Next Article in Journal
Comments on the Mechanism of Detachment in Creep of Dispersion Strengthened Alloys at High Temperatures
Previous Article in Journal
Numerical Investigation of Impact Behavior of Strut-Based Cellular Structures Designed by Spatial Voronoi Tessellation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

In Situ Observation of Tensile Deformation of Ti-22Al-25Nb Alloy and Characterization of Deformation in α2 Phase

1
State Key Laboratory of Solidification Processing, Northwestern Polytechnical University, Xi’an 710072, China
2
Defense Technologies Innovation Center of Precision Forging and Ring Rolling, School of Materials Science and Engineering, Northwestern Polytechnical University, Xi’an 710072, China
3
Shaanxi Key Laboratory of High-Performance Precision Forming Technology and Equipment, Northwestern Polytechnical University, Xi’an 710072, China
4
High Temperature Materials Research Institute, Central Iron and Steel Research Institute, Beijing 100081, China
*
Author to whom correspondence should be addressed.
Metals 2022, 12(7), 1190; https://doi.org/10.3390/met12071190
Submission received: 15 June 2022 / Revised: 6 July 2022 / Accepted: 8 July 2022 / Published: 13 July 2022
(This article belongs to the Section Metal Casting, Forming and Heat Treatment)

Abstract

:
The room temperature tensile deformation of Ti-22Al-25Nb alloy with an equiaxed α2 phase microstructure and the activated slip system of α2 particles were investigated by a combination of in situ tensile tests and electron backscatter diffraction experiments. The results demonstrate that only a few wide and long slip bands occur in the B2 matrix in the initial stage of yielding. With the tensile displacement increases, a large number of slip bands, including multiple- and cross-slip bands, appear in the B2 matrix and the distance between two adjacent slip bands decreases significantly. Meanwhile, the movement of the slip bands is hindered by the α2 particles and the B2 grain boundaries, and the slip bands appear only in a small number of the α2 particles. From the beginning of the tensile process to the final fracture, there are lots of α2 particles without slip bands. The slip bands penetrate the needle-like lamellar O phase without changing the slip direction. Compared with the α2 particles, the hindering effect of needle-like O phases on the motion of the slip bands is quite small. The microcracks nucleated at the α2/B2 phase boundaries or within the α2 particles, and microcracks propagated along the α2/B2 phase boundaries or across the α2 particles. The fracture surface shows the quasi-cleavage feature, which contains a large number of small and shallow dimples on planar facets. The analysis indicates that the plastic deformation of the alloy is mainly contributed by the B2 phase. For room temperature tensile deformation of α2 phase, there are three types of slip systems that can be activated, including the prism <a> type slip, the basal <a> type slip and the pyramidal <a+c> type slip. The prism <a> type slip is most likely to be activated, followed by the basal <a> type slip and finally the pyramidal <a+c> type slip. In addition, the critical resolved shear stress (CRSS) for the pyramidal <a+c> type slip is the highest among the three types of slip systems. Therefore, the deformation in the α2 phase is mainly contributed by the prism <a> type slip and the basal <a> type slip.

1. Introduction

Since the first generation of the Ti2AlNb-based alloy, Ti-25Al-12.5Nb (at.%), was discovered by Banerjee [1]; Ti2AlNb-based alloys have attracted a great deal of interest in the aerospace field as prospective elevated-temperature structural materials due to their high specific strength, good creep resistance and oxidation resistance [2,3,4]. As a second generation of Ti2AlNb-based alloy, the Ti-22Al-25Nb alloy has an excellent balance of room temperature and elevated-temperature properties [5,6,7]. Therefore, a lot of studies on this alloy have been conducted, such as microstructural evolution, phase diffusion and mechanical properties [8,9,10].
There are generally three phases in the Ti2AlNb-based alloy, including the B2 phase as an ordered body-centered cubic structure phase, the α2 phase as a hexagonal close-packed (base on Ti3Al) structure phase and the O phase as an orthorhombic (based on Ti2AlNb) structure phase [11]. The mechanical properties of Ti2AlNb-based alloys are significantly influenced by the microstructures due to the diversity of the phase structure. Boehlert [12] suggested that Ti-25Al-24Nb alloys with single B2 phases presented a poor tensile property (ultimate tensile strength ≤ 672 MPa, elongation ≤ 0.6%), and those with fully O phases exhibited better mechanical property (ultimate tensile strength ≤ 704 MPa, elongation ≤ 1%) than those with fully B2 microstructure. The tensile deformation of Ti-22Al-25Nb with different sizes of lamellar O microstructures was investigated by Zeng et al. [13,14,15]. The ultimate tensile strength and elongation ranged from 873–1065 MPa and 6.5–14.8%, respectively. Wang et al. [16] studied the tensile behavior of a Ti-22Al-25Nb (at.%) orthorhombic alloy with an equiaxed O/α2 + B2 microstructure and found that the alloys with this microstructure had favorable tensile properties (ultimate tensile strength ≤ 1180 MPa, elongation ≤ 8%). In addition, the deformation mechanisms of the different constituent phases have a serious influence on the tensile properties of the alloy. There are a large number of studies focused on the deformation mechanisms of the B2 phase and O phase. The B2 phase contains three different slip systems, which are {110}<111>, {112}<111> and {123}<111>, respectively [17]. The critical resolved shear stress of {123}<111> is higher than that of the other two [18]. Kamat et al. [19] proposed that the B2 phase contributes to the blunting of microcracks and improves the fracture toughness of the alloy. Banerjee et al. [20] investigated the deformation of the O phase in Ti-Al-Nb alloy and found that [100], [110], [102] and [114] are the four slip directions that can be activated at room temperature and elevated temperature. Poplille et al. [21] found that only two types of dislocations were activated in the O phase at room temperature. Zhang et al. [22] found that the O phase and the B2 phase have good slip compatibility and basal <a> type slip, prism <a> type slip and pyramidal <c+a> type slip of the O phase can be activated at room temperature. Studies on activated slip systems for the α2 phase have mainly focused on Ti3Al alloys. There are three types of dislocation for α2-Ti3Al [23]: (I) the basal <a> type with a Burgers vector of <1 1 2 ¯ 0> moving on basal (0001), prism {1 0 1 ¯ 0} or pyramidal { 20 2 ¯ 1 } planes, (II) <c> type with a Burgers vector of [0001] on second-order {1 1 2 ¯ 0} prism planes and (III) <c+a> type with a Burgers vector of <11 2 ¯ 6 > on {1 1 2 ¯ 1} or { 20 2 ¯ 1 } pyramidal planes. The absence of c-component deformation results in the generation of local stresses resulting in brittle cleavage and poor ductility. Djanarthany et al. [24] summarized the mechanical properties and deformation mechanisms of monolithic titanium aluminides based on Ti3Al. They proposed that the absence of <c> slip and the low mobility of <c+a> dislocations contribute to stress concentrations at grain boundaries and twinning deformation is not possible because it leads to a disorder incompatible with the α2-Ti3Al structure. Court et al. [25] investigated the influence of alloying additions on the mechanisms of plastic deformation of Ti3Al alloy containing 4 at.% Nb. They found that the <1 1 2 ¯ 0> and <11 2 ¯ 6 > dislocations become more active with the addition of niobium elements and the pyramidal <c+a> type slip system is activated at room temperature. Banerjee et al. [26] proposed that the increase in the Nb content of the α2 phase leads to increasing basal glide of <a> dislocations and no pyramidal <a> slip was activated during room temperature deformation. Legros et al. [27] studied the deformation process in a Ti-23.7Al-9.4Nb alloy using in situ experiments performed at room temperature. They suggested that the main difference between the α2 phase in Ti-23.7Al-9.4Nb alloy and the single-phase Ti3Al alloy is the higher homogeneity of the basal glide in the α2 phase. As mentioned above, the slip mechanism of the α2 phase changes considerably with the addition of β-stabilizing elements. However, there are relatively few studies on the α2 phase slip deformation mechanism of the Ti-22Al-25Nb alloy. Therefore, it is particularly necessary to study the activated slip systems of the α2 phase due to higher Nb content in the Ti-22Al-25Nb alloy.
This paper reports the research on the room temperature deformation of the α2 phase in Ti-22Al-25Nb (at.%) alloy. The aim of this work is to investigate the deformation characteristics of the α2 phase using in situ tensile tests and electron backscatter diffraction experiments. This information will be beneficial to designing high-performance Ti-22Al-25Nb alloy components in practical production.

