Skip to main content
Erschienen in: Journal of Materials Science 1/2024

Open Access 20.12.2023 | Ceramics

Wear behavior at high temperature of ZrO2–Y2O3 (YSZ) plasma-sprayed coatings

verfasst von: D. Franco, F. Vargas, E. López, H. Ageorges

Erschienen in: Journal of Materials Science | Ausgabe 1/2024

Aktivieren Sie unsere intelligente Suche, um passende Fachinhalte oder Patente zu finden.

search-config
loading …

Abstract

The wear behavior of two plasma-sprayed zirconia–yttria coatings was studied at high temperatures. Agglomerated and sintered, as well as fused and crushed zirconia–yttria feedstock powders were used to manufacture bimodal and monomodal coatings by atmospheric plasma spraying onto an INCONEL 718 substrate previously coated with a NiCrAlY bond coat. The structure of the coatings was analyzed by SEM on their cross section and surface. The samples were subjected to wear conditions by sliding contact through a ball-on-disk test up to 1000 °C, using an alumina ball 6 mm in diameter as the counterbody, on which a load of 5 N was applied. The samples were rotated during 20000 cycles, reaching a speed of 0.10 m·s−1 at the contact area with the counterbody. The porosity, phase, and mechanical properties were determined before and after wear tests. The results indicate that at 25 °C, both coatings have enough mechanical resistance to withstand the tribological conditions they were exposed to. Therefore, low wear rates were produced by ductile deformation. The tribological conditions became more aggressive as the thermal stresses increased with the test temperature, producing cracking, and detaching particles in the coatings tested at 500 and 750 °C. Consequently, high wear rates related to brittle deformation were obtained. However, the transformation of the amorphous phase to the t′-zirconia phase, produced at 1000 °C, increased the hardness of both coatings and, consequently, their wear resistance; thus, the predominant mechanism of damage was ductile deformation, with wear rates similar to those obtained when the coatings were tested at 25 °C.
Hinweise
Handling Editor: David Cann.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Introduction

Atmospheric plasma-sprayed (APS) zirconia–yttria top coatings deposited on NiCrAlY bond coats have been widely used as thermal barrier coatings (TBCs) for gas turbine blades and combustion system components owing to their high thermal stability, low thermal conductivity, and relatively large thermal expansion, which is close to that of a metallic substrate [13]. Under operating conditions, TBCs are exposed to high-temperature oxidation, hot corrosion, and severe wear, such as erosion, adhesion, abrasion, and fretting [3].
The mechanisms and chemical reactions that produce high-temperature oxidation and hot corrosion of yttria-stabilized zirconia (YSZ) TBCs manufactured by APS have been studied to identify ways to improve their performance and useful life [47]. Recently, studies have been conducted to understand the tribological behavior of atmospheric-pressure plasma-sprayed YSZ thermal barrier coatings exposed to conditions similar to those in operation. Specifically, D. Shin et al. [8] have evaluated the erosion resistance of these coatings at temperatures between 537 and 980 °C, using an erosion tunnel to simulate the modern gas turbine operating conditions. The erodent particles were transported by gas at velocities between 122 and 305 m·s−1, impacting the coating surface at angles between 20° and 90°. The effect of the coating structure porosity (porosities of 12.9 ± 0.5% and 19.5 ± 1.2%) on the erosion resistance was evaluated. These results demonstrate that higher wear rates due to erosion are associated with increased porosity in the YSZ coatings [8].
Similarly, Pakseresht et al. [3] studied the wear behavior of atmospheric plasma-sprayed YSZ coatings manufactured from Metco 204NS powder, with and without the addition of alumina whiskers, using a ball-on-disk microtribometer at room temperature to promote abrasive conditions on the contact surface between the coating and counterbody. An alumina ball with a diameter of 5 mm was used as the counterbody, on which normal loads of 7, 10, and 13 N were applied for each test. During the tests, the samples were rotated at a linear speed of 0.5 m·s−1 up to a sliding distance of 500 m. The wear track analysis reported by the authors indicated that a smooth surface and abrasive detachment of particles were produced with wear rates between 4.1 \(\times\) 10–2 and 7.3 \(\times\) 10–2 mm3/N·m for the coating without adding alumina whiskers and between 3.5 \(\times\) 10–2 and 5.3 \(\times\) 10–2 mm3/N·m for specimens reinforced with alumina whiskers [3].
Liang et al. [9], Pawlowski [10], Shi et al. [11], Xiao et al. [12], and Lima et al. [13] reported that plasma thermally sprayed coatings using nanometric and submicrometric feedstock powders improved the mechanical properties by improving the coating’s structure. Additionally, H. Chen et al. [14] reported better wear performance of zirconia coatings manufactured from nanometric feedstock powders than that of the coatings sprayed from micrometric powders, which was attributed in the same way to the optimization of their structure, and therefore, the improvement of their mechanical properties.
L. Bai et al. [1517] studied the tribological performance of YSZ coatings exposed to sliding contact with an alumina ball from 25 to 800 °C. The results obtained by these researchers showed that the alumina ball used as a counterpart produces severe wear on the surface of these coatings, not only at room temperature [17], but also up to 800 °C [16]. Likewise, the results of these studies show that as the test temperature increases, the rate of coating wear tends to increase. However, the formation of a tribolayer produced on the surface of the coatings when they were tested at temperatures above 200 °C protects them from the damage produced by the alumina counterpart, reducing their wear rate [15, 16].
The topcoats of thermal barrier coatings are frequently manufactured by atmospheric plasma spraying from yttria-stabilized zirconia powders, which were previously fused and crushed, or agglomerated and sintered, using powder processing methods. Although fused and crushed powders are usually less expensive, agglomerated and sintered powders are commonly used to manufacture coatings that are exposed to high temperatures because their bimodal structure gives them higher thermal shock resistance than the monomodal structure obtained in coatings sprayed from fused and crushed powders [18]. At room temperature, the mechanical properties (hardness, elastic modulus, and fracture toughness) of monomodal and bimodal structure YSZ coatings could be statistically similar [19]. The phase transformations at high temperatures could change the mechanical performance of these coatings. Additionally, the yttria content used in the feedstock powders plays an essential role in the stability of the t′-ZrO2 phase when zirconia-based materials are exposed to high temperatures for a long time. It has been reported that the complete and maximum stability of the t′-ZrO2 phase is achieved when the content of Y2O3 is above 6 wt% [20].
A few studies have reported the performance of these coatings exposed to sliding contact with an alumina ball at temperatures up to 800 °C [16, 17]. However, their behavior at higher temperatures have not yet been published. For this reason, this work studied the wear performance of YSZ coatings exposed to abrasive conditions at temperatures between 25 and 1000 °C. The coatings studied were manufactured by atmospheric plasma spraying from agglomerated and sintered, as well as fused and crushed ZrO2–Y2O3 feedstock powders, to produce bimodal and monomodal microstructures, respectively, which are specified for thermal barrier coatings in aircraft, stationary gas turbines, and engines with high thermal shock resistance, thermal insulating properties, and hot corrosion resistance [21].

