Next Article in Journal
Data Science in Order and Disorder of High-Entropy Materials
Previous Article in Journal
Advances in Preparation Methods and Numerical Simulation of Composites: Formation and Properties
Previous Article in Special Issue
Interfacial Microstructure and Cladding Corrosion Resistance of Stainless Steel/Carbon Steel Clad Plates at Different Rolling Reduction Ratios
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of Microstructure and Texture on Tensile Properties of an As-Rolled Ti2AlNb-Based Alloy

1
Shanghai Key Laboratory of D&A for Metal-Functional Materials, School of Materials Science & Engineering, Tongji University, Shanghai 201804, China
2
Institute of Metal Research, Chinese Academy of Sciences, Shenyang 110016, China
3
Department of Mechanical, Industrial and Mechatronics Engineering, Toronto Metropolitan University, Toronto, ON M5B 2K3, Canada
*
Authors to whom correspondence should be addressed.
Metals 2025, 15(6), 631; https://doi.org/10.3390/met15060631
Submission received: 24 April 2025 / Revised: 25 May 2025 / Accepted: 28 May 2025 / Published: 3 June 2025
(This article belongs to the Special Issue Numerical Simulation and Experimental Research of Metal Rolling)

Abstract

Ti2AlNb-based alloys are widely used in aerospace applications due to their excellent high-temperature mechanical properties. This study aims to investigate the texture, microstructural evolution, and phase transformation behavior of Ti2AlNb-based alloy sheets during heat treatment and their effects on tensile properties. During heat treatment, B2 → O phase transformation occurs at 550 °C and 650 °C, while Ostwald ripening takes place at 700 °C and 850 °C. The α2 phase undergoes spheroidization around 1000 °C due to grain boundary separation and recrystallization. Additionally, the B2, O, and α2 phases all exhibit strong textures. The B2-phase texture follows a cubic orientation ({100}<001>), rotated ~30° around the normal direction (ND). The O-phase texture consists of a strong {100}<010> rolling texture and a weaker texture component <001>//RD, influenced by the B2-phase texture, rolling deformation, and variant selection during O-phase precipitation. Each B2 grain generates four variants, forming distinct O-phase textures within the same grain. The α2-phase texture exhibits typical rolling textures, [0001]//TD, < 1 ¯ 2 1 ¯ 0>//ND, and {11 2 ¯ 0}<01 1 ¯ 0>, remaining stable after heat treatment. Tensile tests show that the rolled sheet has better ductility along the rolling direction (RD), while the transverse direction (TD) demonstrates higher yield strength (up to 1136 MPa). The anisotropy in tensile properties is mainly attributed to the O-phase texture, with minor contributions from the α2-phase and B2-phase textures. These findings provide a theoretical basis for optimizing the mechanical properties of Ti2AlNb-based alloys.

1. Introduction

In 1988, Banerjee et al. [1] first identified the orthorhombic (O) phase in the Ti-25Al-12.5Nb (at.%) alloy, expanding the application potential of the Ti3Al-Nb alloy system [2,3]. The addition of Nb enhances the mechanical properties of alloys [4,5,6,7]. Among these, Ti2AlNb-based alloys have attracted significant attention in the aerospace field due to their excellent high-temperature strength, fracture toughness, room-temperature ductility, oxidation resistance, and superplastic deformation capability [8,9,10,11,12]. Typically, Ti2AlNb-based alloys consist of primary phases: the B2 phase with a body-centered cubic (BCC) structure, the α2 phase with a hexagonal close-packed (HCP) structure, and the O phase with an orthorhombic structure. As the matrix phase, the B2 exhibits good ductility owing to its numerous slip systems. In contrast, the α2 phase has fewer slip systems and is typically brittle, while the O phase presents intermediate properties between these two phases [13,14,15,16,17].
The mechanical properties of alloys are significantly influenced by composition, microstructural morphology, and microstructural homogeneity. Over the past few decades, researchers have successfully fabricated Ti2AlNb-based alloys with various microstructures, such as equiaxed, bimodal, and lamellar, through thermal processing techniques including casting, isothermal forging, and hot rolling [18,19,20,21,22,23,24]. Previous studies from the research group have established that the fine lamellar O phase in as-cast Ti2AlNb-based alloys not only provides good plasticity and strength but also enhances fatigue resistance [25]. Forged alloys with a bimodal O phase exhibit the highest yield strength and tensile strength but the lowest ductility [2]. The rolling process has proven effective in improving the mechanical properties of as-cast alloys. To address the poor room-temperature ductility of Ti2AlNb-based alloys, Sui et al. [26] and Wang et al. [27] obtained rolled alloys with both ductility and strength via hot-rolling after spark plasma sintering (SPS) of pre-alloyed powders. Microstructural inhomogeneity may significantly affect the reliability of mechanical properties. Zavodov et al. [28] investigated the influence of deformation bands (DB) in Ti-22Al-25Nb (at.%) alloy rolled sheets on the ultimate tensile strength (UTS) in tensile tests. The results indicated that the presence of DB and its area ratio significantly reduced material strength, with single-crystal structures and unfavorable orientations being the primary contributing factors.
The texture-induced anisotropic behavior of Ti2AlNb-based alloys plays a critical role in advancing their application in aeroengine components [29]. Zhang et al. [29] demonstrated that the yield strength anisotropy of Ti-22Al-25Nb alloy, isothermally forged in the B2-phase region, is predominantly governed by the O-phase texture. The deformation temperature and phase composition significantly influence the morphology and texture development of the O phase. Dey et al. [30] observed that during deformation in the two-phase region, the O phase tends to become equiaxed, with its texture intensifying as deformation progresses. Conversely, deformation in the single β-phase region, followed by cooling, results in the formation of coarse lamellar O phases that inherit the deformation texture of the β phase. Rollett et al. [31] investigated the as-rolled Ti-22Al-23Nb (at.%) and found that the BCC phase exhibited a strong (001)<110> texture, while the orthorhombic and hexagonal phases showed weaker textures, suggesting minimal variation in yield strength across different directions. Wang et al. [32] further validated this hypothesis, reporting tensile strengths of 1110 MPa and 1157 MPa in the rolling direction (RD) and transverse direction (TD), respectively, with elongations of 5.7% and 2.4%. Liu et al. [33] investigated the influence of α-phase texture on yield strength through theoretical calculations of Schmid factors (SFs) and critical resolved shear stress (CRSS), concluding that the α phase is responsible for the observed strength anisotropy in the alloy.
Current research has discussed the influence of rolling processes on microstructure and mechanical properties, establishing correlations between texture and mechanical performance using parameters such as SF and CRSS [25,29,30,31,33]. However, few studies have examined how subsequent heat treatment affects the microstructure and texture of rolled sheets, and a detailed explanation of the formation mechanism of O-phase texture remains lacking. This study systematically investigates the evolution of texture during heat treatment in rolled Ti2AlNb sheets with complex microstructures. The formation mechanism of the O-phase texture is analyzed from two perspectives: the inheritance of B2-phase texture and the variant selection of the O phase. Furthermore, the impact of multiphase textures on the anisotropic tensile properties of the alloy is examined, offering novel insights into texture regulation for rolled Ti2AlNb-based alloys.

2. Materials and Methods

2.1. Materials

The present study focused on a Ti2AlNb-based alloy sheet with a nominal composition of Ti-22Al-23Nb (Mo, V, Si) (at.%) (Shanghai, China). The alloy sheet was produced through a series of processing steps: First, the alloy ingot was first fabricated via three cycles of vacuum arc remelting (VAR), followed by hot isostatic pressing (HIP) to homogenize the microstructure. Next, the ingot was subjected to multi-directional isothermal forging to refine the grain structure. Finally, the sheet was manufactured using pack precision hot rolling combined with intermediate heat treatment, yielding a hot-rolled sheet with a thickness of 2.5 mm.

