Sei sulla pagina 1di 8

Effects of phase constituents/microstructure of thermally grown oxide on the

failure of EB-PVD thermal barrier coating with NiCoCrAlY bond coat


J. Liu, J.W. Byeon, Y.H. Sohn

Advanced Materials Processing and Analysis Center and Department of Mechanical, Materials and Aerospace Engineering, University of Central Florida, Orlando,
FL 32816-2455, United States
Received 18 August 2005; accepted in revised form 31 August 2005
Available online 25 October 2005
Abstract
A correlation among thermal cycling lifetime, bond coat surface preparation, phase constituents and microstructure of thermally grown oxide
(TGO) was examined for electron beam physical vapor deposited (EB-PVD) thermal barrier coatings (TBCs) consisting of ZrO
2
7 wt.% Y
2
O
3
(YSZ) ceramic topcoat, NiCoCrAlY bond coats and CMSX-4 superalloy. Variation in the bond coat surface is characterized based on surface
roughness modification and pre-oxidation (1100 C at P
O
2
of 10
8
atm up to 4 h) carried out prior to YSZ deposition by EB-PVD. TBC specimens
with pre-oxidized bond coats exhibited longer lifetimes than those without pre-oxidation, especially for metallographically polished bond coats.
Given a surface roughness of NiCoCrAlY bond coats, TBC lifetime was observed to increase with an increase in the amount of -Al
2
O
3
in the
initial thermally grown oxide (TGO) scale, which was detected by photostimulated luminescence (PL). Using focused ion beam in-situ lift out
technique (FIB-INLO), site-specific preparation of thermally cycled TBC specimens for transmission electron microscopy (TEM) was
successfully carried out. In addition to the presence of -Al
2
O
3
, a small particulate phase identified as cubic Y
2
O
3
was observed within the TGO
scale by electron diffraction analysis on TEM.
2005 Elsevier B.V. All rights reserved.
Keywords: Thermal barrier coatings; Surface modification; Photoluminescence; Thermally grown oxide; Thermal cycling
1. Introduction
Thermal barrier coatings (TBCs) have helped to improve the
high temperature durability and performance of hot section
components in advanced gas turbine engines [15]. Generally,
TBC system consists of four layers: ceramic topcoat, thermally
grown oxide (TGO), metallic bond coat, and superalloy
substrate. Electron beam physical vapor deposition (EB-PVD)
has provided TBCs with advantageous properties, such as
improved durability, smooth surface finish and good erosion
resistance [6]. In EB-PVD TBCs, characteristics of bond coat
surface and the growth of TGO are considered to be the crucial
factor influencing the failure of TBCs [79]. The initial defects
and surface irregularities of the bond coats can give rise to in-
plane and out-of-plane tensile stresses. TGO growth results in a
constrained volume expansion that leads to compressive
growth stresses (<1 GPa) persisting at all temperatures. Upon
cooling, the thermal expansion mismatch between the TGO and
bond coat leads to a high compressive residual stress in the TGO
[1014] that provides the strain energy that can drive spallation,
typically within and/or near the TGO scale.
Phase constituent of the TGO is considered to be a critical
factor influencing the adhesion at the TGO/YSZ interface. The
formation of the metastable - and/or -Al
2
O
3
and its
conversion to the stable -Al
2
O
3
in the TGO has been reported
to have a profound effect on the structural integrity of the TGO
during thermal cycles [15,16]. The polymorphic transformation
of Al
2
O
3
may affect the residual stress in the TGO due to
volumetric constraint and the nucleation of sub-critical cracks
[17,18]. Thus, the formation of an optimum TGO that only
consists of -Al
2
O
3
prior to the deposition of topcoat may help
to improve the durability and reliability of TBCs.
In this study, TBCs, produced by the EB-PVD as a function
of NiCoCrAlY bond coat surface roughness, were examined to
Surface & Coatings Technology 200 (2006) 58695876
www.elsevier.com/locate/surfcoat

Corresponding author. Tel.: +1 407 882 1181; fax: +1 407 882 1462.
E-mail address: ysohn@mail.ucf.edu (Y.H. Sohn).
0257-8972/$ - see front matter 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.surfcoat.2005.08.140
investigate the influence of bond coat surface modification and
pre-oxidation heat treatment on their lifetime, microstructural
development and failure characteristics.
2. Experimental procedure
Twenty-six disk-shaped specimens of CMSX-4 (Ni9.0 Co
6.5 Cr6.5 Ta6.0 W5.6 Al3.0 Re1.0 Ti0.6 Mo0.1 Hf in
wt.%) superalloys coated with low pressure plasma sprayed
(LPPS) NiCoCrAlY (PWA276) bond coat and EB-PVD ZrO
2

7 wt.% Y
2
O
3
(YSZ) topcoat were employed in this study.
Various surface preparation and heat treatment were carried out
before the deposition of the top YSZ coatings as listed in Table
1. The bond coat surface can be categorized based on processing
technique: as-sprayed (as low-pressure plasma sprayed), hand-
polished metallographically down to 0.25 m, and barrel-
finished (media tumble treatment for 90 min). Moreover, after
each kind of the bond coat surface processing, selected
specimens were pre-oxidized in an oxidizing environment
(P
O
2
10
8
atm) at 1100 C for up to 4 h.
Prior to YSZ deposition, bond coat surface roughness and
phase constituents within the initial oxide layer were
examined by optical profilometry (OP) (WYKO
TM
NT 3300
optical profilometry) and photostimulated luminescence spec-
troscopy (PL) (Renishaw System 1000B Ramanscope),
respectively. From the bond coat surface roughness profiles,
the average value of the roughness R
a
was calculated using:
R
a

1
n
X
n
i1
R
i
1
where R
i
is the surface roughness of each lateral resolution
(0.3 m) with vertical resolution of 0.13.0 nm. Each
measurement contained 1000 spatial sampling (n).
After the topcoat deposition, changes in the phase
constituents of the TGO scale were examined by using
photostimulated luminescence (PL). Intensity ratio of -
Al
2
O
3
to the total intensity, I
/T
was determined by:
I
=T

I
R
I
R
I
N
I
m
2
where I
R
, I
N
, and I
m
refer to the integrated luminescence
intensities of -Al
2
O
3
, N-luminescence, and metastable alumina
luminescence, respectively.
Table 1
Summary of bond coat surface roughness, R
a
, the fracture surface characteristics and lifetime of TBCs as a function of bond coat surface preparation
Bond coat surface
preparation
Pre-oxidation
a
Average bond coat surface
roughness (R
a
, nm)
Fracture path Percentage of
TGO on the
exposed bond
coat %
Average lifetime
of TBCs (thermal
cycles)
Average St. Dev.
As-sprayed
b
No 1321.3 601.6 YSZ/TGO and TGO/Bond coat 70.5 87.5
Yes 1321.3 601.6 YSZ/TGO and TGO/Bond coat 58.9 101.0
Hand-polished
c
No 4.3 8.4 * * 0.0
Yes 302.6 21.1 TGO/Bond coat 18.0 95.0
Barrel-finished
d
No 320.1 10.3 YSZ/TGO 98.9 25.0
Yes 491.2 22.9 TGO/Bond coat 24.1 50.0
*No observable TGO. Fracture between the YSZ and bond coats prior to thermal cycling.
a
Heat treatment at P
O
2
10
8
atm and 1100 C for up to 4 h to form continuous TGO layer.
b
Surface finish as low pressure plasma was sprayed.
c
Surface finish by metallographic polishing.
d
Media tumble treatment for 90 min.
Fig. 1. Thermal cycling lifetime of TBCs as a function of bond coat surface finish and pre-oxidation heat treatment. Values in parenthesis represent the number of
specimens tested for the lifetime determination.
5870 J. Liu et al. / Surface & Coatings Technology 200 (2006) 58695876
A furnace thermal cycling test, that consisted of 10-min heat-
up to 1121 C, 40-min hold at 1121 C, followed by 10-min
forced air-quench, was carried out for all TBC specimens.
During the test, specimens were periodically withdrawn for PL
measurements at room temperature to assess the residual stress
and phase transformations within the TGO as a function of
thermal cycles. The specimens were considered to have failed
when the YSZ spallation area was greater than 50%.
Microstructure and phase constituents of the fracture surfaces
(i.e., the top surface of the bond coat where the YSZ had spalled,
and the bottom surface of the spalled YSZ) and cross-sections
were analyzed by scanning electron microscopy (SEM) and X-
ray energy dispersive spectroscopy (XEDS). Focused ion beam
in-situ lift out technique (FIB-INLO) [19] was employed for the
preparation of site-specific transmission electron microscopy
(TEM) specimens with an emphasis on the TGOscale. TEMand
scanning TEM (STEM) analysis included high angle annular
dark field (HAADF) imaging, electron diffraction, and XEDS.
3. Results
Fig. 1 shows the average thermal cycling lifetime of TBC
specimens as a function of bond coat surface preparation. The
reported lifetimes are low for all TBCs when compared to those
available in numerous literatures. In addition, without pre-
oxidation, TBCs with hand-polished bond coats failed before
any thermal cycling (i.e., 0 lifetime). The lifetime of all TBCs
increased after pre-oxidation. The most significant improve-
ment in TBC lifetime was observed for specimens with hand-
polished bond coats that were pre-oxidized prior to the EB-PVD
of YSZ.
Table 1 reports the bond coat surface roughness values. The
roughest surface was observed for as-sprayed bond coat, while
the smoothest surface was observed for hand-polished bond
coat. Fig. 2 presents the typical PL spectra from the TGO scale
developed on the TBC specimen with barrel-finished bond
coats with and without pre-oxidation. Similar spectra were
obtained for TBC specimens with as-sprayed bond coat. The N-
peaks of typical PL spectrum from the TGO scale developed on
the TBC specimen with hand-polished bond coats with pre-
oxidation was also shown in Fig. 2.
Figs. 3 and 4 present the magnitude of compressive residual
stress in the -Al
2
O
3
scale as a function of thermal cycle for
TBCs produced without and with pre-oxidation heat treatment,
respectively. In Fig. 3, the magnitude of the initial compressive
residual stress is higher for TBCs with as-sprayed bond coat,
and lower for the TBCs with barrel-finished bond coat.
However, during the thermal cycling test, the compressive
residual stress in the -Al
2
O
3
scale increased rapidly for TBCs
with barrel-finished bond coats after 5 cycles. In Fig. 4, the
magnitude of the initial compressive residual stress is the
highest for TBCs with hand-polished bond coats, followed by
as-sprayed and barrel-finished. During thermal cycling test, the
magnitude of compressive residual stress in the -Al
2
O
3
scale
Fig. 2. Typical PL spectra from the TGO scale developed on the TBC
specimens with barrel-finished bond coats with and without pre-oxidation
and TBC specimen with hand-polished bond coat and pre-oxidation, showing
N-luminescence.
Fig. 3. Magnitude of compressive residual stress in the -Al
2
O
3
scale for TBCs,
determined from PL measurements, as a function of thermal cycle. The bond
coats for these TBCs were not pre-oxidized.
Fig. 4. Magnitude of compressive residual stress within the -Al
2
O
3
scale for
TBCs, determined from PL measurements, with pre-oxidized bond coats as a
function of thermal cycle.
5871 J. Liu et al. / Surface & Coatings Technology 200 (2006) 58695876
of TBCs with hand-polished and as-sprayed bond coats
remained high compared to that of TBCs with barrel-finished
bond coats as shown in Fig. 4. The standard deviation of the
compressive residual stress during thermal cycling is
significantly larger for TBCs with barrel-finished bond coat.
The initial rise in compressive residual stress after few thermal
cycles, as shown in Figs. 3 and 4 can be attributed to the
coverage of TGO scale from discontinuous to continuous
[20,21]. In Gell's studies, the initial oxide scale did not form a
continuous layer until after at least 10, 1-h thermal cycles at
1121 C.
Cross-sectional microstructure of TBCs with pre-oxidized
bond coats are shown in Fig. 5. Fig. 5(a) presents the cross-
sectional microstructure of failed TBCs with hand-polished
Fig. 5. Cross-sectional backscattered electron micrographs of failed TBC specimens with (a) hand-polished and pre-oxidized bond coat, showing Yand Hf-rich oxides
embedded in the TGO; (b) Cross-sectional backscattered electron micrograph of TBC specimen with as-sprayed and pre-oxidized bond coat; showing Yand Hf or Ni,
Co and Cr-rich particles embedded in the TGO.
