Next Article in Journal
Structural Characteristics and Cementitious Behavior of Magnesium Slag in Comparison with Granulated Blast Furnace Slag
Previous Article in Journal
Stepwise Identification Method of Thermal Load for Box Structure Based on Deep Learning
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Effect of Nb Doping on the Properties of Ti-Al Intermetallic Compounds Using First-Principles Calculations

1
Shanghai Key Laboratory of D&A for Metal-Functional Materials, School of Materials Science & Engineering, Tongji University, Shanghai 201804, China
2
Biomaterials R&D Center, Zhuhai Institute of Advanced Technology, Chinese Academy of Sciences, Zhuhai 519000, China
3
Interdisciplinary Center for Additive Manufacturing, School of Materials and Chemistry, University of Shanghai for Science and Technology, Shanghai 200093, China
4
Department of Mechanical, Industrial and Mechatronics Engineering, Toronto Metropolitan University, Toronto, ON M5B 2K3, Canada
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Materials 2024, 17(2), 358; https://doi.org/10.3390/ma17020358
Submission received: 6 December 2023 / Revised: 6 January 2024 / Accepted: 8 January 2024 / Published: 11 January 2024

Abstract

:
The crystal structures, stability, mechanical properties and electronic structures of Nb-free and Nb-doped Ti-Al intermetallic compounds were investigated via first-principles calculations. Seven components and eleven crystal configurations were considered based on the phase diagram. The calculated results demonstrate that hP8-Ti3Al, tP4-TiAl, tP32-Ti3Al5, tI24-TiAl2, tI16-Ti5Al11, tI24-Ti2Al5, and tI32-TiAl3 are the most stable phases. Mechanical properties were estimated with the calculated elastic constants, as well as the bulk modulus, shear modulus, Young’s modulus, Poisson’s ratio and Pugh’s ratio following the Voigt–Reuss–Hill scheme. As the Al content increases, the mechanical strength increases but the ductility decreases in the Ti-Al compounds. This results from the enhanced covalent bond formed by the continuously enhanced Al-sp hybrid orbitals and Ti-3d orbitals. Nb doping (~5 at.% in this study) keeps the thermodynamical and mechanical stability for the Ti-Al compounds, which exhibit slightly higher bulk modulus and better ductility. This is attributed to the fact that the Nb 4d orbitals locate near the Fermi level and interact with the Ti-3d and Al-3p orbitals, improving the metallic bonds based on the electronic structures.

1. Introduction

Ti-Al alloys possess significant potential for applications due to their high strength, stiffness, hardness, thermal stability, and corrosion resistance (e.g., in Boeing 787) [1,2,3,4]. As the aviation industry continues to advance at a rapid pace, there arises a need to further enhance the strengthening and toughening capabilities of Ti-Al alloys [5,6,7]. Considerable efforts have been dedicated to the development of techniques such as heat treatment, thermomechanical processing, and alloying to effectively control the microstructure of Ti-Al alloys [8,9,10]. Notably, the introduction of specific solute elements, such as Nb, as dopants into Ti-Al binary intermetallic compounds, has been found to induce microstructural alterations and subsequent changes in the properties.
A total of seven components and eleven configurations of Ti-Al binary intermetallic compounds have been reported based on experimental and theoretical Ti-Al binary equilibrium phase diagrams [11,12]. However, only Ti3Al and TiAl are presently employed as base materials for Ti-Al alloys. Other intermetallic compounds, such as TiAl3 with a high aluminum content and low density, have limited usage due to poor ductility and fracture toughness, despite possessing superior resistance to high-temperature oxidation and specific strength. Ghosh and Asta [13] conducted a systematic investigation on the enthalpies of formation of Ti-Al binary intermetallic compounds with varying components and configurations, classifying these compounds into stable, sub-stable, and unstable states. Jian et al. [14] studied the stability, mechanical properties and electronic structures of Ti3Al, TiAl, TiAl2, and TiAl3 based on the first-principles calculation, concluding that with the increase in Al content, the bulk modulus, Poisson’s ratio and ductility decrease while the shear modulus, Young’s modulus and hardness gradually increase. Tang et al. [15] performed the first-principles calculation on the long-period superstructures Al5Ti2 and Al11Ti5 to examine the elastic properties and phonon focusing on electronic structures, reporting that both compounds are mechanistically anisotropic due to strong directional bonding between the Al and Ti atoms induced by strong hybridization between Al-3p and Ti-3d.
In Ti-Al alloys, Nb primarily exists in two forms: as a constituent of the third phase, such as Ti2AlNb, or as a dopant incorporated into the Ti-Al binary intermetallic compound. Chen et al. [16] investigated the mechanical properties and electronic structures of Nb-doped TiAl2, which is a metal-stable phase with a space group of CMMM, and found that Nb doping at Al sites improved ductility more than at Ti sites. Song et al. [17] used the discrete variational cluster method to calculate the compound electronic structure and binding energy to determine the preferred occupancy of various alloying elements in γ-TiAl, reporting that Nb preferentially occupies sites in the Ti sublattice. Based on the first-principles supercell calculations of the electronic structure and total energy, Wolf et al. [18] examined the site preference of Nb atoms in the γ-TiAl and observed that Nb predominantly occupies Ti sites. Recently, Lee et al. [19] investigated the point defect formation energies of the substitutional defects based on the first-principles calculations and found Nb prefers to locate at the Ti sites instead of the Al sites; also, Nb substitution at the Ti sites increases the yield strength of the alloy. As Liu et al. [20] reviewed, in experiments, the addition of Nb in a small amount (2 at.%) can increase the ductility and fracture toughness of γ-TiAl; as the content of Nb increases to 4–10 at.%, the hot workability and creep resistance of γ-TiAl at high temperatures improves.
Up to now, most reports on the Nb doping of Ti-Al intermetallic compounds focus on TiAl2 and TiAl. A comprehensive study on the Nb doping of Ti-Al intermetallic compounds is still lacking [21,22,23]. In this study, we investigated the crystal structure, stability, mechanical properties, and microscopic electronic structure of Nb-doped Ti-Al intermetallic compounds, including 7 components and 11 configurations based on the first-principles calculations. It is found that Nb doping enhances the compressibility of the Ti-Al compounds under hydrostatic pressure as well as their ductility. The partial density of states show that the Nb 4d orbitals locate near the Fermi level and interact with the Ti-3d and Al-3p orbitals, improving the metallic bonds and accounting for improving the mechanical properties.

