Next Article in Journal
A Review of Microstructural Evolution and Modelling of Aluminium Alloys under Hot Forming Conditions
Previous Article in Journal
Co-Conversion Mechanisms of Boron and Iron Components of Ludwigite Ore during Reductive Soda-Ash Roasting
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Precipitation Behavior of Orthorhombic Phase in Ti-22Al-25Nb Alloy during Slow Cooling Aging Treatment and Its Effect on Tensile Properties

1
State Key Laboratory of Solidification Processing, Northwestern Polytechnical University, Xi’an 710072, China
2
Shaanxi Key Laboratory of High-Performance Precision Forming Technology and Equipment, Northwestern Polytechnical University, Xi’an 710072, China
3
School of Materials Engineering, Jiangsu University of Technology, Changzhou 213001, China
*
Author to whom correspondence should be addressed.
Submission received: 2 November 2020 / Revised: 12 November 2020 / Accepted: 13 November 2020 / Published: 16 November 2020

Abstract

:
The precipitation behavior of the orthorhombic (O) phase during the slow cooling aging treatment of Ti-22Al-25Nb (at.%) alloy was investigated by microstructural characterization. Then the effect of O phase precipitation on the tensile properties was studied by room-temperature tensile tests. The results showed that the precipitation of the O phase transformed from both grain boundaries and intragranular to only grain boundaries with the temperature increasing. The nucleation mechanism of the O phase from intergranular is composed of sympathetic nucleation and interface instability nucleation. In addition, the results of tensile tests indicated that the ultimate tensile strength of the alloy decreases as the precipitation of the O phase increases. Meanwhile, from the tensile results, it is concluded that the optimum heat treatment process is slow cooling after aging at 950 °C for 1 h.

Graphical Abstract

1. Introduction

As an important high temperature structural metal, Ti2AlNb-based alloys have low density, high specific strength, high temperature strength, and excellent creep resistance [1,2,3]. In comparison with the conventional γ-TiAl-based alloys which suffer from brittleness and poor fracture toughness [4], Ti2AlNb-based alloys have better plasticity at room temperature, giving rise to outstanding application potential in aero engines and aerospace [4,5].
Generally, Ti2AlNb-based alloys are usually composed of 12–25 at.% Al and 12–38 at.% Nb. This leads to the fairly complicated phase constituents. Except the hcp (hexagonal close-packed) Ti3Al phase and bcc (body-centered cubic) B2 phase [6,7,8], the orthorhombic Ti2AlNb phase (O phase) was observed in the Ti-25Al-12.5Nb (at.%) alloy in 1988 for the first time [3]. Subsequent studies revealed that the O phase has important effects on room temperature plasticity, creep resistance, and high temperature strength of Ti3Al-based alloys [9]. The plastic deformation of O phase is mainly carried by (001) <110> slip systems, as well as twinning on (101) planes [1]. Although the alloys with a single O phase were found to have poor room-temperature ductility and toughness, the O phase as a second phase has a positive effect on the mechanical properties. Therefore, the dual-phase (O + B2) Ti2AlNb alloys have attracted intensive attention in the past two decades due to their superior mechanical properties [10,11,12]. Wang et al. reported that after thermomechanical processing followed by solid solution and aging at the B2 + O region, a lamellar microstructure was obtained which exhibited excellent tensile strength at room and elevated temperatures [13,14,15]. However, due to the complicated phase transition pathways, the microstructure of Ti2AlNb-based alloys is quite sensitive to the thermomechanical parameters. As a consequence, the level of thermomechanical processing maturity of Ti2AlNb-based alloys only reaches 3–5 to date [16]. Former researches have shown that the behaviors and the mechanism of phase transitions are related to the temperatures of processing, solution, and age-treatments [17]. Thus, it is of significance to reveal the relationship among the heat-treatment temperatures, microstructure, and mechanical properties, especially the effect of precipitation behavior of the O phase on mechanical properties.
In this paper, the effect of heat treatment parameters on the precipitation behavior of the O phase during the slow cooling aging treatments of a Ti-22Al-25Nb alloy was studied by the quantitative metallographic statistics based on the microstructure images, and the nucleation mechanism of the O phase was clarified. In addition, the relationship between the content of the O phase and its tensile properties was also established.