Materials and methods

To prepare the substrates, an INCONEL 718 bar was cut into discs with a diameter of 25 mm and a height of 7 mm. The surface to be coated was blasted using a corundum jet of particles, reaching an arithmetic average roughness (Ra) between 4 and 10 μm. Subsequently, the substrates were sonicated in an acetone bath to remove residues from the treatment previously carried out with abrasive particles and other dirt. NiCrAlY Sulzer–Metco Amdry 962 powder was atmospheric plasma sprayed as a bond coat on an INCONEL 718 substrate. Afterward, ZrO2–Y2O3 top coatings were also manufactured by APS from the agglomerated and sintered H.C. Starck Amperit 827.423 and fused and crushed H.C. Starck Amperit 825 powders to produce bimodal and monomodal microstructures, respectively. The bond and top coatings are manufactured using a Sulzer–Metco PTF4 plasma torch according to the parameters listed in Table 1. These parameters were selected from preliminary tests carried out looking for coatings with the crystalline and amorphous phases, as well as the mechanical properties usually required for their use as topcoat in thermal barriers.
Table 1
Plasma-spraying parameters
Parameter
Top coating powders
Bond coating powder
H.C. Starck Amperit 827.423
H.C. Starck Amperit 825.1™
Sulzer–Metco Amdry 962™
Current intensity [A]
650
650
650 
Ar-H2 flow rate [L/min]
45-15
45–15
45-15 
Nozzle internal diameter [mm]
7
7
7
Feeder type
Screw Praxair
Screw Praxair
Screw Praxair 
Powder flow rate [g/min]
22–28
24–30
15–19
Ar carrier gas pressure [bar]
5.0
5.0
5.0 
Ar carrier gas flow rate [L/min]
4.5
4.5
4.5
Spraying distance [mm]
100 ± 1
100 ± 1
100 ± 1 
Sample translation speed [mm/s]
24
24
24 
Sample rotation speed [rpm]
124
124
124 
Cooling air distance [mm]
12
12
12
Preheating temperature [°C]
300
300
300
Surface substrate preheating passes
5–8
5–8
2–3
Spraying time [min]
4
4
2
Number of spraying passes
95
93
55
The chemical compositions of the feedstock powders were measured using a wavelength-dispersive X-ray fluorescence (WD-XRF) spectrometer with commercial reference: Thermo Fisher SCIENTIFIC ARL™ OPTIM'X. In the same way, Horiba PARTITA LA-950V2 laser diffraction (LD) equipment was used to characterize the particle size distribution of these powders. The crystallographic composition of the feedstock powders and the coatings was determined using an X-ray Cu Kα 1 radiation (DRX) Diffractometer with commercial reference: Bruker D8 ADVANCE and the X’Pert Highscore Plus Software following the COD cards: (1) t′-ZrO2 (1525706), (2) c-ZrO2 (1521753), and (3) m-ZrO2 (1010912). Afterward, the Rietveld method was used to quantify the phases, following the same COD cards and the Material Analysis Using Diffraction (MAUD) software. In addition, a JEOL JSM IT-300 LV scanning electronic microscopy (SEM) equipment was used to characterize the morphological features of the ZrO2–Y2O3 feedstock powders particles, the surfaces, and the cross sections of the coatings, as well as the surface of wear tracks. The surfaces and the cross sections of the coatings were ground and polished according to the ASTM E1920 standard [22] to obtain an arithmetic average roughness (Ra) lower than 0.2 µm. The porosity was determined on the cross sections of the coatings from images taken by SEM according to the indications of the ASTM E2109 standard [23] and using the Image J software. On the other hand, the hardness, the elastic modulus, and the fracture toughness of the YSZ coatings were determined from indentations performed on the polished surfaces of all samples using a Shimadzu HMV-G20 equipment following the specifications of ASTM C-1327 [24] and ASTM E-384 [25] standards. The hardness, the elastic modulus, and the fracture toughness were calculated according to Eqs. (1)–(3), respectively:
$$H_{{\text{V}}} = 0.0018544\frac{{P_{{\text{N}}} }}{{d^{2} }}$$
(1)
where \({H}_{{\text{V}}}\) is the Vickers microhardness [GPa], \({P}_{{\text{N}}}\) is the normal load applied to the indenter [N], and \(d\) is the average length of the two diagonals produced during indentation [mm].
$$E = \frac{{ - \alpha H_{{\text{K}}} }}{{\left( {\frac{{b^{\prime } }}{{a^{\prime } }} - \frac{b}{a}} \right)}}$$
(2)
where \(E\) is Young’s modulus [GPa], \(\alpha\) is a constant \((\alpha =0.45)\), \({H}_{{\text{k}}}\) is the Knoop microhardness [Pa], \({a}{\prime}\) and \({b}{\prime}\) are the longer and shorter diagonals, respectively, produced by the indentation [µm], and \(a\) and \(b\) are the geometric constants of the indenter \((b/a=1/7.11)\), as in Fig. 1a.
$$K_{{{\text{IC}}}} = 0.0016\sqrt{\frac{E}{H}} \frac{{P_{{\text{N}}} }}{{C^{3/2} }}$$
(3)
where \({K}_{{\text{IC}}}\) is the fracture toughness [MPa·m1/2],\(E\) is the Young’s modulus [GPa], \(H\) is the Vickers microhardness [GPa], \({P}_{{\text{N}}}\) is the applied normal load on the indenter [N], and \(C\) is the longest radial crack produced during the indentation [mm], as shown in Fig. 1b.
Wear tests were performed at 25, 500, 750, and 1000 °C using a ball-on-disk tribometer under dry sliding contact without eliminating the formed debris. An alumina ball 6 mm in diameter, with a hardness Vickers of 18.0 ± 0.5 GPa, was used as a counter-body, on which a normal load of 5 N was applied. The samples were rotated during 20000 cycles reaching a relative linear speed of 0.1 m·s−1 with respect to the alumina ball, according to some recommendations of the ASTM G-99 standard [26]. Morphological characterization of the wear tracks produced during the tribological tests was performed using SEM with the aforementioned equipment. The wear rate was calculated from the profile curves of the wear tracks measured on the samples (Fig. 2) using a Surtronic S125 profilometer and Eq. (4).
$${\text{WR}}_{Sample} = \frac{{{\text{Volumen}}}}{{{\text{Load}} \times {\text{Distance}}}} = \frac{{A_{s} 2\pi r_{{{\text{wt}}}} }}{{1000P_{{\text{N}}} N_{{\text{c}}} 2\pi r_{{{\text{wt}}}} }}$$
(4)
where \({WR}_{{\text{Sample}}}\) denotes the wear rate [mm3/N·m],\({A}_{s}\) is the wear track cross-sectional area [µm2], \({r}_{{\text{wt}}}\) is the radius of the wear track [mm], \({P}_{{\text{N}}}\) is the applied normal load [N]; and \({N}_{{\text{c}}}\) is the total number of cycles.
In the same way, to calculate the wear rate produced for each counter-body, an electronic micrometer with commercial reference: Mitutoyo and Eq. (5) were used.
$${\text{WR}}_{{{\text{Counterbody}}}} = \frac{{{\text{Volumen}}}}{{{\text{Load}} \times {\text{Distance}}}} = \frac{{\frac{1}{3}\pi h^{2} \left( {3R - h} \right)}}{{P_{{\text{N}}} D_{{\text{T}}} }}$$
(5)
where \({WR}_{{\text{Counterbody}}}\)is the wear rate [mm3/N·m], \(h\) is the spherical cap height [mm], \(R\) is the radius of the counter-body [mm], \({P}_{{\text{N}}}\) is the normal load applied [N], and \({D}_{{\text{T}}}\) is the total distance of the test [m].
After the wear tests, the porosity, crystallographic phases, hardness, elastic modulus, and fracture toughness were reevaluated using the same equipment, standards, and equations mentioned above to compare the values with those obtained before the wear tests. The porosity, mechanical properties, and worn area were measured for three samples, ten times each, guaranteeing statistical reproducibility and repeatability for all measurements.

Results

Feedstock powder characterization

The results of the chemical analysis, the particle size distribution, and the morphological characterization carried out on the Sulzer–Metco Amdry 962 powder used to manufacture the bond coating indicated that it was composed of Ni (~ 67.0 wt%), Cr (~ 22.0 wt%), Al (~ 10.0 wt%), and Y (~ 1.0 wt%), its particle size distribution is between d10 = 63.59 and d90 = 121.73 µm, and its geometry is spheroidal typical of the atomized powders [27]. The results of the chemical analysis carried out on the feedstock powders used to manufacture the top coatings showed that both the H.C. Starck Amperit 827.423 powder and the H.C. Starck Amperit 825 powder were composed mainly of ZrO2 and Y2O3 with Al2O3, Na2O, K2O, TiO2, NiO, MgO, CaO, Fe2O3, and HfO2 in quantities lower than 0.5 wt%. Regarding the particle size distribution, the H.C. Starck Amperit 827.423 powder is significantly coarser (d10 = 20.51 and d90 = 83.17 µm) than the fused and crushed one (d10 = 24.39 and d90 = 49.67 µm). The results of the chemical analyses, phases, and particle size distributions of the powders used to produce the top coatings are listed in Table 2.
Table 2
Physicochemical features of the feedstock powders
Properties
H.C. Starck Amperit 827.423
H.C. Starck Amperit 825.1
Chemical
composition
(wt%)
ZrO2
93.89 ± 0.51
84.90 ± 0.55
Y2O3
3.21 ± 0.62
8.60 ± 0.58
SiO2
1.21 0.08
3.19 ± 0.07
Al2O3
0.46 ± 0.05
0.14 ± 0.06
CaO
0.13 ± 0.01
0.15 ± 0.01
TiO2
0.11 ± 0.01
0.11 ± 0.01
Na2O
0.08 ± 0.01
0.28 ± 0.02
K2O
0.14 ± 0.09
Fe2O3
0.06 ± 0.01
0.06 ± 0.01
NiO
0.18 ± 0.01
MgO
0.11 ± 0.01
HfO2
0.66 ± 0.13
2.05 ± 0.33
Others*
0.19 ± 0.05
0.09 ± 0.01
Phase
analysis
(wt%)
t′-ZrO2
63.1 ± 8.3
93.5 ± 3.8
m-ZrO2
32.4 ± 1.3
YZr8O14
4.5 ± 0.3
Amorphous
6.5 ± 0.4
Particle size
distritubtion
(µm)
d10
20.51
24.39
d50
51.81
34.92
d90
93.17
49.67
*Others: In2O3, WO3, Bi2O3, and Ga2O3
The amounts of yttria in the H.C. Starck Amperit 827.423 powder and the H.C. Starck Amperit 825 one are 3.21 and 8.60 wt%, respectively, which were enough to stabilize the 63.1 wt% and 93.5 wt% of tetragonal phase (t′-ZrO2) in these powders, respectively. The t′-ZrO2 is the characteristic phase of YSZ thermal barrier coatings. For this reason, most of the TBCs are yttria-stabilized zirconia containing ≈ 6.0–8.0 wt% of Y2O3 [28, 29]. It is important to note that under equilibrium conditions, yttria stabilizes a tetragonal phase above about 1050 °C [29]
The morphological analysis of these powders allowed us to identify that H.C. Starck Amperit 827.423 corresponds to spherical granules composed of sub-micrometric particles with nanoparticles in them, as well as some pores on their surface, as manufactured by agglomeration and sintering processes [27, 30]. In contrast, H.C. Starck Amperit 825 comprises particles with irregular morphology and fracture patterns, characteristic of fused and crushed powders [27]. The morphologies of these powders are shown in Fig. 3a, b.
The agglomerated and sintered, and the fused and crushed ZrO2–Y2O3 powders are widely used to manufacture thermal barrier coatings by APS [31, 32]. In order to identify the coatings manufactured from both the agglomerated and sintered powder H. C. Starck Amperit 827.423 and the fused and crushed H. C. Starck Amperit 825, they were codified as CA–S and CF–C, respectively.

Structural characterization of coatings

The structural analysis performed on the surfaces of both the CA–S and CF–C coatings revealed stacking of micrometrical splats, with some pores typical of thermally sprayed coatings [27, 33] (Fig. 4a, b). These porosities were produced mainly by discontinuities among the splats and were slightly more evident in the CA–S coating owing to the partially molten particles present in this sample [33]. It is essential to mention that these partially melted particles contain both submicrometric and nanometric particles, which gives these samples the features of a bimodal structure coating [33]. However, the cross-sectional structure showed good stacking among the lamellas and a homogenous interface between the top and bond coatings (Fig. 4c, d). The thicknesses and porosities of both coatings are listed in Table 3.
Table 3
Thickness and porosity of both ZrO2–Y2O3 coatings
Sample
Thickness [µm]
Wear tests temperature [°C]
Porosity before wear tests [%]
Porosity after wear tests [%]
ANOVA p-value
CA–S coating
275 ± 20
25
6.6 ± 1.2
6.5 ± 0.7
0.823
500
6.5 ± 0.5
0.812
750
6.4 ± 1.1
0.702
1000
6.6 ± 0.7
1.000
CF–C coating
273 ± 22
25
5.9 ± 0.7
5.9 ± 0.9
1.000
500
6.0 ± 0.4
0.701
750
5.9 ± 0.9
1.000
1000
5.9 ± 1.1
1.000
For both coatings, the porosity values before and after the wear tests at the different temperatures evaluated are statistically the same (all p-values are > 0.05), indicating that they do not experience sintering processes. Partially molten particles identified on both the surface and the cross section of CA–S coating are due to the granules of submicrometric and nanometric particles agglomerated and sintered and their consequent low heat transfer when they fly in the plasma jet [33].

Crystallographic characterization

The results of the XRD analysis performed on the manufactured coatings are shown in Fig. 5a, b, and the results of the quantification of phases identified are presented in Table 4. In Fig. 5a, b, the background of the XRD spectra was eliminated in order to compare among the results obtained for each sample tested at different temperatures. The diffraction peaks as well as the broadening and the hump evidenced ~ 30° indicate that both coatings were composed mainly of t′-ZrO2 and the amorphous phase (> 50.0 wt% and > 30.0 wt%, respectively). In addition, lower quantities of m-ZrO2 and c-ZrO2 phases were identified. The broadening and hump in the diffraction peak ~ 30° were previously reported by other authors [34, 35].
Table 4
Crystallographic phases in both ZrO2–Y2O3 coatings before and after wear tests
Sample
Phases
Wear tests temperature [°C]
Before wear tests [wt%]
After wear tests [wt%]
ANOVA p-value
CA–S coating
t′-ZrO2
25
51.6 ± 1.82
51.8 ± 1.4
0.723
500
51.1 ± 1.9
0.597
750
51.4 ± 1.2
0.852
1000
77.2 ± 1.0
0.000
m-ZrO2
25
1.6 ± 0.8
1.8 ± 0.3
0.370
500
1.1 ± 0.2
0.124
750
1.4 ± 0.3
0.644
1000
1.2 ± 0.3
0.204
c-ZrO2
25
2.5 ± 0.7
2.5 ± 0.7
1.000
500
2.5 ± 0.5
0.943
750
2.5 ± 0.5
0.860
1000
2.7 ± 0.7
0.706
YZr8O14
25
2.6 ± 0.5
2.8 ± 0.7
0.527
500
2.9 ± 0.9
0.429
750
2.8 ± 0.9
0.566
1000
2.7 ± 0.9
0.785
Amorphous
25
41.7 ± 1.3
41.0 ± 2.8
0.488
500
42.4 ± 2.0
0.403
750
41.9 ± 2.3
0.885
1000
16.3 ± 2.4
0.000
CF–C coating
t′-ZrO2
25
52.7 ± 1.4
52.8 ± 1.4
0.875
500
52.1 ± 1.9
0.433
750
52.4 ± 1.2
0.614
1000
78.2 ± 1.0
0.000
m-ZrO2
25
1.6 ± 0.8
1.5 ± 0.3
0.718
500
1.6 ± 0.2
1.000
750
1.5 ± 0.4
0.729
1000
1.4 ± 0.3
0.475
c-ZrO2
25
2.1 ± 0.5
2.0 ± 0.6
0.691
500
2.0 ± 0.4
0.628
750
2.1 ± 0.4
1.000
1000
1.8 ± 0.5
0.196
Amorphous
25
43.6 ± 2.9
43.7 ± 3.5
0.945
500
44.3 ± 3.9
0.655
750
44.0 ± 3.2
0.773
1000
21.2 ± 3.1
0.000
The m-ZrO2 and c-ZrO2 phases did not change significantly at different temperatures when the wear tests were performed (p > 0.05). On the other hand, from Table 4 is possible to see for both coatings that statistically, the quantities of the t′-ZrO2 phase and the amorphous phase were steady after the wear test carried out at 25, 500, and 750 °C (p-value > 0.05). However, after the wear tests performed at 1000 °C, the t′-ZrO2 phase increased, and the amorphous phase decreased (p-values < 0.005). It is important to note that other researchers have indicated that owing to the high cooling rate of particles deposited by atmospheric plasma spraying, the tetragonal phase obtained in the structure of coatings manufactured from ZrO2–Y2O3 powders is a metastable phase, called the tetragonal-prime phase (t′-ZrO2) [26, 3538]. For this reason, in this study, the notation t′-ZrO2 phase is used to refer to the t-ZrO2 phase.

Mechanical characterization

The mechanical properties determined before and after the tribological tests are presented in Table 5. Generally, the hardness of both coatings measured at room temperature is similar, and statistically significant changes were not identified (p-value > 0.05) after the wear test carried up to 750 °C. However, the hardness of both coatings after the wear test carried out at 1000 °C increased (p-value < 0.05) with the increase of the t′-ZrO2 phase, which could be produced by the crystallization of the amorphous phase during the wear tests performed at this temperature. The materials based in the t-ZrO2 phase have a higher hardness (~ 12.0 GPa) [38] than that of ZrO2-based materials, which also have an amorphous phase (~ 8.5 GPa) [39, 40].
Table 5
Mechanical properties of both ZrO2–Y2O3 coatings measured before and after the wear tests
Sample
Mechanical property
Wear tests temperature [°C]
Before wear tests [wt%]
After wear tests [wt%]
ANOVA p-value
CA–S coating
Hardness [GPa]
25
8.7 ± 0.5
8.3 ± 0.4
0.291
500
8.5 ± 0.5
1.000
750
8.8 ± 0.5
0.185
1000
9.2 ± 0.5
0.041
Young’s modulus [GPa]
25
82 ± 16
87 ± 7
0.399
500
79 ± 18
0.680
750
71 ± 16
0.115
1000
79 ± 13
0.558
Fracture toughness [MPa·m1/2]
25
3.2 ± 0.1
3.2 ± 0.1
1.000
500
3.2 ± 0.1
1.000
750
3.1 ± 0.2
0.181
1000
3.1 ± 0.2
0.181
CF–C coating
Hardness [GPa]
25
8.7 ± 0.6
8.5 ± 0.6
0.445
500
8.5 ± 0.5
0.561
750
8.6 ± 0.5
0.695
1000
9.3 ± 0.4
0.024
Young’s modulus [GPa]
25
90 ± 14
98 ± 12
0.210
500
101 ± 13
0.099
750
95 ± 19
0.556
1000
95 ± 11
0.386
Fracture toughness [MPa·m1/2]
25
2.9 ± 0.1
3.0 ± 0.4
0.456
500
2.8 ± 0.3
0.379
750
2.7 ± 0.4
0.161
1000
2.6 ± 0.4
0.032
It is also possible to see in Table 5 that at all temperatures evaluated, the values of fracture toughness for the CA–S coating are slightly higher than those for the CF–C coating, as well as that at 1000 °C, the value of fracture toughness for the CA–S coatings remained without any change (p-value > 0.05), while this value in the CF–C coatings decreased (p-value < 0.05), which could be related to the presence of partially melted particles in the CA–S coating. Regarding Young’s modulus values, these did not show statistically significant changes (p-value > 0.05) during the wear tests at all temperatures evaluated.