2. Materials and Methods

The periodic density functional theory (DFT) calculations were performed using the plane-wave pseudo-potential Vienna ab initio simulation package (VASP.5.4.4, Vienna, Austria) [24]. The generalized gradient approximation as formulated by Perdew, Burke and Ernzerhof (GGA-PBE) was employed for the exchange-correlation functional [25]. The projection enhanced wave (PAW) method proposed by Blöchl and implemented by Kresse and Joubert was used with a cutoff energy of 420 eV [26]. A uniform mesh grid with a spacing of 0.03 Å was used to sample the complete Brillouin zone and calculate the density of states [27]. Brillouin zone integrations were carried out with the Methfessel–Paxton technique with a 0.1 eV smearing of the electron levels [28]. The PAW pseudopotentials considered were Ti 3p63d24s2, Al 2s23p1, and Nb 4p64d45s1. The full relaxation structure optimization method was used to obtain the ground-state crystal structure of each compound. The total energy convergence parameter during optimization was 2 × 10−6 eV/atom, the Hermann–Feynman force convergence parameter was 0.01 eV/Å, the tolerance shift was less than 0.002 Å, and the stress deviation per atom was less than 0.1 GPa. In addition, the single-crystal elastic matrix constants of the compounds in the ground-state structures were calculated using the stress–strain method according to the generalized Hooke’s law.
According to experimental and theoretical Ti-Al binary equilibrium phase diagrams [11,12], there are seven chemical compositions and 11 phases for Ti-Al intermetallic compounds, i.e., hP8-Ti3Al, tP4-TiAl, cP2-TiAl, tP32-Ti3Al5, tI24-TiAl2, oC12-TiAl2, tI16-Ti5Al11, tP28-Ti2Al5, tI32-TiAl3, tI8-TiAl3, and cP4-TiAl3, with the person symbols written in front of the formula (Figure 1). Here, cP2-TiAl, oC12-TiAl2, and tI32-TiAl3 were evaluated using thermodynamics calculations [11]. All of the structures were considered in our calculations. Furthermore, we investigated the influences of Nb doping on the mechanical properties of the Ti-Al intermetallic compounds. About 5 at.% Nb atoms were considered in order to avoid too large a change in the crystal structures.

3. Results and Discussion

3.1. Crystal Structure and Phase Stability

Table 1 lists the calculated lattice parameters and zero-temperature formation energy (ΔHr) of all Ti-Al intermetallic compounds including the experimental and other theoretical results for comparison. The formation energy was calculated using
H r Ti x Al y = E tol Ti x Al y     x E coh Ti     y E coh Al x + y
where Etol(TixAly) is the total energy of TixAly (f.u.), and Ecoh(Ti) and Ecoh(Al) are the cohesive energy of the Ti and Al crystals, respectively, which is the difference between the total energy of the Ti/Al crystal and the energy of a single Ti/Al atom [29]. As shown in Table 1, the calculated lattice parameters are quite consistent with the experimental values, with the largest difference of 2.9% for tI8-TiAl3. Compared to the available Δ H r in our experiments (hP8-Ti3Al, tP4-TiAl, and cP4-TiAl3), our results are still consistent with the maximum error of 8% for hP8-Ti3Al. In addition, our results perfectly match other theoretical values using the same PAW-PBE method, showing only a slight difference with other theoretical methods.
Table 1 displays that the most stable structures are tP4 (P4/mmm), tI24 (I41/amd), and tI32 (I4/mmm), for the multiphase TiAl, TiAl2, and TiAl3, respectively. The cP2-TiAl is less stable than the tP4 phase with a formation energy of 13 kJ/mol higher, which is the B2 phase at high temperatures [12]. The oC12-TiAl2, which was reported to be metastable [39], has a formation energy 0.4 kJ/mol higher than the tI24 structure. For TiAl3, cP4 is the most unstable phase, with a much higher formation energy, while tI8 has a formation energy very close to the isomorphic tI32 phase. The most stable phases of tP4-TiAl, tI24-TiAl2, and tI32-TiAl3 were chosen for the subsequent calculations of the mechanical properties and Nb doping.
Furthermore, we performed the Nb-doping calculations for the Ti-Al intermetallic compounds in the most stable phases, including hp8-Ti3Al, tP4-TiAl, tP32-Ti3Al5, tI24-TiAl2, tI16-Ti5Al11, tP28-Ti2Al5, and tI32-TiAl3. A 5% Nb atomic content was considered in order to keep the crystal structures nearly unchanged. Nb atoms were set to occupy Ti sites based on previous reports [17,18]. There were two kinds of components considered for Nb-doped hp8-Ti3Al, i.e., unit-cell components of Ti11Al4Nb (6.25 at.% Nb, 1 × 1 × 2 supercell of hp8-Ti3Al) and Ti23Al8Nb (3.125 at.% Nb, 2 × 2 × 1 supercell hp8-Ti3Al). For tP4-TiAl and tI24-TiAl2, there was only one kind of Nb doping, with the unit-cell components of Ti15Al16Nb (3.125 at.% Nb, 2 × 2 × 2 supercell of tP4-TiAl) and Ti7Al16Nb (4.167 at.% Nb), respectively. For tP32-Ti3Al5 (Ti11Al20Nb, 4.545 at.% Nb), tP28-Ti2Al5 (Ti7Al20Nb, 3.57 at.% Nb), tI16-Ti5Al11 (Ti4Al11Nb, 6.25 at.%), and tI32-TiAl3 (Ti7Al24Nb, 3.125 at.% Nb), there are 2, 5, 3, and 2 kinds of crystallographic sites for Nb doping, respectively, considering the coordination environment of Nb atoms and local symmetry. All of the structures are listed in Figure 2 and are named as follows: person symbol-TixAly-Nb-number.
Table 2 presents the calculated lattice constants and formation energies for all the Nb-doped Ti-Al intermetallic compounds, including the percentage change in the crystal structure parameters relative to the non-doped ones. We found that ~5 at.% Nb doping does not change the crystal shape, and the volume change remains within a very small range (<0.35%). The negative formation energies indicate that the Nb-doped Ti-Al compounds are thermodynamically stable. However, in all systems, only Nb-doped TiAl3 (tI32-TiAl3-Nb-2) has lower formation energies than the non-doped ones by ~0.13 kJ/mol. The formation energy of hp8-Ti3Al-Nb-2 with 3.125 at.% Nb are lower than that with 6.250 at.% Nb. Clearly, Nb doping is not conducive to the thermodynamical stability of the Ti-Al intermetallic compounds. For the same component, such as tI24-Ti2Al5-Nb (3.571 at.% Nb), tI24-Ti2Al5-Nb-1 and -2 have lower formation energies, in which Nb atoms occupy Ti-rich coordinated sites. With increasing Al content, the most stable Nb-doped Ti-Al phases are hp8-Ti3Al-Nb-2 (3.125 at.% Nb), tP4-TiAl-Nb-1 (3.125 at.% Nb), tP32-Ti3Al5-Nb-1 (4.545 at.% Nb), tI24-TiAl2-Nb (4.167 at.% Nb), tI16-Ti5Al11-Nb-3 (3.571 at.% Nb), tP28-Ti2Al5-Nb-1 (3.571 at.% Nb) and tI32-TiAl3-Nb-2 (3.125 at.% Nb), which will be used for the following calculations of the mechanical properties.