2.2. Methods

Rectangular samples measuring 10 × 8 × 2.5 mm (Figure 1a), along with tensile specimens oriented in the rolling direction (RD) and transverse direction (TD), were cut from the fabricated hot-rolled sheet. The rectangular samples underwent heat treatment at temperatures of 550 °C, 650 °C, 700 °C, 850 °C, 900 °C, 950 °C, 1000 °C, 1150 °C, and 1200 °C, respectively. The possible phase compositions at different temperatures were referenced to the Ti-22Al-xNb (at.%) pseudo-binary phase diagram (Figure 1d). To prevent oxidation during the heat treatment process, all samples were hermetically sealed within quartz tubes. Following a 12 h heat treatment, the quartz tubes were carefully fractured to enable water quenching, thereby ensuring the preservation of the high-temperature microstructure. Schematic diagrams illustrating the samples and heat treatment procedures are presented in Figure 1a,b.
To analyze microstructural evolution and texture formation mechanisms, and to predict tensile properties during the heat treatment process from the perspectives of morphology, phase content, texture, and grain orientation, detailed characterization was conducted using scanning electron microscopy (SEM) (Carl Zeiss AG, Jena, Germany), electron backscatter diffraction (EBSD) (Oxford Instruments, Oxford, UK), X-ray diffraction (XRD) (Bruker AXS, Karlsruhe, Germany), and transmission electron microscopy (TEM) (TALOSF200XG2) (Thermo Fisher Scientific, San Diego, CA, USA). The step size for the EBSD scan was configured to 0.05 μm. It should be noted that samples for EBSD analysis were selected at 850 °C, 900 °C, 950 °C, 1000 °C, and 1150 °C, with the aim of observing the changes in the content of the O phase, the grain size of the B2 phase, and the spheroidization and dissolution of the α2 phase.
Sample preparation involved mechanical polishing and electro-polishing. The mechanical polishing utilized Struers OP-S metallographic polishing suspension, while the electro-polishing solution consisted of 6% perchloric acid, 34% n-butanol, and 60% methanol. During electro-polishing, liquid nitrogen was used to cool the electrolyte, maintaining a temperature between −40 °C and −30 °C. The DC power supply voltage was set to a maximum of 60 V with a constant current of 0.8 A, and the polishing duration was controlled to 90 s. After polishing, the samples were thoroughly rinsed with water, followed by ethanol, and then dried. For backscattered electron (BSE) observations, the polished samples were etched using Kroll’s reagent (3% HF + 5% HNO3 + 92% H2O).
The room-temperature tensile properties of the Ti2AlNb-based alloy sheet were evaluated along both the RD and TD using an Instron 5982 tensile testing machine (Instron, Norwood, MA, USA). For each orientation, three parallel specimens were prepared by sequential grinding with abrasive papers followed by mechanical polishing. All the tensile tests were conducted at a constant strain rate of 1 × 10−3 s−1. The orientation, geometry, and dimensions of tensile test specimens are presented in Figure 1c.

3. Results

3.1. Microstructure of the Rolled Sheet and Its Evolution During Heat Treatment

3.1.1. XRD and TEM

The XRD patterns in Figure 2 reveal the presence of three distinct phases in the as-rolled sheet: B2, α2, and O phases. The TEM image in Figure 2c reveals the presence of needle-like and blocky O phase (points 1 and 2), coarse plate-like α2 phase (point 3), and B2 phase intercalated between the O and α2 phases (point 4). Notably, some α2 phase is observed to be intercalated within the needle-like O phase (point 2). Furthermore, the O phase and α2 phase exhibit a well-defined crystallographic orientation relationship, specifically ( 1 ¯ 31)O//(01 1 ¯ 0)α2.
A close examination of the XRD patterns reveals that, as the heat treatment temperature increases, the peaks corresponding to the O phase almost disappear in the sample heat-treated at 900 °C. In the microstructures heat-treated between 900 °C and 1000 °C, only B2 and α2 phases remain, while a single-phase microstructure is observed at 1150 °C. Additionally, by examining the main peaks of the B2 and O phases in the as-rolled state, and after heat treatment at 550 °C, 650 °C, 700 °C, and 850 °C, it is evident that the main peak of the O phase in the as-rolled state is lower than that of the B2 phase. However, at 550 °C and 650 °C, the main peak of the O phase becomes higher than that of the B2 phase. At 700 °C, the main peak of the O phase is again slightly lower than that of the B2 phase, and by 850 °C, the main peak of the B2 phase is significantly higher than that of the O phase. These observations suggest that the B2 → O transformation occurred during heat treatment between 550 °C and 650 °C, while at 700 °C, part of the O phase dissolved back into the B2 phase, and at 850 °C, the dissolution became more pronounced.

3.1.2. SEM

The SEM images of the as-rolled and heat-treated samples are shown in Figure 3. It should be noted that in Ti2AlNb-based alloys, the B2 matrix typically appears light in BSE images, while the α2 phase (stoichiometric chemical formula: Ti3Al) exhibits near-black contrast [24,35]. The O phase (stoichiometric chemical formula: Ti2AlNb) consistently displays a gray contrast in BSE imaging [23]. In the as-rolled Ti2AlNb sheet, the B2 phase serves as the matrix, while the α2 phase appears as primary coarse plates and secondary equiaxed grains. The O phase is distributed in the form of fine laths, edge layers, and a small number of blocky structures. During the rolling deformation process, the plate-like α2 phase was significantly crushed, forming grooves and bent plate-like features. The temperature range of 550–650 °C corresponds to the stable region of the O phase. After 12 h of heat treatment, the lath-like O phase coarsened and became densely distributed within the B2 phase, while the content of the B2 phase decreased. At 700 °C, which lies within the B2 + O two-phase region, the O phase coarsened relative to the as-rolled microstructure but exhibited a sparser distribution compared to samples heat-treated at 550 °C and 650 °C. This indicates that the O phase primarily coarsened due to phase transformation in the single-phase region, while in the two-phase region, it further coarsened and partially dissolved into the B2 phase. This phenomenon became more pronounced in the 850 °C heat treatment condition. The microstructural characteristics observed via SEM fully correspond well to the earlier XRD analysis. After heat treatment at 900 °C, the lath-like O phase nearly completely dissolved. It is necessary to note that due to the high Nb content in Ti2AlNb-based alloys, the diffusion of elements is relatively slow [29]. Once the O phase nucleates, its growth kinetics remain sluggish, resulting in the O phase always being small in size throughout the rolled sheet microstructure. Therefore, high-magnification electron microscopy is required to clearly observe it. The α2 phase remained relatively stable in both the O single-phase region and the B2/β + O two-phase region but partially dissolved upon entering the B2/β + O + α2 three-phase region. At 1000 °C, some of the crushed and bent α2 phases underwent significant fragmentation and exhibited spheroidization, then completely dissolved upon further heating.

3.1.3. EBSD

EBSD characterization was performed on selected heat-treated microstructures of interest (Figure 4). In the EBSD maps, brown, light blue, and yellow represent the B2 phase, α2 phase, and O phase, respectively. In Figure 4d–f, the white lines indicate high-angle grain boundaries (HAGBs). Table 1 summarizes the phase fractions in different heat treatment conditions. The results show that as the heat treatment temperature increased, the B2-phase content gradually increased after 850 °C. Concurrently, the B2 grains exhibited progressive size enlargement when the temperature exceeded 950 °C, as shown in Figure 4f and Figure 5f,g. The grain size of B2 phase increased from approximately 14.5 μm to over 100 μm. In contrast, the O-phase content gradually decreased, with its dissolution occurring between 900 °C and 950 °C. The α2 phase remained relatively stable below 900 °C but began to dissolve beyond this temperature, with significant dissolution of secondary phases observed at 1000 °C. The primary plate-like α2 phase, located at grain boundaries, exhibited a spheroidization trend. No phase transformation between α2 phase and the O phases was observed during the heat treatment process.
Figure 5 presents the statistical results of misorientation angles between adjacent grains and grain sizes for each phase in the Ti2AlNb-based alloy. Figure 5a–e also include the proportions of low-angle grain boundaries (2–15°) and high-angle grain boundaries (>15°). In the as-rolled state, the misorientation angles between adjacent B2 grains are mainly concentrated in the low-angle range, indicating that the B2 phase was deformed during rolling, resulting in the formation of numerous substructures. As the heat treatment temperature increases, the misorientation angles between adjacent B2 grains shift towards high angles, suggesting partial recrystallization of the grains. Heat treatment at 1000 °C led to nearly complete recrystallization of the B2 phase, while at 1150 °C, the grain orientations became entirely random. The B2 grains gradually grew, but when the α2 phase was not fully dissolved, the pinning effect of second-phase particles inhibited the growth of B2 grains. At 950 °C and 1000 °C, the average grain sizes are approximately 14.5 μm and 18.6 μm (Figure 5f,g), respectively. Once the α2 phase was completely dissolved, the B2 grains grew rapidly. Due to the excessively large grain size, the number of complete B2 grains captured was limited. Based on metallographic, SEM, and EBSD data, it can be estimated that after heat treatment at 1150 °C, B2 grain sizes generally exceeded 500 μm, forming equiaxed large grains.
The misorientation angles between adjacent α2 grains are concentrated near 10°, 30°, 60°, and 90° (Figure 5h). In titanium alloys, 12 variants can form within a single β grain, with each variant having an equal probability of formation. These 12 variants exhibit specific misorientation angles between each pair [36,37]. In the Ti-6Al-4V alloy, theoretical calculations predict five types of misorientation angles: 10.5°, 60°, 60.832°, 63.262°, and 90°, with a ratio of 1:2:4:2:2 in the absence of variant selection [38,39,40]. The distribution of misorientation angles in the as-rolled alloy suggests that some form of variant selection may occur during the formation of the α2 phase, leading to the development of α2 texture. At 950 °C, the number of high-angle α2 grains significantly decreases, likely due to the dissolution of secondary equiaxed α2 grains. In the sample heat-treated at 1000 °C, the misorientation angles between adjacent α2 grains show a tendency to shift towards higher angles, indicating that some coarse plate-like α2 grains undergo recrystallization.

