Next Article in Journal
Carbonaceous Decomposition Products at High Temperatures and Their Interfacial Role in the Friction Behaviour of Composite Brake Material
Previous Article in Journal
Lubrication Characteristics of a Warhead-Type Irregular Symmetric Texture on the Stator Rubber Surfaces of Screw Pumps
Previous Article in Special Issue
Hard-Anodized Aluminum Alloy: Wear Properties in Vegetable Oils
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Transformation from D022 to L12 in Al3Ti by Fe Addition for Enhanced Wear Resistance

1
Department of Chemical and Materials Engineering, University of Alberta, Edmonton, AB T6G 2H5, Canada
2
School of Mechatronic Engineering, China University of Mining and Technology, Xuzhou 221116, China
3
School of Optoelectronic Engineering, Chongqing University of Posts and Telecommunications, Chongwen Road, Nanan District, Chongqing 400065, China
4
School of Machinery and Transportation, Southwest Forestry University, Kunming 650224, China
*
Author to whom correspondence should be addressed.
Lubricants 2024, 12(11), 398; https://doi.org/10.3390/lubricants12110398
Submission received: 12 October 2024 / Revised: 13 November 2024 / Accepted: 15 November 2024 / Published: 19 November 2024
(This article belongs to the Special Issue Friction and Wear of Alloys)

Abstract

:
The addition of third elements may help transform brittle D022-structured lightweight Al3Ti to a relatively ductile L12-structured (Al, M)3Ti (where M represents the third elements), thus increasing the ductility at the expense of hardness. Such a transformation could benefit the wear resistance of the alloy due to improved toughness if a proper balance between the hardness and ductility is achieved. In this work, a D022-predominant Al3Ti alloy (S-Al3Ti) and an L12-predominant (Al, Fe)3Ti alloy (S-Al67Ti25Fe8) were fabricated by arc melting. Change in wear resistance, corresponding to a D022-to-L12 transformation, caused by the addition of Fe as a representative third element, was investigated and compared with the wear resistance of a commercial Al-matrix composite reinforced by 30 wt.% SiC particles (S-Al/SiCp) as a reference material. It was observed that wear of the S-Al3Ti resulted from abrasion involving synergistic oxidation, leading to a larger volume loss. In contrast, the softer S-Al67Ti25Fe8 showed enhanced wear resistance, benefiting from improved toughness with reasonable hardness. During the wear testing, both the alloys exhibited better performance than S-Al/SiCp, a well-known lightweight composite. This study highlights that D022-to-L12 transformation enhances wear resistance due to increased toughness which can be adjusted using the addition of a third element.

1. Introduction

Al-Ti intermetallic compounds are very attractive candidates for the structural components used at elevated temperatures due to their low density, high melting point and high specific strength [1,2,3]. The long-range ordering in the intermetallic compounds effectively impedes atomic diffusion and pin dislocations, leading to a high resistance to softening at elevated temperatures, compared to solid-solution alloys [4,5]. According to the Al-Ti binary phase diagram, four Al-Ti intermetallic compounds, i.e., AlTi, Al2Ti, Al3Ti and Ti3Al, are promising for high-temperature applications [6], and Al3Ti is the most appealing due to its advantageous density [7].
Binary Al3Ti may have two commonly encountered crystal structures; one is a stable tetragonal D022 structure with a certain degree of anisotropy due to its low symmetry, and the other is a metastable cubic L12 structure with relatively isotropic properties due to its higher symmetry [8]. Figure 1 illustrates the crystal structures of D022 Al3Ti and L12 Al3Ti. Yamaguchi et al. [9] reported that the primary deformation mode of D022 Al3Ti during compression below 620 °C is <111> [112] twinning, with a dislocation slip activated at higher temperatures, resulting in improved ductility. The brittleness of the D022 structure at room temperature, because of its insufficient slip systems [10], remains a significant challenge for practical applications of Al3Ti [11].
The D022 structure is more ionic with stronger bonding than that of L12 structure. In the D022 structure, Ti atoms gain charges and Al atoms lose charges [12]. Alloying Al3Ti with a small amount of one or multiple transition metal elements, such as Cr, Mn, Fe, Ni, Cu, and Ag [13,14,15,16,17,18,19,20,21], can transform the tetragonal D022 structure into the cubic L12 structure. These less electropositive third elements may replace Al, taking electrons from Al and suppressing the charge transfer from Al to Ti, which stabilizes the L12 structure [22,23]. The cubic L12 structure provides more slip systems for plastic deformation, thus enhancing the ductility. The D022-to-L12 transformation can help achieve improved plasticity, which is beneficial to practical applications of Al3Ti. In general, the improvement in plasticity is achieved at the expense of strength and hardness, which may lead to negative influence on some properties such as the resistance to wear [24]. A theoretical study on tetragonal-D022 and cubic-L12 structures via calculations with density functional theory by Boulechfar et al. [25] demonstrated that the L12 phase showed lowered brittleness and hardness as well, compared to the D022 structure. Milman et al. [26] compared the hardness of tetragonal Al3Ti with that of cubic L12 Cr-doped Al61Cr12Ti27 and Mn-doped Al66Mn11Ti23, and observed that the hardness decreased by about 40% with Cr doping and 60% with Mn doping, respectively.
Wear, as one of the material failure processes, occurs when two surfaces rub each other, causing surface damage or failure. Estimates show that up to 80% of machine parts failure involves wear [27]. The economic loss caused by friction and wear reaches 2~7% of the gross domestic product for industrial countries [27]. Wear is a complex failure process, influenced by many factors, e.g., environmental aggressiveness, the wear loading condition, and microstructure modification, etc. [28,29,30,31,32], which could be mitigated using different approaches, including alloying, microstructure engineering, heat treatment, and texture engineering, etc. [33,34,35,36,37]. Mechanical properties, such as properly balanced hardness/strength and ductility/toughness, are indispensable to desirable wear resistance. Given the improved ductility and reduced strength/hardness by the transformation from the hard but brittle D022 Al3Ti to the ductile L12 (Al, M)3Ti (where M represents ternary elements), the corresponding wear resistance is accordingly affected, which should be adjustable by tuning the balance between hardness and ductility or toughness. However, there are few studies on the wear behaviors of D022 Al3Ti and L12 (Al, M)3Ti reported in the literature. It is, thus, of importance to investigate how the wear resistance is affected by the phase transformation.
The objectives of this study are (1) to investigate the wear behavior of tetragonal D022 Al3Ti and that of cubic L12 (Al, M)3Ti with improved ductility but reduced hardness, and (2) to reveal the underlying mechanism for corresponding variations in microstructure and relevant properties in order to further improve or optimize the alloy system. For details, the following tasks were undertaken in this study:
(a)
Fe was selected as a representative third element to be alloyed into Al3Ti, turning the D022-structured Al3Ti to a ductile L12 structure (Al, M)3Ti. Fe was selected for this study based on the consideration of its abundant resource, low cost, and the industrial relevance or interest in ferrous alloys. It should be indicated that adding Fe would not affect the alloy’s resistance to corrosion, since the alloy contains passive elements Al and Ti of high concentrations, which generate a protective passive film that can prevent continuous corrosion when used in corrosive environments;
(b)
The wear resistances of the two fabricated alloys were evaluated, in comparison to that of a well-known 30 wt.% SiC particle-reinforced Al-matrix composite as a reference material;
(c)
Corresponding changes in the microstructure, mechanical properties, and wear behavior were analyzed in detail in order to elucidate the underlying wear mechanisms for further improvement or optimization.