Fig. 6. (a) HAADF images of Y
2
O
3
embedded in the TGO (-Al
2
O
3
) for failed TBC specimen with hand-polished and pre-oxidized bond coats. Small precipitates at
the Y
2
O
3
/-Al
2
O
3
interface boundaries are very rich in Hf; (b) Diffraction pattern of Y
2
O
3
(cubic) particles marked as (b) in (a); (c) Diffraction patterns of -Al
2
O
3
marked as (c) in (a); (d) HAADF image of TGO of as-sprayed and pre-oxidized bond coat.
5872 J. Liu et al. / Surface & Coatings Technology 200 (2006) 58695876
bond coat. Particles rich in Yand Hf were frequently found in the
TGOscale. Fig. 5(b) shows the cross-sectional microstructure of
failed TBCs with as-sprayed bond coats. Particles rich in Y
and Hf were also frequently found within the TGO, which is
penetrated into the bond coats (i.e., pegging). Oxides rich in
Ni, Co, and Cr were frequently found near the fracture
interfaces, which may have led to the final failure.
Fig. 6(a) shows a HAADF image obtained from STEM of
particles embedded within the TGO layer parallel to the YSZ/
TGO/bond-coat interfaces for hand-polished and pre-oxidized
bond coats. The TGO primarily consisted of -Al
2
O
3
as
confirmed by the diffraction pattern in Fig. 6(c). XEDS shows
that the particles embedded in -Al
2
O
3
are rich in Y or Hf. Y-
rich particles were identified as cubic-Y
2
O
3
by the diffraction
pattern shown in Fig. 6(b). Hf concentration was observed to be
higher at Al
2
O
3
grain boundaries and Al
2
O
3
/Y
2
O
3
interphase
boundaries. Specifically, grain boundaries of Al
2
O
3
gave rise to
brighter contrast in HAADF images in Fig. 6 (a). For TBCs with
as-sprayed and pre-oxidized bond coats, the TGO also primarily
consisted of -Al
2
O
3
. Y
2
O
3
and Cr-rich particles embedded in
-Al
2
O
3
TGO were also observed as shown in Fig. 6(d).
Backscattered electron images from the bottom surface of
the spalled YSZ and the top surface of the bond coats where
YSZ had spalled are presented in Fig. 7. For TBC specimen
with barrel-finished bond coat without any pre-oxidation heat
treatment, the phase constituents on the fracture surfaces
suggest that the spallation occurred primarily along the
interfaces between the YSZ and TGO as shown in Fig. 7(a)
and (b). However, after pre-oxidation, the phase constituents
of the fracture paths changed, along with the increase in
thermal cycling lifetime. For TBC specimens with barrel-
finished bond coat surface with pre-oxidation heat treatment,
the phase constituents on the fracture surfaces indicate that
the spallation occurred primarily along the TGO/bond coat
interface as presented in Fig. 7(c) and (d).
Characteristics of fracture paths and lifetimes of TBCs as a
function of bond coat surface preparation are summarized in
Table 1. For TBCs with hand-polished and barrel-finished bond
coats after pre-oxidation, more TGO remained on the bottom
surface of the spalled YSZ with increases in the thermal cycling
lifetime. The fracture path remained similar for TBCs with as-
sprayed bond coat (i.e., both interfaces of YSZ/TGO and TGO/
bond coat as well as within the TGO) after the pre-oxidation
heat treatment: the improvement in the lifetime of the TBCs
with as-sprayed bond coats was not significant after the pre-
oxidation heat treatment.
Fig. 7. (a) Backscattered electron image of the bottom surface of spalled YSZ and (b) the top surface of the bond coat for TBC specimen with barrel-finished bond coat
without pre-oxidation. The spallation has occurred after 20 cycles along the YSZ/TGOinterface. (c) Backscattered electron image of the bottomsurface of spalled YSZ
and (d) the top surface of the bond coat for TBC specimen with barrel-finished and pre-oxidation bond coat. The spallation has occurred after 75 cycles along the TGO/
bond coat interface.
5873 J. Liu et al. / Surface & Coatings Technology 200 (2006) 58695876
4. Discussion
Fig. 8 correlates the bond coat surface preparation
technique, the average bond coat surface roughness (R
a
), and
the average thermal cycling lifetime of TBCs. For TBCs
without pre-oxidation heat treatment, a longer TBC lifetime
was observed with rougher bond coats as presented in Fig. 8
(a). However, such a simple relation does not exist for TBCs
after the pre-oxidation heat treatment as presented in Fig. 8
(b). TBCs with hand-polished and pre-oxidized bond coats
yielded higher durability than TBCs with barrel-finished and
pre-oxidized bond coat. Clearly, the surface roughness of the
bond coat alone is not the only factor that affects the lifetime
of the TBCs.
For TBC specimens produced without any pre-oxidation
heat treatment, Fig. 9(a) shows the correlation among the initial
residual stresses of the TGO scale (-Al
2
O
3
), bond coat surface
preparation and the thermal cycling lifetime. The magnitude of
initial compressive residual stress in the -Al
2
O
3
is only slightly
higher for TBCs with as-sprayed bond coats, which has the
longer lifetime. For TBCs with pre-oxidized bond coats, the
magnitude of the initial compressive residual stress in the -
Al
2
O
3
was similar for the TBCs with as-sprayed and barrel-
finished bond coats. However, the lifetime of these two types of
specimens had a great difference as shown in Fig. 9(b).
Throughout thermal cycling, for TBCs without pre-oxidation
heat treatment, the magnitude of compressive residual stress
within the TGO was observed to be lower for the specimens
with as-sprayed bond coat. These TBCs had a longer lifetime.
For TBC specimens with barrel-finished bond coat and a shorter
lifetime, the magnitude of the compressive residual stress in the
TGO was higher as shown in Fig. 3. For TBCs with pre-
oxidized bond coats, the magnitude of compressive residual
stress within the TGO remained higher during thermal cycling
for the TBCs with longer lifetime as shown in Fig. 4. Clearly,
the initial compressive residual stress within the TGO is not the
only factor that can be directly and simply related to the lifetime
of the TBCs.
A longer lifetime was obtained for specimens with higher
relative luminescence intensity from the equilibrium-Al
2
O
3
in
the initial TGO layer for a given surface preparation/roughness
as shown in Fig. 10. Phase transformation from the metastable
to stable phase may be accompanied by the formation of voids
due to volumetric constraints. For example, there is about 4.7%
volume contraction for - to -Al
2
O
3
phase transformation [18].
Thus, these phase transformations can create damages at the
YSZ/TGO interface. However, controlled pre-oxidation heat
treatment that can promote the formation of -Al
2
O
3
and stop
premature failure at the YSZ/TGO interface can increase the
lifetime of the TBCs.
According to Czech's study [22], a continuous -Al
2
O
3
layer formed on the hand-polished surface after oxidation at
950 and 1000 C. In this study, N-luminescence from TGO
scale was observed from the pre-oxidized TBC specimens with
Fig. 8. Correlation among the bond coat surface preparation method, the bond
coat surface roughness and the lifetime of EB-PVD TBCs with NiCoCrAlY
bond coat: (a) without pre-oxidation heat treatment; and (b) with pre-oxidation
heat treatment.
Fig. 9. Correlation among the initial residual compressive stress within the -
Al
2
O
3
scale, bond coat surface preparation and the thermal cycling lifetime for
TBCs specimens: (a) without pre-oxidation heat treatment and (b) with pre-
oxidized bond coats.
5874 J. Liu et al. / Surface & Coatings Technology 200 (2006) 58695876
hand-polished bond coat as shown in Fig. 2, presumably due to
a significant presence of Cr
3+
in -Al
2
O
3
. The formation of
Cr
2
O
3
has been reported to assist the nucleation of -Al
2
O
3
[18]. In addition, hand-polishing may introduce residual
stresses near the bond coat surface, which can provide
additional strain energy that can aid the equilibrium phase
nucleation and/or transformation. Moreover, the hand-polished
process can remove the absorbed contaminants that can
adversely affect the phase nucleation/transformation. These
contaminants can also influence the integrity of TGO/bond coat
interface [23].
Initial phase constituents within the TGO, not only affected
the lifetime of TBCs, but also influenced the initial com-
pressive residual stress within the -Al
2
O
3
. Before thermal
cycling, the higher relative luminescence intensity from the
equilibrium -Al
2
O
3
corresponded to the larger magnitude of
initial compressive residual stress for TBCs with pre-oxidized
bond coats as shown in Fig. 11.
A significant amount of spinels were observed at fracture
interfaces as shown in Fig. 5. According to Lee's and Haynes's
studies [24,25], they found that the fracture in the bond coat
oxidation products often takes place in the region of spinels.
Spinels are considered to be brittle material with low fracture
toughness that weakens the interface [26,27].
The difference in the thermal expansion coefficient of the
bond coat and TGO is much larger than that of the TGO and
YSZ.[28] Therefore, after a significant repeated thermal fatigue,
fracture would prefer to propagate at the TGO/bond coat
interface. For specimen with shorter lifetime, spallation occurs
at the interface of YSZ/TGO, due to a poor adhesion between
the YSZ and the TGO. Thus, TBCs with the longest durability
can be achieved by having the fracture paths restricted to the
TGO/bond coat interface when an optimum TGO scale with a
good adhesion between YSZ and TGO is established via pre-
oxidation heat-treatment. Similar spallation results were found
by Mumm and Evans using EB-PVD TBCs [29]. Based on
their research, after short time (<10 h) of isothermal exposure
at 1100 C, delamination occurs primarily within the TGO
and YSZ. After longer exposures (<100 h), delamination
occurs principally along the interface of TGO and bond coat.
According to Wright [30], reactive elements (RE) (e.g., Y)
inhibit S segregation at the metal-oxide interface. Also, RE
segregate to oxide grain boundaries, where they can sig-
nificantly reduce the outward transport of Al, hence decrease
the rate of oxidation (now mostly by oxygen transport),
drastically change the oxide morphology and contribute to the
improved scale adherence (reduced interfacial void formation)
[23]. However, oversupply of the RE can be detrimental.
Increased concentration of RE in the bond coat can cause
internal oxidation and formation of RE oxides [24] (e.g., HfO
2
,
Y
2
O
3
and YAG) within the TGO scale as shown in Fig. 6. Also,
the formation of RE oxides can act as fast diffusion paths for
oxygen and cause mechanical disruption if incorporated into the
TGO scale. Extensive formation of Y
2
O
3
and other RE oxides
may be responsible for overall poor performance of lifetime
observed in this study.
5. Conclusion
Effects of surface preparation methods for a NiCoCrAlY
bond coat on the thermal cyclic lifetime and failure of an EB-
PVD TBC were investigated in this study. Findings from this
investigation are listed below:
A longer lifetime during 1-h thermal cycling at 1121 C was
observed for EB-PVD TBCs with as-sprayed NiCoCrAlY
bond coats regardless of pre-oxidation and TBCs with hand-
polished bond coats only after pre-oxidation.
Fig. 10. Correlation among the relative initial intensity ratio of -Al
2
O
3
, bond
coat surface preparation, and the thermal cycling lifetime of TBCs specimens:
TBCs with (a) as-sprayed bond coat; (b) hand-polished bond coat; (c) barrel-
finished bond coat.
Fig. 11. Correlation among the bond coat surface preparation method, initial
residual stress within the -Al
2
O
3
scale, and initial relative intensity ratio of the
equilibrium -Al
2
O
3
for pre-oxidized TBC specimens.
5875 J. Liu et al. / Surface & Coatings Technology 200 (2006) 58695876
Lifetime of EB-PVD TBCs, in general, was observed to
increase when a pre-oxidation heat treatment at 1100 C with
P
O
2
=10
8
atm was carried out prior to the YSZ deposition.
With pre-oxidation heat treatment, relative photostimulated
luminescence intensity of the -Al
2
O
3
increased. Thus, the
improvement in TBC lifetime can be correlated with the
increase in luminescence intensity of -Al
2
O
3
in the TGO
scale given a surface preparation/roughness. The lifetime
improvement due to the pre-oxidation was particularly
significant to TBCs with hand-polished NiCoCrAlY bond
coat.
Spallation-fracture paths depended on the lifetime of TBCs.
Premature spallation of TBCs occurs at the interface between
the YSZ and TGO. Longer durability can be achieved by
restricting the fracture paths to the TGO/bond coat interface.
Small particulate phase observed through the TGO scale was
identified as Y
2
O
3
(cubic) by diffraction analysis on TEM.
While the addition of Y in the NiCoCrAlY bond coat helps
the adhesion of the TGO scale, excessive alloying can lead to
deleterious effects.
Acknowledgements
The authors would like to thank Drs. F.S. Pettit and G.H.
Meier at University of Pittsburgh for pre-oxidation heat
treatment of TBC specimens, and Mr. K.S. Murphy at Howmet
for processing the TBC. Authors would also like to express their
gratitude to Drs. M. Gell and E.H. Jordan for financial support
of this study, which was carried out as a part of University
Turbine Systems Research (No. 01-01-SR091) awarded by
South Carolina Institute for Energy Studies administered by Dr.
R.A. Wenglarz.
References
[1] N.P. Padture, M. Gell, E.H. Jordan, Science 296 (2002) 280.
[2] D.J. Wortman, B.A. Nagaraj, E.C. Duderstadt, Mater. Sci. Eng., A Struct.
Mater.: Prop. Microstruct. Process. 120121 (1989) 433.
[3] R. Miller, J. Therm. Spray Technol. 6 (1997) 35.
[4] A.G. Evans, D.R. Mumm, J.W. Hutchinson, G.H. Meier, F.S. Pettit, Prog.
Mater. Sci. 46 (2001) 505.
[5] D. Zhu, R.A. Miller, MRS Bull. 25 (2000) 43.
[6] F.C. Toriz, A.B. Thakker, S.K. Gupta, Surf. Coat. Technol. 3940 (1989)
161.
[7] P.K. Wright, Mater. Sci. Eng., A Struct. Mater.: Prop. Microstruct. Process.
245 (1998) 191.
[8] J. Cheng, E.H. Jordan, B. Barber, M. Gell, Acta Mater. 46 (1998) 5839.
[9] A.G. Evans, M.Y. He, J.W. Hutchinson, Prog. Mater. Sci. 46 (2001) 249.
[10] D.M. Lipkin, D.R. Clarke, Oxid. Met. 45 (1996) 267.
[11] V.K. Tolpygo, D.R. Clarke, Oxid. Met. 49 (1998) 187.
[12] C. Mennicke, E. Schumann, C. Ulrich, M. Ruehle, Mat. Sci. Forum 389
(1997) 251.
[13] V. Sergo, D.R. Clarke, J. Am. Ceram. Soc. 81 (1998) 3237.
[14] R.J. Christensen, V.K. Tolpygo, D.R. Clarke, Acta Mater. 45 (1997) 1761.
[15] T.F. An, H.R. Guan, X.F. Sun, Z.Q. Hu, Oxid. Met. 54 (2000) 301.
[16] M.W. Brumm, H.J. Grabke, Corros. Sci. 33 (1992) 1677.
[17] C. Mennicke, D.R. Mumm, D.R. Clarke, Z. Metallkd. 90 (1999) 1079.
[18] D.R. Clarke, V. Sergo, M.Y. He, in: J.M. Hampikian, N.B. Dahotre (Eds.),
Elevated Temperature Coatings: Science and Technology, vol. III, TMS,
Warrendale, PA, 1999, p. 67.
[19] K. Gamo, Nucl. Instrum. Methods Phys. Res., B Beam Interact. Mater.
Atoms 121 (1997) 464.
[20] L.D. Xie, Y.H. Sohn, E.H. Jordan, M. Gell, Surf. Coat. Technol. 176
(2003) 57.
[21] S. Sridharan, L.D. Xie, E.H. Jordan, M. Gell, Surf. Coat. Technol. 179
(2004) 286.
[22] N. Czech, Surf. Coat. Technol. 108109 (1998) 36.
[23] B.A. Pint, Oxid. Met. 45 (1996) 1.
[24] E.Y. Lee, Life prediction and failure mechanisms for thermal barrier
coatings, Ph.D. dissertation, Worcester Polytechnic Institute (1991).
[25] J.A. Haynes, E.D. Rigney, M.K. Ferber, W.D. Porter, Surf. Coat. Technol.
8687 (1996) 102.
[26] C.A. Calow, I.T. Porter, J. Mater. Sci. 6 (1971) 156.
[27] R.G. Vardiman, Mater. Res. Bull. 7 (1972) 699.
[28] A.G. Evans, D.R. Mumm, J.W. Hutchinson, G.H. Meier, F.S. Pettit, Prog.
Mater. Sci. 46 (2001) 505.
[29] D.R. Mumm, A.G. Evans, Acta Mater. 48 (2000) 1815.
[30] I.G. Wright, B.A. Pint, First International Conference on Industrial Gas
Turbine Technologies CAME-GT Brussels, 1011 July, 2003.
5876 J. Liu et al. / Surface & Coatings Technology 200 (2006) 58695876

Potrebbero piacerti anche