3.2. Texture of the B2, O, and α2 Phases

The pole figures of the B2 phase exhibit extensive regions of green, indicating the presence of texture (Figure 6). This texture is characterized by a cubic orientation rotated approximately 30° around the normal direction (ND). Following heat treatment at 900 °C, the texture was observed to weaken; subsequently, a {111}//rolling plane and <111>//ND texture emerged, likely forming during the recovery process of the B2 phase. The texture remains evident after heat treatment at 950 °C but is completely eliminated at 1150 °C as the second phase fully dissolves, leading to random orientations. Throughout the heat treatment process, an increase in temperature correlates with a gradual enhancement in recrystallization of the B2 phase, accompanied by a corresponding weakening of its texture.
The rolled sheet exhibits both coarse plate-like and secondary equiaxed α2 grains, with the coarse plate-like α2 phase demonstrating a pronounced preferred orientation. The pole figures and inverse pole figures presented in Figure 7a illustrate textures such as [0001]//TD, < 1 ¯ 2 1 ¯ 0>//ND, and {11 2 ¯ 0}<01 1 ¯ 0> rolling texture. It should be noted that the obtained texture of the α2 phase differs from that of single-phase hcp metals such as magnesium alloys and titanium alloys after rolling deformation [41,42]. However, Ti2AlNb is a multiphase alloy, and the α2 phase is different from the conventional hcp phase [43]. Its deformation mechanism selection and texture evolution during rolling are strongly influenced by the B2-phase matrix, thus resulting in a texture evolution outcome distinct from that of single-hcp-phase alloys. Similar [0001]//TD oriented texture was also reported in [44,45]. The α2 phase remained relatively stable within the B2 + O two-phase region. Upon entering the three-phase region (~940 °C), the secondary equiaxed α2 phase dissolved first, resulting in a slight increase in texture intensity (Figure 7c,d). Heat treatment does not significantly affect the preferred orientation of the α2 phase.
The consistent green color observed in the lath-like O phase indicates a preferential orientation along the <010> direction (Figure 7e). The O phase displays a pronounced {100}<010> rolling texture, alongside a weaker texture component <001>//RD, which correspond to the green and red regions in the inverse pole figure (IPF), respectively. As the O phase undergoes dissolution, it is evident that the intensity of this texture gradually diminishes. Furthermore, it can be noted that the texture of the O phase exhibits a correlation with both the cubic-like texture of the B2 phase and the texture of the α2 phase.

3.3. Tensile Properties

The rolled sheet exhibits notable anisotropy in its tensile properties. Room-temperature tensile tests (Figure 8, Table 2) indicate that the alloy exhibits greater ductility in the RD while achieving a high yield strength in the TD. The yield strength reaches an average value of 1070 MPa in the TD, while showing a lower average value of 944 MPa in the RD in Table 2.

4. Discussion

4.1. Spheroidization of the α2 Phase and Coarsening of the O Phase

After transitioning from the B2/β + O two-phase region to the B2/β + O + α2 three-phase region, the secondary equiaxed α2 phase was observed to dissolve first. However, following heat treatment at 1000 °C, a substantial amount of equiaxed or nearly equiaxed α2 phase remains in the microstructure, indicating that spheroidization occurred at this temperature. Both prior deformation and heat treatment temperature are critical factors influencing static spheroidization [46]. During the rolling process, the plate-like α2 phase underwent elongation, crushing, or deformation, resulting in the formation of numerous substructures within the plates [47,48]. At elevated temperatures, the B2 phase can readily infiltrate these regions, leading to the disintegration of the plates. In this context, the spheroidization mechanism of the α2 phase is characterized by grain boundary separation [46,49,50,51]. Additionally, the formation of numerous subgrains within the α2 plates during deformation leads to the accumulation of stored deformation energy, which subsequently drives recrystallization during heat treatment. Therefore, recrystallization occurring during heat treatment also plays a significant role in the spheroidization of the α2 phase [46,52]. Another mechanism that facilitates spheroidization in the Ti2AlNb-based alloys involves the mixed growth of the O phase within elongated, plate-like α2 phases, as illustrated in Figure 4b and case ③ of Figure 9a. During heat treatment, the O phase undergoes a transformation into the B2 phase, often occurring in the necking regions of α2 plates. This phenomenon resembles a termination migration mechanism [53], where elements diffuse from high-curvature areas to smoother regions. Figure 9a illustrates three pathways for spheroidization. In practice, the grain boundary separation mechanism typically appears during the early stages of heat treatment, with the release of stored deformation energy at subgrain boundaries driving the transformation of the α2 phase into the B2 phase. The α2 grains that undergo spheroidization via grain boundary separation and element migration exhibit similar orientations. Conversely, α2 grains undergoing spheroidization through recrystallization are characterized by high-angle grain boundaries, leading to significant misorientation. It is crucial to note that complete spheroidization still requires adjustment through the tip migration mechanism [54].
The thickness of the O-phase lamellae exhibits a strong dependence on temperature, increasing as the temperature rises. This phenomenon can be attributed to the fact that higher temperatures supply greater energy for diffusion, thereby promoting lamellar growth [55], as described by the Arrhenius equation. At temperatures of 550 °C and 650 °C, which are close to the O single-phase region, the transformation of the B2 phase into the O phase led to the coarsening of O-phase lamellae. This growth process in the lamellar O phase is governed by energy minimization and the equilibrium of atomic concentration between the lamellar O phase and the B2 matrix. A new stable state is attained by reducing the total interfacial area between O and B2 phases while preserving atomic concentration equilibrium [55,56,57]. When the temperature rises to 700 °C and 850 °C, within the B2/β + O two-phase region, some unstable and smaller O phase dissolved, while larger and more stable lamellae continue to grow, as illustrated in Figure 9b. After a heat treatment duration of 12 h, a portion of the O phase transformed into the B2 phase. Meanwhile, the remaining O phases coalesced with one another, resulting in coarsening and columnar growth of the O phase. This phenomenon can be elucidated by Ostwald ripening.

4.2. Formation Mechanism of Texture

In the rolled Ti2AlNb-based alloys, both B2 and O phases exhibit pronounced texture characteristics, with a specific crystallographic orientation relationship existing between the two phases [36,37]. During the rolling process, the α2 phase precipitates from the B2 phase at elevated temperatures while simultaneously undergoing deformation induced by rolling. It is noteworthy that heat treatment alone does not induce the formation of the {11 2 ¯ 0}<uvw> texture in the α2 phase [46,47]. Therefore, the {11 2 ¯ 0}<uvw> texture represents a hot deformation texture developed during rolling. The O phase characterized by both strong and weaker texture components precipitates from the B2 phase, with textures corresponding to those of the cubic structure of the B2 phase, specifically, {100}O//{001}B2//rolling plane and <100>O//{100}B2//ND. The strongly textured O phase is designated as O1, while that with secondary strength is referred to as O2. Both O1 and O2 primarily maintain a 90°/[100] rotation relationship, as illustrated in Figure 10b,c. For this paper, the remaining O phase was not analyzed for the time being; its content is very small, being mostly edge O phase formed by the peritectic reaction between the B2 phase and the α2 phase, which could introduce interference in the analysis.
The formation of the O-phase texture is influenced by the variant selection effect that occurs during its precipitation from the B2 phase. Under ideal conditions, a single BCC unit cell has the potential to generate 24 variants of the O phase [58]. However, due to the distinct ordered structures of both the B2 and O phases compared to typical crystal structures, only 12 variants are observed within a single BCC unit cell, with each variant manifesting with equal probability [29,59,60,61]. However, in actual production processes, due to the influences such as deformation and thermal effects, variant selection may occur, leading to a significantly higher probability of forming certain specific O-phase variants [40,58,62,63]. This ultimately results in texture developments.
The correspondence between the {110} plane and <111> direction of the B2 phase and the {001} and {110} planes of the O phase, as illustrated in Figure 10c, indicates that during precipitation, the O phase maintains a Burgers vector relationship (BOR) with the B2 phase. However, it is evident that the B2-phase grains experience deformation and rotation during rolling, resulting in non-ideal single grains. Furthermore, since a single BCC grain has six completely equivalent {110} crystal planes and four equivalent <111> crystal directions, the eight focal points of the {110} plane in Figure 10c indicate that the B2 phase with strong texture contains grains of four orientations. To analyze these orientations, this study divides the analysis area into four grain types, illustrated in Figure 10a, with pink, green, blue, and yellow denoting G1, G2, G3, and G4, respectively. Extensive pole figure analysis of the O phase reveals the variant conditions of different types of B2 grains, as summarized in Table 3. Figure 11 illustrates the analysis process of the variants in G1, showing that each type of grain could only precipitate four distinct variants; within a single type of grain, O1 and O2 are composed of two variants each. Although there is inconsistency in the precipitation relationship between various types of B2 grains and the O phase, this relationship is precisely what drives the formation of a strong texture in the precipitated O phase. Overall, rolling induces deformation in the B2 phase, contributing to the development of a robust texture. The O phase subsequently inherits this texture from the B2 phase during its formation. Furthermore, dislocations generated during the rolling process enhance the variant selection effect in the O phase [29,58,64], ultimately contributing to the establishment of a strong texture in this phase.