2. Materials and Methods

Two alloys, with nominal compositions of Al3Ti and Al67Ti25Fe8 (in molar ratio or at.%, denoted as S-Al3Ti and S-Al67Ti25Fe8, respectively), were fabricated using an arc melting process. High-purity Al, Ti and Fe powders (≥99.9%) were mixed and pressed into disk-like bulks, followed by arc melting in a water-cooled copper hearth using an arc melting furnace (MRF Inc., Allenstown, NH, USA) in a protective argon atmosphere. The sample preparation procedures are pictorially illustrated in Figure 2. Both of the samples were re-melted four times to minimize the chemical inhomogeneity.
The crystal structure and phases of the alloys were analyzed using an X-ray diffractometer (Rigaku XRD Ultima IV, Tokyo, Japan) at a scanning speed of 4°/min. The microstructure was characterized using a scanning electron microscope (Zeiss EVO M10 SEM, Carl Zeiss AG, Oberkochen, Germany) equipped with energy-dispersive X-ray spectroscopy (EDS) after mechanically grinding and polishing.
The macro-hardness of the alloys was measured using a ZHR Rockwell hardness tester (ZwickRoell, Kennesaw, GA, USA) with a conical diamond indenter under a load of 60 kgf at room temperature. Each polished sample was tested by indentation in 5 randomly selected positions to obtain the average hardness, and error bars were then determined based on the standard deviation. The indentation marks made by a diamond indenter with a square pyramid shape under a load of 15 kgf were observed to visually compare the toughness of the samples.
Sliding wear tests were performed using a ball-on-disk tribometer (CSEM Instruments, Peseux, Switzerland). A silicon nitride (Si3N4) ball with a diameter of 3 mm was used as the counter-face to wear the alloy samples at room temperature. The samples were ground and polished before the wear tests. The duration of each test was 30 min. The applied normal load was 3 N. The rotary speed was 191 rpm with a circular path of about 2 mm in diameter. The wear test for each sample was repeated three times to obtain the average value of wear volume loss. The volume losses from the wear tracks and the worn surface morphologies of the samples were automatically measured and analyzed, respectively, using a three-dimensional (3D) optical profilometer (ZeGage™ Pro, Zygo, Middlefield, CT, USA) with installed software (version 1.14.38, ZeMaps). The worn surfaces were further characterized using SEM with EDS.

