Vapor deposited samarium zirconate thermal barrier … · Vapor deposited samarium zirconate...

11
Vapor deposited samarium zirconate thermal barrier coatings Hengbei Zhao a, , Carlos G. Levi b , Haydn N.G. Wadley a a Dept. of Materials Science and Engineering, University of Virginia, Charlottesville, VA 22904, USA b Materials Department, University of California, Santa Barbara, CA 93106-5050, USA abstract article info Article history: Received 3 December 2008 Accepted in revised form 24 March 2009 Available online 1 April 2009 Keywords: Samarium zirconate Directed-vapor deposition Thermal barrier coatings The rare earth zirconates (M 2 Zr 2 O 7 , M=LaGd) have a low intrinsic thermal conductivity and high temperature phase stability making them attractive candidates for thermal barrier coating applications. An electron-beam evaporation, directed-vapor deposition (EB-DVD) technique has been used to investigate the synthesis of Sm 2 Zr 2 O 7 (SZO) coatings and to explore the relationships between the deposition conditions and the coating composition, structure, texture, pore morphology, and thermal conductivity. The coatings exhibited signicant uctuations in composition because of differences in the vapor pressures of the constituent oxides and had a metastable uorite structure because kinetic limitations hindered the cation ordering needed to form the equilibrium pyrochlore structure. The growth morphology of the EB-DVD SZO was similar to that of EB-DVD 7YSZ and EB-PVD Gd 2 Zr 2 O 7 . The SZO coatings had as-deposited conductivities of 0.5±0.1 W m 1 K 1 , about one-half of their DVD 7YSZ counterparts. The majority of the conductivity difference appears to be associated with the lower intrinsic conductivity of this zirconate. Published by Elsevier B.V. 1. Introduction Thermal barrier coating (TBC) systems have become an enabling materials technology for the gas turbine engines used for propulsion and power generation [1]. Through their ability to provide thermal and environmental protection, TBCs enhance superalloy component dur- ability and/or allow an increased operating temperature with con- comitant benets for engine performance and efciency [2]. Today's state-of-the-art TBC systems comprise; (i) a thick (100300 μm) thermal barrier layer typically based on 7 wt.% yttria stabilized zirconia (7YSZ), (ii) a much thinner (b 10 μm) environmental barrier layer of thermally grown aluminum oxide (TGO) that protects the superalloy substrate against oxidation, and (iii) a relatively thick (50100 μm) metallic bond coatlayer with a chemical composition formulated to result in growth of a stable TGO over the life of the system [3,4]. A key characteristic of the thermal barrier layer is the presence of substantial porosity of various morphologies and length scales. The primary role of these pores is to increase the in-plane compliance of the thicker oxide and enhance its tolerance to strains arising from its thermal expansion mismatch with the substrate during the thermal cycling typical of engine operation [5,6]. It also results in a reduction in the thermal conductivity of the TBC, especially if thin lamellar pores are present and oriented to impede heat ow [7]. 7YSZ has been used effectively in TBCs for over two decades [8] owing to its combination of low intrinsic thermal conductivity (~2.1 W m 1 K 1 at 1000 T 1200 °C [9]), its thermochemical compatibility with the underlying TGO [10,11], and its durability to thermal cycling [12]. However, its viability has become increasingly challenged as the demand for higher gas path temperatures has continued to increase. The primary concerns arise from; (i) accelerated sintering kinetics that degrade the compliance and insulating efciency of the coating [1316], (ii) de-stabilization of the desirable non- transformable tphase [10,1720] that is responsible for the intrinsic toughness of 7YSZ [21,22] and (iii) its propensity for penetration by molten deposits of calciummagnesiumalumino-silicates (CMAS) that degrade the strain tolerance of the coating and accelerate its de- stabilization [2325]. These concerns, coupled with the potential design benets of further reductions of the thermal conductivity relative to 7YSZ [2628], have motivated a search for alternate thermal barrier oxides [2,10,2933]. Among the numerous oxides that have been explored as alternate thermal barrier materials, the rare earth zirconates (REZ's) M 2 Zr 2 O 7 (M=LaGd), with the pyrochlore structure, have attracted particular interest because of their generally lower thermal conductivity (~1.21.5 W m 1 K 1 at 1000 T 1200 °C) [27,3439]. In addition, these materials are phase-stable at all temperatures relevant for gas turbine operation (T 1500 °C), Fig. 1 [40]. They also exhibit higher resistance to morphological evolution of the pore structure (sintering) than 7YSZ [39,41,42] and, notably show potential for inhibiting CMAS penetration by inducing crystallization of the silicate melt [43]. However, the REZ's suffer two fundamental disadvantages relative to 7YSZ. First, they exhibit a thermochemical incompatiblity with alumina [10,44], tending to form aluminate interphases by diffusional interactions with the TGO [45,46]. Second, as with other cubic ZrO 2 -based phases, they exhibit substantially lower toughness than the tetragonal phase of 7YSZ [3,21,47,48]. This has deleterious consequences for both their erosion resistance and the failure mechanisms that involve crack propagation through the TBC [3,5]. Surface & Coatings Technology 203 (2009) 31573167 Corresponding author. E-mail address: [email protected] (H. Zhao). 0257-8972/$ see front matter. Published by Elsevier B.V. doi:10.1016/j.surfcoat.2009.03.048 Contents lists available at ScienceDirect Surface & Coatings Technology journal homepage: www.elsevier.com/locate/surfcoat

Transcript of Vapor deposited samarium zirconate thermal barrier … · Vapor deposited samarium zirconate...

Page 1: Vapor deposited samarium zirconate thermal barrier … · Vapor deposited samarium zirconate thermal barrier coatings ... Directed-vapor deposition Thermal barrier ... temperature

Surface & Coatings Technology 203 (2009) 3157–3167

Contents lists available at ScienceDirect

Surface & Coatings Technology

j ourna l homepage: www.e lsev ie r.com/ locate /sur fcoat

Vapor deposited samarium zirconate thermal barrier coatings

Hengbei Zhao a,⁎, Carlos G. Levi b, Haydn N.G. Wadley a

a Dept. of Materials Science and Engineering, University of Virginia, Charlottesville, VA 22904, USAb Materials Department, University of California, Santa Barbara, CA 93106-5050, USA

⁎ Corresponding author.E-mail address: [email protected] (H. Zhao).

0257-8972/$ – see front matter. Published by Elsevierdoi:10.1016/j.surfcoat.2009.03.048

a b s t r a c t

a r t i c l e i n f o

Article history:Received 3 December 2008Accepted in revised form 24 March 2009Available online 1 April 2009

Keywords:Samarium zirconateDirected-vapor depositionThermal barrier coatings

The rare earth zirconates (M2Zr2O7, M=La→Gd) have a low intrinsic thermal conductivity and hightemperature phase stability making them attractive candidates for thermal barrier coating applications. Anelectron-beam evaporation, directed-vapor deposition (EB-DVD) technique has been used to investigate thesynthesis of Sm2Zr2O7 (SZO) coatings and to explore the relationships between the deposition conditions andthe coating composition, structure, texture, pore morphology, and thermal conductivity. The coatingsexhibited significant fluctuations in composition because of differences in the vapor pressures of theconstituent oxides and had a metastable fluorite structure because kinetic limitations hindered the cationordering needed to form the equilibrium pyrochlore structure. The growth morphology of the EB-DVD SZOwas similar to that of EB-DVD 7YSZ and EB-PVD Gd2Zr2O7. The SZO coatings had as-deposited conductivitiesof 0.5±0.1 W m−1 K−1, about one-half of their DVD 7YSZ counterparts. The majority of the conductivitydifference appears to be associated with the lower intrinsic conductivity of this zirconate.