3.2. Mechanical Properties

Table 3 shows the elastic matrix constants of hp8-Ti3Al, tP4-TiAl, tP32-Ti3Al5, tI24-TiAl2, tI16-Ti5Al11, tP28-Ti2Al5 and tI32-TiAl3, and the most stable Nb-doped phases of hp8-Ti3Al-Nb-2, tP4-TiAl-Nb-1, tP32-Ti3Al5-Nb-1, tI24-TiAl2-Nb, tI16-Ti5Al11-Nb-3, tP28-Ti2Al5-Nb-1 and tI32-TiAl3-Nb-2. The stiffness-related elastic constants directly reflect the mechanical stability [29], and the elastic matrix constants in Table 3 meet the mechanical stability criteria [45,46,47]. Thus, the Ti-Al compounds and Nb-doped ones are mechanically stable. The tP4-TiAl values in Table 3 are consistent with the experimental report, with a difference of ~10%. Theoretically, the elastic matrix constants are sensitive to the initial calculation parameters in the first-principles calculations, such as the cutoff energy and K-points. For tP4-TiAl, our results, calculated with a cutoff energy of 420 eV, are closer to the values from the same method (PAW-GGA) with a cutoff energy of 450 eV [48]; however, they are somewhat higher than that those with a cutoff energy of 400 eV [49]. In general, our results are consistent with previous theoretical results. As shown in Table 3, when Nb atoms are introduced into the Ti-Al intermetallic compounds, there are some changes on the elastic matrix constants: (1) hP8-Ti3Al-Nb-2 exhibits smaller C11, C22, C44, C55, C66 values, but a larger C33, implying enhanced anisotropy; (2) tI32-TiAl3-Nb-2 possess increased elastic constants, in which C11, C22 and C33 increase over 5 GPa; (3) for all compounds, C33 increases, and even more than 5 GPa for tP28-Ti2Al5-Nb-1 and tI32-TiAl3-Nb-2.
Based on the elastic constants, the bulk modulus (B), shear modulus (G), Young’s modulus (E), Poisson’s ratio ( ν ), and Pugh’s ratio (K, B/G) were calculated using the Voigt–Reuss–Hill (VRH) scheme [52,53,54]. With the calculated bulk modulus and shear modulus, Vickers hardness (Hv) can be calculated according to the empirical formula proposed by Chen et al. [54]. The VRH approximation is known as the best method for the evaluation of the theoretical mechanical properties of polycrystalline materials, taking the value from the average of the Voigt and Reuss approximations [47,53]. In addition, the Debye temperature (ΘD) was evaluated in terms of the sound velocity [55,56]. All of the calculated results are shown in Table 4.
It is known that B reflects the compressibility of a solid under hydrostatic pressure, while G generally indicates the relationship between reversible deformation resistance and shear stress. E is defined as the ratio of stress to strain and is used to measure the stiffness of a material. A larger E means a higher stiffness with more covalent bond characteristics [57,58]. For Ti-Al compounds (Table 4), B decreases as the Al content increases, with the highest value of 116.09 GPa for Ti3Al and the lowest value of 106.86 GPa for TiAl3. G and E continually increase with increasing Al content. The Pugh’s ratio K (B/G) is normally used to reflect the ductility of a compound, with a critical value of 1.75, i.e., being brittle when K < 1.75 and ductile when K > 1.75 [59]. Likewise, Poisson’s ratio ν reflects the chemical bonding characteristics of compounds. Covalent bonds become weaker and metallic bonds become stronger as ν increases, with a critical value of 0.26 [59,60]. Obviously, for Ti-Al compounds, only Ti3Al has a K higher than 1.75 and ν larger than 0.26, showing good ductility and strong metallic bonds. As the Al content increases, ν/K reduces, indicating the presence of reinforcing covalent bonds. This is consistent with the results of Hv and ΘD (Table 4), both of which increase with increasing Al content.
As shown in Table 4, all Nb-doped Ti-Al compounds possess a larger B, in which Ti5Al11 has the largest D-value of 5 GPa. Obviously, Nb doping can strengthen the compressibility of Ti-Al compounds under hydrostatic pressure. After Nb doping, G and E show a non-monotonic change with increasing Al content, i.e., decreasing G and E for Ti3Al, a very small influence on TiAl, Ti3Al5 and TiAl2, and 2 (G) and 6 (E) GPa increase for Ti5Al11. As the Al content further increases, this increment decreases. It can be seen that Nb doping can increase the ductility of Ti-Al compounds, which is reflected in the increased Pugh’s ratio K and Poisson’s ratio ν. Obviously, Nb doping weakens the covalent bonds and strengthens the metallic bonds; thus, the ΘD and Hv of Nb-doped Ti-Al compounds become smaller than the non-doped ones. In addition, Nb doping has a greater influence on low-Al-content systems such as hP8-Ti3Al and tP4-TiAl, with the ν increment being about 0.01.
Figure 3a shows the three-dimensional plots of the Young’s modulus of hp8-Ti3Al, tP4-TiAl, tP32-Ti3Al5, tI24-TiAl2, tI16-Ti5Al11, tP28-Ti2Al5 and tI32-TiAl3. The plots of hp8-Ti3Al, tP4-TiAl, and tI32-TiAl3 are quite similar to previous theoretical reports [14]. The anisotropy of single-crystal structures usually originates from the directional properties of covalent bond. From the visual observation, hp8-Ti3Al seems to have a greater isotropic Young’s modulus. As the Al content increases, the Ti-Al compounds display a greater anisotropic Young’s modulus. As shown in Figure 3b, after Nb doping, the Young’s modulus anisotropy of hP8-Ti3Al-Nb-2 has a considerable change along the [100] and [010] directions, i.e., its anisotropy increases. However, the Young’s modulus of tI32-TiAl3-Nb-2 decreases slightly along the [100] and [010] directions, gently weakening its anisotropy. For other Ti-Al compounds, Nb doping has almost no influence on the anisotropy.

