Introduction

Ti2AlNb-based alloy, sometimes referred to as orthorhombic alloy1,2, is a class of highly promising lightweight high-temperature materials. This type of alloy is considered to partially substitute the high-density (ρ = 8~8.5 g/cm3) Ni-based superalloys in the aerospace industry due to its low density, high strength, superior plasticity, high fracture toughness and excellent creep resistance at elevated temperatures3,4,5,6,7,8,9. In such applications, the operating temperatures could go beyond 600–650 °C10,11, leading to severe oxidation of the alloy surface12,13,14. There are three potential approaches to improve high-temperature oxidation resistance: alloying15,16,17, pre-oxidation18, and coating19,20,21,22,23,24,25,26. For example, Wang et al.27 reported that a two-step voltage-controlled microarc oxidation (MAO) method can be used to produce ceramic coatings on a Ti2AlNb-based alloy. However, after a prolonged exposure to air at elevated temperatures, intermetallics exhibit oxygen-induced embrittlement characteristics such as low ductility and brittle fracture16,28,29. Thus, an understanding of high-temperature oxidation mechanisms is essential for improving the oxidation resistance of materials. In our previous study30, a Ti2AlNb-based alloy was observed to exhibit fairly good oxidation resistance below 750 °C. After reaching 800 °C, the oxidation resistance decreased dramatically. Thus, the oxidation behavior and mechanisms are investigated at a higher temperature of 800 °C in this study.

Mass transfer is known to be the essence of oxidation reaction. During the high-temperature oxidation of a Ti2AlNb-based alloy, O and N elements diffuse inward, whereas Al, Ti and Nb elements diffuse outward. Among the many elements that can improve the oxidation resistance, such as Al, Nb, Mo, Si, Zr, etc., Al and Nb are the most important elements31. While Al atoms and O atoms are able to generate a continuous and dense Al2O3 protective layer on the alloy surface and thus improve the oxidation resistance of alloys, this is not the case in Ti2AlNb-based alloys. The Gibbs free energy of Al2O3 and TiO2 is so similar that both oxides are produced almost simultaneously32. The addition of the element Nb can improve the oxidation resistance of the alloy: Nb substitutes for Ti in TiO2 as a cation with a valence of 5, while no Nb is present in Al2O333,34,35,36,37. The doping of Nb in TiO2 grains reduces oxygen vacancy and Ti cations, which impedes the mass transfer in TiO234. Lu et al.35 observed the substitution of Ti by Nb via high-resolution transmission electron microscopy (HRTEM) Z-contrast imaging, as represented by the Nb enrichment in TiO2 grains of the mixture layer. Vojtěch et al.36 studied the role of the addition of 2 at.% Nb to the eutectic TiAl-Ti5Si3 alloy, and reported that Nb markedly influences oxidation kinetics, with a six-fold decrease of oxidation rate.

Some fundamental aspects of oxidation behavior, such as weight gain, scale morphology, and structure, have been investigated12,21,38,39,40,41. Leyens and Gedanitz12 studied the mass gain and oxidation rate of a Ti-22Al-25Nb alloy in air between 650 °C and 800 °C, and reported a fairly good oxidation resistance at 650 °C up to 4000 hours and at 700 °C up to 500 hours, whereas at 800 °C “breakaway” oxidation occurred after about 100 hours. Wang et al.21 observed layers of TiO2 and a small amount of AlNbO4 with needle-like TiO2 crystals present all over the surface. Ralison et al.38,39 reported a multi-layered scale \(({{\rm{TiO}}}_{2}+\frac{1}{2}{{\rm{AlNbO}}}_{4}/({\rm{eventually}}\,{\rm{Ta}},\,{\rm{Mo}})-{\rm{rich}}\,{{\rm{AlNbO}}}_{4})\) along with an oxygen-affected zone in a Ti-27Al-15Nb alloy at 800 °C in air, and Al2O3/(TiO2 + AlNbO4)/Ta-rich Al2O3/oxygen-affected zone in a Ti-27Al-10Nb alloy at the same temperature of 800 °C. Some cracks were present in the multi-layered scale. Zheng et al.40 studied the oxidation behavior of a Ti-22Al-25Nb alloy at 800 °C for 300 hours and observed the formation of a mixed oxide scale on the alloy surface, which was predominantly composed of TiO2, AlNbO4, and Nb2O5.