2. Materials and Methods

2.1. Material

The as-received Ti2AlNb-based alloy bar was supplied by the High Temperature Materials Research Institute of Central Iron and Steel Research Institute (Beijing, China). Its nominal and chemical composition are shown in Table 1. The chemical composition was determined by inductively coupled plasma atomic emission spectrometry (ICP-AES). The bar was isothermally forged at (O + B2) phase region, and the microstructure after forging is shown in Figure 1. As can be observed in Figure 1b, the microstructure of the forged alloy is composed of the equiaxed α2/O particles, coarse lamellar O phase and B2 matrix. The equiaxed α2/O particles consist of α2 particles and rim O, which are randomly and homogeneously distributed in the B2 matrix. In order to better investigate the deformation behavior of the α2 phase in the Ti-22Al-25Nb alloy during the room temperature tensile process, the alloy was solution treated at 1020 °C for 30 min followed by water quenching. The microstructure of the as-received alloy was observed using optical microscopy (OM, OLYMPUS/PMG3, Olympus Corporation, Tokyo, Japan), and scanning electron microscope (SEM, ZEISS SUPRA55, Carl Zeiss, AG, Jena, Germany). The SEM images were analyzed using the Image-Pro Plus 6.0 software (Media Cybernetics, Rockville, MD, USA) to measure the volume fraction and size of the α2 particles. The volume fraction of α2 particles is approximately 33%. After statistical analysis, the diameter distribution of α2 particles is in accordance with the normal distribution, and the average diameter of α2 particles is 5.2 μm. The 95% confidence interval of the average diameter is 5.05 μm, 5.38 μm.

2.2. In Situ Tensile and EBSD Experiment

The in situ tensile test with electron backscatter diffraction (EBSD) was conducted in the present work to examine the deformation and activated slip systems of α2 phase in Ti-22Al-25Nb at room temperature. The specimens for the in situ tensile test and EBSD experiment were cut from the solution-treated bar using a wire electrical discharge machine (DK7732HB, Posittec Equipment, Suzhou, China), coarsely ground on with 80# metallographic sandpaper, then finely ground with 1000# metallographic sandpaper and finally electrolytic polished with an electrolyte of 10% perchloric acid, 35% n-butyl alcohol and 55% methanol in volume fraction. The electrolytic polishing process lasted for 40 s at a voltage of 40 V and a current of under 1 A at 15 °C. The geometric dimensions of the specimen for the in situ tensile test are shown in Figure 2. The gauge dimensions of in situ tensile specimens are 1.58 mm in length and 1.8 mm in width. The following experiments consist of two steps. For the first step, an area of interest (AOI) near the center of the gauge region of the specimens was selected for EBSD experiments on ZEISS Gemini 500 field emission scanning electron microscope (Carl Zeiss AG, Jena, Germany). The specimens were positioned on a 70° sample stage. The step size, testing voltage and resolution rate of the EBSD analysis were 0.2 μm, 20 kV and 97.84% respectively. The EBSD data were analyzed using HKL-Channel 5 software (HKL, Oxford, UK). For the second step, the specimens used in the EBSD experiments were etched with a corrosive of 11% HF, 33% HNO3 and 56% H2O in volume fraction and then placed in the vacuum chamber of the ZEISS Gemini 500 field emission scanning electron microscope with a servo-hydraulic loading platform (Gatan, Pleasanton, CA, USA) for room temperature in situ uniaxial tensile SEM observation. The rate of tensile displacement is 0.1 mm/min. The loading platform will hold the load during each temporary pause in displacement for SEM observation. To ensure the reliability of the experiment, three specimens were used for the in situ uniaxial tensile test.