3. Results

3.1. Microstructure Characterization

The XRD pattern of S-Al3Ti is displayed in Figure 3a. The main peaks of the D022 Al3Ti phase are accompanied by several small peaks of a second phase which is identified as pure Al, indicating that the D022 Al3Ti phase is predominant in the S-Al3Ti alloy. The microstructure of S-Al3Ti was examined by SEM in the back-scattered electron (BSE) mode, and is illustrated in Figure 3b,c. One may see that the second phase is distributed along grain boundaries, which accounts for approximately 3 vol.%. The EDS results shown in Figure 3d–f confirm the presence of Al as second phase, along with the predominant D022 Al3Ti phase.
The XRD pattern of S-Al67Ti25Fe8 is shown in Figure 4a. The main peaks come from the L12 phase, accompanied by several small peaks of second phase(s), indicating that S-Al67Ti25Fe8 was dominated by the L12 phase. The microstructure of S-Al67Ti25Fe8 observed with the SEM in BSE mode is shown in Figure 4b,c. One may see that there are two second phases: one is a light gray phase and the other is a bright one inside the light gray phase, which account for approximately 7 vol.% in total. The EDS results presented in Figure 4d–g confirm that the L12 predominant phase has a composition of (Al, Fe)3Ti, and show that the second phases are Fe-rich with depleted Ti.
In binary Al3Ti, the D022 structure is more stable, in which the charge transfers from Al to Ti, contributing to strong bonding strength. When the third element, which is less electropositive than Al, is added, it substitutes for Al and takes electrons from Al and consequently suppresses the charge transfer from Al to Ti. As a result, the L12 structure (Al,M)3Ti becomes more stable than D022 (Al,M)3Ti, i.e., the L12 structure is stabilized by adding third elements. More discussions with more details and references have been given in Section 4. The phase transformation from D022 to L12 certainly affects the mechanical properties, i.e., the ductility increases at the expense of hardness. The corresponding change in the balance between the hardness and ductility consequently influences the combination of hardness and toughness and, thus, the wear resistance. The resultant variations in wear resistance caused by the phase transformation are reported in the next section.
Figure 5 illustrates the microstructure of the 30 wt.% SiC particle-reinforced Al-matrix composite (denoted as S-Al/SiCp), which was used as a reference material when evaluating the wear resistance of the D022 Al3Ti-predominant S-Al3Ti alloy and the L12 (Al, Fe)3Ti-predominant S-Al67Ti25Fe8 alloy. As the SEM-BSE image in Figure 5a shows, dense particles in dark were distributed in the light gray matrix. According to the corresponding EDS maps shown in Figure 5b–d, the dark particles were SiC and the light gray matrix was Al. The SiC particles significantly hardened the Al matrix, playing an important role in contributing to the wear resistance of the S-Al/SiCp composite.

3.2. Wear Behavior

The Rockwell hardness values of three samples, S-Al3Ti, S-Al67Ti25Fe8, and S-Al/SiCp as a reference material, were measured at room temperature using a conical diamond indenter under a load of 60 kgf. Results of the tests are displayed in Figure 6. As shown, S-Al3Ti exhibited the highest hardness amongst the three samples. S-Al67Ti25Fe8 showed a slightly lower hardness, compared to that of S-Al3Ti, suggesting that the transformation of the D022 phase predominant in S-Al3Ti to L12 (Al, Fe)3Ti in S-Al67Ti25Fe8 did not decrease the hardness much. Although having densely distributed reinforcing SiC particles, the hardness of the S-Al/SiCp composite was markedly lower than those of the two alloys.
In this study, S-Al/SiCp was used as a reference material for the comparison purpose. S-Al3Ti and S-Al67Ti25Fe8 had relatively low densities (~3.17 and 3.48 g/cm3, respectively). The present study was conducted with the motivation to develop Al3Ti-based lightweight alloys for tribological applications. Since the commercial S-Al/SiCp possessed a good wear resistance with its density around 2.84 g/cm3, close to those of the above-mentioned alloys, selecting this well-known material as a reference material should facilitate this comparison with relatively quantitative information.
Indents caused by a diamond indenter with a square pyramid shape under a load of 15 kgf were observed under SEM in the second electron (SE) mode. The obtained images are displayed in Figure 7, showing that S-Al3Ti (Figure 7a) is the hardest, S-Al67Ti25Fe8 (Figure 7b) is less hard, and Al/SiCp (Figure 7c) is the softest among the three materials. Comparing the morphologies of the indents on S-Al3Ti and S-Al67Ti25Fe8, the zone around the indent on the former showed some wrinkles and micro-cracks, while the latter had much less these features, indicating that S-Al3Ti is brittle or less ductile, compared to S-Al67Ti25Fe8.
The volume losses of the samples after sliding wear testing under 3 N in the ambient environment for 30 min are shown in Figure 8a, and the wear track profiles are displayed in Figure 8b1–d2. In spite of the lower hardness, the alloy S-Al67Ti25Fe8 demonstrated a higher wear resistance than the harder alloy S-Al3Ti, which should benefit from its L12 (Al, Fe)3Ti phase with higher plasticity than that of the harder but less ductile D022 Al3Ti phase in S-Al3Ti. A combination of strength and ductility/plasticity can make a reasonably hard material tougher against wear attacks. The composite S-Al/SiCp showed the lowest wear resistance due to its lowest hardness among the three materials.
In order to elucidate the wear mechanisms for different performances of the three materials during wear tests, their worn surface morphologies were observed and analyzed, as illustrated in Figure 9 and Figure 10. S-Al3Ti showed its worn surface involving wear and oxidation caused by frictional heating [38,39], as evidenced by the detected oxygen distribution shown in the inset of Figure 9a1, and the oxide scale spallation shown in Figure 9a2,a3 (yellow arrow). In addition, surface fragmentation was also observed on S-Al3Ti near the area where a piece of oxide scale was removed, as shown in Figure 9a2 (magenta arrow), due to its lower plasticity or brittleness, compared to that of S-Al67Ti25Fe8. In contrast, the worn surface of S-Al67Ti25Fe8 showed more abrasive wear characteristics with scratching grooves as illustrated in Figure 9b1. Oxidation occurred on the surface as well, but it appeared to be much less, as one may see in the oxygen distribution shown in the inset of Figure 9b1. Very tiny craters were observed on or around the Fe-rich second phase (marked by yellow circles) connecting to the dominant L12 (Al, Fe)3Ti phase, as shown in Figure 9b2, which could be an indication of the brittleness of the Fe-rich Ti-depleted second phases in comparison with that of the L12 phase. Due to the mechanical difference between the second phase and the L12 phase, the local stress concentrations induced by incompatible deformations of the two phases under the wearing stress could induce local beak-off, which could trigger crack initiation. During the continuous wear process, micro-cracks could propagate and lead to eventual wear volume loss.
With higher hardness, S-Al3Ti was supposed to be subject to less abrasive wear damage when compared to S-Al67Ti25Fe8. However, the wear damage to S-Al3Ti is severer than that to S-Al67Ti25Fe8, as demonstrated by Figure 9a1,b1. This happened, as discussed earlier. The less protective oxide scale on S-Al3Ti with obvious spallation and surface fragmentation could make the alloy less resistant to wear attack. The oxide spall-off along with abraded material pieces of material may act as debris to result in additional wear damage [40,41]. Thus, the brittleness of the substrate, oxidation, and the spallation of the oxide scales could be responsible for the larger wear volume loss of S-Al3Ti, compared to that of S-Al67Ti25Fe8.
Regarding the wear of S-Al67Ti25Fe8 with decreased hardness and increased toughness, less oxidation was involved, evidenced by less oxygen contents inside and outside the wear track (see the inserted figure in Figure 9b1), and no oxide scale spallation was observed (see Figure 9b1,b2). The reduced brittleness with maintained reasonable hardness led to reduced abrasive wear. Consequently, S-Al67Ti25Fe8 showed a lower volume loss or higher wear resistance, compared to S-Al3Ti. Regarding oxidation, as shown in Figure 9b3, oxygen was still detected, but oxidation was markedly less, which can be ascribed to two possible reasons: (1) the oxide film was compacted and protective, so the oxide growth was slow, leading to the formation of a thin oxide film; and (2) the thin oxide film was removed during the wear testing, and re-growth was slow due to higher protectiveness of the oxide film. Bear in mind that a protective oxide film can form rapidly in the initial stage and the following growth becomes slow since the protective film more effectively hinders the diffusion of metallic atoms and oxygen, the continuous oxidation can thus be stopped. A protective oxide film helps reduce wear and minimize synergistic attacks by wear and oxidation. However, it is not very clear why the Fe addition improved the oxidation behavior or the protectiveness of the oxide scale on S-Al67Ti25Fe8. Analyzing the influences of Fe on the stability or cohesive energy, the adherence and mechanical properties of the oxide scale on the S-Al67Ti25Fe8 substrate would help understand the role that the Fe addition plays. These will be included in our follow-up studies.
It should also be mentioned that no sign of adhesive wear was noticed. In the obtained EDS maps, Si was not observed due to its insufficient content, implying that the adhesive transfer from the Si3N4 ball to the sample under study was negligible. There were also no noticeable patches of material peel-off on the alloy samples, i.e., alloy transfer. Thus, adhesive wear should have little contribution to wear in the present case, and abrasive wear was the dominant wear mechanism with influences of synergistical oxidation for both S-Al3Ti and S-Al67Ti25Fe8 at different extents.
The worn surface of S-Al/SiCp, as a reference material, which exhibited the lowest wear resistance among the three samples due to its lower hardness, is illustrated in Figure 10. The characteristics of abrasive wear with obvious plastic deformation were observed as shown in Figure 10a. The enlarged image in Figure 10b reveals dense wear debris, abrasive grooves, and some adhesive patches. During the wear test, the soft Al matrix experienced severe plastic deformation under the wearing force, forming flakes that adhered to the surface during repeated sliding when oxidation was also involved, as evidenced by the oxygen distribution shown in the inset of Figure 10a. In S-Al/SiCp, the Al matrix may not be strong enough to effectively maintain the reinforcing SiC particles during the wear process. Thus, SiC particles may be either detached from the Al matrix or extruded along with the plastically deformed Al matrix, as illustrated in Figure 10c,d, thereby failing to effectively protect the surface from wear attack. As a result, S-Al/SiCp showed considerably larger volume loss and lower wear resistance than both S-Al3Ti and S-Al67Ti25Fe8.