These studies revealed a complex scale structure containing oxidation products of Al2O3, TiO2, Nb2O5, AlNbO4, etc., with an outer layer consisting mainly of TiO2. When the content of the element Nb is high enough, Nb2O5 or AlNbO4 would be present, however, they are prone to spall-off and are unfavorable to the oxidation resistance. As for the structure of inner oxide layer, Małecka41 observed that it consists of an Al-rich layer and Ti, Al (Nb, Mo, V)-rich zone. Li et al.42 reported that it has such a structure: TiO2-rich layer/AlNbO4-rich layer/TiO2-rich layer/AlNbO4-rich layer/oxygen and a nitride-enriched zone. Leyens14 reported that when the temperature is above 900 °C there exists a nitrogen-enriched layer underneath the oxide scale, i.e., a nitride-containing layer. However, the questions remain as to how the oxide scale containing various oxides is formed; in what form (crystalline or amorphous) the substance/scale interface region would be; and if different types of Al2O3 can be co-existent in the oxidation of Ti2AlNb-based alloys. The objective of the present study is to address these questions via detailed microstructrual examinations using different advanced techniques along with thermodynamic calculations.

Materials and Methods

The selected material is as-cast Ti2AlNb alloy with a nominal composition of Ti-22Al-20Nb-2V-1Mo-0.25Si (in at.%). The alloy ingot was cut into small plates with a size of 8 × 8 × 3 mm by electro-discharge machining. The surface of the samples was ground with sandpaper from grit #400 to #1200, ultrasonically cleaned in acetone for 15 mins. The dimensions were measured using a Vernier caliper and the samples were weighted using an analytical balance with an accuracy of 0.00001 g. During the isothermal oxidation in air at 800 °C, the samples were taken out of the furnace at intervals of 1, 3, 6, 12, 24, 36, 50, 62, 74, 86, and 100 h, cooled to room temperature and weighed. It should be noted that the Ti2AlNb-based alloy was observed to exhibit a fairly good oxidation resistance below 750 °C, while its oxidation resistance decreased considerably above 800 °C. Thus, the oxidation behavior and mechanism are studied in detail at 800 °C in the present investigation. The time of oxidation experiments was selected according to the standard HB5258-2000 of aerospace sector in China, where a duration of 100 hours is suggested to be sufficient. Also, if the oxidation time was too long, the oxide layer would peel off, causing difficulties for the study of the oxide scale.

A stereoscope was used to observe the grain sizes. XRD (Rigaku D/Max-2550) with a Cu Kα radiation (λ = 1.5418 Å) was used to identify the phases in the oxide scale at 50 kV and 200 mA with a diffraction angle (2θ) from 10° to 100° at a step size of 0.02° and 1 s in each step. SEM (Nova Nano SEM 450) was used to observe the surface morphology and cross-sectional structure of the oxide layer. For the observation of the oxide cross-section, it was necessary to mount the sample with resin, use sandpapers from grit #400 to #1200 to grind, and then diamond paste to polish the sample to a smooth mirror surface. Since the oxide is an insulator, it is necessary to perform carbon deposition on the polished surface before SEM observations. Elemental distribution in different regions on the cross-section of the oxide scale was characterized by using an electron probe microanalyzer (EPMA, Shimadzu 1720) with a resolution of 1 μm and secondary-electron image resolution of 6 nm using a beam current of 10 nA. TEM (FEI TECNAI G2 S-TWIN F20) was used to examine the structures of the oxide scale as well as the scale/substrate interface. To locate the scale/substrate interface more precisely TEM samples were prepared via the cutting of a focused ion beam (FIB, FEI, Helios nanolab 600). The dimension of FIB-TEM samples is: 5 μm in length, 4 μm in width, ~35 nm in thickness for the scale, and ~65 nm in thickness for the scale/substrate interface.