2.3. Method for Identifying the Activated Slip System in the α2 Phase

In previous research, the method to identify different slip systems by the tensile test in α/β titanium alloy (Ti-6Al-4V) [28] and near-β titanium alloy (Ti-17) [29] using the EBSD experiment had been reported. The method was used and considered reliable by many researchers. For example, Tan et al. [30] used this method to determine the activated slip system of the α phase in TC-21. Zhang et al. [22] applied this method to identify the activated slip system of B2 and O phases in Ti-22Al-25Nb. Therefore, it is appropriate to use this method to identify the activated slip system in the α2 phase (hexagonal close-packed phase based on Ti3Al) in the Ti-22Al-25Nb alloy. The method is based on Schmid’s law, that is, for a single crystal subjected to tension and the tensile axis along a certain crystal direction, the slip system starts to activate when the critical resolved shear stress (CRSS) of the external force reaches a certain critical value in the slip direction of a certain slip plane. For an equivalent set of slip systems, they were subjected to the same CRSS. Therefore, the one with the largest Schmid factor (SF) should be the easiest to activate. In the process of calculating the SF for 4-axis coordinates, it is usually transformed to the corresponding 3-axis coordinates first for the convenience of calculation based on the Equations (1) and (2).
[ h k l ] = [ 2 h + k 3 k 3 l ( a / c ) ]
[ u v w ] = [ ( 2 u + v ) 3 / 2 3 v / 2 w c / a ]
The module of the [ h k l ] and [ u v w ] can be expressed by the Equations (3) and (4). Therefore, after normalization of Equations (1) and (2), the 4-axis coordinates in hexagonal are turned into 3-axis coordinates in cubic, which can be expressed by the Equations (5) and (6) [29].
d h k l = [ ( 2 h + k ) 2 + 3 k 2 + 3 l 2 ( a / c ) 2 ] 1 / 2
d u v w = [ 3 ( u + v / 2 ) 2 + 9 v 2 / 4 + w 2 ( c / a ) 2 ] 1 / 2
[ h k l ] = [ 2 h + k 3 k 3 l ( a / c ) ] / d h k l
[ u v w ] = [ ( 2 u + v ) 3 / 2 3 v / 2 w c / a ] / d u v w
where h, k and l are parameters of slip plane; u, v and w are parameters of slip direction; a, b and c are lattice parameters.
Tensile direction T in sample coordinate should be transformed into the corresponding vector in crystal coordinate Tc by the Equations (7) and (8) [31]. In the present research, the specimens are subjected to tensile force only on the X-axis, so tensile direction T = [ 1 0 0 ] T in the sample coordinate.
T c = g T
g = { cos φ 1 cos φ 2 sin φ 1 sin φ 2 cos ϕ sin φ 1 cos φ 2 + cos φ 1 sin φ 2 cos ϕ sin φ 2 sin ϕ cos φ 1 sin φ 2 sin φ 1 sin φ 2 cos ϕ sin φ 1 cos φ 2 + cos φ 1 sin φ 2 cos ϕ cos φ 2 sin ϕ sin φ 1 sin ϕ cos φ 1 sin ϕ cos ϕ }
where φ1, φ2 and ϕ are the crystal orientation Euler angles of the α2 phase in the area of interest (AOI) obtained by the EBSD experiment. After the transformation of the above equation, the SF of the slip system can be expressed by the Equation (9).
SF = ( n T c ) ( s T c )
where n is the normal direction of slip plane, s is the slip direction and Tc is the tensile direction in crystal coordinate.
The slip lines, which are the intersection lines of the slip plane and the specimen surface, are generated when the specimen is continuously subjected to the tensile test. To determine which slip system is most easily to be activated, the theoretical angle (theoretical θ) for every possible activated slip system needs to be calculated. When the theoretical angle is the same or close to the generated angle (which can be measured by the slip line and the tensile direction), the slip system can be considered to be activated. The theoretical θ can be calculated by the Equation (10).
cos θ = ( g e z × n ) T c
where ez is the unit normal vector of sample surface in crystal coordinate, i.e., (001) plane in sample coordinate; n is the normal direction of slip plane; Tc is the corresponding vector in crystal coordinate.

3. Results and Discussion

3.1. Microstructure and In Situ Tensile Observation

3.1.1. Microstructure

The microstructure of the alloy after solution treated at 1020 °C for 30 min is shown in Figure 3. It is a typical equiaxed microstructure, which consists of the equiaxed α2 particles, B2 matrix and a few needle-like lamellar O phases. Compared with the microstructure after forging shown in Figure 1b, the volume fraction and size of the α2 phases after solution treatment remain almost unchanged, while the content and thickness of lamellar O phases are drastically decreased and the volume fraction of the B2 matrix is significantly increased. At the same time, the equiaxed α2/O particles are transformed into equiaxed α2 particles. Based on the phase diagram of Ti-22Al-25Nb established by Raghavan [32], the solution treatment temperature of 1020 °C is in the (α2 + B2) phase region, thus leading to a significant dissolution of the O phases. The presence of a small amount of needle-like lamellar O phases is mainly due to the short solution time, which makes the O lamellae not completely dissolved into the B2 matrix.

3.1.2. In Situ Tensile Observation

The stress-displacement curve of the solution-treated Ti-22Al-25Nb alloy is shown in Figure 4. Each turning point in the curve represents a short pause for observation and the tensile process will be continued after the observation. As can be observed in Figure 4, there are four stopping points at the elastic deformation stage, and only one stopping point (marked as the first) is recorded because of no significant difference in the microstructure among the four turning points before yielding. From the second stop point to the seventh stop point, the alloy is in the yielding stage. According to the in situ tensile curve in Figure 4, the yield strength (YS) and ultimate tensile strength (UTS) of the alloy with equiaxed α2 particles are 574 MPa and 601 MPa, respectively. The overall elongation of the alloy is 8.23%. Microstructure observations of the in situ tensile process according to different stages in Figure 4 are shown in Figure 5. The microstructural morphology has undergone a series of major changes during the in situ tensile deformation. At the elastic deformation stage (Figure 5a), the microstructure had no obvious change from that after solution treatment. When the in situ tensile process reached the second and third stages (Figure 5b,c), a few wide slip bands appeared in the B2 matrix and no slip bands occurred in α2 particles, indicating that deformation preferentially occurred in the B2 matrix to relieve the stress. At that time, the distance (marked by the blue lines) between the two narrower adjacent slip bands is about 3.8 μm. As the tensile displacement increased (Figure 5d), more inhomogeneous slip bands occurred in the B2 matrix and the former slip bands became more well-defined. When the tensile process reached the fifth stage (Figure 5e), a large number of slip bands including cross-slip bands existed in the B2 matrix and the distance (marked by the blue lines) between two adjacent slip bands was reduced significantly to approximately 1.5μm. This indicates that continuous plastic deformation occurs in the B2 matrix and the appearance of cross slip bands is beneficial for releasing the stress caused by dislocation piling up [33]. Meanwhile, some of the slip bands shear the smaller α2 particles, while more slip bands were blocked by α2 particles, especially the larger ones. This indicated that larger α2 particles hinder the movement of the slip bands, while smaller α2 particles have a limited hindering effect on the movement of the slip bands. In other words, while a large number of slip bands appeared in the B2 matrix, only a few appeared in the α2 particles. It revealed that the plastic deformation of the alloy during the tensile process is mainly contributed to by the B2 phase. That is consistent with the results proposed by Boehlert [12] that the B2 matrix can enhance the plastic properties of the Ti2AlNb-based alloy compared with the O phase and the α2 phase. It is known that the B2 phase as a body-centered cubic (BCC) structure phase has three independent slip systems [17], such as {110}<111> slip systems, {112}<111> slip systems and {123}<111> slip systems. Each independent slip system has lots of equivalent slip modes. Therefore, during the tensile deformation of the alloy, there can be more activated slip systems in the B2 phase to relieve the stress concentration. It should be noted that the direction of the slip bands (marked by the red line in Figure 5b) from Figure 5b to Figure 5e appearing in the B2 matrix is basically identical. This is because the deformation occurs within a single B2 grain and only one slip system is activated for dislocation slipping in the B2 grain.
The morphology of slip bands at different locations of the specimen at the 5th stage of the stress-displacement curve is shown in Figure 6. As can be observed in Figure 6a, most of the slip bands are obstructed by α2 particles, some of them bypass the particles and only a few can penetrate the particles. It is indicated that the α2 phases and the B2 phases exist slip incompatibility, and the α2 phase is a source of the alloy strengthening. Figure 6b shows the movement of the slip bands in different B2 grains. The white dotted line represents the B2 grain boundary and solid lines with different colors (green line, red line and yellow line) represent the direction of the slip bands appearing in different B2 grains. The straight and continuous slip bands in a certain B2 grain are equidistant from each other, indicating that even plastic deformation occurs in a single B2 grain during the tensile process. In #5 B2 grain, there are two groups of slip bands in different directions (marked by yellow lines), indicating that multiple slip bands are activated during deformation. The multiple slip planes ({110}, {112}, {123}) and only one slip direction (<111>) for the B2 phase are the key factors for the appearance of multiple- and cross-slip bands [34]. In different B2 grains (marked by the numbers in Figure 6b), the slip bands have different directions, indicating that these B2 grains hold different orientations and therefore different slip systems are activated during the uniaxial tensile process. Meanwhile, the slip bands within the B2 grains are obstructed by the B2 grain boundaries, such as the grain boundary between #1 and #2 B2 grain and the grain boundary between # 2 and #5 B2 grain. For the grain boundary between #1 and #4 B2 grain, there is only a limited distance of the slip bands passing through the grain boundary. This indicates that not only the α2 particles, but also the B2 grain boundaries also have a strengthening effect on the alloy. Figure 6c is a partially enlarged view of Figure 6b. It can be observed from Figure 6c that the slip bands penetrate the needle-like lamellar O phase without changing the slip direction. Compared with the α2 particles (Figure 6a), the hindering effect of O phases on the motion of the slip bands is almost negligible. In other words, the slip transmission between the B2 phase and the needle-like O phases is much easier than that between the B2 phase and the α2 phase. This is mainly due to the multiple <c+a> type slips that were activated even in grain orientations which do not specially favor the slip [26]. Moreover, the activated slip systems in the B2 phase can induce a single and/or double system slip in the O phases [22]. Therefore, the slip compatibility between B2 and the O phases is better than that of the B2 and the α2 phases, resulting in slip bands shearing across the needle-like O phase and being hindered at the α2 particles. The crack initiation and propagation behavior during the in situ tensile test will be discussed in detail in a later subsection.

