You are on page 1of 8

Surface & Coatings Technology 203 (2009) 23072314

Contents lists available at ScienceDirect

Surface & Coatings Technology


j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / s u r f c o a t

Mechanical properties of thermal barrier coatings after isothermal oxidation.


Depth sensing indentation analysis
A. Rico, J. Gmez-Garca, C.J. Mnez, P. Poza , V. Utrilla
Departamento de Ciencia e Ingeniera de los Materiales, Universidad Rey Juan Carlos, Escuela Superior de Ciencias Experimentales y Tecnologa, C/Tulipn s.n. 28933 Mstoles, Madrid, Spain

a r t i c l e

i n f o

Article history:
Received 24 September 2008
Accepted in revised form 22 February 2009
Available online 1 March 2009
Keywords:
Thermal barrier coatings (TBC)
Mechanical properties
Isothermal oxidation
Sensing indentation (nanoindentation)
Scanning electron microscopy (SEM)

a b s t r a c t
Thermal barrier coatings (TBC) are extensively used to protect metallic components in applications where
the operating conditions include aggressive environment at high temperatures. Isothermal oxidation
degrades the performance of these coatings, so this work analyses the mechanical properties (Young's
modulus, E, and hardness, H) of TBC and its evolution after thermal exposure in air. ZrO2(Y2O3) top coat and
NiCrAlY bond coating were air plasma sprayed onto an Inconel 600 Ni base alloy. The TBC were isothermally
oxidized in air at 950 C and 1050 C for 72, 144 and 336 h. Depth sensing indentation tests were carried out
on the ceramic coating to evaluate E and H in the as-sprayed materials and after isothermal oxidation. An
approach based on multiple tests at different loads was used to determine size independent apparent E an H.
These mechanical properties, measured perpendicular to the surface, clearly decreased after isothermal
oxidation as a consequence of microcracking within the ceramic coating.
2009 Elsevier B.V. All rights reserved.

1. Introduction
Thermal barrier coatings (TBC) are widely used in turbine engines for
propulsion and power generation where materials capable of withstanding increasing operating temperatures, mechanical loads and
chemical degradation are required. The aim of these coatings is to
insulate the metallic components from the aggressive environment and
to provide the opportunity of increasing turbine entry gas temperature,
which promotes overall engine efciency. TBC comprise at least two
layers: a ceramic top coating and a metallic bond coat [17].
ZrO2 coatings, in which the high temperature cubic phase is
partially stabilised relative to the lower temperature monoclinic phase
by the addition of typically 8% Y2O3, are widely used in TBC. Partially
stabilised zirconia (PSZ) has the required low thermal conductivity
(0.81.5 W m 1 K 1) together with a relatively high coefcient of
thermal expansion (CTE) of 71010 6 K 1 [8]. This value, although
higher than the CTEs usually found in ceramic materials, is smaller than
the typical CTE for the load bearing metals (approximately 1510 6 K 1
for Ni based alloys or superalloys). Consequently, the direct use of a PSZ
layer onto a metallic component is problematic because the CTE mismatch
between the ceramic coating and the metallic substrate generates high
interfacial shear stresses during coating manufacture and in-service
thermal cycling that leads ceramic failure by spalling. To reduce these
problems arising from CTE mismatch and to protect the metallic
component from chemical attack (ZrO2 is permeable to oxygen and

Corresponding author. Tel.: +34 914 887 179; fax: +34 914 888 150.
E-mail address: pedro.poza@urjc.es (P. Poza).
0257-8972/$ see front matter 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.surfcoat.2009.02.035