4.3. Relationship Between Texture and Tensile Properties

Since the microstructure of the rolled sheet contains a high proportion of textures, it is necessary to explain the influence of texture on its tensile properties. Although α2-Ti3Al, Mg, and Ti all belong to the P63/mmm space group, α2-Ti3Al exhibits a D019 ordered hexagonal structure characterized by highly ordered atomic arrangements [65,66,67]. Twinning involves coordinated shear displacement of atoms along specific crystallographic planes; however, in an ordered structure such as D019, the fixed atomic positions resist such deformation. Twinning deformation disrupts the long-range order, resulting in the formation of high-energy anti-phase boundaries or complex defects, which significantly elevates the energy barrier for twinning [43,68,69,70,71]. Furthermore, as an intermetallic compound, α2-Ti3Al features a combination of metallic and covalent bonding. The strong directional nature of covalent bonds impedes shear deformation, thereby hindering the nucleation and propagation of twins [66,72]. Consequently, compared to twinning, which would destroy the ordered structure, dislocation slip is more energetically favorable and thus more likely to occur [43,69,71]. The slip behavior of each phase in the material plays an important role in determining yield strength. From the perspective of single crystals, the yield strength of a material is governed by the CRSS and the Schmid factor (SF):
σ y = CRSS SF ,
For polycrystalline materials, it can be expressed by the yield strength (σi) and volume fraction (f(Vi)) of each orientation grain:
σ y = i = 1 σ i * f ( V i ) ,
From Equation (1), it is evident that a higher SF and a lower CRSS will result in a lower yield strength. Equivalent slip systems have the same CRSS, whereas the non-equivalent slip systems differ. Based on the method introduced by Zhang et al. [60], we evaluated the yield strength in the TD and RD of the rolled sheet. Since we only focus on the influence of texture and the rolled sheet contains a high proportion of texture, the yield strength ratio between the RD and TD directions for the B2 and α2 phases can be simply determined based on the SF. However, for the O phase, which exhibits a weaker texture component, the differences in grain orientations must also be considered when calculating the yield strength ratio.
Figure 12 and Figure 13 illustrate the distribution of Schmid factors for each phase under tensile loading in different directions, along with the corresponding inverse pole figures. Since the Schmid factors for the {123}<111> slip system in the B2 phase and the {11 2 ¯ 2}< 1 ¯ 1 ¯ 23> pyramidal slip system in the α2 phase are both very small and difficult to activate at room temperature [17,68,73], this study focuses on analyzing the potentially activatable slip systems in the B2 and α2 phases under room-temperature conditions. Moreover, due to the difficulty in fabricating O-phase single crystals, the CRSS values for this phase have not been reported. In this study, the CRSS ratios of O-phase slip systems in Ti-22Al-25Nb alloys, as reported by Zhang et al. [60], are used as a reference. The ratios for basal slip, prismatic slip, first-order pyramidal slip, and second-order pyramidal slip are 1.19:1:1.37:1.39, respectively. The CRSS ratios indicate that the activation of prismatic and basal slip systems holds a significant advantage over the two types of pyramidal slip systems, with first-order pyramidal slip being more easily activated than second-order pyramidal slip.
A comparative analysis demonstrates that in the strongly textured B2-phase microstructure, the Schmid factors for both slip systems exceed 0.45 in different loading directions while showing nearly identical values. Consequently, the yield strength ratio between the RD and TD approaches 1. This indicates that the slip deformation behavior of the B2 phase exhibits negligible dependence on loading direction, contributing minimally to the yield strength anisotropy between the RD and TD, which is beneficial for plastic deformation.
For the α2 phase, the prismatic slip system {10 1 ¯ 0}<11 2 ¯ 0> is preferentially activated in the RD direction due to its strong texture, while both basal and prismatic slip systems exhibit relatively low activation tendencies in the TD. According to Kishida et al. [74], the CRSS ratio between basal and prismatic slip systems reaches approximately 3.26, which would theoretically result in a yield strength ratio of about 0.15 between the RD and TD directions. However, experimental observations contradict this theoretical prediction due to the intrinsic brittleness of the α2 phase. The limited plastic deformability stems from two intrinsic characteristics of the α2 phase: (i) an insufficient number of available slip systems, and (ii) severely constrained dislocation mobility, as evidenced by extremely short slip lines confined within individual grain boundaries even when slip activation occurs [68]. These inherent limitations fundamentally restrict the plastic deformation capacity of the α2 phase, rendering slip system activation largely ineffective in enhancing macroscopic ductility.
The O phase can be considered to have two orientations: the strongly textured O phase (referred to as O1) and the moderately textured O phase (referred to as O2). These two orientations are in a 90°/ND rotational relationship, as shown in Figure 10c. Due to this relationship, the activation of slip systems in the TD for O2 mirrors that in the RD for O1, and vice versa, as illustrated in Figure 13b,d. Consequently, the average Schmid factor of O2 in the TD can be directly represented by the average Schmid factor of O1 in the RD for the same slip system. Based on Figure 13, it can be observed that the most easily activated slip systems for the two orientations of the O phase in the RD are the prismatic {110}<1 1 ¯ 0> and the pyramidal {041}<114> slip systems, respectively. EBSD analysis revealed that the phase fractions of O1 and O2 are 86.8% and 13.2%, respectively. Assuming the CRSS for basal slip to be x and for prismatic slip to be 1.37x (the coefficient 1.37 is the ratio of CRSS of the two slip systems), the yield strengths in the rolling direction ( σ RD ) and transverse direction ( σ TD ) can be expressed as:
σ RD = 86.8 % x 0.46 + 13.2 %   ×   1.37 x 0.47 = 2.27 x ,
σ TD = 86.8 %   ×   1.37 x 0.47 + 13.2 % x 0.47 = 2.82 x ,
where 0.46 and 0.47 are the maximum Schmid factors of x and y, respectively. From Equations (3) and (4), it can be seen that the yield strength in the TD is higher than that in the RD. Based on the actual yield strengths (σRD = 936 MPa, σTD = 1136 MPa), the calculated values of x were 412.3 and 402.8, respectively. The close agreement between these values indicates the validity of the yield strength calculation method employed in this study. A higher yield strength corresponds to lower ductility; thus, the TD exhibits inferior ductility. In the RD, prismatic slip systems with a lower CRSS are more easily activated, whereas in the TD, pyramidal slip systems with a higher CRSS are required for activation. Consequently, the textured O phase contributes to greater elongation in the RD but higher yield strength in the TD. In summary, for Ti2AlNb-based alloy sheets, the B2 phase exhibits superior deformability compared to both the α2 and O phases. The crystallographic texture induces orientation-dependent activation behavior of slip systems in the α2 and O phases, with the texture of the O phase being primarily responsible for the observed anisotropy in tensile properties.

5. Conclusions

This study analyzed the textures, the mechanisms underlying the formation mechanisms, as well as the microstructural evolution and phase transformations during heat treatment of rolled Ti2AlNb-based alloy. Additionally, the influence of textures on tensile properties of the material was investigated. The key conclusions drawn from this study are as follows:
  • In the rolled Ti2AlNb-based alloy sheet, all three phases exhibit significant textures. The BCC phase develops a near-cubic texture; the α2 phase is characterized by textures such as [0001]//TD, < 1 ¯ 2 1 ¯ 0>//ND, and {11 2 ¯ 0}<01 1 ¯ 0>; and the O phase exhibits a strong {100}<010> rolling texture. During heat treatment, the texture of the BCC phase weakens due to recovery and recrystallization, the texture of the α2 phase remains largely unaffected, while the intensity of the O-phase texture shows a weakening trend with a decreasing phase content.
  • The α2 phase exhibits a tendency to spheroidize at 1000 °C, driven by the mechanisms involving grain boundary separation and recrystallization. During heat treatment, the O phase undergoes coarsening. When heat treatment is performed in or near the single-phase region, coarsening occurs due to the transformation from B2 to O phase. In contrast, when heat treatment is conducted in the two-phase region, coarsening is governed by the Ostwald ripening mechanism.
  • The {11 2 ¯ 0}<01 1 ¯ 0> texture of the α2 phase is a hot deformation texture formed during the rolling process. The strong texture of the O phase is influenced by both the texture of the B2 phase and the variant selection effect during its formation. Each B2 grain can only precipitate four variants, with O1 and O2 in the same B2-phase grain each comprising two variants.
  • The Ti2AlNb-based alloy sheet demonstrates enhanced ductility in the RD and superior yield strength in the TD, reaching a maximum value of 1136 MPa. The anisotropy in its tensile properties is primarily attributed to the difference in CRSS values between the prismatic slip system and the first-order pyramidal slip system of the O phase.