3.2. Analysis of Slip System for α2 Phase

Based on the above analysis, it is evident that α2 particles have an inhibitory effect on the movement of the slip band, which significantly affects the mechanical properties of the alloy, particularly plasticity. Therefore, in order to better understand the deformation mechanism of the alloy with the equiaxed microstructure, it is essential to analyze the activated slip systems in the α2 phase. As mentioned in Section 2.2, the specimens were subjected to EBSD experiments prior to the in situ tensile test for obtaining the orientation data (Euler angles) of the α2 particles. The inverse pole figure (IPF) map of α2 particles in the Area of Interest (AOI) is shown in Figure 7. The Euler angles (ϕ1, φ, ϕ2) were collected for each α2 particle to analyze the activated slip system. Figure 8a shows the distribution of slip bands at the fifth stage of the stress-displacement curve. The red line represents the direction of the slip bands within α2 particles and the observed θ represents the angle between the tensile direction and the slip bands. Figure 8b shows the corresponding numbering of the 26 α2 particles in Figure 8a. From Figure 8a, it is found that the planar slip appears in α2 particles and only 15 α2 particles show the slip bands. It should be noted that from this stage to the final fracture, no new slip bands are generated inside the α2 particles. Using the procedure mentioned in Section 2.3, the determination of the activated slip system was analyzed for the #10 α2 particle as an example. As shown in Table 2, there is a maximum value of Schmid Factor (SFmax) in each of the three different types of slip systems, which represents the most possible activated slip system in that type. The three most possible activated slip systems are (0001)[ 2 ¯ 11 0] for basal <a> type, (1 1 ¯ 0 0)[11 2 ¯ 0] for prism <a> type and ( 2 ¯ 111 )[ 2 1 ¯ 1 ¯ 6 ] for pyramidal <a+c> type, respectively. By matching the calculated angle (110.2°) to the observed angle (110°), (0001)[ 2 ¯ 11 0] is considered the activated slip system in the #10 α2 particle. The activated slip systems of all α2 particles in AOI can be identified by this method, as shown in Table 3. There is a small deviation between the observed angle and the calculated angle (less than 4.1°). The distribution of SF derived from EBSD data is shown in Figure 9. In the scale of Figure 9, the lighter the color represents the larger value of SF; that is, the slip system is more likely to be activated with the larger value of SF due to the CRSS being assumed to be identical to the equivalent slip systems. It can be observed that Figure 9 is in accordance with the results in Table 3. Figure 10 shows the number and the SF distribution of the activated slip systems for three different types. It is found that basal <a> type, prism <a> type and pyramidal <a+c> type slip systems are all activated during room temperature deformation. As shown in Figure 10b, three different types of slip systems are activated with a larger SF (above 0.3) and the number of basal <a> type and prism <a> type slip systems significantly increase with increasing SF. Moreover, the prism <a> type slip appears to be more easily activated than the basal <a> type slip due to the larger number of activated slip systems and smaller SF. It should be noted that the activated pyramidal <a+c> type slip system has the largest of all SF (0.496) and only one of them is activated. This suggests that the pyramidal <a+c> type slip system is the hardest slip to be activated among the three slip systems and the conditions for activating the pyramidal <a+c> type slip are severe. Therefore, the prism <a> type slip is most likely to activate for the α2 phase during the room temperature tensile process, followed by the basal <a> type slip and finally the pyramidal <a+c> type slip. A similar phenomenon is also found by Legros et al. [27]. However, the present results are not in agreement with the results in Ti3Al [23] that only <a> type slips are observed in the α2 phase during room and intermediate temperatures deformation. This is mainly because of the higher niobium (Nb) content in Ti2AlNb-based alloys than in Ti3Al alloys. Previous studies [25] suggest that a Ti3Al alloy with 4 at % Nb decreases the Peierls stresses of slip systems and increases the mobility of both <11 2 ¯ 0> and <11 2 ¯ 6 > dislocations, leading to a higher ductility than a Ti3Al alloy. Moreover, the addition of Nb elements reduces the difference in CRSS between pyramidal <a+c> type slips and prism <a> type slips [35], indicating that the anisotropy of deformation is decreased by the addition of Nb elements. For different types of slip systems, i.e., the non-equivalent slip systems, CRSS needs to be taken into account. The CRSS for basal <a> type slips is approximately three times that of the prism <a> slips in Ti3Al alloys, while the addition of Nb increases basal glide, resulting in a closer CRSS for basal <a> type slip and prism <a> type slip [26]. Thus, for basal <a> and prism <a> type slips with the same number of equivalent slip systems, it can be observed from Figure 10a that the number of activated slip systems between these two types of slip systems is quite similar. As shown in Table 2 and Figure 9, SFmax for pyramidal <a+c> type slip ( 2 ¯ 111 )[ 2 1 ¯ 1 ¯ 6 ] is 0.482, which is much larger than SFmax of 0.331 for basal <a> type slip (0001)[ 2 ¯ 11 0] in #10 α2 particle. In addition, the angle between the c-axis and the tensile direction is only about 19°, which is very close to the [1] direction, as shown in Figure 11. However, the slip system is not activated for the pyramidal <a+c> type slip. In contrast, the basal slip system is activated with the smaller SF, indicating that CRSS for the pyramidal <a+c> type slip is much higher than basal <a> type slip. Based on an experimental result, the CRSS for pyramidal <a+c> type is about seven times higher than that of prism <a> slips in a Ti3Al-Nb alloy [35], even though the addition of Nb reduces the ratio of CRSS in pyramidal <a+c> type slips and prism <a> type slips. Therefore, the deformation in the α2 phase is mainly contributed by the prism <a> type slip and the basal <a> type slip.
From Figure 8a, it is also found that the slip band in the B2 phase passed directly through a small number of α2 particles such as #22 α2 particle and #26 α2 particle without deflection. This indicates that there is a good slip compatibility between these two α2 particles and the B2 phase. The pole figures of #22 α2 particle and #26 α2 particle with the B2 phase pole figure are shown in Figure 12 and Figure 13, respectively. The {0001} plane of the #22 α2 particle is parallel to {110} plane of the B2 phase as marked by the red ellipse in Figure 12, and <11 2 ¯ 0> direction of the #22 α2 particle is also parallel to <111> direction of the B2 phase as marked by the dark green triangle in Figure 12. Thus, the #22 α2 particle maintained the specific orientation relationship, i.e., {0001}α2//{110}B2 and <11 2 ¯ 0>α2//<111>B2. Similar to the #22 α2 particle, the #26 α2 particle also maintains specific orientation relationships with the B2 phase marked by the blue ellipse and purple triangle in Figure 13. Therefore, it can be derived that the Burgers orientation relationships (BOR) between B2 and α2 phases is favorable for the slip compatibility. In other words, when the α2 particles maintain a specific orientation relationship with the B2 matrix, the hindrance of the α2 particles to the movement of the slip bands is significantly reduced.