Results

Microstructures of as-cast Ti2AlNb alloy

Ti2AlNb-based alloy contains different volumes of the ordered phases β0 (Strukturbericht: B2, space group: \({\rm{Pm}}\mathop{3}\limits^{-}{\rm{m}}\), Person symbol: cP2), α2-Ti3Al (Strukturbericht: DO19, space group: P63/mmc, Pearson symbol: hP8), and the ordered orthorhombic O-Ti2AlNb phase (Strukturbericht: A2BC, space group: CmCm, Person symbol: oC16)43. There exist crystallographic orientations of these phases1,2,3: \({[1\bar{1}1]}_{B2}//{[11\bar{2}0]}_{{\alpha }_{2}}\), \({(011)}_{B2}//{(0001)}_{{\alpha }_{2}}\), \({[0001]}_{{\alpha }_{2}}//{[001]}_{O}\), \({(10\bar{1}0)}_{{\alpha }_{2}}//{(110)}_{O}\), \({[\bar{1}11]}_{B2}//{[1\bar{1}0]}_{O}\), \({(110)}_{B2}//{(001)}_{O}\). X-ray diffraction pattern, stereoscopic image, back-scattered electron (BSE) SEM micrograph, and TEM bright field image along with the relevant selected area electron diffraction (SAED) patterns of as-cast Ti2AlNb alloy are shown in Fig. 1(a–g). XRD results reveal the presence of O-Ti2AlNb phase, B2 phase, and α2 phase in Fig. 1(a). The stereoscopic image indicates coarse grains in the Ti2AlNb-cast alloy (Fig. 1(b)). The SEM image in Fig. 1(c) shows dark α2 phase, gray O-Ti2AlNb phase and B2 phase, which can be better seen in a magnified TEM image in Fig. 1(d). In Fig. 1(c), a small amount of α2 phase is mainly located at grain boundaries in the Ti2AlNb-based alloy. Then there are only B2 phase (matrix) and O-Ti2AlNb phase (lath) presented in Fig. 1(d–g), with crystallographic orientations between them: \({[\bar{1}11]}_{B2}//{[1\bar{1}0]}_{O}\), \({(110)}_{B2}//{(001)}_{O}\).

Figure 1
figure 1

X-ray diffraction pattern (a) and microstructure (bg) of as-cast Ti2AlNb-based alloy. (b) Stereoscopic image showing coarse grains; (c) SEM back-scattered electron micrograph showing dark α2-Ti3Al, gray B + O-Ti2AlNb phases; (d) TEM bright field image along with (eg) the corresponding SAED patterns of points 1 and 2, and crystallographic orientations, where 1 stands for O-Ti2AlNb phase, and 2 denotes B2 phase.

Isothermal oxidation kinetics

Figure 2(a) shows a curve of isothermal oxidation kinetics of Ti2AlNb-cast alloy at 800 °C. The obtained weight gain of this alloy after 100 h at 800 °C was about 9.3 mg/cm2. The relationship between oxidation and mass gain could be obtained by fitting the experimental data using the following equation,

$${\rm{\Delta }}{M}^{n}=k{}_{n}t,$$
(1)

where ΔM represents the weight gain per unit area (mg/cm2), n is an oxidation exponent (n = 1, liner relationship; n = 2, parabolic relationship), kn is a rate constant (mgn/cm2n·h), and t is oxidation time (h). The obtained oxidation exponent was ~0.83, being close to 1, thus suggesting that the oxidation kinetics of Ti2AlNb-cast alloy at 800 °C obeyed basically a linear relationship and the oxide layer is not protective at this temperature.

Figure 2
figure 2

Oxidation of Ti2AlNb-cast alloy at 800 °C for 100 h. (a) Oxidation kinetics; (b) surface morphology of scale; (c) XRD pattern of the oxide scale on the surface and substrate; (d) SEM back-scattered electron micrograph of cross-structure scale; (e) EPMA mapping of a selected area in (d).