Author Contributions

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

Funding

This study was funded by the National Natural Science Foundation of China (NSFC) (Grants 52271012 and 51871168). D.C. would like to thank the Natural Sciences and Engineering Research Council of Canada (NSERC) for their financial support.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding authors.

Conflicts of Interest

There are no conflicts of interest to declare.

Abbreviations

The following abbreviations are used in this manuscript:
RDRolling Direction
TDTransverse Direction
NDNormal Direction
CRSSCritical Resolved Shear Stress
SFSchmid Factor
IPFInverse Pole Figure
O1The O Phase with Strong Texture
O2The O Phase with Weaker Texture Component
SEMScanning Electron Microscopy
EBSDElectron Backscatter Diffraction
TEMTransmission Electron Microscopy
XRDX-ray Diffraction
HAGBsHigh-angle Grain Boundaries

References

  1. Banerjee, D.; Gogia, A.K.; Nandi, T.K.; Joshi, V.A. A new ordered orthorhombic phase in a Ti3Al-Nb alloy. Acta Metall. 1988, 36, 871–882. [Google Scholar] [CrossRef]
  2. Goyal, K.; Sardana, N. Phase stability and microstructural evolution of Ti2AlNb alloys—A review. Mater. Today Proc. 2021, 41, 951–968. [Google Scholar] [CrossRef]
  3. Hu, K.; Huang, J.; Wei, Z.; Peng, Q.; Xie, Z.; Sa, B.; Wu, B. Elastic and thermodynamic properties of the Ti2AlNb orthorhombic phase from first-principles calculations. Phys. Status Solidi B 2017, 254, 1600634. [Google Scholar] [CrossRef]
  4. Wang, K.; Xiang, H.; Xu, L.; Feng, A.; Qu, S.; Wang, H.; Chen, D. The Effect of Nb Doping on the Properties of Ti-Al Intermetallic Compounds Using First-Principles Calculations. Materials 2024, 17, 358. [Google Scholar] [CrossRef]
  5. Chen, H.; Li, X.; Chen, Z.; Zhang, R.; Ma, X.; Zheng, F.; Ma, Z.; Pan, F.; Lin, X. Investigation on electronic structures and mechanical properties of Nb–doped TiAl2 intermetallic compound. J. Alloys Compd. 2019, 780, 41–48. [Google Scholar] [CrossRef]
  6. Liu, X.; Lin, Q.; Zhang, W.; Horne, C.V.; Cha, L. Microstructure Design and Its Effect on Mechanical Properties in Gamma Titanium Aluminides. Metals 2021, 11, 1644. [Google Scholar] [CrossRef]
  7. Li, Y.; Dai, J.; Song, Y. First-principles investigation on stability and oxygen adsorption behavior of a O/B2 interface in Ti2AlNb alloys. J. Alloys Compd. 2020, 818, 152926. [Google Scholar] [CrossRef]
  8. Feng, A.H.; Chen, Q.; Wang, J.; Wang, H.; Qu, S.J.; Chen, D.L. Thermal Stability of Microstructures in Low-Density Ti2AlNb-Based Alloy Hot Rolled Plate. Acta Metall. Sin. 2023, 59, 777–786. [Google Scholar] [CrossRef]
  9. Chen, W.; Li, J.W.; Xu, L.; Lu, B. Development of Ti2AlNb Alloys: Opportunities and Challenges. Adv. Mater. Process 2014, 172, 23–27. [Google Scholar] [CrossRef]
  10. Xiang, J.M.; Mi, G.B.; Qu, S.J.; Huang, X.; Chen, Z.; Feng, A.H.; Shen, J.; Chen, D.L. Thermodynamic and microstructural study of Ti2AlNb oxides at 800 °C. Sci. Rep. 2018, 8, 12761. [Google Scholar] [CrossRef]
  11. Qu, S.; Feng, A.; Shagiev, M.R.; Xie, H.; Li, B.B.; Shen, J. Superplastic behavior of the fine-grained Ti-21Al-18Nb-1Mo-2V-0.3Si intermetallic alloy. Lett. Mater. 2018, 8, 567–571. [Google Scholar] [CrossRef]
  12. Ayadh, W.; Denand, B.; Halkoum, A.; Boulet, P.; Sennour, M.; Delfosse, J.; Sallot, P.; Esin, V.A. Effect of prior α2 phase on precipitation kinetics of O-phase in advanced Ti2AlNb alloy. Acta Mater. 2023, 252, 118930. [Google Scholar] [CrossRef]
  13. Boehlert, C.J. Part III. The tensile behavior of Ti-Al-Nb O+Bcc orthorhombic alloys. Metall. Mater. Trans. A 2001, 32, 1977–1988. [Google Scholar] [CrossRef]
  14. Boehlert, C.J.; Chen, Z.; Gutiérrez-Urrutia, I.; Llorca, J.; Pérez-Prado, M.T. In situ analysis of the tensile and tensile-creep deformation mechanisms in rolled AZ31. Acta Mater. 2012, 60, 1889–1904. [Google Scholar] [CrossRef]
  15. Jin, S.; Marthinsen, K.; Li, Y. Formation of {11-21} twin boundaries in titanium by kinking mechanism through accumulative dislocation slip. Acta Mater. 2016, 120, 403–414. [Google Scholar] [CrossRef]
  16. Nandy, T.K.; Banerjee, D. Deformation mechanisms in the O phase. Intermetallics 2000, 8, 1269–1282. [Google Scholar] [CrossRef]
  17. Zhang, P.; Zeng, W.; Jia, R.; Kou, Y.; Xu, J.; Liang, X.; Zhao, Y. Tensile behavior and deformation mechanism for Ti–22Al–25Nb alloy with lamellar O microstructures. Mater. Sci. Eng. A 2021, 803, 140492. [Google Scholar] [CrossRef]
  18. Zhang, H.; Yan, N.; Liang, H.; Liu, Y. Phase transformation and microstructure control of Ti2AlNb-based alloys: A review. J. Mater. Sci. Technol. 2021, 80, 203–216. [Google Scholar] [CrossRef]
  19. Hémery, S.; Villechaise, P. In situ EBSD investigation of deformation processes and strain partitioning in bi-modal Ti-6Al-4V using lattice rotations. Acta Mater. 2019, 171, 261–274. [Google Scholar] [CrossRef]
  20. Zheng, Y.; Zeng, W.; Li, D.; Zhao, Q.; Liang, X.; Zhang, J.; Ma, X. Fracture toughness of the bimodal size lamellar O phase microstructures in Ti-22Al-25Nb (at.%) orthorhombic alloy. J. Alloys Compd. 2017, 709, 511–518. [Google Scholar] [CrossRef]
  21. Wei, W.; Weidong, Z.; Chen, X.; Xiaobo, L.; Jianwei, Z. Designed bimodal size lamellar O microstructures in Ti2AlNb based alloy: Microstructural evolution, tensile and creep properties. Mater. Sci. Eng. A 2014, 618, 288–294. [Google Scholar] [CrossRef]
  22. Wang, W.; Zeng, W.; Xue, C.; Liang, X.; Zhang, J. Microstructure control and mechanical properties from isothermal forging and heat treatment of Ti–22Al–25Nb (at.%) orthorhombic alloy. Intermetallics 2015, 56, 79–86. [Google Scholar] [CrossRef]
  23. Zhang, H.; Li, C.; Ma, Z.; Yu, L.; Liu, Y. Effect of dual aging treatments on phase transformation and microstructure in a pre-deformed Ti2AlNb-based alloy containing O + β/B2 structures. Vacuum 2019, 164, 175–180. [Google Scholar] [CrossRef]
  24. Zheng, Y.; Zeng, W.; Zhang, P.; Ma, H.; Li, D.; Xu, J.; Ma, X. Anti-correlation between solution strengthening and precipitation strengthening in lamellar O microstructures of a Ti2AlNb based alloy. J. Alloys Compd. 2020, 847, 156470. [Google Scholar] [CrossRef]
  25. Zhang, Y.L.; Qu, S.J.; Feng, A.H.; Shen, J.; Chen, D.L. Microstructure and low cycle fatigue of a Ti2AlNb-based lightweight alloy. J. Mater. Sci. Technol. 2020, 44, 140–147. [Google Scholar] [CrossRef]
  26. Sui, X.C.; Liu, Q.; Luo, S.Y.; Liu, Y.K.; Li, Z.L.; Kang, Q.X.; Wang, G.F. Achieving high ductility of Ti2AlNb sheet without sacrificing the tensile strength without post heat treatment. Mater. Sci. Eng. A 2022, 840, 142897. [Google Scholar] [CrossRef]
  27. Wang, G.; Sui, X.; Liu, Q.; Liu, Y. Fabricating Ti2AlNb sheet with high tensile strength and good ductility by hot packed rolling the spark plasma sintered pre-alloyed powder. Mater. Sci. Eng. A 2021, 801, 140392. [Google Scholar] [CrossRef]
  28. Zavodov, A.V.; Nochovnaya, N.A.; Lyakhov, A.A.; Filonova, E.V. Effect of deformation band on the strength of a rolled plate of intermetallic titanium alloy based on Ti–22Al–25Nb system. Mater. Charact. 2021, 180, 111438. [Google Scholar] [CrossRef]
