You are on page 1of 13

Corrosion Science 209 (2022) 110803

Contents lists available at ScienceDirect

Corrosion Science
journal homepage: www.elsevier.com/locate/corsci

Grain boundary corrosion mechanism of YSZ thermal barrier oxides under


CMAS attack
Di Wu a, b, Xiao Shan a, Huangyue Cai a, Fan Yang c, *, Xiaofeng Zhao a, *
a
Shanghai Key Laboratory of Advanced High-Temperature Materials and Precision Forming, Shanghai Jiao Tong University, Shanghai 200240, China
b
Aero Engine Corporation of China Commercial Aircraft Engine Co. Ltd, Shanghai 200241, China
c
School of Mechanical Engineering, Shanghai Jiao Tong University, Shanghai 200240, China

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

Keywords: Grain boundary (GB) corrosion of yttria-stabilized zirconia (YSZ) under calcium-magnesium-alumino-silicates
Calcium-magnesium-alumino-silicates (CMAS) (CMAS) attack was investigated. Dense YSZ ceramics with YO1.5 contents varying from 8 to 57 mol.% were
Thermal barrier coatings (TBCs) prepared, and exposed to CMAS at 1300 ºC. The results show preferential GB corrosion occurs for all YSZ ce­
Yttria-stabilized zirconia (YSZ)
ramics, and the extent of corrosion significantly depends on the YO1.5 concentration. For low YO1.5 YSZ (e.g.
Molten silicate corrosion
Grain boundary (GB)
8YSZ and 20YSZ), GB is rapidly dissolved and infiltrated by CMAS melt. In contrast, for YSZ with high YO1.5
contents (e.g. Y2Zr2O7 and Y4Zr3O12), a uniform corrosion through the GB and grains is observed. The reason of
the preferential GB corrosion is attributed to segregation of Y element in YSZ, which promotes the CMAS
reactivity with the GBs. Furthermore, the mechanism of extent difference of GB corrosion in YSZ materials is
discussed. It is proposed that the GB corrosion of YSZ can be mitigated by decreasing the Y segregation,
increasing the relative volume of reaction products, and/or promoting the precipitation kinetic of products.

1. Introduction accepted [8,10]. One is thermal-mechanical degradation. During oper­


ation, when the temperature exceeds its melting point (1200–1250 ºC),
Thermal barrier coating (TBC) is widely applied on the hot sections the CMAS deposit will melt and then fill the open pores of TBC driven by
of a gas turbine engine to insulate the thermal conduction from high- the capillary force [17]. The melt solidifies and remains in the coating
temperature gas and to prolong the lifetime of key components [1–4]. during cooling. This melting-infiltration-solidification behavior of
The state-of-the-art TBC is 7–8 wt% Y2O3 stabilized ZrO2 (8YSZ), which CMAS deposit significantly increases the elastic modulus of the coating,
is usually deposited on metallic substrates by electron beam physical and thereby increases the thermal mismatch stress and degrades the
vapor deposition (EB-PVD) or air plasma spraying (APS) [5–7]. With the strain tolerance of TBC [13,14,18,19]. The second mechanism is
ever-increasing demand of higher engine-operating temperature, the thermal-chemical reaction. At high temperature, the CMAS melt can
calcium-magnesium-alumino-silicates (CMAS) corrosion of TBC has dissolve 8YSZ rapidly to form Y-lean ZrO2, which destroys the micro­
become a more severe issue and needs to be addressed urgently [8–11]. structure integrity of TBC [20,21]. During cooling, the reprecipitated
CMAS from the volcanic ash, fine sand and dust may enter the engines Y-lean ZrO2 can transform to the brittle monoclinic phase with ~5%
with the inlet air and deposit on the surface of TBC [10]. During engine volume expansion, which increases the internal stress of TBC and ex­
operation, the CMAS deposit will melt, infiltrate and react with the TBC, acerbates failure of the coating [22,23].
and thus degrades the thermomechanical properties of TBC and causes However, previous CMAS corrosion mechanism does not consider
the premature failure[11–15]. Tremendous studies have shown that, the the difference between grain boundary (GB) and intra-grain. The reason
state-of-the-art 8YSZ TBC has poor resistance to CMAS corrosion, for may be attributed to that the EB-PVD TBC, which is comprised of
example, the classic APS 8YSZ TBC with a thickness of 250 µm was separated columnar single crystal, has no GB [24]. In contrast, the APS
corroded thoroughly and spalled from the substrate after exposure to TBC, comprised of polycrystal splat, contains plenty of GBs. The recently
CMAS at 1250 ºC for 4 h [16]. developed nano-crystalline TBC, prepared by suspension plasma spray­
Failure of the 8YSZ TBC caused by CMAS attack has been extensively ing or plasma spraying physical vapor deposition technology, has a
investigated, and two mechanisms have been elucidated and well- higher GB density [24]. Previous studies have preliminarily pointed out

* Corresponding authors.
E-mail addresses: fanyang_0123@sjtu.edu.cn (F. Yang), xiaofengzhao@sjtu.edu.cn (X. Zhao).

https://doi.org/10.1016/j.corsci.2022.110803
Received 21 July 2022; Received in revised form 13 October 2022; Accepted 30 October 2022
Available online 1 November 2022
0010-938X/© 2022 Elsevier Ltd. All rights reserved.
D. Wu et al. Corrosion Science 209 (2022) 110803

that APS 8YSZ TBC is subject to preferential GB corrosion under CMAS APS using commercial 8YSZ powder (Metco 204B NS, Oerlikon Metco,
attack [16,25–29], although the systematic experiment evidence and Westbury, NY). The bond coat was prepared by high velocity air fuel
analysis is lacking. Additionally, the nanocrystalline 8YSZ TBC exhibits technology using commercial NiCoCrAlY powders (AMDRY 365–2,
more severe CMAS corrosion than the traditional APS TBC comprised of Oerlikon Metco).
micro-size grains [30]. These results indicate that preferential GB 20YSZ, YZO and Y4Zr3O12 ceramics were prepared by the conven­
corrosion may exacerbate the CMAS attack and cause the terrible CMAS tional solid-state reaction method using ZrO2 (>99.99%, Macklin,
resistance of 8YSZ, however, the underlying mechanism is not clear. Shanghai, China), Y2O3 (>99.99%, Macklin, Shanghai, China) powders
On the other hand, previous studies show ZrO2 doped with high as starting materials. Appropriate amounts of ZrO2 and Y2O3 powders
levels of REO1.5 (RE, rare earth) (e.g. Y2Zr2O7) display a uniform CMAS were weighed according to their chemical formula and then mixed
corrosion through GB and intra-grain [31,32] rather than the severe GB thoroughly by planet ball-milling. The resultant mixture was dried,
infiltration as observed in 8YSZ. These materials have much better sieved and compacted into pellets by uniaxial pressing (150 MPa). The
CMAS resistance than 8YSZ [32–34], suggesting the CMAS resistance of green bodies were sintered in a box furnace at 1500 ºC for 20 h to obtain
TBC materials can be enhanced by inhibiting the preferential GB dense pellets. Both heating and cooling rates were 10 ºC/min.
corrosion, and the CMAS corrosion mode (i.e. preferential GB corrosion 8YSZ pellets were prepared by spark plasm sintering (SPS) using
or uniform corrosion) of YSZ materials may depend on the YO1.5 doping commercial 8YSZ powder (<1 µm, Sigma-Aldrich). The rapid sintering
content. process and high cooling rate of SPS technology can ensure dense 8YSZ
Based on the above considerations, in this work, the CMAS corrosion ceramics with a pure t′ phase [36]. During SPS, green bodies of 8YSZ
behavior of APS 8YSZ coating was firstly investigated to demonstrate the were exposed at 1500 ºC for 5 min under a uniaxial pressure of 50 MPa
GB corrosion issue. And then, the CMAS corrosion behaviors of YSZ with a heating/cooling rate of 100 ºC/min.
ceramics with different YO1.5 contents were investigated with a focus on In addition, a 20YSZ single crystal with (100) orientation (Hefei
the GB corrosion mechanism. Several compositions (i.e. Zr0.92Y0.08O1.96 Kejing Materials Technology Corp, Hefei, China) was also used for
(8YSZ), Zr0.8Y0.2O1.9 (20YSZ), Y2Zr2O7 (YZO) and Y4Zr3O12) were comparison with ceramics.
selected from the ZrO2-YO1.5 binary phase diagram (Fig. 1). Results
show that all YSZ ceramics exist the preferential GB corrosion, and the 2.2. CMAS corrosion test
level of GB corrosion is significantly influenced by the YO1.5 doping
content. The mechanism of preferential GB corrosion in YSZ materials is Based on the average composition of CMAS deposits on TBC of
discussed, and the mitigation strategies are proposed. turboshaft shrouds operated in a desert environment[37], a model
CMAS with a typical composition of 33CaO-9MgO-13AlO1.5-45SiO2
2. Experiment (Ca33Mg9Al13Si45, molar ratio) was prepared and used in this study. The
corresponding oxides were weighed and ball-milled in isopropyl alcohol
2.1. Materials to form a slurry, which was uniformly applied to the clean surfaces of the
YSZ to obtain the CMAS-coated specimens after drying. The CMAS area
The coating specimens used are standard APS 8YSZ TBC, which density was fixed at 20 ± 1 mg/cm2 for all specimens. Subsequently, the
consist of an 8YSZ top coat (~16% porosity), a NiCoCrAlY bond coat, CMAS coated YSZ specimens were heat-treated in a box furnace. YSZ
and a Hastealloy-X superalloy substrate. The top coat was prepared by ceramics were heated to 1300 ºC with a ramping rate of 10 ºC/min, and
exposed for 1, 4, 8, 24, and 50 h. 8YSZ TBC was exposed at 1250 ºC for
only 2 h to avoid damaging the metallic bond coat and substrate. After
high-temperature exposure, the specimens were air-quenched to room
temperature withdrawl from the furnace into air (< 3 min to reach the
room temperature) to maintain the high-temperature composition and
morphology of the CMAS corrosion zone.

2.3. Characterization

The phase constitution and density of the as-prepared YSZ ceramics


were determined by X-ray diffraction (XRD; Ultima IV, Rigaku, Japan)
and image analysis, respectively. The CMAS-corroded specimens were
mounted in epoxy, cross-sectioned using a low-speed diamond saw,
ground and polished to a 0.5 µm diamond finish. Microstructure and
phase constitution of the CMAS reaction zones were characterized by a
combination of scanning electron microscopy (SEM; Mira3, Tescan,
Czech Republic), transmission electron microscopy (TEM; Talos F200X,
FEI, USA) and Raman microscopy (HR evolution, Horiba, 532 nm Argon
laser). Compositional analyses were performed by energy dispersive
spectroscopy (EDS; Aztec X-MaxN80, Oxford Instruments, UK) under
SEM and TEM. TEM specimens were extracted from specific locations of
the polished cross-sections by focused ion beam (FIB, Versa 3D, FEI,
USA).

3. Results

3.1. CMAS corrosion behavior of APS 8YSZ coating

Fig. 1. ZrO2-YO1.5 binary phase diagram (replot after Ref. [35]). Compositions Fig. 2a presents the cross-sectional BSE image of the CMAS-corroded
of interest are indicated by red dots, from left to right, corresponding to 8YSZ, 8YSZ coating. The corresponding EDS Ca mapping (Fig. 2b) shows the
20YSZ, Y2Zr2O7 and Y4Zr3O12, respectively. ~400 µm-thick coating has been infiltrated thoroughly by CMAS melt

2
D. Wu et al. Corrosion Science 209 (2022) 110803

Fig. 2. (a) Cross-sectional BSE images of 8YSZ TBC corroded by CMAS at 1250 ºC for 2 h. (b) EDS Ca mapping of (a). (c-e) High-magnification BSE images of Area1,
Area2 and Area3 in (a), respectively.

after exposure at 1250 ºC for 2 h with plenty CMAS remnants on the the features of reprecipitated reaction products (Y-lean ZrO2 [20,29]), as
coating surface. Expanded views of three selected areas, as indicated by shown in Fig. 2c. The middle region of the coating (Area2) is modestly
Area1, Area2, and Area3 in Fig. 2a, are presented in Fig. 2c-e. Near the corroded. The initial coating structure (e.g. the sprayed splats) is
top surface (Area1), the coating is severely corroded. The original partially remained, and fine grains from the precipitated reaction
coating microstructure has been destroyed completely and replaced by products can be observed (Fig. 2d). The bottom region of the coating (i.

Fig. 3. (a) Bright field (BF) TEM image of the splat in CMAS infiltration front corresponding to line A in Fig. 2e. (b) and (c) HAADF image and the corresponding Ca
mapping of the dashed rectangle area in (a). The table at the bottom-right lists the chemical compositions of points A, B and C in (b).

