You are on page 1of 10

Materials Chemistry and Physics 257 (2021) 123703

Contents lists available at ScienceDirect

Materials Chemistry and Physics


journal homepage: www.elsevier.com/locate/matchemphys

Characterization of the nitrided γ-Ti-46Al–2Nb and


γ-Ti-46Al–2Nb-0.7Cr-0.3Si intermetallic alloys
M.N. Mathabathe a, b, **, A.S. Bolokang a, c, *, G. Govender a, C.W. Siyasiya b, R.J. Mostert b
a
Council for Scientific Industrial Research, Materials Science and Manufacturing, Light Metals, Meiring Naude Road, P O Box 395, Pretoria, South Africa
b
Department of Material Science and Metallurgical Engineering, Faculty of Engineering, Built Environment and Information Technology, University of Pretoria, South
Africa
c
Department of Physics, University of the Western Cape, Private Bag X 17, Bellville 7535, South Africa

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• Annealed Ti–46Al–2Nb and


Ti–46Al–2Nb-0.7Cr-0.3Si induced γ-TiAl
and α2-Ti3Al.
• The β and Ti5Si3 phases were formed in
Ti–46Al–2Nb alloy and Ti–46Al–2Nb-
0.7Cr-0.3Si alloys.
• Cyclic oxidation of nitrided
Ti–46Al–2Nb and Ti–46Al–2Nb-0.7Cr-
0.3Si was improved.
• Thermal properties of Ti–46Al–2Nb-
0.7Cr-0.3Si were better than
Ti–46Al–2Nb alloy.
• TEM analysis confirmed partial
amorphization in Ti–46Al–2Nb.

A R T I C L E I N F O A B S T R A C T

Keywords: The thermal, surface properties and microstructures of the nitrided γ-Ti-46Al–2Nb and Ti–46Al–2Nb-0.7Cr-0.3Si
Nitridation intermetallic alloys were investigated. These vacuum-melted alloys were annealed at 1100 ◦ C to stabilize the
Cyclic oxidation (α2+γ) lamellae structure. Moreover, the annealed alloys were further annealed at 900 ◦ C under nitrogen gas
Intermetallics
(nitridation). The alloys microstructures and nitride surface coatings were characterized. The Ti–46Al–2Nb alloy
XPS
was comprised of γ-TiAl, α2-Ti3Al and β-phases after the annealing, while the Ti–46Al–2Nb-0.7Cr-0.7Si alloy
revealed the Ti5Si3, γ-TiAl and α2-Ti3Al phases. Thin nitride layers formed on both alloys. After cyclic oxidation,
the nitrided alloys exhibited surface spallation at different rates. The X-ray photoelectron spectroscopy (XPS)
revealed the presence of mixed oxide layer of Al2O3, Cr2O3 and SiO2 near the Ti–46Al–2Nb-0.7Cr-0.7Si surface.

* Corresponding author. Council for Scientific Industrial Research, Materials Science and Manufacturing, Light Metals, Meiring Naude Road, P O Box 395, Pretoria,
South Africa.
** Corresponding author. Department of Material Science and Metallurgical Engineering, Faculty of Engineering, Built Environment and Information Technology,
University of Pretoria, South Africa.
E-mail addresses: nmathabathe@csir.co.za (M.N. Mathabathe), sbolokang@csir.co.za (A.S. Bolokang).

https://doi.org/10.1016/j.matchemphys.2020.123703
Received 23 December 2019; Received in revised form 28 April 2020; Accepted 15 August 2020
Available online 31 August 2020
0254-0584/© 2020 Elsevier B.V. All rights reserved.
M.N. Mathabathe et al. Materials Chemistry and Physics 257 (2021) 123703

Fig. 1. Schematic illustration of the thermal treatment process of the vacuum melted Ti–46Al–2Nb and Ti–46Al–2Nb-0.7Cr-0.3Si alloys.