2. Experimental

Table 1 shows the actual composition of Ti-22Al-25Nb (at.%) alloy. The alloy was prepared by hot rolling and then cut into samples with a dimension of 9 mm × 7 mm × 6 mm by wire electrical discharge machining. A tubular furnace was used for heat treatment experiments. The involved parameters are shown in Figure 1. The samples were heated to 1080 °C with a rate of 8 °C/min, held for 15 min, and then cooled slowly to different temperatures (900 °C, 930 °C, 950 °C) with a rate of 5 °C/min. After that, the samples were held for 0.5 h, 1 h, 1.5 h, and 2 h, respectively, and were then water-quenched to room temperature.
The samples were sectioned along the longitudinal direction for microstructure characterization by X-ray diffraction (XRD, PANalytical X’Pert PRO, CuKα, (PANalytical B. V, Almelo, Netherlands), optical microscopy (OM, OLYMPUS/PMG3, Olympus Corporation, Tokyo, Japan), and scanning electron microscope (SEM, ZEISS SUPRA55, Carl Zeiss AG, Jena, Germany) equipped with electron backscatter diffraction (EBSD) system. After mechanically polished, the samples for OM were etched in the etching solutions of HF:HNO3:H2O = 1:3:8. The SEM and EBSD samples were electrochemically polished for 18–30 s at 20–25 V and about 18 °C in the solution of 6 vol% perchloric acid + 34 vol% butanol + 60 vol% methanol. The quantitative metallographic analysis was performed using the Image-Pro Plus 5.0 (IPP) software (Media Cybernetics, MD, USA). The grain size was evaluated by using the line-intercept method according to GB/T 6394-2017. The EBSD analysis was conducted with a step of 0.2 μm and analyzed by the HKL-Channel 5 software (HKL, Oxford, UK).
For tensile tests, the samples with a dimension of 65 mm × 10 mm × 8 mm were machined from the as-rolled sheet and then machined into a gauge dimension of Φ5 mm × 30 mm after various heat treatments, as shown in Figure 2. The tensile tests were carried out on an Instron-3362 machine (Instron, Boston, MA, USA) with a constant speed of rate of 0.001 mm/s. At least three sets of tensile tests were conducted at each condition.

3. Results and Discussion

3.1. Initial Microstructure

Figure 3a shows the XRD patterns of the as-rolled sheet, it can be concluded that the initial microstructure mainly consists of the O phase and B2 phase. In Figure 3b, it can be seen that the initial grain size of the B2 phase is relatively large with an average size of about 160 μm. According to the BSE (back-scattered electron) image in Figure 3c, numerous O phases with acicular shape are precipitated in the matrix of the B2 phase, while the sizes are far from identical.
From Figure 3d–f, it can be seen that the microstructure after solution treatment sample consists of a single B2 phase with grain sizes ranging from 60 to 350 μm. Obviously, the mean grain size is larger than that in the initial state, indicating inevitable grain growth when the solution is treated in the single-phase region.

