Sie sind auf Seite 1von 7

Surface and Coatings Technology 108109 (1998) 7379

Progress in coatings for gas turbine airfoils


G.W. Goward*
40 Uncas Road, Clinton, CT 06413, USA

Abstract
The development of ever more efficient gas turbines has always been paced by the results of research and development in the
concurrent fields of design and materials technology. Improved structural design and airfoil cooling technology applied to higher
strength-at-temperature alloys cast by increasingly complex methods, and coated with steadily improved coating systems, have led to
remarkably efficient turbine engines for aircraft propulsion and power generation. For first stage turbine blades, nickel-based superalloys
in various wrought and cast forms, and augmented by coatings since the 1960s, have been singularly successful materials systems for the
past 50 yearsand still no real world substitutes are on the horizon. This paper traces the history of protective coatings for superalloy
airfoils beginning with the simple aluminides, followed by modifications with silicon, chromium and platinum, then MCrAlY overlay
coatings, and finally the elegant electron beam vapor deposited ceramic thermal barrier coatings recently introduced to service. The
publicly available results of several decades of research supporting these advances are highlighted. These include generic research on
oxidation and hot corrosion mechanisms of superalloys and coatings, the intricacies of protective oxide adherence, mechanisms of low
temperature (Type II) hot corrosion, and of aluminide coating formation and mechanical properties of alloycoating systems. With no
promising turbine materials beyond coated nickel-base superalloys apparent in the foreseeable future, continued progress will likely be
made by further refinement of control of thermally grown oxide adherence, and by more cost effective manufacturing processes for
contemporary types of protective coatings. 1998 Elsevier Science S.A. All rights reserved.
Keywords: Coatings; Gas turbines; Superalloys aluminides; MCrAlY; Thermal barrier

1. Introduction
The efficiency of all types of gas turbine engines,
aircraft, terrestrial and marine, is proportional to firing or
turbine inlet temperature. Increases in temperature are
facilitated by improved structural design and airfoil cooling technology applied to higher strength-at-temperature
alloys cast by increasingly complex processes, and coated
with steadily improved protection systems. First stage
turbine blades, the most critical components of gas turbines, made from nickel-base superalloys in various
wrought and cast forms, and augmented by coatings, have
been singularly successful materials systems for the past
50 years. This paper will trace the history of turbine airfoil
coatings from simple and modified diffusion aluminides,
through MCrAlY overlay systems, and finally, thermal
barrier coatings. Highlights of several decades of research
supporting these advances will be presented. Attempts will
be made to predict future advances in the science and
technology of superalloy coatings. Because of space
*Tel.: 11-1860-6691272;
sailbill35@aol.com

fax:

11-1860-203-4815039;

e-mail:

limitations references have been chosen so as to guide


readers to more detailed bibliographies and relevant photos
and tables of data.

2. Historical

2.1. Diffusion coatings


The first public descriptions of pack cementation
aluminizing were by Van Aller in a US patent filed in 1911
[1] and in a 1914 paper by Allison and Hawkins [2], all of
the General Electric Research Laboratory. Metals were
coated by embedding in a powder mixture of aluminum,
sal ammoniac (NH 4 Cl), and graphite, and heating the
assembly at 4508C (8428F) for 2 h. Later, Gilson, also at
G.E., patented the use of alumina as inert filler [3]; this
was blended with a chloride and aluminum powder and the
mix used to calorize metals to render them inoxidizable. Early uses included coating iron wire or ribbon
heating elements, and copper for power plant steam
condenser tubes. Ruder [4] summarized uses of aluminizing in 1915, including coating steel furnace fixtures and

0257-8972 / 98 / $ see front matter 1998 Elsevier Science S.A. All rights reserved.
PII: S0257-8972( 98 )00667-7

74

G.W. Goward / Surface and Coatings Technology 108 109 (1998) 73 79

nickel combustion screens. These pioneers correctly attributed the enhanced oxidation resistance of coated metals to
the selective formation of alumina scales. In about 1942
Anselm Franz in Germany [5] used aluminized low alloy
steels for combustion chambers, and possibly blades, in the
Jumo 004 engine to avoid the use of short supplies of
nickel and chromium in highly alloyed steels. The engine
powered the Messerschmidt ME 262 fighter, which entered
service in Germany in mid-1944. In the post-war period,
additional development and uses of the process up to the
first practical use of pack cementation aluminizing of
cobalt-base gas turbine vane airfoils in 1957 [6] are not
well documented. The first aluminizing of nickel-base
turbine blades may have been by hot dip processes at
Allison [7] and Curtiss Wright in about 1952. Pratt and
Whitney introduced nickel-base blades aluminized by a
slurry-fusion process in about 1963 [8]. Since about 1970
most vane and blade coatings have been applied by pack
cementation and the more recently developed out-ofcontact or chemical vapor deposition (CVD) processes
[911]. The surfaces of internal cooling passages are
coated by slurry aluminizing [12], or more effectively, by
forced flow gas phase aluminizing [11,13] and by vacuum
pulse aluminizing [14].
Kelley described his invention of pack cementation
chromizing of steels in 1923 [15]. One of the first uses was
for the protection of steam turbine buckets. Samuel and
Lockington [16] published a comprehensive review of
chromize coatings on steel in 19511952, and Drewett
[17] another in 1951. In 1953 Gibson patented aluminizing
of chromized steels with aluminum paints and heat treatment to improve high-temperature oxidation resistance
[18]. Time of first use of chromizing of gas turbine airfoils
is obscureit may have been in Europe for protection of
industrial gas turbine airfoils in the early 1960s [19]. Such
applications continue for protection of later stage airfoils in
both ground and aircraft engines. Chromizing, followed by
aluminizing is used on early stage turbine airfoils in many
older aircraft engines. From the 1970s on, developments in
the field of diffusion coatings include modification of
aluminide coatings with chromium [20], silicon [21], and
platinum [22]. In the 1990s, aluminide coatings have been
recognized as useful bond coats for some types of thermal
barrier coatings [23]. While contemporary commercial, and
some military gas turbines usually have more advanced
coating systems, e.g. platinum aluminides, MCrAlYs, etc.,
on first stage blades, it is estimated that greater than 80%
of all coated airfoils are coated by pack cementation or the
more recently developed out-of-contact aluminizing and / or
chromizing processes. Incorporation of so-called active
elements in diffusion aluminide coatings to enhance adherence of protective alumina appears to have been successful
on an experimental basis [24], but no such coatings are in
commercial production. References [25,26] provide more
details on the above items up until 1995.

2.2. Overlay coatings


Pratt and Whitney initiated a program in the late 1960s
to develop coatings with compositions nominally independent of substrates, and with capabilities for tailoring to the
wide range of requirements of gas turbine applications.
Research on the effect of reactive elements in improving
oxide adherence led to exploitation of the phenomenon in
new coating compositions. In 1937 Griffiths and Pfeil
patented [27] the use of active element additions, cerium,
mischmetal, etc., to heater element alloys to enhance
resistance to thermal cycling. In 1964 an alloy with the
composition Fe25%Cr4%Al1.0%Y emerged from
G.E. nuclear programs [28]. Pratt and Whitney research
confirmed that the alloy exhibited exceptional alumina
adherence in cyclic oxidation. A model composition
FeCrAlY [29], with 1015% aluminum, was applied as a
coating to nickel-base superalloys by electron beam physical vapor deposition (EBPVD)thickness about 125 mm
(5 mils)in a cooperative program with Airco Temescal
Corp. The coating was not practically useful at high
temperatures because nickel in the substrate alloys reacted
with aluminum in the coating to form a layer of NiAl
between the coating and the alloy. Next in the series was
CoCrAlY [30], which in special composition ranges, has
useful hot corrosion and oxidation resistance in some
applications. Within these ranges, however, it did not meet
ductility requirements for airfoils in high-performance
commercial and military engines. Since NiCrAlY has
limited hot corrosion resistance a compromise was sought
by adding cobalt. This led to a NiCoCrAlY coating of
exceptional ductility within a useful range of compositions
[31]. The nominal composition Ni23%Co15%Cr
12.5%Al0.5%Y had adequate oxidation and hot corrosion
properties and proved to be satisfactory for protection of
blade airfoils in then contemporary Pratt and Whitney
commercial aircraft engines. A follow on with additions of
silicon and hafnium [32] is still used on early stage turbine
airfoils, either alone or as a thermal barrier coating
bondcoat, in Pratt and Whitney commercial engines.
Throughout the industry many more complex NiCoCrAlY
compositions containing additions of tantalum, tungsten,
titanium, niobium, rhenium, zirconium, etc., singly or in
combination, have been developed for specific purposes.
Many of these are for protection of early stage airfoils in
industrial turbines. Initial emphasis was on improved
resistance to hot corrosion, including the low-temperature
Type II variety, by the use of up to 40% chromium [33].
The higher temperatures in newer industrial turbines fueled
with natural gas have caused more emphasis to be placed
on oxidation resistance and some coatings are now augmented by over-aluminizing [34]. There are now at least
40 patented variations of the original MCrAlY concept.
Manufacture of overlay coatings by EBPVD dominated
through the 1970s. Low-pressure plasma spray (LPPS),

G.W. Goward / Surface and Coatings Technology 108 109 (1998) 73 79

and lately the Union Carbide (Praxair) shrouded plasma


torch, are now favored for more complex formulations.
High-velocity oxy-fuel (HVOF) deposition is also being
tested.

2.3. Thermal barrier coatings


The authors first experience with thermal barrier coatings was with a system for protection of a TD nickel
transition duct in a military engine in 1970. A NiCr bond
coat and MgOZrO 2 top coat, both applied by air plasma
spray, doubled the thermal fatigue life of the duct. Subsequent use of the coatings on burner liners and development work at Pratt and Whitney and NASA, led to the use
of MCrAlY bond coats [35,36] and topcoats of zirconia
partially stabilized with 8% yttria [37]. Major programs
were initiated in the late 1970s to solve top coat spalling
problems with the goal of using insulating properties to
lower temperature and cyclic thermal stresses of turbine
airfoils. Control of crack formation and substrate temperature during coating application provided beneficially micro-cracked and / or segmented structures in the ceramic
layer [38]. The improved coatings are useful for vane
platforms and / or airfoils but durability sufficient for use
on blades proved elusive. A breakthrough came in the late
1970s [39] when it was demonstrated that columnar, low
modulus zirconia structures applied by EBPVD had thermal cyclic lives of the order of a factor of 10 over the best
plasma-sprayed coatings. The early work used EBPVD
NiCoCrAlY as bond coats but either low-pressure plasmasprayed NiCoCrAlY types [36], or diffusion aluminide
coatings [23], modified with platinum, are now in use.
Considering that these coatings can lower the temperature
of a cooled blade by up to 1708C (3068F), while simultaneously reducing cyclic thermal strains, the importance
of this technology cannot be underestimated.

3. Science and technology base

3.1. Diffusion coatings


In general, practical processes and applications long
preceded development of a significant science and technology base for these coatings. In 1971, research by Goward
and Boone [40] provided a qualitative description of
diffusion mechanisms for the formation of aluminide
coatings on nickel-base superalloys. The results enabled
classification of the coatings as either inward diffusion
types, based on the singular motion of aluminum in Ni 2 Al 3
and high aluminum NiAl, or outward diffusion types,
based on the higher rate of diffusion of nickel in low
aluminum NiAl. While some microstructural variations are

75

obtainable by varying the activities of source material [25],


this simple classification has found broad use in the
coating community. Research by Janssen and Rieck [41]
and Shankar and Seigle [42] quantified these unique
diffusion mechanisms. Shankar and Seigle demonstrated
with coatings that diffusion of nickel predominates in
nickel-rich NiAl; rates of diffusion of nickel and aluminum
are equal at about 51 at.% of aluminum; and aluminum
diffusion predominates in aluminum-rich NiAl as observed
by Goward and Boone. Work by Walsh [43], Levine and
Caves [44] and Seigle and co-workers [26,45,46] on the
thermodynamics and kinetics of pack aluminizing provided
a firm foundation for further technological progress. Gupta
and Seigle [47] reported on fundamentals for the selection
of the most effective activators for aluminizing and
chromizing, and refined details of the kinetics of coating
processes. Johnson and Komarek [48] published data on
activity coefficients of aluminum in NiAl, CoAl and
FeAl alloys which aids in the selection of source compositions for pack cementation and out-of-contact processes. Rapp and co-worker [24] explored thermodynamic
details for processes for co-deposition of other elements,
chromium, silicon, yttrium, hafnium, etc., with aluminum.
With regard to mechanisms of protection and degradation, Pettit defined mechanisms of oxidation of nickel
aluminum [49] and nickelchromiumaluminum alloys
[50]. Research at NASA [51] established the importance of
interdiffusion of aluminides with base alloys in decreasing
coating life. Aldred [52] disclosed that small amounts of
hafnium in a substrate alloy aided adherence of protective
alumina on coatings. Research on hot corrosion [53]
helped to define which substrate elements would benefit or
decrease the hot corrosion resistance of aluminide coatings. Godlewska and Godlewski [54] explored the hot
corrosion behavior of chromium modified aluminide coatings and found that the best coatings contained chromium,
from chromizing, in the outer layer of inward diffusion
aluminide coatings. Wu et al. [55], using electrochemical
molten sulfate salt techniques, showed that two-phase
(PtAl 2 in NiPtAl) platinum-modified aluminide coatings
had good resistance to high temperature (Type I) but lesser
resistance to low temperature (Type II) hot corrosion.
Barkalow and Pettit [56] found that a thin layer of PtAl 2
on the outer surface of such coatings provided much
improved resistance to Type II hot corrosion.
With the discovery that diffusion aluminide coatings can
serve as bond coats for thermal barrier coatings [23], the
goal is now to provide perfect resistance to spalling of
alumina to allow use of prime reliant thermal barrier
coatings with the capability of decreasing blade airfoil
cooling air flow to gain still higher turbine efficiency.
Lower sulfur contents of diffusion coatings appear to
improve alumina adherence [57]. Understanding of the
effects of platinum and various forms of active metals and
their oxides [58] will lead to further improvements in the

76

G.W. Goward / Surface and Coatings Technology 108 109 (1998) 73 79

use of diffusion aluminide systems as thermal barrier bond


coats.

3.2. Overlay coatings


Most research in this area has centered around the
active element effect. Lustman [59] summarized knowledge prior to 1950 on active element-doped nickel
chromium alloys, and from cyclic oxidation tests, postulated that the Cr 2 O 3 scales spalled less because of
pegging by active element oxides. Tien and Pettit [60],
came to similar conclusions from oxidation of FeCrAl
alloys doped with yttrium and scandium. Later work by
Giggins and Pettit [61] continued to support the pegging
mechanism and also showed that active elements minimize
void formation at the scale metal interface. They demonstrated that fine dispersions of active element oxides also
improve alumina adherence to NiCrAl and CoCrAl. In
1986, Ashary et al. [62] apparently abandoned pegging
explanations, and postulated only that active elements may
assist in the formation of strong chemical bonds between
oxide scales and the metal substrates. In 1983 Ikeda et al.
[63], and later Funkenbusch et al. [64], showed that small
amounts of impurities, such as sulfur, can segregate to the
aluminametal interface and lower the interface bond
strength. It followed that active elements must react with
tramp sulfur to prevent this segregation. Sulfur segregation
effects have since been confirmed by many independent
investigations. Pint [65], however, in a review of the
subject, contended that such effects are not consistent with
the fact that active element oxide dispersions also improve
scale adhesion. Pint also showed that active elements, such
as yttrium, diffuse through grain boundaries in alumina
scales, and lower scale growth rates. Pint proposed a
dynamic segregation theory to explain lower scale
growth rates and improved scale adhesionthis involves
segregation of active element ions to scale boundaries and
to the oxidemetal interface. Thus, the search proceeds
unabated while the practical benefits of the active element
effect have been commercially exploited for over 50 years.

3.3. Hot corrosion mechanisms


The importance of this phenomena within all gas
turbines warrants a separate, if brief, overview. It was
recognized in the early 1960s that surface wastage of
superalloy airfoils by hot corrosion was probably caused
by deposition of sodium sulfate (Na 2 SO 4 ), a model but
typical compound, formed by the reaction of sodium
chloride, from e.g. sea salt, and sulfur oxide byproducts of
combustion of sulfur-bearing fuels. Because microstructures representative of the process contained sulfides, for
many years the process was called sufidation. For this
reason much early work centered on the effects of sulfur
but it proved difficult to develop a coherent description of
the process on this basis alone. It was not until about 1969

that the beginnings of logical mechanisms which recognized the acidbase properties of molten Na 2 SO 4 began to
emerge [53,66,67]. Molten Na 2 SO 4 containing dissolved
sulfur trioxide (SO 3 ), from sulfur in fuels, is acidic,
whereas that deficient in SO 3 or high in oxide ion
concentration, is basic. Dissolution of protective oxides on
the acidic side is termed acidic fluxing and on the basic
side basic fluxing.
Zhang and Rapp [68] and others measured the solubilities in Na 2 SO 4 of oxides important to surface protection as functions of acidity, that is, activity of SO 3 in the
molten medium. While not absolutely complete, workable
hot corrosion mechanisms based on these phenomena are
now available to guide coating and superalloy development. These include the recognition of so called good
elements, chromium, platinum, silicon, and bad elements
molybdenum, tungsten, vanadium [53], to use or avoid in
development of hot corrosion resistant superalloys and
coatings. Some of these fit the acidbase theoryfor
example MoO 3 , WO 3 and V2 O 5 are acidic in molten
Na 2 SO 4 , and SiO 2 is relatively insoluble in acidic media.
Logical explanations of the beneficial effects of platinum
in coatings have not yet emerged. There is thus a working
science and technology base to guide the amelioration of
turbine corrosion in land and marine machines involving
intake air filtration, control of detrimental fuel impurities
(sulfur, vanadium, sodium, lead, phosphorus, etc.), and the
use of improved alloys protected by more corrosion
resistant coatings such as CoCrAlY and platinum-modified
aluminides.
In spite of this knowledge, the discovery in 1975 of
accelerated corrosion of CoCrAlY in a marine gas turbine
running at low power, and low airfoil metal temperatures,
was an unpleasant surprise [69]. Although it had been
anticipated by Cutler [67], who recognized that molten
sulfates, rendered acidic by SO 3 from fuel combustion,
cause severe corrosion of boiler and superheater steels at
about 6008C (11128F), much research was done [56,70]
before it was recognized that corrosion in turbines and
boilers had much in common. The phenomenon is now
universally recognized by the definitions of low-temperature Type II and high-temperature Type I hot corrosion.
Although Type II hot corrosion in aircraft gas turbines is
rare, it has been observed by the author, it being easily
identified by its unique microstructure [71]. It frequently
occurs in engines on aircraft operated over coastal and so
called island hopping routes. It should be widely recognized by now that chromium [33] and silicon [21] are
particularly beneficial in coatings [72] for protection
against Type II corrosion. It should also be universally
accepted that laboratory and burner rig tests involving
deposition of Na 2 SO 4 , which do not control partial pressure of SO 3 in the gas phase, are essentially useless for
evaluation of corrosion resistance of superalloys and
coatings. This is particularly true in the Type II temperature range: 600 7508C (110013508F).

G.W. Goward / Surface and Coatings Technology 108 109 (1998) 73 79

77

3.4. Thermal barrier coatings

3.5. Mechanical properties of coatings

In contrast to the science base of diffusion and overlay


coatings, all of the early development work to improve the
durability of these coatings was by empirical experimentation and evaluation by testing in burner rigs and actual
engines. Only in the last 510 years has research been
initiated to understand thermal barrier systems to aid in the
development of more durable coatings. Early improvements in durability involving the use of MCrAlY bond
coats [35] were based on prior knowledge of the better
oxidation and hot corrosion resistance, and oxide adherence of these coating alloys. Experimental work at NASA
[37] showed the improved durability of partially stabilized
zirconia (8%Y 2 O 3 ZrO 2 ) and Ruckle [38] demonstrated
the beneficial effects of micro-cracked and / or segmented
ceramic layers. The fact that EBPVD produces columnar
grain structures in MCrAlY coatings led Strangman [39] to
explore the properties of columnar grained EBPVD zirconia-based thermal barrier coatings. Knowledge that
active elements in nickel-base superalloys and / or platinum
modifications caused improved scale adherence in diffusion aluminide coatings must have surely led to the use of
these coatings as bond coats [23,52]. The large beneficial
effects possible in gas turbine efficiency by the use of
thermal barriers, e.g. up to 1708C (3068F) lowering of
blade metal temperature, have sparked a resurgence in
research on all aspects of scale adherence and failure
mechanisms involved in loss of ceramic layers. It is
anticipated that the extensive research now in motion will
yield incremental improvements in scale adherence toward
the goal of perfection required to attain prime reliant
status necessary to take full advantage of saved cooling air.
The latest summaries of some of this work are collected in
recent TBC Workshop Proceedings [73,74]. Near perfect
oxide adherence is implied both in a recent patent [75], and
an announcement by Pratt and Whitney [76] that a new
engine in development will use an advanced yttriumcontaining superalloy to which the TBC is applied directly
without the need for a bond coat.
Some research has been performed on possible effects of
hot corrosion on the ceramics used as thermal barriers, first
by Barkalow and Pettit [77] and also by Jones et al. [78].
In conditions most likely to be encountered in industrial
turbines, destabilization of the preferred crystal structures
occur by leaching out of stabilizers, such as Y 2 O 3 , by
molten acidic sodium sulfate (Na 2 SO 4 SO 3 ). This does
seem less likely to occur in industrial engines fired with
clean natural gas at the higher temperatures required for
improved efficiency.
Because of the dependence of airfoil integrity on TBCs
there has been considerable effort to develop life prediction systems [7981] to aid in airfoil design. Parallel
efforts for development of design systems, based on
constitutive properties of TBCs, are subjects of several
current programs [73].

Early work on mechanical properties focussed on possible effects of coatings on superalloy properties such as
stress rupture, and isothermal low and high cycle fatigue.
With the exception of high cycle fatigue debits [82], few
real effects on these properties have been observed. Even
though some such debits have been observed in laboratory
tests, this author knows of no high cycle fatigue blade
failures directly assignable to coatings. Indeed, real high
cycle fatigue failures signal design problems, and are
alleviated when the problems are corrected. Instances of
debits in stress rupture properties have been documented
for some aluminide coateds on nickel-base alloys, such as
Udimet 520 and Rene 80 [83,84] which are particularly
dependent on grain boundary carbides for strength. The
debits are caused by depletion of carbon from M 23 C 6 grain
boundary carbides by strong carbide formers, e.g. tantalum, tungsten, molybdenum, which concentrate in the
diffusion zone by rejection from NiAl during the formation
of that zone [40]. While the phenomenon is real in
laboratory tests, no blade failures attributed to the effects
have been reported in the literature. The author is aware of
failure of coated blades made from a similar alloybut it
was not documented in the open literature. The effect is
minimal in stronger cast polycrystalline alloys and has not
been observed in directionally solidified and, of course,
single-crystal alloys.
Significant thermal fatigue problems are rare in uncooled airfoilsin some cases coatings increase thermal
fatigue resistance of polycrystalline superalloys in temperature regions above about 8608C (14728F) [85]. This is
attributed to coating protection of grain boundary crack
initiation sites. With the advent of highly cooled blades and
vanes, however, serious thermal fatigue problems emerged.
Early in the use of diffusion aluminide coatings, it was
recognized that they were inherently brittlean expected
property of body-centered cubic NiAl, the predominant
phase in the coatings. In 1970 the concept of transition
from brittle to ductile behavior of these coatings as
temperature is increased was introduced [86]. Aluminide
and some CoCrAlY coatings exhibited brittle to ductile
transitions at around 7008C (13008F). It was therefore
anticipated that if thermalmechanical tensile strains were
large enough below this transition temperature, brittle
cracking of coatings could occur, followed by crack
propagation into base alloys. In a seminal paper, Linask
[87] described the results of an analytical study using
fracture mechanics principles to model turbine airfoil
cracking. It was shown that cracks initiated in a CoCrAlY
coating when tensile strains, due to thermalmechanical
stresses, peaked at about 200C (3928F) in a thermally
cycled TF30 engine air cooled first stage blade. The strains
peaked upon engine deceleration, at well below the brittleto-ductile transition temperature of the CoCrAlY coating,
and exceeded the value required to crack the coating in a

78

G.W. Goward / Surface and Coatings Technology 108 109 (1998) 73 79

brittle fracture mode. The cracks then propagated into the


substrate alloy. The fracture mechanics approach and
thermalmechanical fatigue (TMF) testing have been
further refined over a period of years as airfoil design tools
and as guides to coating and alloy development [88].
NiCoCrAlY overlay coatings, formulated by adding critical
amounts of cobalt to NiCrAlY, were specifically developed
to have some ductility in the 3008C (6008F) temperature
range. Improvement in thermal fatigue resistance of
NiCoCrAlY-coated superalloys was confirmed with TMF
testing and by subsequent engine tests [31,89]. Although
not immune to TMF cracking in ductile modes, the
coatings provide significant life extension over prior
compositions.
Holmes and McClintock [90] and Busso and McClintock [91] devised a useful laboratory thermal fatigue test
using induction heating and air-blast cooling of stepped
disc specimens to induce controllable thermal strains in
coated superalloys. Application of the method provided
much useful information on crack initiation and propagation, along with interacting effects of coatingalloy interdiffusion and oxidative degradation of diffusion aluminide
[90] and NiCoCrAlY overlay coatings [91] on singlecrystal nickel-base superalloys.
Because thermal expansion mismatch between
NiCoCrAlY and superalloys, and relatively low yield and
creep strength are important deficiencies of NiCoCrAlY in
thermal-mechanical fatigue, some efforts have been made
to improve these properties by additions of refractory
elements and / or oxide dispersions [9294]. Experimental
results seem promising but no practical applications have
so far been reported. As previously described, thermal
barrier coatings significantly reduce TMF strains in cooled
airfoils and it is therefore from improvements in durability
of these coatings that major improvements in thermal
fatigue life result.

4. Summary and conclusions


The use of coatings on gas turbine airfoils, in concert
with improved alloys and design technology, has unfailingly supported increases in either or both efficiency and
service life in varieties of turbomachinery. Early coatings
were simple diffusion aluminides applied principally by
pack cementation and later by gas-phase processes. Additions of silicon, chromium and platinum improved the
service lives of these coatings. Parallel research in oxidation and hot corrosion mechanisms, coating processing,
and mechanical properties of coatingalloy systems have
provided sound bases to guide the intelligent use of the
coatings over several decades. Research aimed at understanding protective oxide adherence supported the development of MCrAlY overlay coatings which offered the
ability to tailor properties for practical uses in a variety of
gas turbine applications. Early uses of zirconia-based

thermal barrier coatings on burner components signaled


possibilities of the coatings to substantially extend turbine
airfoil lives and / or improve engine efficiencies. MCrAlY
compositions as bond coat under controlled ceramic structures in plasma-deposited thermal barriers have durability
sufficient for extending stationary airfoil service lives. In
the late 1980s, electron beam vapor-deposited zirconia
coatings over MCrAlY or diffusion aluminide bondcoats
provided durability sufficient for use on rotating blades. A
complete understanding of oxide adherence is still elusive
and is one of the most important areas of research, for both
coatings and superalloys, to provide further advances in
engine efficiency and service life. Research and development on coating processing for diffusion, overlay, and
thermal barrier systems deserves equally strong support.
As the millennium approaches the future for coating
science and technology is brighter than ever.

References
[1] T. Van Aller, US Patent 1,155,974 (1915).
[2] H.B.C. Allison, L.A. Hawkins, General Electric Rev. 17 (1914)
947951.
[3] E.G. Gilson, US Patent 1,091,057 (1914).
[4] E. Ruder, Trans. Am. Electrochem. Soc. 21 (1915) 253261.
[5] C.B. Meher-Homji, Mechanical Eng. September (1997) 8891.
[6] R.P. Seelig, R.J. Steuber, High Temp.-High Press. 10 (1978) 207
213.
[7] E.S. Nichols, J.A. Burger, D.K. Hanink, Mech. Eng. (1965) 5256.
[8] A.D. Joseph, US Patent 3,102,044 (1960).
[9] R.S. Parzukowski, Thin Solid Films 45 (1977) 349355.
[10] G. Gauje, R Morbioli, in: S.C. Singhal (Ed.), High Temperature
Protective Coatings, The Metallurgical Society of AIME, Atlanta,
GA, 1983, pp. 1326.
[11] B.M. Warnes, D.C. Punola, Surf. Coat. Technol. 9495 (1997) 16.
[12] P.M. Galmiche, US Patent 3,900,613 (1975).
[13] R.S. Parzuchowski, R.B Benden, US Patent 4,148,275 (1979).
[14] J.E. Restall, M.I. Wood, Mater. Sci. Tech. 2 (1986) 225231.
[15] F.D. Kelley, Trans. Am. Electrochem. Soc. 43 (1923) 351370.
[16] R.L. Samuel, M.L. Lockington, Metal Treatment and Drop Forging
18 (1951) 354359, 407414, 440444, 495506, 543548; and 19
(1952) 2732, 8185.
[17] R. Drewett, Anti-Corrosion Methods Mater. 16 (1951) 543548.
[18] T. Gibson, US Patent 2,809,127 (1957).
[19] R. Burgel, Mater. Sci. Tech. 2 (1986) 302308.
[20] H.W. Brill-Edwards, US Patent 3,801,353 (1974).
[21] R. Bauer, H. Grunling, Thin Solid Films 1 (1982) 320.
[22] G. Lehnardt, H. Meinhardt, Electrodeposition Surf. Treatment 1
(1972) 189193.
[23] T.E. Strangman, US Patent 5,514,482 (1996).
[24] R. Bianco, R.A. Rapp, in: K.H. Stern (Ed.), Metallurgical and
Ceramic Coatings, ch. 9, Chapman and Hall, London, 1993.
[25] G.W. Goward, L. Cannon, J. Eng. Gas Turbines Power 110 (1998)
150154.
[26] G.W. Goward, L.L. Seigle, ASM Handbook, vol. 5, Surface Engineering, 1994, pp. 611620.
[27] W.T. Griffiths, L.B. Pfeil, UK Patent 459 (1937) 848.
[28] C.S. Wukusik, J.F. Collins, Mater. Res. Standards 4 (1964) 637
646.
[29] F.P. Talboom, J. Grafwallner, US Patent 3,542,5330 (1970).
[30] D.J. Evans, R.C. Elam, US Patent 3,676,085 (1972).

G.W. Goward / Surface and Coatings Technology 108 109 (1998) 73 79


[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]
[42]
[43]
[44]
[45]
[46]
[47]
[48]
[49]
[50]
[51]
[52]
[53]
[54]
[55]
[56]
[57]
[58]
[59]
[60]
[61]
[62]
[63]
[64]

R.J. Hecht, G.W. Goward, R.C. Elam, US Patent 3,928,026 (1975).


D.K. Gupta, D.S. Duval, US Patent 4,585,481 (1986).
K.L. Luthra, J.H. Wood, Thin Solid Films 119 (1984) 271280.
N.S. Cheruvu, T.J. Carr, J. Dworak, J. Coyle, J. Metals 48 (1996)
3438.
G.W. Goward, D.A. Grey, R.C. Krutenat, US Patent 4,248,940
(1981); Re. 33,876 (1992).
S.M. Meier, D.K. Gupta, Trans. AIME 116 (1994) 250257.
S. Stecura, NASA TM-78976 (1979).
D.L. Ruckle, Thin Solid Films 64 (1979) 327.
T.E. Strangman, US Patent 4,321,311 (1982).
G.W. Goward, D.H. Boone, Oxid. Met. 3 (1971) 475495.
M.M.P. Janssen, G.D. Rieck, Trans. Met. Soc. AIME 239 (1967)
13721385.
S. Shankar, L.L. Seigle, Met. Trans. A 9A (1978) 14681476.
P.N. Walsh, in: Chemical Vapor Deposition, 4th International
Conference, The Electrochemical Society, 1973, pp. 147168.
S. Levine, R.M. Caves, J. Electrochem. Soc. 121 (1974) 1051
1064.
L.L. Seigle, B.K. Gupta, S. Shankar, A.K. Sarkel, NASA Contract
Report 2939 (1978).
N. Kandasamy, L.L. Seigle, Thin Solid Films 84 (1981) 1727.
B.K. Gupta, L.L. Seigle, Thin Solid Films 83 (1980) 365371.
W. Johnson, K. Komarek, E. Miller, Trans. Met. Soc. AIME 242
(1968) 16851690.
F.S. Pettit, Trans. Met. Soc. AIME 239 (1967) 12961305.
C.S. Giggins, F.S. Pettit, J. Electrochem Soc. 118 (1971) 1782
1790.
J.J. Smialek, C.E. Lowell, NASA TM X68274 (1974).
P. Aldred, Natl. Aerospace Eng. and Mfg. Conf., SAE Paper
751049, Los Angeles 1975.
J.A. Goebel, F.S. Pettit, G.W. Goward, Met. Trans. 4 (1973) 261
278.
E. Godlewska, K. Godlewski, Oxid. Met. 22 (1984) 117131.
W.T. Wu, A. Rahmel, M. Schorr, Oxid. Met. 22 (1984) 5981.
R.H. Barkalow, F.S. Pettit, US Navy Contract Report FR-10225,
1978.
W.Y. Lee, Y. Zhang, I.G. Wright, B.A. Pint, P.K. Liaw, Met. Trans. A
29 (1998) 833.
B.A. Pint, Mater. High Temp. 14 (1997) 403412.
B. Lustman, Trans. AIME 188 (1950) 995.
J. Tien, F.S. Pettit, Met. Trans. 3 (1972) 15871599.
C.S. Giggins, F.S Pettit, Air Force Report, ARL TR 75-0234, June,
1975.
A. Ashary, G.H. Meier, F.S. Pettit, Air Force Contract 80-0089,
April, 1986.
Y. Ikeda, Y.K. Nii, K. Yoshihara, Trans. Jpn. Inst. Met. 24 (1983)
207.
A.W. Funkenbusch, J.G. Smeggil, N.S. Bornstein, Met. Trans. 16A
(1985) 1164.

[65]
[66]
[67]
[68]
[69]

[70]
[71]

[72]
[73]
[74]
[75]
[76]
[77]

[78]
[79]
[80]
[81]
[82]
[83]
[84]
[85]
[86]
[87]
[88]
[89]
[90]
[91]
[92]
[93]
[94]

79

B.A. Pint, Oxid. Met. 45 (1996) 137.


N.S. Bornstein, M.A. DeCrescente, Trans. AIME 245 (1969) 1947.
A.J.B. Cutler, J. Appl. Electrochem. 1 (1971) 1926.
Y.S. Zhang, R.A. Rapp, Corrosion 43 (1987) 348352.
D.J. Wortman, R.E. Fryxell, I.I. Bessen, in: Proceedings of the Third
Conference on Gas Turbine Materials in a Marine Environment,
Section V, Paper 11, Bath, England, 1976.
K.L. Luthra, Met. Trans. A 13A (1982) 18531864.
R.H. Barkalow, G.W. Goward, in: NACE 6, High Temperature
Corrosion, National Association of Corrosion Engineers, San Diego,
CA, 1981, pp. 502506.
G.W. Goward, J. Eng. Gas Turbines Power 108 (1986) 421425.
Proceedings of Thermal Barrier Coating Workshop, NASA Conference Publication 3312, Cleveland, OH, 1995.
Proceedings of Thermal Barrier Coating Workshop, TBC Interagency Coordination Committee, Cincinnati, OH, 1997.
J.C. Schaffer, W.H. Murphy, W.B. OConnor, B.J. Nagaraj, H.V.
Vakil, US Patent 5,538,796 (1996).
S.W. Kandebo, Aviation Week 148 (1998) 3435.
R.H. Barkalow, F.S. Pettit, in: J.W. Fairbanks, J. Stringer (Eds.),
Proceedings of First Conference on Advanced Materials for Alternate-Fuel Capable Directly Fired Heat Engines, Castine, ME, 1979,
p. 704.
R.L. Jones, S.R. Jones, C.E. Williams, J. Electrochem. Soc. 132
(1985) 14981501.
T.E. Strangman, J. Neumann, A. Liu, NASA Contractor Report
179648 (1987).
R.V. Hillery, B.H. Pilsner, R.L. McKnight, T.S. Cook, M.S. Hartle,
NASA Contractor Report 180807 (1988).
S.M. Meier, D.M. Nissley, K.D. Sheffler, NASA Contractor Report
189111 (1991).
H.W. Grunling, K. Schneider, L. Singheiser, Mater. Sci. Eng. 88
(1987) 177189.
R. Castillo, K.P. Willett, Met. Trans.A 15A (1984) 229236.
S.B. Kang, Y.G. Kim, Mater. Sci. Eng. 83 (1986) 7588.
A.T. Santhananam, C.G. Beck, Thin Solid Films 73 (1980) 387.
G.W. Goward, J. Met. (1970) 19.
I. Linask, ASME Paper 75-GT-79, 1975.
T.E. Strangman, S.W. Hopkins, Am. Ceram. Soc. Bull. 55 (1976)
304307.
T.E. Strangman, Ph.D. Thesis, University of Connecticut, 1978.
J.W. Holmes, F.A. McClintock, Met. Trans. A 31A (1990) 1209
1222.
E.B. Busso, F.A. McClintock, Mater. Sci. Eng. A161 (1993) 165
179.
T.E. Strangman, S.J. Vonk, Canadian Patent 1,294,155 (1992).
N. Czech, F. Schmitz, W. Stamm, Surf. Coat. Technol. 76 (1995)
2833.
T.A. Taylor, D.F. Betteridge, Surf. Coat. Technol. 8788 (1996)
914.

Das könnte Ihnen auch gefallen