3.3. Electronic Structures

In order to gain an insight into the physical mechanisms, the calculations of the electronic structures were performed for the Ti-Al compounds of hp8-Ti3Al, tP4-TiAl, tP32-Ti3Al5, tI24-TiAl2, tI16-Ti5Al11, tP28-Ti2Al5 and tI32-TiAl3, and the most stable Nb-doped phases of hp8-Ti3Al-Nb-2, tP4-TiAl-Nb-1, tP32-Ti3Al5-Nb-1, tI24-TiAl2-Nb, tI16-Ti5Al11-Nb-3, tP28-Ti2Al5-Nb-1 and tI32-TiAl3-Nb-2. Figure 4 presents the total density of states (TDOSs) and the partial density of states (PDOSs). The TDOS displays a large distribution across the Fermi energy level (EF), indicating that all compounds show a metallic behavior. In addition, a pseudo-energy gap (a pronounced valley near EF) can be clearly observed from the TDOSs in Figure 4, which exists in the bonding and anti-bonding regions. The stability of a compound can be assessed based on the relative position of the EF and the pseudo-energy gap. When the EF lies to the right of the pseudo-energy gap, the electrons occupy the bonding region, indicating a stable structure. Conversely, when the EF lies to the left of the pseudo-energy gap, the electrons occupy the anti-bonding region, resulting in a less stable structure. The width of the pseudo-energy gap serves as an indicator of the strength of the covalent bond, and a wider gap suggests a stronger covalent interaction [61].
As shown in Figure 4a, the pseudo-energy gap width increases with increasing Al content, implying an enhancement of the covalent bond. Thus, hP8-Ti3Al has the lowest pseudo-energy gap, in agreement with its best ductility and lowest Debye temperature of 496 K (Table 4). The PDOSs in Figure 4a show that the Al 3s and 3p orbitals are almost completely separated in hP8-Ti3Al. Near the Fermi level, Al-3p and Ti-3d form strong metallic bonds, accounting for the good ductility of hP8-Ti3Al. As the Al content increases, the Al 3s orbitals widen and hybridize with the Al-3p orbitals; moreover, this hybridization gradually strengthens. The enhanced Al-sp-hybridizing orbitals form strong covalent bonds with the Ti-3d orbitals, accounting for the enhancing mechanical strength and lower ductility with increasing Al content. After Nb doping, the pseudo-energy gap width reduces (Figure 4b). This indicates that Nb doping weakens the character of the covalent bond, being consistent with the results of the Poisson’s ratio and Debye temperature, as shown in Table 4. The electronic structures display that for hP8-Ti3Al-Nb-2 and tp4-TiAl-Nb-1, in which Al-sp hybridization is weak, the Nb 4d orbitals locate near the Fermi level (>−4 eV) and interact with the Ti-3d and Al-3p orbitals, strengthening the metallic bonds. This is consistent with the result that Nb doping increases the Poisson’s ratio ν more significantly for hP8-Ti3Al and tP4-TiAl than for other Ti-Al intermetallic compounds. As the Al content increases, although some Nb 4d electrons participate in the formation of covalent bonds because of the enhanced Al-sp hybridization, the introduction of Nb 4d electrons improves the metallicity of the Ti-Al compounds.

4. Conclusions

The first-principles density functional theory (DFT) was employed to study the crystal structures, stability, mechanical properties, anisotropy, and electronic structures of Nb-free and Nb-doped Ti-Al intermetallic compounds, including seven components and eleven crystal configurations based on the phase diagrams. The calculated total energies reveal that hP8-Ti3Al, tP4-TiAl, tP32-Ti3Al5, tI24-TiAl2, tI16-Ti5Al11, tI24-Ti2Al5, and tI32-TiAl3 are the most stable phases. Mechanical properties were estimated using the calculated elastic constants, as well as the bulk modulus, shear modulus, Young’s modulus, Poisson’s ratio and Pugh’s ratio following the Voigt–Reuss–Hill scheme. As the Al content increases, the bulk, shear and Young’s modulus increase but the Poisson’s ratio decreases for Ti-Al compounds, indicating the strengthened mechanical properties and weakened ductility. This is due to the enhanced covalent bonds, which are formed by the continuously enhanced Al-sp hybrid orbitals and Ti-3d orbitals. Nb doping (~5 at.% used in this study) maintains thermodynamic and mechanical stability for the Ti-Al compounds. Moreover, Nb-doped tI32-TiAl3 has a lower formation enthalpy than the non-doped ones. The mechanical results show that Nb doping brings a slightly larger bulk modulus and better ductility for Ti-Al compounds. The electronic structures display that the Nb 4d orbitals locate near the Fermi level and interact with the Ti-3d and Al-3p orbitals, strengthening the metallic bonds in the Ti-Al compounds. Nb doping also increases the mechanical anisotropy of hP8-Ti3Al.

Author Contributions

Conceptualization, H.X. and A.F.; Methodology, L.X.; Software, K.W. and H.X.; Validation, H.X. and A.F.; Formal analysis, L.X. and S.Q.; Investigation, K.W., H.X. and A.F.; Resources, S.Q. and D.C.; Data curation, L.X. and H.W.; Writing—original draft, K.W. and H.X.; Writing—review & editing, A.F., S.Q., H.W. and D.C.; Visualization, H.W.; Supervision, D.C.; Project administration, A.F.; Funding acquisition, H.X. and S.Q. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundations of China (Grant Nos: 51971159, 52271012, and 51871168), Zhuhai Science Technology Department Project (2220004000049).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