4. Discussion

The D022 structure is proven to be the most stable structure of stoichiometric Al3Ti alloy, while the L12 structure is a metastable structure which may be obtained under non-equilibrium processing condition, such as rapid solidification [42,43] and mechanical alloying [44,45]. The metastable L12 structure prepared by mechanical alloying may transform (through D023 structure) to stable D022 structure by heating processes [44,46]. First-principals calculations suggest that, in the D022 structure, Ti atoms gain charges and Al atoms lose charges, while in L12 structure, the situation is opposite [12]. As indicated earlier, adding an appropriate amount of less electropositive third elements such as Fe and Cr can replace Al, take electrons from Al, and consequently suppress the charge transfer from Al to Ti, thus stabilizing the cubic L12 structure and leading to the transformation of the D022 structure to the L12 one [22,23]. The resultant L12 structure keeps stable in an as-cast and homogenized state [16], indicating that adding proper third elements is an effective approach to obtain a stable L12 structure near stoichiometric Al3Ti composition, leading to improved plasticity. The above points have also been confirmed by the present study on the modification of Al3Ti with Fe addition for enhanced wear resistance.
For dry sliding wear of materials in the ambient environment, mechanical properties, such as strength and ductility that, along with hardness, determine the toughness of a material, are considered to be the most important factors which govern the wear resistance. The D022-to-L12 transformation from S-Al3Ti to S-Al67Ti25Fe8 improved the ductility (revealed by indentation marks in Figure 7) without losing much hardness (see Figure 6), leading to enhanced wear resistance, as shown in Figure 8. In addition to Fe, other elements such as Cr, Mn, Ni, Cu, and Ag can also stabilize L12 (Al, M)3Ti due to the difference in electronegativity between the constituent atoms, and the energy difference between the D022 and L12 structures [22,47]. Although little research on the wear behavior of D022 Al3Ti and L12 (Al, M)3Ti is reported, the beneficial effects of these elements on the wear resistance of Al3Ti can be expected based on the present study on the role of Fe addition in improving the wear resistance of the Al3Ti alloy.
In this work, both S-Al3Ti and S-Al67Ti25Fe8 were subject to surface abrasion and oxidation to different extents. Although S-Al3Ti is harder than S-Al67Ti25Fe8, it showed lower wear resistance than the latter, ascribed to the fact that its higher brittleness and formation of a less protective oxide scale resulted in more wear damage, as described earlier [48,49]. It should also be mentioned that a brittle substrate may deteriorate the adherence of its oxide scale due to lowered flexibility that may elevate the interfacial stress if the chemical interaction and lattice mismatch between the oxide scale and substrate remain the same or similar, thus accelerating the oxide scale spallation and promoting the wear-oxidation synergy, leading to more wear for the brittle or less tough S-Al3Ti, compared to that of the more plastic S-Al67Ti25Fe8. In addition, the oxide scale spall-off mixed with wear debris could result in extra wear damage [41].