Surface morphology and structure of scale

Secondary electron micrograph, XRD pattern, back-scattered electron SEM micrograph, and EPMA mapping are shown in Fig. 2(b–e). Randomly-oriented and fairly-dense laminar-shaped oxides of about 2~3 μm long are observed to cover the alloy surface, as seen in Fig. 2(b).

X-ray diffraction patterns of oxide scale on Ti2AlNb alloy are shown in Fig. 2(c). It is seen that a large amount of AlNbO4 and TiO2 is present in the scale after oxidation at 800 °C. According to our previous studies30, the laminar-shaped oxide is AlNbO4. The cross-section of the overall scale can be seen from an SEM back-scattered electron (BSE) image shown in Fig. 2(d). The scale appears dense with a thickness of ~60 μm, and it consists of three layers based on the EPMA mapping in Fig. 2(e). The structure can be deduced to be (Al,Nb)-rich mixed oxide layer (I)/mixed oxide layer (II)/oxygen-rich layer (III) from the outside to inside. However, the presence of some Al2O3 is also found in layer I through TEM investigations.

EPMA point microanalyses were used to reveal the chemical composition in various locations of the cross-sectional scales in Fig. 2(d). According to XRD results in Fig. 2(c) and the chemical composition of points 1–10 in Table 1, AlNbO4 and TiO2 can be confirmed to be the main oxides in the scale. However, Al and Nb are richer at points 1–2 (in the outermost scale of ~5 μm thick) than at points 3–6, i.e., the (Al,Nb)-rich mixed outer layer (I) and mixed mid-layer (II). This should be a result of the rapid growth of TiO2. The consumption of oxygen leads to a reduction in the oxygen partial pressure, but there is still sufficient oxygen (compared to the interior of the alloy) in the O-rich zone (III), as represented by points 7–10 where the content of oxygen was determined to be ~15 at.% or higher, as shown in Table 1. Also, a low content of nitrogen indicates that there is no nitride present in the oxide scale (points 1–6). N and O penetrated into O-rich zone (III) will cause an environmental brittleness and reduce the mechanical properties of the alloy28.

Table 1 Elemental composition (at.%) determined via EPMA at points 1–10 in Fig. 2(d).

TEM study of oxide layer

TEM samples were taken from the outer scale (right) and at the scale/substrate interface (left) via FIB, as shown in Fig. 3(a). Figure 3(b) shows a TEM bright-field image, with the SAED pattern of point 1 given in Fig. 3(c), indicating the presence of AlNbO4 phase. Figure 3(d) is an HRTEM image of point 2 in Fig. 3(b), where Fourier transformations of 2A, 2B, and 2 C were performed to obtain diffraction patterns shown in Fig. 3(e) to (g), corresponding to α-Al2O3, δ-Al2O3 and α-Al2O3. This suggests the co-existence of different types of Al2O3 oxides. It should be noted that, to the best of the authors’ knowledge, such a co-existent phenomenon of different forms of Al2O3 oxides observed via HRTEM has not been reported in the literature, where α-Al2O3 is a thermodynamically stable form (corundum form) while δ-Al2O3 is one of metastable transition forms/polymorphs of alumina. Figure 3(h) is a SAED pattern of point 3 in Fig. 3(b), which corroborates the presence of TiO2.

Figure 3
figure 3

Analysis of oxide scale after oxidation at 800 °C for 100 h. (a) Position of TEM samples prepared via FIB; (b) TEM bright field image; (c) and (h) the corresponding SAED patterns of point 1 and 3 in (b), and (dg) the corresponding HRTEM and FFT images of point 2 in (b), where 1 represents AlNbO4, 2 A denotes α-Al2O3, 2B stands for δ-Al2O3, 2 C signifies α-Al2O3, and 3 indicates rutile-TiO2.