D.C. would like to thank the Natural Sciences and Engineering Research Council of Canada for their financial support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zope, R.R.; Mishin, Y. Interatomic potentials for atomistic simulations of the Ti-Al system. Phys. Rev. B 2003, 68, 024102. [Google Scholar] [CrossRef]
  2. Ding, J.; Zhang, M.; Ye, T.; Liang, Y.; Ren, Y.; Dong, C.; Lin, J. Microstructure stability and micro-mechanical behavior of as-cast gamma-TiAl alloy during high-temperature low cycle fatigue. Acta Mater. 2018, 145, 504–515. [Google Scholar] [CrossRef]
  3. Yang, T.; Cao, B.X.; Zhang, T.L.; Zhao, Y.L.; Liu, W.H.; Kong, H.J.; Luan, J.H.; Kai, J.J.; Kuo, W.; Liu, C.T. Chemically complex intermetallic alloys: A new frontier for innovative structural materials. Mater. Today 2022, 52, 161–174. [Google Scholar] [CrossRef]
  4. Qu, S.J.; Tang, S.Q.; Feng, A.H.; Feng, C.; Shen, J.; Chen, D.L. Microstructural evolution and high-temperature oxidation mechanisms of a titanium aluminide based alloy. Acta Mater. 2018, 148, 300–310. [Google Scholar] [CrossRef]
  5. Cui, S.; Cui, C.; Xie, J.; Liu, S.; Shi, J. Carbon fibers coated with graphene reinforced TiAl alloy composite with high strength and toughness. Sci. Rep. 2018, 8, 2364. [Google Scholar] [CrossRef]
  6. Chen, J.; Chen, Q.; Qu, S.J.; Xiang, H.P.; Wang, C.; Gao, J.B.; Feng, A.H.; Chen, D.L. Oxidation mechanisms of an intermetallic alloy at high temperatures. Scr. Mater. 2021, 199, 113852. [Google Scholar] [CrossRef]
  7. Wu, G.D.; Cui, G.R.; Qu, S.J.; Feng, A.H.; Cao, G.J.; Ge, B.H.; Xiang, H.P.; Shen, J.; Chen, D.L. High-temperature oxidation mechanisms of nano-/submicro-scale lamellar structures in an intermetallic alloy. Scr. Mater. 2019, 171, 102–107. [Google Scholar] [CrossRef]
  8. Niu, H.Z.; Chen, Y.Y.; Xiao, S.L.; Xu, L.J. Microstructure evolution and mechanical properties of a novel beta γ-TiAl alloy. Intermetallics 2012, 31, 225–231. [Google Scholar] [CrossRef]
  9. Brotzu, A.; Felli, F.; Pilone, D. Effect of alloying elements on the behaviour of TiAl-based alloys. Intermetallics 2014, 54, 176–180. [Google Scholar] [CrossRef]
  10. Kim, S.-W.; Hong, J.K.; Na, Y.-S.; Yeom, J.-T.; Kim, S.E. Development of TiAl alloys with excellent mechanical properties and oxidation resistance. Mater. Des. 2014, 54, 814–819. [Google Scholar] [CrossRef]
  11. Murray, J.L. Calculation of the titanium-aluminum phase diagram. Metall. Trans. A 1988, 19, 243–247. [Google Scholar] [CrossRef]
  12. Schuster, J.C.; Palm, M. Reassessment of the binary Aluminum-Titanium phase diagram. J. Phase Equilibria Diffus. 2006, 27, 255–277. [Google Scholar] [CrossRef]
  13. Ghosh, G.; Asta, M. First-principles calculation of structural energetics of Al–TM (TM = Ti, Zr, Hf) intermetallics. Acta Mater. 2005, 53, 3225–3252. [Google Scholar] [CrossRef]
  14. Jian, Y.; Huang, Z.; Xing, J.; Sun, L.; Liu, Y.; Gao, P. Phase stability, mechanical properties and electronic structures of TiAl binary compounds by first principles calculations. Mater. Chem. Phys. 2019, 221, 311–321. [Google Scholar] [CrossRef]
  15. Tang, P.Y.; Huang, G.H.; Xie, Q.L.; Guan, N.N.; Ning, F. Elastic Properties, Phonon Focusing and Electronic Structures of Typical Long-Period Superstructures Al5Ti2 and Al11Ti5. Mater. Sci. Forum 2017, 898, 1026–1035. [Google Scholar] [CrossRef]
  16. Chen, H.; Li, X.; Chen, Z.; Zhang, R.; Ma, X.; Zheng, F.; Ma, Z.; Pan, F.; Lin, X. Investigation on electronic structures and mechanical properties of Nb–doped TiAl2 intermetallic compound. J. Alloys Compd. 2019, 780, 41–48. [Google Scholar] [CrossRef]
  17. Song, Y.; Yang, R.; Li, D.; Hu, Z.Q.; Guo, Z.X. A first principles study of the influence of alloying elements on TiAl: Site preference. Intermetallics 2000, 8, 563–568. [Google Scholar] [CrossRef]
  18. Wolf, W.; Podloucky, R.; Rogl, P.; Erschbaumer, H. Atomic modelling of Nb, V, Cr and Mn substitutions in γ-TiAl. 2: Electronic structure and site preference. Intermetallics 1996, 4, 201–209. [Google Scholar] [CrossRef]
  19. Lee, T.; Kim, S.-W.; Kim, J.Y.; Ko, W.-S.; Ryu, S. First-principles study of the ternary effects on the plasticity of γ-TiAl crystals. Sci. Rep. 2020, 10, 21614. [Google Scholar] [CrossRef]
  20. Liu, X.; Lin, Q.; Zhang, W.; Horne, C.V.; Cha, L. Microstructure Design and Its Effect on Mechanical Properties in Gamma Titanium Aluminides. Metals 2021, 11, 1644. [Google Scholar] [CrossRef]
  21. Zhang, Z.; Qu, S.; Cui, G.; Feng, A.; Shen, J.; Chen, D. A New Mechanism of Dynamic Phase Transformations in An Isothermal Forged Beta–Gamma Intermetallic Alloy. Materials 2019, 12, 2787. [Google Scholar] [CrossRef]
  22. Zhang, Y.; Feng, A.; Qu, S.; Shen, J.; Chen, D. Microstructure and low cycle fatigue of a Ti2AlNb-based lightweight alloy. J. Mater. Sci. Technol. 2020, 44, 140–147. [Google Scholar] [CrossRef]
  23. Xiang, J.M.; Mi, G.B.; Qu, S.J.; Huang, X.; Chen, Z.; Feng, A.H.; Shen, J.; Chen, D.L. Thermodynamic and microstructural study of Ti2AlNb oxides at 800  °C. Sci. Rep. 2018, 8, 12761. [Google Scholar] [CrossRef]
  24. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef] [PubMed]
  25. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  26. Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  27. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  28. Methfessel, M.; Paxton, A.T. High-precision sampling for Brillouin-zone integration in metals. Phys. Rev. B 1989, 40, 3616–3621. [Google Scholar] [CrossRef]