Tribological characterization

The wear track analysis of both YSZ coatings (Fig. 6a–p) demonstrated the development of different tribological mechanisms depending on the test temperature. In particular, both coatings have shown fuzzy wear tracks with friction marks and some slight spalling (Fig. 6a–d) after the tribological tests performed at 25 °C. Although the wear track of the coatings tested at 500 °C remains diffuse, the analysis performed at higher magnifications revealed the onset of both cracks and particle detachment from the coating (Fig. 6e–h), which are more evident on the samples tested at 750 °C in which the wear tracks are clearly evidenced (Fig. 6i–l). Other authors have reported comparable results with the occurrence of grooves and spalling pits in YSZ coatings exposed to tribological conditions similar to those used for the test performed at 500 °C, as well as delamination and ejection of wear particles owing to brittle fracture in coating performed at 800 °C [16]. Furthermore, regular wear tracks were observed in both coatings tested at 1000 °C and in a continuous layer consisting of debris particles where the plastic flow was identified (Fig. 6m–p).
The wear rate results are presented in Fig. 7a. They evidenced that the samples tested at 25 °C showed the lowest wear rate. In comparison, at 500 and 750 °C, the wear rates increased with the tribological test temperatures. Similar results were reported by other authors for YSZ coatings manufactured by APS and tested under similar tribological conditions [16]. For the samples tested at 1000 °C, the wear rates were decreased and showed values lower than those at 500 and 750 °C. For its part, Fig. 7b presents the wear rates produced on alumina counterbodies used to wear both coatings, showing an increase in the wear rate with the test temperature.
From Fig. 8a–h, it is possible to identify traces of friction on the worn counterbody surfaces, which is typical of abrasive wear, with no signs of wear by adhesion with the surface of coatings.
The friction coefficient values shown in Fig. 9, measured during the tribological tests for both coatings, increased with the increase of temperature until 750 °C. Then, they decreased for the samples evaluated at 1000 °C.

Discussion

In ceramic materials, as YSZ coatings manufactured by APS, the tribological performance is influenced mainly by their hardness and fracture toughness [41, 42], which depend, among other factors, on the crystalline and amorphous phases of which they are composed. It is important to note in Table 4, the increase of the t′-ZrO2 phase at the expense of the amorphous phase after the wear tests developed at 1000 °C (p-value < 0.05), which other researchers have previously reported, but for Al2O3–ZrO2 plasma sprayed coatings [43]. However, the quantity of the m-ZrO2 and c-ZrO2 phases did not change with the heating of the samples during the tribological tests at high temperatures. This could be linked to both the temperature and time at which these tests were carried out are not enough for the transformation of these phases. During the tribological evaluation, the samples were exposed to each test temperature for 4 h, which could be insufficient for the diffusion of an additional amount of Y2O3 in the t′-ZrO2 phase, as well as for Y2O3 diffusion from t′-ZrO2 to produce a mixture of stable tetragonal phases with monoclinic or cubic phases. Other authors have reported the transformation of t′-ZrO2 to m-ZrO2 in ZrO2-8 wt% Y2O3 coatings manufactured by APS and heated at 1100 °C for more than 800 h [20], as well as the transformation from t′-ZrO2 to c-ZrO2 in a single crystal of similar chemical composition heated at 1600 °C during 50 h [44]. Likewise, it has been reported that the transformation from t′-ZrO2 to m-ZrO2 in coatings manufactured by APS from powders with less than 6 wt% Y2O3 occurred slowly at 1300 °C [20]. In this order of ideas, since the m-ZrO2 and c-ZrO2 phases did not change and any decrease in porosity of the coatings was evidenced by sintering during tribological tests (Table 3), it was possible to establish that the key for the good wear behavior of coatings tested at 1000 °C, was the transformation from amorphous to t′-ZrO2 phase, thanks the higher hardness of this crystalline phase regarding the amorphous one [45].
The fracture toughness of the CA–S coating was slightly higher than that of the CF–C coating at all the temperatures evaluated. This indicates that the fine particles remaining inside the partially molten granules detected in the structure of the CA–S coating (Fig. 4a) can relax the stress and then arrest the cracks produced by the microindentations carried out to measure this mechanical property, as reported by other authors [18, 46]. Additionally, the fracture toughness of the CA–S coating remained constant after the wear tests performed at different temperatures. At the same time, for the CF–C coating, this property decreased slightly after the wear test was carried out at 1000 °C. Although in Eq. (3), the increase in microhardness (\(H\)) could promote the decrease in the \(E/H\) ratio, the decrease in the crack length (\({C}^{3/2}\)) owing to the presence of partially molten particles in the CA–S coating, in turn prompted an increase in the \({P}_{N}/{C}^{3/2}\) ratio, maintaining the fracture toughness at the end. On the other hand, in the CF–C coating, the absence of partially molten particles in its structure does not allow the reduction of the \(E/H\) ratio to be compensated by the increase of the \({P}_{N}/{C}^{3/2}\) ratio, due to the reduction of crack length after the tribological test carried out at 1000 °C.
The typical tribological mechanisms of ceramic materials under sliding contact conditions are functions of speed and load [46]. It is essential to highlight that when a ceramic material can withstand the mechanical stress applied by the counterbody, it produces a wear mechanism called “ductile deformation,” which usually shows features such as friction marks, and plastic flow, and, therefore, low wear rates, as obtained for coatings tested at 25 and 1000 °C [47]. Moreover, suppose the ceramic material cannot withstand the mechanical stress applied by the counterbody. In that case, it produces another wear mechanism called “brittle deformation,” which typically shows features such as fracture, cracks, and excessive detachment of particles, and therefore, high wear rates, as obtained for coatings tested at 500 and 750 °C [47]. Ductile and brittle deformation are wear mechanisms applicable only to ceramic materials [47].
It is important to note that despite using two powders with different chemical compositions (especially in terms of Y2O3 content) and morphologies, the results in Table 45 show that the manufactured coatings have similarities in the type and percentage of crystalline and amorphous phases, as well as in their mechanical properties, which gives them similar tribological behaviors at each temperature evaluated. However, this does not rule out a possible difference between the two types of coatings if they are evaluated at high temperatures for longer durations. In the coating manufactured from the powder containing 3 wt% Y2O3, the t′-ZrO2 phase could reach a transformation to the m-ZrO2 phase [37], while in that containing 8 wt% Y2O3, the t′-ZrO2 phase will be completely stable [20]. Thus, it is possible to establish that the decrease in the wear resistance of the two coatings is due to the increase in thermal stress with the temperature of the tests, promoting cracks and the detachment of particles by brittle deformation. However, when the two coatings were tested at 1000 °C, the amorphous to t′-ZrO2 phase transformation occurred. Their hardness increased, and a protective debris layer was produced, promoting the wear by ductile deformation.
The particle detachment evidenced on the wear tracks of samples tested at 500 °C, which was more notorious for the coating performed at 750 °C, is characteristic of wear produced by brittle deformation because the stresses applied by the hard alumina ball (~ 19 GPa) used as a counterbody during sliding contact are substantially higher than the mechanical resistance of the coatings, whose hardness is ~ 9 GPa  and their fracture toughness is 2.9–3.2 MPa·m1/2 [47]. On the other hand, the microcracks produced by the sliding of the counterbody on the surface of the samples are due to fatigue fracture, as has been previously reported by other authors [48, 49]. These cracks were more evident in the samples tested at 500 and 750 °C owing to the thermal stresses produced as the test temperature increased. In the coatings tested at 1000 °C, the fine debris could have decreased the stress contact between their surface and the alumina counterbody [50] and produced a continuous layer densified by this contact.
The transformation of the wear mechanism from ductile deformation produced at room temperature to brittle deformation at temperatures up to 800 °C was reported by J. H. Ouyang et al. [48] for ZrO2–Y2O3 coatings manufactured by low-pressure plasma spraying; however, the new transformation of wear mechanism toward ductile deformation occurring at 1000 °C that is presented in this study for ZrO2–Y2O3 coatings manufactured by atmospheric plasma spraying is unpublished. This transformation from wear by brittle deformation produced in the samples tested between 500 and 750 °C to wear by ductile deformation at 1000 °C (Fig. 6m–p) was mainly due to the increase in hardness produced by the crystallization of the amorphous phase, as well as the formation of a protective layer from debris. Figure 6n–p indicates the protective layer produced on the wear tracks of coatings tested at 1000 °C, evidenced by plastic flow described by other authors for wear with ductile deformation [47].
From Fig. 6a–p, it is possible to see similar wear behaviors for both coatings, starting with wear by ductile deformation, followed by wear by fragile deformation, and finally wear by ductile deformation. These wear behaviors have been reported in [51, 52] for ZrO2–Al2O3 and Al2O3 coatings, respectively. It is also important to note that the wear rate values reported in Fig. 7a are comparable to those reported in [16] for YSZ coatings tested at similar tribological conditions at 25 and 500 °C, as well as those reported in [51, 52] at high temperatures for Al2O3 and Al2O3–ZrO2 coatings, respectively.
As shown in Fig. 7a, the samples tested at 25 °C showed the lowest wear rate, which was likely due to the slight damage caused by ductile deformation during the tribological contact between the surface of the coating, the alumina ball, and the low quantity of debris (Fig. 6a–d). At 500 and 750 °C, the wear rates increased with the tribological test temperatures, probably because of the contribution of thermal stresses to the wear tests (Fig. 6e–l). For the samples tested at 1000 °C, the wear rate was lower than that at 500 °C and 750 °C, probably due to the reduction in the severity of the tribological contact promoted by the protective layer formed from debris. In the same way, Fig. 7b shows the increase of the counterbodies wear rates as the temperature increased in all samples (Fig. 8a–h), which was probably related to: (i) the hardness decrease of alumina ball when exposed to high temperatures [53], (ii) the increase in the severity of the contact conditions due to the increase in thermal stresses, and (iii) the increase of hardness in both coatings heated at 1000 °C owing to their crystallographic changes mentioned above.
The highest wear rate in the alumina ball used as the counter-body in tests performed at 1000 °C confirms that the increase in the hardness of the coatings was a driving factor for the change in the wear mechanism from brittle deformation at 750 °C to ductile deformation at 1000 °C. Wear-by-ductile deformation occurs when the tested material has sufficient mechanical resistance to withstand the contact conditions to which it is exposed, and a controlled quantity of rounded debris produces a layer that protects the sample [47, 50]. However, it can increase the severity of the damage produced to the counter-body by harder particles. The decrease in the friction coefficient measured in the tests performed at 1000 °C indicated that the debris that acted as a third body tended to be more rounded than those produced in the tests performed at 500 and 750 °C, where the wear was due to brittle deformation.
The friction coefficient values (Fig. 9) measured during the tribological tests for both coatings increased with the temperature increase until 750 °C. They then decreased for the samples evaluated at 1000 °C, which could be linked to the fine particles of debris in the protective layer, whose morphology tended to be mainly spherical, thus reducing this coefficient. The obtained coefficients of friction were similar to those previously reported for YSZ and PSZ materials against alumina [16, 54].

Conclusions

  • It was studied the wear behavior up to 1000 °C of two YSZ coatings widely used to manufacture thermal barrier coatings, which were manufactured by APS from both an agglomerated and sintered, and a fused and crushed ZrO2–Y2O3 feedstock powders aiming to produce bimodal and monomodal microstructures, respectively. The wear behavior is correlated with the mechanical properties, which depend on the crystalline phases. The results allowed us to identify the changes in the wear mechanism as a function of temperature.
  • Both ZrO2–Y2O3 coatings have shown no significant differences in their hardness before and after the wear tests up to 750 °C. However, the increase of the t′-ZrO2 phase at the expense of the amorphous phase during the tests performed at 1000 °C contributed to increasing the hardness and, therefore, the wear performances of the coatings. This increase in hardness prevented the severe cracking and particle detachment produced in the coatings tested at 500 and 750 °C due to high thermal stresses; therefore, only a limited amount of fine debris was produced, forming a protective layer on the contact surface of samples.
  • The wear mechanisms identified in both atmospheric plasma sprayed ZrO2–Y2O3 coatings were at 25 °C, ductile deformation, at 500 and 750 °C, brittle deformation, and at 1000 °C, ductile deformation again. This behavior in all the samples evaluated was strongly related to the thermal stresses and changes in their mechanical properties owing to their crystallographic phases.

Acknowledgements

The authors are grateful to Departamento Administrativo de Ciencia, Tecnologia e Innovacion—Colciencias—(Bogotá—Colombia). Convocatoria Doctorado Nacional—647 de 2014 for the Doctoral Fellowship, awarded to David Franco and the CODI-Committee of the University of Antioquia for its economic support given to both the GIPIMME—GIMMACYR research group in 2019 and for FIT 1-1-01 project.

Declarations

Conflict of interest

The authors declare that they have no known competing financial interests or personal relationships or conflict of interests that could have appeared to influence the work reported in this paper. The authors declare the following financial interests/personal relationships which may be considered as potential competing interests:

Ethical approval

Not Applicable.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Literatur
22.
Zurück zum Zitat ASTM E-1920 (2003) Standard guide for metallographic preparation of thermal spraying coatings ASTM E-1920 (2003) Standard guide for metallographic preparation of thermal spraying coatings
23.
Zurück zum Zitat ASTM E-2109 (2014) Standard test methods for determining area percentage porosity in thermal sprayed coatings ASTM E-2109 (2014) Standard test methods for determining area percentage porosity in thermal sprayed coatings
24.
Zurück zum Zitat ASTM C-1327 (2015) Standard test method for vickers indentation hardness of advanced ceramics ASTM C-1327 (2015) Standard test method for vickers indentation hardness of advanced ceramics
25.
Zurück zum Zitat ASTM E-384 (2017) Standard test method for microindentation hardness of materials ASTM E-384 (2017) Standard test method for microindentation hardness of materials
26.
Zurück zum Zitat ASTM G-99 (2017) Standard test method for wear testing with a pin-on-disk apparatus ASTM G-99 (2017) Standard test method for wear testing with a pin-on-disk apparatus
44.
Zurück zum Zitat Lanteri V, Heuer AH, Mitchell TE (1985) Tetragonal phase in the system ZrO2–Y2O3. J Am Ceram Soc 12:118–130 Lanteri V, Heuer AH, Mitchell TE (1985) Tetragonal phase in the system ZrO2–Y2O3. J Am Ceram Soc 12:118–130
53.
Metadaten
Titel
Wear behavior at high temperature of ZrO2–Y2O3 (YSZ) plasma-sprayed coatings
verfasst von
D. Franco
F. Vargas
E. López
H. Ageorges
Publikationsdatum
20.12.2023
Verlag
Springer US
Erschienen in
Journal of Materials Science / Ausgabe 1/2024
Print ISSN: 0022-2461
Elektronische ISSN: 1573-4803
DOI
https://doi.org/10.1007/s10853-023-09204-w

Weitere Artikel der Ausgabe 1/2024

Journal of Materials Science 1/2024 Zur Ausgabe

    Marktübersichten

    Die im Laufe eines Jahres in der „adhäsion“ veröffentlichten Marktübersichten helfen Anwendern verschiedenster Branchen, sich einen gezielten Überblick über Lieferantenangebote zu verschaffen.