic deformation under this condition. This result shows that due to the low critical shear stress, 86◦ tensile twins were preferentially generated at the initial stage of plastic deformation. Grain boundaries and twin boundaries hindered the movement of dislocations during plastic deformation, resulting in the accumulation of dislocations; furthermore, the twins generated at the initial stage of plastic deformation changed the grain orientation, thereby activating the nonbasal surface slip system and promoting interactions among the grain boundaries, twin boundaries, and dislocations in the subsequent deformation process, which enhanced the room-temperature tensile plastic deformation ability of the alloy. The ultralow-temperature rolled alloy promoted the formation of a wide range of low-angle grain boundaries due to the interaction of high-density twin boundaries and adjacent dislocations. Statistics of the high-angle and low-angle grain boundaries under different conditions showed that the proportion of low-angle grain boundaries in the ultralow-temperature rolled alloy was 54.9% after annealing at different times. The proportion of high-angle grain boundaries increased, in contrast to the decrease in the proportion of low-angle grain boundaries. In addition, after annealing at 350 ◦C for 3, 6, and 10 min, the proportions of the low-angle grain boundaries were 43.8%, 34.8%, and 33.4%respectively. Combined with Figure6a, it can be seen that a large number of dislocation entanglements and dislocation cells were generated inside the grains of the ultralow-temperature rolled alloy. The walls of the dislo- cation cells were entangled with a large number of dislocations, forming a low-angle grain boundary by dislocation plugging. The accumulation of the dislocation density resulted in the formation of a dislocation packing area, which increased the internal distortion of the material, and abundant dislocations entangled into a subcrystalline structure. During annealing, subgrains with high dislocation migrated to surrounding subgrains with larger orientation differences such that the adjacent deformed matrix was swallowed into subgrain growth; thus, the phase difference between adjacent subgrains increased and gradually transformed into high-angle grain boundaries that served as recrystallization nucleation sites and grew further, increasing the proportion of high-angle grain bound- aries. Consequently, the proportion of low-angle grain boundaries decreased, that is, the dislocation density decreased. The ultralow-temperature rolling of the alloy increased the critical shear stress (CRSS) for basal slip such that the twin deformation became the main deformation mechanism, and the large number of twins produced promoted further plastic deformation of the alloy. Figure 10 shows that the distribution of twin boundaries of the alloy before and after annealing at 350 ◦C for different times. The four types of twins shown in four different − − colors were mainly 86◦ <1120> tensile twins, a few 56◦ <1120> compression twins, 64◦ − − <1120> compression twins, and 38◦ <1120> secondary twins. A large number of twins formed during ultralow-temperature deformation were swallowed and consumed during recrystallization; thus, the twin content gradually decreased. Channel 5 software was used to calculate the proportion of twin boundaries and it was found that the ultralow- − temperature rolled alloy 86◦ <1120> tensile twin content was 14.09%, followed by 6.57%, 5.25%, and 1.22% after annealing. Rolling at an ultralow temperature pulled the alloy material along the c-axis direction, making it easy to reach the small required critical − shear stress of {1012} tensile twins; consequently, the magnesium alloys in this study were − dominated by 86◦ <1120> tensile twins during the deformation process. Figure 10 also Metals 2021, 11, x FOR PEER REVIEW 11 of 16

sulted in the formation of a dislocation packing area, which increased the internal distor- tion of the material, and abundant dislocations entangled into a subcrystalline structure.

Metals 2021, 11, 315 During annealing, subgrains with high dislocation densities migrated to surrounding sub- 11 of 15 grains with larger orientation differences such that the adjacent deformed matrix was swallowed into subgrain growth; thus, the phase difference between adjacent subgrains increased and gradually transformed into high-angle grain boundaries that served as re- − crystallizationdemonstrates nucleation sites that and the gr 86e◦w< further,1120> tensile increasing twins the accounted proportion for theof high highest-angle proportion of all grain boundaries.twins Consequently, before and after the annealing,proportion and of low this-angle gradually grain decreased boundaries in decrease number asd, the annealing that is, the dislocationtime increased. density decreased.