  29. Liu, Y.; Chong, X.; Jiang, Y.; Zhou, R.; Feng, J. Mechanical properties and electronic structures of Fe-Al intermetallic. Phys. B Condens. Matter 2017, 506, 1–11. [Google Scholar] [CrossRef]
  30. Asta, M.; de Fontaine, D.; van Schilfgaarde, M. First-principles study of phase stability of Ti–Al intermetallic compounds. J. Mater. Res. 1993, 8, 2554–2568. [Google Scholar] [CrossRef]
  31. Watson, R.E.; Weinert, M. Transition-metal aluminide formation: Ti, V, Fe, and Ni aluminides. Phys. Rev. B 1998, 58, 5981–5988. [Google Scholar] [CrossRef]
  32. Hayes, F.H. The Al-Ti-V (aluminum-titanium-vanadium) system. J. Phase Equilibria 1995, 16, 163–176. [Google Scholar] [CrossRef]
  33. Sridharan, S.; Nowotny, H.; Wayne, S.F. Investigations within the quaternary system titanium-nickel-aluminium-carbon. Monatshefte Chem. Chem. Mon. 1983, 114, 127–135. [Google Scholar] [CrossRef]
  34. Kubaschewski, O.; Dench, W.A. The heats of formation in the systems titanium-aluminium and titanium-iron. Acta Metall. 1955, 3, 339–346. [Google Scholar] [CrossRef]
  35. Kubaschewski, O.; Heymer, G. Heats of formation of transition-metal aluminides. Trans. Faraday Soc. 1960, 56, 473–478. [Google Scholar] [CrossRef]
  36. Nassik, M.; Chrifi-Alaoui, F.Z.; Mahdouk, K.; Gachon, J.C. Calorimetric study of the aluminium–titanium system. J. Alloys Compd. 2003, 350, 151–154. [Google Scholar] [CrossRef]
  37. Braun, J.; Ellner, M. Phase equilibria investigations on the aluminum-rich part of the binary system Ti-Al. Metall. Mater. Trans. A 2001, 32, 1037–1047. [Google Scholar] [CrossRef]
  38. Braun, J.; Ellner, M. X-ray high-temperature in situ investigation of the aluminide TiAl2 (HfGa2 type). J. Alloys Compd. 2000, 309, 118–122. [Google Scholar] [CrossRef]
  39. Schuster, J.; Ipser, H. Phases and phase relations in the partial system TiAl3-TiAl. Int. J. Mater. Res. 1990, 81, 389–396. [Google Scholar] [CrossRef]
  40. Raman, A.; Schubert, K. The constitution of some alloy series related to TiAl3. I. Investigations in some T4-Zn-Al, T4-Zn-Ga, and T4-Ga-Ge systems. Acta Metall. 1965, 56, 99–104. [Google Scholar]
  41. Menand, A.; Huguet, A.; Nérac-Partaix, A. Interstitial solubility in γ and α2 phases of TiAl-based alloys. Acta Mater. 1996, 44, 4729–4737. [Google Scholar] [CrossRef]
  42. Amador, C.; Hoyt, J.J.; Chakoumakos, B.C.; de Fontaine, D. Theoretical and Experimental Study of Relaxations in Al3Ti and Al3Zr Ordered Phases. Phys. Rev. Lett. 1995, 74, 4955–4958. [Google Scholar] [CrossRef]
  43. 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]
  44. Meschel, S.V.; Kleppa, O.J. The Standard Enthalpies of Formation of Some 3d Transition Metal Aluminides by High-Temperature Direct Synthesis Calorimetry. In Metallic Alloys: Experimental and Theoretical Perspectives; Faulkner, J.S., Jordan, R.G., Eds.; Springer: Dordrecht, The Netherlands, 1994; pp. 103–112. [Google Scholar]
  45. Patil, S.K.R.; Khare, S.V.; Tuttle, B.R.; Bording, J.K.; Kodambaka, S. Mechanical stability of possible structures of PtN investigated using first-principles calculations. Phys. Rev. B 2006, 73, 104118. [Google Scholar] [CrossRef]
  46. Nye, J.F.; Lindsay, R.B. Physical Properties of Crystals: Their Representation by Tensors and Matrices. Phys. Today 1957, 10, 26. [Google Scholar] [CrossRef]
  47. Wu, Z.-j.; Zhao, E.-j.; Xiang, H.-p.; Hao, X.-f.; Liu, X.-j.; Meng, J. Crystal structures and elastic properties of superhard IrN2 and IrN3 from first principles. Phys. Rev. B 2007, 76, 054115. [Google Scholar] [CrossRef]
  48. Hu, H.; Wu, X.; Wang, R.; Li, W.; Liu, Q. Phase stability, mechanical properties and electronic structure of TiAl alloying with W, Mo, Sc and Yb: First-principles study. J. Alloys Compd. 2016, 658, 689–696. [Google Scholar] [CrossRef]
  49. Junhua, T.; Kaijin, Z.; Junhui, P. First-Principles Simulation on Structure-Property of Ti-Al Intermetallic Compounds. Chin. J. Comput. Phys. 2017, 34, 365–373. [Google Scholar] [CrossRef]
  50. He, Y.; Schwarz, R.B.; Migliori, A.; Whang, S.H. Elastic constants of single crystal γ—TiAl. J. Mater. Res. 1995, 10, 1187–1195. [Google Scholar] [CrossRef]
  51. Tang, P.-Y.; Tang, B.-Y.; Su, X.-P. First-principles studies of typical long-period superstructures Al5Ti3, h-Al2Ti and r-Al2Ti in Al-rich TiAl alloys. Comput. Mater. Sci. 2011, 50, 1467–1476. [Google Scholar] [CrossRef]
  52. Anderson, O.L. A simplified method for calculating the debye temperature from elastic constants. J. Phys. Chem. Solids 1963, 24, 909–917. [Google Scholar] [CrossRef]
  53. Li, Y.; Gao, Y.; Xiao, B.; Min, T.; Fan, Z.; Ma, S.; Xu, L. Theoretical study on the stability, elasticity, hardness and electronic structures of W–C binary compounds. J. Alloys Compd. 2010, 502, 28–37. [Google Scholar] [CrossRef]
  54. Chen, X.-Q.; Niu, H.; Li, D.; Li, Y. Modeling hardness of polycrystalline materials and bulk metallic glasses. Intermetallics 2011, 19, 1275–1281. [Google Scholar] [CrossRef]
  55. Feng, J.; Xiao, B.; Zhou, R.; Pan, W.; Clarke, D.R. Anisotropic elastic and thermal properties of the double perovskite slab–rock salt layer Ln2SrAl2O7 (Ln = La, Nd, Sm, Eu, Gd or Dy) natural superlattice structure. Acta Mater. 2012, 60, 3380–3392. [Google Scholar] [CrossRef]
  56. Feng, J.; Xiao, B.; Chen, J.; Du, Y.; Yu, J.; Zhou, R. Stability, thermal and mechanical properties of PtxAly compounds. Mater. Des. 2011, 32, 3231–3239. [Google Scholar] [CrossRef]