5. Conclusions

In this work, two Al3Ti-based alloys, S-Al3Ti containing ~97 vol.% of D022 Al3Ti phase, and S-Al67Ti25Fe8 containing ~93 vol.% of L12 (Al, Fe)3Ti phase, were fabricated using an arc melting process, with the main objective being to investigate how the wear resistance of Al3Ti was influenced by the transformation from the D022 to the L12 structure caused by alloying with a third element, Fe. Their wear resistances were evaluated and compared with a commercial Al-matrix composite reinforced by 30 wt.% SiC particles (S-Al/SiCp) as a reference material. The following conclusions were drawn:
(a)
Alloying Al3Ti with Fe, which may partially substitute Al, successfully transformed the D022 Al3Ti to L12 (Al, Fe)3Ti, leading to improved ductility or toughness at the expense of hardness;
(b)
S-Al3Ti experienced oxide scale spallation and surface fragmentation due to its brittleness, resulting in higher wear volume loss;
(c)
S-Al67Ti25Fe8 exhibited lowered abrasive wear due to a reasonable combination of hardness and ductility/toughness, thus increasing its wear resistance;
(d)
Both the S-Al3Ti and S-Al67Ti25Fe8 alloys exhibited higher wear resistances than the commercial S-Al/SiCp composite. The (Al, M)3Ti alloy series can be further optimized and is promising as lightweight materials for tribological applications.
The reported in this article is a preliminary study showing the potential anti-wear or tribological applications of lightweight Al3Ti-based L12-structure material fabricated by arc melting. In the follow-up work, efforts will be made to further improve or optimize the base alloy by adding several third alloying elements with appropriate compositions to improve the toughness. In addition, potential combinations of L12-structure Al3Ti-based lightweight material with other materials such as carbides as the reinforcing phases should be explored to develop composites with a superior strength–toughness combination for tribological applications under specific or extreme conditions.

Author Contributions

Conceptualization, G.D. and D.L.; methodology, J.Y., A.H., A.K., R.F., A.V., W.C. and D.L.; validation, J.Y., A.H., D.Z., A.K., R.F., A.V., W.C. and D.L.; formal analysis, G.D.; investigation, G.D. and D.Z.; resources, R.F., A.V., W.C. and D.L.; data curation, A.H. and D.Z.; writing—original draft, G.D.; writing—review and editing, R.F., A.V., W.C. and D.L.; supervision, D.L.; project administration, D.L.; funding acquisition, W.C. and D.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science and Engineering Research Council of Canada (NSERC AMA ALLRP 567506-21 Li), NSERC-NRCan (ALLRP 586454-23), Trimay and Mitacs (MI MA IT29134 Kumar/Xu/Li), and the Academician Workstation in Yunnan Province (Grant No. 202305AF150019).