3.3. Crack Initiation and Propagation

Figure 14 and Figure 15 show the process of crack nucleation and propagation of Ti-22Al-25Nb during the room temperature deformation. As shown in Figure 14a, a few microcracks are generated at the α2/B2 phase boundaries in the 5th stage of the stress-displacement curve. The area marked by the purple ellipse and the orange rectangle in Figure 14 is where we focus our attention on the continuous tensile process. As the tensile process continues, microcracks nucleate not only at the α2/B2 phase boundaries (marked by the red and black arrows in Figure 14) but also within the α2 phase (Figure 15a). With the further increase in tensile displacement, the microcracks propagate along the α2/B2 phase boundaries (Figure 14c) and/or across the α2 particles (Figure 15b). However, no microcracks are initiated in the B2 matrix due to the multiple activated slip systems of the B2 phase, which also indicates that the overall plastic deformation of the alloy is controlled by the B2 phase. A similar phenomenon was found by Wu et al. [36] in a study of the tensile deformation of Ti-24Al-14Nb-3V-0.5Mo alloy. Therefore, the α2/B2 phase boundaries and the interior of α2 particles are the favored sites for microcracks’ formation. According to the analysis in the previous section, the deformation of the α2 phase is mainly concentrated only in prism <a> type slip and basal <a> type slip systems during the room temperature tensile deformation. Moreover, only three equivalent slip modes are included for each type of slip system. Therefore, it is difficult for the α2 phase to release the stress through deformation due to a few activated slip systems, which leads to the nucleation of microcracks within the α2 particles. In addition, the different crystal structures of the α2 and B2 phases make the stress tend to be concentrated at the α2/B2 phase boundary, leading to the nucleation of microcracks at the α2/B2 phase boundary. As can be observed in Figure 15, there is a lot of crack propagation within the α2 particle that is not fully developed. When a crack is initiated on the basal plane and activates co-planar <a> slip, large normal stresses near the crack tip are generated. The stress can only be relieved when the <a+c> slip system is activated [23]. However, the <a+c> slip system is difficult to be activated due to the larger CRSS and specific orientation relationship with the c-axis. Therefore, the cracks continue to grow within the α2 particle. When the crack moves into the B2 matrix, the propagation of the crack is not fully developed because the B2 phase is beneficial to the blunt of the microcracks [12]. According to the above analysis, the schematic diagram of the initiation and propagation of the crack is shown in Figure 16. Based on previous discussions, the Ti-22Al-25Nb alloy with the addition of more Nb elements did not lead to the initiation of more types of slip systems in the α2 phase compared to the Ti3Al-Nb alloy. Due to seldom activated slip systems, most of the slip bands in the B2 matrix are obstructed by α2 particles, except for those that maintain a specific orientation relationship with the B2 matrix. Hall [37] pointed out that the length of the slip band is proportional to the grain size. Therefore, the presence of α2 particles leads to a decrease in the mean free path of slip (the size of B2 grain) in the B2 phase. According to the Hall–Petch Equation (11), the relationship between grain size and yield strength (YS) is expressed as follows [38]:
σ y = σ 0 + k g b d 1 / 2
where σ0 is the lattice friction stress against the movement of a single dislocation, kgb is the constant related to the material properties, σy is the yield strength, and d is the average diameter of the grain. Therefore, in the case of a two-phase Ti-22Al-25Nb alloy with equiaxed α2 particles, the yield strength of the alloy is increased due to the fact that the α2 particles reduce the mean free path of the slip (the size of B2 grain). In general, the B2 phase provides the main plastic deformation and the α2 phase is the main strengthening phase of the alloy. Therefore, in order to improve the plasticity of the two-phase alloy with equiaxed α2 particles, we can reduce the volume fraction of α2 particles and keep the α2 particles in a specific orientation relationship with the B2 matrix. On the other hand, we can increase the volume fraction of α2 particles and break the specific orientation relationship of particles with the B2 matrix to enhance the alloy strength. The fracture morphology of the alloy after the in situ room temperature tensile test is shown in Figure 17. The fracture surface shows the quasi-cleavage feature that contains a large number of small and shallow dimples on planar facets, indicating good ductility of the alloy. The shallow dimples indicate that the fracture originated at the α2/B2 phase boundary where microvoids initiated. The growth and the coalescence of the microvoids resulted in the propagation of microcracks.