other aggressive species at high temperature) an intermediate metallic


bond coat is used. To improve the oxidation and corrosion resistance of
the protected component, these overlay coatings should develop a surface
oxide layer thermodynamically stable, slowly growing and adherent.
MCrAlY type (M=Ni, Co or Ni+Co) coatings are currently used in
systems processed by thermal spraying [1,57]. During service at high
temperatures a thermally grown oxide (TGO) layer is formed between the
MCrAlY bond coat and the ceramic top coat. The main constituent of this
TGO layer should be a protective -Al2O3 scale that inhibits continued
ingress of active oxygen and other species. However, the development of
this alumina scale reduces the Al content in the bond coating and other
oxides, like (Ni, Co)Al2O4 spinels, could be formed. The formation of
spinels may compromise TBC durability because of their relatively fast
growth rate and associated volume changes [1,6,9,10].
TBC are exposed to high temperature during service, by isothermal
oxidation or thermal cycling. Several processes, including crack
nucleation and propagation, take place within the system until failure
by spallation of the top coat. Failure mechanisms depend on the specic
TBC and the operating conditions. Coatings used in turbines for
propulsion, like aeronautical engines, are mainly exposed to thermal
fatigue as they operate at full capacity during relatively short periods of
time resulting in numerous thermal cycles. In contrast, power generation engines often operate for long periods of time with only a few shutdown cycles and they are mainly degraded by isothermal oxidation.
Despite these complexities some generic understanding of TBC failure
mechanisms is emerging [5,10,11]. TGO growth is the most important
phenomenon controlling TBC durability. This layer is under compression
because TGO growth results in a constrained volume expansion,
depending on the oxides formed during service. In addition, thermal
expansion mismatch between TGO and bond coat leads to very high

2308

A. Rico et al. / Surface & Coatings Technology 203 (2009) 23072314

tested at the sub-micron scale. However, this approach estimates the


sample-indenter contact area during the test from the theoretical
indenter geometry and any deviation in the real geometry leads to
deviations in the estimated mechanical properties. In addition, the
indentation size could also affect the hardness and modulus measured
by effects intrinsic to the material tested or related to the load
displacement curve analysis. As a consequence, the scattering
observed in the mechanical properties of TBC, measured by sensing
indentation, could be important. For these reasons, the aim of this
paper is to evaluate the mechanical properties of the top coating, used
in plasma sprayed TBC, and its evolution after isothermal oxidation by
sensing indentation using an alternative approach, which has been
previously used for bulk ceramics [2629] and, recently, extended to
Al2O3 coatings [30,31] although, to the authors' knowledge, it has not
been implemented to analyse thermal barrier coatings. This method
estimates the coating stiffness and hardness from several tests
performed at different loads. This analysis about the evolution of the
top coating mechanical properties, during isothermal oxidation, could
be used in life prediction models, based on the nite element method
[11,15,3234], to extend the actual knowledge about the failure
mechanisms in TBC.
2. Materials and experimental techniques

Fig. 1. Indentations location. a) Schematic of the top-coat side indented. b) Indentation


carried out at 500 mN placed inside a Y-PSZ splat. It should be noticed the absence of
cracks at the indent tip.

compressive residual stresses in the TGO upon cooling [5,1013]. Finally,


roughening and undulations at the interface, intrinsic to thermal spray
processing, promotes out of plane stresses normal to the bond coattop
coat interface. These stresses are tensile at the undulation crests in the
TGO-bond coat and TGO-top coat interfaces being primarily responsible
of TBC failure [5,1012,14,15]. Tensile stresses increases as the TGO
thickens causing fracture at the interfaces and cracking within the brittle
ceramic coating. These failure mechanisms are inuenced by the
stiffness and hardness of the TBC constituents (mainly the top coating)
and their evolution during exposure at high temperature.
The ceramic coating mechanical properties have been measured by
different methods like bending tests [16,17] and indentation tests,
including conventional Vickers microindentation [18] and sensing
indentation [16,1923]. Apparent modulus, E, and hardness, H, of
detached top coatings increased as they were heat treated due to
resintering at high temperature [16,23]. However, an imposed tensile
stress retards sintering and the stiffness of the ceramic layer in TBC
exposed to thermal fatigue decreased signicantly as the number of
cycles increased [17]. Sensing indentation has been extensively used
for the mechanical characterization of coatings because of its ability to
probe the mechanical properties of materials at very small scale
[24,25]. This method uses high resolution sensors and actuators to
continuously monitor the loads and displacements on an indenter as it
is driven into the material studied. Usually, the loaddisplacement
curve is analysed using the method proposed by Oliver and Pharr in
1992 [24,25] to obtain hardness and elastic modulus of the material

The TBC were air plasma sprayed (APS) onto a nickel-base alloy,
Inconel 600 (Ni 16Cr 8Fe, wt.%). AMDRY 962 (Ni 22Cr 10Al 1 Y, wt.%)
and METCO 204, (ZrO2 8Y2O3, wt.%) powders were used for the bond
coat and for the top coating, respectively. The sprayed samples were
supplied by REMESA (Recargues y Mecanizados, S.A.; Gijn-Spain).
Some of the coated materials were oxidised in air at 950 C and
1050 C for 72, 144, and 336 h, in a furnace (AGNI FHT 4/165).
Metallographic samples from the as-sprayed and oxidized coatings
were prepared in the longitudinal section, parallel to the spraying
direction. The samples were cut with a diamond saw and grinded using
1200 grit SiC paper. This was followed by polishing in a diamond slurry
up to a 1 m nish. Care was taken to reduce polishing times and
pressures to minimize enlarging any surface porosity or the excessive
removal any poorly bonded regions of the coating (commonly referred
to as pull-out). Polished surfaces were cleaned in deionized water,
and afterward by ultrasound in acetone and propanol. The microstructure was observed in a LEICA DMR optical microscope and it was
also analysed by electron microscopy. A JEOL JSM-6300 scanning
electron microscope (SEM), and a Philips XL30, environmental
scanning electron microscope (ESEM) were used. Both were equipped
with energy dispersive X-ray microanalysis (EDX). Secondary (SEI)

Fig. 2. BEI of a typical cross section of the as-spayed TBC system.

A. Rico et al. / Surface & Coatings Technology 203 (2009) 23072314

2309

and backscattered electron images (BEI), were obtained. Au/Pd ultra


thin lms were deposited to the samples analysed by SEM and water
vapour was used (partial pressure 0.6 Torr) in ESEM observations to
minimize charging effects.
Cracking within the ceramic coating was quantied in the assprayed sample and after oxidation at 950 C and 1050 C using an
image analysis system (Image Pro Plus) in two zones: close to the bond
coatceramic layer interface (zone 1) and at the middle of the top coat
(zone 2). Cracks were dened as pores with an aspect ratio equal or
higher than three.
The top surface of several as-sprayed and oxidized samples was
polished following the procedure described previously, to analyse
the mechanical properties of the ceramic coating. Depth sensing
indentation tests were carried out with a diamond Berkovich
indenter, with a nominal edge radius of 100 nm, using a
Nanoindenter XP (MTS System Co.). Load was applied via a
calibrated electromagnetic coil with a resolution of 50 nN. The
displacement of the indenter was measured using a capacitive
transducer with a resolution of 0.01 nm. Ten series of indentation
tests and ten tests per each one were carried out at different
maximum loads ranging from 1 mN to 500 mN for the as sprayed
and oxidized samples. Fig. 1a shows schematically the indentation
location. As it is shown, imprints were located on the plan view
section. Indentation size was clearly smaller than the Y-PSZ splats,

Fig. 4. a) BEI showing oxidation within the NiCrAlY and at the bond coattop coat interface
in the TBC oxidized at 950 C for 144 h. b) BEI showing the TGO at the bond coattop coat
interface in the TBC oxidized at 950 C for 72 h. c) EDX prole corresponding to the TGO. Al,
Cr and O peaks are clearly shown suggesting Al2O3 and Cr2O3 formation.

even for those corresponding to the maximum load used (Fig. 1b),
and the indentations were placed inside the splats. Young's modulus
and hardness could be estimated from the loaddisplacement curve
using the Oliver and Pharr method [24,25] for a Berkovich indenter
according to the following expressions:
Fig. 3. a) BEI of the as-sprayed NiCrAlY bond coat. Porosity and dark oxidized areas are
marked. b) EDX prole corresponding to the oxidized areas with a grey contrast. Al, Cr
and O peaks are clearly shown suggesting Al2O3 and Cr2O3 formation.

H=

Pmax
A

2310

A. Rico et al. / Surface & Coatings Technology 203 (2009) 23072314

Table 1
TGO thickness after isothermal oxidation.
Exposure time (h)

Thickness @ 950 C (m)

Thickness @ 1050 C (m)

72
144
336

2.3
2.5
2.8

2.7
2.8
3.0

where, Pmax, is the maximum load and A the projected area of contact
at maximum load. On the other hand,
p
1 1
E = S p
2 A

where, S, is the slope of the loaddepth of penetration curve at the


beginning of the elastic unloading (the penetration depth, h, starts to
decrease from its maximum value, ht)
S=

 
dP
dh h = ht

, is a constant dependent on the indenter geometry (=1.034 for a


Berkovich indenter), and, E, the reduced or effective modulus determined by:

diamond indenter and was considered 0.28 for the ceramic coating. In
the preceding expressions, the projected contact area, A, is evaluated
through the indenter shape function, A=A(hp), a relationship between A
and the contact depth, hp (A=24.5 h2p for an ideal Berkovich tip). Finally,
this last magnitude can be calculated from the expression:
hp = ht e

Pmax
S

where, , is again a constant dependent on the indenter geometry


(e = 2d 2 for a Berkovich indenter).
This methodology provides E and H from the loaddisplacement
curve. However, several effects could take place leading to deviations
in the measured mechanical properties, as it was explained
previously and will be evidenced in the following section. One
possibility to solve these problems is to consider Eq. (1) as a linear
relationship between Pmax and A with the hardness as the slope.
Similarly, Eq. (2) represents a linear relationship between S and A1/2
(or hp) where the slope is proportional to the reduced modulus, E.
These relationships will provide an approach to estimate E and H
from several tests carried out at different loads as it has been
proposed previously [30,31].

1
1 2
1 V
+
=

E
EV
E

where E, , E and represent Young's modulus and Poisson ratio of the


material and the indenter, respectively. E=1141 GPa and =0.07 for a

Fig. 5. a) SEI of the TGO at the bond coattop coat interface in the TBC oxidized at 1050 C
for 336 h. Top coat spallation and cracking within the ceramic coating are shown. b) SEI
showing microcracking within the top coat in the TBC oxidized at 950 C for 144 h.

Fig. 6. Quantication of cracking within the top coating, close the bond coat (zone 1)
and at the middle of the ceramic coating (zone 2). a) Optical image treated for cracking
measurement. b) Crack density as a function of exposure time.

A. Rico et al. / Surface & Coatings Technology 203 (2009) 23072314

2311

3. Results
3.1. Microstructure
The TBC system (Fig. 2) comprised a 150 m thick NiCrAlY bond
coat and a 300 m thick Y-PSZ top coating. The as-sprayed
microstructure of both coatings was inhomogeneous containing thin
and large splats associated with the deposition of individual molten
droplets, with porosity between splats, and un-melted particles
(Fig. 3a). The bond coat microstructure was mainly formed by Ni
splats which contained Cr, Al, and Y. Dark areas reach in O with Al, Cr
and some Y were observed next to splats suggesting the formation of
Al2O3, Cr2O3 and possibly Y2O3 during spraying (Fig. 3b).
During thermal exposure the bond coat was oxidized and those dark
areas observed between the splats were increased, indicating oxygen
penetrated through the interconnected porosity up to the bond coat
(Fig. 4a). As it was expected, a TGO layer (Fig. 4b), mainly formed by Al
and Cr according to EDX analysis (Fig. 4c), grew between the NiCrAlY
bond coat and the ceramic coating. Fig. 4b shows a detail of the TGO
formed in the TBC treated at 950 C for 72 h. TGO thickness was 2.3 m
and it increased up to 2.8 m after oxidation for 336 h at 950 C. As
expected, TGO thickness, after the same oxidation time, was higher at
1050 C than at 950 C as it is shown in Table 1. As a consequence of TGO

Fig. 8. Mechanical properties vs. total penetration depth for the as sprayed and oxidized
samples. a) Young's modulus. b) Hardness.

forming, and subsequent thickening, out of plane tensile stresses were


developed in the interface promoting fracture at the top coat-TGO
interface and cracking within the ceramic layer in the vicinity of the
interface (Fig. 5a). These cracks propagated through the top coat
increasing the discontinuities in the ceramic, which will affect the
mechanical performance of this layer (Fig. 5b).
Fig. 6a shows a treated image, corresponding to the coating exposed
at 1050 C for 72 h, showing the areas where the cracks were measured
and Fig. 6b presents cracking evolution after oxidation. Crack density in
the vicinity of the interface increased quickly with the exposure time up
to ~20%. This value was reached after 144 h at 1050C and after 336 h at
950 C. At this point delaminations were observed in the TGO-top
coating interface (Fig. 5). Cracking increased lineally at the middle of the
ceramic layer as the exposure time was extended. Cracks propagated
more rapidly at 1050 C than at 950 C, so the differences in cracks
density were higher as the exposure time was increased.
3.2. Mechanical properties

Fig. 7. Loaddisplacement curves. a) As sprayed coating. b) TBC oxidized at 1050 C for 366 h.

Sensing indentation tests were performed on the ceramic coating


to estimate hardness and Young's modulus and its evolution after
isothermal oxidation. Fig. 7 shows several examples of typical load

2312

A. Rico et al. / Surface & Coatings Technology 203 (2009) 23072314

displacement curves measured on the as-sprayed (Fig. 7a) and heat


treated at 1050 C for 336 h (Fig. 7b) samples. It is remarkable that no
uncontrolled phenomena, like failure, cracking or pop-in, were
detected during indentation tests. The curves presented in Fig. 7
exhibit a behaviour characteristic from a sprayed coating, formed by
splats, and it is possible to distinguish a residual penetration depth
after the sample was unloaded.
Young's modulus and hardness could be estimated using directly
the Oliver and Pharr approach and they are plotted versus the
maximum penetration depth reached during the indentation tests in
Fig. 8. Size effect is observed, i. e. the mechanical properties increase as
the total depth decreases. This phenomenon is inuenced by a great
number of factors affecting the sensing indentation data: surface
roughness, geometry deviations, indenter deformation due to the high
stiffness of the sample, etc. In addition, Fig. 8 points out a general
trend for the mechanical properties evolution during thermal
treatment. Hardness and Young's modulus decreased as the severity
of the oxidation was increased for all penetration depth considered.
Another methodology, developed to obtain size independent
properties on stiff materials from several tests performed at different
maximum loads, was carried out. A linear relationship between the

Table 2
Linear ts obtained from Fig. 9a.
S = M + N hp
T (C)

t (h)

950

M (mN nm 1)

N (mN nm 2)

0.096
0.111
0.087
0.076
0.082
0.090
0.117

8.81 10 4
8.40 10 4
7.74 10 4
8.63 10 4
6.96 10 4
5.23 10 4
9.45 10 4

0.99800
0.99812
0.99843
0.99815
0.99849
0.99677
0.99444

72
144
336
72
144
336
As sprayed

1050

slope of the loaddepth of penetration curve at the beginning of the


elastic unloading, S, and the contact depth, hp, could be established
from Eq. (2):
S = 2E

r
24:5
hp

The slope of this straight line is proportional to the reduced


modulus and Fig. 9a shows how the S values measured for several
contact depths (corresponding to the different maximum loads used)
arrange in a linear t. Table 2 shows the linear t equations presented
in Fig. 9a. The slope of these linear t equations will be used to
estimate the reduced modulus which is related to the top coating
Young's modulus by Eq. (4).
The size independent hardness could be determined in a similar
way. A linear relationship between the maximum load, Pmax, and the
contact area between the indenter and sample at maximum load, A,
could be established from Eq. (1):
7

Pmax = HA

Therefore, H is the slope of the PmaxA plots which are shown in Fig. 9b.
The experimental data arrange in a linear t making possible to determine
characteristic mechanical properties that are independent from indentation size. Table 3 includes the linear t parameters from Fig. 9b. Table 4
summarizes the size independent properties, E and H, obtained from
Fig. 9 for the as-sprayed and oxidized samples. Apparent Young's modulus
decreased from 190 GPa in the as-sprayed condition to 93 GPa after 336 h
at 1050 C. Similarly, apparent hardness decreased from 11.4 GPa in the assprayed condition to 7.1 GPa after 336 h at 1050 C. These results clearly

Table 3
Linear ts obtained from Fig. 9b.
P=U+V A
T (C)
950

1050

t (h)

U (mN)

72
144
336
72
144
336
As sprayed

V (mN m 2)

10.77
9.07
8.52
10.00
8.09
7.13
11.37

0.99990
0.99970
0.99968
0.99985
0.99996
0.99990
0.99982

3.26
5.37
5.42
3.95
2.52
2.97
3.80

Table 4
Independent size E and H of the as sprayed and oxidized top coatings.
T (C)
950

1050
Fig. 9. a) Slope of the unload branch of the loaddisplacement curve, S, vs. the contact
depth, hp, for the as sprayed and oxidized samples. b) Maximum load of indentation,
Pmax, vs. the contact area, A, for the as sprayed and oxidized samples.

t (h)
72
144
336
72
144
336
As sprayed

E (GPa)
168
160
144
164
130
93
190

H (GPa)
10.8
9.1
8.5
10.0
8.3
7.1
11.4

A. Rico et al. / Surface & Coatings Technology 203 (2009) 23072314

2313

treatment which reduces round porosity and damage tolerance [16,23].


Cracking within the ceramic layer further propagate along the top coat
due to tensile stresses in the ceramic layer. The consequence is that TGO
thickening induces a more compliant structure of the ceramic coating.
However, this higher compliance does not improve top coating damage
tolerance as this is due to the presence of cracks, which act as stress
concentrators favouring damage propagation.
The mechanical response of the top coat is, then, affected by
degradation at high temperature. Microcracking through the ceramic
coating is more important as the oxidation severity (higher time or
temperature) is increased. This is shown in Fig. 6 where it could be
observed how cracking at the middle of the top coating, where the
mechanical properties have been measured, increases lineally with the
exposure time and this increment is quicker at 1050 C than at 950 C. In
a rst approximation, Young's modulus and hardness are properties only
dependent on the material tested. Nevertheless, the presence of defects,
such as porosity or microcracks, reduces the stiffness and strength of the
structure diminishing the properties measured by sensing indentation,
which are, consequently, strongly affected by the volume of the sample
affected by the indentation. If the void fraction in the coating increases,
the probability that a pore or a crack is included in the indentation
volume also increases. Therefore, the apparent mechanical properties
take into account TBC degradation.
Fig.10a shows Young's modulus evolution with exposure time for the
two temperatures considered. The graph gives the ratio between the
apparent Young's modulus, E, and the apparent Young's modulus
measured before the oxidation tests, E0. The same plot is presented in
terms of the ratio between apparent hardness, H, and apparent hardness
before the oxidation, H0 in Fig. 10b. These ratios quantify the level of
cracking within the top coat as a consequence of TGO thickening. As it
can be seen, these ratios decrease as exposure time was increased.

Fig. 10. a) Ratio between the independent size Young's modulus of the oxidized samples, E,
and the independent size Young's modulus of the as sprayed sample, E0, vs. exposure time.
b) Ratio between the independent size hardness of the oxidized samples, H, and the
independent size hardness of the as sprayed sample, H0, vs. exposure time.

show how the apparent mechanical properties decreased as the oxidation


conditions were more severe, thus mechanical measurements seem to
take into account the top coating cracking due to TBC degradation by TGO
thickening.
4. Discussion
TBC exposed at high temperature experienced several degradation
processes up to failure by top coating spallation. TGO formation and
growth is the most important phenomenon controlling TBC durability.
This layer should be thermodynamically stable, slowly growing and
adherent. Initially, TGO thickness is xed below a critical value and this
layer acts as a protective scale due to its good adherence to the metallic
bond coat and its low porosity. Ingress of oxygen and other species to the
metallic bond coat is, then, reduced by the TGO and continuous
oxidation of the metal is impeded. However, as soon as a critical TGO
thickness is reached during oxidation, out of plane stresses, developed at
TGO-top coat and bond coat interfaces [5,1012,14,15], are high enough
to promote top coat spallation and cracking within the brittle ceramic
coating in the vicinity of the interface undulations (Figs. 5 and 6). These
stresses are also induced by ceramic coating resintering during the heat

Young's modulus and hardness decrement were more important at


the highest temperature. In addition, the difference between the
mechanical properties measured after oxidation at 950 C and 1050 C
was higher as the exposure time increased. This behaviour could be
explained by the evolution of cracking at the middle of the top coat,
where E and H were measured, presented in Fig. 6. Microcracking is
always higher after oxidation at 1050 C and the difference increased
with the exposure time. TGO thickens quicker during oxidation at
1050 C than at 950 C, as it can be observed in Table 1, driving to higher
stresses in the TGO-bond coat and top coat interfaces for the same
exposure time. The nal effect is a more important decrement in Young's
modulus and hardness at high temperature as it is shown in Fig. 10.
5. Conclusions
TBC's comprising a Ni22Cr10Al1Y bond coat and a ZrO2(8Y2O3) top
coat were air plasma sprayed onto an Inconel 600 Ni base alloy. The
sprayed samples were isothermally oxidized at 950 C and 1050 C for
72, 144 and 336 h. The mechanical properties, Young's modulus and
hardness, of the top coating and its evolution after isothermal
oxidation were analysed leading to the following conclusions:
(1) During oxidation in air a TGO layer was formed at the bond
coatceramic coating interface. TGO thickening promoted out
of plane stresses at the TGO-bond coat and ceramic coating
interfaces leading to failure by top coat spallation and cracking
within the ceramic layer in the vicinity of the interface.
(2) Crack density close to the interface reached a maximum value
~20% after 144 h at 1050 C and this value was maintained after
subsequent oxidation. However, the coatings treated at 950 C
reached this maximum value after 336 h. Cracking was extended
through the top coating. Crack density increased lineally at the
middle of the ceramic layer and this increment was quicker at
1050 C than at 950 C.

2314

A. Rico et al. / Surface & Coatings Technology 203 (2009) 23072314

(3) Depth sensing indentation has been able to measure size


independent apparent mechanical properties, E and H. The
measured values were signicantly reduced as the exposure
time was increased. Decrement of E and H was more important at
1050 C than at 950 C and this difference increased with the
exposure time due to a quicker TGO thickening at higher
temperature.
(4) The mechanical properties analysed in this study could be used
in life prediction models, based on the nite element method,
extending the actual knowledge about the failure mechanisms
in TBC.
Acknowledgements
The authors would like to thank the Spanish government CICYT
through grants MAT 2003-06147 and MAT 2007-64433 and Comunidad
de Madrid through the program ESTRUMAT-CM (reference MAT/77) for
nancial support.
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]

M.J. Pomeroy, Mater Design 26 (2005) 223.


R.A. Miller, J. Thermal Spray Technol. 6 (1) (1997) 35.
S. Bose, J. De Masi-Marcin, J. Thermal Spray Technol. 6 (1) (1997) 99.
V. Sergo, D.R. Clarke, J. Am. Ceram. Soc. 81 (1998) 3237.
N.P. Padture, M. Gell, E.H. Jordan, Science 296 (2006) 280.
G.W. Goward, Surf. Coat. Technol. 108109 (1998) 73.
M.J. Stiger, N.M. Yanar, M.G. Topping, F.S. Pettit, G.H. Meier, Z. Metallkd. 90 (1999)
1069.

[8] W.E. Lee, W.M. Rainforth, Ceramic microstructures, Chapman and Hall, London,
1994.
[9] P. Poza, P.S. Grant, Surf. Coat. Tech. 201 (2006) 2887.
[10] A.G. Evans, D.R. Mumm, J.W. Hutchinson, G.H. Meier, F.S. Pettit, Prog. Mater. Sci. 46
(2001) 505.
[11] A.G. Evans, M.Y. He, J.W. Hutchinson, Prog. Mater. Sci. 46 (2001) 249.
[12] M. Ahrens, R. Vaen, D. Stver, Surf. Coat. Technol. 161 (2002) 26.
[13] X. Wang, G. Lee, A. Atkinson, Acta Mater. 57 (2009) 182.
[14] M. Bialas, Surf. Coat. Technol. 202 (2008) 6002.
[15] R. Vaen, G. Kerchoff, D. Stver, Mater. Sci. Eng. A 303 (2001) 100.
[16] J.A. Thompson, T.W. Clyne, Acta Mater. 49 (2001) 1565.
[17] F. Tang, J.M. Shoenung, Scripta Mater. 54 (2006) 1587.
[18] J.Y. Kwon, K.H. Dong, J.H. Lee, Y.G. Jung, U. Paik, C.Y. Jo, Prog. Org. Coat. 61 (2008)
300.
[19] J. Lian, J.E. Garay, J. Wang, Scripta Mater. 56 (2007) 1095.
[20] C.C. Raceck, Berndt, Surf. Coat. Technol. 202 (2007) 362.
[21] S. Guo, Y. Kagawa, Ceram. Int. 32 (2006) 263.
[22] J.Y. Kwon, J.H. Lee, H.C. Kim, Y.G. Jung, U. Paik, K.S. Lee, Mater. Sci. Eng. A 429
(2006) 173.
[23] B. Siebert, C. Funke, R. Vaen, D. Stver, J. Mater. Process. Technol. 9293 (1999)
217.
[24] W.C. Oliver, G.M. Pharr, J. Mater. Res. 7 (4) (1992) 1564.
[25] A.C. Fisher-Cripps, Vacuum 58 (2000) 569.
[26] Jianghong Miao, Zhijian Peng, Longhao Qi, Mater. Sci. Eng. A 354 (2003) 140.
[27] Zhijian Gong, Hezhou Miao, J. Eur. Ceram. Soc. 24 (2004) 2193.
[28] J. Woirgard, C. Tromas, J.C. Girard, V. Audurier, J. Eur. Ceram. Soc. 18 (1998)
22972305.
[29] A. Krell, S. Schdlich, Mater. Sci. Eng. A 307 (2001) 172.
[30] J. Rodrguez, A. Rico, E. Otero, W.M. Rainforth, Acta Mater. (in press) DOI:10.106/j.
actamat.2009.03.020.
[31] A. Rico, M.A. Garrido, J. Rodrguez, Bol. Soc. Esp. Ceram. V. 47 (2008) 110.
[32] F. Traeger, M. Ahrens, R. Vaen, D. Stver, Mater. Sci. Eng. A 358 (2003) 255.
[33] J. Aktaa, K. Sfar, D. Munz, Acta Mater. 53 (2005) 4399.
[34] E.P. Busso, L. Wright, H.E. Evans, L.N. McCartney, S.R.J. Saunders, S. Osgerby, J. Nunn,
Acta Mater. 55 (2007) 1491.

You might also like