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. He, Z.Y.; Wang, Z.X.; Liu, X.P.; Han, P.D. Preparation of TiAl–Cr surface alloy by plasma-surface alloying technique. Vacuum 2013, 89, 280–284. [Google Scholar] [CrossRef]
  2. Uenishi, K.; Kobayashi, K.F. Processing of intermetallic compounds for structural applications at high temperature. Intermetallics 1996, 4, S95–S101. [Google Scholar] [CrossRef]
  3. Wei, N.; Han, X.; Zhang, X.; Cao, Y.; Guo, C.; Lu, Z.; Jiang, F. Characterization and properties of intermetallic Al3Ti alloy synthesized by reactive foil sintering in vacuum. J. Mater. Res. 2016, 31, 2706–2713. [Google Scholar] [CrossRef]
  4. Russell, A.M. Ductility in Intermetallic Compounds. Adv. Eng. Mater. 2003, 5, 629–639. [Google Scholar] [CrossRef]
  5. Bochenek, K.; Basista, M. Advances in processing of NiAl intermetallic alloys and composites for high temperature aerospace applications. Prog. Aerosp. Sci. 2015, 79, 136–146. [Google Scholar] [CrossRef]
  6. Li, Z.; Gao, W. High temperature corrosion of intermetallics. In Intermetallics Research Progress; Nova Science Publisher: Hauppauge, NY, USA, 2008; pp. 1–64. [Google Scholar]
  7. Nayak, S.S.; Murty, B.S. Synthesis and stability of L12–Al3Ti by mechanical alloying. Mater. Sci. Eng. A 2004, 367, 218–224. [Google Scholar] [CrossRef]
  8. Schoenitz, M.; Zhu, X.; Dreizin, E.L. Mechanical alloys in the Al-rich part of the Al-Ti binary system. J. Metastable Nanocryst. Mater. 2004, 20, 455–461. [Google Scholar] [CrossRef]
  9. Yamaguchi, M.; Umakoshi, Y.; Yamane, T. Plastic deformation of the intermetallic compound Al3Ti. Philos. Mag. A 1987, 55, 301–315. [Google Scholar] [CrossRef]
  10. Huang, Y.; Jiang, J. A Critical Review of von Mises Criterion for Compatible Deformation of Polycrystalline Materials. Crystals 2023, 13, 244. [Google Scholar] [CrossRef]
  11. Schwarz, R.B.; Desch, P.B.; Srinivasan, S. Phase Stability of Al3X Alloys (X = Ti, Zr, Hf). In Statics and Dynamics of Alloy Phase Transformations; Turchi, P.E.A., Gonis, A., Eds.; Springer: Boston, MA, USA, 1994; pp. 81–101. [Google Scholar]
  12. Hussain, A.; Mehmood, S.; Rasool, N.; Li, N.; Dharmawardhana, C.C. Electronic structure, mechanical and optical properties of TiAl3 (L12 & D022) via first-principles calculations. Chin. J. Phys. 2016, 54, 319–328. [Google Scholar] [CrossRef]
  13. Nakayama, Y.; Mabuchi, H. Formation of ternary L12 compounds in Al3Ti-base alloys. Intermetallics 1993, 1, 41–48. [Google Scholar] [CrossRef]
  14. Mabuchi, H.; Kito, A.; Nakamoto, M.; Tsuda, H.; Nakayama, Y. Effects of manganese on the L12 compound formation in Al3Ti-based alloys. Intermetallics 1996, 4, S193–S199. [Google Scholar] [CrossRef]
  15. Gengxiang, H.; Shipu, C.; Xiaohua, W.; Xiaofu, C. Plastic deformation and fracture behavior of a Fe-modified Al3Ti-base L12 intermetallic alloy. J. Mater. Res. 2011, 6, 957–963. [Google Scholar] [CrossRef]
  16. Mabuchi, H.; Tsuda, H.; Matsui, T.; Morii, K. Microstructure and mechanical properties of ternary L12 intermetallic compound in Al–Ti–Cr system. Mater. Trans. JIM 1997, 38, 560–565. [Google Scholar] [CrossRef]
  17. Prakash, U.; Buckley, R.A.; Jones, H.; Sellars, C.M. DO22 to L12 transition in intermetallic systems. J. Mater. Sci. 1992, 27, 2001–2004. [Google Scholar] [CrossRef]
  18. Varin, R.A.; Zbroniec, L.; Chiu, C. Observations of crack tip process zones in cubic titanium trialuminide intermetallics. Intermetallics 2001, 9, 937–941. [Google Scholar] [CrossRef]
  19. Tian, W.H.; Nemoto, M. Phase decomposition and hardening of Ag-modified L12 Al3Ti. Intermetallics 1998, 6, 193–200. [Google Scholar] [CrossRef]
  20. Heilmaier, M.; Handtrack, D.; Varin, R.A. Creep properties of boron-doped dual-phase Ti-rich L12-based trialuminide intermetallics. Scr. Mater. 2003, 48, 1409–1414. [Google Scholar] [CrossRef]
  21. Varin, R.A.; Zbroniec, L.; Wang, Z.G. Fracture toughness and yield strength of boron-doped, high (Ti+Mn) L12 titanium trialuminides. Intermetallics 2001, 9, 195–207. [Google Scholar] [CrossRef]
  22. George, E.P.; Pope, D.P.; Fu, C.L.; Schneibel, J.H. Deformation and Fracture of L12 Trialuminides. ISIJ Int. 1991, 31, 1063–1075. [Google Scholar] [CrossRef]
  23. Ghosh, G.; van de Walle, A.; Asta, M. First-Principles Phase Stability Calculations of Pseudobinary Alloys of (Al,Zn)3Ti with L12, D022, and D023 Structures. J. Phase Equilibria Diffus. 2007, 28, 9–22. [Google Scholar] [CrossRef]
  24. Xu, X.; Xu, W.; Ederveen, F.H.; van der Zwaag, S. Design of low hardness abrasion resistant steels. Wear 2013, 301, 89–93. [Google Scholar] [CrossRef]
  25. Boulechfar, R.; Sayad, D.; Khenioui, Y.; Meradji, H.; Ghemid, S.; Khenata, R.; Bin-Omran, S.; Bouhemadou, A.; Goumri-Said, S. Theoretical investigations of Zr-concentration influence on the thermodynamic, elastic, electronic, and structural stability of D022/L12-Al3Ti. Eur. Phys. J. B 2024, 97, 1. [Google Scholar] [CrossRef]