3.2. Precipitation Behavior of the O Phase

To understand the precipitation behavior of the O phase, the heat treatment of Ti-22Al-25Nb alloy in Figure 1, named as the slow cooling aging, was carried out. Obviously, the heat-treated microstructure at 900–950 °C/0.5 h is mainly composed of the B2 phase and O phase as shown in Figure 4a–c. The precipitation of the O phase transformed from both grain boundaries and intragranular to only grain boundaries with the temperature increase. When the aging temperature is below 950 °C, the acicular O phase is distributed in the intragranular and grain boundaries. According to the diagram of Ti-22Al-xNb phase [18], the approximate phase regions of the alloy were B2 + O < 955 °C < B2 + O + α2 < 1010 °C < B2 + α2 < 1065 °C < B2, which indicates that 900 °C and 930 °C is far from the transition point of the B2 + O two-phase region and the B2 + O + α2 three-phase region (950–960 °C). The higher degree of subcooling can achieve the higher driving force. Accordingly, the O phase tends to nucleate simultaneously in the grain boundary and intragranular. The average size of the O phase at 930 °C (75 μm) is larger than that at 900 °C (45 μm). With the increase of temperature, the O phase nucleates preferentially at grain boundary. The driving force for coarsening derived from the interfacial energy [19,20]. However, the globular O phase only exists at grain boundaries when the aging temperature reaches 950 °C. The driving force of spheroidization is provided by the reduction in interface energy [21]. The volume fraction of precipitated O phase is counted by IPP. The statistical results are shown in Figure 4d. It can be seen that the volume fraction of the O phase reached more than 50% at 900 °C, while it is less than 5% at 950 °C, indicating that the precipitation of the O phase decreased significantly as the temperature increased.
Figure 5 shows the heat-treated microstructure under various conditions. Distinctly, as the aging time increases, the precipitation of the O phase is promoted. The nucleation and growth of the O phase can occur both in the intragranular and grain boundaries. The quantitation metallographic statistic was performed based on the images. The results in Figure 6 demonstrated that the volume fraction of the acicular O phase increases with the decrease of aging temperature and the increase of time. In addition, the volume fraction of the O phase increases obviously at 900 °C during the initial 0.5 h. Then the growth rate decreases quickly during the subsequent aging process. Similarly, the volume fraction of the O phase increases steadily at 930 °C with the increasing time. However, the growth rate is constant and is lower than that at 900 °C/0.5 h. When the aging temperature increased to close to the transition point of the B2 + O two-phase region and the B2 + O + α2 three-phase region, defined as T2→3, such as 950 °C, there is nearly no nucleation of O phase occurring at the initial aging stage until the aging time extended to 1.5 h. The volume fraction of the O phase increases heavily during the subsequent process. Furthermore, the precipitation rate of the O phase reached the highest point at 900 °C/0.5 h.
Due to the limitation of OM magnification, the heat-treated samples at 900–950 °C/1–2 h were selected for SEM observations. The corresponding microstructure is shown in Figure 7. It is found that the acicular O phase which precipitates at 900 °C/1 h can grow up into the lath cluster with thicknesses of 10–50 μm when holding at 900 °C for 2 h. At 930 °C/1 h, the O phase precipitates along the grain boundaries at a certain angle and then continues to nucleate and grow up nearly perpendicular to the O phase which has already precipitated. When the aging time increases to 2 h, it can be seen that the O phase precipitates discontinuously. The O phase is broken and refined due to the collision of O phase laths. Moreover, almost all O phases precipitate along the grain boundaries at 950 °C.

3.3. Nucleation Mechanism of the O Phase

In this work, two kinds of nucleation mechanisms of the O phase from grain boundaries are observed—sympathetic nucleation and interface instability nucleation [22]. The continuous α layers along the β/β grain boundary denoted αGB, and Widmanstätten plates colonies denoted αWGB [23,24]. Sympathetic nucleation will lead to different orientations between αWGB and αGB, while αWGB grows up from αGB grain with the same orientation due to interface instability. The orientation relationship (OR) between the O phase and the B2 phase satisfies the well-known Burgers OR, {001} O // {110} B2, <110>O // <111>B2 [8,25]. Under ideal conditions, there are 12 variants of the O phase with different orientations that can precipitate in the B2 matrix [26]. However, these variants will not precipitate with the same probability due to various reasons, and the variants selection of the O phase occurs during the process of nucleation and growth [27,28]. The EBSD measurements of the orientation parameters were performed on a triangle grain boundary of B2 phase obtained at 900 °C/0.5 h. Figure 8a,b shows the inverse pole figures (IPF) of the O phase and B2 matrix respectively, and Figure 8c shows the pole figures (PF) of the sites marked in Figure 8a,b. The O phase precipitated from the grain boundaries is named OGB. OG1, OG2, and OG3 refer to the OGB. Moreover, it was found that some O phase variants nucleate in the vicinity of OGB and then grow into the β grains. O1 is a variant nucleated at OG1, and O2, O3 are the variants nucleated at OG2. Obviously, the orientation of OG1 is consistent with that of O1 as shown in Figure 8c. This is a typical interface instability nucleation behavior. Furthermore, the ORs of O1, OG1, and b1 satisfy the Burgers ORs. However, the O phase variants precipitated in b2, such as O2 and O3, have different orientations with OG2. It is also strong evidence that there are two kinds of nucleation mechanisms of the O phase from grain boundaries—sympathetic nucleation and interface instability nucleation. Through the analysis, six kinds of variants precipitated in b2 have a certain orientation relationship with b2 grains, (001)O//(110)B2,[111]B2//[110]O. The OG3 and b3 grains also satisfy the above orientation relationship.

3.4. Tensile Properties

Based on the above analysis, the volume fraction of the O phase increases with the decrease of aging temperature and the increase of time, as shown in Figure 6. To figure out the effect of the O phase on mechanical properties of Ti-22Al-25Nb alloy, room-temperature tensile tests of the heat-treated samples were carried out. Figure 9a–c shows the tensile curve of the samples at different conditions. It can be seen that the elongation of the alloy at 900 °C is lower than that of at 930 °C and 950 °C after aging for 0.5 h. As it was known, the O phase suffers from a lower number of available slip systems compared with the B2 structure [29]. The high-volume fraction of the O phase was considered to be the probable reason for this exhibition at 900 °C [30]. It can be seen that the elongation tends to increase with time at 900 °C. It can be interpreted that most of the O phase preferentially precipitates at the grain boundaries and then grows into the B2 matrix. During the subsequent aging process, the preferentially precipitated O phase grows up and gathers into a lath cluster. This O phase cluster is larger in size than the individual O phase, and the plasticity of the large-size O phase is superior to that of the small-size O phase [13,31]. The elongation of the alloy decreases with the prolongation of aging time due to the gradual precipitation of O phase at 930 °C and 950 °C. Compared with this, the initial precipitation amount of the O phase at 900 °C has reached 50%, and then the growth of the O phase is the main process. At 930 °C and 950 °C, the precipitation process of the O phase is the dominant, the size of the precipitated O phase is relatively large, but its volume fraction is lower, which cannot grow into clusters. Thus, the elongation of the alloy decreases with the increase of O phase precipitation. Figure 9d shows the relationship between the ultimate strength (UTS) and the temperature of the alloy. It can be seen that the UTS of the samples heat-treated at 900 °C is the lowest. On the contrary, the volume fraction of the O phase at 900 °C is the highest in the same aging time. With the increase in the aging temperature, the volume fraction of the O phase decreased. At the early stage of aging, the UTS of the alloy at different temperatures varies slightly, such as 0.5 h. With the extension of aging time, the precipitation of the O phase increased. In the subsequent aging process, the decrease rate of UTS at 900 °C is greater than that of at 930 °C. Boehlert [30] found that when the volume fraction of the O phase reaches 28%, the UTS of the alloy with the B2+O microstructure is higher than that of the full B2 microstructure for the Ti-12Al-38Nb alloy. The precipitation of the O phase can strengthen the matrix, but the corresponding plasticity will lose a little. When the volume fraction of the O phase reaches 79%, both strength and plasticity will reduce significantly. In addition, the thickness of lamellar O affects the tensile properties, as has been known that the coarse lamellar O exhibits poor strengthening effect compared with the fine lamellar O [32]. It can be seen from Figure 7a, the thickness of lamellar O gradually increased with the increase of time at 900 °C. Therefore, although the volume fraction of the O phase increased with the increase of time, the UTS of the alloy decreased. When aging at 950 °C, the content of the O phase is very small at the initial aging stage. As mentioned previously, a small amount of the O phase can strengthen the matrix, but when the volume fraction of the O phase increases to a certain extent, the strength of the alloy will be sacrificed. This is the reason why the strength experiences the first increase then decrease process at 950 °C. The strength of the alloy reaches the highest at 950 °C/1 h.

4. Conclusions

(1)
The acicular O phase precipitated from both the grain boundaries and the intragranular when aging at 900 °C. Increasing the temperature to 930 °C, the acicular O phase tends to precipitate from the grain boundaries. In addition, the average size of the O phase at 930 °C is larger than that at 900 °C. When aging at 950 °C, only a few globular O phases could be found at grain boundaries.
(2)
Under the condition of B2 slow cooling aging, there are two modes of nucleation mechanism of O phase from grain boundaries—sympathetic nucleation and interface instability nucleation.
(3)
With the increase of aging temperature, the volume fraction of the O phase decreased and the strength of the alloy increased. The strength of the alloy decreases with the increase of aging time when aging at 900 °C and 930 °C. When aging at 950 °C, strength experiences a first increase then decrease process. The reasonable match between strength could be obtained at 950 °C/1 h.

Author Contributions

B.W. interpreted data and wrote the paper. B.T. performed experiments and analysed the data. X.C. analysed the data. Q.X. processed the figures. S.Z. revised the language of the paper. H.K. and J.L. provided funds for the study. All authors discussed the results and commented on the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Aeronautical Science Foundation of China (No. 201936053001), the National Natural Science Foundation of China (No.51771150), the Natural Science Foundation of Shaanxi Province (2018JM5174), the Research Fund of the State Key Laboratory of Solidification Processing (NPU), China (Grant No. 2019-TS-07), and the Postgraduate Research & Practice Innovation Program of Jiangsu Province (SJCX19_0732).

Acknowledgments

We would like to thank the Analytical & Testing Center of Northwestern Polytechnical University for SEM testing.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Huang, S.H.; Shao, B.; Xu, W.C.; Shan, D.B.; Guo, B.; Zong, Y.Q. Deformation Behavior and Dynamic Recrystallization of Ti–22Al–25Nb Alloy at 750–990 °C. Adv. Eng. Mater. 2019, 22, 1901231. [Google Scholar] [CrossRef]
  2. Gogia, A.K.; Nandy, T.K.; Banerjee, D.; Carisey, T.; Strudel, J.L. Microstructure and mechanical properties of orthorhombic alloys in the Ti-Al-Nb system. Intermetallics 1998, 6, 741–748. [Google Scholar] [CrossRef]
  3. Banerjee, D.; Gogia, A.K.; Nandy, T.K.; Joshi, V.A. A new ordered orthorhombic phase in a Ti3Al-Nb alloy. Acta Metall. 1988, 36, 871–882. [Google Scholar] [CrossRef]
  4. Clemens, H.; Mayer, S. Design, Processing, Microstructure, Properties, and Applications of Advanced Intermetallic TiAl Alloys. Adv. Eng. Mater. 2013, 15, 191–215. [Google Scholar] [CrossRef]
  5. Zhang, Y.L.; Feng, A.H.; Qu, S.J.; 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]
  6. Pathak, A.; Singh, A.K. A first principles study of Ti2AlNb intermetallic. Solid State Commun. 2015, 204, 9–15. [Google Scholar] [CrossRef]
  7. 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]
  8. Muraleedharan, K.; Nandy, T.K.; Banerjee, D.; Lele, S. Transformations in a Ti-24Al- 15Nb alloy: Part II. a composition invariant βo → O transformation. Metall. Mater. Trans. A 1992, 23, 417–431. [Google Scholar] [CrossRef]
  9. Banerjee, D. The intermetallic Ti2AlNb. Prog. Mater. Sci. 1997, 42, 135–158. [Google Scholar] [CrossRef]
  10. Cai, Q.; Li, M.C.; Zhang, Y.R.; Liu, Y.C.; Li, H.J. Precipitation behavior of Widmansttten O phase associated with interface in aged Ti2AlNb-based alloys. Mater. Charact. 2018, 145, 413–422. [Google Scholar] [CrossRef]
  11. Germann, L.; Banerjee, D.; Guédou, J.Y.; Strudel, J.L. Effect of composition on the mechanical properties of newly developed Ti2AlNb-based titanium aluminide. Intermetallics 2005, 13, 920–924. [Google Scholar] [CrossRef]
  12. Xue, C.; Zeng, W.D.; Wang, W.; Liang, X.B.; Zhang, J.W. 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] [CrossRef]
  13. Wang, W.; Zeng, W.D.; Xue, C.; Liang, X.B.; Zhang, J.W. 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]
  14. Wang, W.; Zeng, W.D.; Xue, C.; Liang, X.B.; Zhang, J.W. 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]
  15. He, Y.-S.; Hu, R.; Luo, W.-Z.; He, T.; Lai, Y.-J.; Du, Y.-J.; Liu, X.-H. Microstructure and mechanical properties of a new Ti2AlNb-based alloy after aging treatment. Rare Met. 2018, 37, 942–951. [Google Scholar] [CrossRef]
  16. He, L.J.; Su, H.H.; Xu, J.H.; Zhang, L. Study on dynamic chip formation mechanisms of Ti2AlNb intermetallic alloy. Int. J. Adv. Manuf. Technol. 2017, 92, 4415–4428. [Google Scholar] [CrossRef]
  17. Wang, W.; Zeng, W.D.; Li, D.; Zhu, B.; Zheng, Y.P.; Liang, X.B. Microstructural evolution and tensile behavior of Ti2AlNb alloys based α2-phase decomposition. Mater. Sci. Eng. A 2016, 662, 120–128. [Google Scholar] [CrossRef]
  18. Kumpfert, J.; Leyens, C. Orthorhombic Titanium Aluminides: Intermetallics with Improved Damage Tolerance; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2003; pp. 59–88. [Google Scholar]
  19. He, Y.S.; Hu, R.; Luo, W.Z.; He, T.; Lai, Y.J.; Du, Y.J.; Liu, X.H. Microstructural evolution and creep deformation behavior of novel Ti−22Al−25Nb−1Mo−1V−1Zr−0.2Si (at.%) orthorhombic alloy. Trans. Nonferrous Met. Soc. China 2019, 29, 313–321. [Google Scholar] [CrossRef]
  20. Gustafson, A.; Hättestrand, M. Coarsening of precipitates in an advanced creep resistant 9% chromium steel—quantitative microscopy and simulations. Mater. Sci. Eng. A 2002, 333, 279–286. [Google Scholar] [CrossRef]
  21. Zhang, H.Y.; Li, C.; Ma, Z.Q.; Yu, L.M.; Liu, Y.C. 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]
  22. Tang, B.; Kou, H.C.; Zhang, W.; Gao, P.Y.; Li, J.S. Study on the formation mechanism of α lamellae in a near β titanium alloy. Prog. Nat. Sci. 2016, 26, 385–390. [Google Scholar] [CrossRef] [Green Version]
  23. Da Costa Teixeira, J.; Appolaire, B.; Aeby-Gautier, E.; Denis, S.; Bruneseaux, F. Modeling of the effect of the β phase deformation on the α phase precipitation in near-β titanium alloys. Acta Mater. 2006, 54, 4261–4271. [Google Scholar] [CrossRef] [Green Version]
  24. Appolaire, B.; Héricher, L.; Aeby-Gautier, E. Modelling of phase transformation kinetics in Ti alloys–Isothermal treatments. Acta Mater. 2005, 53, 3001–3011. [Google Scholar] [CrossRef] [Green Version]
  25. Muraleedharan, K.; Gogia, A.K.; Nandy, T.K.; Lele, D.B. Transformations in a Ti-24AI-15Nb alloy: Part I. Phase equilibria and microstructure. Metall. Mater. Trans. A 1992, 23, 401–415. [Google Scholar] [CrossRef]
  26. Zhao, Z.B.; Wang, Q.J.; Hu, Q.M.; Liu, J.R.; Yu, B.B.; Yang, R. Effect of β (110) texture intensity on α-variant selection and microstructure morphology during β→α phase transformation in near α titanium alloy. Acta Mater. 2017, 126, 372–382. [Google Scholar] [CrossRef]
  27. Wang, S.C.; Aindow, M.; Starink, M.J. Effect of self-accommodation on α/α boundary populations in pure titanium. Acta Mater. 2003, 51, 2485–2503. [Google Scholar] [CrossRef]
  28. Van Bohemen, S.M.C.; Kamp, A.; Petrov, R.H.; Kestens, L.A.I.; Sietsma, J. Nucleation and variant selection of secondary α plates in a β Ti alloy. Acta Mater. 2008, 56, 5907–5914. [Google Scholar] [CrossRef]
  29. Popille, F.; Douin, J. The dislocation microstructure in orthorhombic O Ti2AlNb deformed between room temperature and 800 °C. Philos. Mag. A 1996, 73, 1401–1418. [Google Scholar] [CrossRef]
  30. 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]
  31. Xue, C.; Zeng, W.D.; Wang, W.; Liang, X.B.; Zhang, J.W. 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]
  32. Cowen, C.J.; Boehlert, C.J. Microstructure, creep, and tensile behavior of a Ti–21Al–29Nb(at.%) orthorhombic+B2 alloy. Intermetallics 2006, 14, 412–422. [Google Scholar] [CrossRef]
Figure 1. The schematic illustration of the heat treatment process.
Figure 1. The schematic illustration of the heat treatment process.
Metals 10 01515 g001
Figure 2. Machining drawing of the tensile specimens.
Figure 2. Machining drawing of the tensile specimens.
Metals 10 01515 g002
Figure 3. Initial microstructure (ac) and the one after solution treatment at 1080 °C for 15 min (df) for the rolled Ti-22Al-25Nb alloy. (a,d) XRD patterns, (b,e) optical microscopy (OM) images, (c,f) back-scattered electron (BSE) images.
Figure 3. Initial microstructure (ac) and the one after solution treatment at 1080 °C for 15 min (df) for the rolled Ti-22Al-25Nb alloy. (a,d) XRD patterns, (b,e) optical microscopy (OM) images, (c,f) back-scattered electron (BSE) images.
Metals 10 01515 g003
Figure 4. Microstructure of Ti-22Al-25Nb alloy held for 30 min at different temperatures. (a) 900 °C, (b) 930 °C, (c) 950 °C, (d) statistical results for the volume fraction of the orthorhombic (O) phase.
Figure 4. Microstructure of Ti-22Al-25Nb alloy held for 30 min at different temperatures. (a) 900 °C, (b) 930 °C, (c) 950 °C, (d) statistical results for the volume fraction of the orthorhombic (O) phase.
Metals 10 01515 g004
Figure 5. The OM images of the heat-treated Ti-22Al-25Nb alloy at (a) 900 °C/1 h, (b) 900 °C/1.5 h, (c) 900 °C/2 h, (d) 930 °C/1 h, (e) 930 °C/1.5 h, (f) 930 °C/2 h, (g) 950 °C/1 h, (h) 950 °C/1.5 h, and (i) 950 °C/2 h.
Figure 5. The OM images of the heat-treated Ti-22Al-25Nb alloy at (a) 900 °C/1 h, (b) 900 °C/1.5 h, (c) 900 °C/2 h, (d) 930 °C/1 h, (e) 930 °C/1.5 h, (f) 930 °C/2 h, (g) 950 °C/1 h, (h) 950 °C/1.5 h, and (i) 950 °C/2 h.
Metals 10 01515 g005
Figure 6. The volume fraction of the acicular O phase varies with temperature and time.
Figure 6. The volume fraction of the acicular O phase varies with temperature and time.
Metals 10 01515 g006
Figure 7. The BSE images of the heat-treated Ti-22Al-25Nb alloy at (a) 900 °C/1 h, (b) 930 °C/1 h, (c) 950 °C/1 h, (d) 900 °C/2 h, (e) 930 °C/2 h, (f) 950 °C/2 h.
Figure 7. The BSE images of the heat-treated Ti-22Al-25Nb alloy at (a) 900 °C/1 h, (b) 930 °C/1 h, (c) 950 °C/1 h, (d) 900 °C/2 h, (e) 930 °C/2 h, (f) 950 °C/2 h.
Metals 10 01515 g007
Figure 8. The inverse pole figures (IPF) maps (a) of O phase and (b) of B2 phase. (c) The {110}, <111> pole figures of B2 phase and the {001}, <110> pole figures of O variants.
Figure 8. The inverse pole figures (IPF) maps (a) of O phase and (b) of B2 phase. (c) The {110}, <111> pole figures of B2 phase and the {001}, <110> pole figures of O variants.
Metals 10 01515 g008
Figure 9. The effect of B2 slow cooling to aging treatment to ambient tensile curve for Ti-22Al-25Nb alloy: (a) 900 °C/0–2 h, (b) 930 °C/0–2 h, (c) 950 °C/0–2 h, (d) the ultimate strength (UTS) of the alloy at different temperatures.
Figure 9. The effect of B2 slow cooling to aging treatment to ambient tensile curve for Ti-22Al-25Nb alloy: (a) 900 °C/0–2 h, (b) 930 °C/0–2 h, (c) 950 °C/0–2 h, (d) the ultimate strength (UTS) of the alloy at different temperatures.
Metals 10 01515 g009
Table 1. The actual composition of Ti-22Al-25Nb alloy used in the present work (wt.%).
Table 1. The actual composition of Ti-22Al-25Nb alloy used in the present work (wt.%).
Rolled PlateTiAlNbCNHO
wt.%Bal.10.1141.250.00510.00310.0020.0471
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

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. https://0-doi-org.brum.beds.ac.uk/10.3390/met10111515

AMA Style

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(11):1515. https://0-doi-org.brum.beds.ac.uk/10.3390/met10111515

Chicago/Turabian Style

Wei, Beibei, Bin Tang, Xiaofei Chen, Qin Xu, Shuaijin Zhang, Hongchao Kou, and Jinshan Li. 2020. "Precipitation Behavior of Orthorhombic Phase in Ti-22Al-25Nb Alloy during Slow Cooling Aging Treatment and Its Effect on Tensile Properties" Metals 10, no. 11: 1515. https://0-doi-org.brum.beds.ac.uk/10.3390/met10111515

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