1. Introduction the applied stress induces large residual stress in the nitride scale owing
to the mismatch of creep rates between the metal and the nitride scale.
The titanium-aluminium (Ti–Al) binary alloys are composed of The large stress influences the diffusional flux of nitrogen vacancies,
intermetallic phases such as α-(Ti3Al), (TiAl), TiAl2 and TiAl3 depend­ leading to deviation from the parabolic growth law [9]. The aim of the
ing on the alloying and heat treatment employed. In particular, the current work is to investigate the cyclic oxidation properties of the
γ-TiAl alloys have attracted research interest in the structural engi­ nitrided γ-Ti-46Al–2Nb and Ti–46Al–2Nb-0.7Cr-0.3Si intermetallic al­
neering applications. Production of these γ-TiAl based intermetallic al­ loys developed by vacuum melting. Structural analysis and surface layer
loys are associated with machining difficulties. These problems led to characterization are discussed in detail.
extensive research aiming to modify the γ-TiAl intermetallic alloys by
alloying and developments of new ternary, quaternary or quinary 2. Materials and experimental procedure
compounds. The alloying elements improves the high-temperature
applicability of Ti–Al alloys especially those developed for aerospace The nominal composition of the γ-based alloys employed in the study
and automotive industries [1]. The ease to deform and machine the was Ti–48Al–2Nb, and Ti–48Al–2Nb-0.7Cr-0.7Si (at. %). Vacuum arc
γ-TiAl intermetallic alloys is enhanced by producing a class of beta melted buttons of 50 g each were initially produced from a proportional
(β)/gamma (γ) alloys with superior properties [2]. Since these alloys mixture of high purity Ti, Al, Nb, Cr and Si, blended and cold-pressed,
operates at high temperatures, their surface thermal oxidation proper­ respectively. The melting and casting of the alloys were done accord­
ties are crucial. Alternatively, the surface properties of the γ-TiAl ing to the previous study [10]. Schematic illustration of the experi­
intermetallic alloys are improved by coating the alloy with a mental procedure is displayed in Fig. 1. The vacuum melted alloys were
high-temperature resistant layer achievable by nitrogen surface diffu­ homogenized at 1100 ◦ C under Ar atmosphere for 3.6 ks at Ar flow rate
sion (nitridation) [3,4]. These surface modifications are commercial for of 1 L per minute (1LPM) before nitridation and cyclic oxidation ex­
Ti alloys such as Ti–6Al–4V [5] and γ-TiAl alloys [6–8]. Nonetheless, periments. These Ti–48Al–2Nb and Ti–48Al–2Nb-0.7Cr-0.7Si alloys
Zhao et al. [7], revealed the benefits and limitations of the nitrided were annealed at 900 ◦ C in a flowing nitrogen gas environment (Nitri­
quinary Ti–47Al–2Nb–2Cr–0.2Si intermetallic alloy. It has emerged that dation). Other samples were characterized while others were annealed
the non-nitrided Ti–47Al–2Nb–2Cr–0.2Si alloys displayed better in air at 900 ◦ C for 900 h for cyclic oxidation test. Nine cycles were
oxidation resistance than the nitrided alloys in the range of 800–1000 ◦ C conducted (each cycle represents 100 h) at 900 ◦ C with subsequent
due to the dominant development of Al2O3 on the outer layer of the cooling at ambient temperature and returned to the furnace after
scales [7]. It was noticed that the nitrided alloy exhibited increased recording their mass. For both Ti–48Al–2Nb and
oxidising rate with prolongation of nitridation time at 800 ◦ C but Ti–48Al–2Nb-0.7Cr-0.7Si, one sample for each alloy was used until the 9
showed signs of better oxidation resistance at more severe oxidising cycles were completed. The elaborate prefabrication methodology is
conditions when nitrided at 940 ◦ C for 50 h [7]. This alloy disobeyed the outlined in Refs. [10,11]. Microstructural examinations were adminis­
parabolic law at 900 and 1000 ◦ C, which is an indication of the variation tered using conventional methods for γ-TiAl based intermetallic alloys,
in the uniformity of scale [7]. Furthermore, the Al-rich composition was on as-stabilized, pre-nitridated and cyclically oxidized specimens. The
detected below the nitride/sub-layer interface whereby TiN and Ti2AlN, microstructure and the morphology of the phase structures were
ƞ-Ti3N2-x and Ti2N and Al5Ti2 intermetallic phases were detected [8]. determined by the scanning electron microscope (SEM) which is fur­
Interventions were made to develop a model reference to the develop­ nished with energy dispersion spectroscopy (EDS) for microanalysis.
ment of a TiN scale on a γ-TiAl substrate [9]. This model revealed that The X-ray photoelectron spectroscopy (XPS) was employed to determine

2
M.N. Mathabathe et al. Materials Chemistry and Physics 257 (2021) 123703

Table 1 3.1. Annealing of γ-TiAl based alloys


Final chemical composition of the vacuum melted Ti–48Al–2Nb, and
Ti–48Al–2Nb-0.7Cr-0.7Si intermetallic alloys in (at. %). The SEM image of a stabilized ternary γ-Ti-46Al–2Nb alloy is shown
Alloys Final melted alloy composition (at. %) References in Fig. 2a. Dominating phases on the microstructure are γ-(TiAl) and
Elements Ti Al Nb Cr Si
α2-(Ti3Al). The EDS analysis reveal similarities in Table 2 (1100 ◦ C-
Ternary 52.38 45.82 1.8 [11] annealed Ti–46Al–2Nb and Ti–46Al–2Nb-0.7Cr-0.3Si alloys) and the as-
Quinary 51.68 45.54 1.92 0.73 0.27 [13] cast composition in Table 1, which signify alloy homogeneity. Addi­
tionally, residual β-phase was detected as indicated by the double red
arrow. These phase is enriched with beta-stabilizer Nb in γ-Ti-46Al–2Nb
the varying composition of oxide scales. Atomic force microscope (AFM)
alloy [14]. A microstructure of the Ti–46Al–2Nb-0.7Cr-0.7Si alloy is
was utilized to evaluate the effect of surface roughness and wear on the
shown in Fig. 2b comprising of γ-(TiAl), α2-(Ti3Al) phases and Ti5Si3
nitride/oxidized alloys. For transmission electron microscopy (TEM)
precipitates. Similar to the γ-Ti-46Al–2Nb, the Ti–46Al–2Nb-0.7Cr-0.7Si
analysis, a small spatula tip of the oxidized sample was dispersed in
alloy has lamellar structure with different grain orientations attributed
ethanol to yield a lightly dispersed solution and ultrasonicated for ± 30
to alloying elements and heat treatment. Fig. 2c and d shows the
min at ambient temperature. The copper grid coated with carbon was
as-nitrided Ti–46Al–2Nb and Ti–46Al–2Nb-0.7Cr-0.7Si intermetallic
dipped into the dispersed solution and left to dry overnight. Observa­
alloys which were processed at 900 ◦ C before cyclic oxidation tests. The
tions were conducted in a JEOL JEM 2100 high-resolution electron
surface nitride layers of both alloys were very thin. Upon nitridation,
transmission microscopy (HRTEM) operating at 200 kV. The macro and
precipitation of phases was observed. EDS chemical composition of
micro-hardness profile values were obtained using a Vickers hardness
tester according to ASTM standard E384-11. Applied loads of 20 Kgf and
500 g were conducted, as reported by Ref. [11,12]. Density measure­ Table 2
ments of the as-stabilized annealed, nitrided and cyclically oxidized EDS spot analysis of the 1100 ◦ C-annealed Ti–46Al–2Nb, and Ti–46Al–2Nb-
specimens were obtained using a pycnometer according to ASTM D854. 0.7Cr-0.3Si alloys (at. %).
Alloy Ti Al Nb Cr Si
3. Results and discussion
TiAlNb 1. 53.55 ± 44.34 ± 2.11 ± – –
0.18 0.08 0.03
The final chemical composition of the as-cast vacuum melted 2. 50.04 ± 48.27 ± 1.69 ±
Ti–48Al–2Nb, and Ti–48Al–2Nb-0.7Cr-0.7Si intermetallic alloys are 0.06 0.10 0.14
shown in Table 1. These results are in agreement with the previous 3. 50.06 ± 48.28 ± 1.66 ±
0.21 0.13 0.05
reporting in Refs. [11,13]. The difference between nominal and final
TiAlNbCrSi 1. 50.92 ± 43.99 ± 1.92 ± 0.57 ± 2.60 ±
composition is due to varying vapor pressures of different components in 0.20 0.17 0.14 0.06 0.21
the alloys, and volatile Al that suffers a burning loss [13]. 2. 50.50 ± 45.79 ± 2.15 ± 0.60 ± 0.96 ±
0.18 0.15 0.09 0.04 0.13