  26. Milman, Y.V.; Miracle, D.B.; Chugunova, S.I.; Voskoboinik, I.V.; Korzhova, N.P.; Legkaya, T.N.; Podrezov, Y.N. Mechanical behaviour of Al3Ti intermetallic and L12 phases on its basis. Intermetallics 2001, 9, 839–845. [Google Scholar] [CrossRef]
  27. Luo, J.; Liu, M.; Ma, L. Origin of friction and the new frictionless technology—Superlubricity: Advancements and future outlook. Nano Energy 2021, 86, 106092. [Google Scholar] [CrossRef]
  28. Xu, Z.; Li, D.Y.; Chen, D.L. Effect of Ti on the wear behavior of AlCoCrFeNi high-entropy alloy during unidirectional and bi-directional sliding wear processes. Wear 2021, 476, 203650. [Google Scholar] [CrossRef]
  29. Akonko, S.; Li, D.Y.; Ziomek-Moroz, M. Effects of cathodic protection on corrosive wear of 304 stainless steel. Tribol. Lett. 2005, 18, 405–410. [Google Scholar] [CrossRef]
  30. Yin, S.; Li, D.Y. Effects of prior cold work on corrosion and corrosive wear of copper in HNO3 and NaCl solutions. Mater. Sci. Eng. A 2005, 394, 266–276. [Google Scholar] [CrossRef]
  31. Faghihi, S.; Li, D.; Szpunar, J.A. Tribocorrosion behaviour of nanostructured titanium substrates processed by high-pressuretorsion. Nanotechnology 2010, 21, 485703. [Google Scholar] [CrossRef]
  32. Hutchings, I.; Shipway, P. Tribology: Friction and Wear of Engineering Materials; Butterworth-Heinemann: Oxford, UK, 2017. [Google Scholar]
  33. He, H.-B.; Han, W.-Q.; Li, H.-Y.; Li, D.-Y.; Yang, J.; Gu, T.; Deng, T. Effect of deep cryogenic treatment on machinability and wear mechanism of TiAlN coated tools during dry turning. Int. J. Precis. Eng. Manuf. 2014, 15, 655–660. [Google Scholar] [CrossRef]
  34. Budinski, K.G.; Budinski, S.T. Tribomaterials: Properties and Selection for Friction, Wear, and Erosion Applications; ASM International: Almere, The Netherlands, 2021. [Google Scholar]
  35. Tong, X.; Zhang, H.; Li, D.Y. Effect of annealing treatment on mechanical properties of nanocrystalline α-iron: An atomistic study. Sci. Rep. 2015, 5, 8459. [Google Scholar] [CrossRef] [PubMed]
  36. Li, D.Y.; Szpunar, J.A. A possible role for surface packing density in the formation of {111} texture in solidified FCC metals. J. Mater. Sci. Lett. 1994, 13, 1521–1523. [Google Scholar] [CrossRef]
  37. Bunge, H.J. Texture Analysis in Materials Science: Mathematical Methods; Elsevier: Amsterdam, The Netherlands, 2013. [Google Scholar]
  38. Knothe, K.; Liebelt, S. Determination of temperatures for sliding contact with applications for wheel-rail systems. Wear 1995, 189, 91–99. [Google Scholar] [CrossRef]
  39. Quinn, T.F.J. Oxidational wear modelling: I. Wear 1992, 153, 179–200. [Google Scholar] [CrossRef]
  40. Garza-Montes-de-Oca, N.F.; Rainforth, W.M. Wear mechanisms experienced by a work roll grade high speed steel under different environmental conditions. Wear 2009, 267, 441–448. [Google Scholar] [CrossRef]
  41. Cheng, X.; Jiang, Z.; Wei, D.; Wu, H.; Jiang, L. Adhesion, friction and wear analysis of a chromium oxide scale on a ferritic stainless steel. Wear 2019, 426–427, 1212–1221. [Google Scholar] [CrossRef]
  42. Majumdar, A.; Muddle, B.C. Microstructure in rapidly solidified Al-Ti alloys. Mater. Sci. Eng. A 1993, 169, 135–147. [Google Scholar] [CrossRef]
  43. Guo, J.Q.; Ohtera, K.; Kita, K.; Shibata, T.; Inoue, A.; Masumoto, T. New metastable phases in rapidly solidified Al-Zr and Al-Ti alloys with high solute contents. Mater. Sci. Eng. A 1994, 181–182, 1397–1404. [Google Scholar] [CrossRef]
  44. Srinivasan, S.; Desch, P.B.; Schwarz, R.B. Metastable phases in the Al3X (X = Ti, Zr, and Hf) intermetallic system. Scr. Metall. Mater. 1991, 25, 2513–2516. [Google Scholar] [CrossRef]
  45. Klassen, T.; Oehring, M.; Bormann, R. Microscopic mechanisms of metastable phase formation during ball milling of intermetallic TiAl phases. Acta Mater. 1997, 45, 3935–3948. [Google Scholar] [CrossRef]
  46. Colinet, C.; Pasturel, A. Ab initio calculation of the formation energies of L12, D022, D023 and one dimensional long period structures in TiAl3 compound. Intermetallics 2002, 10, 751–764. [Google Scholar] [CrossRef]
  47. Carlsson, A.E.; Meschter, P.J. Relative stability of LI2, DO22, and DO23 structures in MAl3 compounds. J. Mater. Res. 1989, 4, 1060–1063. [Google Scholar] [CrossRef]
  48. Qiu, X.W.; Zhang, Y.P.; Liu, C.G. Effect of Ti content on structure and properties of Al2CrFeNiCoCuTix high-entropy alloy coatings. J. Alloys Compd. 2014, 585, 282–286. [Google Scholar] [CrossRef]
  49. Li, Y.; Liang, H.; Nie, Q.; Qi, Z.; Deng, D.; Jiang, H.; Cao, Z. Microstructures and Wear Resistance of CoCrFeNi2V0.5Tix High-Entropy Alloy Coatings Prepared by Laser Cladding. Crystals 2020, 10, 352. [Google Scholar] [CrossRef]
Figure 1. Schematic illustrations of crystal structures of (a) D022 Al3Ti and (b) L12 Al3Ti.
Figure 1. Schematic illustrations of crystal structures of (a) D022 Al3Ti and (b) L12 Al3Ti.
Lubricants 12 00398 g001
Figure 2. The preparation procedures of S-Al3Ti and S-Al67Ti25Fe8.