4. Conclusions

In the present work, the characterization of tensile deformation of Ti-22Al-25Nb alloy with equiaxed α2 particles was investigated through the in situ tensile test. The activated slip system and deformation characteristics of α2 particles at room temperature were also examined by a combination of in situ tensile tests and EBSD experiments. The main conclusions of this work are as follows:
  • With the tensile displacement increases, a large number of slip bands including multiple- and cross-slip bands appear in the B2 matrix and the distance between two adjacent slip bands decreases significantly. Meanwhile, the movement of the slip bands is hindered by the α2 particles and the B2 grain boundaries. From the beginning of the tensile process to the final fracture, there are lots of α2 particles without slip bands.
  • The slip bands penetrate the needle-like lamellar O phase without changing the slip direction. Compared with the α2 particles, the hindering effect of needle-like O phases on the motion of the slip bands is quite small.
  • For room temperature tensile deformation of α2 phase, there are three types of slip systems that can be activated, including the prism <a> type slip, the basal <a> type slip and the pyramidal <a+c> type slip. The prism <a> type slip is most likely to be activated, followed by the basal <a> type slip and finally the pyramidal <a+c> type slip. The critical resolved shear stress (CRSS) for the pyramidal <a+c> type slip is the highest among the three types of slip systems.
  • The microcracks nucleated at the α2/B2 phase boundaries or within the α2 particles, and microcracks propagated along the α2/B2 phase boundaries or across the α2 particles. The fracture surface shows the quasi-cleavage feature, which contains a large number of small and shallow dimples on planar facets.

Author Contributions

Conceptualization, D.L.; methodology, P.Z.; formal analysis, H.M.; investigation, D.L.; resources, X.M.; writing—original draft preparation, D.L.; supervision, J.X.; project administration, W.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Program of National Key Research and Development Plan of China (2016YFB0301203). In addition, this work was supported by the Natural Science Foundation of Shaanxi province of China (2020JQ-156) and the Fundamental Research Funds for the Central Universities (3102019TS0404).

Data Availability Statement

The data are not publicly available due to the fact that these data are part of an ongoing study.