Fig. 2. SEM-SEI micrographs of the as-stabilized γ-TiAl (a) Ternary Ti–46Al–2Nb (at. %), (b) Quinary Ti–46Al–2Nb-0.7Cr-0.7Si (at. %) alloys with their corre­
sponding nitrided images are shown in Fig. 2c and d, respectively.

3
M.N. Mathabathe et al. Materials Chemistry and Physics 257 (2021) 123703

Table 3
The precipitated particles beneath the nitride layer (a) β-phase and (b) Ti5Si3 of the Ti–46Al–2Nb, and
Ti–46Al–2Nb-0.7Cr-0.3Si alloys, respectively.

3.2. Cyclic oxidation behavior of the γ-TiAl alloys

Fig. 3 is the cyclic oxidation curves demonstrating the Ti–46Al–2Nb


and Ti–46Al–2Nb-0.7Cr-0.3Si alloy behavior at high temperature. The
cyclic oxidation trends of the nitrided alloys compared with those
unnitrided conducted revealed different thermal properties. The unni­
trided Ti–48Al–2Nb alloy initially show a sharp negative mass loss
during the first 100 h cycle. It steadily continues to show a decrease-
increase in mass until the 900 h cycle. This mass loss is indicative of
rapid oxide growth shown by initial positive weight gain and spallation
on the sample surface. Upon extensive exposure times, erratic specimen
weight gain and loss was observed throughout oxidation. It became
stable between the 900–1000 h. In contrast, the nitrided curve of the
Ti–46Al–2Nb alloy shows improved oxidation resistance due to inhibi­
tion of the fast oxide growth during the initial stages of cyclic oxidation.
It implies that the nitride coating played a significant role in prohibiting
the rapid development of the oxide layer despite both samples revealing
Fig. 3. Cyclic oxidation conducted in air at 900 ◦ C for 900 h on the pre-nitrided a mass loss. The unnitrided Ti–46Al–2Nb-0.7Cr-0.3Si alloy illustrates a
Ti–46Al–2Nb and the Ti–46Al–2Nb-0.7Cr-0.3Si alloys. Each cycle represents
narrow initial negative mass loss compared to the Ti–46Al–2Nb alloy in
100 h.
the initial 100 h cycle. On the other hand, the nitrided Ti–46Al–2Nb-
0.7Cr-0.3Si alloy shows a steady increase in mass. The show better
phases (Fig. 2c and d) circled point 1 is given in Table 3 for both the performance than the as-cast alloy counterpart. Despite the negative
ternary and quinary alloy. The composition differ for γ-Ti-46Al–2Nb and mass, Ti–46Al–2Nb-0.7Cr-0.3Si alloy performed better than all alloys
Ti–46Al–2Nb-0.7Cr-0.7Si alloys. until 900 ◦ C. The Ti–46Al–2Nb-0.7Cr-0.3Si alloy is composed of Ti5Si3
phase attributable to better thermal properties. Accordingly, at tem­
peratures slightly below 900 ◦ C, a mixed crystalline TiO2 (rutile) and

Fig. 4. SEM-SEI images and EDS analysis of the surface scales of (a, c) Ti–46Al–2Nb and (b, d) Ti–46Al–2Nb-0.7Cr-0.3Si intermetallic alloys after cyclic oxidation.

4
M.N. Mathabathe et al. Materials Chemistry and Physics 257 (2021) 123703

Fig. 5. AFM surface topography images of the (a–c) Ti–46Al–2Nb and (d–f) Ti–46Al–2Nb-0.7Cr-0.3Si intermetallic alloys, respectively, after cyclic oxidation.

SiO2 scale formed. This scale grows inwardly by diffusion of oxygen presence Ti5Si3 precipitates in Ti–46Al–2Nb-0.7Cr-0.3Si alloy was
along grain boundaries or through oxygen vacancies in rutile structure enhanced by (N) surface diffusion. Furthermore, the
[15]. Moreover, a linear mass gain rate of 0.02 mg/cm2/hr controlled Ti–46Al–2Nb-0.7Cr-0.3Si alloy showed a marginal positive mass gain
growth rate and inner mixed scale of TiOx or TiN and Si might have throughout the 600–900 h range cycle, compared to the Ti–46Al–2Nb
formed [16]. It implies that the oxidation resistance contributed by the alloy at these intervals. This behavior is attributed to Cr contributing to

Fig. 6. XPS survey scan of the (a) Ti46Al–2Nb intermetallic alloy after cyclic oxidation at 900 ◦ C for 900 h, and the corresponding experimental and fitted curves for
normalized (b) Al 2p (c) N 1s, (d) Ti 2p, and (e) O 1s XPS spectra.