Figure 2. The preparation procedures of S-Al3Ti and S-Al67Ti25Fe8.
Lubricants 12 00398 g002
Figure 3. Microstructure characterization for S-Al3Ti. (a) XRD pattern; (b,c) SEM-BSE images; (d,e) elemental distributions of the same area in (c); (f) elemental distribution along the blue line in (c).
Figure 3. Microstructure characterization for S-Al3Ti. (a) XRD pattern; (b,c) SEM-BSE images; (d,e) elemental distributions of the same area in (c); (f) elemental distribution along the blue line in (c).
Lubricants 12 00398 g003
Figure 4. Microstructure characterization for S-Al67Ti25Fe8. (a) XRD pattern; (b,c) SEM-BSE images; (df) elemental distributions of the same area shown in (c); (g) elemental distribution along the blue line shown in (c).
Figure 4. Microstructure characterization for S-Al67Ti25Fe8. (a) XRD pattern; (b,c) SEM-BSE images; (df) elemental distributions of the same area shown in (c); (g) elemental distribution along the blue line shown in (c).
Lubricants 12 00398 g004
Figure 5. Microstructure of S-Al/SiCp composite. (a) SEM-BSE image in secondary electron mode. (bd) Elemental distribution of the same area shown in (a).
Figure 5. Microstructure of S-Al/SiCp composite. (a) SEM-BSE image in secondary electron mode. (bd) Elemental distribution of the same area shown in (a).
Lubricants 12 00398 g005
Figure 6. Rockwell hardness using a conical diamond indenter under load of 60 kgf.
Figure 6. Rockwell hardness using a conical diamond indenter under load of 60 kgf.
Lubricants 12 00398 g006
Figure 7. Indents caused by a diamond indenter with square pyramid shape under load of 15 kgf: (a) S-Al3Ti. (b) S-Al67Ti25Fe8. (c) S-Al/SiCp.
Figure 7. Indents caused by a diamond indenter with square pyramid shape under load of 15 kgf: (a) S-Al3Ti. (b) S-Al67Ti25Fe8. (c) S-Al/SiCp.
Lubricants 12 00398 g007
Figure 8. Volume losses caused by wear under 3 N for 20 min at ambient. (a) Volume losses; 3D and cross-sectional wear track profiles of (b1,b2) S-Al3Ti, (c1,c2) S-Al67Ti25Fe8, and (d1,d2) S-Al/SiCp.
Figure 8. Volume losses caused by wear under 3 N for 20 min at ambient. (a) Volume losses; 3D and cross-sectional wear track profiles of (b1,b2) S-Al3Ti, (c1,c2) S-Al67Ti25Fe8, and (d1,d2) S-Al/SiCp.
Lubricants 12 00398 g008
Figure 9. SEM-SE images and EDS mapping analysis of the worn surfaces of (a1a5) S-Al3Ti and (b1b6) S-Al67Ti25Fe8. (a1,a2) SEM-SE images of worn surface morphology; (a3a5) O, Al, and Ti distribution of the same area in (a2). (b1,b2) SEM-SE images of worn surface morphology; (b3b6) O, Al, Ti, and Fe distribution of the same area in (b2). The insets in (a1,b1) are the oxygen distribution of the surface area in (a1,b1), respectively.
Figure 9. SEM-SE images and EDS mapping analysis of the worn surfaces of (a1a5) S-Al3Ti and (b1b6) S-Al67Ti25Fe8. (a1,a2) SEM-SE images of worn surface morphology; (a3a5) O, Al, and Ti distribution of the same area in (a2). (b1,b2) SEM-SE images of worn surface morphology; (b3b6) O, Al, Ti, and Fe distribution of the same area in (b2). The insets in (a1,b1) are the oxygen distribution of the surface area in (a1,b1), respectively.
Lubricants 12 00398 g009
Figure 10. SEM-SE images and EDS mapping of the worn surface of the S-Al/SiCp composite. (ac) Worn surface morphology with different magnifications; (d) silicon distribution of the same area in (c). The inset in (a) shows the oxygen distribution on the worn surface area marked by a blue rectangular frame.
Figure 10. SEM-SE images and EDS mapping of the worn surface of the S-Al/SiCp composite. (ac) Worn surface morphology with different magnifications; (d) silicon distribution of the same area in (c). The inset in (a) shows the oxygen distribution on the worn surface area marked by a blue rectangular frame.
Lubricants 12 00398 g010
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Diao, G.; Yuan, J.; He, A.; Zhang, D.; Kumar, A.; Fang, R.; Vorobyev, A.; Chen, W.; Li, D. Transformation from D022 to L12 in Al3Ti by Fe Addition for Enhanced Wear Resistance. Lubricants 2024, 12, 398. https://doi.org/10.3390/lubricants12110398

AMA Style

Diao G, Yuan J, He A, Zhang D, Kumar A, Fang R, Vorobyev A, Chen W, Li D. Transformation from D022 to L12 in Al3Ti by Fe Addition for Enhanced Wear Resistance. Lubricants. 2024; 12(11):398. https://doi.org/10.3390/lubricants12110398

Chicago/Turabian Style

Diao, Guijiang, Junfeng Yuan, Anqiang He, Dong Zhang, Aakash Kumar, Ranran Fang, Anatoliy Vorobyev, Wengang Chen, and Dongyang Li. 2024. "Transformation from D022 to L12 in Al3Ti by Fe Addition for Enhanced Wear Resistance" Lubricants 12, no. 11: 398. https://doi.org/10.3390/lubricants12110398

APA Style

Diao, G., Yuan, J., He, A., Zhang, D., Kumar, A., Fang, R., Vorobyev, A., Chen, W., & Li, D. (2024). Transformation from D022 to L12 in Al3Ti by Fe Addition for Enhanced Wear Resistance. Lubricants, 12(11), 398. https://doi.org/10.3390/lubricants12110398

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