  57. Shein, I.R.; Ivanovskii, A.L. Elastic properties of mono- and polycrystalline hexagonal AlB2-like diborides of s, p and d metals from first-principles calculations. J. Phys. Condens. Matter 2008, 20, 415218. [Google Scholar] [CrossRef]
  58. Young, A.F.; Sanloup, C.; Gregoryanz, E.; Scandolo, S.; Hemley, R.J.; Mao, H.-k. Synthesis of Novel Transition Metal Nitrides IrN2 and OsN2. Phys. Rev. Lett. 2006, 96, 155501. [Google Scholar] [CrossRef]
  59. Pugh, S.F. XCII. Relations between the elastic moduli and the plastic properties of polycrystalline pure metals. Lond. Edinb. Dublin Philos. Mag. J. Sci. 1954, 45, 823–843. [Google Scholar] [CrossRef]
  60. Lakes, R.; Wojciechowski, K.W. Negative compressibility, negative Poisson’s ratio, and stability. Phys. Status Solidi 2008, 245, 545–551. [Google Scholar] [CrossRef]
  61. Kirihara, K.; Nagata, T.; Kimura, K.; Kato, K.; Takata, M.; Nishibori, E.; Sakata, M. Covalent bonds and their crucial effects on pseudogap formation in α-Al(Mn, Re)Si icosahedral quasicrystalline approximant. Phys. Rev. B 2003, 68, 014205. [Google Scholar] [CrossRef]
Figure 1. The crystal structures of the Ti-Al intermetallic compounds: (a) hP8-Ti3Al, (b) tP4-TiAl, (c) cP2-TiAl, (d) tP32-Ti3Al5, (e) tI24-TiAl2, (f) oC12-TiAl2, (g) tI16-Ti5Al11, (h) tP28-Ti2Al5, (i) tI32-TiAl3, (j) tI8-TiAl3, (k) cP4-TiAl3. The person symbols are written in front of the formula, and the brown and orange balls represent Ti and Al, respectively.
Figure 1. The crystal structures of the Ti-Al intermetallic compounds: (a) hP8-Ti3Al, (b) tP4-TiAl, (c) cP2-TiAl, (d) tP32-Ti3Al5, (e) tI24-TiAl2, (f) oC12-TiAl2, (g) tI16-Ti5Al11, (h) tP28-Ti2Al5, (i) tI32-TiAl3, (j) tI8-TiAl3, (k) cP4-TiAl3. The person symbols are written in front of the formula, and the brown and orange balls represent Ti and Al, respectively.
Materials 17 00358 g001
Figure 2. The crystal structures of the Nb-doped Ti-Al binary intermetallic compounds: (a) hP8-Ti3Al-Nb, (b) tP4-TiAl-Nb, (c) tP32-Ti3Al5-Nb, (d) tI24-TiAl2-Nb, (e) tI16-Ti5Al11-Nb, (f) tI24-Ti2Al5-Nb, and (g) tI32-TiAl3-Nb. The numbers of (1)–(5) represent the Nb doping at different crystallographic sites; the brown, orange, and green balls represent Ti, Al, and Nb, respectively.
Figure 2. The crystal structures of the Nb-doped Ti-Al binary intermetallic compounds: (a) hP8-Ti3Al-Nb, (b) tP4-TiAl-Nb, (c) tP32-Ti3Al5-Nb, (d) tI24-TiAl2-Nb, (e) tI16-Ti5Al11-Nb, (f) tI24-Ti2Al5-Nb, and (g) tI32-TiAl3-Nb. The numbers of (1)–(5) represent the Nb doping at different crystallographic sites; the brown, orange, and green balls represent Ti, Al, and Nb, respectively.
Materials 17 00358 g002
Figure 3. The 3D projection of the Young’s modulus of (a) Ti-Al and (b) Nb-doped Ti-Al compounds.
Figure 3. The 3D projection of the Young’s modulus of (a) Ti-Al and (b) Nb-doped Ti-Al compounds.
Materials 17 00358 g003
Figure 4. The total density of states (TDOSs), partial density of states (PDOSs) and trend of the pseudo-energy gap width with increasing Al content of the (a) Ti-Al and (b) Nb-doped Ti-Al compounds.
Figure 4. The total density of states (TDOSs), partial density of states (PDOSs) and trend of the pseudo-energy gap width with increasing Al content of the (a) Ti-Al and (b) Nb-doped Ti-Al compounds.
Materials 17 00358 g004
Table 1. The calculated lattice parameters (Å) and formation enthalpy (kJ/mol) of the Ti-Al intermetallic compounds.
Table 1. The calculated lattice parameters (Å) and formation enthalpy (kJ/mol) of the Ti-Al intermetallic compounds.
Pearson Symbol
(Space Group)
Lattice Parameters and Formation EnthalpyMethod and Reference
abcΔHr
Ti3AlhP85.759 4.655−27.086PAW-GGA
(P63/mmc)5.759 4.655−26.827US-PP-GGA [14]
Ti3AlhP8
(P63/mmc)
5.7372 4.6825−27.395US-PP-GGA [13]
5.6496 4.5706−28.70FP-LMTP-LDA [30]
5.6136 4.6649−26.979FLASTO-LDA [31]
5.775 4.655 Experiment [32]
TiAltP4
(P4/mmm)
3.9893 4.074−39.23PAW-GGA
3.994 4.079−38.431US-PP-GGA [14]
3.9814 4.0803−39.712US-PP-GGA [13]
3.9921 4.04−42.00FP-LMTP-LDA [30]
3.9716 4.051−42.00FLASTO-LDA [31]
4.001 4.071 Experiment [33]
−40.1 ± 1Experiment [34]
−36.4 ± 1Experiment [35]
−35.1 ± 0.5Experiment [36]
cP2
(Pm-3m)
3.1865 −26.154PAW-GGA
3.1854 −25.876US-PP-GGA [13]
3.1529 −25.052FLASTO-LDA [31]
Ti3Al5tP32
(P4/mmm)
11.283 4.0305−41.25PAW-GGA
11.286 4.0311−41.640US-PP-GGA [13]
11.293 4.0381 Experiment [37]
TiAl2tI24
(I41/amd)
3.967 24.306−41.73PAW-GGA
3.9658 24.321−42.370US-PP-GGA [13]
3.9628 24.068−42.396FLASTO-LDA [31]
3.9711 24.313 Experiment [38]
oC12
(Cmmm)
12.1493.93054.0067−41.346PAW-GGA
12.1643.9364.011−40.896US-PP-GGA [14]
12.1613.93224.0018−42.013US-PP-GGA [13]
Ti5Al11tI16
(I4/mmm)
3.926 16.517−39.519PAW-GGA
3.9239 16.52−40.18US-PP-GGA [13]
3.923 16.519 PAW-GGA [16]
3.917 16.524 Experiment [39]
3.923 16.535 Experiment [40]
Ti2Al5tP28
(P4/mmm)
3.9132 29.019−39.808PAW-GGA
3.9114 29.023−39.398US-PP-GGA [13]
3.912 29.004 PAW-GGA [16]
3.905 29.196 Experiment [40]
TiAl3tI32
(I4/mmm)
3.8732 33.841−38.846PAW-GGA
3.875 33.84 Experiment [41]
tI8
(I4/mmm)
3.9664 8.4797−38.37PAW-GGA
3.76 8.4976−41.44FP-LMTO-LDA [42]
3.799 8.5174−39.51FLASTO-LDA [31]
3.8400–3.8537 8.5600–8.6140 Experiment [43]
−36.6 ± 1.3Experiment [44]
−39.2 ± 1.8Experiment [31]
cP4
(Pm-3m)
3.9807 −35.616PAW-GGA
3.981 −36.583US-PP-GGA [13]
3.9800–4.0500 −36.907Experiment [43]