  29. Zhang, P.; Zeng, W.; Ma, H.; Zhang, F.; Xu, J.; Liang, X.; Zhao, Y. Research on tensile anisotropy of Ti-22Al-25Nb alloy isothermally forged in B2 phase region related with texture and variant selection. Mater. Charact. 2023, 201, 112899. [Google Scholar] [CrossRef]
  30. Dey, S.R.; Suwas, S.; Fundenberger, J.J.; Ray, R.K. Evolution of crystallographic texture and microstructure in the orthorhombic phase of a two-phase alloy Ti–22Al–25Nb. Intermetallics 2009, 17, 622–633. [Google Scholar] [CrossRef]
  31. Rollett, A.D.; Smith, P.R.; James, M.R. Texture and anisotropy of Ti-22Al-23Nb foil. Mater. Sci. Eng. A 1998, 257, 77–86. [Google Scholar] [CrossRef]
  32. Wang, S.; Xu, W.; Zong, Y.; Zhong, X.; Shan, D. Effect of initial microstructures on hot deformation behavior and workability of Ti2AlNb-based alloy. Metals 2018, 8, 382. [Google Scholar] [CrossRef]
  33. Liu, Z.; Zhao, Z.; Liu, J.; Wang, Q.; Guo, Z.; Wang, J.; Yang, G.; Gong, S. Anisotropy in microstructure and tensile properties of Ti–5Al–5Mo–5V–1Cr–1Fe produced via additive manufacturing. Mater. Sci. Eng. A 2020, 798, 140093. [Google Scholar] [CrossRef]
  34. Boehlert, C.J.; Majumdar, B.S.; Seetharaman, V.; Miracle, D.B. Part I. The microstructural evolution in Ti-Al-Nb O+Bcc orthorhombic alloys. Metall. Mater. Trans. A 1999, 30, 2305–2323. [Google Scholar] [CrossRef]
  35. Liu, S.; Cao, J.; Zhou, Y.; Hou, C.; Dai, S.; Huang, X. Microstructure evolution and mechanical properties of a high-strength Ti2AlNb alloy processed by large-strain two-pass hot extrusion. J. Alloys Compd. 2022, 925, 166715. [Google Scholar] [CrossRef]
  36. Zhang, Y.L.; Chen, Z.; Qu, S.J.; Feng, A.H.; Mi, G.B.; Shen, J.; Huang, X.; Chen, D.L. Microstructure and cyclic deformation behavior of a 3D-printed Ti–6Al–4V alloy. J. Alloys Compd. 2020, 825, 153971. [Google Scholar] [CrossRef]
  37. Zhang, Y.; Chen, Z.; Qu, S.; Feng, A.; Mi, G.; Shen, J.; Huang, X.; Chen, D. Multiple α sub-variants and anisotropic mechanical properties of an additively-manufactured Ti-6Al-4V alloy. J. Mater. Sci. Technol. 2021, 70, 113–124. [Google Scholar] [CrossRef]
  38. Ma, J.; Zhang, Y.; Li, J.; Wang, Z.; Wang, J. Variant selection within one β grain in laser solid formed Ti-6Al-4V alloys. Mater. Charact. 2022, 185, 111744. [Google Scholar] [CrossRef]
  39. Shi, R.; Dixit, V.; Fraser, H.L.; Wang, Y. Variant selection of grain boundary α by special prior β grain boundaries in titanium alloys. Acta Mater. 2014, 75, 156–166. [Google Scholar] [CrossRef]
  40. Li, W.; Zheng, K.; Zhao, G.; Zhang, Y.; Jin, Y. Variant Selection in the Thermally Treated Ti-6Al-4V Processed by Laser Metal Deposition. Metallogr. Microstruct. Anal. 2023, 12, 645–650. [Google Scholar] [CrossRef]
  41. Okayasu, M.; Ogawara, T.; Matsushima, M. Formability Improvement for Bending Magnesium Alloy AZ31 Sheet via Microstructural Control. J. Mater. Eng. Perform. 2024. [Google Scholar] [CrossRef]
  42. Won, J.W.; Suh, B.-C.; Kim, J.H.; Lee, D.W.; Hyun, Y.-T. Texture development of pure Ti by rolling at various temperatures and its effect on sheet formability. J. Mater. Res. Technol. 2024, 32, 2214–2223. [Google Scholar] [CrossRef]
  43. Froes, F.H.; Suryanarayana, C.; Eliezer, D. Synthesis, properties and applications of titanium aluminides. J. Mater. Sci. 1992, 27, 5113–5140. [Google Scholar] [CrossRef]
  44. Xu, Y.; Xu, L.; Ding, L.; Zhang, S.; Zhang, B.; Pan, H.; Lu, X.; Jiang, P. Microstructure and microtexture evolution of hot-rolled Ti-6321 alloy at different post-annealing temperatures. J. Alloys Compd. 2022, 902, 163842. [Google Scholar] [CrossRef]
  45. Wang, Z.; Chen, Y.; Wang, X.; Wang, J. Texture of Ti2AlNb sheet and its effect on anisotropy of tensile properties. Acta Metall. Sin. 2024, 1–13. Available online: https://www.ams.org.cn/CN/10.11900/0412.1961.2024.00074 (accessed on 27 May 2025).
  46. Zherebtsov, S.; Murzinova, M.; Salishchev, G.; Semiatin, S.L. Spheroidization of the lamellar microstructure in Ti–6Al–4V alloy during warm deformation and annealing. Acta Mater. 2011, 59, 4138–4150. [Google Scholar] [CrossRef]
  47. Dong, X.j.; Lu, S.; Zheng, H. Dynamic spheroidization kinetics behavior of Ti–6.5Al–2Zr–1Mo–1V alloy with lamellar microstructure. Trans. Nonferr. Met. Soc. China 2016, 26, 1301–1309. [Google Scholar] [CrossRef]
  48. Shell, E.B.; Semiatin, S.L. Effect of initial microstructure on plastic flow and dynamic globularization during hot working of Ti-6Al-4V. Metall. Mater. Trans. A 1999, 30, 3219–3229. [Google Scholar] [CrossRef]
  49. Semiatin, S.L.; Seetharaman, V.; Weiss, I. Flow behavior and globularization kinetics during hot working of Ti–6Al–4V with a colony alpha microstructure. Mater. Sci. Eng. A 1999, 263, 257–271. [Google Scholar] [CrossRef]
  50. Stefansson, N.; Semiatin, S.L.; Eylon, D. The kinetics of static globularization of Ti-6Al-4V. Metall. Mater. Trans. A 2002, 33, 3527–3534. [Google Scholar] [CrossRef]
  51. Xu, J.W.; Zeng, W.D.; Jia, Z.Q.; Sun, X.; Zhou, J.H. Microstructure coarsening behavior of Ti-17 alloy with equiaxed alpha during heat treatment. J. Alloys Compd. 2015, 618, 343–348. [Google Scholar] [CrossRef]
  52. Salishchev, G.A.; Zerebtsov, S.V.; Mironov, S.Y.; Semiatin, S.L. Formation of grain boundary misorientation spectrum in alpha-beta titanium alloys with lamellar structure under warm and hot working. Mater. Sci. Forum 2004, 467–470, 501–506. [Google Scholar] [CrossRef]
  53. Wang, K.; Wu, M.; Ren, Z.; Zhang, Y.; Xin, R.; Liu, Q. Static globularization and grain morphology evolution of α and β phases during annealing of hot-rolled TC21 titanium alloy. Trans. Nonferr. Met. Soc. China 2021, 31, 2664–2676. [Google Scholar] [CrossRef]
  54. Sharma, G.; Ramanujan, R.V.; Tiwari, G.P. Instability mechanisms in lamellar microstructures. Acta Mater. 2000, 48, 875–889. [Google Scholar] [CrossRef]
  55. Chen, X.; Zeng, W.; Wang, W.; Liang, X.; Zhang, J. Coarsening behavior of lamellar orthorhombic phase and its effect on tensile properties for the Ti–22Al–25Nb alloy. Mater. Sci. Eng. A 2014, 611, 320–325. [Google Scholar] [CrossRef]
  56. Lin, P.; He, Z.; Yuan, S.; Shen, J.; Huang, Y.; Liang, X. Instability of the O-phase in Ti–22Al–25Nb alloy during elevated-temperature deformation. J. Alloys Compd. 2013, 578, 96–102. [Google Scholar] [CrossRef]
  57. Wang, W.; Zeng, W.; Sun, Y.; Zhou, H.; Liang, X. Microstructure, Tensile, and Creep Behaviors of Ti-22Al-25Nb (at.%) Orthorhombic Alloy with Equiaxed Microstructure. Materials 2018, 11, 1224. [Google Scholar] [CrossRef]
  58. Zheng, Y.; Zeng, W.; Li, D.; Xu, J.; Ma, X.; Liang, X.; Zhang, J. Orthorhombic precipitate variant selection in a Ti2AlNb based alloy. Mater. Des. 2018, 158, 46–61. [Google Scholar] [CrossRef]
  59. Wu, Y.; Kou, H.; Tang, B.; Liu, D.; Li, J.; Li, J. O Phase Precipitation and Variant Selection in Ti–22Al–25Nb Alloy during the Hot Shear Spinning. Adv. Eng. Mater. 2018, 20, 1800153. [Google Scholar] [CrossRef]