Figure 4 shows the microstructures of interface between Ti2AlNb substrate and oxide scale, from a TEM sample taken at the interface of oxide layer/substrate (i.e., the left TEM sample shown in Fig. 3(a)). Figure 4(a,b) present a TEM bright-field image and a high angle annular dark-field (HAADF) image, respectively, where the left side represents the substrate and the right side represents the oxide scale. It should be noted that the white spots in the oxide layer in Fig. 4(a) are pores. The rutile-TiO2 is found at point 1 in Fig. 4(c,d), with an amorphous surrounding. Nb content at point 2 is very high, and it has been identified as AlNb2 (Fig. 4(e)), which is also one of the common oxidation products of the Ti2AlNb-based alloy. Point 3 shows a TiO polycrystalline ring (Fig. 4(f)), which is further oxidized to become TiO2. In Fig. 4(g,h) for point 4, Nb2O5 is revealed to be present at the boundary of two phases. This is due to the fact that the diffusion of oxygen at the phase boundary is faster as a result of the presence of phase (or grain) boundary energy, along with the reduced reaction activity of Ti and Al. Point 5 shows a lamellar O-Ti2AlNb phase where brookite TiO2 and O-Ti2AlNb phases are present, which are shown in Fig. 4(i). In Fig. 4(j), TiO2, γ-Al2O3, and Nb2O5 are observed to co-exist in the O-Ti2AlNb laths, with the orientation relationships of γ-Al2O3 and TiO2: \({(211)}_{\gamma -A{l}_{2}{O}_{3}}//{(220)}_{rutile-Ti{O}_{2}}\), \({(003)}_{\gamma -A{l}_{2}{O}_{3}}//{(112)}_{rutile-Ti{O}_{2}}\), \(\,{[1\bar{2}0]}_{\gamma -A{l}_{2}{O}_{3}}//{[1\bar{1}0]}_{rutile-Ti{O}_{2}}\).

Figure 4
figure 4

Microstructures of interface between as-cast Ti2AlNb-based alloy substrate and oxide scale. (a) TEM bright field image of interface, where white regions reflect pores, (b) high angle annular dark field (HAADF) image of (a); (c) HRTEM of point 1, and (d) FFT image of (c); (g) HRTEM of point 4, and (h) FFT image of (g); (e), (f), (i) and (j) the corresponding SAED patterns of points 2, 3, 5, 6, where point 1 represents rutile-TiO2, point 2 represents AlNb2, point 3 represents TiO, point 4 represents Nb2O5, point 5 represents brookite TiO2 and O-Ti2AlNb phase, point 6 represents rutile TiO2, Nb2O5 and γ-Al2O3.

Discussion

Evolution of B2 phase and O-Ti2AlNb phase

As mentioned above, the Ti2AlNb-based alloy consists of 3 phases, i.e., α2-Ti3Al, B2, and O-Ti2AlNb. In the present Ti2AlNb-cast alloy there is fewer α2-Ti3Al phase existing in the vicinity of grain boundaries and the O-Ti2AlNb laths are present in the original coarsened B2 grains. The diffusion rate of oxygen atoms in the B2 phase is larger than that in the other two phases44. It follows that no B2 phase is present at the interface. This means that oxidation reaction occurs first in the B2 phase, producing metastable TiO, which is further oxidized to TiO2. The brightest regions on the HAADF image in Fig. 4(b) are Nb-rich areas and have been identified to be AlNb2. It is sandwiched between the complete and residual O-Ti2AlNb laths, which corresponds to the former position of the B2 phase. AlNb2 is considered as a by-product during the oxidation of the Nb-containing TiAl-based alloy, which is due to the reaction of enriched Nb and Al elements45. There are a large number of Nb element in the B2 phase43. During oxidation, the Ti element in the B2 phase reacts with the O element to form TiO/TiO2. The consumption of Ti causes the reaction of Al and Nb. As a result, AlNb2 is surrounded by gray TiO/TiO2 in Fig. 4(b). AlNb2 is stably present at the interface, indicating that its oxidation resistance is relatively strong45, however, the discontinuous nature of AlNb2 at the interface makes the effect limited.

