You are on page 1of 13

Surface & Coatings Technology 327 (2017) 126–138

Contents lists available at ScienceDirect

Surface & Coatings Technology

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

CMAS behavior of yttrium aluminum garnet (YAG) and yttria-stabilized


zirconia (YSZ) thermal barrier coatings
Rishi Kumar a,⁎, Eric Jordan a, Maurice Gell a, Jeffrey Roth a, Chen Jiang b, Jiwen Wang b, Sarshad Rommel a
a
Institute of Materials Science, 97 North Eagleville Road, University of Connecticut, Storrs, CT 06269-3136, USA
b
HiFunda LLC, 270 Middle Turnpike, Storrs, Mansfield, CT 06269, USA

a r t i c l e i n f o a b s t r a c t

Article history: Calcium magnesium aluminosilicate (CMAS) that is formed from the ingested deposits in gas turbines degrades
Received 10 March 2017 thermal barrier coatings (TBCs), especially for the most widely used material; yttria-stabilized zirconia (YSZ). In
Revised 20 July 2017 the present work, we examine the behavior of yttrium aluminum garnet (YAG) as an alternative material for
Accepted in revised form 8 August 2017
TBCs. CMAS interaction studies were conducted by making composite pellets of YAG-CMAS and YSZ-CMAS pow-
Available online 09 August 2017
ders. These pellets, after being subjected to heat treatment between 1100 °C and 1500 °C were characterized by
Keywords:
X-ray Diffraction (XRD), scanning electron microscopy (SEM) and Energy-dispersive X-ray spectroscopy (EDS),
Thermal barrier coatings which showed YAG to be almost inert to CMAS whereas YSZ exhibited significant phase changes. To test the be-
Yttrium aluminum garnet havior of TBCs with YAG and 8YSZ as the topcoat material in a CMAS environment, cyclic furnace tests were con-
CMAS resistance ducted in which a controlled amount of CMAS was applied and then the samples were cycled to failure. In
Optical basicity addition, to simulate the continuous accumulation of CMAS expected in service, a cyclic furnace test was devised
Solution precursor plasma spray in which a small dose of aqueous solution of CMAS was applied on TBC specimens at the start of every cycle until
the samples were cycled to failure. In all these tests YAG TBCs outperformed YSZ in terms of durability. The mech-
anisms of CMAS attack are described and the relative resistance of YAG and YSZ is shown to be consistent with
the Optical Basicity (OB) theory.
© 2017 Elsevier B.V. All rights reserved.

1. Introduction monoclinic phase, accompanied by a detrimental volumetric expansion


and spallation [6], b) high sintering rates of YSZ which causes coating
Thermal barrier coatings (TBCs) are widely employed for thermal densification and loss of strain tolerance resulting in spallation [7] and
protection of gas turbine engines to improve their efficiency, durability c) attacks by molten calcium magnesium-aluminosilicate in the form
and performance in their high temperature operation [1–3]. Generally, a of dust, sand and runway debris. This infiltrated CMAS degrades the du-
TBC comprises four layers - a ceramic top coat, a thermally grown oxide rability and thermal characteristics of the coating through
(TGO), a bond coat and finally a superalloy substrate. TBCs have proven thermomechanical [8,9] as well as thermochemical effects [10,11]. Infil-
to be a critical component in hot sections of the gas turbines in order to trated CMAS causes high internal stresses in YSZ TBC microstructures
increase the turbine inlet temperature (TIT) and attain improved engine due to loss of strain tolerance and a decrease in the effective coefficient
efficiency. This has inevitably led to an increasing demand for yet higher of expansion, and consequent thermal cycling eventually leads to spall-
TIT, and hence, for TBCs that can effectively protect turbine components ation. Molten CMAS is also known to partially dissolve YSZ coatings,
at these higher temperatures. The state of art material used for TBCs is while selectively removing yttrium ions which destabilizes the metasta-
6–8 wt% yttria stabilized zirconia (YSZ) because of its favorable proper- ble tetragonal phase and results in phase transformation to monoclinic
ties such as high thermal expansion coefficient, high fracture toughness of YSZ accompanied by a 2–5% volumetric expansion causing spallation
[4] and low thermal conductivity [1,5]. However, use of YSZ is limited to in the coating [12,13].
a maximum temperature of 1200 °C because of a) transformation of A plethora of alternate material systems have been proposed and ex-
metastable tetragonal phase into tetragonal, and further into amined as prospective TBC candidates including pyrochlores, fluorites,
garnets and zirconates [14–17]. One such material has been gadolinium
zirconate which has shown excellent properties such as high melting
⁎ Corresponding author. point (N 2000 °C) [18], high temperature phase stability (≤ 1550 °C)
E-mail addresses: rishi.kumar@uconn.edu (R. Kumar), eric.jordan@uconn.edu
(E. Jordan), maurice.gell@uconn.edu (M. Gell), jeffrey.roth@uconn.edu (J. Roth),
[19], lower thermal conductivity than 8YSZ [5] and excellent CMAS re-
cjiang@hifundallc.com (C. Jiang), jwang@hifundallc.com (J. Wang), sistance by forming reaction products capable of blocking further
sarshad.rommel@uconn.edu (S. Rommel). CMAS infiltration in the top coat of TBC [20]. Despite having these

http://dx.doi.org/10.1016/j.surfcoat.2017.08.023
0257-8972/© 2017 Elsevier B.V. All rights reserved.
R. Kumar et al. / Surface & Coatings Technology 327 (2017) 126–138 127

exceptional properties, gadolinium zirconate has some weaknesses as pre-made solution, and was then mixed with the nitrate solution as a
well. Not only is it commercially expensive because of high stoichiomet- source of sulfur. After this step titanium sulfate solution and colloidal sil-
ric amounts of gadolinium, but also it lacks in some other critical mate- ica was added. pH of the solution was constantly measured and nitric
rial properties. These include poor thermodynamic compatibility with acid was used to keep the pH between 4 and 7 to ensure stability of col-
Al2O3 from the TGO layer [21], worse erosion resistance than air plasma loidal silica as per manufacturer's recommendation. Both CMAS solu-
sprayed (APS) 8YSZ coatings and lower fracture toughness than YSZ tions were dried overnight at 120 °C to convert the solution to gels,
[22]. which were later heat treated at 600 °C in a tube furnace (CM 1600
In our previous work [23,24], we demonstrated that yttrium alumi- Tube Furnace, Bloomfield N.J.). The chemical compositions of the two
num garnet (YAG) TBCs deposited by Solution Precursor Plasma Spray CMAS are given in Table 1. Melting points of the two CMAS were deter-
(SPPS) have superior properties compared to Air Plasma Sprayed mined using simultaneous differential scanning calorimetry (DSC) and
(APS) 8YSZ. These desirable properties include phase stability up to thermal gravimetric analysis (TGA), (Q600 SDT, 10 °C/min,
1500 °C, low thermal conductivity of 0.95 W/m·k at 1300 °C, improved 100 ml/min N2 flow, TA Instruments, New Castle DE). Viscosity of the
erosion resistance, higher thermal cycling lives and increased sintering two CMAS was calculated by Giordano [29] and Fluegel [30] model
resistance than APS YSZ [23]. In the current work, we discuss the reac- and has been shown in Table 2.
tivity of YAG with CMAS, while 8YSZ is used as the reference material.
In this paper, we have also compared the performance of SPPS YAG 2.2. Powder pellet testing
TBCs with APS 8YSZ coatings that are the current industrial standard.
Ideally, APS YAG coating would provide a head on comparison however, YAG powder (particle size b 50 μm) was obtained by mechanical
previous work show for APS YAG the formation of amorphous YAG [25] grinding of YAG agglomerate (Alfa Aesar, Ward Hill, MA) and then siev-
phase occurs which will be prone to crystallization during service and ing using a 50 μm wire mesh. YAG and 8YSZ (Metco 204 NS) powders
thus cracking. Also, an APS YAG is devoid of stress relieving vertical were then mixed with 9-CMAS powder (composition listed in Table 1)
cracks that are found in SPPS coatings thus will have a much larger ther- in a 1:1 ratio by weight. 5 wt% of PVA (low molecular weight, Alfa
mal expansion mismatch with the bottom TBC layers. A head to head Aesar, Ward Hill, MA) was added to the mixture to provide strength
comparison with the APS industrial standard is not done here. The aim to the green body. The mixture was ball milled, dried and compressed
of the study is to look at the application of a new TBC material processed to form powder pellets. The pellets were heat treated in a box furnace
by a relatively young SPPS process and show that it is a candidate to re- (CM Furnaces Inc., Bloomfield, N.J.) for 24 h at different temperatures
place the current industry standard APS YSZ in the presence of CMAS. At ranging from 1100 °C–1500 °C. Then the pellets were analyzed for reac-
the same time, we are looking for the first time at the nature of YAG in- tion using X-ray Diffractometer (D2 Phaser, Bruker AXS, Madison, WI).
teraction with CMAS. We have also explored the theory of Optical Basic- For SEM, a small piece of the heat-treated pellets was embedded in
ity which has been recently used to predict the reactivity between the epoxy resin, polished and finally sputtered with Au/Pd coatings. Field
top coat materials with CMAS [26,27]. This theory will be shown to be emission scanning electron microscopes (JSM-6350/5F, JEOL USA, Pea-
consistent with the current results. In the following, the CMAS behavior body, MA) and energy dispersive X-ray spectrometer (Noran system
of YAG TBCs will be studied. In each case YSZ samples are tested in the six EDS, Thermo, Waltham, MA) were used to further analyze the reac-
same experiment to provide a comparison basis using identical tions between ceramics and 9-CMAS.
conditions.
2.3. Thermal cycling
2. Materials and methods
Thermal cycling experiments were simultaneously conducted on
2.1. CMAS powder synthesis and characterization SPPS YAG and APS 8YSZ baseline samples alongside CMAS paste test
that is described in Section 2.4. The temperature profile is shown in
Four and nine components CMAS (4-CMAS, 9-CMAS) powders were Fig. 1. These tests will help evaluate the performance of both the coat-
synthesized using stoichiometric amounts of metal salts along with col- ings with and without CMAS in a test related to expected turbine engine
loidal silica and a titanium sulfate solution. 9-CMAS composition was performance. For the thermal cycling experiment and also the forth-
provided to us by our industrial partner while 4-CMAS composition coming CMAS experiments, three samples each of SPPS YAG and APS
was adopted from the study conducted by Drexler et al. [28]. Synthesis YSZ baselines were tested. This was done to account for the dispersion
of 4-CMAS aqueous solution was done by adding stoichiometric in the cyclic lives of the samples which is generally observed in furnace
amounts of hydrated nitrates salts of calcium, aluminum and magne- testing of TBCs.
sium (Ca(NO3)2·4H2O, Al(NO3)3·9H2O, Mg(NO3)2·6H2O, Alfa Aesar, Only SPPS YAG coatings (~225 μm thickness) were deposited at our
Ward Hill, MA), and Silica (LUDOX® TMA colloidal silica, Sigma-Aldrich spray booth on superalloy substrates provided to us by one of our indus-
Co. LLC. Spruce Street, MO) in de-ionized (DI) water. For 9-CMAS, ni- trial partners which already had a 100 μm MCrAlY bond coat and ~ 25
trates of calcium, aluminum, magnesium, iron, potassium and sodium μm APS YSZ as the inner layer. The spray conditions have been adopted
were dissolved in DI water. L-Cystine was dissolved separately in a from previous study [24] and are presented in Table 3. The APS 8YSZ

Table 1
Chemical compositions of 4 and 9 component CMAS used in paste and spritz testing.

Constituents 4-CMAS 4-CMAS 9-CMAS 9-CMAS 9-CMAS after sulfate decomposition 9-CMAS after sulfate decomposition
(wt%) (mol%) (wt%) (mol%) (wt%) (mol%)

CaSO4 – – 42.00 26.8 0.0 0.0


SiO2 52.30 51.5 31.57 45.6 41.9 45.6
Al2O3 7.10 4.1 16.50 14.0 21.9 14.0
MgO 3.50 5.2 2.66 5.7 3.5 5.7
CaO 37.10 39.2 1.54 2.4 25.0 29.1
Fe2O3 – – 1.54 0.8 2.0 0.8
K2O – – 1.54 1.4 2.0 1.4
TiO2 – – 1.33 1.4 1.8 1.4
Na2O – – 1.32 1.8 1.8 1.8
128 R. Kumar et al. / Surface & Coatings Technology 327 (2017) 126–138

Table 2 Table 3
CMAS viscosity calculated using Giordano and Fluegel model. Spraying parameters for SPPS YAG samples.

Viscosity (Pa ∗ s) 4-CMAS 9-CMAS Parameters

Temperature (°C) Giordano Fluegel Giordano Fluegel Plasma gun Metco 9 MB, GP gun nozzle
Gun power 45.5 kW
1180 376.6 33.2 11.9 4.4
Primary gas/secondary Ar/H2
1200 252.9 21.5 8.4 1.9
gas
Gas flow rate Ar: 52–57 l/min, H2:8–10 l/min
Precursor injection mode Atomization using BETE FC4 spray nozzles, 15 psi
baseline samples were provided to us by the same industrial partner pressure
and were tested in ‘as received’ condition. These samples had a 250 Precursor feed rate 18–20 ml/min
μm top coat of APS 8YSZ and same bondcoat and substrate as the SPPS Standoff distances 35–37 mm
YAG, allowing a fair comparison of coatings. For all the tests half of a Gun scan speed 550 mm/s
Raster scan step size 2 mm
sample was used which initially had a diameter of 25.4 mm.

2.4. CMAS paste test


3. Results
To evaluate the performance of TBCs, burner rig tests are widely
3.1. CMAS powder characterization
employed [31] to create a temperature gradient, thereby simulating ser-
vice conditions. In the current study, a simpler test was conceived to test
Fig. 2 shows the simultaneous DSC and TGA results of the two CMAS
the interaction of different CMAS compositions (Table 1) with SPPS YAG
compositions. The 4-CMAS exhibited a complete melting at 1203 °C but
and APS YSZ baseline samples described in Section 2.3. CMAS paste with
the melting started much earlier from around ~1165 °C. 9-CMAS melted
a concentration of 10 mg/cm2 was applied on the TBCs that were moist-
completely at 1174 °C·in order to run tests with both CMAS composi-
ened with DI water to aid paste homogeneity and spreading. This partic-
tions simultaneously, a temperature of 1180 °C was chosen for both
ular CMAS concentration was chosen based on the reported minimum
paste and spritz testing. We note that a weight change in both 4 and
level needed to initiate TBC damage by Wellman et al. [32]. These sam-
9-CMAS was observed. In case of 9-CMAS this can be attributed to the
ples were isothermally cycled in a bottom-loading isothermal furnace
loss of sulfur however a weight change in 4-CMAS was not expected.
(CM Furnaces Inc., Bloomfield, NJ) to 1180 °C (above the CMAS melting
It must be the case that volatile species (nitrates) survived the heat
temperature, Section 3.1) using 1-hour cycles (Refer to Fig. 1), with
treatment during synthesis of 4-CMAS. Table 1 shows the expected
10 min of ramp up time, 40 min of dwell time and 10 min of expedited
composition of 9-CMAS assuming sulfur volatilizes at and above 1100
cooling time using an external fan, until their failure. Failure was de-
°C.
fined by 50% of coating delamination from the substrate. Once the coat-
Temperature-dependent CMAS viscosity is an important parameter
ings failed, XRD was conducted on the surface of the coatings to study
that affects infiltration of coatings thus an attempt has been made to cal-
the reaction between the CMAS and TBCs; SEM and EDS was performed
culate CMAS viscosity using existing models and has been shown in
on the coating cross-sections to investigate the infiltration of CMAS in
Table 2. Previous study by Wiesner et. Al [33]. shows that Fluegel
the TBCs.
model predicted CMAS viscosity which was in reasonable agreement
with the experimental data, however it has compositional restrictions
2.5. CMAS spritz test which may provide inaccurate results for the 4-CMAS. Thus Giordano
model was used for the 4 component CMAS despite the fact that it pre-
Another CMAS test was carried out where 0.1 ml of 1 wt% aqueous dicted more than an order of magnitude higher than the experimental
solutions of CMAS precursor chemicals (4 and 9 Component) were values for a CMAS like composition. Because of this we also used the
sprayed on the TBC samples before the start of every cycle. The temper- Fluegel model to calculate CMAS viscosity. The temperature for the vis-
ature profile of the experiment was kept similar to the paste test and the cosity calculation (1180 °C) was chosen to match the temperature of
samples were cycled to failure. SEM and EDS were used to study the CMAS-TBC interaction in paste and spritz tests. Both the models predict
coating cross sections and investigate the infiltration of CMAS into the that 4-CMAS has a higher viscosity than 9-CMAS. The viscosity for 4-
TBCs. Synthesis of the two CMAS solutions was similar to as described CMAS was higher by a factor of 31 and 8 as predicted by Giordano
in Section 2.1 while keeping the equivalent solid loading to 1 wt% in [29] and Fluegel [30] models respectively.
DI water.
3.2. Powder pellet testing

Pellet testing was done using 9 component CMAS only. The cylindri-
cal shaped pellets after a 24-h exposure changed their shape depending
upon the temperature and the material. This change, captured in Fig. 3
was exacerbated in the case of 8YSZ + CMAS pellets which melted
completely whereas the YAG + CMAS pellet only became rounded.
Such a macroscopic behavior indicates reaction between CMAS and
8YSZ which led to dissolution and re-precipitation of 8YSZ particles
thus forming a ‘puddle’ of reaction products. The minimal change in
the shape of YAG based pellets establish the lack of interaction between
YAG and CMAS and the deformation resulted from preferential or local-
ized melting of CMAS in the pellets but held intact due to YAG matrix.
This theory can be further confirmed by Fig. 4 which shows the SEM im-
ages and EDS of the reaction zone in both 8YSZ and YAG during the heat
treatment at 1500 °C. While edge dissolution of ceramic particles can be
observed in both cases, the YAG particles are still rounded and discrete.
Fig. 1. Temperature profile of elevator furnace during CMAS paste and spritz testing. 8YSZ particles have changed their shape from rounded to irregular,
R. Kumar et al. / Surface & Coatings Technology 327 (2017) 126–138 129

Fig. 2. Simultaneous DSC and TGA data of a) 4-CMAS and b) 9-CMAS.

primarily due to dissolution in CMAS and re-precipitation. Some CMAS Interestingly, at 1100 °C–1300 °C some anorthite (CaAl2Si2O8) phase
is also trapped within the particles in the process. X-ray diffraction re- can also be seen with maximum intensity at 1300 °C. At higher temper-
sults for both YAG and YSZ based pellets are shown side by side for com- atures, anorthite phase is absent and the same anorthite observation
parison in Fig. 5. YSZ based pellets showed primarily tetragonal phase at holds true in YAG + CMAS pellets. With an increase in temperature
1100 °C and 1200 °C with small peaks of monoclinic phase. for YSZ, the tetragonal to monoclinic phase transformation increases,
as indicated by an increase in monoclinic peak intensity. At 1400 °C
and 1500 °C most of the tetragonal phase has converted to a combina-
tion of cubic and monoclinic. YAG based pellets show almost no reactiv-
ity with CMAS at all temperatures. Crystallization of anorthite is
observed between 1100 °C–1300 °C which disappears at higher temper-
atures. At 1400 °C peaks of Ca4Y6O(SiO4)6 is observed which does indi-
cate some reactions between CMAS and YAG but the phase seems to
disappear at 1500 °C.

3.3. Microstructure of SPPS YAG and APS 8YSZ baseline samples


Fig. 3. Pictures of powder pellets (a) Before heat treatment (b) YAG + 9-CMAS (1:1 by wt)
after heat treatment for 24 h at 1500 °C, and (c) 8YSZ + 9-CMAS (1:1 by wt) after heat The microstructure of “as received” APS 8YSZ sample is shown in
treatment for 24 h at 1500 °C. Fig. 6a which shows three different layers of the coatings. The top
130 R. Kumar et al. / Surface & Coatings Technology 327 (2017) 126–138

Fig. 4. SEM images of mixed powder pellets post sintering at 1500 °C for 24 h with chemical analysis data acquired at the ceramic-CMAS interface. (a) 8YSZ + 9-CMAS (1:1 by wt), (b) EDS
spectrum at 8YSZ − 9-CMAS interface, (c) YAG + 9-CMAS (1:1 by wt) and (d) EDS spectrum at YAG − 9-CMAS interface.

layer, deposited via APS process, is of 8YSZ and has a thickness of precursor is pyrolyzed on the sample which causes shrinkage and for-
~ 250 μm, followed by a 100 μm layer of bondcoat and finally the su- mation of horizontal pores [24] while the over spray which results in
peralloy substrate. Fig. 6b shows the magnified image of APS 8YSZ isolated splats adds additional porosity in that layer.
coating which has a typical microstructure obtained through APS
process with uniformly distributed splat boundaries and porosity 3.4. Thermal cycling without CMAS and failure modes
(~ 15%, calculated using ImageJ software).
Fig. 7a is a cross section SEM image of “in house” deposited SPPS YAG Fig. 8 shows the average thermal cycling lives of SPPS YAG and APS
with a thickness of ~225 μm on superalloy substrates pre-coated with 8YSZ baseline samples with standard deviation indicated through
25 μm of APS 8YSZ and a 100 μm bondcoat layer. The SPPS YAG coating bars. SPPS YAG samples lasted 22% longer than the baseline 8YSZ sam-
has approximately uniform vertical cracks which extend through the ples. To analyze the failure modes, BSE images of the failed coatings
entire coating thickness and offers stress relieving during thermal cy- were taken. Fig. 9a shows the failed baseline sample with spallation at
cling experiments. Fig. 7b shows a magnified SEM image of YAG coat- YSZ-TGO interface. Fig. 9b shows a magnified image of the failure inter-
ings which depicts the aforementioned vertical cracks and also reveals face with elemental mapping. The failure is a mixture of failure in the
horizontal pores that form between the subsequent gun passes by trap- top coat and through the TGO. A distinct TGO layer (10–15 μm) can be
ping of semi-pyrolyzed precursor an overspray. The under pyrolyzed observed. The relatively large TGO thickness enables a failure

Fig. 5. X-ray diffraction of mixed powder pellets after heat treatments at temperatures from 1100 °C to 1500 °C. (a) 8YSZ + 9-CMAS (1:1 by wt) (b) YAG + 9-CMAS (1:1 by wt)
R. Kumar et al. / Surface & Coatings Technology 327 (2017) 126–138 131

Fig. 8. Thermal cyclic performance of SPPS YAG and APS 8YSZ TBCs.

TGO growth while separation at YSZ-YAG interface is most likely due


to differential sintering [24].

3.5. CMAS paste test

CMAS paste tests were done with both 4 component and 9 compo-
nent CMAS and the general failure modes were the same for the two
CMAs compositions however the chemical interactions were different.
Fig. 6. (a) SEM image of as received APS 8YSZ TBC showing topcoat, bondcoat and The lives of both YAG and 8YSZ coatings decreased drastically with
substrate layers. (b) Magnified image of the 8YSZ topcoat showing typical CMAS application but the reduction in life of 8YSZ sample was more
microstructure generated by an APS process with horizontal splats and pores. dramatic. 8 YSZ samples showed a 147 times reduction due to CMAS ex-
posure where SPPS YAG samples showed ~20 times reduction. Fig. 11
mechanism than includes shape change imposed on the ceramic associ- shows the cyclic lives of the two coatings. SPPS YAG coatings performed
ated with TGO growth [7,34]. much better than APS 8YSZ coatings, with a cycling life greater by at
In the case of SPPS YAG the failure occurs in YSZ interlayer-TGO in- least a factor of 8. The mode of failure was quite different in the two
terface and failure at YAG-YSZ interface is also observed as shown in coatings. APS 8YSZ coatings failed in the form of flakes, where layer
Fig. 10a. A magnified image of the failed region is shown in Fig. 10b after layer of the coating came off until no coating remained on the sub-
with the elemental analysis. Failure at the YSZ TGO interface probably strate. On the other hand, SPPS YAG coatings failed in one piece at the
has a contribution to stress in the coating due to shape change due to SPPS YAG to APS YSZ interface (shown in Fig. 12). On further examina-
tion, as shown in Fig. 13, it appears that delamination of flakes happens
in relatively larger horizontal pores in the APS 8YSZ coatings. This is

Fig. 7. SEM image of SPPS YAG TBC showing “in house” deposited YAG topcoat with
vertical cracks, a thin APS 8YSZ inner layer, bondcoat and substrate layers. (b) Magnified
image of the SPPS YAG topcoat showing typical microstructure generated by SPPS Fig. 9. (a) Cross section of a failed APS 8YSZ coating in thermal cycling test showing
process with vertical cracks for strain tolerance and horizontal pores called inter pass spallation at YSZ-TGO interface. (b) Magnified cross section of the spallation region with
boundaries. elemental mapping showing growth of TGO and separation at YSZ-TGO interface.
132 R. Kumar et al. / Surface & Coatings Technology 327 (2017) 126–138

A similar trend in CMAS reactivity is observed in the cases of SPPS YAG


samples, where stronger peaks of calcium yttrium oxide silicate (apa-
tite) as well as anorthite peaks are observed in case of 4-CMAS. Forma-
tion of apatite phase is an important observation as it has been shown to
block the infiltration of CMAS in other studies. However, despite the for-
mation of apatite phase in SPPS YAG, we see infiltration of CMAS
through the YAG layer and reaching the bottom APS YSZ inner layer.
We believe that the microstructure of SPPS YAG has cracks that are
too open and wide that prevents the apatite phase to seal the cracks.
Thus, controlling the crack width of SPPS coatings becomes an impor-
tant parameter for future work. The results for 9-CMAS were consistent
with the powder pellet testing, where YAG did not show any major re-
action with CMAS. Minor apatite peaks were observed in paste test
while no peaks were observed in paste test at low temperature (1100
°C and 1200 °C). In both paste and pellet test (1100 °C–1300 °C) with
9-CMAS anorthite phase was present. CMAS Spritz test.
In gas turbine engines, intake of CMAS most often happens in small
amounts as a continuous process. The paste test is a severe test, which
might not simulate real environments. In an attempt to test samples
in a more realistic environment, the spritz testing was developed (see
experimental Section 2.5. The cyclic lives of the sample in the spritz
test using both types of CMAS are shown in Fig. 17 where lower cyclic
lives were observed as compared to thermal cycling but a significant im-
provement was observed compared to the paste test for both YAG and
Fig. 10. (a) Cross section of a failed SPPS YAG coating in thermal cycling test showing YSZ. The longer lives, as compared to paste test, can be attributed to
spallation at YAG-YSZ interface and YSZ-TGO interface. (b) Magnified cross section of
the spallation region with elemental showing growth of TGO and separation at both
the lower dosages of CMAS applied in the spritz test compared to the
YAG-YSZ and YSZ-TGO interfaces. paste test. If one would calculate the equivalency between the two
tests, 25 doses of CMAS from the ‘spritzer’ is equivalent to the amount
of CMAS applied on the samples before the start of the paste test. Fig.
likely due to a mixture of stress generated as CMAS melt infiltrated in 18 shows the cross-sectional images, depicting the failure mode in the
the pores by capillary flow and solidified on cooling thereby causing samples. The inset has the macro images of the failed coupons. It is in-
loss of strain compliance, and also due to the reaction between 8YSZ teresting to note that the failure modes are consistent with the paste
and CMAS which leads to dissolution and precipitation in the form of test, despite the method and amount of CMAS application being differ-
globules also seen in previous studies [9]. ent. APS YSZ coatings show the exfoliation behavior whereas SPPS YAG
Fig. 14 show the results of X-ray diffraction studies that were per- coatings failed from the APS YSZ inner layer. Figs. 19 and 20 show the
formed on the top surface of failed APS YSZ coatings for both types of failed cross-sections of APS YSZ and SPPS YAG. EDS spectral imaging
CMAS. A reactivity difference between the two CMAS is observed, was done on the APS YSZ coatings which show infiltration of CMAS ele-
where 4-CMAS resulted in stronger peaks of calcium zirconium oxide ments in the failed layer. In case of YAG, CMAS elements were detected
and formation of additional products that can be identified as calcium near the bottom proving the infiltration of CMAS through the vertical
silicate and calcium oxide. Minor anorthite phase is observed in 9- cracks in the YAG layer and reaching the APS YSZ inner layer. In the
CMAS which is consistent with the pellet test. Fig. 15 shows the cross past, similar tests using pure water were run to show that the thermal
section of SPPS YAG coating with both types of CMAS and its failure orig- shock of the application of the precursor liquid resulted in only a small
inating in APS 8YSZ layer. EDS data, in the failed layer, confirms the reduction (7%) of the cyclic life compared to a test with no liquid appli-
presence of CMAS elements (also distinguishable as grey areas sur- cation [34].
rounding the bright ceramic) which suggests that CMAS infiltrated the
YAG coating through vertical cracks and went all the way down to the 4. Discussion
bottom of APS 8YSZ layer. The infiltration process is widely regarded
to be due to capillary flow [35]. Fig. 16 shows the results of X-ray diffrac- 4.1. Optical basicity and CMAS viscosity
tion studies that were performed on the top of failed SPPS YAG coatings.
The concept of Optical Basicity (OB) is reliant on the Lewis concept of
acids and bases. For oxides, the amount of negative charge on the oxy-
gen atoms or ions determines their acidity or basicity. The OB value of
an oxide is representative of this charge, and, hence, is also a compara-
tive indicator of the relative electronegativity and polarizability values
between oxides [36]. This theory has recently been applied to CMAS re-
activity [26,27]. When calculated, OB of CMAS falls on the lower end of
the spectrum rendering them acidic and of the TBC materials are on the
higher end, thereby rendering them basic. The theory predicts greater
reactivity if the difference in OB values of CMAS and TBC materials are
higher. To test the theory, OB of the two TBC materials and the two
CMAS compositions were calculated according to Eq. 1 and are shown
in Table 4. It is observed that the differences in the values of YAG and
CMAS are lower than that of 8YSZ and CMAS thus predicting lesser reac-
tivity between YAG and CMAS. With the experimental evidence, it can
be confirmed that YAG in all the previous tests shows little to negligible
Fig. 11. Cyclic lives of APS YSZ baselines and SPPS YAG TBCs during paste test. reaction compared to YSZ in 4 and 9-CMAS respectively. Also, the
R. Kumar et al. / Surface & Coatings Technology 327 (2017) 126–138 133

Fig. 12. TBCs, (a), (b) SEM images showing regions of failure in APS YSZ samples tested with 4 and 9-CMAS respectively. Inset showing macro pictures of failed samples. (c), (d) SEM images
showing regions of failure in SPPS YAG samples with 4 and 9-CMAS respectively. Inset showing macro pictures of failed samples.

difference in the OB of the ceramics and 9-CMAS is less than that of 4- As discussed in Section 3.1 CMAS viscosity was calculated using
CMAS, which is also consistent with the greater reaction with the 4- existing models and it was shown that 4-CMAS has a higher viscosity
CMAS as seen Figs. 14 and 16. 4-CMAS reacts with both YSZ and YAG than the 9-CMAS with nearly an order of magnitude difference. Thus,
forming different reaction products. This is not so visible in the 9- we would expect 9-CMAS to infiltrate the coatings faster and thereby
CMAS, which in fact, does not significantly react with YAG. Thus, OB the- leading to shorter cyclic lives in paste and spritz test. Contrary to the hy-
ory predicts all the reactivity trends in the current experiments and can pothesis, we don't see this trend in the cyclic lives of samples in either of
be a very useful tool in choosing TBC materials for different strategies of the aforementioned tests. Only SPPS YAG samples in spritz test show
enhancing CMAS resistance, either by aiming for no reaction or by vig- marginally shorter life with 9-CMAS.
orous reactions forming crack arresting reaction products like in the
case of Gadolinium Zirconate.
Formula for calculation of optical basicity: 4.2. Differences in failure modes and lives

Optical Basicity ¼ X 1  OB1 þ X 2  OB2 þ X 3  OB3 þ … ð1Þ APS 8YSZ samples show distinctive failure characteristics in the form
of flakes. We propose that this can be attributed to rapid infiltration of

Fig. 13. (a), (d) Backscattered electron images showing cross-section of failed APS YSZ coatings with 4, 9-CMAS infiltration respectively (b), (e) showing the regions of attack of 4, 9-CMAS
respectively on YSZ and (c), (f) showing EDS spectra of 4, 9-CMAS infiltrated APS YSZ coatings respectively.
134 R. Kumar et al. / Surface & Coatings Technology 327 (2017) 126–138

Fig. 14. XRD pattern from failed APS YSZ baselines during CMAS paste test showing phase changes in 8YSZ due to its reaction with (a) 4-CMAS (b) 9-CMAS.

Fig. 15. (a), (b) Backscattered electron images showing cross-sections of failed SPPS YAG coatings with 4, 9-CMAS infiltration. (b) Backscattered electron images showing cross-sections of
failed SPPS YAG coatings with 9-CMAS infiltration (c EDS spectra of APS YSZ inner layer confirming the presence of 4, 9-CMAS. (d) EDS spectra of APS YSZ inner layer confirming the
presence of 9-CMAS.
R. Kumar et al. / Surface & Coatings Technology 327 (2017) 126–138 135

Fig. 16. XRD pattern from failed SPPS YAG samples, during CMAS paste test, showing phase changes in YAG because of its reaction with (a) 4-CMAS (b) 9-CMAS.

CMAS melt in the horizontal pores in APS coatings generated between prevents accumulation of CMAS in the YAG layer, however once CMAS
the splats. The infiltrated CMAS in a molten state, causes dissolution of comes in contact with APS 8YSZ inner layer it causes failure for the rea-
8YSZ, in which yttrium ions were leached out of the YSZ matrix sons described earlier. Another question that arises is the differences in
resulting in detrimental phase change from metastable tetragonal to cyclic life of the coatings. APS 8YSZ coatings offer no resistance to CMAS
monoclinic. Also, solidification of the CMAS in the horizontal pores is infiltration with readily available horizontal pores. Failure lives in SPPS
likely to result in stress generation and loss of strain compliance. YAG YAG coatings is primarily governed by the infiltration of CMAS into
samples show failure at the interface in the 8YSZ inner layer when the YAG layer with minimal chemical interaction and then its contact
CMAS fully infiltrates the top YAG layer. Vertical cracks in SPPS YAG with the YSZ inner layer. The infiltration kinetics of CMAS can be safely
coatings serve as channels for CMAS infiltration. While YAG being less assumed to depend on the vertical crack geometry which includes
reactive or inert to CMAS (depending on the composition) and does width of the cracks and its branching to form narrower channels. Fur-
not undergo detrimental phase changes. The lack of horizontal pores ther in-depth quantification of SPPS coating microstructure and its ef-
fect on CMAS infiltration kinetics will be conducted.

5. Conclusions

In previous work [23] YAG TBCs were shown to have very good cy-
clic furnace durability and erosion resistance. In this study CMAS resis-
tance was investigated and YAG powder shows limited to almost no
reactivity with CMAS. 8YSZ shows dissolution and re-precipitation in
the presence of CMAS accompanied by the undesirable phase changes
from metastable tetragonal to monoclinic and cubic. Post sintering con-
ditions of the pellets emphasized this behavior on a macroscopic scale,
where YAG pellets only changed its shape due to localized melting,
while YSZ pellets melted and formed a puddle of the reaction products.
In the tests where CMAS was applied on TBCs, the cyclic lives of YAG
coatings were at least 8 times longer in the paste test and 2 times longer
in the spritz test compared to the YSZ baselines and the failure occurred
only when CMAS melt, infiltrated through the coating and reached the
Fig. 17. Cyclic lives of YSZ baseline and SPPS YAG samples during Spritz test. APS YSZ inner layer. Failure in APS YSZ samples occurred in one cycle
136 R. Kumar et al. / Surface & Coatings Technology 327 (2017) 126–138

Fig. 18. (a), (b) SEM images showing regions of failure in APS YSZ samples tested with 4 and 9-CMAS respectively. Inset showing macro pictures of failed samples. (c), (d) SEM images
showing regions of failure in SPPS YAG samples with 4 and 9-CMAS respectively. Inset showing macro pictures of failed samples.

in the paste due to infiltration of CMAS in the coatings which acted both TBCs showed superior durability in the Spritz test compared to the paste
chemically and mechanically to disrupt the structure. Similar trends test, by a factor of 2–3, and the total number of cycles was much greater.
were observed in Spritz test, which was conducted to simulate realistic The concept of optical basicity theory (OB) was employed to successful-
steady ingestion of CMAS in gas turbines. Longer lives were experienced ly explain the differences in reactivity between two TBC materials (YAG
in the Spritz test resulting from the lower doses of CMAS. The SPPS YAG and YSZ) and two CMAS compositions (4-CMAS and 9-CMAS). These

Fig. 19. (a), (c) SEM images showing cross-section of failed APS YSZ Baseline coatings with 4, 9-CMAS infiltration respectively (b), (d) showing elemental mapping of 4, 9-CMAS infiltrated
APS YSZ baselines respectively.
R. Kumar et al. / Surface & Coatings Technology 327 (2017) 126–138 137

Fig. 20. (a), (b) Backscattered electron images showing cross-sections of failed SPPS YAG coatings with 4, 9-CMAS infiltration respectively (c), (d) show EDS spectra of SPPS YAG coatings
near APS YSZ inner layer confirming the presence of 4, 9-CMAS respectively.

Table 4 [4] C. Mercer, J.R. Williams, D.R. Clarke, A.G. Evans, On a ferroelastic mechanism
Calculated values of optical basicity of 8YSZ, YAG and 4,9-CMAS. Differences between the governing the toughness of metastable tetragonal-prime (t′) yttria-stabilized zirco-
nia, Proc. R. Soc. London A Math. Phys. Eng. Sci. (2007) 1393–1408, http://dx.doi.
optical basicity of ceramics and CMAS have been calculated to predict chemical reactivity
org/10.1098/rspa.2007.1829.
between them.
[5] R. Vassen, A. Stuke, D. Stöver, Recent developments in the field of thermal barrier
TBC materials Optical basicity 4-CMAS Difference 9-CMAS Difference coatings, J. Therm. Spray Technol. 18 (2009) 181–186, http://dx.doi.org/10.1007/
s11666-009-9312-7.
YAG 0.70 0.63 0.07 0.75 0.05 [6] R.A. Miller, J.L. Smialek, R.G. Garlick, Phase stability in plasma sprayed partially
8YSZ 0.87 0.24 0.12 stabilized zirconia-yttria, Adv. Ceram, Sci. Technol. Zirconia, Vol. 3, 1981,
pp. 242–253.
[7] A.G. Evans, D.R. Mumm, J.W. Hutchinson, G.H. Meier, F.S. Pettit, Mechanisms con-
trolling the durability of thermal barrier coatings, Prog. Mater. Sci. 46 (2001)
results provide support to the usefulness of using the OB theory to pre- 505–553, http://dx.doi.org/10.1016/S0079-6425(00)00020-7.
[8] C. Mercer, S. Faulhaber, A.G. Evans, R. Darolia, A delamination mechanism for ther-
dict the CMAS resistance of TBCs. Lastly, viscosity of the two CMAS was
mal barrier coatings subject to calcium–magnesium–alumino-silicate (CMAS) infil-
calculated using existing models which predicted 4-CMAS to have an tration, Acta Mater. 53 (2005) 1029–1039, http://dx.doi.org/10.1016/j.actamat.
order of magnitude higher viscosity than 9-CMAS but no trend was ob- 2004.11.028.
[9] S. Krämer, S. Faulhaber, M. Chambers, D.R. Clarke, C.G. Levi, J.W. Hutchinson, A.G.
served in TBCs cyclic lives in paste and spritz tests consistent with the
Evans, Mechanisms of cracking and delamination within thick thermal barrier sys-
higher viscosity CMAS being less damaging. tems in aero-engines subject to calcium-magnesium-alumino-silicate (CMAS) pen-
etration, Mater. Sci. Eng. A 490 (2008) 26–35, http://dx.doi.org/10.1016/j.msea.
2008.01.006.
Acknowledgements [10] 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.
This work was supported by the U.S. Department of Energy [STTR Ceram. Soc. 89 (2006) 3167–3175, http://dx.doi.org/10.1111/j.1551-2916.2006.
01209.x.
Grant # DE-SC0007544]. The authors would like to thank Dr. Patcharin [11] A. Aygun, A.L. Vasiliev, N.P. Padture, X. Ma, Novel thermal barrier coatings that are
Burke, DOE Program Manager for STTR Phase I, II & IIA Programs on SPPS resistant to high-temperature attack by glassy deposits, Acta Mater. 55 (2007)
YAG TBCs. We are also thankful to our industrial partners for DOE STTR 6734–6745, http://dx.doi.org/10.1016/j.actamat.2007.08.028.
[12] M.P. Borom, C.A. Johnson, L.A. Peluso, Role of environment deposits and operating
Phase IIA namely, Solar Turbines, Siemens Energy, Pratt & Whitney, surface temperature in spallation of air plasma sprayed thermal barrier coatings,
Rolls-Royce, Praxair, Progressive Surface, Metco-Oerlikon, Chromalloy Surf. Coat. Technol. 86 (1996) 116–126, http://dx.doi.org/10.1016/S0257-
and GE Power. 8972(96)02994-5.
[13] L. Li, N. Hitchman, J. Knapp, Failure of thermal barrier coatings subjected to CMAS
attack, J. Therm. Spray Technol. 19 (2010) 148–155, http://dx.doi.org/10.1007/
References s11666-009-9356-8.
[14] R. Vaßen, M.O. Jarligo, T. Steinke, D.E. Mack, D. Stöver, Overview on advanced ther-
[1] N.P. Padture, M. Gell, E.H. Jordan, Thermal barrier coatings for gas-turbine engine mal barrier coatings, Surf. Coat. Technol. 205 (2010) 938–942, http://dx.doi.org/10.
applications, Science 296 (2002) 280–284 80-. http://dx.doi.org/10.1126/science. 1016/j.surfcoat.2010.08.151.
1068609. [15] G. Mauer, M.O. Jarligo, D.E. Mack, R. Vaßen, Plasma-sprayed thermal barrier coat-
[2] M.J. Stiger, N.M. Yanar, M.G. Topping, F.S. Pettit, G.H. Meier, Thermal barrier coatings ings: new materials, processing issues, and solutions, J. Therm. Spray Technol. 22
for the 21st century, Zeitschrift Fur Met. 90 (1999) 1069–1078. (2013) 646–658, http://dx.doi.org/10.1007/s11666-013-9889-8.
[3] R.A. Miller, Thermal barrier coatings for aircraft engines: history and directions, J. [16] N.P. Padture, P.G. Klemens, Low thermal conductivity in garnets, J. Am. Ceram. Soc.
Therm. Spray Technol. 6 (1995) 35–42, http://dx.doi.org/10.1007/BF02646310. 80 (1997) 1018–1020, http://dx.doi.org/10.1111/j.1151-2916.1997.tb02937.x.
138 R. Kumar et al. / Surface & Coatings Technology 327 (2017) 126–138

[17] D.R. Clarke, S.R. Phillpot, Thermal barrier coating materials, Mater. Today 8 (2005) part I, optical basicity considerations and processing, J. Am. Ceram. Soc. 97 (2014)
22–29, http://dx.doi.org/10.1016/S1369-7021(05)70934-2. 3943–3949, http://dx.doi.org/10.1111/jace.13210.
[18] E. Bakan, D.E. Mack, G. Mauer, R. Vaßen, Gadolinium zirconate/YSZ thermal barrier [28] J.M. Drexler, A.L. Ortiz, N.P. Padture, Composition effects of thermal barrier coating
coatings: plasma spraying, microstructure, and thermal cycling behavior, J. Am. ceramics on their interaction with molten Ca–Mg–Al–silicate (CMAS) glass, Acta
Ceram. Soc. 97 (2014) 4045–4051, http://dx.doi.org/10.1111/jace.13204. Mater. 60 (2012) 5437–5447, http://dx.doi.org/10.1016/j.actamat.2012.06.053.
[19] S. Lakiza, O. Fabrichnaya, C. Wang, M. Zinkevich, F. Aldinger, Phase diagram of the [29] D. Giordano, J.K. Russell, D.B. Dingwell, Viscosity of Magmatic Liquids: a Model,
ZrO2–Gd2O3–Al2O3 system, J. Eur. Ceram. Soc. 26 (2006) 233–246, http://dx.doi. 2008http://dx.doi.org/10.1016/j.epsl.2008.03.038.
org/10.1016/j.jeurceramsoc.2004.11.011. [30] A. Fluegel, Glass viscosity calculation based on a global statistical modelling ap-
[20] S. Krämer, J. Yang, C.G. Levi, Infiltration-inhibiting reaction of gadolinium zirconate proach, Glass Technol. Eur. J. Glass Sci. Technol. Part A 48 (2007) 13–30 http://
thermal barrier coatings with CMAS melts, J. Am. Ceram. Soc. 91 (2008) 576–583, www.ingentaconnect.com/content/sgt/gta/2007/00000048/00000001/art00003
http://dx.doi.org/10.1111/j.1551-2916.2007.02175.x. (accessed July 7, 2017).
[21] R.M. Leckie, S. Krämer, M. Rühle, C.G. Levi, Thermochemical compatibility between [31] T. Steinke, D. Sebold, D.E. Mack, R. Vaßen, D. Stöver, A novel test approach for plas-
alumina and ZrO2–GdO3/2 thermal barrier coatings, Acta Mater. 53 (2005) ma-sprayed coatings tested simultaneously under CMAS and thermal gradient cy-
3281–3292, http://dx.doi.org/10.1016/j.actamat.2005.03.035. cling conditions, Surf. Coat. Technol. 205 (2010) 2287–2295, http://dx.doi.org/10.
[22] G. Dwivedi, V. Viswanathan, S. Sampath, A. Shyam, E. Lara-Curzio, Fracture tough- 1016/j.surfcoat.2010.09.008.
ness of plasma-sprayed thermal barrier ceramics: influence of processing, micro- [32] R. Wellman, G. Whitman, J.R. Nicholls, CMAS corrosion of EB PVD TBCs: identifying
structure, and thermal aging, J. Am. Ceram. Soc. 97 (2014) 2736–2744, http://dx. the minimum level to initiate damage, Int. J. Refract. Met. Hard Mater. 28 (2010)
doi.org/10.1111/jace.13021. 124–132, http://dx.doi.org/10.1016/j.ijrmhm.2009.07.005.
[23] E.H. Jordan, M. Gell, C. Jiang, J. Wang, B. Nair, High temperature thermal barrier coat- [33] V.L. Wiesner, U.K. Vempati, N.P. Bansal, High temperature viscosity of calcium-mag-
ing made by the solution precursor plasma spray process, Proc. ASME Turbo Expo nesium-aluminosilicate glass from synthetic sand, Scr. Mater. 124 (2016) 189–192,
Turbine Tech. Conf. Expo, 6, 2014. http://dx.doi.org/10.1016/j.scriptamat.2016.07.020.
[24] M. Gell, J. Wang, R. Kumar, J. Roth, E.H. Jordan, Higher temperature thermal barrier [34] C. Jiang, E.H. Jordan, A.B. Harris, M. Gell, J. Roth, Double-layer gadolinium zirconate/
coatings with the combined use of yttrium aluminum garnet and the solution pre- yttria-stabilized zirconia thermal barrier coatings deposited by the solution precur-
cursor plasma spray process. (Unpublished results- J. Therm. Spray Technol. (n.d.). sor plasma spray process, J. Therm. Spray Technol. 24 (2015) 895–906, http://dx.doi.
[25] Y.J. Su, R.W. Trice, K.T. Faber, H. Wang, W.D. Porter, Thermal conductivity, phase sta- org/10.1007/s11666-015-0283-6.
bility, and oxidation resistance of Y3Al5O12 (YAG)/Y2O3—ZrO2 (YSZ) thermal-barri- [35] R. Naraparaju, M. Hüttermann, U. Schulz, P. Mechnich, Tailoring the EB-PVD colum-
er coatings, Oxid. Met. 61 (2004) 253–271, http://dx.doi.org/10.1023/B:OXID. nar microstructure to mitigate the infiltration of CMAS in 7YSZ thermal barrier coat-
0000025334.02788.d3. ings, J. Eur. Ceram. Soc. 37 (2017) 261–270, http://dx.doi.org/10.1016/j.
[26] A.R. Krause, H.F. Garces, B.S. Senturk, N.P. Padture, 2ZrO2·Y2O3 thermal barrier jeurceramsoc.2016.07.027.
coatings resistant to degradation by molten CMAS: part II, interactions with sand [36] J.A. Duffy, Optical basicity: a practical acid-base theory for oxides and oxyanions, J.
and fly ash, J. Am. Ceram. Soc. 97 (2014) 3950–3957, http://dx.doi.org/10.1111/ Chem. Educ. 73 (1996) 1138, http://dx.doi.org/10.1021/ed073p1138.
jace.13209.
[27] A.R. Krause, B.S. Senturk, H.F. Garces, G. Dwivedi, A.L. Ortiz, S. Sampath, N.P. Padture,
2ZrO2·Y2O3 thermal barrier coatings resistant to degradation by molten CMAS:

You might also like