Published by Elsevier B.V.

1. Introduction

Thermal barrier coating (TBC) systems have become an enablingmaterials technology for the gas turbine engines used for propulsion andpower generation [1]. Through their ability to provide thermal andenvironmental protection, TBCs enhance superalloy component dur-ability and/or allow an increased operating temperature with con-comitant benefits for engine performance and efficiency [2]. Today'sstate-of-the-art TBC systems comprise; (i) a thick (100–300 μm)thermal barrier layer typically based on 7 wt.% yttria stabilized zirconia(7YSZ), (ii) a much thinner (b10 μm) environmental barrier layer ofthermally grown aluminum oxide (TGO) that protects the superalloysubstrate against oxidation, and (iii) a relatively thick (50–100 μm)metallic “bond coat” layer with a chemical composition formulated toresult in growth of a stable TGO over the life of the system [3,4]. A keycharacteristic of the thermal barrier layer is the presence of substantialporosity of various morphologies and length scales. The primary role ofthese pores is to increase the in-plane compliance of the thicker oxideand enhance its tolerance to strains arising from its thermal expansionmismatch with the substrate during the thermal cycling typical ofengine operation [5,6]. It also results in a reduction in the thermalconductivity of the TBC, especially if thin lamellar pores are present andoriented to impede heat flow [7].

7YSZ has been used effectively in TBCs for over two decades [8]owing to its combination of low intrinsic thermal conductivity(~2.1 W m−1 K−1 at 1000≤ T≤1200 °C [9]), its thermochemicalcompatibility with the underlying TGO [10,11], and its durability to

B.V.

thermal cycling [12]. However, its viability has become increasinglychallenged as the demand for higher gas path temperatures hascontinued to increase. The primary concerns arise from; (i) acceleratedsintering kinetics that degrade the compliance and insulating efficiencyof the coating [13–16], (ii) de-stabilization of the desirable non-transformable t′ phase [10,17–20] that is responsible for the intrinsictoughness of 7YSZ [21,22] and (iii) its propensity for penetration bymolten deposits of calcium–magnesium–alumino-silicates (CMAS)that degrade the strain tolerance of the coating and accelerate its de-stabilization [23–25]. These concerns, coupledwith the potential designbenefits of further reductions of the thermal conductivity relative to7YSZ [26–28], have motivated a search for alternate thermal barrieroxides [2,10,29–33].

Among the numerous oxides that have been explored as alternatethermal barrier materials, the rare earth zirconates (REZ's) M2Zr2O7

(M=La→Gd), with the pyrochlore structure, have attracted particularinterest because of their generally lower thermal conductivity (~1.2–1.5 W m−1 K−1 at 1000≤T≤1200 °C) [27,34–39]. In addition, thesematerials are phase-stable at all temperatures relevant for gas turbineoperation (T≤1500 °C), Fig. 1 [40]. They also exhibit higher resistance tomorphological evolution of the pore structure (sintering) than 7YSZ[39,41,42] and, notably show potential for inhibiting CMAS penetrationby inducing crystallization of the silicate melt [43]. However, the REZ'ssuffer two fundamental disadvantages relative to 7YSZ. First, theyexhibita thermochemical incompatiblity with alumina [10,44], tending to formaluminate interphases by diffusional interactions with the TGO [45,46].Second, aswith other cubic ZrO2-based phases, theyexhibit substantiallylower toughness than the tetragonal phase of 7YSZ [3,21,47,48]. This hasdeleterious consequences for both their erosion resistanceandthe failuremechanisms that involve crack propagation through the TBC [3,5].

Page 2: Vapor deposited samarium zirconate thermal barrier … · Vapor deposited samarium zirconate thermal barrier coatings ... Directed-vapor deposition Thermal barrier ... temperature

Fig. 1. The ZrO2–SmO1.5 phase diagram (adapted from reference [40]). When pyrochlore (P)phase ordering is kinetically suppressed, as in the current work, the extended fluorite solidsolution would be bound by the extrapolated M+F and F+B boundaries (denoted by thedashed lines). The shaded area in the middle of the metastable fluorite field shows thedeposition temperature range and the maximum variation in SmO1.5 observed in theexperiments. Note that the minimum fluctuation observed (±5%) would fall mostly withinthe single phase pyrochlore upon ordering, whereas larger fluctuations could lead to two-phase F+P and P+B mixtures.

Fig. 2. Schematic diagram of the directed-vapor deposition (DVD) apparatus used todeposit SZO coatings.

3158 H. Zhao et al. / Surface & Coatings Technology 203 (2009) 3157–3167

To alleviate some of these problems, zirconate TBCs are typicallydeposited on a tougher, TGO-compatible 7YSZ underlayer [10,27,45,49],although that does not address the erosion issue.

Full exploitation of REZ's as thermal barrier oxides requires notonly strategies to circumvent their limitations relative to 7YSZ, butalso the development of suitable processing approaches. In general,the deposition of rare earth zirconate coatings is more challengingthan 7YSZ. The atmospheric plasma spray (APS) process mustconfront the higher melting temperatures of the stoichiometricphases while the electron-beam physical vapor deposition (EB-PVD)approach must overcome the sometimes significant differences invapor pressure of the constituent oxides [50] which hinders achieve-ment of the desired composition during evaporation [51]. Experiencesuggests that Gd2Zr2O7 (GZO) is more amenable to EB-PVD thanLa2Zr2O7, (LZO) and is consistent with the relative magnitudes of theirvapor pressures [50]. Samarium zirconate, Sm2Zr2O7 (SZO), isreported to be suitable for vapor deposition [42] but its behaviorduring vapor phase processing has not been well documented.Moreover, the high cost of the rare-earth oxides and their highcontent in zirconate TBCs exacerbates concerns about the lowdeposition efficiency (a few percent) of conventional EB-PVD.

Gas-jet based evaporation techniques such as electron-beamdirected-vapor deposition (EB-DVD) offer high material depositionefficiency [52] and their multi-source capability facilitates thedeposition of the preferred multilayer (e.g. REZ/7YSZ) [10,27,45,49]and/or the option of co-evaporation and deposition of the constituentoxides if challenges posed by their vapor pressure differences areinsurmountable. This investigation is motivated by the need todevelop a processing science for the deposition of REZ TBCs using aDVD approach. The specific emphasis is on samarium zirconate. Theviability of depositing uniform compositions is assessed and theinfluence of substrate rotation rate and deposition temperature on thecoating's density, pore morphology, texture and microstructure, andtheir related impacts on coating thermal conductivity are investigated.The observations are compared with those reported recently foridentically grown 7YSZ coatings [53].

2. Experimental methods

2.1. Samarium zirconate deposition

Samarium zirconate coatings were deposited on both stationaryand rotated substrates using an EB-DVD technique [52] with theconfiguration illustrated in Fig. 2. The deposition conditions werenominally identical to those used for 7YSZ which are describedelsewhere [53]. The coatings were applied to 25.4 mm diameterHastelloy-X substrates (provided by GE Aviation, Evendale, OH)oriented as shown in Fig. 2. Prior to deposition the substrates wereEB-PVD coated with a 100-200 μm thick Ni–22Co–20Cr–10Al–0.5Y (inwt.%) alloy, and subsequently hand polished using 1200 grit paper andthen 1 μm diamond finish. The evaporation source was a 12.5 mmdiameter SZO rod (TCI Ceramics Inc., MD), positioned in the throat of awater-cooled copper nozzle with a coaxial He/O2 gas jet at itsperiphery, as illustrated in Fig. 2. The chemical composition of the rodmeasured by energy-dispersive X-ray spectroscopy (EDS) was48ZrO2–52SmO1.5 (in mol%). The electron-beam accelerating voltagewas 70 kV and the incident beam power was 2.45 kW. The substratetemperature was monitored and maintained at a prescribed levelduring deposition using a thermocouple placed very close to thesubstrate. We note that there is a difference between the actualtemperature of the deposition surface and that measured by thethermocouple, but for the purposes of this comparative analysis theyare assumed to be the same. A systematic investigation of the effect ofthe putative substrate temperature (900–1100 °C) and rotation rate(0–20 rpm) upon the coating morphology, density and thermalconductivity was then conducted. The deposition rate was 10±2 μm/min for the samples deposited without rotation and approximately 3–4 μm/min for all others.

2.2. Sample characterization

The microstructure of the coatings was characterized by scanningelectron microscopy (SEM) of the growth surfaces as well as through-thickness polished cross-sections and fracture surfaces. The chemical

Page 3: Vapor deposited samarium zirconate thermal barrier … · Vapor deposited samarium zirconate thermal barrier coatings ... Directed-vapor deposition Thermal barrier ... temperature

3159H. Zhao et al. / Surface & Coatings Technology 203 (2009) 3157–3167

composition was analyzed by EDS in the SEM. The composition of thecompound in this paper is given in a stoichiometric compound. Theorientation and phase constitution of the coatings were determinedby X-ray diffraction (XRD) using a XDS 2000 powder diffractometer(Scintag Inc., USA), CuKα radiation, a scan angle range 10°≤2θ≤90°,and a step size of 0.02°. Lattice parameters were determined fromrefined cell dimensions using a software package based on Cohen'smethod of least squares [54]. A full texture analysis was conductedusing an X1 Texture instrument (Scintag Inc., USA) by varying the tiltangle α from 0 to 90° and rotating the sample over a range0°≤ψ≤360° about the substrate normal. The step sizes for bothrotation and tilting were 5°. The thermal conductivities of the coatingswere measured at room temperature by the phase-of-thermal-emission spectroscopy method [55], as described in the earlier studyon 7YSZ [53].

3. Results

Typical microstructures of the EB-DVD SZO coatings are illustratedin Fig. 3.Measured thicknesses ranged from~65 to 110 μm. The averagebulk composition was Sm2.1Zr1.9O7: slightly off stoichiometry butrelatively consistentwith thatmeasured for the source rod. In contrast,the backscattered electron images in Fig. 3 reveal fluctuations from theaverage SmO1.5 content within each coating. The amplitude of thesefluctuations varied from sample to sample. The lowest variation of

Fig. 3. Back scattered electron images of typical SZO coating microstructures depositedat 1000 °C with superposed composition profiles determined by EDS. The banding is amanifestation of variation in the samarium content through the thickness. (a) Coatingdeposited without rotation, exhibiting a samarium fluctuation of b±5 at.%Sm. (b) Acoating deposited at a high rotation rate where the fluctuation was up to ±12 at.%Sm.

±5% was found for a non-rotated specimen, Fig. 3(a), but some of therotated coatings had higher variations (up to ±12% for the specimenrotated at 20 rpm), Fig. 3(b). There was no apparent correlationbetween the deposition conditions and the extent of compositionalheterogeneity. Rather, it appears that the significant vapor pressuredifference of the constituent species has created melt instabilitiesleading to unsteady evaporation of the two species. As with 7YSZ, thetexture and morphology of the SZO coatings grown by DVD werehighly sensitive to the growth conditions. The salient observations aredescribed below.

3.1. Deposition temperature effects

The morphology of the SZO coatings was strongly affected by thesubstrate temperature during deposition. Fig. 4 shows surface andthrough-thickness fracture surface images of coatings deposited at 900°,1000° and 1100 °Cusing afixed substrate rotation rate of 6 rpm.All threecoatings were composed of relatively tightly packed growth columnswith faceted tips. At the higher temperatures, Fig. 4(c), the tips had awell-defined 4-sided pyramidal shape typical of fluorite derivativestructures and were similar to those seen in 7YSZ [56,57] and otherzirconates [51]. The growth columns had a strong ⟨200⟩ preferredorientation. As the deposition temperature decreased, secondarycrystals started to appear on the growth surface, e.g. Fig. 4(b), andthese became sufficiently numerous at 900 °C to make the shape of thetips appear rough and indistinct, Fig. 4(a).

Examination of the through-thickness fracture surfaces, Fig. 4(d–f),reveals a fairly typical columnar structure for coatings grown at thehigher temperatures. These are again reminiscent of those seen in EB-PVD coatings [58]. The structure formed at 1100 °C, Fig. 4(f), appearsmore regular than one might infer from the top view, Fig. 4(c), withcolumns separated by distinct intercolumnar gaps andwell-developedfeathery pores within the columns. These have been designated inearlier papers as Type I and II porosity respectively [53,59]. As thetemperature decreases to 1000 °C the fracture surface reveals incipientbranches growing off the column sides, Fig. 4(e), but themain columnsare still clearly identifiable. Further reduction in deposition tempera-ture to 900 °C, however, leads to highly branched columns with anapparent loss of the feathery morphology. The distinction betweenintra- and intercolumnar pores on the fracture surface also becomesblurred, as seen by comparing Fig. 4(d) and (f). Onemight envisage thestructure in Fig. 4(d) as containing two main columns each consistingof multiple sub-columns oriented on average along the vertical axis.Examination of polished cross-sections corresponding to this samplerevealed that the scale of the branching increased with distance fromthe substrate (Fig. 5). Even so, some of the coarser “branches” underthe surface still exhibited the fine feathery pore structure associatedwith a high through-thickness thermal resistance.

XRD patterns of coatings deposited at the different substratetemperatures (all at 6 rpm) are shown in Fig. 6. All of the peak positionsare consistent with the fluorite (disordered) form of Sm2Zr2O7 [60]with a lattice parameter 0.5285 nm. When doubled to account for theordering in the pyrochlore structure, this is slightly lower than thereported value of 1.0594 nm for samarium zirconate [60]. The peakintensities reveal a predominant ⟨200⟩1 out of plane texture at alltemperatures, with minor amounts of secondary orientations presentat the higher temperatures (most notably ⟨111⟩). The consistency ofthe strong ⟨200⟩ texture across the temperature range is ratherremarkable in the context of the substantial changes in the columnarstructure in Fig. 4. Pole figures (not shown) also suggested preferredin-plane orientations but these are less distinct than those reported forconventional EB-PVD 7YSZ [56,58] and gadolinium zirconate [48].

1 Crystallographic notation referred to the fluorite structure. The equivalentpyrochlore planes have doubled indices, i.e. {200}F={400}Py.

Page 4: Vapor deposited samarium zirconate thermal barrier … · Vapor deposited samarium zirconate thermal barrier coatings ... Directed-vapor deposition Thermal barrier ... temperature

Fig. 4. Surface topology (a–c) and through-thickness fracture surface images (d–f) of SZO coatings deposited at temperatures of (a and d) 900°, (b and e) 1000° and (c and f) 1100°Con substrates rotated at 6 rpm. Note the increase propensity for growth column branching as the deposition temperature is reduced.

3160 H. Zhao et al. / Surface & Coatings Technology 203 (2009) 3157–3167

The density (ρ) and through-thickness thermal conductivity (k) ofthe SZO coatings were measured as a function of the depositiontemperature and are compared on a normalized basis with data forsimilarly grown 7YSZ coatings in Fig. 7. The conductivities (k0) of thetheoretically fully dense (ρ0) 7YSZ and SZOat ambient temperature aretaken as 3 and2Wm−1 K−1 based on [38]. Both normalizedpropertiesfor SZO are lower than for 7YSZ, by ~5% for ρ/ρ0 and ~8% by k/k0. Theporosity content of the zirconate is therefore slightly higher andslightly more effective at hindering heat transport in the through-thickness direction. One should note, however, that these differencesare close to the typical scatter of the measurement techniques utilized[55]. Both properties increase with coating deposition temperature,reflecting a concomitant decrease in porosity. However, the relativeconductivity for SZO rises comparatively faster than its relative density

with increasing deposition temperature. Relative conductivities afteraging (Fig. 7b) are also lower for the zirconate, and it is notable that thevalues for the higher temperature are now distinctly separated,whereas those for the lower temperature are closer together.

3.2 Rotation rate effects

Examples of themorphology of columnar coatings grown at 1000 °Con stationary substrates as well as those corresponding to the slowest(0.5 rpm)andhighest (20 rpm) rotation rates are shown in Fig. 8. (Thosefor 6 rpm were shown in Fig. 4b and e). Polished cross-sections arechosen for Fig. 8(d–f) instead of fracture surfaces (cf. Fig. 4(d–f))because the latter were less informative, especially for the lowestrotation rate.

Page 5: Vapor deposited samarium zirconate thermal barrier … · Vapor deposited samarium zirconate thermal barrier coatings ... Directed-vapor deposition Thermal barrier ... temperature

Fig. 5. A polished cross-section of the sample deposited at 900°C that shows themultiply branched columns near the coating surface. Fine scale “feathery” porositywithin the growth columns is responsible for the chevron shading.

Fig. 6. X-ray diffraction patterns for SZO coatings deposited at nominal substratetemperatures of (a) 900°, (b) 1000° and (c) 1100°C. The rotation rate was 6 rpm.

3161H. Zhao et al. / Surface & Coatings Technology 203 (2009) 3157–3167

When the coatings were grown on stationary substrates undernominally normal incidence, as in Fig. 8(a), the growth surface wascomposed of column tips bound by triangular facets andwasmuchmoreplanar than those observed inYSZgrownunder the same conditions [53],or on the samarium zirconate coatings grownwith rotation, Fig. 8(b) and(c), which exhibited square-based pyramidal column tips bound bycurved triangular faces. Some pyramidal tips are also evident in Fig. 8(a),but themajority of the triangular facets were oriented closer to the planeof the substrate. It is also noted that these facets show ledge-like featurespredominantly parallel to one of the sides.

The cross-sectional views in Fig. 8 show that the coating grownwithout rotation comprises closely packed columns with muchnarrower intercolumnar gaps than those grown with rotation.Interestingly, the columns in Fig. 8(d) still show “feathery” branchesthat are more similar to those found in rotated EB-PVD coatings thanto those grown with stationary substrates [56].

Slow rotation at 0.5 rpm resulted in pronounced C-shapedcolumnar structures with a wavelength of ~10.8 μm and similarlyshaped intercolumnar gaps, Fig. 8(e). The wavelength is consistentwith the rate of deposition and the period of a single revolution of thesubstrate. The columns are clearly defined and very thin near thesubstrate, but they become less distinct as they grow. This is a result ofthe concomitant widening and segmentation of the column tips,evident in Fig. 8(e), apparently occurring as the vapor flux orientationtransitions from the “sunset” to “sunrise” part of the next rotationcycle [58].With increasing rotation rate the C-shaped features becomesmaller and rapidly lose their identity as the growth front moves awayfrom the substrate. The columns then adopt the more typicalstructure, as shown for the 20 rpm case in Figs. 3(b) and 8(f). Theintercolumnar gaps are aligned close to the substrate normal, withtypical feathery pores within each column.

XRD patterns in Fig. 9 are again indexable to the (disordered)fluorite form of the zirconate, regardless of the rotation rate. However,the relative intensity of the diffraction peaks changed with rotationrate, indicative of a change in the out-of-plane texture. Coatingsgrown on stationary substrates exhibit a dominant ⟨111⟩ peak,consistent with prior observations on similarly grown 7YSZ coatings[53] but with significant amounts of ⟨220⟩ and ⟨311⟩ reflections, whichwere not present in the latter. The implication is that the texture ismixed, as suggested by the top views of the growth surface in Fig. 8(a).The ⟨200⟩ reflections are much smaller in this case, but grow inintensity relative to ⟨111⟩ as the rotation rate increases, while the⟨220⟩ and ⟨311⟩ peaks rapidly disappear. This was accompanied by achange in surface morphology from triangular facets in Fig. 8(a) to thepyramidal column tips characteristic of ⟨200⟩ growth in 7YSZ undersimilar conditions [53].

The relative density and normalized thermal conductivity for theSZO coatings are shown as a function of the rotation rate in Fig. 10, andcompared with similar data for 7YSZ [53]. Both materials exhibitsimilar trends with rotation rate in the as-deposited condition, withthe individual property values being consistently lower for the

Page 6: Vapor deposited samarium zirconate thermal barrier … · Vapor deposited samarium zirconate thermal barrier coatings ... Directed-vapor deposition Thermal barrier ... temperature

Fig. 7. Effect of deposition temperature on the (a) density and (b) thermal conductivityof SZO DVD coatings, compared with those of YSZ deposited under the same conditions.The properties are normalized in both cases by those of the dense oxides. The rotationrate was 6 rpm. AD and HT denote as-deposited and heat treated for 10 h at 1100 °C.

3162 H. Zhao et al. / Surface & Coatings Technology 203 (2009) 3157–3167

zirconate for all rotation rates, (although we note that some of thedifferences are within the error of the measurements). The relativeorder seems to be reversed for the stationary samples, but thedifferences in conductivity are within the experimental scatter. Thedifferences in conductivity are somewhat reduced after heat treat-ment (Fig. 10c) but the same trends persist.

2 Note that the curve for LaOLZO does not bend at the liquidus temperature of LZO,suggesting the calculation in [61] ignores the solid/liquid transition.

4. Discussion

Salient findings from this study include; (i) confirmation that SZOcan be deposited by DVD with microstructures suitable for thermalbarrier coating applications albeit with some composition variabilitythat requires further study; (ii) SZO is deposited with a metastablefluorite structure instead of the stable pyrochlore form; (iii) thereexists overall similarity in crystallographic texture with 7YSZ grownunder comparable conditions but the pore morphology appears to bemore sensitive to deposition temperature in SZO; and (iv) there aredifferences in porosity content and its evolution during hightemperature aging that impact the thermal conductivity. Thesefindings are further discussed below.

4.1. Coating composition

In principle, the composition of a multi-component vapor shouldbe different from that of the melt fromwhich it emanates, as dictatedby the vapor pressures of the pure components and their activities inthe melt. In the case of oxides, there are additional concerns aboutdissociation reactions, e.g. the vapor of a ZrO2–MO1.5 (M=Y or rareearth) alloy may contain Zr, ZrO, ZrO2, M, MO, O and O2 [61]. It is alsoknown that one may establish a steady state evaporation conditionwherein a solid is continuously fed into its melt and the vapor formingfrom the melt has the same composition as the solid, albeit not that ofthemelt. This is claimed to be feasible as long as the vapor pressures ofthe components are not too dissimilar (reportedly by no more thantwo orders of magnitude) [62].

Information on the vapor pressures of the relevant species overZrO2–MxOy melts is generally not available in the literature, althoughestimates could be made based on data for the corresponding solids.(The slope should change moderately at the melting point but thecurve should be essentially continuous.) The available (calculated)data for the relevant solids is summarized in Fig. 11, plotted as ratios tothe total vapor pressure of the two dominant Zr-bearing species, i.e.pMO/p(ZrO+ZrO2). It is reported that for all the other oxides in thisfigure the dominant species is the sub-oxide MO [61]. Also plotted arethe vapor pressures of MO and ZrO over two equi-molar compositions,namely Zr0.5M0.5O1.75, where M=Y, La. (La occurs as a pyrochlorecompound that melts congruently, whereas the equivalent Y compo-sition is a fluorite solid solution and melts incongruently.) Therelevant temperature range is above the liquidus, TL, which is markedin Fig. 11 for 7YSZ and the three zirconates that have attracted themost interest for TBC technology.2

Fig. 11 reveals that the partial pressure of the dopant monoxide isusually higher than the total of the Zr-bearing species, but is within afactor of ≤102 for the relevant temperature range, and also that pMO/p(ZrO+ZrO2)→1 for all MO at some critical temperature above TL. Theimplication is that it should be possible, to evaporate and depositcombinations of these oxides with ZrO2 and preserve the original com-position. Upon closer examination it is evident that pMO/p(ZrO+ZrO2) ~1for Y at the 7YSZ liquidus but diverges progressively from unity in theorder YObGdObSmObLaO. This is qualitatively consistent with theobserved difficulty in preserving composition uniformity duringdeposition here and elsewhere [45,51]. The reality is, however, morecomplicated. Note that pMO diverges more from p(ZrO+ZrO2) for theZr0.5M0.5O1.75 compositions than for MO above the pure M2O3 oxides,suggesting a significant effect of the relative activities of the oxide/sub-oxide species in themelt. Concomitantly, there is a change in the relativepopulations of ZrO and ZrO2 with the addition of the second oxide.Unfortunately, thepaucityof thermodynamic informationprecludes anyfurther analysis of the Zr0.5Sm0.5O1.75 system. Moreover, while a fullsteady state is arguably not achieved, further understanding is neededofthe interplay between melting and evaporation with variations in thebeam power input and/or ingot feeding to explain the oscillations incomposition along the thickness noted in Fig. 3. Process models thatcapture the complexity of the phenomena involved are necessary forthis but do not exist at the present time.

4.2. Metastable fluorite structure

Deposition of SZO from the vapor occurs in a temperature rangewhere the stable structure should be the ordered pyrochlore, asevident in Fig. 1(b), even when accounting for some deviation fromstoichiometry as in the present case. However, XRD reveals no signs ofordering in the as-deposited material, Figs. 6 and 9, although the

Page 7: Vapor deposited samarium zirconate thermal barrier … · Vapor deposited samarium zirconate thermal barrier coatings ... Directed-vapor deposition Thermal barrier ... temperature

Fig. 8. Surface topology (a–c) and polished cross-section (d–f) of SZO coatings deposited at rotation rates of (a and d) zero, (b and e) 0.5 and (c and f) 20 rpm at a substratetemperature of 1000 °C.

3163H. Zhao et al. / Surface & Coatings Technology 203 (2009) 3157–3167

characteristic pyrochlore reflections (331, 511, etc.) are relatively weakand could be further obscured by texturing. This behavior is similar tothat observed during EB-PVD deposition of GZO [48]. Absent cationordering, the structure is a metastable defect fluorite, as shown byremoving the pyrochlore phase field from Fig. 1 and extending therelevant phase boundaries of the fluorite field. Note that in spite of thecompositional heterogeneity the coatings can all be single phasefluorite, although only those with narrower fluctuations can even-tually order to single phase pyrochlore.

The lack of ordering may be understood from consideration of theatomic arrangements needed to generate the pyrochlore structureduring growth, Fig. 12, and the relevant kinetic constraints. Regardlessof texture, active growth takes place on {111} facets and the behavior

is broadly similar to that observed in 7YSZ [53,56,58,63], with someimportant differences. One is that 7YSZ grows directly in the t′ form,where all {111} planes are equivalent (although slightly asymmetricbecause of the tetragonality). In contrast, the cubic pyrochlorestructure has two types of {111} planes with different stoichiometry,namely A3B1 and A1B3, Fig. 12(a,b). The pyrochlore pattern isestablished by stacking six alternating A3B1 and A1B3 planes in aspecific sequence, normal to the ⟨111⟩ directions.When growing alongthe ⟨100⟩ axes with a pyramidal habit, Fig. 12(c,d), three of the four{111} facets bounding the tip have one stoichiometry while the fourthhas the conjugate composition. The distribution is then reversed foreach alternating set of layers. Because the vapor is in principlestoichiometric, A and B atoms (with their bound oxygen's) arrive in

Page 8: Vapor deposited samarium zirconate thermal barrier … · Vapor deposited samarium zirconate thermal barrier coatings ... Directed-vapor deposition Thermal barrier ... temperature

Fig. 9. X-ray diffraction patterns of SZO coatings deposited at a temperature of 1000 °Cwith rotation rates of (a) 0 rpm, (b) 0.5 rpm, and (c) 20 rpm. Compare alsowith coatingsdeposited at 6 rpm in Fig. 6(b).

Fig. 10. Effect of rotation rate on the (a) relative density, and relative thermalconductivity of (b) as-deposited coatings and (c) coatings heat-treated 10 h at 1100 °C.Values for 7YSZ and SZO coatings deposited at the same substrate temperature(1000 °C) are compared, normalized by the corresponding property for the dense oxide.

3164 H. Zhao et al. / Surface & Coatings Technology 203 (2009) 3157–3167

equal proportions at the surface. For the pyrochlore pattern to evolve,the crystallization mechanismmust involve the cooperative growth ofseveral layers and concomitant diffusional rearrangement of thecations. The XRD results indicate that these processes are kinetically

insufficient within the time scale of the deposition. It is possible thatsome local short-range ordering occurs, whereupon the as-depositedcolumns would be conglomerates of small coherent domains withdifferent degrees of partial order.

Page 9: Vapor deposited samarium zirconate thermal barrier … · Vapor deposited samarium zirconate thermal barrier coatings ... Directed-vapor deposition Thermal barrier ... temperature

Fig.11. Reported vapor pressures for oxide species expected during evaporation of Yandrare-earth oxides of interest and the corresponding zirconates.

3165H. Zhao et al. / Surface & Coatings Technology 203 (2009) 3157–3167

4.3. Texture and morphology

In general, the crystallographic aspects of the texture developed inDVD SZO coatings resemble those reported for DVD 7YSZ [53], andPVD GZO [48] albeit not for LZO [51,64]. The latter, however, were

Fig. 12. Atomic arrangement on the two types of (111) planes present in the pyrochlore strcolumn tips for growth along the (001) direction. The darker circles represent the larger (SmNote that the stacking sequence along the (111) direction in the fully ordered pyrochlore in

affected by much larger compositional heterogeneity (up to ±40%)and atypical columnar structures, so they are not suitable forcomparison. DVD SZO coatings rotated at conventional rates(≥6 rpm) exhibit predominantly ⟨200⟩ out-of-plane textures thathave been explained in YSZ as a result of evolutionary selection [65] ofspecial orientations wherein the tip facets can capture equal amountsof flux from the vapor during a rotation cycle [56,58]. These coatingsalso exhibit preferred ⟨002⟩ in-plane orientations, justifiable onsimilar grounds. The texture of the rotated SZO coatings was generallyless developed than that observed for YSZ at the same temperaturesand rotation rates, i.e. minor amounts of other orientations werefound mixed in with the dominant ⟨200⟩ columns. It is likely that thisis due to the significantly smaller thickness of the SZO coatings, ~70±5% of their 7YSZ counterparts, implying that the evolutionaryselection process was more advanced in the latter and the minororientations had been screened out.

Texture selection under stationary deposition is less understoodand more sensitive to temperature than in rotated specimens. Earlierwork on PVD 7YSZ revealed the changes in the population oforientations from ⟨111⟩≫⟨220⟩ at 900 °C to ⟨111⟩≫⟨220⟩N⟨311⟩ at1000 °C, to ⟨220⟩≫⟨311⟩N⟨111⟩ at 1100 °C [56]. By comparison, thestationary DVD YSZ deposited at 1000 °C had a predominant ⟨111⟩texture [53], closer to that seen in PVD coatings at 900 °C, while SMOdeposited under the same conditions exhibits nearly the same patternof orientations as PVD YSZ at 1000 °C. Some of the variability in 7YSZmight be ascribed to differences in the experimental set up, but it ismore likely to result from the higher deposition rate of DVD thatarguably has a similar effect on the time available for rearrangements

ucture (a, b), and the alternating configurations of (111) planes forming the pyramidal3+) cations, and the lighter ones are the (Zr4+) cations. Oxygens are omitted for clarity.volves 6 alternating planes of the types in (a, b) with different in-plane shifts.

Page 10: Vapor deposited samarium zirconate thermal barrier … · Vapor deposited samarium zirconate thermal barrier coatings ... Directed-vapor deposition Thermal barrier ... temperature

Fig. 13. (a) The thermal conductivity and (b) the K factor in Eq. (1) as a function of porevolume fraction (f) for SZO and 7YSZ coatings. The data for 7YSZ coatings are fromreferences [66] and [67].

3166 H. Zhao et al. / Surface & Coatings Technology 203 (2009) 3157–3167

at the surface than a lower temperature.3 The same argument couldnot be used for the differences between DVD 7YSZ and SZO, but in thiscase the coating is significantly thinner in the zirconate, suggestingthat the texture is not as developed as that in 7YSZ, an observationalso made in the rotated specimens.

The samples deposited at very low rotation rates represent a specialcase because of the adjustments that are needed at the column tipfacets as the growth direction changes over a 180° range and a longwavelength each revolution of the substrate. The tips (Fig. 8e) undergoexacerbated segmentation because the features on one facet build upsubstantially on the “sunrise” part of the cycle before the “sunset”portion occurs, biasing the shadowing process. The evolutionaryselection process is arguably delayed, and the segmentation addscomplexity to the possible spectrum of orientations at the surface.

Branching of the columns, particularly evident at the lowerdeposition temperatures (Fig. 4) also complicates texture development.Similareffects havebeenobserved in LZOdeposition [51] andascribed tothe occurrence of re-nucleation, but the coatings in that case did notappear to have any preferred orientation. That is certainly not the casehere, although there are obviously multiple smaller crystallites oftenobserved on the tips of columns, e.g. Fig. 8(b) and (c). Several scenariosare thenpossible. Inone, the “nuclei” formbycluster condensation in thevapor and subsequent landing on the growth surface. In that case theclusters may give rise to mis-oriented branches that are occluded bythe growth of the larger columns, cf. Fig. 4(b) and (c). Alternatively, theclustersmay be able to re-orient themselves shortly after landing on thesurface and give rise to a new but oriented growth center, producing abranch because of shadowing but still being competitive in anevolutionary selection process. Finally, it is also possible that there isheterogeneous nucleation of a new crystal with a twin orientation, butthatwould require local ordering to develop as a separate growth center.

4.4. Porosity and thermal conductivity

With the exception of the coatings on stationary substrates, therelative density (and pore fraction) depends rather weakly ondeposition temperature, Fig. 7(a) and rotation rate, Fig. 10(a). However,the thermal conductivities of both SZO and 7YSZ exhibit a quite strong,nonlinear dependence upon pore volume fraction, Fig. 13(a). The SZOcoatings grown here have thermal conductivities that range from 0.4 to1.1 W/m·K; much lower than that of identically grown 7YSZ coatings.Data for the thermal conductivity of EB-PVD YSZ coatings is shown onFig. 13(a) as a comparison [66,67]. Most of these coatings have a higherdensity and thermal conductivity than the EB-DVD coatings eventhough similar deposition temperatures were used. The differences inconductivity appear to arise almost entirely from the substantiallydifferent intrinsic thermal conductivities of the two materials. Toillustrate this, the thermal conductivity dependence upon pore fractionhas been fitted to an expression of the form:

kcoating = K � kintrinsic � 1− fð Þn ð1Þ

where, κcoating is the coating thermal conductivity, κintrinsic is theintrinsic thermal conductivity of the material from which it is made[38], f is the pore volume fraction and n is an exponent. The ambienttemperature thermal conductivity of fully dense SZO has been mea-sured to be ~2.0 W/mK [38], while that for bulk 7YSZ is ~3.0 W/mK[38]. Combining DVD and PVD data, it is found that the 7YSZ and SZObehaviors in Fig. 13(a) are best fitted by an exponent, n=2.5.

3 Note that the concept of a “homologous” temperature does not work well heresince the melting temperature of SZO is significantly lower than that of 7YSZ, so thesame substrate temperature would imply a higher T/TM and “faster” diffusion. Theargument is obviously too simplistic because diffusion is slower in the pyrochlore, butone must remember that the structure growing is fluorite, where the cationarrangement is random and therefore the effects on diffusion should be diminished.

A plot of the K factor against the pore fraction, Eq. (1), is shown inFig. 13(b). It clearly reveals that K is reasonably close to unity for bothmaterials, from which one may infer that the differences in themeasured conductivity to first order arise from kintrinsic. The separationof the two curves in Fig. 13(b) as the pore fraction increases is likely theresult of the sensitivity of the pore morphology to temperature androtation rate. Note, for example, that the lowest density (highestporosity) corresponds to the lowest rotation rate in Fig. 10(a), whereinthe heavily curved columns near the bottom provide much largerbarriers to heat transfer than the straight intercolumnar gaps in conven-tional structures.

Finally, it is worth noting that the heat treatment seems to have asstrong an effect upon the density and thermal conductivity of the SZODVD coatings as it does on their 7YSZ counterparts. In principle, this isinconsistent with the reportedly higher resistance to sintering ofzirconates [49]. The implication is that the driving force for sinteringmust be dominant, presumably because of a substantially higher initialporosity contents introduced by the DVD process (cf. Figs. 7 and 10).Although the heat treatment tends to increase the thermal conductiv-ity of the coating, the pore shape, distribution and fraction is stillstrongly dependent upon the as-deposited microstructure [68]. Thedifference in thermal conductivity of heat-treated samples stillremains to be relatively larger as the as-deposited samples have.

Page 11: Vapor deposited samarium zirconate thermal barrier … · Vapor deposited samarium zirconate thermal barrier coatings ... Directed-vapor deposition Thermal barrier ... temperature

3167H. Zhao et al. / Surface & Coatings Technology 203 (2009) 3157–3167

5. Conclusion

Sm2Zr2O7 coatings with microstructures suitable for thermalbarrier coatings have been successfully synthesized by EB-DVD. Thecoatings exhibit compositional variation across the thickness of up to±12%SmO1.5, indicative of source instabilities induced in part by thedifference in vapor pressures of the constituent oxides.

SZO coatings grow with a metastable fluorite structure due tokinetically constrained cation ordering within the time scale of thedeposition process. This observation is in agreement with observa-tions on other rare earth zirconates. The coating morphology isbroadly consistent with those produced in 7YSZ and GZO by EB-DVDand EB-PVD, respectively, but becomes increasingly atypical withdecreasing deposition temperature and/or at low rotation rates, bothof which induce extreme branching of the columns. The origin of thebranching is associated with shadowing in the low rotation speci-mens, but is less clear in the case of the lower deposition temperature.

The thermal conductivity of SZO DVD coatings is substantiallylower than that measured on DVD 7YSZ or on EB-PVD Gd zirconate.The former is primarily due to the lower intrinsic conductivity of theSZO, whereas the latter is a result of the much higher porosity contentproduced by the DVD process. This high level of porosity, however,appears to drive pore reconfiguration during high temperatureexposure at the same rate in SZO and 7YSZ, notwithstanding theexpected lower diffusional rates in SZO.

Implementation of DVD SZO in thermal barrier systems appearsviable but is likely to require improved control of the compositionduring growth, as well as a study of the optimum levels of porosity andthe processing parameters needed to achieve them.

Acknowledgements

The Office of Naval Research funded this investigation underContract #N00014-03-1-0297 monitored by Dr. David Shifler. Addi-tional support for activities at UCSB was provided by the NationalScience Foundation under grants DMR-0099695 and DMR-0605700.The authors would like to thank Prof. Ted Bennett and Mr. TylerKakuda of UCSB for making the thermal conductivity measurement,and DavidWortman of GE Aviation for kindly providing the substrates.

References

[1] R.A. Miller, Journal of Thermal Spray Technology 6 (1997) 35.[2] D.R. Clarke, C.G. Levi, Annual Review of Materials Research 33 (2003) 383.[3] A.G. Evans, D.R. Clarke, C.G. Levi, Journal of the European Ceramic Society 28 (2008)

1405.[4] N.P. Padture, M. Gell, E.H. Jordan, Science 296 (2002) 280.[5] A.G. Evans, D.R. Mumm, J.W. Hutchinson, G.H. Meier, F.S. Pettit, Progress in

Materials Science 46 (2001) 505.[6] T.E. Strangman, Thin Solid Films 127 (1985) 93.[7] T.J. Lu, C.G. Levi, H.N.G. Wadley, A.G. Evans, Journal of the American Ceramic

Society 84 (2001) 2937.[8] R.A. Miller, Surface and Coatings Technology 30 (1987) 1.[9] M. Fevre, A. Finel, R. Caudron, R. Mevrel, Physical Review B (Condensed Matter and

Materials Physics) 72 (2005) 104118.[10] C.G. Levi, Current Opinion in Solid State and Materials Science 8 (2004) 77.[11] O. Fabrichnaya, F. Aldinger, Zeitschrift für Metallkunde 95 (2004) 27.[12] S. Stecura, Optimization of the NiCrAl–Y/ZrO2–Y2O3 Thermal Barrier System,

Cleveland, OH, 1985.[13] U. Schulz, Journal of the American Ceramic Society 83 (2000) 904.[14] V. Lughi, V.K. Tolpygo, D.R. Clarke, Materials Science and Engineering A 368 (2004)

212.[15] R.G. Hutchinson, N.A. Fleck, A.C.F. Cocks, Acta Materialia 54 (2006) 1297.[16] D. Zhu, R.A. Miller, Journal of Thermal Spray Technology 9 (2000) 175.[17] J.M. Cairney, N.R. Rebollo, M. Rüler, C.G. Levi, International Journal of Materials

Research 12 (2007) 1177.[18] R.A. Miller, J.L. Smialek, R.G. Garlick, Phase Stability in Plasma-Sprayed, Partially

Stabilized Zirconia-Yttria, in Science and Technology of Zirconia, The AmericanCeramic Society, Inc., Columbus, OH, 1981.

[19] J.R. Brandon, R. Taylor, Surface and Coatings Technology 46 (1991) 75.[20] J.R. VanValzah, H.E. Eaton, Surface and Coatings Technology 46 (1991) 289.

[21] C. Mercer, J.R. Williams, D.R. Clarke, A.G. Evans, Proceedings of the Royal SocietyA—Mathematical Physical and Engineering Sciences, 2007, p. 1393.

[22] T.A. Schaedler, R.M. Leckie, S. Kramer, A.G. Evans, C.G. Levi, Journal of the AmericanCeramic Society 90 (2007) 3896.

[23] M.P. Borom, C.A. Johnson, L.A. Peluso, Surface and Coatings Technology 86–87 (1996)116.

[24] C. Mercer, S. Faulhaber, A.G. Evans, R. Darolia, Acta Materialia 53 (2005) 1029.[25] S. Krämer, J. Yang, C.G. Levi, C.A. Johnson, Journal of the American Ceramic Society

89 (2006) 3167.[26] D.R. Clarke, Surface and Coatings Technology 163–164 (2003) 67.[27] M.J. Maloney, Thermal barrier coating systems and materials, United States Patent

6177200, United Technologies Corporation (Hartford, CT), 2001.[28] J.R. Nicholls, K.J. Lawson, A. Johnstone, D.S. Rickerby, Surface and Coatings

Technology 151–152 (2002) 383.[29] R.L. Jones, Experiences in Seeking Stabilizers for Zirconia Having Hot Corrosion-

Resistance and High Temperature Tetragonal (t') Stability. NRL/MR/6170-96-7841,Naval Research Laboratory, Washington, D.C., 1996.

[30] U. Schulz, C. Leyens, K. Fritscher,M. Peters, B. Saruhan-Brings, O. Lavigne, J.-M. Dorvaux,M. Poulain, R. Mevrel, M. Caliez, Aerospace Science and Technology 7 (2003) 73.

[31] D. Zhu, R.A. Miller, Ceramic Engineering and Science Proceedings (2002).[32] X.Q. Cao, R. Vassen, D. Stoever, Journal of the EuropeanCeramic Society 24 (2004) 1.[33] D.R. Clarke, S.R. Phillpot, Materials Today 8 (2005) 22.[34] W. Pan, C.L. Wan, Q. Xu, J.D. Wang, Z.X. Qu, Thermochimica Acta 455 (2007) 16.[35] H. Lehmann, D. Pitzer, G. Pracht, R. Vassen, D. Stover, Journal of the American

Ceramic Society 86 (2003) 1338.[36] G. Suresh, G. Seenivasan, M.V. Krishnaiah, P.S. Murti, Journal of Alloys and

Compounds 269 (1998) L9.[37] G. Suresh, G. Seenivasan, M.V. Krishnaiah, P. Srirama Murti, Journal of Nuclear

Materials 249 (1997) 259.[38] J. Wu, X. Wei, N.P. Padture, P.G. Klemens, M. Gell, E. Garcia, P. Miranzo, M.I. Osendi,

Journal of the American Ceramic Society 85 (2002) 3031.[39] R. Vassen, X. Cao, F. Tietz, D. Basu, D. Stover, Journal of the American Ceramic

Society 83 (2000) 2023.[40] C. Wang, Experimental and Computational Phase Studies of the ZrO2-based

Systems for Thermal Barrier Coatings, Stuttgart, 2006.[41] X.Q. Cao, R. Vassen, W. Jungen, S. Schwartz, F. Tietz, D. Stover, Journal of the

American Ceramic Society 84 (2001) 2086.[42] R. Subramanian, Thermal barrier coating having high phase stability, US Patent

6258467, Siemens Westinghouse Power Corporation (Orlando, FL), 2001.[43] S. Krämer, J. Yang, C.G. Levi, Journal of the American Ceramic Society 91 (2008)

576.[44] S. Lakiza, O. Fabrichnaya, C. Wang, M. Zinkevich, F. Aldinger, Journal of the

European Ceramic Society 26 (2006) 233.[45] R.M. Leckie, S. Kramer, M. Ruhle, C.G. Levi, Acta Materialia 53 (2005) 3281.[46] J. Wu, N.P. Padture, M. Gell, Scripta Materialia 50 (2004) 1315.[47] U. Bast, E. Schumann, Development of Novel Oxide Materials for TBC's, 2002, p. 525.[48] R.M. Leckie, Fundamental Issues Regarding the Implementation of Gd Zirconate in

Thermal Barrier Systems, University of California, Santa Barbara, Santa Barbara,CA, 2006.

[49] R. Vassen, F. Traeger, D. Stöver, International Journal of Applied CeramicTechnology 1 (2004) 351.

[50] U. Schulz, B. Saruhan, K. Fritscher, C. Leyens, International Journal of Applied CeramicTechnology 1 (2004) 302.

[51] B. Saruhan, P. Francois, K. Fritscher, U. Schulz, Surface and Coatings Technology182 (2004) 175.

[52] J.F. Groves, H.N.G. Wadley, Composites Part B: Engineering 28 (1997) 57.[53] H. Zhao, F. Yu, T.D. Bennett, H.N.G. Wadley, Acta Materialia 54 (2006) 5195.[54] M.U. Cohen, Review of Scientific Instruments 6 (1935) 68.[55] T.D. Bennett, F. Yu, Journal of Applied Physics 97 (2005) 013520.[56] S.G. Terry, Evolution of Microstructure during the Growth of Thermal Barrier

Coatings by Electron-Beam Physical Vapor Deposition, University of California,Santa Barbara, 2001.

[57] U. Schulz, H. Oettel, W. Bunk, Z. Metalkd. 87 (1996) 488.[58] U. Schulz, S.G. Terry, C.G. Levi, Materials Science and Engineering A 360 (2003)

319.[59] D.D. Hass, A.J. Slifka, H.N.G. Wadley, Acta Materialia 49 (2001) 973.[60] M.A. Subramanian, G. Aravamudan, G.V. Subba Rao, Progress in Solid State

Chemistry 15 (1983) 55.[61] N.S. Jacobson, Thermodynamic Properties of Some Metal Oxide-Zirconia Systems,

Lewis Research Center, Cleveland, Ohio, 1989 62.[62] P.G. Partridge, C.M. Ward-Close, International Materials Reviews 38 (1993) 1.[63] U. Schulz, M. Schmucker, Materials Science and Engineering A 276 (2000) 1.[64] Z. Xu, X. Zhong, J. Zhang, Y. Zhang, X. Cao, L. He, Surface and Coatings Technology

202 (2008) 4714.[65] A. Van der Drift, Philips Research Reports 22 (1967) 267.[66] B.K. Jang, H. Matsubara, Scripta Materialia 52 (2005) 553.[67] F. Yu, The Nondestructive Evaluation of Thermal Barrier Coatings: Measurements

of Thermal Properties and Associated Defects, Santa Barbara, University ofCalifornia, 2005.

[68] H.J. Rätzer-Scheibe, U. Schulz, T. Krell, Surface and Coatings Technology 200 (2006)5636.