−36.614Experiment [31]
Table 2. The calculated lattice parameters (Å) and formation enthalpy (ΔHr, kJ/mol) of the Nb-doped Ti-Al intermetallic compounds, and the percentage change in the structural parameters relative to those of the non-doped ones. x% is the atomic content of Nb in percentage. Boldface denotes the most stable structure in the same Ti-Al component compounds.
Table 2. The calculated lattice parameters (Å) and formation enthalpy (ΔHr, kJ/mol) of the Nb-doped Ti-Al intermetallic compounds, and the percentage change in the structural parameters relative to those of the non-doped ones. x% is the atomic content of Nb in percentage. Boldface denotes the most stable structure in the same Ti-Al component compounds.
x%acc/aVΔaΔcΔ(c/a)ΔVΔHr
hP8-Ti3Al-Nb-16.250%5.7299.3271.628267.107−0.327%0.266%1.189%0.349%−26.304
hP8-Ti3Al-Nb-23.125%11.5314.6580.404533.2710.301%0.150%−0.075%0.171%−26.784
tP4-TiAl-Nb-13.125%7.9808.1571.022519.5090.021%0.116%0.095%0.159%−38.784
tP32-Ti3Al5-Nb-14.545%11.2974.0350.357514.0120.124%0.102%−0.022%0.171%−40.608
tP32-Ti3Al5-Nb-24.545%11.2894.0340.357514.0130.051%0.075%0.024%0.172%−40.512
tI24-TiAl2-Nb-14.167%3.96424.3836.151383.283−0.069%0.318%0.387%0.203%−40.512
tI16-Ti5Al11-Nb-16.250%3.92816.5714.218255.7360.063%0.326%0.262%0.452%−38.4
tI16-Ti5Al11-Nb-26.250%3.92816.5514.213255.3690.052%0.204%0.151%0.307%−37.152
tI16-Ti5Al11-Nb-36.250%3.92016.5824.230254.817−0.151%0.394%0.546%0.090%−38.592
tP28-Ti2Al5-Nb-13.571%3.91229.0967.437445.385−0.018%0.264%0.282%0.228%−38.784
tP28-Ti2Al5-Nb-23.571%3.92028.9827.394445.3290.171%−0.127%−0.297%0.215%−37.92
tP28-Ti2Al5-Nb-33.571%3.91129.1077.442445.276−0.05%0.304%0.355%0.203%−36.96
tP28-Ti2Al5-Nb-43.571%3.91229.0927.436445.270−0.025%0.252%0.277%0.202%−36.96
tP28-Ti2Al5-Nb-53.571%3.91429.0957.434445.6720.016%0.261%0.246%0.292%−36.48
tI32-TiAl3-Nb-13.125%3.87433.8718.772508.2530.013%0.414%0.401%0.112%−38.208
tI32-TiAl3-Nb-23.125%3.87133.8898.754507.911−0.047%0.141%0.188%0.045%−38.976
Table 3. The elastic constants (GPa) of the Ti-Al and Nb-doped Ti-Al intermetallic compounds.
Table 3. The elastic constants (GPa) of the Ti-Al and Nb-doped Ti-Al intermetallic compounds.
Cij (GPa)
C11C12C13C22C23C33C44C55C66
Ti-Al compounds
hP8-Ti3Al193.984.166.5193.966.522363.563.554.9 a
192.278.266.8192.266.8234.261.661.657.0 b [14]
202.667.678.9202.678.9202.961.661.667.5 a [49]
tP4-TiAl17188.785.917185.9165.5114.1114.169.8 a
168.688.380.9168.680.9174.1111.8111.873.7 a [14]
166.49688.1166.488.1179.6119.2119.276.0 a [49]
1738384 168111 75 a [48]
1867274 176101 77 d [50]
tP32-Ti3Al52155071.121571.1180.1104.8104.869.7 a
213.752.772.1 181.8101.4 65.8 c [51]
tI24-TiAl2199.269.558.4199.258.4214.688.588.598.7 a
tI16-Ti5Al11201.668.856.6201.656.6208.988.588.593.9 a
200.671.858.8 208.587.6 92.6 a [16]
tP28-Ti2Al5206.168.154206.154205.584.584.5100.2 a
218.562.948.8 221.1102.3 117.0 a [16]
tI32-TiAl3208.771.347.1208.747.1215.889.389.3116.2 a
Nb-doped Ti-Al compounds
hP8-Ti3Al-Nb-2189.991.368.218566.4226.860.762.653.6
tP4-TiAl-Nb-1171.393.587.4171.387.4167.7114.7114.773.8
tP32-Ti3Al5-Nb-1216.551.673.2218.272.2181.1104.7104.767.5
tI24-TiAl2-Nb-1200.87361201.361.3216.188.588.5100.2
tI16-Ti5Al11-Nb-3208.374.161.3208.361.3212.392.292.299
tP28-Ti2Al5-Nb-1209.274.452.4209.252. 4212.786.286.2104.6
tI32-TiAl3-Nb-22147447.321447.3221.290.690.6117.8
a PAW-GGA-PBE; b USPP-GGA; c PAW-GGA-PW91; d experiment.
Table 4. The bulk modulus B (GPa), shear modulus G (GPa), Young’s modulus E (GPa), Poisson’s ratio ν, Pugh’s ratio K, Vickers hardness Hv, and Debye temperature ΘD (K) of the Ti-Al and Nb-doped Ti-Al intermetallic compounds.
Table 4. The bulk modulus B (GPa), shear modulus G (GPa), Young’s modulus E (GPa), Poisson’s ratio ν, Pugh’s ratio K, Vickers hardness Hv, and Debye temperature ΘD (K) of the Ti-Al and Nb-doped Ti-Al intermetallic compounds.
BGEHvνKΘD
Ti-Al compounds
hP8-Ti3Al116.162.2158.37.70.2731.87496.5
tP4-TiAl114.369.1172.510.20.2401.65554.7
tP32-Ti3Al5110.581.4196.115.20.2041.36620.0
tI24-TiAl2109.582.7198.215.90.1981.32631.7
tI16-Ti5Al11108.582.2196.918.20.1971.32632.8
tP28-Ti2Al5107.782.7197.418.40.1941.30640.2
tI32-TiAl3106.989.7210.319.50.1721.19673.1
Nb-doped Ti-Al compounds
hP8-Ti3Al-Nb-2118.260.4154.87.00.2821.96481.8
tP4-TiAl-Nb116.369.4173.610.00.2511.67545.6
tP32-Ti3Al5-Nb-1112.180.9195.714.70.2091.38601.5
tI24-TiAl2-Nb112.182.6199.015.40.2041.35614.9
tI16-Ti5Al11-Nb-3113.684.6203.315.70.2011.34617.6
tP28-Ti2Al5-Nb-1109.984.7202.116.60.1931.29628
tI32-TiAl3-Nb-2109.590.6213.019.20.1761.21662.1
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

Wang, K.; Xiang, H.; Xu, L.; Feng, A.; Qu, S.; Wang, H.; Chen, D. The Effect of Nb Doping on the Properties of Ti-Al Intermetallic Compounds Using First-Principles Calculations. Materials 2024, 17, 358. https://doi.org/10.3390/ma17020358

AMA Style

Wang K, Xiang H, Xu L, Feng A, Qu S, Wang H, Chen D. The Effect of Nb Doping on the Properties of Ti-Al Intermetallic Compounds Using First-Principles Calculations. Materials. 2024; 17(2):358. https://doi.org/10.3390/ma17020358

Chicago/Turabian Style

Wang, Kun, Hongping Xiang, Lin Xu, Aihan Feng, Shoujiang Qu, Hao Wang, and Daolun Chen. 2024. "The Effect of Nb Doping on the Properties of Ti-Al Intermetallic Compounds Using First-Principles Calculations" Materials 17, no. 2: 358. https://doi.org/10.3390/ma17020358

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