5
M.N. Mathabathe et al. Materials Chemistry and Physics 257 (2021) 123703

Fig. 7. XPS survey scan of the (a) nitrided Ti46Al–2Nb-0.7Cr-0.3Si intermetallic alloy after cyclic oxidation at 900 ◦ C for 900 h, and the corresponding experimental
and fitted curves for normalized (b) C 1s, (c) Al 2p, (d) Ti 2p, (e) O 1s and.(f) Si 2p.

the oxidation properties [17]. multiple slip-stepped surface structure comprising different orienta­
Although nitridation was induced in Ti–46Al–2Nb and Ti–46Al–2Nb- tions. The deformed surface layers of the γ-TiAl based alloys are due to
0.7Cr-0.3Si alloys, it should be noted that static laboratory air is cyclic oxidation. Fig. 5a and b illustrates a slipping pattern with steps
comprised of a large quantity of nitrogen. Both Ti and Al have higher along a hill-valley groove in the Ti–46Al–2Nb surface. These are the
affinity for O than N. Therefore, N occupy O vacancies due to structural alternating layers of the oxide/oxynitride composition. Fig. 5c clearly
instability created by vibrations caused by high temperature on oxides shows slip steps of thin layers. The slip steps are due to deformation
and forms the titanium oxynitride (TiO2-xNx). The oxynitride layer for­ instead of crystallographic planes observed after spalling of the surface
mation reduces further oxygen diffusion into the surface [11]. The linear scale. The spectral RMS amplitude of the Ti–46Al–2Nb oxide layer is
oxidation at high temperature is time-dependent and yields the change 106 nm (inset of Fig. 5c) and a maximum fringe height of about 240 nm.
in the weight as described by Eq. (1): AFM surface topography of the Ti–46Al–2Nb-0.7Cr-0.3Si oxide films is
shown in Fig. 5d–f. On the contrary, the Ti–46Al–2Nb-0.7Cr-0.3Si oxide
Δm = KL t (1)
film is comprised of a slip-step pattern of intermixed scale layers of
Δm is the change in weight monitored with time (t) with a positive kL different morphology (Fig. 5d and e). The 3D topography in Fig. 5e has
when weight is gained [18]. peripheral slip fringe with a maximum fringe height of 198 nm. The
spectral RMS amplitude of the Ti–46Al–2Nb-0.7Cr-0.3Si alloy is 70.2 nm
(inset of Fig. 5f) smoother than the Ti–46Al–2Nb surface film. This
3.3. Characterization of oxide layers
behavior is attributed to intermixed layers of oxides. This further sug­
gests that there may be more passive layer formation, which may enable
The SEM images of the Ti–46Al–2Nb and Ti–46Al–2Nb-0.7Cr-0.3Si
ease in the restriction of oxygen diffusion in the Ti–46Al–2Nb-0.7Cr-
intermetallic alloys are shown in Fig. 4a and b. The black, red, yellow,
0.3Si alloy in comparison to the Ti–46Al–2Nb alloy.
purple and blue arrows illustrate the EDS composition of the passive
scales. The Ti–46Al–2Nb scale in Fig. 4a confirms the formation of
multiple layers as well as spallation. Dark outer surface scale (black 3.4. XPS spectra analysis
arrow) peels off, leaving the underneath white scale exposed (red
arrow). The formation of Al2O3 is thermodynamically more feasible than 3.4.1. Ternary Ti–46Al–2Nb alloy
the formation of TiO2 at 900 ◦ C [11]. The same principle applies to the Fig. 6a shows a fully scanned spectrum of the x-ray photoelectron
Ti–46Al–2Nb-0.7Cr-0.3Si alloy in Fig. 4b. However, due to prolonged spectroscopy (XPS) chemical composition of the Ti46Al–2Nb alloy. It
exposure in the air at elevated temperatures, γ-alumina phase absorbs confirms the presence of Ti, Nb, Al, O, N, and C. The detected carbon is
the nitrogen ions to form a non-stoichiometric AlON compound rich in associated with the carbon tape used during the sample preparation
defect structures (purple arrow). SEM-SEI images of the Ti–46Al–2Nb analysis. In addition, the Nb3d and 4p XPS spectrum was observed at
and Ti–46Al–2Nb-0.7Cr-0.3Si surface oxides are displayed in Fig. 4c and 209.8 eV and 50 eV, corresponding to the Nb5+ 3d3/2 and Nb5+ 3d5/2
d. The oxide layer of Ti–46Al–2Nb alloy in Fig. 4c shows a spin-orbit peaks of the Nb2O5 phase, respectively [19,20]. Fig. 6b–e
fibrous-morphology with interconnected particles. It is comprised of displays the XPS spectra of the Ti2p core-level of the nitrided
valleys and multiple layers formed during oxidation. The oxide film of Ti46Al–2Nb after cyclic oxidation with Fig. 6b showing the peak of
the cyclically oxidized Ti–46Al–2Nb-0.7Cr-0.3Si in Fig. 4d revealed a 75.38 eV which is associated with Al–O (α-Al2O3) bonds [21]. Another
mixture of a rods and fibrous type structures. peak of 397.3 eV shown in Fig. 6c is related to N–Ti–O (TiO2-xNx) bonds
AFM surface topography of the oxide films shown in Fig. 5a–f reveals [20–23]. This suggests that more oxygen vacancies are produced during

6
M.N. Mathabathe et al. Materials Chemistry and Physics 257 (2021) 123703

Fig. 8. TEM images of the γ-Ti-46Al–2Nb alloy shown in (a) with a corresponding insert indicated by a zoomed-out area marked with a red circle, (b) elemental
mapping viz. (N, O, Ti) of the coating, (c) the high-resolution faceted morphology indicated by the red circle in Fig. 8a (d) with the corresponding selected area
electron diffraction (SAED). (For interpretation of the references to colour in this figure legend, the reader is referred to the Web version of this article.)

annealing for N atoms to occupy the interstitial position in TiO2. Partial of Ti, which yielded a pair of spin-orbit viz. (Ti 2p3/2 and Ti 2p1/2) in
Ti–O bonds on the TiO2 surface break due to annealing and oxygen is the Ti 2p spectrum [21]. The spectrum consisted of broad peaks
released from the TiO2 lattice creating vacancies for N to occupy. Fig. 6d considering that the Ti-bonds occurred at the same time. According to
shows the peaks centered at around 459.8 eV and 465.5 eV assigned to Badini et al. [21], the Ti–O bonds in TiOx, Ti2O3 and TiO2 show bind
Ti4+(2p3/2), and Ti4+(2p1/2) for TiO2 layer [24], while Fig. 6e presents energy (BE) that increases with titanium oxide state viz. Ti 2+, Ti 3+ and
the O1s core level spectra of the TiO2. Each O1s signal of the TiO2 scale Ti 4+. The TiOx is a non-stoichiometric compound with composition
can be fitted into two symmetric peaks. The peaks located at 531.59 and ranging from TiO0.7 to TiO1.3. The TiO2 spin-orbit doublet showed
533.31 eV corresponds to the oxygen vacancy-Ti4+ and oxygen binding energies of 458.8 eV (Ti 2p3/2) and 464.6 eV (Ti 2p1/2). On the
vacancy-Ti3+. other hand, the XPS peaks at 457.0 eV and 463.1 eV are attributed to
Ti2O3 and TiOx and previously found at 455.3 eV and 461.5 eV [21]. The
3.4.2. Quinary Ti–46Al–2Nb-0.7Cr-0.3Si alloy O 1s XPS depth profile in Fig. 7e shows that the oxygen is present in
Fig. 7 illustrates the XPS high-resolution surface spectra of the cyclic Al2O3 at the binding energy of 532.0 eV near the surface of the oxide
oxidized Ti46Al–2Nb-0.7Cr-0.3S intermetallic alloy at 900 ◦ C for 900 h. layer. Moreover, minor SiO2 at 532.9 eV, and Cr2O3 at the binding en­
The Cr 2p spectrum in Fig. 7a is characterized by the Cr 2p 3/2 binding ergy of 530.9 eV were located near the surface of the oxide [25]. Fig. 7f
energy at 576.8 eV. According to Chang et al., [25] this peak is associ­ shows Si peaks appearing in the spectrum and a small fraction of SiO2
ated with the presence of Cr2O3. Simialr to Ti–46Al–2Nb alloy, the near the surface from the Si 2p peak at the binding energy of 103.0 eV
detected carbon in Fig. 7b is associated with the carbon tape used during [25]. Therefore, a mixed oxide layer including Al2O3, Cr2O3 and SiO2
sample preparation. The peak at 75.38 eV in Fig. 7c is associated with was found near the surface [25].
Al–O (α-Al2O3) bonds [21]. Fig. 7d represents the chemical composition Fig. 8a shows the TEM image with corresponding zoomed out area

7
M.N. Mathabathe et al. Materials Chemistry and Physics 257 (2021) 123703

Fig. 9. TEM images of the cyclically oxidized γ-Ti-46Al–2Nb-0.7Cr-0.3Si alloy at 900 ◦ C for 900 h. The corresponding insert of a zoomed-out particle is shown in (a),
the elemental mapping of (N, O, Al, Si, Ti, Cr, Cu, Nb) of the interface domain between the coating and the alloy (b), the high-resolution TEM indicating a faceted
particle (c) with the corresponding SAED (d).

indicating a γ-Ti-46Al–2Nb alloy faceted morphology. The micrograph assume a regular morphology. Fig. 9c illustrates the HRTEM of the
reveals the particles agglomerated on the carbon film and abroad the faceted particle. The red arrow designates the ordered lattice fringes.
margin of the holes within the film. This makes the process of attaining The diffraction pattern in Fig. 8d was taken from the dark ordered
high-resolution image complex because particles of interest are often particles in Fig. 8c. The SAED validated the crystalline nature of the
unnoticeable by the overlapping particles [26]. The EDS results in particles. Similar findings by Rozita et al. [26], revealed facets that
Figs. 8b, 9b and 10b was done to analyse the interfacial reaction be­ terminate with both (111) and (100) planes confirmed by partially
tween the oxide film and the alloy after cyclic oxidation at 900 ◦ C for disordered sub-surface visible in Fig. 9c. The phase indexed by (300) and
100 h. The TEM results justifies the oxide XPS spectra analysis. The (113) planes belongs to the alumina phase. The measured (100) plane
elemental mapping in Fig. 8b confirm that the scale is an oxynitride. The d-value is shown in the lattice fringe.
scanning transmission electron microscope (STEM) did not pick-up Nb Fig. 10 shows the density graphs of the annealed, nitrided and cyclic
nor Al. Besides, the HRTEM Fig. 8c revealed evidence of surface reform oxidized Ti–46Al–2Nb and Ti–46Al–2Nb-0.7Cr-0.3Si intermetallic al­
on a single-crystal faceted particle with the d-spacing of 0.3259 nm. The loys. The annealed Ti–46Al–2Nb alloy has a density of 3.954 g/cm3
corresponding SAED Fig. 8d suggests that the overall oxide layer is while nitridation at 900 ◦ C it decreased to 3.909 g/cm3. A slight increase
crystalline, despite the weak brightness and intensity of the other in the density of Ti–46Al–2Nb alloy to 3.912 g/cm3 was recorded after
polymorphic rings due to partial amorphous structure of the oxide film. cyclic oxidation attributable to oxide, nitride and oxynitride scale for­
Fig. 9a shows the TEM image of the cyclically oxidized γ-Ti- mations. This behavior is agreement with the slight mass gain shown in
46Al–2Nb-0.7Cr-0.3Si alloy with the complementary zoomed-out par­ Fig. 3. The 3.946 g/cm3 and 3.919 g/cm3 densities were measured for
ticle insert. The particle was characterized by STEM-EDS, which is the annealed and cyclically oxidized Ti–46Al–2Nb-0.7Cr-0.3Si inter­
represented by elemental mapping in Fig. 9b. The elemental particles metallic alloys, respectively. The cyclically oxidized Ti–46Al–2Nb-

8
M.N. Mathabathe et al. Materials Chemistry and Physics 257 (2021) 123703

Fig 12. Micro-hardness profile of the γ-TiAl alloys.

structure (B1-face centered cubic (FCC), which decreases the grain size,
by modifying its preferred crystalline orientation from (200) to (111)
Fig. 10. Density graphs of the annealed, nitrided and cyclic oxidized
Ti–46Al–2Nb and Ti–46Al–2Nb-0.7Cr-0.3Si intermetallic alloys.
[27], and inhibit the formation of hexagonal chromium nitride (Cr2N)
phase [23,24]. The previous study on the Ti–46Al–2Nb and
Ti–46Al–2Nb-0.7Cr-0.3Si alloys in as-cast and cyclically oxidized con­
ditions without nitridation treatment indicated a slight increase in
hardness [11] in comparison with the current findings. This behavior
was attributed to the brittleness prompted by the adsorption of inter­
mixed oxides on the surface of the alloys [11], contrary to the nitrided
scales that induced softness in the alloys in the current study. The
hardness values measured after Ti–46Al–2Nb alloy cyclic oxidation
without nitridation was 262 HV while that of Ti–46Al–2Nb-0.7Cr-0.3Si
alloy was 298HV [11], which is a decrease of 1.9%, and 8.7%, respec­
tively, to the nitridated alloys.
Fig. 11 illustrates the micro-hardness profile of the nitrided
Ti–46Al–2Nb and Ti–46Al–2Nb-0.7Cr-0.3Si alloys. The average profile
evaluation was conducted on the nitride surfaces and core of the sam­
ples. Generally, the hardness measurements obtained near the nitrided
surfaces are slightly higher than the core matrix. This is attributed to the
difference in the cooling effect between the core and the case of the
sample. Upon annealing, more Al was dissolved in the core structure due
Fig. 11. Micro-hardness profile of the Ti–46Al–2Nb and Ti–46Al–2Nb-0.7Cr- to its greater mobility than Ti. The effect of vacancies within the
0.3Si alloys.
structure promotes steadfast oxide growth throughout the cavities [28],
and accordingly the reduction in hardness in the core of the alloys is due
0.7Cr-0.3Si alloy also showed a slight rise in density compared to the to adsorption of oxides on the surface matrix causing deterioration/­
nitrided sample. Moreover, the density variation between the nitrided brittleness of the structure [11]. Abd El-Rahman et al. [29], reported
and cyclic oxidized alloys is very negligible in comparison to the nitrided that surface hardness primarily depends on the heat treatment temper­
and cyclic oxidized Ti–46Al–2Nb alloys. This behavior implies that the ature. Moreover, in nitrided samples, the surface enhancement is
Ti–46Al–2Nb-0.7Cr-0.3Si alloy has shown better stability after cyclic attributed to the formation of the nitride layer on the surface. The sur­
oxidation than the Ti–46Al–2Nb alloy where spallation was profound in face hardness of the Ti–46Al–2Nb-0.7Cr-0.3Si alloy is greater than
Fig. 3 between 700 and 900 h. Ti–46Al–2Nb alloy in a nitrided condition because the alloy is comprised
of the Cr contribution to the protective scale as indicated by the
composition in Figs. 7a and 9b. The presence of Cr2N in the γ-TiAl based
3.5. Macro and micro hardness profile of the γ-TiAl based alloys alloys display superior hardness although inclined to oxidation damage
resulting in degradation of the thermal stability of the scale [28].
Fig. 10. Macro-hardness profile of the γ-TiAl based alloys. Accordingly, the variation in surface hardness of the alloys is due to the
Fig. 10 illustrates the macro-Vickers hardness conducted on the relative substrate effect on the treated layer; along with small variations
cross-sections of the Ti–46Al–2Nb and Ti–46Al–2Nb-0.7Cr-0.3Si alloys in the microstructure as a consequence of different alloying elements
in annealed, nitrided and cyclically oxidized samples. The annealed [25,26,30]. The Al and Si metals in Ti–46Al–2Nb-0.7Cr-0.3Si alloy hard
samples had high hardness values when compared to all the other coatings have improved oxidation resistance due to the formation of
samples. Moreover, the bulk hardness results show that Ti–46Al–2Nb protective oxides on the surface layer and hindered ionic diffusion.
alloy is low in hardness. The nitrided Ti–46Al–2Nb-0.7Cr-0.3Si alloy Similar findings by Menand et al. [31] showed that high O and C affinity
hardness is also lower than that of the annealed sample. On the contrary, to γ-TiAl based alloys leads to an increase in the hardness values.(see fig
cyclic oxidation improved the hardness of the nitrided Ti–46Al–2Nb but 12)
softened that of the Ti–46Al–2Nb-0.7Cr-0.3Si alloy. Chang et al. [25]
reported comparable results and found that the coexisting of Al2O3 and 4. Conclusion
SiO2 near the surface retards the diffusion of oxygen into the
Cr0.38Al0.56Si0.06N coating, hence resulted in the improved oxidation The thermal properties, macro and micro-hardness, and micro­
behavior. Furthermore, in case of the Ti–46Al–2Nb-0.7Cr-0.3Si alloy, it structure of the vacuum melted, annealed and cyclically oxidized
is believed that the Al may have been incorporated into the CrN

9
M.N. Mathabathe et al. Materials Chemistry and Physics 257 (2021) 123703

Ti–46Al–2Nb and Ti–46Al–2Nb-0.7Cr-0.3Si alloys were studied. The [5] W. Zhu, D. Shi, Z. Zhu, J. Sun, Effect of the surface oxidization and nitridation on
the normal spectral emissivity of titanium alloys Ti-6Al-4V at 800-1100 K at a
following conclusions are drawn:
wavelength of 1.5 μm, Infrared Phys. Technol. 76 (2016) 200–205.
[6] K. Sopunna, T. Thongtem, M. McNallan, S. Thongtem, Surface modification of the
• The annealing of Ti–46Al–2Nb and Ti–46Al–2Nb-0.7Cr-0.3Si alloys γ-TiAl alloys by the nitridation, Surf. Sci. 566–568 (2004) 810–815.
has stabilized γ-TiAl +α2-Ti3Al phases. Moreover, the residual [7] B. Zhao, J. Wu, J. Sun, B. Tu, F. Wang, Oxidation kinetics of the nitrided TiAl-based
alloys, Mater. Lett. 56 (4) (2002) 533–538.
β-phase in Ti–46Al–2Nb, and the Ti5Si3 precipitates in Ti–46Al–2Nb- [8] B. Zhao, J. Sun, J.S. Wu, Z.X. Yuan, Gas nitriding behavior of TiAl based alloys in
0.7Cr-0.3Si alloys have emerged. an ammonia atmosphere, Scripta Mater. 46 (2002) 581–586.
• Due to surface nitridation, the oxidation properties of Ti–46Al–2Nb [9] A.M. Limarga, D.S. Wilkinson, A model for the effect of creep deformation and
intrinsic growth stress on oxide/nitride scale growth rates with application to the
and Ti–46Al–2Nb-0.7Cr-0.3Si alloys slightly improved but suffered nitridation of γ-TiAl, Mater. Sci. Eng. 415 (1–2) (2006) 94–103A.
the oxide spallation. [10] M.N. Mathabathe, A.S. Bolokang, G. Govender, C.W. Siyasiya, R.J. Mostert, Cold-
• The density of Ti–46Al–2Nb and Ti–46Al–2Nb-0.7Cr-0.3Si interme­ pressing and vacuum arc melting of γ-TiAl based alloys, Adv. Powder Technol. 30
(12) (2019) 2925–2939.
tallic alloys were decreased due to surface oxides, nitride and oxy­ [11] M.N. Mathabathe, A.S. Bolokang, G. Govender, R.J. Mostert, C.W. Siyasiya, The
nitride formations. vacuum melted ɣ -TiAl ( Nb, Cr, Si) -doped alloys and their cyclic oxidation
• TEM analysis confirmed partial amorphization in the Ti–46Al–2Nb properties, Vacuum 154 (2018) 82–89.
[12] M.N. Mathabathe, S. Govender, A.S. Bolokang, R.J. Mostert, C.W. Siyasiya, Phase
alloy and crystalline in the Ti–46Al–2Nb-0.7Cr-0.3Si alloy. transformation and microstructural control of the alpha-solidifying ɣ-Ti-45Al-2Nb-
• XPS chemical composition confirmed the presence of mixed oxide 0.7Cr-0.3Si intermetallic alloy, J. Alloys Compd. 757 (2018) 8–15.
layer of Al2O3, Cr2O3 and SiO2 near the Ti–46Al–2Nb-0.7Cr-0.7Si [13] M.N. Mathabathe, A.S. Bolokang, G. Govender, R.J. Mostert, C.W. Siyasiya,
Structure-property orientation relationship of a γ/α2/Ti5Si3 in as-cast Ti-45Al-2Nb-
surface.
0.7Cr-0.3Si intermetallic alloy, J. Alloys Compd. 765 (2018) 690–699.
[14] C. Yang, D. Hu, A. Huang, M. Dixon, Solidification and grain refinement in
CRediT authorship contribution statement Ti45Al2Mn2Nb1B subjected to fast cooling, Intermetallics 32 (2013) 64–71.
[15] J.J. Williams, Structure and High-Temperature Properties of Ti5Si3 with Interstitial
Additions, PhD thesis, 1999, p. 2494.
M.N. Mathabathe: Formal analysis, Writing - original draft, con­ [16] A. Abba, A. Galerie, M. Caillet, High-temperature oxidation of titanium silicide
ducted the experiments and prepared the manuscript. A.S. Bolokang: coatings on titanium, Oxid. Metals 17 (1) (1982) 43–54.
[17] C.M. Austin, T.J. Kelly, Gas turbine engine implementation of gamma titanium
Formal analysis, Writing - original draft, conducted the experiments and aluminide, Superalloys 1996 (1996) 539–543.
prepared the manuscript. C.W. Siyasiya: Formal analysis, Writing - [18] D.R. Askeland, The Science and Engineering of Materials, Springer US, 1996.
original draft, conducted the TEM and XPS discussions. R.J. Mostert: [19] K. Li, Y. Li, X. Huang, D. Gibson, Y. Zheng, J. Liu, L. Sun, Y.Q. Fu, Surface
microstructures and corrosion resistance of Ni-Ti-Nb shape memory thin films,
Formal analysis, Writing - original draft, conducted the TEM and XPS Appl. Surf. Sci. 414 (2017) 63–67.