Metals 2021, 11, x FOR PEER REVIEW 12 of 16

Figure 9. ImageFigure of 9. misorientation Image ofand misorientation after distribution annealing, distribution of ultralow-temperature and this of ultralow gradually-temperature rolled decrease alloy rolledd (a in) before numberalloy annealing(a) beforeas the an- and annealing (b–d) after time in- annealing fornealing different and times: (b–dcrease)( bafter) 3 min, annealingd. (c) 6 min, for different and (d) 10 times: min. (b) 3 min, (c) 6 min, and (d) 10 min. The ultralow-temperature rolling of the alloy increased the critical shear stress (CRSS) for basal slip such that the twin deformation became the main deformation mech- anism, and the large number of twins produced promoted further plastic deformation of the alloy. Figure 10 shows that the distribution of twin boundaries of the alloy before and after annealing at 350 °C for different times. The four types of twins shown in four differ-  ̄  ̄ ent colors were mainly 86° <1120> tensile twins, a few 56° <1120> compression twins, 64°  ̄  ̄ <1120> compression twins, and 38° <1120> secondary twins. A large number of twins formed during ultralow-temperature deformation were swallowed and consumed during recrystallization; thus, the twin content gradually decreased. Channel 5 software was used to calculate the proportion of twin boundaries and it was found that the ultralow-temper-  ̄ ature rolled alloy 86° <1120> tensile twin content was 14.09%, followed by 6.57%, 5.25%, and 1.22% after annealing. Rolling at an ultralow temperature pulled the alloy material along the c-axis direction, making it easy to reach the small required critical shear stress  ̄ of {1012} tensile twins; consequently, the magnesium alloys in this study were dominated  ̄ by 86° <1120> tensile twins during the deformation process. Figure 10 also demonstrates  ̄ that the 86° <1120> tensile twins accounted for the highest proportion of all twins before

FigureFigure 10. 10.Twin Twin boundary boundary distribution distribution of of ultralow-temperature ultralow-temperature rolled rolled alloy alloy (a ()a before) before annealing annealing and and (b (–bd–)d after) after annealing annealing

forfor different different times: times: (b ()b 3) 3 min, min, (c ()c 6) 6 min, min, and and (d ()d 10) 10 min. min.

3.4. Mechanical Properties Figure 11 shows a graph of the trend in microhardness change for the alloy before and after annealing at 350 °C for different times. As the annealing time increased, the av- erage hardness of the alloy gradually decreased by 10.32% from the original value of 63.12 to 56.6 HV. When the alloy was deformed at ultralow temperature, it was difficult for dislocations to slip and climb, which promoted work hardening. Work hardening pro- duced relatively large stresses and strain accumulation in the ultralow-temperature rolled alloy, resulting in many twins and a relatively high dislocation density. Dislocations moved and entangled into cells, resulting in a relatively high hardness. Annealing the sample for different times released the internal stress in the material and reduced the dis- location density and twin content, thereby weakening the work hardening and texture; static recrystallization and nucleation of crystal grains in regions of high-distortion energy produced new undistorted grains such that the hardness gradually decreased. After the annealing time exceeded 10 min, the recrystallized grains began to grow. The grain size had little effect on the hardness, which was therefore maintained within a certain range. The interaction between dislocations and between dislocations and twins, the texture, and the accumulated strain were all factors that significantly changed the microhardness of the ultralow-temperature rolled magnesium alloy after annealing.

Metals 2021, 11, 315 12 of 15

3.4. Mechanical Properties Figure 11 shows a graph of the trend in microhardness change for the alloy before and after annealing at 350 ◦C for different times. As the annealing time increased, the average hardness of the alloy gradually decreased by 10.32% from the original value of 63.12 to 56.6 HV. When the alloy was deformed at ultralow temperature, it was difficult for dislocations to slip and climb, which promoted work hardening. Work hardening produced relatively large stresses and strain accumulation in the ultralow-temperature rolled alloy, resulting in many twins and a relatively high dislocation density. Dislocations moved and entangled into cells, resulting in a relatively high hardness. Annealing the sample for different times released the internal stress in the material and reduced the dislocation density and twin content, thereby weakening the work hardening and texture; static recrystallization and nucleation of crystal grains in regions of high-distortion energy produced new undistorted grains such that the hardness gradually decreased. After the annealing time exceeded 10 min, the recrystallized grains began to grow. The grain size Metals 2021, 11, x FOR PEER REVIEW had little effect on the hardness, which was therefore maintained within a certain13 of 16 range. The interaction between dislocations and between dislocations and twins, the texture, and the accumulated strain were all factors that significantly changed the microhardness of the ultralow-temperature rolled magnesium alloy after annealing.

64

62

60

58

Hardness (HV)

56

0 10 20 30 40 50 60 Annealing time (min)

Figure 11. Microhardness of ultralow-temperature rolled alloy before and after annealing for different times. Figure 11. Microhardness of ultralow-temperature rolled alloy before and after annealing for dif- ferent times.In the annealing process, the recrystallization grains are refined by increasing the recrystallization nucleation rate and hindering the growth of grains. The alloy is deformed In the annealing process, the recrystallization grains are refined by increasing the by ultralow-temperature rolling, and a large number of defects, such as dislocations and recrystallization nucleation rate and hindering the growth of grains. The alloy is deformed twins, are formed, that is, deformation storage energy, which provides a greater driving by ultralowforce for-temperature static recrystallization rolling, and during a large the number annealing of defects process,, such thereby as dislocations increasing the and recrys- twinstallization, are formed, nucleation that is, rate.deformation At the same storage time, energy, the formation which provides of a large a numbergreater ofdriving twins can forceincrease for static the recrystallization grain boundary during energy the inside annealing the grains process, so that thereby there increasing is not enough the drivingre- crystallizationforce for the nucleation growth ofrate. the At grains, the same which time, effectively the formation hinders of thea large growth number of the of grainstwins and can increaseobtains the fine grain recrystallized boundary grains. energy Figureinside the12 andgrains Table so that1 shows there the is not corresponding enough driv- data ing forcestatistics for the for growth the alloy of beforethe grains, and which after annealing effectively at hinders 350 ◦C the for differentgrowth of times. the grains It can be and obtainsseen from fine Figure recrystallized 12 that thegrains. ultralow-temperature Figure 12 and Table rolled 1 shows alloy the had corresponding high ultimate data tensile statisticsstrength for the (UTS) alloy and before low elongationand after annealing (EL). On theat 350 one °C hand, for different in conjunction times. withIt can the be alloy seen shownfrom Figure in Figure 12 4thata, it the can ultralow be seen that-temperature after four rolled passes alloy of ECAP had andhigh ultralow-temperature ultimate tensile strengthrolling, (UTS ultrafine) and low crystal elongation twins were (EL). formed On the to one produce hand, finein conjunction grain strengthening with the andalloy grain shownboundary in Figure strengthening. 4a, it can be seen The that twin after boundaries four passes could of beECAP equivalent and ultralow to the- effecttempera- of grain ture rolling,boundaries, ultrafine effectively crystal hinderingtwins were the formed movement to produce of dislocations fine grain andstrengthening preventing and plastic graindeformation boundary strengthening. of the material; The when twin the boun dislocationsdaries could moved be equivalent to the twin to grain the effect boundary of or grainthe boundaries continuous, effectively accumulation hinder ofing the the grain movement boundary, of the dislocations linear defect and density prevent insideing the plasticmaterial deformation increased, of the thereby material; promoting when the improvement dislocations move of thed materialto the twin strength. grain bound- Therefore, ary orthe the ultralow-temperature continuous accumulation rolling of formed the grain a large boundary, number the of linear twin graindefect boundaries density inside and pro- the material increased, thereby promoting improvement of the material strength. There- fore, the ultralow-temperature rolling formed a large number of twin grain boundaries and produced grain boundary strengthening. On the other hand, combined with Figure 6a,c, it can be seen that due to multiple deformations of the alloy, a large number of dis- locations were formed to form dislocation entanglement to produce dislocation strength- ening. In general, fine grain strengthening, dislocation strengthening, and grain boundary strengthening all make contributions to high tensile strength. With the extension of the annealing time, the tensile strength of the alloy decreased and the elongation increased. After annealing of ultralow-temperature rolled alloy, its tensile strength reduced because of reduction in the dislocation density and twin boundary content. The alloy recrystallized during the annealing process, the material became soft, and its elongation increased. Com- bining the recrystallized grain distribution in Figure 5, it can be seen that when the recrys- tallization fraction was low (11.62%), the corresponding elongation was very low (11%). The elongation of the alloy was significantly improved when the recrystallization fraction reached 39.55%, increasing from 11% to 29%. After the alloy was annealed at 350 °C for 10 min, the tensile strength decreased from the original 245.89 to 186.15 MPa.

Metals 2021, 11, 315 13 of 15

duced grain boundary strengthening. On the other hand, combined with Figure6a,c, it can be seen that due to multiple deformations of the alloy, a large number of dislocations were formed to form dislocation entanglement to produce dislocation strengthening. In general, fine grain strengthening, dislocation strengthening, and grain boundary strengthening all make contributions to high tensile strength. With the extension of the annealing time, the tensile strength of the alloy decreased and the elongation increased. After annealing of ultralow-temperature rolled alloy, its tensile strength reduced because of reduction in the dislocation density and twin boundary content. The alloy recrystallized during the annealing process, the material became soft, and its elongation increased. Combining the recrystallized grain distribution in Figure5, it can be seen that when the recrystallization Metals 2021, 11, x FOR PEER REVIEW fraction was low (11.62%), the corresponding elongation was very low (11%). The14 of elonga-16 tion of the alloy was significantly improved when the recrystallization fraction reached 39.55%, increasing from 11% to 29%. After the alloy was annealed at 350 ◦C for 10 min, the tensile strength decreased from the original 245.89 to 186.15 MPa.

250

200

150

100

Stress (MPa) Ultra-low-temperature-rolled state Annealing for 3 min 50 Annealing for 6 min Annealing for 10 min

0 0 5 10 15 20 25 30 35 40 Strain (%)

Figure 12. Stress–strain curves of ultralow-temperature rolled alloy before and after annealing for Figure 12. Stress–strain curvesdifferent of ultralow times.-temperature rolled alloy before and after annealing for different times.

Table Table1. Mechanical 1. Mechanical properties properties of ultralow of ultralow-temperature-temperature rolled rolled alloy alloybefore before and after and annealing after annealing for different for different times. times.

SampleSample Ultimate Ultimate Tensile Strength Strength (UTS/MPa) (UTS/MPa) Elongation Elongation (EL/ (EL/%)%) Ultralow-temperature rolled state 245.89 11 Ultralow-temperature rolled state 245.89 11 AnnealingAnnealing for for3 min 3 min 198.8198.8 26 26 AnnealingAnnealing for for6 min 6 min 192.64192.64 25 25 AnnealingAnnealing for for10 10min min 186.15186.15 29 29

4. Conclusions4. Conclusions 1. The1. alloyThe was alloy treated was treated with ECAP with ECAP to obtain to obtain fine grains fine grains.. Ultralow Ultralow-temperature-temperature rolling rolling introduceintroducedd a large a largenumber number of high of- high-energyenergy defects, defects, such such as grain as grain boundaries boundaries and anddis- dislo- locations,cations, which which serve servedd as preferential as preferential nucleation nucleation sites sites during during subsequent subsequent annealing annealing for for recrystallized grain formation. The twins formed by tension exerted along the c- recrystallized grain formation. The twins formed by tension exerted ̄  ̄ along the c-axis axis during rolling under ultralow temperatures were mainly {10−12} −<1011> tensile during rolling under ultralow temperatures were mainly {1012} <1011> tensile twins. twins. 2. Increasing the annealing time at 350 ◦C resulted in the gradual completion of the 2. Increasing the annealing time at 350 °C resulted in the gradual completion of the static recrystallization process. The walls of the cells formed by dislocation entangle- static recrystallization process. The walls of the cells formed by dislocation entangle- ment flattened and formed subcrystalline grains during annealing, thereby gradually ment flattened and formed subcrystalline grains during annealing, thereby gradually decreasing the twin content and the proportion of small-angle grain boundaries. decreasing3. By increasing the twin thecontent annealing and the time proportion at 350 ◦C, of the small hardness-angle grain decreased boundaries. by 10.32% from

3. By increasing63.12 to 56.6the HV,annealing the tensile time strength at 350 °C, of thethe alloy hardness was reduced decrease fromd by 245.89 10.32% to 186.15from MP, 63.12 andto 56.6 the HV, plasticity the tensile of the strength alloy significantly of the alloy increased was reduced by 34.48% from 245.89 from 11% to 186.15 to 29%. MP, and the plasticity of the alloy significantly increased by 34.48% from 11% to 29%.

Author Contributions: H.Y conceived and designed the experiment, collected the data, analyzed the data, and wrote the paper; Y.T. (Yun Tan) contributed to retrieving documents; and W.L. and Y.T. (Yuanbiao Tan) checked and revised the paper. All authors have read and agreed to the pub- lished version of the manuscript. Funding: This work was supported by the National Natural Science Foundation of China (51661007), the Science and Technology Plan Project of Guizhou Province ([2019] 5303), and the Cul- tivation Project of Guizhou University ([2019] 15). Data Availability Statement: No new data were created or analyzed in this study. Data sharing is not applicable to this article. Conflicts of Interest: The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Metals 2021, 11, 315 14 of 15

Author Contributions: H.Y. conceived and designed the experiment, collected the data, analyzed the data, and wrote the paper; Y.T. (Yun Tan) contributed to retrieving documents; and W.L. and Y.T. (Yuanbiao Tan) checked and revised the paper. All authors have read and agreed to the published version of the manuscript. Funding: This work was supported by the National Natural Science Foundation of China (51661007), the Science and Technology Plan Project of Guizhou Province ([2019] 5303), and the Cultivation Project of Guizhou University ([2019] 15). Data Availability Statement: No new data were created or analyzed in this study. Data sharing is not applicable to this article. Conflicts of Interest: The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Abbreviations

ECAP equal channel angular pressing OM optical microscopy EBSD electron backscatter diffraction technology TEM transmission electron microscopy CYS compress yield strength UTS ultimate tensile strength EL elongation HCP hexagonal close packed SPD severe plastic deformation UFG ultrafine grain HPT high-pressure torsion MAF multiaxial forging ARB accumulative roll-bonding YS yield strength ND normal direction ED extrusion direction RD rolling direction TD transverse direction θ the phase difference between adjacent grains LAGB low-angle grain boundaries (2◦ < θ < 10◦) HAGB high-angle grain boundaries (θ > 10◦) f LAGB the proportion of low-angle grain boundary to total grain boundary

References 1. Adel, A.; Danil, P.; Ivan, E.; Mohamed, T.; Mahmoud, S.; Magdy, E.R. ANN Surface Roughness Optimization of AZ61 Magnesium Alloy Finish Turning: Minimum Machining Times at Prime Machining Costs. Materials 2018, 11, 808. 2. Jing, T.W.; Yin, D.L.; Jin, Q.L.; Tao, J.; Yan, L.S.; Xiang, Z. Effect of grain size on mechanical property of Mg–3Al–1Zn alloy. Scr. Mater. 2008, 59, 63–66. 3. Liu, Q.; Zhou, X.; Zhou, H.; Fan, X.; Liu, K. The effect of extrusion conditions on the properties and textures of AZ31B alloy. J. Magn. Alloy 2017, 5, 202–209. [CrossRef] 4. Reed-Hill, R.E.; Robertson, W.D. Additional modes of deformation twinning in magnesium. Acta Metall. 1957, 5, 717–727. [CrossRef] 5. Atik, K.; Efe, M. Twinning-induced shear banding and its control in rolling of magnesium. Mater. Sci. Eng. A 2018, 725, 267–273. [CrossRef] 6. Edalati, K.; Yamamoto, A.; Horita, Z.; Ishihara, T. High-pressure torsion of pure magnesium: Evolution of mechanical properties, microstructures and hydrogen storage capacity with equivalent strain. Scr. Mater. 2011, 64, 880–883. [CrossRef] 7. Schafler, E.; Pippan, R. Effect of thermal treatment on microstructure in high pressure torsion (HPT) deformed nickel. Mater. Sci. Eng. A 2004, 387, 799–804. [CrossRef] 8. Bhowmik, A.; Biswas, S.; Dhinwal, S.S.; Sarkar, A.; Ray, R.K.; Bhattacharjee, D.; Suwas, S. Microstructure and Texture Evolution in Interstitial-Free (IF) Processed by Multi-Axial Forging. Mater. Sci. Forum 2012, 702–703, 774–777. [CrossRef] 9. Shamsolhodaei, A.; Zarei-Hanzaki, A.; Moghaddam, M. Structural and functional properties of a semi equiatomic NiTi shape memory alloy processed by multi-axial forging. Mater. Sci. Eng. A 2017, 700, 1–9. [CrossRef] Metals 2021, 11, 315 15 of 15