Acknowledgments

We would like to thank the Analytical & Testing Center of Northwestern Polytechnical University for the convenience of experimental tests.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Banerjee, D.; Gogia, A.; Nandy, T.; Joshi, V. A new ordered orthorhombic phase in a Ti3AlNb alloy. Acta. Metall. 1988, 36, 871–882. [Google Scholar] [CrossRef]
  2. Kumpfert, J. Intermetallic alloys based on orthorhombic titanium aluminide. Adv. Eng. Mater. 2001, 3, 851–864. [Google Scholar] [CrossRef]
  3. Mao, Y.; Hagiwara, M.; Emura, S. Creep behavior and tensile properties of Mo- and Fe-added orthorhombic Ti-22Al-11Nb-2Mo-1Fe alloy. Scripta Mater. 2007, 57, 261–264. [Google Scholar] [CrossRef]
  4. Yang, S.; Nam, S.; Hagiwara, M. Phase identification and effect of W on the microstructure and micro-hardness of Ti2AlNb-based intermetallic alloys. J. Alloys Compd. 2003, 350, 280–287. [Google Scholar] [CrossRef]
  5. Wang, W.; Zeng, W.; Xue, C.; Liang, X.; Zhang, J. 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]
  6. Xue, C.; Zeng, W.; Wang, W.; Liang, X.; Zhang, J. The enhanced tensile property by introducing bimodal size distribution of lamellar O for O+B2 Ti2AlNb based alloy. Mater. Sci. Eng. A 2013, 587, 54–60. [Google Scholar]
  7. Germann, L.; Banerjee, D.; Guédou, J.; Strudel, J. Effect of composition on the mechanical properties of newly developed Ti2AlNb-based titanium aluminide. Intermetallics 2005, 13, 920–924. [Google Scholar] [CrossRef]
  8. Zou, G.; Xie, E.; Bai, H.; Wu, A.; Wang, Q.; Ren, J. A study on transient liquid phase diffusion bonding of Ti-22Al-25Nb alloy. Mater. Sci. Eng. A 2009, 499, 101–105. [Google Scholar] [CrossRef]
  9. Leyens, C.; Gedanitz, H. Long-term oxidation of orthorhombic alloy Ti-22Al-25Nb in air between 650 and 800 °C. Scripta Mater. 1999, 41, 901–906. [Google Scholar] [CrossRef]
  10. Wei, B.; Tang, B.; Chen, X.; Xu, Q.; Zhang, S.; Kou, H.; Li, J. Precipitation behavior of orthorhombic phase in Ti-22Al-25Nb alloy during slow cooling aging treatment and its effect on tensile properties. Metals 2020, 10, 1515. [Google Scholar] [CrossRef]
  11. Boehlert, C.; Majumdar, B.; Seetharaman, V.; Miracle, D. Part I. The microstructural evolution in Ti-Al-Nb O + Bcc orthorhombic alloys. Metall. Mater. Trans. A 1999, 30, 2305–2323. [Google Scholar] [CrossRef]
  12. Boehlert, C. Part III. The tensile behavior of Ti-Al-Nb O+Bcc orthorhombic alloys. Metall. Mater. Trans. A 2001, 32, 1977–1988. [Google Scholar] [CrossRef]
  13. Wang, W.; Zeng, W.; Xue, C.; Liang, X.; Zhang, J. Quantitative analysis of the effect of heat treatment on microstructural evolution and microhardness of an isothermally forged Ti-22Al-25Nb (at.%) orthorhombic alloy. Intermetallics 2014, 45, 29–37. [Google Scholar] [CrossRef]
  14. Xue, C.; Zeng, W.; Wang, W.; Liang, X.; Zhang, J. Quantitative analysis on microstructure evolution and tensile property for the isothermally forged Ti2AlNb based alloy during heat treatment. Mater. Sci. Eng. A 2013, 573, 183–189. [Google Scholar] [CrossRef]
  15. Wang, W.; Zeng, W.; Liu, Y.; Xie, G.; Liang, X. Microstructural evolution and mechanical properties of Ti-22Al-25Nb (at.%) orthorhombic alloy with three typical microstructures. J. Mater. Eng. Perform. 2018, 27, 293–303. [Google Scholar] [CrossRef]
  16. Wang, W.; Zeng, W.; Xue, C.; Liang, X.; Zhang, J. Microstructural evolution, creep, and tensile behavior of a Ti-22Al-25Nb (at%) orthorhombic alloy. Mater. Sci. Eng. A 2014, 603, 176–184. [Google Scholar] [CrossRef]
  17. 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]
  18. Dey, S.; Suwas, S.; Fundenberger, J.; Zou, J.; Grosdidier, T.; Ray, R. Evolution of hot rolling texture in β (B2)-phase of a two-phase (O+B2) titanium-aluminide alloy. Mater. Sci. Eng. A 2008, 483–484, 551–554. [Google Scholar] [CrossRef]
  19. Kamat, S.; Gogia, A.; Banerjee, D. Effect of alloying elements and treatment on the fracture toughness of Ti-Al-Nb alloys. Acta Mater. 1998, 46, 239–251. [Google Scholar] [CrossRef]
  20. Banerjee, D.; Rowe, R.; Hall, E. Deformation of the orthorhombic phase in Ti-Al-Nb alloys. Mater. Res. Soc. 1990, 213, 285–290. [Google Scholar] [CrossRef]
  21. Popille, F.; Douin, J. The dislocation microstructure in orthorhombic O Ti2AlNb deformed between room temperature and 800 °C. Phil. Mag. A 1996, 73, 1401–1418. [Google Scholar] [CrossRef]
  22. 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]
  23. Froes, F.; Suryanarayana, C.; Eliezer, D. Synthesis, properties and applications of titanium aluminides. J. Mater. Sci. 1992, 27, 5113–5140. [Google Scholar] [CrossRef]
  24. Djanarthany, S.; Viala, J.; Bouix, J. An overview of monolithic titanium aluminides based on Ti3Al and TiAl. Mater. Chem. Phys. 2001, 72, 301–319. [Google Scholar] [CrossRef]
  25. Court, S.; Löfvander, J.; Loretto, M.; Fraser, H. The influence of temperature and alloying additions on the mechanisms of plastic deformation of Ti3Al. Phil. Mag. A 1990, 61, 109–139. [Google Scholar] [CrossRef]
  26. Banerjee, D. Deformation of the O and α2 phases in the Ti-Al-Nb system. Phil. Mag. A 1995, 72, 1559–1587. [Google Scholar] [CrossRef]
  27. Legros, M.; Caillard, D.; Couret, A. An in situ study at room temperature of deformation processes in a Ti-23.7Al-9.4Nb alloy. Intermetallics 1996, 4, 387–401. [Google Scholar] [CrossRef]
  28. Bridier, F.; Villechaise, P.; Mendez, J. Analysis of the different slip systems activated by tension in a α/β titanium alloy in relation with local crystallographic orientation. Acta Mater. 2005, 53, 555–567. [Google Scholar] [CrossRef]
  29. Zhang, S.; Zeng, W.; Zhao, Q.; Ge, L.; Zhang, M. In situ SEM study of tensile deformation of a near-β titanium alloy. Mater. Sci. Eng. A 2017, 708, 574–581. [Google Scholar] [CrossRef]
  30. Tan, C.; Sun, Q.; Xiao, L.; Zhao, Y.; Sun, J. Characterization of deformation in primary α phase and crack initiation and propagation of TC21 alloy using in-situ SEM experiments. Mater. Sci. Eng. A 2018, 725, 33–42. [Google Scholar] [CrossRef]
  31. Engler, O.; Randle, V. Introduction to Texture Analysis: Macrotexture, Microtexture, and Orientation Mapping, 2nd ed.; CRC Press: Boca Raton, FL, USA, 2009. [Google Scholar]
  32. Raghavan, V. Al-Nb-Ti (Aluminum-Niobium-Titanium). J. Phase. Equilib. Diff. 2005, 26, 360–368. [Google Scholar] [CrossRef]
  33. Majumdar, B.; Boehlert, C.; Rai, A.; Miracle, D. Structure-property relationships and deformation mechanisms in an orthorhombic based Ti-25Al-17Nb alloy. Mater. Res. Soc. 1995, 364, 1259–1264. [Google Scholar] [CrossRef]
  34. Zhang, K.; Ni, L.; Lei, Z.; Chen, Y.; Hu, X. In situ investigation of the tensile deformation of laser welded Ti2AlNb joints. Mater. Char. 2017, 123, 51–57. [Google Scholar] [CrossRef]
  35. Nakano, T.; Yanagisawa, E.; Umakoshi, Y. Effect of addition of β-phase stabilizing elements (Nb, Mo and V) on the plastic behaviour of Ti3Al single crystals with the D019 Structure. ISIJ Int. 1995, 35, 900–907. [Google Scholar] [CrossRef] [Green Version]
  36. Wu, Y.; Zhen, L.; Yang, D.; Kim, M.; Hwang, S.; Umakoshi, Y. In situ tensile deformation and fracture behavior of Ti-24Al-14Nb-3V-0.5Mo alloy with various microstructures. Intermetallics 2004, 12, 43–53. [Google Scholar] [CrossRef]