3
D. Wu et al. Corrosion Science 209 (2022) 110803

e. the CMAS infiltration front, Area3) is slightly corroded to preserve the initial specimen surface (red dash line) to the CMAS infiltration front
initial coating structure without any notable precipitation of the reac­ (black dash line). Evolution of the CMAS corrosion depth with exposure
tion products. This changes of coating structure under CMAS attack are time is plotted in Fig. 6. The sintered 8YSZ and 20YSZ ceramics show
also reported elsewhere [38,39]. Furthermore, it is worth mentioning similar corrosion behaviors that the corrosion depth increases almost
that the GBs of the 8YSZ coating, as indicated by the black arrows in linearly with exposure time showing a corrosion rate of ~15 µm/h. After
Fig. 2e, show different image contrast with the coating, suggesting they 50 h exposure, the corrosion depth is ~ 750 µm. Y2Zr2O7 and Y4Zr3O12
may have been corroded by CMAS. To confirm the GB corrosion show almost identical corrosion behavior, which shows a rapid increase
behavior, the CMAS infiltration front is further characterized by TEM. with exposure time in the initial stage (0–8 h) with a corrosion rate of
Fig. 3a presents the bright field TEM image of the splat in the CMAS 13 µm/h, followed by a slow increase to the maximum depth (~130 µm)
infiltration front. The corresponding HAADF image (Fig. 3b) and EDS Ca after 24 h exposure. Comparing with the sintered ceramics, the 20YSZ
mapping (Fig. 3c) confirm presence of CMAS melt in the GBs of the splat. single crystal (20YSZ-SC) shows the smallest corrosion depth with an
Grains in the splat still keep the columnar morphology and their initial corrosion rate of 4 µm/h at the initial stage (0–8 h), followed by a
chemical compositions are close to the as-sprayed 8YSZ coating slow increase to ~70 µm after 50 h exposure. The above results suggest:
(Table in Fig. 3c). The above results indicate that in the reaction front, (1) YSZ ceramics with high YO1.5 contents (i.e., Y2Zr2O7 and Y4Zr3O12)
GBs of the 8YSZ coating are dissolved and infiltrated by CMAS melt, have better CMAS corrosion resistance than those with low YO1.5 con­
whereas the grains are not corroded. tents (i.e., 8YSZ and 20 YSZ); (2) single-crystal 20 YSZ has a significantly
Based on the above observations, the CMAS corrosion behavior of the enhanced corrosion resistance compared with the sintered 20YSZ. To
APS 8YSZ coating can be summarized as the following procedure. unveil the reasons behind the observations, further characterizations of
Firstly, the CMAS melt fills the open pores of the coating and infiltrates the reaction layer is performed and presented below.
along the GBs of 8YSZ. Subsequently, the CMAS melt gradually dissolves
the 8YSZ grains and precipitates the reaction products (Y-lean ZrO2), 3.2.3. Morphology of the reaction layer
which destroys the coating structure and changes the phase composi­ Fig. 7 presents the high-magnification BSE images of the reaction
tion. The above results demonstrate that severe GB corrosion exists in layer in the YSZ specimens. The most striking feature is that 8YSZ and
8YSZ, which significantly reduces the CMAS resistance of the coating. 20YSZ ceramics exhibit apparent GB corrosion, as indicated by the black
arrows in Fig. 7a-d. In contrast, Y2Zr2O7 and Y4Zr3O12 show a uniform
3.2. CMAS corrosion behavior of YSZ ceramics with different YO1.5 corrosion through GB and grain, and a dense reaction product layer is
contents formed near the reaction front, as indicated by the double-head arrows
in Fig. 7g and i. In addition, 20YSZ-SC also shows a uniform corrosion
3.2.1. Characterization of pristine materials behavior, as shown in Fig. 7e and f. Comparison between the two 20YSZ
Fig. 4a presents the XRD patterns of the as-sintered YSZ ceramics. As samples (ceramic and single-crystal) indicates the critical role of GBs on
expected, the SPS 8YSZ presents a pure t′ phase, and the other samples the CMAS corrosion behavior of YSZ with low YO1.5 doping levels. The
(20YSZ, Y2Zr2O7 and Y4Zr3O12) display a pure cubic phase. SEM images above results also highlight the importance of YO1.5 content. To further
(Figs. 4b-4e) reveal a dense microstructure (relative density of ~98%) of understand the GB corrosion mechanism, 20YSZ and Y2Zr2O7 ceramics
these ceramics. The average grain sizes are 2.6, 4.0, 1.1 and 1.6 µm for are selected for in-depth investigation in Sections 3.2.4 and 3.2.5. The
8YSZ, 20YSZ, Y2Zr2O7 and Y4Zr3O12, respectively. The difference of reasons for selecting these two compositions are 1) 20YSZ and Y2Zr2O7
grain size is assumed to have a negligible effect on the CMAS corrosion represent two distinct CMAS corrosion behaviors (i.e. preferential GB
behavior of YSZ ceramics. corrosion versus uniform corrosion through GB and grain); and 2) they
both present cubic structure to minimize the possible impact from
3.2.2. Overall CMAS corrosion behavior different crystal structures.
Fig. 5 presents the low-magnification, cross-sectional BSE images of
the sintered YSZ ceramics and the 20YSZ single-crystal after reacting 3.2.4. Identification of the reaction layer in 20YSZ ceramic and single
with CMAS at 1300 ºC for various holding times, from which the CMAS crystal
corrosion depth can be obtained by measuring the distance from the As shown in Figs. 5b-b3 and c-c3, both ceramic and single-crystal

Fig. 4. (a) XRD patterns of the as-sintered YSZ ceramics with different YO1.5 contents. (b-e) SEM images of the thermally-etched surfaces of 8YSZ, 20YSZ, Y2Zr2O7
and Y4Zr3O12 ceramics, respectively.

4
D. Wu et al. Corrosion Science 209 (2022) 110803

Fig. 5. Low-magnification cross-sectional BSE images of the CMAS-corroded YSZ bulk specimens: (a-a3) 8YSZ, (b-b3) 20YSZ polycrystal, (c-c3) 20YSZ single crystal,
(d-d3) Y2Zr2O7 and (e-e3) Y4Zr3O12. The CMAS corrosion depth was measured from the initial specimen surface (red dash line) to the infiltration front (black
dash line).

20YSZ show double-layer reaction zone after exposure to CMAS. The layers for 20YSZ and 20YSZ-SC are presented in Fig. 8a-b and c-d,
upper reaction layer thickens initially (0–8 h), and then gradually re­ respectively. In the upper reaction layer, precipitations of the reaction
duces in thickness with further exposure and finally disappears after product (labeled as product #1) in both samples present irregular
50 h corrosion. The lower reaction layer, on the contrary, keeps thick­ morphology. In the lower reaction layer, 20YSZ and 20YSZ-SC show
ening during CMAS corrosion. The above observations indicate the different features. For 20YSZ, CMAS melt dissolves the GB region of the
upper reaction layer is temporary and metastable, whereas the lower lower layer, as indicated by EDS Ca mapping (Fig. 8e). Furthermore, EDS
layer is the main reaction zone. line profile of Ca across a grain (L1 in Fig. 8c) reveals a core-shell
High-magnification BSE images of the upper and the lower reaction structure (Fig. 8g). The Ca-rich shell is labelled as product #2 for later

5
D. Wu et al. Corrosion Science 209 (2022) 110803

Fig. 6. (a) The evolution of CMAS corrosion depth in all YSZ bulks with exposure time. (b) Expanded view of the relationship between corrosion depth and time in
20YSZ-SC, Y2Zr2O7 and Y4Zr3O12 samples.

identification. For 20YSZ-SC, the lower reaction layer presents a indicate that the YZO also exists the preferential GB corrosion. However,
columnar structure (Fig. 8d). The gap between two columns is rich in Ca, the extent of GB corrosion is much lower compared with 20YSZ ceramic,
as suggested by EDS Ca mapping (Fig. 8f). Line scan along L2 in Fig. 8d and GB corrosion depth is less than 1 µm. High resolution TEM analysis
shows a uniform distribution of Ca inside the columnar “grain” (Fig. 8h), shows a crystallized apatite is precipitated in the GB region of YZO. Note
indicating the “grain” is a reaction product between 20YSZ and CMAS. that the GB region is less than 20 nm (see Fig. 11b), but the products
To identify the reaction products, Raman spectra analysis was carried (apatite) can still precipitate in GB, which indicates that the apatite has
out. As shown in Fig. 9, Raman spectra for product #1 in both samples much high precipitation rate. Therefore, it is preliminarily proposed that
show typical features for tetragonal ZrO2, whereas those for product #2 the rapid precipitation of apatite seals the GB and therefore inhibits the
present typical cubic ZrO2 structure. EDS analysis indicates both product GB corrosion. In summary, the YZO also exists the preferential GB
#1 and #2 are CaO/YO1.5 doped ZrO2 with the average composition of corrosion, but the GB corrosion extent is much limited, and the reason is
Zr0.917Y0.041Ca0.042O1.938 and Zr0.818Y0.142Ca0.04O1.889, respectively preliminarily attributed to the rapid precipitation of apatite.
(Table 1).
Based on the above results, the CMAS corrosion behavior of 20YSZ 4. Discussion
ceramic can be summarized as follows. At the initial stage (0–8 h),
CMAS melt dissolves 20YSZ and precipitates the tetragonal products The major finding of this work is YSZ ceramics present preferential
(#1); subsequently, CMAS mainly dissolves GB and precipitates the GB CMAS corrosion, and the YO1.5 content has a dramatic impact on the
cubic products (#2) on the 20YSZ grain to form a core-shell structure. GB corrosion behavior of YSZ. In this section, mechanism for the pref­
In essence, CMAS corrosion is the chemical reaction between CMAS erential GB corrosion in YSZ, and the reasons for its dependence on the
melt and specimen. Besides identifying reaction products, it is also YO1.5 content, are discussed. Based on the findings and discussions,
necessary to understand the compositional evolution of the CMAS melt some strategies for mitigating GB corrosion are proposed.
(the dark top area in Figs. 5b-b3). Here we focus on CaO, YO1.5 and ZrO2
in CMAS as the reaction products are comprised of these oxides 4.1. Preferential GB corrosion mechanism of YSZ materials
(Table 1). As shown in Fig. 10, the CaO content in CMAS decreases
continuously with increasing exposure time, which is attributed to the It is generally accepted that GB of YSZ has more defects than intra-
consumption of CaO to form the reaction products. ZrO2 content rapidly grain, thereby can be easily corroded by CMAS melt. However, this
reaches the saturated value (~4 mol%) within 1 h and then remains cannot explain the different CMAS corrosion behaviors between 20YSZ
stable, which indicates that the reaction products containing ZrO2 will and Y2Zr2O7. Previous studies indicate that YSZ materials, whether
be precipitated. Additionally, YO1.5 content keeps increasing during tetragonal or cubic phase, exhibit GB segregation of Y element [40–44].
corrosion, which indicates that YO1.5 is not saturated in melt and For examples a tetragonal YSZ with 2 mol% Y2O3 doping has a Y2O3
therefore CMAS melt tends to dissolve more YO1.5. content of ~5 mol% in GB[42]. For a cubic Zr0.8Y0.2O1.9 (20YSZ), the GB
composition is quantified as Zr0.67±0.06Y0.33±0.06O1.11±0.25[44]. GB
3.2.5. Identification of the reaction layer in YZO segregation of Y is also observed in YZO, as demonstrated in Fig. 13.
For bulk YZO, the BSE image (Fig. 7g) shows YZO is uniformly Compositions of intra-grain and GB determined by EDS are Zr0.5±0.03Y0.5
corroded by CMAS through GB and grain. To unveil the uniform ±0.02O1.75±0.03 and Zr0.4±0.03Y0.6±0.02O1.49±0.03, respectively. The reason
corrosion mechanism, the reaction layer of YZO is characterized by of GB segregation of Y is related to the uneven distribution of oxygen
TEM, as shown in Fig. 11. Two types of reaction products can be iden­ occupy between GB and intra-grain [40,44]. First-principles calculation
tified at the corrosion front, as shown by the HAADF image and the indicates that Y segregation in GB is energetically favorable, which can
corresponding EDS Ca/Zr mapping in Fig. 11b. Selected area diffraction mitigate the lattice distortion and reduce the internal energy of YSZ[40].
patterns indicate the products have apatite (Ap) and cubic ZrO2 (C-ZrO2) Here, it is proposed that the GB segregation of Y causes the prefer­
structure, respectively (Fig. 11d and e). EDS analysis suggests the ential GB CMAS corrosion of YSZ. The reasons are demonstrated below.
average compositions for Ap and C-ZrO2 are Zr0.654Y0.323Ca0.023O1.816 The composition analysis of reaction products (Tables 1 and 2) indicates
and Ca0.157Y0.442Si0.378Zr0.023O1.622, respectively (Table 2). The above that the reaction between YSZ and CMAS mainly involves the CaO, SiO2,
composition and phase results of reaction products are consistent with ZrO2 and YO1.5. CMAS composition evolution (Fig. 10b) shows that the
previous reports [32,33]. CaO and SiO2 amount are sufficient during reaction, ZrO2 content
Furthermore, the GB region of YZO in the corrosion front is also rapidly reaches the saturated state within 1 h exposure, and only YO1.5
characterized. The HAADF and EDS map analyses (Fig. 12a and b) content in CMAS keeps increasing with time. This suggests that YO1.5 is

6
D. Wu et al. Corrosion Science 209 (2022) 110803

Fig. 7. High-magnification BSE images of the reaction layer in various YSZ samples after CMAS corrosion at 1300 ºC for 8 h (left column) and 50 h (right column).

7
D. Wu et al. Corrosion Science 209 (2022) 110803

Fig. 8. High-magnification BSE images of (a, b) upper and (c, d) lower reaction layer in 20YSZ pellets following 8 h and 50 h CMAS corrosion, respectively. (e, f) The
corresponding EDS Ca elemental maps of (c) and (d), respectively. (g, h) EDS line profile of Ca element in location L1 and L2 of (c) and (d), respectively.

not saturated and CMAS melt tends to dissolve more YO1.5. Therefore, it electrons, thereby causing a high OB of oxides (e.g. CaO)[49]. In
is speculated that the higher YO1.5 content in YSZ, the more severe contrast, the cations with high polarizability will draw electrons away
CMAS corrosion. Y segregation intrinsically causes the increasing of from oxygen, and the oxygen is therefore less able to donate electrons,
YO1.5 content in GB of YSZ, therefore GB region is easier to be corroded. which causes a low OB of oxides (e.g. SiO2). OB theory indicates that the
In order to verify this speculation, further discussions based on optical larger OB difference between reactants, the higher reactivity between
basicity (OB) consideration and experimental results are performed. both[50]. The OB value can be obtained by calculation or experimental
The OB concept has been widely used to classify the high- measurement as described elsewhere[49–52]. Here, the OB(Λ) and the
temperature reactivity of oxides (e.g. metallurgical slags and lining re­ relative OB(ΔΛ) of CMAS and YSZ are calculated and summarized in
fractory) in melt glass[45–48]. Specifically, OB is based on lewis Table 3. The results show that both OB of YSZ and the relative OB of YSZ
acid-base theory and represents the ability of oxygen anion in oxides to and CMAS increase with an increase of YO1.5 content in YSZ. Therefore,
donate electrons for cations[49]. OB of oxides depends on the cations in view of the OB theory, the CMAS reactivity of YSZ will be promoted
type, the lowly polarizing cations allow oxygen anion to donate more with increasing the YO1.5 doping content.

8
D. Wu et al. Corrosion Science 209 (2022) 110803

mitigates the CMAS infiltration and makes the corrosion become diffu­
sion control gradually. In summary, the CMAS reactivity of various YSZ
can be evaluated by the corrosion rate in initial stage (0–8 h). Here, the
20YSZ-SC, Y2Zr2O7 and Y4Zr3O12 are selected to compare the initial
corrosion rate. It is because those samples all show uniform corrosion
and therefore the measured corrosion rate can represent the intrinsic
corrosion behavior of materials. However, the polycrystalline 8YSZ and
20YSZ show severe GB corrosion, which suggests that the measured
corrosion rates mainly represent the GB corrosion rather than the
intrinsic corrosion behavior of materials. The results indicate that in
initial stage (0–8 h), the corrosion rate of 20YSZ single crystal is ~
4 µm•h− 1 (Fig. 6b) which is much less than that in Y2Zr2O7 and
Y4Zr3O12 (13 µm•h− 1). Therefore, in view of the corrosion rate, it also
can be proposed that the higher YO1.5 content in YSZ, the higher reac­
tivity of CMAS and YSZ. In summary, both OB consideration and
experiment result confirm that the preferential GB corrosion of YSZ can
be attributed to the YO1.5 segregation in GB region.

Fig. 9. Raman spectra of the CMAS reaction products in 20YSZ. Product #1 4.2. The extent difference ofGB corrosion in YSZ materials
and #2 correspond to the reaction products in the upper and lower reaction
layer, respectively. Spectra for 20YSZ ceramic and single crystal are also
As described, all YSZ materials exist preferential GB corrosion, and
included for comparison.
the reason is attributed to the Y segregation in GB. However, the GB
corrosion extent has a significant difference in various YSZ. The reason is
Table 1 discussed below. Firstly, all YSZ exist GB segregation of Y, but the extent
Composition and crystal structure of the reaction products between 20YSZ and of Y segregation is different for various YSZ. For example, the GB
CMAS. Data for 20YSZ are included for comparison. composition in 4YSZ, 20YSZ and Y2Zr2O7 is qualified as Zr0.9Y0.1O1.95,
Name Chemical composition (mol. %) Crystal structure Zr0.67Y0.33O1.11 and Zr0.4Y0.6O1.5, respectively, as summarized in
Table 4. Therefore, the GB segregation extent of Y element can be
CaO YO1.5 ZrO2
calculated as 150%, 65% and 20%, respectively. This indicates that the
Product #1 4.2 ± 0.7 4.1 ± 0.7 91.7 ± 1.3 tetragonal Y segregation extent gradually decrease with the increasing of YO1.5
Product #2 4.0 ± 0.1 14.2 ± 0.5 81.8 ± 0.6 cubic
content in YSZ. Additionally, the first-principles calculation indicates
20YSZ 0 20.0 ± 0.7 80.0 ± 0.7 cubic
that the driving force of Y segregation will decrease with the increasing
of YO1.5 content in YSZ. For example, the driving force of Y atom
On the other hand, the reactivity of YSZ and CMAS can be evaluated segregation in GB is calculated as − 5.08 ± 0.88 eV, − 2.88 ± 0.88 eV,
by CMAS corrosion rate shown in Fig. 6. CMAS corrosion contains two − 1.59 ± 0.10 eV in 6YSZ, 24YSZ and 60YSZ, respectively[40]. This
steps: (1) chemical reaction at the interface, and (2) diffusion or trans­ also indicates that the segregation extent of Y will decrease with the
port of reacting species toward interface. The relative slow process is the increasing of YO1.5 content in YSZ. The decreasing of Y segregation
rate determining step, which controls the corrosion rate. Therefore, the extent explains the lower GB corrosion extent of YSZ with higher YO1.5
corrosion rate obtained from the experiment results can be used to content.
reflect the chemical reaction rate, i.e. reactivity. Specifically, at the Additionally, it is proposed that volume change of crystalline ma­
initial stage (0–8 h), the CMAS is sufficient, and the reaction layer is terials during CMAS corrosion also has an important effect on the GB
relative thin, therefore the corrosion is mainly controlled by chemical corrosion extent. Here, the volume of dissolved YSZ and precipitated
reaction. With further exposure, the CMAS amount obviously decreases reaction products is defined as VD and VP, respectively. If VP/VD > 1, the
as well as the reaction layer thickens and becomes dense (Fig. 5), which dense reaction product layer will tend to form, which can mitigate the
CMAS infiltration effectively. The CMAS corrosion process of 20YSZ and

Fig. 10. (a) Compositional evolution of the CMAS melt with corrosion time in 20YSZ. (b) Expanded view of the data for ZrO2 and YO1.5 in (a).

9
D. Wu et al. Corrosion Science 209 (2022) 110803

Fig. 11. (a) Bright field TEM image of the CMAS infiltration front in Y2Zr2O7 after 8 h corrosion. (b) HAADF image and EDS Ca/Zr EDS maps of the CMAS infiltration
front corresponding to the red rectangle in (a). (c-e) Selected area diffraction patterns of YZO, C and Ap, respectively.

Y2Zr2O7 is qualifiedly described by the following equations. For 20YSZ,


Table 2
two reaction products (product #1 and #2) are subsequently precipi­
Composition and phase of the reaction products between YZO and CMAS. Data
tated during corrosion, as demonstrated in Eqs. (1) and (2). Balancing
for YZO are included for comparison.
those equations is based on the ZrO2 in 20YSZ are fully precipitated in
Name Chemical composition (mol. %) Crystal
the reaction products. It is because the ZrO2 content in CMAS melt
structure
CaO YO1.5 ZrO2 SiO2 rapidly reaches the saturated state within 1 h corrosion and then keep no
Apatite 15.7 44.2 2.3 ± 0.2 37.8 apatite changes shown in Fig. 10. For Y2Zr2O7, apatite and C-ZrO2 are co-
± 0.9 ± 1.7 ± 2.0 precipitated during reaction, as described by Eq. (3). Balancing Eq. (3)
C-ZrO2 2.3 ± 0.2 32.3 65.4 0 cubic is based on both YO1.5 and ZrO2 in Y2Zr2O7 are fully precipitated in the
± 1.1 ± 1.4
reaction products, which is also described elsewhere[53].
YZO 0 49.8 50.2 0 cubic
± 0.1 ± 0.1 Y0.2 Zr0.8 O1.9 + 0.04CaO(CMAS)→0.894Zr0.917 Y0.041 Ca0.042 O1.938 (#1)
+ 0.146YO1.5 (CMAS) (1)

Fig. 12. (a) High-magnification HAADF image of the CMAS infiltration front corresponding to the blue rectangle in Fig. 11a. (b, c) HAADF and high-resolution TEM
images of the YZO grain boundary in the CMAS infiltration front, respectively. The color figures below are the corresponding EDS element maps of Zr, Ca, Si and Y.

10
D. Wu et al. Corrosion Science 209 (2022) 110803

Fig. 13. (a) TEM BF image of the grain boundary area in bulk Y2Zr2O7. (b, c) The corresponding Zr and Y EDS maps of (a), respectively. (d) EDS line profile of Zr and
Y along Line A in (a).

Finally, the precipitation kinetics of reaction product also affects the


Table 3
GB corrosion extent. As described, Y2Zr2O7 shows slight GB corrosion in
Calculated optical basicities (Λ) of YSZ and CMAS. The difference between
contrast to the severe GB corrosion in 20YSZ. An important reason is that
YSZ and CMAS (ΔΛ) is also listed.
the reaction products (i.e. apatite) are rapidly precipitated in the
Composition Λ ΔΛ Y2Zr2O7 GB of corrosion front, which seals the GB and inhibit the further
8YSZ 0.86 0.23 CMAS corrosion, as shown in Fig. 12. For 20YSZ, the GB region is rapidly
20YSZ 0.88 0.25 dissolved, and then the reaction products with cubic structure are slowly
Y2Zr2O7 0.99 0.33
precipitated on the residual 20YSZ grain to form a core-shell structure
Y4Zr3O12 0.99 0.33
CMAS 0.63 / (Fig. 8). This indicates that the reaction products in 20YSZ have a lower
precipitation rate, and thus cannot seal the GB effectively. In view of
this, the faster precipitation rate of reaction products, the slighter GB
corrosion of YSZ.
Table 4
The crystal structure, average chemical composition of intra-grain and grain
boundary, and the Y segregation extent in GB of different YSZ materials. 4.3. Implication for mitigating GB corrosion

Name Crystal Composition Composition Y segregation


Based on the above discussions, some mitigation strategies are pro­
structure of bulk of GB extent
posed. Firstly, the GB corrosion can be mitigated by decreasing the GB
4YSZ Tetragonal Zr0.96Y0.04O1.98 Zr0.9Y0.1O1.95 150%
segregation of Y element. The reason of Y segregation is related to the
[42]
20YSZ Cubic Zr0.8Y0.2O1.7 Zr0.67Y0.33O1.11 65% oxygen vacancy induced by YO1.5 doping in ZrO2 [40,44]. Therefore,
[44] elimination of the oxygen vacancies may suppress the Y segregation and
YZO Cubic Zr0.5Y0.5O1.75 Zr0.4Y0.6O1.5 20% thereby decrease the GB corrosion. For example, equimolar trivalent and
pentavalent cations (e.g. Y3+ and Ta5+) co-dope ZrO2 without oxygen
vacancies is expected to have enhanced CMAS corrosion resistance. On
Y0.2 Zr0.8 O1.9 + 0.039CaO(CMAS)→0.978Zr0.818 Y0.148 Ca0.04 O1.889 (#2) the other hand, a large relative volume of rection products can induce
+ 0.06YO1.5 (CMAS) (2) the formation of dense reaction layer, and therefore mitigate the GB
corrosion. For YSZ with high YO1.5 content, co-precipitation of c-ZrO2
Y0.5 Zr0.5 O1.75 + 0.109CaO(CMAS) and apatite can form the dense layer. For other TBC materials, the ideal
+ 0.22SiO2 (CMAS)0.744Zr0.654 Y0.323 Ca0.023 O1.816 (CZrO2 ) reaction products will be further investigated. Additionally, the rapid
precipitation of reaction products can seal the GB and inhibit the further
+ 0.588Ca0.157 Y0.442 Zr0.023 Si0.378 O1.622 (Apatite) (3)
corrosion. Therefore, it is helpful to promote the precipitation rate or
With the combination of those equations and the unit-cell volume of develop the new product with high precipitation rate. For example, the
YSZ and reaction products, the volume change of crystalline materials (i. precipitation kinetics of reaction product can be enhanced by adding
e. VP/VD) during corrosion can be estimated. For 20YSZ, VP/VD in extra oxides in TBC as nucleating agent, just like the TiO2 in precipita­
product #1 and #2 is 0.89 and 0.98, respectively. This indicates that the tion of anorthite (CaAl2Si2O8)[54].
volume of precipitated products is less than or close to the volume of
dissolved bulk materials, which explains the lack of dense reaction 5. Conclusion
products layer in 20YSZ, as shown in Fig. 7c-f. For Y2Zr2O7, VP/VD is
calculated as 1.3, which indicates that the precipitation of products This study shows that the severe GB corrosion causes the rapid CMAS
leads to a volume expansion of ~30%. This explains the formation of infiltration in APS 8YSZ TBC. Further investigation shows that all YSZ
dense reaction product layer in Y2Zr2O7 shown in Fig. 7g and i. The materials exist the preferential GB corrosion, whereas the GB corrosion
formation of dense reaction layer inhibits the CMAS infiltration and extent depends on the YO1.5 content in YSZ. The mechanism of GB
therefore decreases the overall corrosion rate. Correspondingly, the GB corrosion is elucidated, and the extent difference of GB corrosion in
corrosion rate or extent also decreases in the YSZ with high YO1.5 con­ various YSZ are discussed. Some key conclusions are as follows:
tent (e.g., Y2Zr2O7).

11
D. Wu et al. Corrosion Science 209 (2022) 110803

• Preferential GB corrosion is observed in all YSZ ceramics, which is [15] W. Song, S. Yang, M. Fukumoto, Y. Lavallée, S. Lokachari, H. Guo, Y. You, D.
B. Dingwell, Impact interaction of in-flight high-energy molten volcanic ash
attributed to the GB segregation of Y element, which promotes the
droplets with jet engines, Acta Mater. 171 (2019) 119–131.
CMAS reactivity of GB region. [16] X. Shan, Z. Zou, L. Gu, L. Yang, F. Guo, X. Zhao, P. Xiao, Buckling failure in air-
• GB corrosion is more pronounced in YSZ ceramics with low YO1.5 plasma sprayed thermal barrier coatings induced by molten silicate attack, Scr.
contents (e.g. 8YSZ and 20YSZ) than those with high YO1.5 content Mater. 113 (2016) 71–74.
[17] X. Zhang, X. Shan, P.J. Withers, X. Zhao, P. Xiao, Tracking the calcium-magnesium-
(e.g. Y2Zr2O7), which can be related to the different levels of Y alumino-silicate (CMAS) infiltration into an air-plasma spray thermal barrier
segregation, and the relative volume as well as precipitation kinetics coating using X-ray imaging, Scr. Mater. 176 (2020) 94–98.
of reaction products. [18] X. Chen, Calcium–magnesium–alumina–silicate (CMAS) delamination mechanisms
in EB-PVD thermal barrier coatings, Surf. Coat. Technol. 200 (11) (2006)
• GB corrosion can be mitigated by decreasing Y segregation, 3418–3427.
increasing the relative volume of products, and/or promoting pre­ [19] S. Krämer, S. Faulhaber, M. Chambers, D.R. Clarke, C.G. Levi, J.W. Hutchinson, A.
cipitation rate of products. G. Evans, Mechanisms of cracking and delamination within thick thermal barrier
systems in aero-engines subject to calcium-magnesium-alumino-silicate (CMAS)
penetration, Mater. Sci. Eng.: A 490 (1) (2008) 26–35.
CRediT authorship contribution statement [20] S. Krämer, J. Yang, C.G. Levi, C.A. Johnson, Thermochemical interaction of
thermal barrier coatings with molten CaO-MgO-Al2O3-SiO2(CMAS) deposits, J. Am.
Ceram. Soc. 89 (10) (2006) 3167–3175.
Di Wu: Conceptualization, Methodology, Formal analysis, Investi­ [21] M.H. Vidal-Setif, N. Chellah, C. Rio, C. Sanchez, O. Lavigne,
gation, Data curation, Writing – original draft. Xiao Shan: Resources, Calcium–magnesium–alumino-silicate (CMAS) degradation of EB-PVD thermal
Validation, Writing – review & editing, Funding acquisition. Huangyue barrier coatings: Characterization of CMAS damage on ex-service high pressure
blade TBCs, Surf. Coat. Technol. 208 (2012) 39–45.
Cai: Resources, Writing – review & editing, Funding acquisition. Fan [22] H.F. Garces, B.S. Senturk, N.P. Padture, In situ Raman spectroscopy studies of high-
Yang: Writing – review & editing, Supervision. Xiaofeng Zhao: Su­ temperature degradation of thermal barrier coatings by molten silicate deposits,
pervision, Funding acquisition. Scr. Mater. 76 (2014) 29–32.
[23] A.R. Krause, H.F. Garces, G. Dwivedi, A.L. Ortiz, S. Sampath, N.P. Padture, Calcia-
magnesia-alumino-silicate (CMAS)-induced degradation and failure of air plasma
sprayed yttria-stabilized zirconia thermal barrier coatings, Acta Mater. 105 (2016)
Declaration of Competing Interest 355–366.
[24] S. Sampath, U. Schulz, M.O. Jarligo, S. Kuroda, Processing science of advanced
thermal-barrier systems, MRS Bull. 37 (10) (2012) 903–910.
The authors declare that they have no known competing financial
[25] S. Morelli, V. Testa, G. Bolelli, O. Ligabue, E. Molinari, N. Antolotti, L. Lusvarghi,
interests or personal relationships that could have appeared to influence CMAS corrosion of YSZ thermal barrier coatings obtained by different thermal
the work reported in this paper. spray processes, J. Eur. Ceram. Soc. 40 (12) (2020) 4084–4100.
[26] D.E. Mack, R. Laquai, B. Müller, O. Helle, D. Sebold, R. Vaßen, G. Bruno, Evolution
of porosity, crack density, and CMAS penetration in thermal barrier coatings
Data Availability subjected to burner rig testing, J. Am. Ceram. Soc. 102 (10) (2019) 6163–6175.
[27] C.S. Holgate, G.G.E. Seward, A.R. Ericks, D.L. Poerschke, C.G. Levi, Dissolution and
The raw/processed data required to reproduce these findings cannot diffusion kinetics of yttria-stabilized zirconia into molten silicates, J. Eur. Ceram.
Soc. 41 (3) (2021) 1984–1994.
be shared at this time as the data also forms part of an ongoing study. [28] F.H. Stott, D.J. de Wet, R. Taylor, Degradation of thermal-barrier coatings at very
high temperatures, MRS Bull. 19 (10) (2013) 46–49.
[29] L. Guo, Y. Gao, Y. Cheng, J. Sun, F. Ye, L. Wang, Microstructure design of the laser
Acknowledgement
glazed layer on thermal barrier coatings and its effect on the CMAS corrosion,
Corros. Sci. 192 (2021), 109847.
The authors thank National Natural Science Foundation of China [30] X. Zhou, T. Chen, J. Yuan, Z. Deng, H. Zhang, J. Jiang, X. Cao, Failure of plasma
(No. 51971139, 52102071 and 52102072). sprayed nano-zirconia-based thermal barrier coatings exposed to molten CaO-
MgO-Al2O3-SiO2 deposits, J. Am. Ceram. Soc. 102 (10) (2019) 6357–6371.
[31] D.L. Poerschke, C.G. Levi, Effects of cation substitution and temperature on the
References interaction between thermal barrier oxides and molten CMAS, J. Eur. Ceram. Soc.
35 (2) (2015) 681–691.
[32] A.R. Krause, H.F. Garces, B.S. Senturk, N.P. Padture, D.J. Green, 2ZrO2⋅Y2O3
[1] D.R. Clarke, M. Oechsner, N.P. Padture, Thermal-barrier coatings for more efficient
thermal barrier coatings resistant to degradation by molten CMAS: part II,
gas-turbine engines, MRS Bull. 37 (10) (2012) 891–898.
interactions with sand and fly ash, J. Am. Ceram. Soc. 97 (12) (2014) 3950–3957.
[2] N.P. Padture, M. Gell, E.H. Jordan, Thermal barrier coatings for gas-turbine engine
[33] J.M. Drexler, A.L. Ortiz, N.P. Padture, Composition effects of thermal barrier
applications, Science 296 (5566) (2002) 280–284.
coating ceramics on their interaction with molten Ca–Mg–Al–silicate (CMAS) glass,
[3] R.A. Miller, Thermal barrier coatings for aircraft engines: history and directions,
Acta Mater. 60 (15) (2012) 5437–5447.
J. Therm. Spray. Technol. 6 (1) (1997) 35–42.
[34] J.M. Drexler, A.D. Gledhill, K. Shinoda, A.L. Vasiliev, K.M. Reddy, S. Sampath, N.
[4] B. Gleeson, Thermal barrier coatings for aeroengine applications, J. Propuls. Power
P. Padture, Jet engine coatings for resisting volcanic ash damage, Adv. Mater. 23
22 (2) (2006) 375–383.
(21) (2011) 2419–2424.
[5] D.R. Clarke, S.R. Phillpot, Thermal barrier coating materials, Mater. Today 8 (6)
[35] E.R. Andrievskaya, Phase equilibria in the refractory oxide systems of zirconia,
(2005) 22–29.
hafnia and yttria with rare-earth oxides, J. Eur. Ceram. Soc. 28 (12) (2008)
[6] C.G. Levi, Emerging materials and processes for thermal barrier systems, Curr.
2363–2388.
Opin. Solid State Mater. Sci. 8 (1) (2004) 77–91.
[36] X. Ren, W. Pan, Mechanical properties of high-temperature-degraded yttria-
[7] R. Vassen, M.O. Jarligo, T. Steinke, D.E. Mack, D. Stoever, Overview on advanced
stabilized zirconia, Acta Mater. 69 (2014) 397–406.
thermal barrier coatings, Surf. Coat. Technol. 205 (4) (2010) 938–942.
[37] M.P. Borom, C.A. Johnson, L.A. Peluso, Role of environment deposits and
[8] C.G. Levi, J.W. Hutchinson, M.-H. Vidal-Sétif, C.A. Johnson, Environmental
operating surface temperature in spallation of air plasma sprayed thermal barrier
degradation of thermal-barrier coatings by molten deposits, MRS Bull. 37 (10)
coatings, Surf. Coat. Technol. 86- 87 (1996) 116–126.
(2012) 932–941.
[38] X. Shan, H. Cai, L. Luo, F. Guo, X. Zhao, Influence of pore characteristics of air
[9] W. Song, Y. Lavallée, K.-U. Hess, U. Kueppers, C. Cimarelli, D.B. Dingwell, Volcanic
plasma sprayed thermal barrier coatings on calcia-magnesia-alumino-silicate
ash melting under conditions relevant to ash turbine interactions, Nat. Commun. 7
(CMAS) attack behavior, Corros. Sci. 190 (2021), 109636.
(2016) 10795.
[39] X. Shan, W. Chen, L. Yang, F. Guo, X. Zhao, P. Xiao, Pore filling behavior of air
[10] D.L. Poerschke, R.W. Jackson, C.G. Levi, Silicate deposit degradation of engineered
plasma spray thermal barrier coatings under CMAS attack, Corros. Sci. 167 (2020),
coatings in gas turbines: progress toward models and materials solutions, Annu.
108478.
Rev. Mater. Res. 47 (2017) 297–330.
[40] B. Feng, T. Yokoi, A. Kumamoto, M. Yoshiya, Y. Ikuhara, N. Shibata, Atomically
[11] W.S. Walsh, K.A. Thole, C. Joe, Effects of sand ingestion on the blockage of film-
ordered solute segregation behaviour in an oxide grain boundary, Nat. Commun. 7
cooling holes. Power for Land, Sea, and Air, ASME, Turbo Expo, 2006, p. 2006.
(2016) 11079.
[12] E.M. Zaleski, C. Ensslen, C.G. Levi, Melting and crystallization of silicate systems
[41] K. Matsui, H. Yoshida, Y. Ikuhara, Isothermal sintering effects on phase separation
relevant to thermal barrier coating, Damage, J. Am. Ceram. Soc. 98 (5) (2015)
and grain growth in yttria-stabilized tetragonal zirconia polycrystal, J. Am. Ceram.
1642–1649.
Soc. 92 (2) (2009) 467–475.
[13] C. Mercer, S. Faulhaber, A.G. Evans, R. Darolia, A delamination mechanism for
[42] K. Matsui, H. Yoshida, Y. Ikuhara, Grain-boundary structure and microstructure
thermal barrier coatings subject to calcium–magnesium–alumino-silicate (CMAS)
development mechanism in 2–8mol% yttria-stabilized zirconia polycrystals, Acta
infiltration, Acta Mater. 53 (4) (2005) 1029–1039.
Mater. 56 (6) (2008) 1315–1325.
[14] R.W. Jackson, E.M. Zaleski, D.L. Poerschke, B.T. Hazel, M.R. Begley, C.G. Levi,
Interaction of molten silicates with thermal barrier coatings under temperature
gradients, Acta Mater. 89 (2015) 396–407.

12
D. Wu et al. Corrosion Science 209 (2022) 110803

[43] B. Feng, N.R. Lugg, A. Kumamoto, Y. Ikuhara, N. Shibata, Direct observation of degradation by molten CMAS: part I, optical basicity considerations and
oxygen vacancy distribution across yttria-stabilized zirconia grain boundaries, ACS processing, J. Am. Ceram. Soc. 97 (12) (2014) 3943–3949.
Nano 11 (11) (2017) 11376–11382. [49] J.A. Duffy, M.D. Ingram, An interpretation of glass chemistry in terms of the optical
[44] Y.Y. Lei, Y. Ito, N.D. Browning, T.J. Mazanec, Segregation effects at grain basicity concept, J. Non-Cryst. Solids 21 (3) (1976) 373–410.
boundaries in fluorite-structured ceramics, J. Am. Ceram. Soc. 85 (9) (2002) [50] J.A. Duffy, Acid–base reactions of transition metal oxides in the solid state, J. Am.
2359–2363. Ceram. Soc. 80 (6) (1997) 1416–1420.
[45] J.A. Duffy, Optical basicity analysis,of glsses containing trivalent scandium, [51] T. Nanba, Y. Miura, S. Sakida, Consideration on the correlation between basicity of
yttrium, gallium and indium, Phys. Chem. Glass 46 (5) (2005) 500–504. oxide glasses and O1s chemical shift in XPS, J. Ceram. Soc. Jpn. 113 (1313) (2005)
[46] L.S. Dent-Glasser, J.A. Duffy, Analysis and prediction of acid–base reactions 44–50.
between oxides and oxysalts using the optical basicity concept, J. Chem. Soc. [52] V. Dimitrov, S. Sakka, Electronic oxide polarizability and optical basicity of simple
Dalton Trans. (10) (1987) 2323–2328. oxides. I, J. Appl. Phys. 79 (3) (1996) 1736–1740.
[47] D. Ghosh, V.A. Krishnamurthy, S.R. Sankaranarayanan, Application of optical [53] S. Krämer, J. Yang, C.G. Levi, Infiltration-inhibiting reaction of gadolinium
basicity to viscosity of high alumina blas furnace slags, J. Min. Metall. Sect. B- zirconate thermal barrier coatings with CMAS melts, J. Am. Ceram. Soc. 91 (2)
Metall. 46 (1) (2010) 41–49. (2008) 576–583.
[48] A.R. Krause, B.S. Senturk, H.F. Garces, G. Dwivedi, A.L. Ortiz, S. Sampath, N. [54] A. Aygun, A.L. Vasiliev, N.P. Padture, X. Ma, Novel thermal barrier coatings that
P. Padture, D.J. Green, 2ZrO2⋅Y2O3 thermal barrier coatings resistant to are resistant to high-temperature attack by glassy deposits, Acta Mater. 55 (20)
(2007) 6734–6745.

13

You might also like