1. Introduction
Lightweight metals and alloys are highly valued in the automotive, aerospace, and medical industries for their excellent strength-to-weight ratios. These properties are ideal for applications that prioritize fuel efficiency and energy conservation. Titanium and magnesium alloys have been studied extensively for lightweight applications. Magnesium is the lightest of all structural metals, with a density that is one-quarter that of steel [
1]. Although Mg-based alloys typically exhibit lower strength and ductility compared to other structural metals, the addition of alloying elements can yield materials with an excellent strength-to-weight ratio, good fatigue and impact strengths, and relatively high thermal and electrical conductivities [
2]. Titanium (Ti) alloys also present favorable characteristics, including high strength, corrosion resistance, and reliable mechanical properties at elevated temperatures.
Aluminum is a common alloying element in both magnesium and titanium. Magnesium–aluminum alloys are valued for their low density, making them ideal for lightweight applications. Adding aluminum to magnesium increases hardness, strength, and formability while having a minimal impact on density [
3]. When aluminum dissolves in α-Mg, it induces solid solution strengthening. However, a higher aluminum content leads to the formation of the hard and brittle intermetallic phase Al
12Mg
17, which is undesirable due to its hardness, brittleness, and low thermal stability [
4,
5]. The fragility at the Mg/Al
12Mg
17 interface stems from a mismatch in crystal structures: α-Mg has a hexagonal close-packed (HCP) arrangement, whereas γ-Al
12Mg
17 has a body-centered cubic (BCC) structure [
6].
Ti-6Al-4V is a dual-phase (α + β) titanium alloy that incorporates aluminum and vanadium as substitutional elements. The α and β of this alloy have different crystal structures. The α phase Ti-6Al-4V is stable under 650 ℃ and has an HCP structure while the β phase occurs at temperatures above 900 ℃ and is of BCC structure [
7]. Compared to pure titanium, Ti-6Al-4V exhibits greater hardness, yield strength, and tensile strength, with a tensile yield strength of 900 MPa [
8]. This alloy has a superior strength-to-weight ratio, excellent corrosion resistance, and ease of machinability [
8,
9]. However, due to its high strength and limited formability, Ti-6Al-4V is typically processed at high temperatures, which can be costly.
Enhancing the strength and ductility of lightweight alloys through plastic deformation processes can improve material sustainability and durability while maintaining a low weight. Grain refinement, a common method to enhance material properties, can be achieved through severe plastic deformation techniques such as conventional rolling, cold rolling, hard-plate rolling, and ultrasonic surface rolling. Bulk forming processes like rolling produce deformation textures influenced by the strain path during processing. Rolling processes typically result in a texture similar to plain strain tension, with compression being the primary stress state.
The cold rolling of magnesium alloys leads to a high density of dislocations and deformation twins, significantly contributing to strain hardening. This process can adversely affect the anisotropic behavior due to the development of a strong basal texture [
10,
11]. However, the strain hardening from cold rolling also increases the yield strength and tensile strength of the Mg alloys. Additionally, cold rolling increases the hardness of Mg alloys, making them more susceptible to brittle cracking. Ti alloys processed through cold rolling experience significant deformation as well as a high density of dislocations. The process of strain hardening the alpha phase in titanium alloys results in a more refined microstructure, which in turn leads to a rise in dislocation density. Similar to magnesium, titanium alloys sacrifice ductility through cold rolling in exchange for increased strength and hardness. Hard plate rolling, being a cold rolling process, exhibits similar effects on both magnesium and titanium alloys to those observed in traditional cold rolling techniques.
Ultrasonic surface rolling (USRP) is similar to conventional rolling but incorporates ultrasonic vibration. This technique aims to enhance the fatigue performance of metallic materials by improving their surface properties. The plastic deformation induced by USRP results in grain refinement and increased dislocation density. In magnesium alloys, USRP creates ultra-fine grain structures on the surface, leading to a more uniform microstructure. This process improves the surface hardness by up to 50% in some Mg alloys and by more than 50% in some Ti alloys [
12]. The enhancements in fatigue performance are due to the compressive residual stress induced by USRP. In titanium alloys, USRP promotes the formation of a nanocrystalline surface layer, resulting in reduced surface roughness. The improvement in the surface hardness and fatigue life of titanium alloys after USRP can be attributed to the compressive residual stresses introduced by the process. These stresses make titanium alloys less susceptible to crack initiation and propagation under cyclic loading [
13].
Conventional rolling is a plastic deformation process in which the dislocation of polycrystalline material structure is initiated through the stress applied by two compressive rolls [
14,
15]. In plastic deformation processes, continuous dislocation induces changes in the grain morphology, ultimately leading to an augmentation in the overall grain boundary area. The increase in dislocations can also be attributed to internal structures within the grains such as secondary phases [
16]. These secondary phases deform less readily and sometimes not at all and tend to be more brittle in nature. The crystal orientation tends to correspond to the direction of the applied stress establishing the preferred orientation and generating texture. The energy stored from dislocation and increased grain boundary area is what causes the strengthening of material through plastic deformation [
17]. For this study, conventional rolling was the preferred deformation process because it minimizes the critical shear stress and is less likely to initiate stress-induced phase transformation.
Deformation mechanisms in metallic structures are often linked to the strain rate sensitivity, defined as the variation of flow stress with increasing strain rate. Key parameters influencing the strain rate sensitivity include the deformation temperature, texture or grain size, and deformation velocity. As the deformation temperature rises, the strain rate sensitivity increases more rapidly compared to room temperature conditions. Dislocations, which cause the reorientation of crystals within a metallic structure, generate texture in plastically deformed alloys. The crystallographic orientation, deformation modes, and slip systems significantly influence the strain. Complex slip systems limit dislocations, and fewer dislocation sites increase the likelihood of concentrated strain in plastically deformed regions [
18].
During severe plastic deformation, dislocations tend to move along the slip plane’s direction, and this movement also relies on the direction in which the stress is applied. The direction of slip typically corresponds with the most densely packed plane. Typically, under ambient conditions, both magnesium and titanium alloys exhibit a hexagonal close-packed (HCP) crystal structure. HCP polycrystalline materials have limited deformation systems, which has a negative impact on material ductility. HCP structures are sensitive to the crystallographic texture because of their low symmetry and anisotropic behavior [
19,
20]. This results in a significant effect on texture evolution for HCP material during plastic deformation. Material texture generated for rolling HCP crystals is heavily dictated by the level of impurities, the effect of the c/a ratio of the crystal cell, and the initial texture.
The HCP crystal structure has a limited number of slip systems, which leads to reduced ductility when subjected to severe plastic deformation. For material like Ti that have a lower than ideal (1.6333) c/a ratio, the preferred slip system during deformation is along the prismatic plane. Whereas for Mg alloys, basal slip (0001) is the preferred direction of deformation because of its high c/a ratio [
16,
21,
22]. During plastic deformation, mechanical twinning of Mg occurs on the
planes in the
direction, specifically observed under c-axis tension. The formability of Mg is limited by its insufficient slip systems necessary to meet the criteria for homogeneous plastic deformation. However, as the temperature rises above 225 °C, the formability of Mg increases due to a decrease in the critical resolved shear stress [
23].
The texture formed during the rolling process offers valuable insights into the material’s behavior and mechanical properties. This research aims to examine these characteristics using electron backscatter diffraction (EBSD) and X-ray diffraction (XRD). XRD has been widely used in various studies to analyze the lattice behavior of crystallographic materials and to determine the phase composition through diffraction intensity. In this study, XRD will help to identify whether a preferred orientation is achieved through conventional rolling and verify the presence of stress-induced secondary phases in the rolled samples as a result of plastic deformation. EBSD is an excellent tool for providing detailed crystallographic data and orientation. Within the parameters of this research, we will use EBSD to observe changes in the material’s texture and orientation induced by the rolling process. This combination of XRD and EBSD techniques will allow for a comprehensive analysis of the material’s behavior and texture following rolling.
As for the present investigation, the aim is to evaluate the effects of the rolling parameters on the structure and strain rate of lightweight materials, with the objective of developing a process map that prioritizes strain rate control over the traditionally used reduction percentage. By focusing on the strain rate control, this study seeks to preserve the ductility of lightweight materials, mitigate strain hardening effects, and promote texture development while enhancing the strength properties. The findings of this research could improve material performance in various applications, particularly in industries in which weight reduction and strength are crucial. Advancing the understanding of strain rate control addresses a specific need in materials science and engineering.
The novelty of this research lies in its approach to understanding the texture and mechanical properties of lightweight alloys, specifically Mg-Al and Ti-6Al-4V, using advanced analytical techniques like electron backscatter diffraction (EBSD) and X-ray diffraction (XRD). These techniques provide detailed insights into crystallographic texture and phase composition changes induced by the rolling process, which are critical for optimizing the material performance. By focusing on Mg-Al and Ti-6Al-4V, this research addresses materials that are important for the automotive, aerospace, and medical industries. Understanding how rolling processes affect these lightweight alloys could lead to enhanced performance in applications in which weight and strength are paramount.
Controlling the rolling temperature is crucial to avoid the formation of undesirable phases, such as the brittle γ-Al12Mg17 phase in Mg-Al alloys and the martensitic transformation in Ti-6Al-4V. Careful temperature management is essential for preserving material integrity while investigating deformation mechanisms. Enhancing the strength and ductility of lightweight alloys through plastic deformation processes can improve material sustainability and durability while maintaining a low weight. By addressing the gap in the current literature regarding the optimization of rolling processes for lightweight materials, this study contributes valuable insights to the field, offering a pathway to optimizing these alloys for high-performance applications in which weight reduction is critical.
2. Materials and Methods
This study utilized three different lightweight alloys with varying compositions. T5 commercial-grade titanium (Ti-6Al-4V) was cut from a commercially procured rolled sheet into strips measuring 25.4 × 152.4 mm with a thickness of 1.6 mm. The other two specimen types were Mg-Al alloys with aluminum contents of 6 and 9 wt%. Both Mg-6Al and Mg-9Al were cast as slabs in-house, with their chemical compositions listed in
Table 1. To avoid the formation of the brittle γ-Al
12Mg
17 phase, the Mg-Al alloys were fully solutionized. Four strips measuring 21.6 × 119.4 mm were cut from Mg-6Al, and four strips measuring 19.1 × 106.7 mm were cut from Mg-9Al, both with an approximate thickness of 3 mm. All the samples were roughly polished to remove surface roughness and sharp edges from the cutting process.
After polishing, measurements of each specimen were recorded and listed in
Table 1. Each specimen underwent warm working through conventional rolling, using work rolls with a diameter of 50 mm, a rolling speed of 1.0 m/min, and a 1:1 speed ratio between the rolls. The rolling parameters were preset for each specimen based on the initial thickness, reduction percentage per pass, and material properties. The target final thickness for each specimen was 1 mm. The initial thicknesses were 2.3 mm for Mg-6Al, 2.4 mm for Mg-9Al, and 1.6 mm for Ti-6Al-4V. The reduction percentages per pass were 15% for Mg-Al alloys and 5% for Ti-6Al-4V, set to prevent defects such as cracking or spallation [
24,
25]. These parameters determined the number of passes needed for each specimen: 5 passes for Mg-Al alloys and 9 for the Ti alloy.
Prior to each pass, the specimens were preheated for 5 min to maintain the rolling temperature. Mg-Al specimens were solution-treated by heating above the γ phase region to avoid undesirable phases, with preheating temperatures of 315 °C for Mg-6Al and 400 °C for Mg-9Al, to prevent the Al
12Mg
17 phase [
2]. The Ti-6Al-4V specimens were preheated at 650 °C to prevent martensitic phase transformation [
9]. The conventional roller used in this experiment had a rolling temperature limit of 300 °C. Consequently, the temperature during the rolling process was maintained at a constant 300 °C.
By controlling the rolling temperature, the formation of undesirable phases, such as the brittle γ-Al12Mg17 phase in Mg-Al alloys and martensitic transformation in Ti-6Al-4V, are avoided. This careful temperature management is essential for preserving the material integrity while investigating deformation mechanisms.
One of the four Mg-6Al samples was used as a dummy sample to calibrate, preheat, and prep the roller for experimentation, leaving the remaining three samples for data analysis. The sample thickness of each specimen type was measured after each pass and averaged. The final pass for each sample was a double pass in which the thickness was reduced to the final size, and then the specimen was flipped and passed through again without additional thickness reduction. After rolling, each specimen was quenched in room-temperature water for 5 s to preserve the material’s microstructure. It is important to note that there was no significant change in the chemical composition of each sample after rolling.
Table 2 shows the chemical composition of each alloy before and after rolling.
The post-rolled samples were prepared for microscopic and spectroscopic analysis. This involved a series of grinding and polishing steps, conducted using a MultiPrep System polisher (Allied High Tech Products, Inc, Compton, CA, USA). For etching, the Mg-Al samples were treated with a 2.5% nitric acid solution for 30 s, while the Ti samples were etched using a 10% HF solution for 10 s. The rolling normal direction of each sample was then examined using a Zeiss Imager M2 optical microscope.
To observe the phase composition, a Hitachi SU8000 scanning electron microscope (SEM, Tokyo, Japan) equipped with an Oxford Instruments Aztec electron backscattering diffraction (EBSD, Abingdon, UK) CMOS EBSD detector was used. EBSD data analysis was carried out with Oxford Instruments Aztec software (version 6.0).
For the analysis of the crystal direction and orientation, a Bruker D8 Discover X-ray diffraction (XRD, Billerica, MA, USA) machine was employed. The XRD measurements were taken over a range of 30° to 80° with a scanning speed of 2 s per step and a step size of 0.0172°. The resulting XRD data were processed using Bruker Diffrac Eva software (version 5.2) to accurately determine the crystal structures present in the samples.
4. Discussion
The expectation was to keep a constant percentage of reduction for each pass; however, the reduction percentage varied and spiked on the last pass of each specimen. This anomaly can be explained by the heightened dislocation density resulting from the final pass being a double pass, leading to increased strain hardening due to a flipped specimen orientation. It was also observed that increases in the percentage of reduction with each pass were correlated with higher strain rates. With each pass, the ability to stay true to the constant reduction percentage became more of a challenge because, with each pass, the strain increases, which limits dislocation. It was also concluded through calculation that, even if a constant percentage were maintained, the strain rate would still increase with each pass. Increasing the strain rate for each material increases the likelihood of failure. From this, it can be hypothesized that, if the rolling process is controlled keeping a constant strain rate, it can ensure that no damage occurs during the process.
A process map was developed in order to control the strain rate for each pass. The variables that affect the true strain rate are the temperature, roll speed, and the initial and final thickness of the material. The temperature must remain constant in order to avoid phase transformation. Therefore, for the purposes of developing the process map, the factors that may vary are the roll velocity and the material thickness or draft. If the roll speed is too fast, there is more possibility of a fracture due to the spring back effect as well as the formation of twin structure especially for the much softer Mg-Al alloys. Because of this, the velocity domain was observed within the ranges of 1–2.5 m/min. The developed process map is depicted in
Figure 6a. It can be observed that, by maintaining a constant strain rate, we can avoid failure and have more control over the rolling process.
Figure 6 also shows a breakdown of the process map in terms of the daft and the reduction percentage, respectively.
The proposed model will help control the strain rate by selecting the appropriate velocity and draft for each pass. By selecting a velocity and draft that corresponds to the desired strain rate, the strain rate can be held constant retaining more ductility and reducing the likelihood of failure during processing. This ensures that we can strengthen material through a cold working process without compromising the material’s integrity.