After this the O-Ti2AlNb phase is oxidized. Combined with Fig. 4(f–j), there are brookite-TiO2, rutile-TiO2, γ-Al2O3, and Nb2O5 in the O-Ti2AlNb phase. It can thus be considered that the O-Ti2AlNb phase has been oxidized and decomposed into oxides of TiO2, Al2O3, and Nb2O5. Al2O3 blocks the transport channel, hinders further growth of TiO2, and improves oxidation resistance. As one of the oxide products, Al2O3 has been reported extensively in the literature. During the early stage of oxidation, the oxides at the interface exist in an amorphous form. Lu et al.29 reported that polycrystalline TiO2 and amorphous Al2O3 coexist in the scale. This is similar to the structure in Fig. 4(c). It should be noted that TiO2 and Al2O3 occur almost simultaneously because of their similar Gibbs free energy43, however, the energy is not high enough at an oxidation temperature of 800 °C, resulting in the presence of metastable alumina. To the best of the authors’ knowledge, there is no report about the structure of Al2O3 in the oxidation of a Ti2AlNb-based alloy. In the present work, the unique formation of alumina is observed via HRTEM in the alloy after oxidation at 800 °C for 100 h. That is, the presence of three adjoining oxide grains and oxides in O-Ti2AlNb laths, as shown in Figs 3(d) and 4(j), suggests the occurrence of phase change among three types of Al2O3, i.e., α-Al2O3, δ-Al2O3, and γ-Al2O3. The γ-Al2O3, θ-Al2O3, and κ-Al2O3 are common metastable phases in TiAl alloys at 900 °C29,46,47. According to Yang et al.48, γ-Al2O3 formed with twins in the oxidation of NiAl, and γ-Al2O3 twins were observed to play an important role in the scale growth. Cowley et al.49 reported that {111} γ-Al2O3 twin boundaries provide a fast diffusion path for Al cations. This would improve the oxidation resistance of an alloy. However, in the present study of the Ti2AlNb-based alloy at 800 °C, there exists δ-Al2O3 (monoclinic, a = 11.74 Å, b = 5.72 Å, c = 11.24 Å, β = 103.34°50), while no γ-Al2O3 twins are observed. As reported by Levin and Brandon51, there is a route for the formation of Al2O3: Amorphous (anodic film) → γ → δ → θ → α-Al2O3. This would be the phase change route of Al2O3 in the present Ti2AlNb-based alloy. Further studies in this aspect are needed at different oxidation temperatures.

Formation of AlNbO4

As described above, there are TiO2, Al2O3, and AlNbO4 in the outer oxide layer, as presented in Fig. 3. Zheng et al.40 assumed that the fast-growing Nb2O5 could react with Al2O3 that developed at the early stage of oxidation to form AlNbO4. In the present work, Dmol3 module in Materials Studio 6.0 is used to calculate the Gibbs free energy change of reaction: Nb2O5 + Al2O3 → 2AlNbO4, however, the temperature in the software is up to 1000 K only, which is lower than the temperature of the present oxidation experiment, 1073.15 K. Thus, the following equation is used to fit the relationship between the Gibbs free energy and temperature,

$$G={H}_{0}+aT\,\mathrm{ln}\,T+b{T}^{2}+\frac{c}{T}+IT,$$
(2)

where H0, a, b, c, and I are fitting parameters. The results show an excellent matching close to 100%, as shown in Fig. 5 and Table 2. The relationships between the Gibbs free energy and temperature for Al2O3, Nb2O5, and AlNbO4 can thus be expressed as follows,

$${{\rm{G}}}_{{{\rm{Al}}}_{2}{{\rm{O}}}_{3}}=1.18577-0.00368T\,\mathrm{ln}\,T-1.4271\times {10}^{-6}{T}^{2}+\frac{3.88868}{T}+0.01932T,$$
(3)
$${{\rm{G}}}_{{{\rm{Nb}}}_{2}{{\rm{O}}}_{5}}=2.07990-0.01388T\,\mathrm{ln}\,T-2.54067\times {10}^{-6}{T}^{2}+\frac{10.8132}{T}+0.05475T,$$
(4)
$${{\rm{G}}}_{{{\rm{AlNbO}}}_{4}}=6.74954-0.01287T\,\mathrm{ln}\,T-8.12388\times {10}^{-6}{T}^{2}+\frac{16.08122}{T}+0.06692T.$$
(5)
Figure 5
figure 5

Gibbs free energy curves of Al2O3, Nb2O5 and AlNbO4 before and after fitting.

Table 2 Fitting results of free energy curve of Al2O3, Nb2O5 and AlNbO4 in Fig. 5.

Substituting T = 1073.15 K into the above equations yields the Gibbs free energy values for the formation of Al2O3, Nb2O5, and AlNbO4, respectively, in the present oxidation condition of a higher temperature. As a result, the Gibbs free energy change of the reaction Nb2O5 + Al2O3 → 2AlNbO4 becomes −1.00873 kJ/mol. This means that the reaction is thermodynamically possible. Also, Ai et al.52 reported that Nb2O5 reacted completely with Al2O3 to form AlNbO4 in Nb2O5-Al2O3 ceramics. While Nb2O5 is consumed as a reactant, the formed AlNbO4 makes the scale unprotective as well14. Furthermore, Nb2O5 is not observed in the XRD results (Fig. 2(c)), which can be mainly attributed to its reaction with Al2O3 as discussed above, in conjunction with the lower diffusion coefficient of Nb, the rapid growth of TiO2, and the hindering effect of alumina.

Oxidation process at the interface

Based on the above observations and analyses, the high-temperature oxidation process of the Ti2AlNb-based alloy can be summarized below and schematically shown in Fig. 6. Stage 1: Oxygen absorbs on the surface of the Ti2AlNb-based alloy, which later penetrates into it. The B2 phase is oxidized to produce TiO, and oxygen dissolves in O-Ti2AlNb phase. Stage 2: TiO is transformed into rutile-TiO2 and AlNb2 is formed in the areas of the B2 phase. Oxidation occurs, i.e., brookite-TiO2 is generated inside the O-Ti2AlNb phase and Nb2O5 outside the O phase. Stage 3: O-Ti2AlNb phase breaks down to TiO2, Nb2O5, and Al2O3. In the scale there is a reaction: Al2O3 + Nb2O5 → 2AlNbO4, and N atoms are dissolved in the alloy because of the consumption of O atoms. The corresponding crystallographic structures of the related oxides of the Ti2AlNb-based alloy are summarized in Table 3.

Figure 6
figure 6

Schematic diagram showing a summary of high-temperature oxidation process of Ti2AlNb-based alloy. Stage 1, by inward diffusion of oxygen B2 phase is transformed into TiO, and Ο-Ti2AlNb is abound with oxygen. Stage 2, TiO is transformed into rutile-TiO2 and AlNb2 is formed in the areas of B2 phase. Oxidation occurs in O-Ti2AlNb phase. Stage 3, O-Ti2AlNb phase breaks down into Al2O3, Nb2O5 and TiO2. Al2O3 reacts with Nb2O5 to form (AlNbO4 + TiO2) mixed oxide layer. At last, nitrogen dissolves in the alloy.

Table 3 Crystallographic structure of the related oxides.

Conclusions

  1. 1.

    After being exposed at 800 °C in static air for 100 h, the Ti2AlNb-based alloy followed an almost liner kinetic law of oxidation and exhibited a multi-layered structure consisting of an (Al,Nb)-rich mixed oxide layer (I), mixed oxide layer (II), and oxygen-rich layer (III) from the outside to inside.

  2. 2.

    In the mixed outer scale, there existed α-Al2O3 and δ-Al2O3. Al2O3 reacted with Nb2O5 to form AlNbO4, however, Nb2O5 and AlNbO4 were not able to hinder the diffusion of oxygen.

  3. 3.

    The B2 phase was oxidized to form TiO2, where Nb and Al were transformed into AlNb2 at the interface during oxidation. AlNb2 could hinder the diffusion of oxygen and improve the oxidation resistance of the Ti2AlNb-based alloy, but its discontinuous nature allowed only a limited effect.

  4. 4.

    After long-term oxidation at 800 °C, O-Ti2AlNb was oxidized to form TiO2, Al2O3 and Nb2O5. Al2O3 could hinder the growth of TiO2 in the O-Ti2AlNb laths and form a compact scale. Increasing the amount of the O-Ti2AlNb phase in the alloy contributes to the improvement in its high-temperature oxidation resistance.