  37. Hall, E. The deformation and ageing of mild steel: III discussion of results. Proc. Phys. Soc. B. 1951, 64, 747–753. [Google Scholar] [CrossRef]
  38. Bata, V.; Pereloma, E. An alternative physical explanation of the Hall-Petch relation. Acta Mater. 2004, 52, 657–665. [Google Scholar] [CrossRef]
Figure 1. Microstructure of the alloy after forging: (a) OM; (b) SEM.
Figure 1. Microstructure of the alloy after forging: (a) OM; (b) SEM.
Metals 12 01190 g001
Figure 2. Geometric dimensions of specimen for in situ uniaxial tensile observation.
Figure 2. Geometric dimensions of specimen for in situ uniaxial tensile observation.
Metals 12 01190 g002
Figure 3. Microstructure of the alloy after solution treatment.
Figure 3. Microstructure of the alloy after solution treatment.
Metals 12 01190 g003
Figure 4. In situ room temperature tensile stress-displacement curve of the solution-treated Ti-22Al-25Nb alloy.
Figure 4. In situ room temperature tensile stress-displacement curve of the solution-treated Ti-22Al-25Nb alloy.
Metals 12 01190 g004
Figure 5. The SEM images during the in situ tensile test according to different stages in Figure 4: (a) 1st, (b) 2nd, (c) 3rd, (d) 4th and (e) 5th.
Figure 5. The SEM images during the in situ tensile test according to different stages in Figure 4: (a) 1st, (b) 2nd, (c) 3rd, (d) 4th and (e) 5th.
Metals 12 01190 g005aMetals 12 01190 g005b
Figure 6. The morphology of slip bands at different locations of the specimen: (a), (b) the 5th stage of the stress-displacement curve; (c) a higher magnification of (b).
Figure 6. The morphology of slip bands at different locations of the specimen: (a), (b) the 5th stage of the stress-displacement curve; (c) a higher magnification of (b).
Metals 12 01190 g006
Figure 7. IPF map of α2 particles in the Area of Interest (AOI).
Figure 7. IPF map of α2 particles in the Area of Interest (AOI).
Metals 12 01190 g007
Figure 8. (a) Slip bands and observed θ of α2 particles in the AOI; (b) schematic diagram of particle position.
Figure 8. (a) Slip bands and observed θ of α2 particles in the AOI; (b) schematic diagram of particle position.
Metals 12 01190 g008
Figure 9. The distribution of SF for different types of slip systems (a) basal <a> type, (b) prism <a> type and (c) pyramidal <a+c> type.
Figure 9. The distribution of SF for different types of slip systems (a) basal <a> type, (b) prism <a> type and (c) pyramidal <a+c> type.
Metals 12 01190 g009
Figure 10. The activated slip systems of α2 particles (a) number of slip systems; (b) distribution of SF in different slip systems.
Figure 10. The activated slip systems of α2 particles (a) number of slip systems; (b) distribution of SF in different slip systems.
Metals 12 01190 g010
Figure 11. The crystallographic orientation of #10 α2 grain.
Figure 11. The crystallographic orientation of #10 α2 grain.
Metals 12 01190 g011
Figure 12. (a) {110} and <111> pole figures of the B2 phase, (b) {0001} and <11 2 ¯ 0> pole figures of the #22 α2 particle.
Figure 12. (a) {110} and <111> pole figures of the B2 phase, (b) {0001} and <11 2 ¯ 0> pole figures of the #22 α2 particle.
Metals 12 01190 g012
Figure 13. (a) {110} and <111> pole figures of the B2 phase, (b) {0001} and <11 2 ¯ 0> pole figures of the #26 α2 particle.
Figure 13. (a) {110} and <111> pole figures of the B2 phase, (b) {0001} and <11 2 ¯ 0> pole figures of the #26 α2 particle.
Metals 12 01190 g013
Figure 14. The SEM images during in situ tensile test according to different stages of Figure 4: (a) 5th, (b) 6th and (c) 7th.
Figure 14. The SEM images during in situ tensile test according to different stages of Figure 4: (a) 5th, (b) 6th and (c) 7th.
Metals 12 01190 g014
Figure 15. The SEM micrographs of (a) crack nucleation within the α2 particle; (b) crack propagation shear across the α2 particle.
Figure 15. The SEM micrographs of (a) crack nucleation within the α2 particle; (b) crack propagation shear across the α2 particle.
Metals 12 01190 g015
Figure 16. The schematic diagram of the initiation and propagation of the crack [36].
Figure 16. The schematic diagram of the initiation and propagation of the crack [36].
Metals 12 01190 g016
Figure 17. The SEM micrographs of the fracture morphology: (a) side view; (b) front view.
Figure 17. The SEM micrographs of the fracture morphology: (a) side view; (b) front view.
Metals 12 01190 g017
Table 1. Chemical composition of the as-received alloy (at.%).
Table 1. Chemical composition of the as-received alloy (at.%).
NominalAnalysis
TiAlNbO (ppm)N (ppm)H (ppm)
Ti-22Al-25NbBal.22.325.7430529
Table 2. The analysis of the activated slip system for #10 α2 grain.
Table 2. The analysis of the activated slip system for #10 α2 grain.
Slip TypeSlip Plane (hkil)Slip Direction
[uvtw]
SFCalculated
θ (°)
Basal slip
<a> type
000111 2 ¯ 00.211
1 2 ¯ 1 00.121
2 ¯ 11 00.331
-SFmax
69.8/110.2
Prism slip
<a> type
1 1 ¯ 0 011 2 ¯ 00.063
-SFmax
29.1/150.9
1 0 1 ¯ 01 2 ¯ 1 00.043
01 1 ¯ 0 2 ¯ 11 00.019
Pyramidal slip
<a+c> type
11 2 ¯ 1 1 ¯ 1 ¯ 26 0.061
1 ¯ 1 ¯ 21 11 2 ¯ 6 0.405
1 2 ¯ 11 1 ¯ 2 1 ¯ 6 0.145
1 ¯ 2 1 ¯ 1 1 2 ¯ 16 0.342
2 ¯ 111 2 1 ¯ 1 ¯ 6 0.482
-SFmax
41.9/138.1
2 1 ¯ 1 ¯ 1 2 ¯ 116 0.330
Table 3. The activated slip system of α2 particles in the AIO.
Table 3. The activated slip system of α2 particles in the AIO.
No.Observed θ (°)Calculated θ (°)SFActivated Slip SystemSlip Type
15957.30.312 ( 1 1 ¯ 0 [ 11 2 ¯ 0]Prism <a> type
2-- No slip system activated
3-- No slip system activated
44547.40.455 ( 0001 ) [ 2 ¯ 11 0]Basal <a> type
55052.30.473 ( 0001 ) [ 2 ¯ 11 0]Basal <a> type
6-- No slip system activated
77070.40.463 ( 10 1 ¯ 0 ) [ 1 2 ¯ 1 0]Prism <a> type
8-- No slip system activated
9-- No slip system activated
10110110.20.331 ( 0001 ) [ 2 ¯ 11 0]Basal <a> type
117071.20.453 ( 0001 ) [ 1 2 ¯ 1 0]Basal <a> type
12-- No slip system activated
134145.10.496 ( 2 ¯ 111 ) [ 2 1 ¯ 1 ¯ 6 ]Pyramidal <a+c> type
146562.80.478 ( 0001 ) [ 2 ¯ 11 0]Basal <a> type
15-- No slip system activated
16-- No slip system activated
17-- No slip system activated
18110114.10.473 ( 10 1 ¯ 0 ) [ 1 2 ¯ 1 0]Prism <a> type
19109111.50.437 ( 1 1 ¯ 0 0 ) [ 11 2 ¯ 0]Prism <a> type
20117112.60.467 ( 10 1 ¯ 0 ) [ 1 2 ¯ 1 0]Prism <a> type
21104104.90.443 ( 10 1 ¯ 0 ) [ 1 2 ¯ 1 0]Prism <a> type
227066.30.472 ( 1 1 ¯ 0 0 ) [ 11 2 ¯ 0]Prism <a> type
23-- No slip system activated
24-- No slip system activated
257674.20.468 ( 0001 ) [ 1 2 ¯ 1 0]Basal <a> type
267069.30.458 ( 1 1 ¯ 0 0 ) [ 11 2 ¯ 0]Prism <a> type
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, D.; Zeng, W.; Zhang, P.; Ma, H.; Xu, J.; Ma, X. In Situ Observation of Tensile Deformation of Ti-22Al-25Nb Alloy and Characterization of Deformation in α2 Phase. Metals 2022, 12, 1190. https://doi.org/10.3390/met12071190

AMA Style

Li D, Zeng W, Zhang P, Ma H, Xu J, Ma X. In Situ Observation of Tensile Deformation of Ti-22Al-25Nb Alloy and Characterization of Deformation in α2 Phase. Metals. 2022; 12(7):1190. https://doi.org/10.3390/met12071190

Chicago/Turabian Style

Li, Dong, Weidong Zeng, Penghui Zhang, Haoyuan Ma, Jianwei Xu, and Xiong Ma. 2022. "In Situ Observation of Tensile Deformation of Ti-22Al-25Nb Alloy and Characterization of Deformation in α2 Phase" Metals 12, no. 7: 1190. https://doi.org/10.3390/met12071190

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