10. Duan, J.Q.; Quadir, M.Z.; Xu, W.; Kong, C.; Ferry, M. Texture balancing in a fcc/bcc multilayered composite produced by accumulative roll bonding. Acta Mater. 2017, 123, 11–23. [CrossRef] 11. Saito, Y.; Utsunomiya, H.; Tsuji, N.; Sakai, T. Novel ultra-high straining process for bulk materials—Development of the accumulative roll-bonding (ARB) process. Acta Mater. 1999, 47, 579–583. [CrossRef] 12. Matsubara, K.; Miyahara, Y.; Horita, Z.; Langdon, T.G. Developing superplasticity in a magnesium alloy through a combination of extrusion and ECAP. Acta Mater. 2003, 51, 3073–3084. [CrossRef] 13. Tham, Y.W.; Fu, M.W.; Hng, H.H.; Pei, Q.X.; Lim, K.B. Microstructure and Properties of Al-6061 Alloy by Equal Channel Angular Extrusion for 16 Passes. Adv. Manuf. Processes 2007, 22, 819–824. [CrossRef] 14. Beyerlein, L.J.; Tóthb, L.S. Texture evolution in equal-channel angular extrusion. Prog. Mater. Sci. 2009, 54, 427–510. [CrossRef] 15. Jahadi, R.; Sedighi, M.; Jahed, H. ECAP effect on the micro-structure and mechanical properties of AM30 magnesium alloy. Mater. Sci. Eng. A 2014, 593, 178–184. [CrossRef] 16. Garces, G.; Muñoz-Morris, M.A.; Morris, D.G.; Adeva, P.; Perez, P. Optimization of strength by microstructural refinement of MgY_2Zn_1 alloy during extrusion and ECAP processing. Mater. Sci. Eng. A 2014, 614, 96–105. [CrossRef] 17. Sarker, D.; Friedman, J.; Chen, D.L. De-twinning and Texture Change in an Extruded AM30 Magnesium Alloy during Compression along Normal Direction. J. Mater. Sci. Technol. 2015, 31, 264–268. [CrossRef] 18. Sarker, D.; Friedman, J.; Chen, D.L. Twin Growth and Texture Evolution in an Extruded AM30 Magnesium Alloy During Compression. J. Mater. Sci. Technol. 2014, 30, 884–887. [CrossRef] 19. Sarker, D.; Chen, D.L. Detwinning and strain hardening of an extruded magnesium alloy during compression. Scr. Mater. 2012, 67, 165–168. [CrossRef] 20. Sarker, D.; Chen, D.L. Dependence of compressive deformation on pre-strain and loading direction in an extruded magnesium alloy: Texture, twinning and de-twinning. Mater. Sci. Eng. A 2014, 596, 134–144. [CrossRef] 21. Sarker, D.; Friedman, J.; Chen, D.L. Influence of pre-strain on de-twinning activity in an extruded AM30 magnesium alloy. Mater. Sci. Eng. A 2014, 605, 73–79. [CrossRef] 22. Sarker, D.; Chen, D.L. Texture transformation in an extruded magnesium alloy under pressure. Mater. Sci. Eng. A 2013, 582, 63–67. [CrossRef] 23. Lei, L.; Shen, Y.; Chen, X.; Qian, L.; Lu, K. Ultrahigh Strength and High Electrical Conductivity in . Science 2004, 304, 422–426. 24. Zhang, D.D.; Yang, Q.; Guan, K.; Li, B.S.; Wang, N.; Qin, P.F.; Jiang, B.; Sun, C.; Qin, X.; Tian, Z.; et al. A high-strength low-rare-earth-alloyed magnesium alloy via traditional hot-extrusion. J. Alloy Compd. 2019, 810, 151967. [CrossRef] 25. Cepeda-Jiménez, C.M.; Molina-Aldareguia, J.M.; Pérez-Prado, M.T. Origin of the twinning to slip transition with grain size refinement, with decreasing strain rate and with increasing temperature in magnesium. Acta Mater. 2015, 88, 232–244. [CrossRef] 26. Drouven, C.; Basu, I.; Al-Samman, T.; Korte-Kerzel, S. Twinning effects in deformed and annealed magnesium–neodymium alloys. Mater. Sci. Eng. A 2015, 647, 91–104. [CrossRef] 27. Sarker, D.; Friedman, J.; Chen, D.L. Influence of pre-deformation and subsequent annealing on strain hardening and anisotropy of AM30 magnesium alloy. J. Alloy Compd. 2014, 611, 341–350. [CrossRef] 28. Kim, Y.J.; Lee, J.U.; Kim, S.H.; Kim, Y.M.; Park, S.H. Grain size effect on twinning and annealing behaviors of rolled magnesium alloy with bimodal structure. Mater. Sci. Eng. A 2019, 754, 38–45. [CrossRef] 29. Song, D.; Li, C.; Liang, N.N.; Yang, F.L.; Jiang, J.H.; Sun, J.P.; Wu, G.S.; Ma, A.B.; Ma, X.L. Simultaneously improving corrosion resistance and mechanical properties of a magnesium alloy via equal-channel angular pressing and post water annealing. Mater. Des. 2019, 166, 107621. [CrossRef] 30. Chen, W.Z.; Ma, L.M.; Chen, X.M.; Cui, G.R.; Zhang, W.C.; Wang, E.D. Microstructure evolution and mechanical performance improvement of thin ZK61 magnesium alloy sheets subjected to cumulative cold rolling and intermediate annealing. Mater. Sci. Eng. A 2018, 733, 350–360. [CrossRef] 31. Guo, F.; Pei, R.S.; Jiang, L.Y.; Zhang, D.F.; Korte-Kerzel, S.; Al-Samman, T. The role of recrystallization and grain growth in optimizing the sheet texture of magnesium alloys with calcium addition during annealing. J. Magn. Alloy 2020, 8, 252–268. [CrossRef] 32. Matsuda, M.; Ii, S.; Kawamura, Y.; Ikuhara, Y.; Nishida, M. Interaction between long period stacking order phase and deformation twin in rapidly solidified Mg97Zn1Y2 alloy. Mater. Sci. Eng. A 2004, 386, 447–452. [CrossRef] 33. Christian, J.W.; Mahajan, S. Deformation twinning. Prog. Mater Sci. 1995, 39, 1–157. [CrossRef]