Next Article in Journal
Analysis of the State of Building Conservation through Study of Damage and Its Evolution with the State of Conservation Assessment BIM Model (SCABIM)
Next Article in Special Issue
Microstructural Variation and Evaluation of Formability According to High-Temperature Compression Conditions of AMS4928 Alloy
Previous Article in Journal
Research on Older Persons’ Access and Use of Technology in the Arab Region: Critical Overview and Future Directions
Previous Article in Special Issue
Process to Establish the Enhance of Fatigue Life of New Mechanical System Such as a Drawer by Accelerated Tests
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Molybdenum Content on Microstructure and Mechanical Properties of Ti-Mo-Fe Alloys by Powder Metallurgy

1
Department of Materials Science and Metallurgical Engineering, Sunchon National University, Suncheon 57922, Korea
2
Material Technical Innovation Group (MTIG), Hwaseong-si 18574, Korea
*
Author to whom correspondence should be addressed.
Appl. Sci. 2022, 12(14), 7257; https://doi.org/10.3390/app12147257
Submission received: 30 June 2022 / Revised: 13 July 2022 / Accepted: 15 July 2022 / Published: 19 July 2022
(This article belongs to the Special Issue Selected Papers from ICAMSME 2022)

Abstract

:
Titanium has many limitations in coverage and frequency of application due to its expensive alloying elements and complex manufacturing process. The biocompatible Ti-Mo-Fe ternary beta titanium alloys were designed by replacing high-cost beta-stabilizer elements (V, Nb, Zr, etc.) with low-cost Mo and Fe elements. In addition, it was attempted to obtain a low-cost, high-strength beta-titanium alloy with 800 MPa or more by applying the powder metallurgy process technology to the Ti-Mo-Fe alloy system. The added Mo element has the effect of reducing the elastic modulus of the titanium alloy without reducing its strength. In this study, Ti-Mo-Fe alloys designed with different Mo contents were fabricated using a powder metallurgy process and analyzed in connection with microstructural properties, phase changes, and mechanical properties. As Mo contents are increased, the α-lath thickness of Widmanstätten decreases and the size of prior β grain decreases. It was confirmed that the hardness and tensile strength were excellent and were compared with the ingot material of the same alloy system.

1. Introduction

Titanium has excellent specific strength and corrosion resistance and is widely used in aerospace, the defense industry, and biomaterials [1,2,3,4,5]. In particular, titanium has lower elastic modulus and higher specific strength than cobalt alloys, stainless steel, alumina, etc., and is widely used as biomaterials. However, commercially pure titanium (c.p. Ti) has comparatively low mechanical properties, such as tensile strength, making it unsuitable for certain applications. Further, it has been reported that the aluminum contained in a commercial titanium alloy, Ti-6Al-4V, can cause Alzheimer’s disease [6], and vanadium can cause allergies and serious complications due to the release of toxic ions [7]. Therefore, it is required to develop biocompatible beta-titanium alloys in order to replace these components [8] and, at this time, niobium, zirconium, tantalum, and molybdenum are attracting attention as alloy elements that are not harmful to the human body [9]. Actually, many beta-titanium alloys, such as Ti-3Zr-5Fe-5Mo, Ti-14Mn, Ti-10Fe-10Ta, Ti-27Nb-7Fe-8Cr, Ti-33Zr-3Fe-4Cr, and Ti-7.5Mo, have been recently developed.
In fact, a beta-titanium alloy has excellent heat treatment characteristics compared to the α + β alloy and represents an increase in elongation due to its BCC structure, including a low elastic modulus, and then is used not only in biomaterials but also in various other fields. However, the β -stabilizer elements, tantalum, niobium, and zirconium, inevitably increase costs due to their expensive price, resulting in many restrictions on the use of beta titanium [10,11]. Thus, several studies have been actively conducted to replace expensive alloy elements, such as tantalum, niobium, and zirconium, with inexpensive elements (Fe, Mo, Mn, etc.) [12,13].
Iron, an inexpensive element, has a strong β -stabilizer effect and a good reinforcing effect on non-toxic and non-allergic titanium alloys [14]. In addition, it is reported to improve the sintering reaction of titanium alloys [15] because it represents a faster diffusion rate than the atom of titanium itself in the β phase as a rapidly diffused element [16]. Likewise, molybdenum is a biocompatible β-stabilizer element and when molybdenum is added to a titanium alloy, it has been reported to have high corrosion resistance [17,18,19,20], lower the elastic modulus, and increase the strength [21].
The titanium alloy obtained through the existing ingot metallurgy (IM) process results in significant process costs, such as multiple vacuum remelting, alloying additions, homogenization, hot deformation and annealing, etc. [22]. However, it is relatively easy to manufacture complex parts in a powder metallurgy (PM) process [23], and simpler work processes and facilities are used for comparing it with the ingot process. In addition, unlike the ingot process, Ti and alloy elements are heated to a temperature lower than the melting point, which has the advantage of significantly reducing costs [24,25].
Therefore, this study designed a biocompatible Ti-Mo-Fe high-strength beta-titanium alloy and applied the powder metallurgy process technology using hydrogenation–dehydrogenation (HDH) titanium powder to ensure excellent price competitiveness. Further, changes in the microstructure of the Ti-xMo-4Fe (x = 3.4, 5) powder alloys with different Mo contents were examined and then it was evaluated and analyzed in connection with mechanical properties. The results were compared with the properties of ingot alloys of the same alloy system.