  60. Zhang, P.; Zeng, W.; Zhang, F.; Ma, H.; Xu, J.; Liang, X.; Zhao, Y. In-situ investigation of tensile anisotropy mechanism in an advanced Ti2AlNb-based alloy associated with CRSS ratio and damage model. Mater. Sci. Eng. A 2024, 890, 145894. [Google Scholar] [CrossRef]
  61. Zhang, F.; Zeng, W.; Ma, H.; Zhang, P.; Liang, X. Variant selection of orthorhombic phase precipitation and its effect on colony side plates formation in Ti–22Al–25Nb alloy. J. Mater. Res. Technol. 2023, 25, 4686–4699. [Google Scholar] [CrossRef]
  62. Banerjee, R.; Bhattacharyya, D.; Collins, P.C.; Viswanathan, G.B.; Fraser, H.L. Precipitation of grain boundary α in a laser deposited compositionally graded Ti–8Al–xV alloy—An orientation microscopy study. Acta Mater. 2004, 52, 377–385. [Google Scholar] [CrossRef]
  63. Xu, S.; Gao, C.; Xiao, N.; Zhang, H.; Cui, Z. Sensitivity of heterogeneous microstructure evolution and preferential variant selection to thermomechanical conditions in a near α titanium alloy. J. Alloys Compd. 2023, 947, 169623. [Google Scholar] [CrossRef]
  64. Kato, M.; Onaka, S.; Fujii, T. Energy consideration for variant selection in precipitation on dislocations. Sci. Technol. Adv. Mater. 2001, 2, 375–380. [Google Scholar] [CrossRef]
  65. Jain, A.; Duygulu, O.; Brown, D.W.; Tomé, C.N.; Agnew, S.R. Grain size effects on the tensile properties and deformation mechanisms of a magnesium alloy, AZ31B, sheet. Mater. Sci. Eng. A 2008, 486, 545–555. [Google Scholar] [CrossRef]
  66. Liu, Z.G.; Frommeyer, G.; Kreuss, M. Field ion microscopy and atom probe investigations on two phase intermetallic TiAlTi3Al alloy. Scr. Metall. Mater. 1991, 25, 1205–1210. [Google Scholar] [CrossRef]
  67. Zhang, H.; Wei, B.; Ou, X.; Ni, S.; Yan, H.; Song, M. Atomic-level study of {10 1 ¯ 1} deformation twinning in pure Ti and Ti-5at.% Al alloy. Int. J. Plast. 2022, 153, 103273. [Google Scholar] [CrossRef]
  68. Shao, B.; Shan, D.; Guo, B.; Zong, Y. Plastic deformation mechanism and interaction of B2, α2, and O phases in Ti-22Al-25Nb alloy at room temperature. Int. J. Plast. 2019, 113, 18–34. [Google Scholar] [CrossRef]
  69. Imayev, R.; Gabdullin, N.; Salishchev, G. Effect of grain size on superplasticity of an intermetallic Ti3Al compound. Intermetallics 1997, 5, 229–236. [Google Scholar] [CrossRef]
  70. Song, L.; Wang, L.; Oehring, M.; Hu, X.; Appel, F.; Lorenz, U.; Pyczak, F.; Zhang, T. Evidence for deformation twinning of the DO192 phase in a high Nb containing TiAl alloy. Intermetallics 2019, 109, 91–96. [Google Scholar] [CrossRef]
  71. Inui, H.; Toda, Y.; Yamaguchi, M. Plastic deformation of single crystals of a DO19 compound with an off-stoichiometric composition (Ti-36·5 at.% Al) at room temperature. Philos. Mag. A 1993, 67, 1315–1332. [Google Scholar] [CrossRef]
  72. Zhang, C.; Hou, H.; Zhao, Y.; Yang, X.; Guo, Y. First-principles study on structural, elastic and thermal properties of γ-TiAl and α2-Ti3Al phases in TiAl-based alloy under high pressure. Int. J. Mod. Phys. B 2017, 31, 1750079. [Google Scholar] [CrossRef]
  73. Court, S.A.; Löfvander, J.P.A.; Loretto, M.H.; Fraser, H.L. The influence of temperature and alloying additions on the mechanisms of plastic deformation of Ti3Al. Philos. Mag. A 1990, 61, 109–139. [Google Scholar] [CrossRef]
  74. Kishida, K.; Takahama, Y.; Inui, H. Deformation twinning in single crystals of a DO19 compound with an off-stoichiometric composition (Ti–36.5at.%Al). Acta Mater. 2004, 52, 4941–4952. [Google Scholar] [CrossRef]
Figure 1. (a) Image of the heat-treated Ti2AlNb-based alloy samples; (b) schematic diagram of the heat treatment procedure; (c) orientation, geometry, and dimensions of tensile test specimens; (d) phase diagram of Ti-22Al-27Nb vertical section Reprinted with permission from ref. [34]. 1999, Springer Nature.
Figure 1. (a) Image of the heat-treated Ti2AlNb-based alloy samples; (b) schematic diagram of the heat treatment procedure; (c) orientation, geometry, and dimensions of tensile test specimens; (d) phase diagram of Ti-22Al-27Nb vertical section Reprinted with permission from ref. [34]. 1999, Springer Nature.
Metals 15 00631 g001
Figure 2. (a) XRD patterns of as-rolled Ti2AlNb-based alloy and 550–1000 °C heat-treated alloy samples; (b) XRD patterns of Ti2AlNb-based alloy at 1150 °C; (c) TEM images of as-rolled Ti2AlNb-based alloy, along with the selected area diffraction patterns.
Figure 2. (a) XRD patterns of as-rolled Ti2AlNb-based alloy and 550–1000 °C heat-treated alloy samples; (b) XRD patterns of Ti2AlNb-based alloy at 1150 °C; (c) TEM images of as-rolled Ti2AlNb-based alloy, along with the selected area diffraction patterns.
Metals 15 00631 g002
Figure 3. SEM-BSE (compositional mode) images of Ti2AlNb-based alloy: (a) as-rolled; heat-treated at (b) 550 °C; (c) 650 °C; (d) 700 °C; (e) 850 °C; (f) 900 °C; (g) 950 °C; (h) 1000 °C; (i) 1150 °C.
Figure 3. SEM-BSE (compositional mode) images of Ti2AlNb-based alloy: (a) as-rolled; heat-treated at (b) 550 °C; (c) 650 °C; (d) 700 °C; (e) 850 °C; (f) 900 °C; (g) 950 °C; (h) 1000 °C; (i) 1150 °C.
Metals 15 00631 g003
Figure 4. EBSD phase maps of Ti2AlNb-based alloy (where the white lines indicate high-angle grain boundaries (>15°)): (a) as-rolled; heat-treated at (b) 850 °C; (c) 900 °C; (d) 950 °C; (e) 1000 °C; (f) 1150 °C.
Figure 4. EBSD phase maps of Ti2AlNb-based alloy (where the white lines indicate high-angle grain boundaries (>15°)): (a) as-rolled; heat-treated at (b) 850 °C; (c) 900 °C; (d) 950 °C; (e) 1000 °C; (f) 1150 °C.
Metals 15 00631 g004
Figure 5. Grain misorientation analysis of various phases in Ti2AlNb-based alloy: (a) as-rolled B2 phase; (b) B2 phase at 900 °C; (c) B2 phase at 950 °C; (d) B2 phase at 1000 °C; (e) B2 phase at 1150 °C; (f) grain size of B2 phase at 950 °C; (g) grain size of B2 phase at 1000 °C; (h) as-rolled α2 phase; (i) α2 phase at 1000 °C.
Figure 5. Grain misorientation analysis of various phases in Ti2AlNb-based alloy: (a) as-rolled B2 phase; (b) B2 phase at 900 °C; (c) B2 phase at 950 °C; (d) B2 phase at 1000 °C; (e) B2 phase at 1150 °C; (f) grain size of B2 phase at 950 °C; (g) grain size of B2 phase at 1000 °C; (h) as-rolled α2 phase; (i) α2 phase at 1000 °C.
Metals 15 00631 g005
Figure 6. EBSD orientation map of the B2 phases along with their pole figures and inverse pole figures (IPFs) during heat treatment: (a) as-rolled; (b) 850 °C; (c) 900 °C; (d) 950 °C; (e) 1000 °C; (f) 1150 °C.
Figure 6. EBSD orientation map of the B2 phases along with their pole figures and inverse pole figures (IPFs) during heat treatment: (a) as-rolled; (b) 850 °C; (c) 900 °C; (d) 950 °C; (e) 1000 °C; (f) 1150 °C.
Metals 15 00631 g006
Figure 7. EBSD orientation map of the α2 phases and O phase along with their pole figures and inverse pole figures (IPFs) during heat treatment. (ad) IPF map of the α2 phase in the as-rolled state and after heat treatment at 900 °C, 950 °C, and 1000 °C; (e,f) IPF map of the O phase in the as-rolled state and after heat treatment at 900 °C, the red dashed circle pointed by the arrow in Figure (e) encloses the weaker texture component.
Figure 7. EBSD orientation map of the α2 phases and O phase along with their pole figures and inverse pole figures (IPFs) during heat treatment. (ad) IPF map of the α2 phase in the as-rolled state and after heat treatment at 900 °C, 950 °C, and 1000 °C; (e,f) IPF map of the O phase in the as-rolled state and after heat treatment at 900 °C, the red dashed circle pointed by the arrow in Figure (e) encloses the weaker texture component.
Metals 15 00631 g007
Figure 8. Engineering stress–strain curves of the as-rolled Ti2AlNb-based alloy sheet in the rolling direction (RD-1) and transverse direction (TD-1).
Figure 8. Engineering stress–strain curves of the as-rolled Ti2AlNb-based alloy sheet in the rolling direction (RD-1) and transverse direction (TD-1).
Metals 15 00631 g008
Figure 9. (a) Schematic illustration of α2 phase spheroidization, the yellow dashed box in the SEM image highlights the manifestation of α2 phase spheroidization; (b) schematic illustration of O phase coarsening.
Figure 9. (a) Schematic illustration of α2 phase spheroidization, the yellow dashed box in the SEM image highlights the manifestation of α2 phase spheroidization; (b) schematic illustration of O phase coarsening.
Metals 15 00631 g009
Figure 10. O-phase variant selection analysis in the as-rolled Ti2AlNb-based alloy. (a) Distribution of B2-phase grains with four orientation types, represented by the pink, green, blue, and yellow regions corresponding to G1, G2, G3, and G4, respectively; (b) IPF of O phase with strong texture and weaker texture components; (c) the corresponding pole figures along with a schematic diagram showing the orientation relationship between the textured B2 phase and O phase.
Figure 10. O-phase variant selection analysis in the as-rolled Ti2AlNb-based alloy. (a) Distribution of B2-phase grains with four orientation types, represented by the pink, green, blue, and yellow regions corresponding to G1, G2, G3, and G4, respectively; (b) IPF of O phase with strong texture and weaker texture components; (c) the corresponding pole figures along with a schematic diagram showing the orientation relationship between the textured B2 phase and O phase.
Metals 15 00631 g010
Figure 11. Analysis of O-phase variants in as-rolled Ti2AlNb-based alloy plates: example of G1 analysis.
Figure 11. Analysis of O-phase variants in as-rolled Ti2AlNb-based alloy plates: example of G1 analysis.
Metals 15 00631 g011
Figure 12. Schmid factor analysis of B2 and α2 phases in the as-rolled Ti2AlNb-based alloy sheets. (a) Statistical analysis of Schmid factors for room-temperature activated slip systems in the B2 phase; (b) distribution of Schmid factors for slip systems in the textured B2 phase in inverse pole figures; (c) statistical analysis of Schmid factors for room-temperature activated slip systems in the α2 phase; (d) distribution of Schmid factors for slip systems in the textured α2 phase in inverse pole figures.
Figure 12. Schmid factor analysis of B2 and α2 phases in the as-rolled Ti2AlNb-based alloy sheets. (a) Statistical analysis of Schmid factors for room-temperature activated slip systems in the B2 phase; (b) distribution of Schmid factors for slip systems in the textured B2 phase in inverse pole figures; (c) statistical analysis of Schmid factors for room-temperature activated slip systems in the α2 phase; (d) distribution of Schmid factors for slip systems in the textured α2 phase in inverse pole figures.
Metals 15 00631 g012
Figure 13. (a) Statistical distribution of Schmid factors for O1-phase basal slip systems; (b) IPF mapping of Schmid factors for O1-phase basal slip systems; (c) statistical distribution of Schmid factors for O1-phase prismatic slip systems; (d) IPF mapping of Schmid factors for O1-phase prismatic slip systems; (e) first-order <a + c> equivalent slip systems in O1 phase; (f) second-order <a + c> equivalent slip systems in O1 phase.
Figure 13. (a) Statistical distribution of Schmid factors for O1-phase basal slip systems; (b) IPF mapping of Schmid factors for O1-phase basal slip systems; (c) statistical distribution of Schmid factors for O1-phase prismatic slip systems; (d) IPF mapping of Schmid factors for O1-phase prismatic slip systems; (e) first-order <a + c> equivalent slip systems in O1 phase; (f) second-order <a + c> equivalent slip systems in O1 phase.
Metals 15 00631 g013
Table 1. Phase content of Ti2AlNb-based alloy after heat treatment at different temperatures.
Table 1. Phase content of Ti2AlNb-based alloy after heat treatment at different temperatures.
PhaseAs-Rolled850 °C900 °C950 °C1000 °C1150 °C
B235.8%45.5%70.3%86.7%91.3%100%
O41.3%19%2.6%000
α218.7%22.8%22.8%12.5%7%0
Table 2. Room-temperature tensile properties of three parallel samples of Ti2AlNb-based alloy: yield strength (YS), ultimate tensile strength (UTS), and elongation (El).
Table 2. Room-temperature tensile properties of three parallel samples of Ti2AlNb-based alloy: yield strength (YS), ultimate tensile strength (UTS), and elongation (El).
Strain Rate, s−1YS, MPaUTS, MPaEl%
RD-11 × 10−393610323.1
RD-21 × 10−395310634.0
RD-31 × 10−394210442.9
TD-11 × 10−3113611941.8
TD-21 × 10−3101811501.5
TD-31 × 10−3105511941.5
Table 3. Precipitation of O-phase variants in B2 grains with different orientations in the as-rolled Ti2AlNb-based alloy sheets.
Table 3. Precipitation of O-phase variants in B2 grains with different orientations in the as-rolled Ti2AlNb-based alloy sheets.
O1O2
G1V1( 1 ¯ 10)B2//(001)O, [11 1 ¯ ]B2//[1 1 ¯ 0]OV3(110)B2//(001)O, [ 1 ¯ 11]B2//[1 1 ¯ 0]O
V2( 1 ¯ 10)B2//(001)O, [111]B2//[110]OV4(110)B2//(001)O, [1 1 ¯ 1]B2//[110]O
G2V1( 1 ¯ 10)B2//(001)O, [111]B2//[1 1 ¯ 0]OV3(110)B2//(001)O, [ 1 ¯ 11]B2//[110]O
V2( 1 ¯ 10)B2//(001)O, [11 1 ¯ ]B2//[110]OV4(110)B2//(001)O, [1 1 ¯ 1]B2//[1 1 ¯ 0]O
G3V1(110)B2//(001)O, [1 1 ¯ 1]B2//[110]OV3( 1 ¯ 10)B2//(001)O, [111]B2//[1 1 ¯ 0]O
V2(110)B2//(001)O, [ 1 ¯ 11]B2//[1 1 ¯ 0]OV4( 1 ¯ 10)B2//(001)O, [11 1 ¯ ]B2//[110]O
G4V1(110)B2//(001)O, [ 1 ¯ 11]B2//[110]OV3( 1 ¯ 10)B2//(001)O, [111]B2//[110]O
V2(110)B2//(001)O, [1 1 ¯ 1]B2//[1 1 ¯ 0]OV4( 1 ¯ 10)B2//(001)O, [11 1 ¯ ]B2//[1 1 ¯ 0]O
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

Jing, C.; Qu, S.; Feng, A.; Wang, H.; Chen, D. Influence of Microstructure and Texture on Tensile Properties of an As-Rolled Ti2AlNb-Based Alloy. Metals 2025, 15, 631. https://doi.org/10.3390/met15060631

AMA Style

Jing C, Qu S, Feng A, Wang H, Chen D. Influence of Microstructure and Texture on Tensile Properties of an As-Rolled Ti2AlNb-Based Alloy. Metals. 2025; 15(6):631. https://doi.org/10.3390/met15060631

Chicago/Turabian Style

Jing, Caihong, Shoujiang Qu, Aihan Feng, Hao Wang, and Daolun Chen. 2025. "Influence of Microstructure and Texture on Tensile Properties of an As-Rolled Ti2AlNb-Based Alloy" Metals 15, no. 6: 631. https://doi.org/10.3390/met15060631

APA Style

Jing, C., Qu, S., Feng, A., Wang, H., & Chen, D. (2025). Influence of Microstructure and Texture on Tensile Properties of an As-Rolled Ti2AlNb-Based Alloy. Metals, 15(6), 631. https://doi.org/10.3390/met15060631

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop