Journal of Materials Research and Technology: Tiancheng Cui, Shujiang Geng, Minghui Chen, Fuhui Wang
Journal of Materials Research and Technology: Tiancheng Cui, Shujiang Geng, Minghui Chen, Fuhui Wang
Journal of Materials Research and Technology: Tiancheng Cui, Shujiang Geng, Minghui Chen, Fuhui Wang
A R T I C L E I N F O A B S T R A C T
Keywords: DD5 single-crystal alloy samples with different surface orientations were obtained by machining the alloy ingot
DD5 SC alloy perpendicular (PP) and parallel (P), respectively, to its solidification direction (SD). The corrosion behaviors of
Surface orientation the PPSD and PSD alloy samples were investigated in the atmosphere of moist air + NaCl aerosol at 750 ◦ C.
NaCl aerosol atmosphere
Results showed that during initial corrosion, γ′ phase developed a thin NiO layer with Al internal corrosion
High temperature corrosion
product. However, a NiO nodule with an internal Cr2O3 layer was formed on γ phase. As corrosion progressed,
the corrosion product gradually grew to a NiO/(Al, Cr)2O3 bilayer, accompanied by internal corrosion consisting
of Al2O3. It was also found that the PSD sample with a lower fraction and smaller size of γ’ phases exhibited
better corrosion resistance than the PPSD sample. The effect of alloy surface orientation on the corrosion
mechanism was discussed.
1. Introduction [10,11]. Several studies revealed that a dense and continuous α-Al2O3
layer will be formed on the surface of the Ni-based SC alloys during high
Over the past few decades, the gas turbine has become the key power temperature corrosion, thereby effectively reducing the corrosion rate at
unit in diverse applications including industrial production, power elevated temperatures [12–14]. However, during gas turbine operation,
generation, and marine propulsion systems [1]. Researchers have con NaCl carried by the coolant significantly accelerates the corrosion of the
ducted extensive studies to enhance blade properties, aiming to main Cr-containing alloys, including Ni-based SC alloys [15–17]. Conse
tain optimal mechanical performance of gas turbines in harsh operating quently, it is essential to further investigate the corrosion mechanisms of
environments. On the one hand, hollow structures with film cooling Ni-based SC alloys in marine environments. Moreover, the complex
channels have been adopted to enhance the heat transfer performance channel geometry results in surfaces exposed to corrosive atmospheres
2–4. On the other hand, due to the high temperature capabilities such as at varying angles to the typical [001] solidification direction of Ni-based
fatigue resistance, creep strength, and structural stability, Ni-based SC alloys, which is crucial for enhancing the mechanical properties [18,
single-crystal (SC) alloys have been widely applied in turbine blades 19]. This orientation effect makes the corrosion behavior more
to extend the high-temperature service life 5–7. Film cooling is a tech complicated. As highlighted by Montero et al., corrosion caused by
nique that introduces ambient air as coolant to high-temperature turbine invasive fluid corrosive media such as molten salt exhibits a strong
blade surfaces, which will effectively reduce the operating temperature orientation dependence [20]. Despite its importance in guiding material
of gas turbine components. However, the temperature of the internal optimization, the effect of surface orientation on the corrosion behavior
cooling channel surfaces can still reach 750 ◦ C [8,9]. In recent years, gas of internal cooling channel surfaces exposed to NaCl aerosol atmo
turbines have become increasingly applied in offshore environments like spheres remains insufficiently studied.
ships, wind power platforms, and oil drilling rigs. Therefore, the service Current paper aims to provide a detailed characterization and
conditions of gas turbine blades become more complex. Coolants drawn investigation of the corrosion behaviors of the DD5 SC alloys in the
from the outside NaCl-contained atmosphere can cause significant simulated corrosion environment of gas turbine blades, with a focus on
corrosion of the internal surfaces of cooling channels at such high the corrosion mechanism of the alloys with different surface orientations
temperatures. As a result, corrosion on the cooling channel surfaces has to understand the effect of surface orientations on corrosion behavior.
been a key challenge limiting the service life of the gas turbine blades The alloys with different surface orientations in this work were obtained
* Corresponding author.
E-mail address: gengsj@smm.neu.edu.cn (S. Geng).
https://doi.org/10.1016/j.jmrt.2024.09.212
Received 10 August 2024; Received in revised form 19 September 2024; Accepted 24 September 2024
Available online 25 September 2024
2238-7854/© 2024 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-
nc-nd/4.0/).
T. Cui et al. Journal of Materials Research and Technology 33 (2024) 2199–2209
2. Experimental procedures
2.1. Materials
The raw materials were Ni-based SC alloy ingots with the brand DD5
provided by the Institute of Metal Research, Chinese Academy of Sci
ences (CAS). The chemical compositions of the alloy are listed in Table 1.
Fig. 1. Schematic diagram of alloy machining direction.
For the convenience of testing, the DD5 alloy ingot was machined
into 15 × 10 × 2 mm3 samples with surfaces oriented perpendicular and
parallel to the solidification direction, as presented in Fig. 1. These alloy
samples were designated as PPSD (Perpendicular to Solidification Di
rection) and PSD (Parallel to Solidification Direction) samples, respec
tively. All the surfaces of these alloy samples were ground evenly with
grit-1000 abrasive paper.
Table 1
The chemical compositions of the DD5 single-crystal alloy.
Element Ni Co Cr Al Ta W Re Mo Hf
Content (wt.%) Bal 7.76 6.99 6.43 6.42 5.12 3.05 1.52 0.14
2200
T. Cui et al. Journal of Materials Research and Technology 33 (2024) 2199–2209
Fig. 3. (a) Schematic diagram of alloy machining direction (a1, a2) SEM (BSE) and phase statistical distribution images of the surface morphology of the PPSD and
PSD alloy samples (b) XRD result of PPSD and PSD alloys (c1, c2) Statistical fraction of the γ′ phase with different average sizes on the surface of PPSD and PSD
alloy samples.
only consist of γ/γ′ phase with a crystal plane index of (200). The BSE
images of the PPSD and PSD samples are illustrated in Fig. 3 (a1, a2). It ± 4.2% and 72.6 ± 5.1%, respectively.
can be seen that the cubic-shaped γ′ phases coherently embed in the γ Fig. 4 (a) depicts the mass gains of the alloys corroded in the two
substrate. The statistical fractions of γ′ phases with different sizes on the different test environments (air + water vapor and air + NaCl + water
surfaces of the PPSD and PSD samples are shown in Fig. 3 (c1, c2), which vapor) at 750 ◦ C. The PSD and PPSD samples both exhibit continuous
were measured with the software Nano Measurer. The average sizes of mass gains over time. After 168 h of corrosion, the total mass gain of the
the γ′ phases on the surfaces of the PPSD and PSD samples are about 560 PSD sample is about 0.13 mg/cm2, and the PPSD sample shows a slightly
± 48 nm and 430 ± 64 nm, respectively. In addition, the chemical higher mass gain, reaching 0.16 mg/cm2. Mass gains become signifi
compositions of γ and γ’ detected with the Energy Dispersive Spectros cantly larger when the alloy samples are exposed to air + NaCl + water
copy (EDS) are listed in Table 2. vapor due to the inducement of NaCl. During the short-term corrosion
Furthermore, the distributions of the γ′ phase on the surface of the process (0–48 h), the samples exhibit a rapid mass gain rate. As the test
PSD and the PPSD samples are different. The software Image Pro was time extends over 48 h, the mass gain rates of the alloy samples grad
used to count and calculate the fraction of pixels occupied by the γ’ ually slow down, and the total mass gains of the PSD and PPSD samples
phases, Cγ’ , and the value can be calculated as follows: are about 0.79 mg/cm2 and 0.91 mg/cm2, respectively. The PPSD
Aγ’ sample still exhibited slightly higher mass gain than that of the PSD
Cγ’ = × 100% (1) sample. The mass gains of the alloy samples corroded in the two
Aγ’ + Aγ
corrosion environments both conform to the parabolic law [22]:
Where Aγ’ and Aγ denote the area occupied by γ′ and γ phases in the SEM (Δm)2
pictures, respectively. The values of Cγʹ are calculated using Image Pro to = kp ⋅t (2)
A
count the proportion of pixels occupied by the γ′ phases. As calculated,
Cγʹ on the surfaces of the PPSD and PSD samples are approximately 84.2 Where Δm is the mass gain, A is the alloy surface area, t is the corrosion
Table 2
Chemical compositions of γ and γ’ (at.%).
Element Ni Co Cr Al Ta W Re Mo Hf
2201
T. Cui et al. Journal of Materials Research and Technology 33 (2024) 2199–2209
Fig. 4. Corrosion kinetics of the samples exposed to air + water vapor and air + NaCl aerosol + water vapor at 750 ◦ C (a) specific mass gains and (b) the linear fitting
curves of squared mass gains.
time, and kp is the parabolic rate constant. The linear fitting curves of the after being corroded for 24 h at 750 ◦ C (Fig. 7 (b)). This phenomenon
squared mass gains of the alloy samples are plotted in Fig. 4 (b). It can be also corroborates the higher mass gain of the PPSD sample as presented
seen that the calculated kp values of the PSD and PPSD samples are 1.69 in Fig. 4. After 168 h of corrosion (Fig. 7 (d)), the oxide scale grows
× 10− 4, 2.51 × 10− 4 mg2cm− 4h− 1 in the atmosphere of air + water thicker and becomes an obvious bilayer structure. Internal corrosion
vapor and 6.12 × 10− 3, 8.50 × 10− 3 mg2cm− 4h− 1 in the atmosphere of with a maximum depth of 5.2 μm can also be observed at the interface
air + NaCl + water vapor during the initial 48 h, respectively. During between the internal oxide layer and substrate, which is slighter than
48–168 h of corrosion, the corresponding kp values decrease signifi that of the PPSD alloy. Based on the EDS mappings, the PSD sample
cantly, with the values of 6.78 × 10− 5, 1.06 × 10− 4 mg2cm− 4h− 1 in the shows a similar element distribution to the PPSD sample. The results
air + water vapor environment, and 2.76 × 10− 3, 3.31 × 10− 3 above show that surface orientation has a significant effect on the
mg2cm− 4h− 1 in the atmosphere of air + NaCl + water vapor. The results corrosion of the alloy samples. Corrosion of the PPSD sample is more
show that the PPSD samples have higher kp values than those of the PSD severe than that of the PSD sample. Combining the XRD analysis of the
samples when the alloy samples are corroded in both two environments. alloy samples after corrosion in the atmosphere of air + NaCl + water
Furthermore, the different kp indicates distinct oxide scale formation vapor (Fig. 6 (b)), the corrosion products mainly consist of NiO, Al2O3,
mechanisms for different corrosion processes, which will be discussed and Cr2O3.
later. To study the initial corrosion behaviors of the alloy samples, both the
Fig. 5 illustrates the cross-sectional morphologies and elements dis PPSD and PSD samples were exposed to the air + NaCl + water vapor
tribution of the tested alloys after corrosion in the atmosphere of air + atmosphere at 750 ◦ C for 10 min. Prior to the corrosion test, all the
water vapor at 750 ◦ C. After 24 h of corrosion, both PPSD and PSD surfaces of the alloy samples were polished with the diamond polishing
samples form uniform and dense oxide scales with a thickness of about slurry to obtain a smooth surface, eliminating the influence of surface
180 nm (Fig. 5 (a and b)). As the corrosion duration extends to 168 h, the roughness on corrosion behavior. The surface morphologies of the alloy
oxide scales uniformly grow thicker to about 370 nm (Fig. 5 (c and d)). samples after corrosion are illustrated in Fig. 8. It can be seen that
Corresponding EDS results reveal that the oxide scales mainly consist of numerous oxide nodules with NiO as the main composition are formed
external Ni oxide layer and internal Al oxide layer. In addition, a small on the γ phase surfaces of both PPSD and PSD samples. The average
amount of Cr oxides can also be observed in the internal oxide layer after width of these oxide nodules on the surface of the PPSD sample is about
168 h of corrosion. 340 nm (Fig. 8 (a)), which is slightly wider than the average width of the
Fig. 6 shows the XRD analysis of the alloy samples after being oxide nodules on the surface of the PSD sample (~220 nm) as shown in
corroded at 750 ◦ C for 168 h. After 168 h of the corrosion in the air + Fig. 8 (b).
water vapor environment (Fig. 6 (a)), NiO and Al2O3 are the main Fig. 9 demonstrates the TEM analysis of the PPSD and PSD samples
corrosion products. Meanwhile, a small amount of Cr2O3 is also formed. after corrosion in the atmosphere of air + NaCl + water vapor for 10
The cross-sectional morphologies and elements distribution of the min. It can be seen from Fig. 9 (a) that the PPSD sample shows a γ/γ′
samples after corrosion in the atmosphere of air + NaCl + water vapor phase coherent structure. EDS mappings show that the Al element is
are presented in Fig. 7. The results reveal that the presence of NaCl mainly enriched in the γ′ phase, while the Cr and Co elements are pri
significantly accelerates the corrosion of the alloy samples. After 24 h of marily segregated in the γ phase. After 10 min of corrosion, oxide
exposure, the PPSD sample develops an oxide scale with internal nodules consisting of external NiO layer and internal Cr2O3 layer are
corrosion of about 3.2 μm (Fig. 7 (a)). EDS analysis reveals a bilayer formed on the surface of γ phase. Obviously, no internal corrosion occurs
structure of the oxide scale: an external Ni oxide layer and an internal in the γ phase. However, the γ′ phase surfaces mainly develop relatively
Al/Cr oxide layer, with additional internal corrosion product of Al ox thin NiO layers with Al2O3 inner layers besides some areas with the thick
ides. After 168 h of exposure, corrosion becomes far more severe, as NiO layer marked by the dashed rectangle in Fig. 9 (a). Moreover, severe
evidenced by the thicker oxide scale and deeper internal corrosion with internal corrosion, reaching a depth of approximately 480 nm, occurs in
a maximum depth of about 6.2 μm as shown in Fig. 7 (c). In addition, the the γ′ phase beneath the NiO layer. The corresponding EDS mappings
internal corrosion at the interface between the internal oxide layer and reveal that the primary internal corrosion product is Al2O3. In addition,
alloy substrate also becomes more severe. Notably, the internal corro a small amount of Cr oxides can be observed at the NiO/Al2O3 interface.
sion mainly occurs in the γ′ phase, which indicates that the γ’ phase is HRTEM analysis was conducted on two selected areas (marked by white
easier to be corroded than the coherent γ phase. Regarding the PSD rectangles in Fig. 9 (a)) to distinguish the corrosion product composi
sample, the corrosion is less severe than that on the PPSD sample surface tions. Areas 1 and 2 (Fig. 9 (a1, a2)) correspond to the corrosion
2202
T. Cui et al. Journal of Materials Research and Technology 33 (2024) 2199–2209
Fig. 5. Cross-sectional morphologies and the corresponding EDS mappings of the alloy samples exposed to air + water vapor at 750 ◦ C (a) PPSD sample exposed for
24 h, (b) PSD sample exposed for 24 h, (c) PPSD sample exposed for 168 h, (d) PSD sample exposed for 168 h.
products of the γ and γ′ phases, respectively. In area 1, NiO (d = 2.09 Å TEM results above show that the PPSD sample exhibits more severe
and 2.41 Å) is distributed in the oxide nodule on the γ phase surface, and corrosion than the PSD sample after 10 min of NaCl-induced corrosion.
Cr2O3 (d = 1.67 Å, 2.08 Å, and 2.41 Å) mainly exists at the NiO layer/γ
interface; In area 2, NiO (d = 2.09 Å and 2.41 Å) and θ-Al2O3 (d = 2.01 4. Discussion
Å) are the primary compositions of the oxide scale and internal corrosion
products of the γ’ phase, respectively. 4.1. Corrosion in the air + water vapor
Fig. 9 (b) presents the TEM images of the PSD sample. It can be seen
that the corrosion products of the PSD sample exhibit similar mor During the corrosion in the atmosphere of air + water vapor, the
phologies to those of the PPSD sample. Nevertheless, the PSD sample, external NiO layer is formed first on the sample surface, followed by the
which develops smaller oxide nodules and shallower internal corrosion formation of the internal Al2O3 layer as shown in Fig. 5(c and d).
(maximum depth ~270 nm), shows slighter corrosion compared to the Furthermore, a small amount of Cr2O3 is also formed in the internal
PPSD sample. The HRTEM images of the corrosion products on the γ and oxide layer based on the XRD results (Fig. 6 (a)). In order to clearly
γ′ phases marked by two white rectangles are shown in Fig. 9 (b1, b2). explain the formation sequence of corrosion product, the Ellingham
The distribution of the corrosion products also resembles that of the diagram of the possible oxidation reactions during corrosion is pre
PPSD sample. Specifically, NiO/Cr2O3 nodules are formed on the γ phase sented in Fig. 10. It can be seen that the comparison of oxygen affinity of
surfaces, and relatively thin NiO/Al2O3 bilayers grow on the γ’ phase the elements at 750 ◦ C is listed as follows: Al (− 902 kJ/mol) > Cr (− 576
surfaces accompanied by the internal corrosion product of θ-Al2O3. The kJ/mol) > Co (− 323 kJ/mol) > Ni (− 292 kJ/mol) > Re (− 211 kJ/mol).
2203
T. Cui et al. Journal of Materials Research and Technology 33 (2024) 2199–2209
Fig. 6. XRD results of the alloy samples exposed to (a) air + water vapor and (b) air + NaCl + water vapor at 750 ◦ C for 168h
It is obvious that Al has a higher oxidation tendency compared to Cr and the volatilized NaCl will attack the alloy samples in the aerosol state. It
Ni. However, as observed in Fig. 5, the oxide scale of the alloy sample is has been reported that NaCl will first react with the metallic elements in
composed of an external NiO layer and an internal (Al, Cr)2O3 layer. the alloy samples and convert to other corrosive media such as Cl2 at
According to Wagner’s theory, there exists a critical concentration, 750 ◦ C [32–34]. The equations of the possible corrosion reactions and
NA(c) , for the formation of the continuous oxide layer of element A their corresponding Gibbs free energy at 750 ◦ C are listed as follows [35,
during corrosion [23]: 36]:
( )1 Ni + H2O(g) + 2NaCl(s/g) + O2(g) = NiO + 2NaOH + Cl2(g)
πgNO DO Vm 2
NA(c) = (3)
2εDA VmAO ΔG750 ◦
C = 103.7 kJ/mol (NaCl(s)) or − 51.1 kJ/mol (NaCl(g)) (6)
Where ε and g are constants; NO is the oxygen permeability at the scale/ 0.8Al + 0.8H2O(g) + 1.6NaCl(s/g) + O2(g) = 0.4Al2O3 + 1.6NaOH +
alloy interface; DO is the diffusion coefficient of oxygen in the alloy; Vm 0.8Cl2(g)
AO
and Vm are the molar volume of alloy and the oxide of A, respectively;
ΔG750 ◦
C = − 341.1 kJ/mol (NaCl(s)) or − 464.9 kJ/mol (NaCl(g)) (7)
DA is the diffusion coefficient of the element A in the alloy. At the
beginning of the corrosion, NAl (the relative concentration of Al) < NAl(c) 0.8Cr + 0.8H2O(g) + 1.6NaCl(s/g) + O2(g) = 0.4Cr2O3 + 1.6NaOH +
(the critical concentration of Al), and NiO is the primary corrosion 0.8Cl2(g)
product on the surface due to the high concentration of Ni. During this
ΔG750 C = − 146.7 kJ/mol (NaCl(s)) or − 269.8 kJ/mol (NaCl(g)) (8)
stage, Ni consumption leads to an increase in NAl at the NiO layer/alloy
◦
interface. Concurrently, NiO layer formation on the surface decreases It can be seen from the above equations that Ni is the most stable
the oxygen partial pressure. This process reduces the NO in Eq. (3), element under the inducement of either solid or gaseous NaCl at 750 ◦ C
subsequently lowering NAl(c). The two factors promote the formation of a because of the higher Gibbs free energy. In contrast to Ni, Al and Cr in
continuous Al2O3 layer beneath the NiO layer, accompanied by the the alloy sample are more prone to react with the NaCl aerosol pro
formation of a small amount of Cr2O3 as the corrosion progresses. ducing Cl2. The trace of Cl2 reacts with the metal elements in the alloy
Combining the effect of the water vapor, the equations of the possible sample forming MClx, which are prone to be volatilized at high tem
reaction occurring during the corrosion process are listed as follows peratures. These volatile M chlorides gradually migrate outward,
[24–26]: reacting with ambient oxygen to form porous corrosion products and
release Cl2. This oxychlorination process can be described via the
M + xH2O(g) = M(OH)x + x/2 H2(g) (4)
following reactions:
M(OH)x = MOx/2 +x/2 H2O(g) (5)
2 M + xCl2(g) = 2MClx(g) (9)
Where M represents the metallic elements in the alloys such as Ni, Cr,
2MClx(g) + x/2O2(g) = 2MOx/2 + xCl2(g) (10)
and Al. The above reactions show that water vapor can directly partic
ipate in the corrosion reaction and form corrosion products. Moreover, Where M represents the metallic elements in the alloy sample such as Ni,
the interstitial protons generated by water dissociation will also dissolve Al, and Cr. It can be seen that Cl2 will be released again and continuously
in the oxide scale, reducing the vacancy formation energy [27]. There take part in the corrosion reaction simultaneously, further accelerating
fore, the vacancy density of the oxide scale increases, enhancing the the corrosion rate. It is reasonable to presume that Cl2 plays a catalytic
lattice diffusion of the metallic element atoms [28]. role during the entire corrosion process [37,38].
Fig. 11 shows the initial NaCl-induced corrosion mechanism of the
alloy samples. According to the discussion above, it can be assumed that
4.2. Corrosion in the air + NaCl + water vapor γ and γ′ phases are corroded independently with negligible interaction
during the initial corrosion. As described in Fig. 11 (a), Ni in the alloy is
When exposing in the atmosphere of air + NaCl + water vapor, both first oxychlorinated to form NiO on both the surfaces of the γ and γ′
PPSD and PSD samples develop bilayer oxide scales consisting of an phase during the initial stage of the corrosion. Notably, it can be seen
external NiO layer and an internal (Al, Cr)2O3 layer. Unlike the typical that NiO tends to accumulate on the surface of the γ phase than the γ′
solid salt corrosion [29–31], when exposed to air + NaCl + water vapor,
2204
T. Cui et al. Journal of Materials Research and Technology 33 (2024) 2199–2209
Fig. 7. Cross-sectional morphologies and the corresponding EDS mappings of the alloy samples exposed to air + NaCl + water vapor at 750 ◦ C. (a) PPSD sample
exposed for 24 h, (b) PSD sample exposed for 24 h, (c) PPSD sample exposed for 168 h, (d) PSD sample exposed for 168 h.
phase as shown in the TEM results, despite the higher Ni concentration layer grows on the γ′ phase surface. Furthermore, higher Ni consumption
in the γ′ phase. Wu et al. reported that in Ni-based single-crystal (SC) rate in the γ phase (where NiO remains the primary corrosion product at
alloys, the γ phase exists as a disordered solid solution, while the γ′ phase this stage) causes a Ni concentration gradient between γ′ and γ phases.
exhibits an ordered structure with specific stoichiometry. The diffusion This promotes the diffusion of Ni from γ’ to γ phase [41], as proved by
rate of Ni in the disordered γ phase is higher than that in the ordered γ′ the EDS mapping of Ni in Fig. 9.
phase [39]. Consequently, the formation rate of NiO on the γ phase is As corrosion progresses, the oxide nodules on the γ phases gradually
faster than that on the γ′ phase. Additionally, Fig. 9 (a) shows that a enlarge and coalesce, forming a porous external oxide scale. The cor
comparatively thick NiO layer is formed on the γ′ phase at some area. rosive medium can easily diffuse through the loose oxide scale and
This may be resulted from the corrosion of the lateral γ phase on the continue to corrode the sample, leading to the formation of the internal
sample surface. Meanwhile, the internal corrosion occurring beneath the oxide scale. With the further development of corrosion, the oxide scale
NiO layer is also more severe. As the NiO layer grows on the γ phase eventually develops into an obvious bilayer structure of an external NiO
surface, oxygen partial pressure at the NiO layer/γ interface decreases. layer and an internal (Al, Cr)2O3 layer. This corrosion product formation
Simultaneously, the relative Cr concentration gradually increases, mechanism also aligns with the fitting curves of the squared mass gains
leading to the formation of the Cr2O3 layer on the surface of γ phase. as shown in Fig. 4 (b). Since the mass gain conforms to the parabolic law,
Similarly, θ-Al2O3, which is prone to be formed in environments with the corrosion is controlled by the diffusion of the elements, especially Ni,
low oxygen partial pressure and relatively low temperatures [40], be Al and O [42]. During the initial 48 h of the corrosion, the oxide scale
comes the primary internal corrosion product in the γ′ phase as the NiO formed by nodule coalescence remains porous, so that it can not
2205
T. Cui et al. Journal of Materials Research and Technology 33 (2024) 2199–2209
Fig. 8. Surface morphologies of the (a) PPSD sample (b) PSD sample exposed to air + NaCl + water vapor at 750 ◦ C for 10 min.
sufficiently prevent the inward diffusion of the corrosive media [43,44]. same corrosion condition. Therefore, DPSD PPSD
Alγ’ and DAlγ’ (the diffusion co
This results in a relatively higher parabolic rate constant kp at this stage. efficients of Al in the γ′ phases of the PPSD and PSD samples) determine
As the oxide nodules grow larger to form a continuous external oxide the mathematical relationship of NAlγ’(c) and NʹAlγ’(c) of the PPSD and PSD
scale, the inward diffusion of oxygen is also effectively impeded.
samples. For the common polycrystalline Ni-based alloys, the overall
Therefore, kp also decreases significantly during the long-term
diffusion rate of the element in the alloy is the sum of the rates through
corrosion.
the bulk lattices and grain boundaries [47]. Whereas, the Ni-based SC
alloy exhibits a distinct microstructure comprising coherent γ and γ′
4.3. Effect of the surface orientation on corrosion behaviors phases because of their close lattice parameters [48]. Despite the co
herency, the γ/γ′ interfaces in the Ni-based SC alloy still contain plenty
It can be concluded from the above results that the surface orienta of misfit dislocations [49,50]. Consequently, these interfaces resemble
tion has an obvious effect on the corrosion behavior. PPSD sample ex sub-grain boundaries, a type of low-angle grain boundary that can
hibits a faster corrosion rate, especially during the NaCl-induced enhance Al diffusion rates [51]. Thus, the diffusion behavior of Al in
corrosion, as evidenced by the higher mass gain (Fig. 4) and more severe Ni-based SC alloys is similar to those in polycrystalline alloys, and the
corrosion (Figs. 7 and 9). The key distinctions between PSD and PPSD
total Al diffusion rate in the γ′ phase (DAlγ’ ) includes the diffusion
samples are the different fractions (Cγʹ ) and the size of the γ′ phases on
through the γ′ phase and the interface diffusion along the γ/γ’ phases
the surfaces as shown in Fig. 3. As the results mentioned above, the γ′
boundaries, which can be expressed by the following equation [52]:
phase is more prone to be corroded than the γ phase. It is believed that
the corrosion of the γ′ phase is crucial to determine the corrosion rate of DAlγ’ = Dγ’
Al (1 − f) + DAl f
γ/γʹ
(13)
the alloy sample. The formation mechanism of the main internal
corrosion products, Al2O3, in the γ′ phases during the initial corrosion Where Dγ’
γ/γʹ
Al and DAl are the diffusion coefficients of Al through the γ′
can be explained by Wagner’s theory. Firstly, the critical concentration phase and the γ/γ′ interfaces at 750 ◦ C, respectively. f is the fraction of
of Al in the γ′ phases for the formation of the continuous Al2O3 layer can the volume occupied by γ/γ′ interfaces.
be rewritten as [45]: Based on the aforementioned discussion, the γ′ phases and γ/γ′ in
( )1 terfaces can be analogously regarded as the grain lattices and grain
2
πgNO DO Vmγ’ boundaries, respectively. Fig. 3 (a1) and (a2) reveal that γ′ phases in the
NAlγ’(c) = (11)
2εDAlγ’ VmAl2 O3 DD5 SC alloy generally exhibit a cubic distribution, bordered by γ phase
matrix on all the four sides. Hence, f is approximately equal to the width
As mentioned above, NAlγ’ (the concentration of Al in the γ′ phase) is
fraction of γ/γ’ interfaces [53]. It can be obtained by the
lower than NAlγ’(c) during the initial stage of corrosion, and NiO is first two-dimensional calculation method applied in polycrystalline alloy, as
formed on the surface of the γ′ phase. With the consumption of the Ni, follows:
NAlγ’ also gradually increases above NAlγ’(c) at the interface of the thin NiO
layer and γ′ phase, leading to the formation of the continuous Al2O3 4σ
f= (14)
layer. Once the Al2O3 layer is formed, Al element in the γ′ phase is d
consumed to maintain the growth of the Al2O3 layer continuously. The Where σ is the γ/γ′ interface width (~1.3 nm for Ni-based SC alloys after
critical concentration of Al in the γ’ phases that can maintain the growth
solution heat treatment [54,55]), and d is the size of γ′ phase on the
of the continuous Al2O3 layer, NʹAlγʹ(c) , should conform to the equation
surface of the alloy samples as shown in Fig. 2(c1, c2). Regarding the
given by Ref. [46]: PSD samples, Eq. (13) can be rewritten as follows:
( )1
Vmγ’ πkAl 2 O3 2 dPSD − 4σ γ/γʹ 4σ
Nʹ
p
(12) DPSD
Alγ’ = DAl
γ’
+ DAl (15)
Alγʹ(c) =
DAlγ’ dPSD dPSD
VmAl2 O3
Further simplifications yield Eq. (15), we can obtain:
Where kAl
p
2 O3
is the parabolic rate constant of Al2O3, Vm
γ’ Al2 O3
and Vm are ( ) 4σ
γ/γʹ
the molar volumes of γ′ phase and Al2O3, respectively; DAlγ’ is the DPSD
Alγ’ = DAl +
γ’
DAl − Dγ’ (15-1)
Al
dPSD
diffusion coefficient of Al in the γ′ phase at 750 ◦ C.
Al2 O3
Similarly, for the PPSD samples, the sum diffusion coefficient of Al
In Eq. (11) and Eq. (12), Vm
γ’
, Vm , and kAl
p
2 O3
are constant in the
2206
T. Cui et al. Journal of Materials Research and Technology 33 (2024) 2199–2209
we have:
( )( 4σ 4σ
)
γ/γʹ
DPSD − DPPSD
= D − Dγ’
− (17)
Alγ’ Alγ’ Al Al
dPSD dPPSD
Poletaev found that low angle grain boundary can also enhance the
diffusion of the Al element [56]. Therefore, it is convinced that the
diffusion coefficient of Al is higher at the γ/γ′ interface, which means
ʹ
Dγ/γ
Al > DAl in Eq. (17). Meanwhile, notice that the average size of γ′
γ’
phases on the surface of the PSD sample is about 0.43 μm, smaller than
those on the PPSD sample surface (~0.56 μm) as depicted in Fig. 3 (c1,
c2), which means dPSD < dPPSD . Therefore, we have DPSD
Alγ’ > DAlγ’ from Eq.
PPSD
PSD
(17). Finally, it can be calculated that NAlγ’(c)
PSD
< NPPSD
Alγ’(c) and N Alγ’(c) <
ʹ
PPSD
NʹAlγ’(c) from Eq. (11) and Eq. (12), respectively. As a result, the Al2O3
formed in the γ′ phases of the PSD sample exhibits better continuity and
protective properties than that of the PPSD sample. This is due to the
lower NAlγ’(c)
PSD
and NʹPSD
Alγ’(c) , which benefit the original formation and the
subsequent growth maintenance of the continuous Al2O3 layer, respec
tively. The insufficient protection of the Al2O3 layer results in more
severe corrosion in the γ′ phases of the PPSD sample than PSD sample, as
displayed in the TEM images in Fig. 9. Consequently, the
corrosion-induced Al depletion elevates the relative Ni concentration in
γ′ phases of the PPSD sample, thereby increasing the Ni concentration
gradient between γ and γ′ phases. This enhances the diffusion of Ni from
γ′ phases to the γ phases, promoting the formation of oxide nodules on
the γ phase surfaces of the PPSD sample. Furthermore, the higher γ’
phase area fraction of the PPSD sample, which can be observed in Fig. 3
(a1, a2), indicates a larger area of severe internal corrosion than the PSD
sample. The overall corrosion of the PPSD sample is subsequently
accelerated. As the corrosion progresses, the PPSD sample finally forms
a thicker oxide scale compared to the PSD sample, as demonstrated in
Fig. 7.
Fig. 9. TEM-HAADF images and the HRTEM patterns of the selected areas of
the as-corroded (a,a1,a2) PPSD and (b, b1, b2) PSD sample exposed to air + 5. Conclusion
NaCl + water vapor at 750 ◦ C for 10 min.
2207
T. Cui et al. Journal of Materials Research and Technology 33 (2024) 2199–2209
Fig. 11. Formation schematic diagram of the corrosion products on the surface of alloy samples during initial corrosion in the atmosphere of air + NaCl +
water vapor.
the atmospheres of air + water vapor and air + NaCl + water vapor at [12] Li M, Sun X, Li J, Zhang Z, Jin T, Guan H, Hu Z. Oxidation behavior of a single-
crystal Ni-base superalloy in air. I: at 800 and 900 C. Oxid Metals 2003;59:
750 ◦ C, respectively. The conclusions can be drawn as follows.
591–605.
[13] Liu C, Ma J, Sun X. Oxidation behavior of a single-crystal Ni-base superalloy
1. After being tested in the air + water vapor environment at 750 ◦ C for between 900 and 1000 C in air. J Alloys Compd 2010;491:522–6.
168 h, uniform and dense oxide scales comprising external NiO [14] Sato A, Chiu Y-L, Reed R. Oxidation of nickel-based single-crystal superalloys for
industrial gas turbine applications. Acta Mater 2011;59:225–40.
layers and internal Al2O3 layers are formed on both the PPSD and [15] Shinata Y. Accelerated oxidation rate of chromium induced by sodium chloride.
PSD alloy samples. Oxid Metals 1987;27:315–32.
2. Corrosion of the alloy samples is more severe after being tested in the [16] Shinata Y, Hara M, Nakagawa T. Accelerated oxidation of chromium by trace of
sodium chloride vapor. Mater Trans, JIM 1991;32:969–72.
air + NaCl + water vapor environment at 750 ◦ C for 168 h. During [17] Shinata Y, Nishi Y. NaCl-induced accelerated oxidation of chromium. Oxid Metals
the initial corrosion, NiO nodules with Cr2O3 internal layers are 1986;26:201–12.
formed on the surfaces of the γ phase. However, relatively thin NiO [18] Pettit F, Giggins C. Hot corrosion, wiley-interscience, john wiley and sons,
superalloys II–High temperature materials for aerospace and industrial power.
layers with Al2O3 inner layers are formed on surfaces of the γ′ phase 1987. p. 327–58.
where internal corrosion occurs. As the corrosion progresses, the [19] Qin J, Yang W, Wang Q, Zhou Y, Fu H, Zhang J, Liu L. Orientation control of
oxide scale of the alloy develops a bilayer structure of NiO/(Al, multiple single crystal blades using a novel high-throughput mold via seeding-
grain selection technique. J Mater Res Technol 2024;29:4845–53.
Cr)2O3, accompanied by the internal corrosion product of Al2O3. [20] Montero X, Ishida A, Meißner TM, Murakami H, Galetz MC. Effect of surface
3. Surface orientation has a significant effect on the corrosion behaviors treatment and crystal orientation on hot corrosion of a Ni-based single-crystal
of alloy samples. The PPSD alloy samples exhibit more severe superalloy. Corrosion Sci 2020;166:108472.
[21] Hu Q, Geng S, Wang J, Wang F, Sun Q, Xia S, Wu Y. Effects of solid NaCl and water
corrosion than the PSD alloy samples in both two corrosive atmo
vapour on hot corrosion behaviour of Inconel 718 superalloy and its aluminide
spheres. This difference is more obvious in NaCl-contained coating. Mater Chem Phys 2023;309:128416.
environments. [22] Liu D, Geng S, Chen G, Wang F. NiO/NiFe2O4 dual-layer coating on pre-oxidized
SUS 430 steel interconnect. Int J Hydrogen Energy 2022;47:21462–71.
[23] Wagner C. Reaktionstypen bei der Oxydation von Legierungen, Zeitschrift für
Elektrochemie. Ber Bunsen Ges Phys Chem 1959;63:772–82.
Declaration of competing interest
[24] Wu D, Liu S, Yuan Z, Cao P, Wei X, Zhang C. Influence of water vapor on the
chlorine-induced high-temperature corrosion behavior of nickel aluminide
The authors declare that they have no known competing financial coatings. Corrosion Sci 2021;190:109689.
interests or personal relationships that could have appeared to influence [25] Wu D, Wu H, Yuan Z, Wei X, Zhang C. Effect of Mo addition and water vapor on the
high-temperature corrosion performance of APS Ni-Al-based coatings. J Therm
the work reported in this paper. Spray Technol 2023;32:2460–77.
[26] Saunders SRJ, Monteiro M, Rizzo F. The oxidation behaviour of metals and alloys
at high temperatures in atmospheres containing water vapour: a review. Prog
Acknowledgement
Mater Sci 2008;53:775–837.
[27] Luo L, Su M, Yan P, Zou L, Schreiber DK, Baer DR, Zhu Z, Zhou G, Wang Y,
The authors would like to appreciate the financial support from the Bruemmer SM, Xu Z, Wang C. Atomic origins of water-vapour-promoted alloy
National Natural Science Foundation of China (Grant No. U2241251). oxidation. Nat Mater 2018;17:514–8.
[28] Zhu D, Wang X, Zhao J, Lu J, Zhou Y, Cai C, Huang J, Zhou G. Effect of water vapor
on high-temperature oxidation of NiAl alloy. Corrosion Sci 2020;177:108963.
References [29] Eliaz N, Shemesh G, Latanision R. Hot corrosion in gas turbine components. Eng
Fail Anal 2002;9:31–43.
[30] Pettit F. Hot corrosion of metals and alloys. Oxid Metals 2011;76:1–21.
[1] Han J-C. Turbine blade cooling studies at Texas A&M University: 1980-2004.
[31] Rapp RA. Hot corrosion of materials: a fluxing mechanism? Corrosion Sci 2002;44:
J Thermophys Heat Tran 2006;20:161–87.
209–21.
[2] Han J-C, Dutta S, Ekkad S. Gas turbine heat transfer and cooling technology. CRC
[32] Chen L, Lan H, Huang C, Yang B, Du L, Zhang W. Hot corrosion behavior of porous
press; 2012.
nickel-based alloys containing molybdenum in the presence of NaCl at 750◦ C. Eng
[3] Ito S, Goldstein RJ, Eckert ER. Film cooling of a gas turbine blade. 1978.
Fail Anal 2017;79:245–52.
[4] Morgan W, Morse C. An experimental investigation of hollow turbine blades for
[33] Liu Y, Sun J, Pei Z, Li W, Liu J, Gong J, Sun C. Oxidation and hot corrosion
expandable jet engines. 1954.
behavior of NiCrAlYSi+ NiAl/cBN abrasive coating. Corrosion Sci 2020;167:
[5] Caron P. High y’solvus new generation nickel-based superalloys for single crystal
108486.
turbine blade applications. Superalloys 2000;2000:737–46.
[34] Liu YD, Sun J, Pei ZL, Li W, Liu JH, Gong J, Sun C. Oxidation and hot corrosion
[6] Hashizume R, Yoshinari A, Kiyono T, Murata Y, Morinaga M. Development of Ni
behavior of NiCrAlYSi+NiAl/cBN abrasive coating. Corrosion Sci 2020;167:
based single crystal superalloys for power generation gas turbine blades. Energy
108486.
materials 2007;2:5–12.
[35] Wang Q, Zhou D, Yu M, Shi L, Li X, Sun Q. Oxidation and hot corrosion behaviors
[7] Yamagata T, Harada H, Nakazawa S, Yamazaki M, Nakagawa Y. Alloy design for
of Mo-doped NiMoAlY alloys at 750 ◦ C. Corrosion Sci 2022;201:110262.
high strength nickel–base single crystal alloys. Superalloys 1984:157–66.
[36] Yu M, Zhou D, Pu J, Cui T, Li C. Effect of Zr, Ti, Ta and Mo addition on high-
[8] Han J-C. Recent studies in turbine blade cooling. Int J Rotating Mach 2004;10:
temperature oxidation and hot corrosion behavior of NiAlY alloys. J Alloys Compd
443–57.
2022;908:164614.
[9] Moskalenko A, Kozhevnikov A. Estimation of gas turbine blades cooling efficiency.
[37] Abu-warda N, Tomás LM, López AJ, Utrilla MV. High temperature corrosion
Procedia Eng 2016;150:61–7.
behavior of Ni and Co base HVOF coatings exposed to NaCl-KCl salt mixture. Surf
[10] Kohlscheen J, Stock H. Deposition of silicon enriched nickel aluminide coatings on
Coating Technol 2021;418:127277.
internally cooled airfoils. Surf Coating Technol 2008;203:476–9.
[11] Smith A, Kempster A, Smith J. Vapour aluminide coating of internal cooling
channels, in turbine blades and vanes. Surf Coating Technol 1999;120:112–7.
2208
T. Cui et al. Journal of Materials Research and Technology 33 (2024) 2199–2209
[38] Yang Z, Zhang J, Luo C, Yu C, Li M, Han W. Effect of pre-oxidation and sea salt on [48] Luo L, Ai C, Ma Y, Li S, Pei Y, Gong S. Influence of temperature on the lattice misfit
the hot corrosion behavior of MCrAlY coatings and AlSi coatings. Surf Coating and elastic moduli of a Ni based single crystal superalloy with high volume fraction
Technol 2024;477:130354. of γ′ phase. Mater Char 2018;142:27–38.
[39] Wu Y, Liu Y, Li C, Xia X, Wu J, Li H. Coarsening behavior of γ′ precipitates in the γ’ [49] Probst-Hein M, Dlouhy A, Eggeler G. Interface dislocations in superalloy single
+γ area of a Ni3Al-based alloy. J Alloys Compd 2019;771:526–33. crystals. Acta Mater 1999;47:2497–510.
[40] Yu M, Sun Q, Wang Q, Li X, Zhou D, Pu J, Chen B, Li C. Effect of Pt-doping on the [50] Yoo J, Kwon H, Song S, Do J, Yun DW, Kim HS, Lee S-g, Kim IS, Choi B-G. Effects of
oxidation behaviors of the γ’-Ni3Al and β-NiAl phases in the NiSiAlY alloy. lattice misfit of γ/γ′ phases on hydrogen embrittlement behavior in Ni-based single
Corrosion Sci 2022;200:110224. crystal superalloy. J Mater Res Technol 2024;30:5040–55.
[41] Chen Z, Zhao J-C. Recommendation for reliable evaluation of diffusion coefficients [51] Guo HY, Tan ZH, Li YM, Zou MK, Tao Y, Wang XG, Liu JD, Liu JL, Yang YH, Li JG,
from diffusion profiles with steep concentration gradients. Materialia 2018;2:63–7. Zhou YZ, Sun XF. Diffusion of alloying elements during high temperature oxidation
[42] Qin L, Ren P, Yi Y, Deng C, Hu L, Chen D, Lu Y, Zhou S. Effect of temperature on the in a low-cost third generation Ni-based single crystal superalloy. Mater Char 2024;
oxidation behavior of Al2O3 reinforced CoCrAlYTa coating by laser-induction 211:113913.
hybrid cladding. Surf Coating Technol 2023;473:130038. [52] Zhao Z, Wang J, Chen M, Zhang J, Wang F, Young DJ. Comparative study on the
[43] Zhou S, Xiong Z, Lei J, Dai X, Zhang T, Wang C. Influence of milling time on the initial oxidation behavior of conventional and nanocrystalline MCrAlY coatings-
microstructure evolution and oxidation behavior of NiCrAlY coatings by laser effect of microstructure evolution and dynamic mechanisms. Acta Mater 2022;239:
induction hybrid cladding. Corrosion Sci 2016;103:105–16. 118264.
[44] Gao D, Shen Z, Chen K, Zhou X, Liu H, Wang J, Li Y, Liu Z, Deng H, Wang WY. [53] Chen Y, Zhao X, Xiao P. Effect of microstructure on early oxidation of MCrAlY
Review of progress in calculation and simulation of high-temperature oxidation. coatings. Acta Mater 2018;159:150–62.
Prog Mater Sci 2024:101348. [54] He R, Li M, Han X, Feng W, Zhang H, Xie H, Liu Z. Experimental study of as-cast
[45] Qin L, Ren P, Yi Y, Xie T, Hu Y, Chen D, Zhou S. Effect of Al2O3 content on the and heat-treated single-crystal Ni-based superalloy interface using TEM.
high-temperature oxidation behaviour of CoCrAlYTa coatings produced by laser- Nanomaterials 2023;13:608.
induction hybrid cladding. Corrosion Sci 2022;209:110739. [55] Zhang H, Wen H, Peng R, He R, Li M, Feng W, Zhao Y, Liu Z. Experimental study at
[46] Wagner C. Theoretical analysis of the diffusion processes determining the oxidation the phase interface of a single-crystal Ni-based superalloy using TEM. Materials
rate of alloys. J Electrochem Soc 1952;99:369. 2022;15:6915.
[47] Wei Y, Zheng S, Pu J, Zhou D, Wang L, Guo W, He G. Effect of aluminum [56] Poletaev GM, Zorya IV, Starostenkov MD, Rakitin RY, Tabakov PY. Molecular
concentration on microstructure and evolution behavior of oxide layer on NiAlSiY dynamics simulation of the migration of tilt grain boundaries in Ni and Ni3Al.
coating at 500◦ C. Corrosion Sci 2020;165:108400. J Exp Theor Phys 2019;128:88–93.
2209