2. Materials and Experiment

2.1. Alloy Design and Preparation

The Ti-xMo-4Fe (x = 3.4, 5) designed in this study is presented in the phase stability map ( B o ¯ M d ¯ map), Figure 1 [26]. The B o ¯ M d ¯ map is a theoretical approach to the design of titanium alloys established by Morinaga et al. [27] and consists of two parameters. In addition, it can be clearly divided into α phase, α + β , and β phase according to the alloy composition, and when the value of Bo increases in the β phase and the value of Md decreases, it becomes a stable state. The Bond order ( B o ¯ ) is a value representing the covalent bonding strength between the titanium and alloying elements and it is reported that as B o ¯ increases, the elastic modulus tends to decrease [28]. The Metal d-orbital energy level ( M d ¯ ) is correlated with the electronegativity in the element. These two parameters can be calculated using the following Equation (1).
M d ¯ = i = 1 n X i · ( M d ) i ,   B o ¯ = i = 1 n X i · ( B o ) i
Using Equation (1), the values of the Metal d-orbital energy level ( M d ¯ ) and Bond order ( B o ¯ ) of the Ti-3.4Mo-4Fe (hereinafter referred to as TMF 34) alloys were 2.39 and 2.79, respectively. Further, the values of the Ti-5Mo-4Fe (hereinafter referred to as TMF 54) alloys were 2.38 and 2.79, respectively.
The objective of this study is to compare the alloy designed in this study by applying the powder metallurgy (hereinafter referred to as P/M) and ingot metallurgy (hereinafter referred to as I/M) processes. In the case of the powder alloy, the HDH titanium powder was used to further reduce the cost of the titanium alloy. The powder used in P/M represented sizes of Ti (<25 μm), Mo (1~2 μm), and Fe (5~7 μm), which were blended and bound according to Ti-xMo-4Fe (x = 3.4, and 5 wt.%). It was blended using a mixer, pressed with a force of 200 kg/cm2 for 15 s using a hydraulic press, and then the resulting bulk compacts were sintered in a sintering furnace for 5 h at 1250 °C. It was then furnace cooled to room temperature. Further, in order to compare it with ingots having the same alloy composition, ingots having a composition of Ti-xMo-4Fe (x = 3.4, 5) with a diameter of 14 mm were fabricated through applying a vacuum arc remelting method using Ti with 99.9% purity, Mo with 99.95% purity, and Fe with 99.5% purity. In the case of the porosity of the alloy applied to P/M, it was derived using Equation (2) [29].
P = [ 1 ( ρ ρ 0 ) ] × 100
where ρ and ρ 0 represent the apparent density and theoretical density in the P/M alloy measured using the Archimedean method. The number of pores was measured using an optical microscope.

2.2. Microstructural Characterization

In order to compare and analyze the microstructure of the designed alloys, Ti-xMo-4Fe (x = 3.4, 5 wt.%), the specimen was cut perpendicular to the longitudinal direction. The specimen was hot mounted for fine polishing and was finely polished using abrasive papers of #220~#2000, abrasives of 6 μm, 3 μm, and 1 μm, and colloid silica. Then, the specimen was etched by mixing 100 mL of distilled water (H2O) with 2 mL of nitric acid (HNO3) and 2 mL of hydrofluoric acid (HF). Using an optical microscope (OM, BX53M, Olympus, Tokyo, Japan) and an electron back-scattered diffraction pattern analyzer (EBSD, JSM-7100F, JEOL, Tokyo, Japan), the changes in microstructures according to the process method and Mo content were observed.

2.3. Mechanical Properties

The Vickers hardness test and room temperature tensile test were conducted to evaluate mechanical properties at room temperature. The hardness test was conducted using a Vickers hardness tester (HM-200, Mitutoyo, Kawasaki, Japan) and the average value excluding the maximum and minimum values was calculated by measuring 12 points while maintaining 1 kgf outward from the center of the designed alloy specimen for 15 s. The room temperature tensile test was conducted with a strain rate of 1 × 10 3 /s by processing it with a tensile specimen standard of ASTM subsize using a room temperature tensile tester (Room Temperature Tensile Machine, BESTUTM-10MD, Ssaul Bestech, Seoul, Korea).

3. Results and Discussion

3.1. Microstructural Characterization

The microstructures of the designed P/M alloys are presented in Figure 2a,b. Due to the furnace cooling after the sintering process, Widmanstätten, in which the dark α and light β phases were generated in a layered pattern in all P/M alloys and pores, were also observed. In the case of the inside of TMF 34 I/M, α + β lamellar Widmanstätten structures were observed and the α phases precipitated within the equiaxed β grain inside of the TMF 54 I/M.
In the case of the I/M alloy, the lamellar Widmanstätten microstructure with a α + β was observed in TMF 34 I/M, and in the case of TMF 54 I/M, the α-phase precipitated inside equiaxed β grain had different structures (Figure 2c,d). In the case of the Widmanstätten structure, cracks are first generated at the boundary of colony or grain boundary α-phase and it is reported that cracks are propagated in this way and its tensile ductility is low [30]. In the equiaxed structure, cracks occur mainly at the interface of α/β, resulting in high strain. Therefore, it is expected that TMF 54 I/M with the equiaxed structure represents higher strain than TMF 34 I/M with the α + β lamellar Widmanstätten.
Figure 3 shows an observation of the Inverse pole figure (IPF) map for analyzing the Widmanstätten observed in the P/M alloys. The thickness of the Widmanstätten α -lath of TMF 34 P/M was measured to be 12.289 μm and the thickness of the Widmanstätten α -lath of TMF 54 P/M was measured to be 10.547 μm. It can be seen that as the Mo content increases, the thickness of the Widmanstätten α -lath becomes thinner. Mo is one of the β -stabilizer elements and the temperature in the β transformation point decreases. In this study, when furnace cooling is performed after the sintering process at 1250 °C, the exposure time to the β transformation temperature increases according to the increase in the Mo content. Therefore, the fraction rate of the β phase increases, and accordingly, it is determined that the thickness of the Widmanstätten α -lath decreased as the time of exposure to the α phase decreased. In the practical observation of the EBSD phase map of the TMF 34 and TMF54 alloys, it was confirmed that the fraction rate of the α phase decreased from 56% to 50.4% and the fraction of the β phase increased from 44% to 49.6% (Figure 4).
Figure 5 shows the pore analysis table observed in the P/M alloys. Small pores below 10 μm are mainly due to the ‘Kirkendal’ effect in powder metallurgy and are caused by atomic mutual diffusion during a solid-phase sintering process [31]. While pores with sizes of 40 μm or more were not observed in TMF 34 P/M, pores with sizes of 40 μm to 45 μm were observed in TMF 54 P/M. As a result of analyzing the pore distribution, the pore fraction rate increased by about 1.4% from 2.1% to 3.5% as the content of Mo increased. This is due to the low-diffusion coefficient of Mo and the diffusion rate of Mo in Ti particles is slower than that of Ti itself. Therefore, the addition of Mo interferes with the sintering reaction and is sensitive to the formation of Kirkendall pores [31,32]. Thus, the pore fraction rate of the TMF 54P/M alloy was measured to be 1.4% higher because the Mo content of TMF 54P/M was increased more than that of TMF 34P/M. In addition, it is known that the low diffusivity of Mo interferes with the movement of the particle boundary and affects grain refinement [33].
According to the IPF map in Figure 6, schematized by measuring the misorientation of the TMF design alloy, the size of the prior β grain decreased from 226.55 μm to 211.99 μm when the Mo content increased from the actual TMF 34 P/M to TMF 54 P/M. The prior β grain size is an important factor that can significantly affect mechanical properties. Because it relies significantly on the prior β grain size, the grain control is essential. In fact, it is known that the maximum tensile strength and yield strength increase according to the decrease in the prior β grain size [34].

3.2. Mechanical Properties

Figure 7 shows the results of measuring Vickers hardness for four types of TMF β alloys manufactured according to the alloy composition and process method. In the case of the P/M alloy, it was determined that the hardness values of TMF 34 P/M and TMF 54 P/M were 397.5 Hv and 396.5 Hv, respectively, which slightly decreased when the Mo content increased. In general, it is known that as the Mo content increases, the hardness value increases due to the solid-solution strengthening effect [35], but when the Mo content increases, the hardness value increases due to the decrease in the thickness of Widmanstätten α -lath and the pore fraction increases (2.1% → 3.5%). As a result, the effect of decreasing the hardness value was largely affected and it is considered that the hardness value of TMF 54 P/M was similar to the hardness value of TMF 34 P/M.
In addition, in the case of the I/M alloy, the hardness values of TMF 34 I/M and TMF 54 I/M are 422.1 Hv and 414.8 Hv, respectively, where the hardness value of TMF 34 I/M is slightly higher than that of TMF 54 I/M. In the case of TMF 34 I/M, it is an α + β lamellar Widmanstätten microstructure. In the case of TMF 54 I/M, it is a relatively soft α-phase precipitated in the equiaxed β grain. Therefore, in the case of the I/M alloy, the hardness value of TMF 34 I/M was measured higher than the hardness value of TMF 54 I/M. In addition, the P/M alloy has a hardness value lower than that of I/M with the same composition and the hardness of the P/M alloy is decreased due to pores in the P/M alloy. Thus, it is considered that the hardness value of the P/M alloy is lower than that of the ingot alloy.
The tensile stress–strain curves obtained from the room temperature tensile test conducted on four types of the TMF beta alloys are shown in Figure 8 and the tensile characteristic values (yield strength, ultimate tensile strength, and elongation) measured from the curve are shown in Table 1. All P/M alloys were fractured before the yield point, and mechanical strength and elongation were measured lower than those of the I/M alloys. This shows that the fracture occurred before the yield point due to pores and the fracture surface shows a cleavage fracture form in all P/M alloys. It can be seen that cracks started from these pores (Figure 9a,b). In addition, the ultimate tensile strength of TMF 54 P/M was 815.27 MPa, which was higher than that of TMF 34 P/M. The cause is considered to be the influence of the prior β grain size and the strength increases according to the Hall–Petch equation because the prior β grain size is small [36].
In the case of the I/M alloy, TMF 54 I/M has the equiaxed β grain with a relatively soft microstructure. As mentioned above, in the equiaxed microstructure, it is reported that the slips are strengthened due to the interference in the movement caused by the α phases precipitating inside the equiaxed β grain, and not only slips but also twins act as a deformation mechanism, contributing to the increase in elongation [34]. Therefore, the TMF 54 I/M alloys with the equiaxed microstructure represents higher elongation than the TMF 34 I/M alloys with α + β lamellar Widmanstätten microstructure. As shown in Figure 9c,d, some dimple structures, which indicate a ductile fracture behavior, were observed on the fracture surface of TMF 54 I/M. However, in the case of TMF 34 I/M, the cleavage fracture surface, a characteristic of brittle fracture behavior, was observed.

4. Conclusions

In this study, ternary beta-titanium alloys with a tensile strength of 800 MPa or higher were designed by replacing alloy elements, Al and V, with biocompatible Mo and Fe. Ti-Mo-Fe alloys were fabricated by applying a powder metallurgy process, and microstructure and mechanical properties according to Mo content were considered. The major results are summarized as follows.
1.
It is possible to secure excellent price competitiveness by successfully developing Ti-Mo-Fe P/M alloys by adding low-cost Mo and Fe, which can fabricate complex parts and reduce costs. The cost of alloying elements is reduced by 81–84% and 63–67% compared to the widely used Ti-6Al-4V and Ti-13Nb-13Zr alloys, respectively. Further, the cost may be further reduced by using HDH powder.
2.
In the case of TMF 34 I/M, an α + β lamellar Widmanstätten microstructure was observed, and in the case of TMF 54 I/M, an α phase precipitating inside the equiaxed β grain was observed. In the equiaxed microstructure, where cracks occur mainly at the α / β interface, deformation can be generated by slips and twins. Further, it is reinforced by the precipitated fine α phase according to an increase in Mo content, showing high tensile strength and elongation in the TMF 54 I/M alloys.
3.
According to the increase in Mo content, TMF 34 P/M → TMF 54 P/M, the thickness of the Widmanstätten α -lath decreases (12.289 μm → 10.547 μm) and the size of prior β grains decrease (226.55 μm → 211.99 μm). For this reason, it was possible to obtain excellent tensile strength of 800 MPa or more, even though the pore fraction increased (2.1% → 3.5%) as the Mo content increased.
4.
Tensile fractures occurred in the elastic region before the yield point due to pores in both the TMF 34 P/M and TMF 54 P/M alloys, and the change in Mo content was analyzed with I/M alloys by the E and σ reinforcing mechanisms.

Author Contributions

Conceptualization, H.-W.H. and D.-G.L.; methodology, H.-W.H. and D.-G.L.; validation, H.-W.H., J.-H.P. and D.-G.L.; formal analysis, H.-W.H., J.-H.P. and D.-G.L.; investigation, H.-W.H. and D.-G.L.; writing—original draft preparation, H.-W.H.; writing—review and editing, D.-G.L.; supervision, D.-G.L.; funding acquisition, J.-H.P. and D.-G.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data are available within the manuscript.

Acknowledgments

This work was supported by the Korean government MOTIE (the Ministry of Trade, Industry and Energy), the Korea Evaluation Institute of Industrial Technology (KEIT) (No. 20010047), and the Korea Institute for Advancement of Technology (KIAT) (No. P0002019).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Leyens, C.; Peters, M. Titanium and Titanium Alloys: Fundamentals and Applications, 1st ed.; betz-druck GmbH: Darmstadt, Germany, 2004. [Google Scholar]
  2. Rehman, A.U.; Kishore Babu, N.; Talari, M.K.; Usmani, Y.; Alkhalefah, H. Characterisation of Microstructure and Mechanical Properties of Linear Friction Welded α+β Titanium Alloy to Nitinol. Appl. Sci. 2021, 11, 10680. [Google Scholar] [CrossRef]
  3. Lütjering, G.; Williams, J.C. Titanium, 2nd ed.; Springer: Berlin/Heidelberg, Germany; New York, NY, USA, 2007; pp. 383–415. [Google Scholar]
  4. Hwang, Y.-J.; Park, Y.-K.; Kim, C.-L.; Kim, J.-Y.; Lee, D.-G. Mechanical Properties Variation of Ti-6Al-4V Alloys by Microstructural Control. J. Korean Soc. Heat Treat. 2016, 29, 220–226. [Google Scholar] [CrossRef] [Green Version]
  5. Yang, S.; Su, S.; Liu, X.; Han, P. Study on Milling Temperature of Titanium Alloy with Micro-Textured Ball End Milling Cutter under Radius of Blunt Edge. Appl. Sci. 2020, 10, 587. [Google Scholar] [CrossRef] [Green Version]
  6. Walker, P.R.; LeBlanc, J.; Sikorska, M. Effects of aluminum and other cations on the structure of brain and liver chromatin. Biochemistry 1989, 28, 3911–3915. [Google Scholar] [CrossRef]
  7. Manam, N.; Harun, W.; Shri, D.; Ghani, S.; Kurniawan, T.; Ismail, M.H.; Ibrahim, M. Study of corrosion in biocompatible metals for implants: A review. J. Alloys Compd. 2017, 701, 698–715. [Google Scholar] [CrossRef] [Green Version]
  8. Hwang, Y.J.; Lee, D.G. Effect of Al Addition on Corrosion Protection of Ti-39Nb-6Zr Alloy for Biological Applications. J. Nanosci. Nanotechnol. 2019, 19, 3811–3815. [Google Scholar] [CrossRef] [PubMed]
  9. Li, Y.; Yang, C.; Zhao, H.; Qu, S.; Li, X.; Li, Y. New developments of Ti-based alloys for biomedical applications. Materials 2014, 7, 1709–1800. [Google Scholar] [CrossRef] [Green Version]
  10. Allen, P.G.; Bania, P.J.; Hutt, A.J.; Combres, Y. Titanium ‘95 Science and Technology. In Proceedings of the Eighth World Conference on Titanium, International Convention Centre, Birmingham, UK, 22–26 October 1995; Blenkinsop, P.A., Evans, W.J., Flower, H.M., Eds.; The Institute of Metals: London, UK, 1996. [Google Scholar]
  11. Lee, Y.T. Titanium; Steel & Metal News Co., Ltd.: Seoul, Korea, 2009; pp. 61–128. [Google Scholar]
  12. Gepreel, M.; Niinomi, M. Biocompatibility of Ti-alloys for long-term implantation. J. Mech. Behav. Biomed. Mater. 2013, 20, 407–415. [Google Scholar] [CrossRef]
  13. Ehtemam-Haghighi, S.; Attar, H.; Okulov, I.V.; Dargusch, M.S.; Kent, D. Microstructural evolution and mechanical properties of bulk and porous low-cost Ti–Mo–Fe alloys produced by powder metallurgy. J. Alloys Compd. 2021, 853, 156768. [Google Scholar] [CrossRef]
  14. Bao, Y.; Zhang, M.; Liu, Y.; Yao, J.; Xiu, Z.; Xie, M.; Sun, X. High strength, low modulus and biocompatible porous Ti–Mo–Fe alloys. J. Porous Mater. 2014, 21, 913–919. [Google Scholar] [CrossRef]
  15. Nakajima, H.; Koiwa, M. Diffusion in titanium. ISIJ Int. 1991, 31, 757–766. [Google Scholar] [CrossRef] [Green Version]
  16. Jindal, V.; Sarda, A.; Degnah, A.; Chandran, K.R. Effect of iron & boron content onthe Spark Plasma Sintering of Ti-B-Fe alloys. Adv. Powder Technol. 2019, 30, 423–427. [Google Scholar]
  17. Ho, W.-F. A comparison of tensile properties and corrosion behavior of cast Ti–7.5Mo with c.p. Ti, Ti–15Mo and Ti–6Al–4V alloys. J. Alloys Compd. 2008, 464, 580–583. [Google Scholar] [CrossRef]
  18. Sugano, M.; Tsuchida, Y.; Satake, T.; Ikeda, M. A microstructural study of fatigue fracture in titanium–molybdenum alloys. Mater. Sci. Eng. A 1998, 243, 163–168. [Google Scholar] [CrossRef]
  19. Kumar, S.; Tsn, S.N. Electrochemical characterization of β-Ti alloys in Ringer’s solution for implant application. J. Alloys Compd. 2009, 479, 699–703. [Google Scholar] [CrossRef]
  20. Lin, D.-J.; Chuang, C.-C.; Chern Lin, J.-H.; Lee, J.-W.; Ju, C.-P.; Yin, H.-S. Bone formation at the surface of low modulus Ti–7.5Mo implants in rabbit femur. Biomaterials 2007, 28, 2582–2589. [Google Scholar] [CrossRef]
  21. Santos, P.F.; Niinomi, M.; Liu, H.; Cho, K.; Nakai, M.; Trenggono, A.; Champagne, S.; Hermawan, H.; Narushima, T. Improvement of microstructure, mechanical and corrosion properties of biomedical Ti-Mn alloys by Mo addition. Mater. Des. 2016, 110, 414–424. [Google Scholar] [CrossRef]
  22. Carman, A.; Zhang, L.; Ivasishin, O.; Savvakin, D.; Matviychuk, M.; Pereloma, E. Role of alloying elements in microstructure evolution and alloying elements behaviour during sintering of a near-beta titanium alloy. Mater. Sci. Eng. A 2011, 528, 1686–1693. [Google Scholar] [CrossRef]
  23. Menhal Shbeh, M.; Yerokhin, A.; Goodall, R. Microporous Titanium through Metal Injection Moulding of Coarse Powder and Surface Modification by Plasma Oxidation. Appl. Sci. 2017, 7, 105. [Google Scholar] [CrossRef] [Green Version]
  24. Santos, D.; Pereira, M.d.S.; Cairo, C.; Graca, M.; Henriques, V. Isochronal Sintering of the Blended Elemental Ti-35Nb Alloy. Mater. Sci. Eng. A 2008, 472, 193–197. [Google Scholar] [CrossRef]
  25. Bolzoni, L. Low-cost Fe-bearing powder metallurgy Ti alloys. Met. Powder Rep. 2019, 74, 308–313. [Google Scholar] [CrossRef]
  26. Kuroda, D.; Niinomi, M.; Morinaga, M.; Kato, Y.; Yashiro, T. Design and mechanical properties of new β type titanium alloys for implant materials. Mater. Sci. Eng. A 1998, 243, 244–249. [Google Scholar] [CrossRef]
  27. Morinaga, M.; Kato, M.; Kamimura, T.; Fukumoto, M.; Harada, I.; Kubo, K. Theoretical design of beta-type titanium alloys. In Titanium’92: Science and Technology; The Minerals, Metals & Materials Society: Pittsburgh, PA, USA, 1993; pp. 217–224. [Google Scholar]
  28. Abdel-Hady, M.; Hinoshita, K.; Morinaga, M. General approach to phase stability andelastic properties of β-type Ti-alloys using electronic parameters. Scr. Mater. 2006, 55, 477–480. [Google Scholar] [CrossRef]
  29. Xu, J.; Bao, L.; Liu, A.; Jin, X.; Tong, Y.; Luo, J.; Zhong, Z.; Zheng, Y. Microstructure, mechanical properties and super elasticity of biomedical porous Ni Ti alloy prepared by microwave sintering. Mater. Sci. Eng. C 2015, 46, 387–393. [Google Scholar] [CrossRef]
  30. Bae, K.-M.; Park, C.-G.; Chang, Y.-W.; Lee, C.-S. The Effect of α-Phase Morphology on the Microcrack initiation Behavior of a Ti-6Al-4V Alloy. Korean J. Met. Mater. 1993, 31, 105–112. [Google Scholar]
  31. Tang, C.Y.; Zhang, L.N.; Wong, C.T.; Chan, K.C.; Yue, T.M. Fabrication and characteristics of porous NiTi shape memory alloy synthesized by microwave sintering. Mater. Sci. Eng. A 2011, 528, 6006–6011. [Google Scholar] [CrossRef]
  32. Liu, Y.; Chen, L.; Tang, H.; Liu, C.T.; Liu, B.; Huang, B. Design of powder metallurgy titanium alloys and composites. Mater. Sci. Eng. A 2006, 418, 25–35. [Google Scholar] [CrossRef]
  33. Zhang, W.; Xie, H.; Hui, S.; Ye, W.; Yu, Y.; Song, X.; Peng, L.; Huang, G.; Yang, Z. In-Situ SEM Observation on Fracture Behavior of Titanium Alloys with Different Slow-Diffusing β Stabilizing Elements. Materials 2020, 13, 1848. [Google Scholar] [CrossRef] [Green Version]
  34. Tiley, J.; Searles, T.; Lee, E.; Kar, S.; Banerjee, R.; Russ, J.C.; Fraser, H.L. Quantification of microstructural features in α/β titanium alloys. Mater. Sci. Eng. A 2004, 372, 191–198. [Google Scholar] [CrossRef]
  35. Lee, E.B.; Han, M.K.; Kim, B.J.; Song, H.J.; Park, Y.J. Effect of molybdenum on the microstructure, mechanical properties and corrosion behavior of Ti alloys. Int. J. Mater. Res. 2014, 105, 847–853. [Google Scholar] [CrossRef]
  36. Chong, Y.; Bhattacharjee, T.; Shibata, A.; Tsuji, N. Investigation of the grain size effect on mechanical properties of Ti-6Al-4V alloy with equiaxed and bimodal microstructures. IOP Conf. Ser. Mater. Sci. Eng. 2017, 219, 012013. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Bo and Md stability phase map showing the positions of the designed alloys [26].
Figure 1. Bo and Md stability phase map showing the positions of the designed alloys [26].
Applsci 12 07257 g001
Figure 2. Optical micrographs of (a) TMF 34P/M, (b) TMF 54 P/M, (c) TMF 34 I/M, and (d) TMF 54 I/M.
Figure 2. Optical micrographs of (a) TMF 34P/M, (b) TMF 54 P/M, (c) TMF 34 I/M, and (d) TMF 54 I/M.
Applsci 12 07257 g002
Figure 3. Inverse pole figure (IPF) maps of (a) TMF 34 P/M and (b) TMF 54 P/M.
Figure 3. Inverse pole figure (IPF) maps of (a) TMF 34 P/M and (b) TMF 54 P/M.
Applsci 12 07257 g003
Figure 4. EBSD analysis of phase map of (a) TMF 34 P/M and (b) TMF 54 P/M.
Figure 4. EBSD analysis of phase map of (a) TMF 34 P/M and (b) TMF 54 P/M.
Applsci 12 07257 g004
Figure 5. Distribution of pore size and pore amount in (a) TMF 34 P/M and (b) TMF 54 P/M alloys.
Figure 5. Distribution of pore size and pore amount in (a) TMF 34 P/M and (b) TMF 54 P/M alloys.
Applsci 12 07257 g005
Figure 6. IPF maps of prior β grains reconstructed by misorientation criterion and denoising method (a) TMF 34 P/M and (b) TMF 54 P/M.
Figure 6. IPF maps of prior β grains reconstructed by misorientation criterion and denoising method (a) TMF 34 P/M and (b) TMF 54 P/M.
Applsci 12 07257 g006
Figure 7. Vickers hardness of TMF 34 and TMF 54 alloys.
Figure 7. Vickers hardness of TMF 34 and TMF 54 alloys.
Applsci 12 07257 g007
Figure 8. Tensile stress–strain curve of TMF 34 and TMF 54 alloys.
Figure 8. Tensile stress–strain curve of TMF 34 and TMF 54 alloys.
Applsci 12 07257 g008
Figure 9. SEM images of the fractured surface of (a) TMF 34 P/M, (b) TMF 54 P/M, (c) TMF 34 I/M, and (d) TMF 54 I/M.
Figure 9. SEM images of the fractured surface of (a) TMF 34 P/M, (b) TMF 54 P/M, (c) TMF 34 I/M, and (d) TMF 54 I/M.
Applsci 12 07257 g009
Table 1. Tensile properties of TMF 34 and TMF 54 alloys.
Table 1. Tensile properties of TMF 34 and TMF 54 alloys.
Yield Strength (MPa)Ultimate Strength (MPa)Elongation (%)
TMF 34 P/M-692.690.60
TMF 54 P/M-815.270.79
TMF 34 I/M-877.150.82
TMF 54 I/M821.29942.514.20
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hwang, H.-W.; Park, J.-H.; Lee, D.-G. Effect of Molybdenum Content on Microstructure and Mechanical Properties of Ti-Mo-Fe Alloys by Powder Metallurgy. Appl. Sci. 2022, 12, 7257. https://doi.org/10.3390/app12147257

AMA Style

Hwang H-W, Park J-H, Lee D-G. Effect of Molybdenum Content on Microstructure and Mechanical Properties of Ti-Mo-Fe Alloys by Powder Metallurgy. Applied Sciences. 2022; 12(14):7257. https://doi.org/10.3390/app12147257

Chicago/Turabian Style

Hwang, Hyo-Woon, Ji-Hwan Park, and Dong-Geun Lee. 2022. "Effect of Molybdenum Content on Microstructure and Mechanical Properties of Ti-Mo-Fe Alloys by Powder Metallurgy" Applied Sciences 12, no. 14: 7257. https://doi.org/10.3390/app12147257

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