discussions. All authors contributed in the analysis of the data, editing [20] Q. Li, L. Zhang, X. Chen, D. Wei, P. Zhang, Y. Chen, X. Xu, Characterization of
and commented on the manuscript. plasma rotating electrode atomized Nb-Ti based alloy powder, Met. Powder Rep
(2019) 1–10, https://doi.org/10.1016/j.mprp.2019.04.064.
[21] C. Badini, S.M. Deambrosis, O. Ostrovskaya, V. Zin, E. Padovano, E. Miorin, Cyclic
oxidation in burner rig of TiAlN coating deposited on Ti-48Al-2Cr- 2Nb by reactive
Declaration of competing interest HiPIMS, Ceram. Int. 43 (7) (2017) 5417–5426.
[22] A.S. Bolokang, Z.P. Tshabalala, G.F. Malgas, I. Kortidis, H.C. Swart, D.E. Motaung,
Room temperature ferromagnetism and CH4 gas sensing of titanium oxynitride
The authors, M. N. Mathabathe, A. S. Bolokang, G. Govender, C.W. induced by milling and annealing, Mater. Chem. Phys. 193 (2017) 512–523.
Siyasiya, R.J. Mostert declare that they have no known competing [23] J. Sjölén, L. Karlsson, S. Braun, R. Murdey, A. Hörling, L. Hultman, Structure and
financial interests or personal relationships that could have appeared to mechanical properties of arc evaporated Ti-Al–O–N thin films, Surf. Coating.
Technol. 201 (2007) 6392–6403.
influence the work reported in this paper. There is no conflict of interests [24] A.S. Bolokang, F.R. Cummings, B.P. Dhonge, H.M.I. Abdallah, T. Moyo, H.C. Swart,
for this article. C.J. Arendse, T.F.G. Muller, D.E. Motaung, Characteristics of the mechanical
milling on the room temperature ferromagnetism and sensing properties of TiO2
nanoparticles, Appl. Surf. Sci. 331 (2015) 362–372.
Acknowledgement [25] Y.Y. Chang, C.P. Chang, D.Y. Wang, S.M. Yang, W. Wu, High temperature oxidation
resistance of CrAlSiN coatings synthesized by a cathodic arc deposition process,
This work was supported by the Department of Science and Inno­ J. Alloys Compd. 461 (1–2) (2008) 336–341.
[26] Y. Rozita, R. Brydson, A.J. Scott, An investigation of commercial gamma-Al2O3
vation (DSI), Republic of South Africa through Grant number : nanoparticles, J. Phys. Conf. Ser. 241 (2010), 012096.
HCARD08. [27] E. Mohammadpour, Z.-T. Jiang, M. Altarawneh, N. Mondinos, M.M. Rahman, H.
N. Lim, N.M. Huang, Z. Xie, Z.-F. Zhou, B.Z. Dlugogorski, Experimental and
predicted mechanical properties of Cr1− xAIxN thin films, at high temperatures,
References incorporating in situ synchrotron radiation X-ray diffraction and computational
modelling, RSC Adv. 7 (36) (2017) 22094–22104.
[1] C.G. McKamey, S.H. Whang, C. Liu, Microstructural characterization of a γ-TiAl-Ni [28] M.N. Mathabathe, G. Govender, C.W. Siyasiya, R.J. Mostert, A.S. Bolokang, Surface
alloy produced by rapid solidification techniques, Scripta Metall. 32 (3) (1995) characterization of the cyclically oxidized γ-Ti-48Al-2Nb-0.7Cr alloy after
383–388. nitridation, Mater. Char. 154 (2019) 94–102.
[2] D.J. Kim, D.Y. Seo, X. Huang, Q. Yang, Y.W. Kim, Cyclic oxidation behavior of a [29] A.M. Abd El-Rahman, M.F. Maitz, M.A. Kassem, F.M. El-Hossary, F. Prokert,
beta gamma powder metallurgy TiAl-4Nb-3Mn alloy coated with a NiCrAlY H. Reuther, M.T. Pham, E. Richter, Surface improvement and biocompatibility of
coating, Surf. Coating. Technol. 206 (13) (2012) 3048–3054. TiAl24Nb10 intermetallic alloy using rf plasma nitriding, Appl. Surf. Sci. 253 (23)
[3] A.S. Bolokang, D.E. Motaung, C.J. Arendse, T.F.G. Muller, Formation of the (2007) 9067–9072.
metastable FCC phase by ball milling and annealing of titanium-stearic acid [30] A.I.P. Nwobu, R.D. Rawlings, D.R.F. West, Nitride formation in titanium based
powder, Adv. Powder Technol. 26 (2) (2015) 632–639. substrates during laser surface melting in nitrogen-argon atmospheres, Acta Mater.
[4] M. Song, M. Xiang, Y. Yang, Q. Zhu, C. Hu, P. Lv, Synthesis of stoichiometric TiN 47 (2) (1999) 631–643A.
from TiH2 powder and its nitridation mechanism, Ceram. Int. 44 (14) (2018) [31] A. Menand, A. Huguet, A. Nerac-Partaix, Interstitial solubility in γ and α2 phases of
16947–16952. TiAl-based alloys, Acta Mater. 44 (1996) 4729–4737.

10

You might also like