throbber
UTC 2007
`General Electric v. United Technologies
`IPR2016-01289
`
`1
`
`

`
`384
`
`CLARKE
`
`
`Turbine
`
`
`
`AirfoilGas,Ternperatu1'eCapability
`
`_
`COHVECEIOH 4“
`impingement
`+film cooling
`film
`mating \
`
`I
`convectimi
`_
`Coflling
`
`
`
`
`
`
`
`
`
`
`Coo1ing+
`EB_PVD TBC
`
`L
`
`L
`
`I
`I
`..iEfi'e’C10fc
`
`:
`
`I
`
`I
`
`
`
`us
`
`33‘
`
`tlncuoieci
`SW4
`
`
`
`I
`
`
`
`I 2000
`
`
`‘caslt
`I
`~
`
`
`/‘T? eqtniaxed
`
`
`wrought
`
`i940
`1960
`
`Llncooled Superalloya
`
`I 1980
`
`Year
`
`Increase in turbine airfoil temperature over the last six decades through com-
`Figure 1
`binations of materials advances and associated developments in cooling techniques.
`Since this diagram was constructed, the shaded region has extended to the present year,
`and the use of uncooled silicon nitride remains for the future.
`
`due to flame out, and as a means to even out local temperature gradients. Indeed,
`in some cases, the use of a TBC has simplified the design of blades by minimizing
`thermal distortions ofthe blade. However, undoubtedly the biggest benefit of TBCS
`has been to extend the life of alloy components in the hottest sections in an engine
`by decreasing their surface temperatures.
`Present day TBCs generally consist of a yttria—stabilized zirconia (YSZ) coat-
`ing deposited onto an oxidation—resistant bond—coat alloy that is first applied to a
`nicl<el—based sup eralloy component (Figure 2). In diesel engine applications where
`the temperatures are usually lower, the YSZ coating is generally applied directly
`onto the alloy. Two main types of coating are in use. For relatively small compo-
`nents such as blades and vanes in aerospace turbines, the coatings can be applied
`by electron—beam physical vapor deposition (EB—PVD). For larger components
`such as the combustion chambers and the blades and vanes of power generation,
`stationary turbines, the coatings are usually applied by plas1na—spraying (PS). In
`many respects, the choice of materials and their production represent a mature
`materials technology. While improvements in their capabilities continue, there is
`a growing realization that new TBC systems will be required for the next genera-
`tion turbines presently being designed. To set the stage for coming developments,
`we first review the selection of materials used in present YSZ coatings, some
`of the new insights that have been gained in understanding how YSZ coatings
`
`2
`
`

`
`THERMAI %ARRTER COATINGS‘
`
`“I385
`
`
`
`(F
`{L/t
`I}
`5%
`i E
`
`£3
`
`Q;
`
`5
`g_
`Q}
`2 E 2
`
`Schematic illustration of a 1‘ SC
`Figure 2
`and the associated bond—coat on a superalloy
`in a thermal gradient.
`
`Superalltéji
`
`PF???‘
`Elmiriliiig Air
`
`fail, and then describe approaches to the development of the next generation TBC
`systems.
`
`PRINCIPAL REQUIREMENTS OF A THERMAL
`BARRIER COATING
`
`The turbine designers’ primary requirement of a TBC is that it have a low ther-
`mal conductivity and, for rotating components, preferably also a low density to
`minimize centrifugal loads. At the materials design level this translates into three
`additional requirements. First, the material must have strain compliance so as to
`withstand the strains associated with thermal expansion mismatch between the
`coating and the underlying alloy on thermal cycling. The use cycle, both the 1nax—
`imum temperature and the times at temperature, of course, varies between aircraft
`and power generation turbines, but nevertheless the coating must accommodate
`the large strains associated with thermal cycling. The need for strain compliance
`is illustrated in Figure 3, where the thermal expansion coefficients of zirconia,
`
`3
`
`

`
`386
`
`CLARKP I LJVI
`
` Q3E
`
`NJ3
`
`
`
`ThermalExpansionCoefficient(ppmC"')
`
`
`
`
`
` 10
`1000
`
`Thermal Conductivity (WlmK)
`
`Figure 3 The thermal expansion coeflicients and thermal conductivity of
`a range of materials illustrating the differences in thermal expansion and
`conductivity of the principal components in TBC systems.
`
`alumina, and a number of alloys including nickel—based superalloys are cross-
`plotted against thermal conductivity. In the absence of any strain compliance, for
`instance due to a decreased elastic modulus, the large elastic mismatch would gen-
`erate Veiy large stresses and lead to spontaneous failure on cooling. Second, the
`coating material must exhibit thermodynamic compatibility with the oxide, usually
`aluminum oxide, formed on the bond-coat alloy at high temperatures. Third, with
`the continual quest to run engines at higher temperatures and the increasing dif-
`ficulty of increasing the metal temperature, it is increasingly likely that designers
`will seek “prime reliant” coatings, namely ones that can be used with assurance
`that they will not fail. Prime reliant thermal coatings are ones that are necessary
`to prevent the temperature of the metal from exceeding its maximum temperature,
`much in the same way that the tiles on the space shuttle prevent the underlying
`aluminum airframe from being exposed to temperatures in excess of their melting
`temperature on re~ent1y.
`Because Weight is at a premium in aircraft engines, thin coatings with the lowest
`possible thermal conductivity are required. In contrast, in stationary, ground—bas ed
`engines where weight is less of a consideration, a desired temperature drop can be
`
`4
`
`

`
`TH *RMAlI BARRIER COATINGS
`
`387
`
`achieved through simply increasing the '1' ’C thickness. In practice, in components
`in both types of engine, the thickness of the T 3C usually is varied from place to
`place to provide the desired thermal insulation.
`Jrosion of the coating by both ingested pa ticles, such as sand, from the op-
`erating environment and particles that come loose from the combustor liners as it
`degrades is a perennial source of concern, especially when the particles are large
`enough to cause impact damage of the coating. In some cases, inborn fine parti-
`cles, primarily dust and sand, melt into the coating as a wetting silicate while it is
`hot and can degrade the coating. These silicates, usually variants of Si-Al-Mg—Ca
`oxides that are the principal elements in sands, are often referred to collectively as
`CMAS.
`
`A recently recognized requirement of many materials exposed to high tempera-
`tures in gas turbines is a long-tenn stability in the presence of steam. This is partly
`a direct result of the generation of water during the combustion process, but in a
`number of designs it is a consequence of the use of steam injection to enhance
`turbine efliciency. Little is known about the effects of long-term exposure to steam
`on turbine materials. However, tests have revealed that many silicon—based com-
`pounds, including SiC, are unstable to the formation of volatile SiO, which results
`in the slow retraction of the material as evidenced by the reduction in thickness
`of components over long operating periods. This active oxidation and evaporation
`phenomenon precludes the use of silicon compounds in coatings unless protected
`by another coating.
`More difiicult to design against are the effects of corrosion, especially airborne
`species and those, such as sulfur and vanadium, in the fuel itself. The majority of
`land-based turbines operate on natural gas, but there is increasing interest in using
`alternative fuels, such as coal gas, that are much dirtier. The consequences ofusing
`such alternative fuels and their effects on coatings are only now beginning to be
`investigated.
`
`THE THERMAL BARRIER COATING SYSTEM
`
`From a materials engineering perspective, it is necessary to consider the TBC as
`an integrated materials system rather than simply a thermally insulating material
`coating on a structural alloy component. A representative cross-section of a com-
`mercial coating, shown in Figure 4, illustrates the multilayered nature of a coating
`after high—te1nperature exposure. There are three principal layers in addition to the
`superalloy and the low—conductivity coating. Between the alloy and the coating is
`the bond—coat, so called because in the initial development stages in producing a
`viable coating, it was found that the superalloy had to be first covered with a bond-
`coat to ensure that the YSZ coating remained adherent upon oxidation. Between
`the bond—coat and the YSZ coating—sometimes referred to as the overcoat—is the
`oxide formed during high—temperature exposure. Finally, during the formation of
`the bond—coat and the YSZ coating as well as subsequently during use, a reaction
`layer forms as a result of inter—diffusion between the bond—coat and the superalloy.
`
`5
`
`

`
`CI ARI 4
`
`T EVI
`
`
`
`Figure 4 Cross—section of a TBC deposited by electron beam evaporation. Note the
`columnar microstructure of the zirconia coating. The white band is a reaction layer
`formed by interdiffusion during use between the Al-rich bond—coat above and the Ni-
`
`rich superalloy, below. The thermally grown oxide (TGO) is too thin to be discernible
`in this micrograph.
`
`In practice, as mentioned above, there are two distinct types of zirconia coat-
`ings reflecting different approaches to creating the strain compliance essential to
`withstand thermal cycling. The two types of coatings are EB—PVD coatings and
`plasma-sprayed coatings. In EB—PVD coatings, the lateral strain compliance re-
`sults from the columnar structure and inter—columnar gaps produced by rotation of
`the component during deposition. The columnar structure can be seen in the mi-
`crograph of Figure 4. Transmission electron microscopy reveals that the individual
`columns also contain microscopic porosity that reduces the thennal conductivity
`of the coating. In plasma-sprayed coatings, the lateral strain compliance and re-
`duced thermal conductivity is conferred by the incorporation of porosity between
`“splats” of successively deposited material. This porosity is illustrated in Figure 5,
`where the splats in a plasma spray bond coat have preferentially oxidized and
`consequentially appear as dark veins.
`Two major classes of bond—coat alloys have also evolved over the years, but
`both were developed to form an aluminum oxide (o¢—Al2O3) on exposure to air
`at high temperatures. This is important for several reasons. One is that A1203
`is phase compatible with YSZ, ensuring long—term theimodynamic stability of
`the coating. Uncoated, the majority of nicl<el—based superalloys form complex,
`
`6
`
`

`
`TH ~RMAl %ARRIER COATINGS
`
`389
`
`
`
`Figure 5 Cross—section of a TBC deposited by plas1na—spraying. The plate-
`like porosity is evident in the coating as the dark veins in the center of the
`micrograph.
`
`multilayered nickel oxide, nickel—chromiu1n spinels and chromium oxide, in addi-
`tion to alumina, and these are not thermodynamically stable with YSZ (2). Further-
`more, alumina is usually considered to be the slowest growing high—te1nperature ox-
`ide on account of it having the smallest oxygen diffusivity (3). The rationale for the
`selection ofbond-coat alloys is really a subj ect of another review, but the bond-coat
`has to perform a number of disparate functions. It must provide a bond between the
`deposited TBC and the underlying alloy. In the early days of TBC development, the
`bonding to the alloy was a major concern, particularly plasma—sprayed, hence
`the term bond-coat. Because zirconia is a fast—ion oxygen conductor, the bond-coat
`must also be able to form a protective, stable, and slow—growing oxide to prevent
`oxidative attack of the alloy. As is described below, one of the principal forms of
`failure is associated with failure of the protective aluminium oxide. The bond-coat
`must also have suflicient morphological stability so that on heating and cooling,
`as well as at high temperature, it does not distort and introduce incompatibilities
`that can also cause the introduction of interface defects.
`
`The two classes of bond-coat alloys that have been developed are the platinu1n—
`modified nickel aluminide (PtNiAl) and MCrAlY alloys (M here refers to one or
`more of the elements Co, Ni, and Fe). The selection of these two classes of alloys
`is largely based on their prior use as oxidation— and co1rosion—resistant coatings for
`protecting high—temperature alloys before the advent of TB Cs. For instance, the
`
`7
`
`

`
`390
`
`CLARKE : LEVI
`
`PtNiAl was originally developed as an alternative oxidation—resistant coating for
`protecting alloys at higher—temperature operation than the MCrAlY alloys available
`at the time.
`
`Different methods of applying the bond-coat alloys have been developed largely
`to meet production goals. Typically, PtNiAl bond-coats are formed by first electro-
`depositing Pt onto the superalloy component and then annealing it in an alu1ninu1n—
`rich vapor atmosphere. In this second step, aluminum diffuses into the surface of
`the alloy while nickel diffuses out where it reacts with the aluminum and platinum
`to form the PtNiAl aluminide coating. Depending on the quality of the coating
`required, the aluminum is provided in a pack—process or in a CVD reactor from an
`AlCl3 source. In contrast, the MCrAlY coatings are commonly deposited by one of
`a number ofvariants ofplasma—spraying. These processes are particularly attractive
`for coating large components and are, of course, cheaper than EB deposition. Also,
`as plasma-spraying does not involve a diffusion process, thicker bond-coats can be
`deposited than with the aluminizing process used to form the PtNiAl bond coats.
`It remains uncertain at this time which ofthese coating types is best for different
`applications. In large part this is because it is not yet known which combination
`of materials properties leads to the longest, high-temperature life of the coating.
`To provide the largest reservoir of aluminum one would expect that the thicker the
`bond-coat and the higher its aluminum content the better. However, one would also
`expect that the bond-coat should have as large a yield stress as possible at high tem-
`perature with as closely matched thermal expansion mismatch with the superalloy
`as possible to avoid thermal expansion mismatch stresses on thermal cycling.
`
`FAILURE MECHANISMS
`
`Investigation of the ways in which present YSZ coatings fail has provided consid-
`erable insight into the underlying mechanisms that limit coating life. Although, as
`with failure analysis in other areas of complex material systems, there are many
`complications, the findings nevertheless point toward methods of producing coat-
`ings that can withstand longer lives at temperature and higher use temperatures.
`One of the chronic problems is that the life of present TBC coatings invariably
`shows a wide distribution, with a high proportion of the population clustered about
`a median value but with a significant proportion failing at much earlier times.
`There is substantial circumstantial evidence to suggest that many of the TBC
`failures are associated with the oxidation of the bond-coat (4). Indeed, a number
`ofmanufacturers are believed to use an oxidation criterion as a basis for predicting
`average life. One such criterion is the combination of time and temperature to lead
`to a critical thickness of the TGO1. Another, embodied in the Coatlife software,
`is an aluminum depletion criterion based on the combined time and temperature
`
`1The concept of a critical thickness, of the order of 6 mn at ll00“C, appears to have
`originated from observations of coatings that failed under isothermal testing conditions.
`
`8
`
`

`
`TH * RMAL BARRIER COATINGS
`
`391
`
`for the concentration of aluminum at the bond—coat surface to fall below a critical
`
`value. In the case of MCrAlY bond-coats, the rationale for this is that when the
`Al concentration falls to ~23 a/o, aluminum oxide is no longer the thermodynamic
`preferred phase and other oxides, notably spinels, form (5). These other oxides
`do not form such a protective scale, and consequently the alloy oxidizes faster. In
`addition, the formation of these oxides is associated with an increase in volume
`that can be disruptive and possibly have lower fracture energies, although this has
`yet to be unequivocally demonstrated. Nevertheless, there are reports that when
`the bond—coat is porous, and at low—oxidation temperatures, failure follows such
`aluminum depletion (6).
`Although related to the oxidation behavior of the bond—coat, neither the con-
`cept of a critical thickness nor aluminum depletion can account for the wide dis-
`tribution in failure lives, especially under thermal cycling conditions. Indeed, in
`the majority of materials examined after failures above about l000°C, the alu-
`rninum concentration, although depleted somewhat, has not fallen to the critical
`Value (7). Similarly, the short—lived coatings have failed before the TGO thickness
`has reached the thickness of its counterparts that have shown the longest lives.
`Together these findings indicate that failure occurs due to extrinsic factors arising
`during oxidation.
`The prevailing mode of failure is one in which part of the coating buckles and
`spalls away from the alloy, typically on cooling down to room temperature (8, 9).
`A typical buckling failure, in this case nucleated from the edge of a test coupon, is
`illustrated in Figure 6. Such buckling and subsequent spallation is a common mode
`of failure of all films and coatings under compression, generally associated with
`the development of compressive residual stresses in the coatings as a result of the
`difference in thermal expansion coefiicient between the coating and the underlying
`alloy. The mechanics of the failure by buckling of a thin, elastically isotropic
`film under compression from a flat surface is well understood (10), provided an
`unbonded region of a critical size, db, exists at the interface (Figure 7). For a fixed
`film thickness and residual stress, the stress at which buckling will occur is given
`by the relation:
`
`/E 48
`
`0 :. ——
`db
`
`h
`
`2
`
`.
`
`1
`
`.
`
`Thin film buckling is entirely analogous to the standard Euler buckling condition
`ofa column—a bifurcation phenomenon. The striking feature ofthis relation is that
`the flaw size depends linearly on the thickness of the film.2 Since even the thinnest
`of TBC is over l00 um thick, the critical size to which an interface flaw must grow
`before buckling can occur can be several millimeters. As interface separations of
`this large size are not usually present after coating, one of the major unresolved
`
`2Because they are designed to have in—plane strain compliance, TBCs are not usually
`isotropic elastic solids. The buckling condition is modified from that in Equation l
`to
`account for the elastic anisotropy.
`
`9
`
`

`
`392
`
`CLARKE E T
`
`(Top) Incipient buckling of a TBC coating viewed under reflected light.
`Figure 6
`(Bottom) The failure surface revealed by spallation of the TBC consists of a mixture
`
`of local failure between the TGO and the bond—coat (appearing dark) and in the TBC
`itself (light regions).
`
`questions is how interface separations first form and then grow to such a large
`size. Such progressive failure consisting of nucleation of local interface separation
`and their subsequent growth has indeed been observed (1 l). Recent mechanics
`calculations have shown that interface perturbations from flatness can decrease
`the critical size at which buckles can initiate and then grow in size to form a spall
`
`10
`
`10
`
`

`
`TH *RMAll %ARRI * R COAIINGS
`
`393
`
`After;
`
`Substrate
`
`
`Figure 7 Schematic illustration of the buckling of a compressed film above
`
`a pre—existing defect of diameter db.
`
`(12). Nevertheless, localized flaws must first initiate and then grow for spallation
`failure to occur. Understanding the nucleation of these flaws, their growth, and
`progressive linking together is essential before realistic models for predicting life
`can be developed.
`Insight into the formation of flaws comes from microstructural examination of
`coating cross-sections after high—temperature exposure but prior to spallation. Four
`examples are shown in Figure 8, each from a YSZ TBC—coated PtNiAl bond—coat
`(13). In each case, the coating was deposited conformally onto the surface of a
`flat bond-coat so that the coating/bond—coat interface was initially flat and intact.
`As three of the inicrographs illustrate, the surface of the bond—coat roughens
`and separations form with the TBC even though the bottom surface of the TBC
`remains flat. These separations are the interface flaws that progressively grow in
`size and link together with adjacent ones to allow buckling and spallation (9, ll).
`Roughening is more pronounced with thermal cycling but also occurs, albeit more
`slowly, on isothermal exposures (14). The growth ofthese separations with thermal
`cycling can now be monitored by luminescence piezospectroscopy, as described
`
`11
`
`11
`
`

`
`mM..u
`
`W!) ‘H Jsmrn
`
`12
`
`12
`
`

`
`TH *‘RMAlI BARRIER COATINGS
`
`395
`
`in the next section, which suggests that it has the potential to be used as a viable
`N )J:’ tool (1 1, 15).
`The micrographs in Figure 8 raise the question as to the underlying mechanisms
`responsible for the observed roughening. At least two new mechanisms have now
`been identified that can lead to such roughening. The roughening has been at-
`tiibuted to a “ratcheting” phenomenon motivated by the lateral compressive stress
`in the growing TGO and facilitated by thermal cycling (9). Measurements indicate
`that as the TGO grows in thickness with oxidation, it also concurrently develops
`a compressive stress (l6, 17). If it were free to expand it would decrease its coin-
`pressive stress but because it is attached to the bond-coat, the only way in which
`it can decrease its elastic strain energy is by undulating (Figure 9). In this way, its
`length increases and it remains attached to the alloy. This undulation requires the
`alloy to deform to accommodate the undulation, and the oxide must also deform
`concurrently. According to the ratcheting mechanism, this accommodation is by
`plastic deformation of both the TGO and bond-coat during thermal cycling. As
`the lateral growth of the thickening oxide continues during the high-temperature
`portion of the thermal cycles, it continues to generate compressive stress that is
`relaxed by ratcheting during the thermal cycle so the process is ongoing. Many of
`the essential features of the mechanism have been substantiated by finite element
`computations (18) and are consistent with observations of the increase in length
`of the TGO as the surfaces roughen.
`Another new mechanism shown to cause roughening is the surface displace-
`ment associated with volumetric changes in the bond-coat as aluminum depletion
`occurs. This roughening is illustrated in Figure 10, together with etched cross-
`sections revealing the presence of both y’ and ,6 phases in the bond-coat (14).
`After aluminizing and after YSZ deposition, the PtNiAl bond-coat is chemically
`homogeneous and has the ,8—NiAl (B2) crystal structure. After high—temperature
`exposure, the initially flat bond-coat is rumpled and etching reveals that the bond-
`coat has partially transformed to y ’ —Ni3Al. In addition, the remaining fl —NiAl phase
`regions often have the characteristic lath structure of a maitensite. These two ob-
`servations can be understood as being the result of aluminum depletion from the
`bond-coat and concurrent enrichment of nickel from the underlying superalloy, a
`classic example of interdiffusion. This change in composition is illustrated using
`the binary NiAl diagram in Figure 11. As aluminum is depleted, the average com-
`position of the bond-coat becomes increasingly enriched in nickel until reaching
`
`< F
`
`igure 8 Cross-section of four TBCs illustrating different forms of the local sep-
`aration between the TBC and the bond-coat after thermal cycling. In each case, the
`coating was deposited conformally on the bond-coat so interface separations such as
`these indicate that the underlying bond-coat has changed its surface morphology dur-
`ing high—temperature exposure and thermal cycling. In example ([9), no separation has
`occurred and it exhibits the longest life.
`
`13
`
`13
`
`

`
`396
`
`CI ARK *
`
`L *‘Vl
`
`
`
`Schematic illustration of how an initially flat but compressed film (left)
`Figure 9
`can lower its elastic strain energy by rumpling (right). The amplitude of rumpling
`is enhanced by thermal cycling and can cause interface separation if a superimposed
`coating cannot deform to follow the displacements of the film.
`
`
`asdemsitédiii
`
`I after cyclic oxirlationf, x
`“j,1200”C,“5Bx_1hr
`r
`a
`
`Figure 10 Microstiucture of an initially flat as-aluminized bond coat after 50 1-11
`cycles at l200°C: (a) surface rumpling; (b) cross—section showing a rather uniform
`oxide layer and strong surface undulations (y’—phase is revealed by etching); (c, (1')
`optical micrographs showing etched cross—section before and after cyclic oxidation.
`Dark areas on the optical images coirespond to the ;3—phase, whereas the y’-phase in
`the coating appears white.
`
`14
`
`14
`
`

`
`TH RMAT ARKIER COATINGS
`
`397
`
`Composifort in wt.% Ni
`
`0
`
`was
`
`20
`e
`
`e
`
`40
`
`u
`
`so
`
`so
`
`100
`.
`
`1800
`
`1460
`
`1200
`
`inC
`Temperature
`
`Al
`
`Composition in at.% Ni
`
`Ni
`
`Figure 11 The Ni—Al pseudo—binary phase diagram illustrating the compositional
`range of the ,8—NiAl phase and the direction of the change in composition as the bond-
`coat is depleted of Al by interdiffusion and selective oxidation.
`
`the single—phase boundary at which point fiirther depletion leads to the formation of
`1/—Ni3Al. (At even later times, the composition can extend into the y region of the
`phase diagram.) Martensitic structures within the ;3—NiAl phase field also form as
`the phase boundary is approached. Whereas the martensite start temperature, Ms,
`of the pure ,8—NiAl compositions is known to be generally around room tempera-
`ture to 300°C (20), the additional Pt, Co, and Cr present in the PtNiAl bond-coat
`increase the Mg temperature, and Hemker et al. have reported Mg temperatures of
`~600°C (21).
`Substantial progress has been made in the past few years in understanding some
`ofthe mechanisms that lead to flaw initiation and growth during use (9, 14, 18, 22).
`These provide the basis for developing life—prediction models, but nevertheless a
`number ofunanswered questions remain. For instance, the micrographs in Figure 8
`were obtained from nominally the same superalloy, with the same bond-coat and
`the same YSZ coating all made by the same manufacturer in the same process
`manner. This difference in interface separation and roughening is particularly
`marked in this figure, but it does suggest that even small, but as yet unidentified,
`concentrations of dopants can have a large effect on life (23).
`Insights gained in the past few years into some of the important processes
`occurring within the coating during use are summarized in Figure 12. Essentially,
`
`15
`
`15
`
`

`
`398
`
`CLARKE I LEVI
`
`L
`
`'
`
`L
`
`
`
`L
`
`
`
`""* L
`,
`rireepsandjalsne
`densifies by eintering
`L
`L
`A Phase lransformaticun
`,
`Q; .
`Ira TGO during uxldattiszni
`Emile? interface
`
`pmsétien a A
` Separaticn
` T
`J
`M
`
`T
`\
`JA‘
`#3.!
`@ 3'
`g
`.1
`Y;
`Local mughnees grows.
`by growth strain induced
`Qxgde gmws in
`ratcheting and bond
`Wckmsa amj
`carat vclume decreases
`expands Iaterauy
`
`Elxidatioh causes
`
`Al-depleticzn of
`I‘_)()nd—{;u:aa?(
`
`fi__W .
`
`N
`
`.L,
`
`LLLL
`
`_,
`
`L-
`
`_N
`
`i, W’ etc
`
`Localized Densification
`Due to [3——;s3:‘Trans'forn1afinn
`
`baweeu
`‘L bond recall and underlying euperaltcy l
`4+“
`-IA!
`
`_
`
`Figure 12
`
`Schematic summary of the concurrent processes occurring in the bond-
`
`coat, TGO and TBC, during use at high temperatures. The complexity in failure times
`and failure modes is believed to reflect the competition between hese individual pro-
`cesses.
`
`the TBC system is one that evolves with time at temperature and its evolution
`depends in detail on not only the temperature but also on the thermal cycle history
`and heating and cooling rates, as well as composition.
`
`NON—DESTRUCTIVE EVALUATION
`
`As the design of coatings shifts to a philosophy of prime reliance, the ability to
`non—destructively monitor the coating, identify defects, and evaluate its remaining
`life becomes more important. In addition, there is a growing economic pressure to
`defer maintenance and replacement of parts until really necessaiy. (The costs are
`staggering: The cost, in replacement electricity alone, oftaking a power generation
`turbine out ofoperation can be ofthe order of $1 million a day. The cost ofreplacing
`a single, first—stage turbine blade can also be very high. Depending on its size the
`cost can be as high as $10,000.)
`In the case of large area separations, several millimeters to centimeters, infrared
`imaging provides a direct means of visualizing incipient coating failure provided
`the blade can be accessed. Usually, though, a coating has failed by the time the
`separation reaches such a large size, and thus methods of identifying damage at
`an earlier time, and hence smaller size, are required. No single solution appears
`practical at this stage, although laser-induced acoustic sounding (from Lasson
`
`16
`
`16
`
`

`
`THERMAL BARRIER COATING S
`
`399
`
`Technologies, personal communication) and higher—spatia resolution imaging us-
`ing polarized scattered light have shown promise in detectng deliberately created
`interface flaws (25). An alternative method is one that utilizes piezospectroscopy,
`the strain—induced shift of luminescence and Raman lines. When illuminated with
`
`a laser having an appropriate wavelength, luminescence from the aluminum oxide
`TGO fonned by oxidation on the bond—coat can be detected through the thickness
`ofthe TBC (26, 27). The luminescence spectrum, from Cr3+ ions incorporated into
`the alumina TGO as it grows, is sensitive to the stress state in the TGO. (Because
`zirconia is a wide band—gap material, it is transparent in the visible so illu1nina—
`tion in the blue or green, e.g., as an argon ion laser, can penetrate to the TGO,
`and the stimulatedluminescence, which is in the red, is transmitted back through
`the coating.) The key to the use of photoluminescence as an NDE tool is that the
`frequency of the luminescence lines shifts with mean stress and the shape of the
`luminescence lines is a direct measure of the stress distribution within the region
`ofthe TGO probed by the laser. As described in detail elsewhere, the luminescence
`lines can be deconvoluted and the degree of damage and local interface separation
`evaluated (1 1, 15). Since this can be performed using focused lasers and is also
`nondestructive, piezospectroscopy—based methods show particular promise as a
`NDE tool for evaluating coatings. An example is illustrated in Figure l3a,b, that is
`a comparison of spectral data from two coatings. In Figure 13a, there little change
`in spectral shift with thermal cycling, the TGO remains flat and does not fail within
`the time of the experiments (curve A in Figure 130) In Figure 131), a continuous
`change in frequency (curve B in Figure 13c) is shown that exhibited substantial
`spectral broadening and failed after several hundred cycles, an average value but
`shorter than the other.
`
`SINTERING AND DENSIFICATION
`
`One of the concerns in developing a reliable and robust coating is how the coating
`changes during use at high temperatures. By analogy with the behavior of other
`porous ceramics, it might be expected that the coating will densify by the reduction
`in surface energy associated with the excess surface area of the pores. This process
`is commonly referred to as sintering. There are two principal concerns. One is
`that densification inevitably increases the elastic modulus and thereby decreases
`the strain compliance of the coating. The other is that densification decreases
`the volume fraction of porosity which, in tum, causes the thermal conductivity
`to increase. Although the kinetics of densification of bulk, free—standing zirconia
`ceramics is well characterized, evaluation of these effects in actual coatings is
`not. In part this is because of the peculiar geometry of the porosity in plasma-
`sprayed and electron-beam deposited coatings. In a bulk ceramic, porosity tends
`to be spherical in shape, whereas it is plate—like in plasma—sprayed coatings and
`columnar-like in EB~PVD TBCs. These differences in pore shape and topology
`are compounded by the fact that the densification of a TBC is constrained by the
`presence of the underlying alloy, which has not yet been investigated in detail.
`
`17
`
`17
`
`

`
`
`
`
`
`$2:.wSamM.o.nWnWM%M‘M.3_3:;MWmgm.82:Mm2:;m.
`.J.-MaTmm
`5..57uh.mum._Hu3:;E92;.\
`
`
`
`..T_.53W.Tumzmauma
`
`Lam:M.£3;aM.W.
`
`9.8:,anmo_u>o:4we..0nE:ZATE3._mnE:_._m>m__S
`
`
`
`
`
`Emmomsawsawa82:3.3..33:.ommz.aomi
`
`W
`
`4SS.338%:3
`
`menu88
`
`C3%
`ISawW.33MW83WE2.2Mfl82.m,
`L88....nw
`
`
`
`02:.aam
`
`
`
`gm».3:;054:.ommfi08:cam?33;.83;ammflSufiao
`
`
`
`
`
`
`
`I253E3:;.95Sara!ill.1:,,.55.....mm
`
`
`
`
`
`A7E3EaE::m>m>>9.E8.._maE=_._m>m_.S
`
`Zrflnnt.
`
`wE...n|
`
`22.3.aSaga3
`
`18
`
`18
`
`
`
`

`
`THERMAL BARRIER COATINGS
`
`401
`
`In fact, it is quite likely that the densification of plas1na—sprayed and * —PVD
`TBCS might be quite different from one another and also from bulk zirconia
`ceramics on account of the constraint imp

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket