a mechanistic understanding on rumpling of a nicocraly ...€¦ · the bulk nicocraly alloy with...

28
The University of Manchester Research A mechanistic understanding on rumpling of a NiCoCrAlY bond coat for thermal barrier coating applications DOI: 10.1016/j.actamat.2017.02.003 Document Version Accepted author manuscript Link to publication record in Manchester Research Explorer Citation for published version (APA): Chen, Y., Zhao, X., Bai, M., Yang, L., Li, C., Wang, Y. L., Carr, J., & Xiao, P. (2017). A mechanistic understanding on rumpling of a NiCoCrAlY bond coat for thermal barrier coating applications. Acta Materialia, 128, 31-42. https://doi.org/10.1016/j.actamat.2017.02.003 Published in: Acta Materialia Citing this paper Please note that where the full-text provided on Manchester Research Explorer is the Author Accepted Manuscript or Proof version this may differ from the final Published version. If citing, it is advised that you check and use the publisher's definitive version. General rights Copyright and moral rights for the publications made accessible in the Research Explorer are retained by the authors and/or other copyright owners and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights. Takedown policy If you believe that this document breaches copyright please refer to the University of Manchester’s Takedown Procedures [http://man.ac.uk/04Y6Bo] or contact [email protected] providing relevant details, so we can investigate your claim. Download date:02. Jun. 2021

Upload: others

Post on 26-Jan-2021

3 views

Category:

Documents


0 download

TRANSCRIPT

  • The University of Manchester Research

    A mechanistic understanding on rumpling of a NiCoCrAlYbond coat for thermal barrier coating applicationsDOI:10.1016/j.actamat.2017.02.003

    Document VersionAccepted author manuscript

    Link to publication record in Manchester Research Explorer

    Citation for published version (APA):Chen, Y., Zhao, X., Bai, M., Yang, L., Li, C., Wang, Y. L., Carr, J., & Xiao, P. (2017). A mechanistic understandingon rumpling of a NiCoCrAlY bond coat for thermal barrier coating applications. Acta Materialia, 128, 31-42.https://doi.org/10.1016/j.actamat.2017.02.003

    Published in:Acta Materialia

    Citing this paperPlease note that where the full-text provided on Manchester Research Explorer is the Author Accepted Manuscriptor Proof version this may differ from the final Published version. If citing, it is advised that you check and use thepublisher's definitive version.

    General rightsCopyright and moral rights for the publications made accessible in the Research Explorer are retained by theauthors and/or other copyright owners and it is a condition of accessing publications that users recognise andabide by the legal requirements associated with these rights.

    Takedown policyIf you believe that this document breaches copyright please refer to the University of Manchester’s TakedownProcedures [http://man.ac.uk/04Y6Bo] or contact [email protected] providingrelevant details, so we can investigate your claim.

    Download date:02. Jun. 2021

    https://doi.org/10.1016/j.actamat.2017.02.003https://www.research.manchester.ac.uk/portal/en/publications/a-mechanistic-understanding-on-rumpling-of-a-nicocraly-bond-coat-for-thermal-barrier-coating-applications(95359148-6362-42bf-889c-2822d605c27b).html/portal/ying.chen-2.html/portal/p.xiao.htmlhttps://www.research.manchester.ac.uk/portal/en/publications/a-mechanistic-understanding-on-rumpling-of-a-nicocraly-bond-coat-for-thermal-barrier-coating-applications(95359148-6362-42bf-889c-2822d605c27b).htmlhttps://www.research.manchester.ac.uk/portal/en/publications/a-mechanistic-understanding-on-rumpling-of-a-nicocraly-bond-coat-for-thermal-barrier-coating-applications(95359148-6362-42bf-889c-2822d605c27b).htmlhttps://doi.org/10.1016/j.actamat.2017.02.003

  • 1

    A mechanistic understanding on rumpling of a NiCoCrAlY bond coat for thermal barrier coating applications

    Y. Chen1, X. Zhao2, M. Bai1, L. Yang2, C. Li1, L. Wang1, J. A. Carr1, P. Xiao 1, 2†

    1 School of Materials, University of Manchester, Manchester M13 9PL, UK

    2 Shanghai Key Laboratory of Advanced High-Temperature Materials and Precision Forming,

    Shanghai Jiao Tong University, Shanghai, 200240, PR China

    Abstract

    We present a series of experimental observations on surface rumpling of an initially flat NiCoCrAlY

    coating deposited on a Ni-based superalloy during cyclic oxidation at 1150 °C. The extent of rumpling

    of the coating depends on the thermal history, coating thickness and exposed atmosphere. While the

    coating surface progressively roughens with cyclic oxidation, the bulk NiCoCrAlY alloys with the same

    nominal composition are much less susceptible to rumpling under the same oxidation conditions. The

    coatings, especially the thin ones, experience substantial degradation (e.g. β to γ phase transformation)

    induced by oxidation and coating/substrate interdiffusion. The observations together suggest that

    rumpling of the NiCoCrAlY coating is driven by a combination of the lateral growth of the thermally

    grown oxide and coating/substrate thermal mismatch. The results in this work are further discussed

    and compared with the rumpling behaviour of a β-(Ni,Pt)Al bond coat reported in the literature to

    illustrate the importance of possible factors in governing the development of rumpling in the

    NiCoCrAlY coating.

    Keywords: Rumpling; NiCoCrAlY; Oxidation; Microstructure

    1. Introduction

    Thermal barrier coatings (TBCs) are used in hot sections of gas-turbine engines to protect the

    superalloy components and improve the engine efficiency [1-3]. A typical TBC system consists of a

    ceramic topcoat (typically made of 7-8 wt. % yttria-stabilised zirconia) and an intermediate metallic

    bond coat (made of β-(Ni,Pt)Al, NiCoCrAlY or Pt-diffused γ/γ’ alloys) deposited on a superalloy

    substrate. Upon exposure at operational temperatures, a thermally grown oxide (TGO), predominately

    α-Al2O3, develops between the bond coat and the topcoat. The topcoat gives thermal insulation; the

    † Corresponding author: P. Xiao ([email protected]).

    mailto:[email protected]

  • 2

    TGO provides oxidation resistance; the bond coat promotes the formation of the protective TGO and

    enhances bonding to the substrate; the superalloy bears the loads. The whole TBC system is dynamic

    and all the components interact with each other to control the performance and durability of the TBCs.

    One of the most important forms of TBC degradation is associated with progressive roughening, also

    termed rumpling, of the bond coat surface along with the TGO scale during the course of cyclic

    oxidation [4-9]. The out-of-plane displacement accompanying rumpling gives rise to a tensile stress

    around the undulation peaks and a compressive stress around the undulation valleys, respectively, across

    the TGO/bond coat interface [5], where the TGO scale is compressively stressed in the lateral

    direction. The tensile stress can initiate interfacial cracking and eventually leads to failure of the TBC.

    Surface rumpling of β-(Ni,Pt)Al coatings induced by cyclic oxidation has been extensively investigated

    by experiments and simulations in past decades. These studies have shown that the evolution of

    rumpling depends on oxidation time, oxidation temperature, oxidation mode (cyclic or isothermal),

    oxidation atmosphere (in air or vacuum), sample configuration (bulk alloys or coatings), coating

    thickness, content of reactive elements (e.g. yttrium and hafnium) in the coatings and external load [10-

    16]. Balint and Hutchinson have developed a comprehensive analytical model (Balint-Hutchinson

    model) to simulate rumpling development of the β-(Ni,Pt)Al bond coat during thermal exposure [17].

    The mechanism of rumpling growth in the model is bond coat creep driven by the laterally compressed

    TGO, facilitated by both thermal and stress activation of the creep mechanism, the lateral resulting

    from strain mismatch between the bond coat and substrate and the phase transformation of the bond

    coat. The predictions of the model have been found to be in agreement with most of the experimental

    observations reported in the literature, although the correspondence between the experiments and

    simulations remains to be validated.

    Another type of bond coat, made of NiCoCrAlY alloys, also shows a propensity to rumple after

    exposure at high temperature [18-23], although the coating shows other concomitant degradation

    mechanisms apart from rumpling (e.g. thickness heterogeneities, often referred to as “pegs”,

    accompanied by localised TGO cracks [23-27]). However, few experimental studies have been carried

    out to systematically investigate rumpling of NiCoCrAlY coatings and how their rumpling could be

    affected by experimental variables. The objective of this work, therefore, is to provide a systematic

  • 3

    study of rumpling of NiCoCrAlY coatings. The materials considered in this study include a

    NiCoCrAlY bond coat and a NiCoCrAlY bulk alloy with the same nominal chemical composition.

    Surface rumpling of these materials under a variety of experimental conditions is examined, together

    with the associated TGO stress development and microstructural evolution. The observations are

    further discussed and compared with the rumpling behaviour of a β-(Ni,Pt)Al bond coat reported in

    the literature to illustrate the rumpling mechanism of the NiCoCrAlY coating and the importance of

    possible factors in controlling the rumpling development of the NiCoCrAlY coating.

    2. Materials and methods

    2.1 Sample preparation

    The NiCoCrAlY overlay bond coat was sprayed on one side of Hastelloy® X polycrystalline superalloy

    strips (55× 8 × 1.6mm3) using high velocity oxygen fuel (HVOF) spraying. Table 1 shows the nominal

    compositions of the as-sprayed coating and the Hastelloy® X superalloy substrate. The thickness of

    the as-deposited coating varies between 200 and 230 μm from place to place (Fig.1a), owing to the

    extensive surface roughness formed during the spraying process. The coating consists of two phases

    (the inset at the top right corner of Fig.1a), generally identified as the β-phase (grey contrast) and the γ-

    phase (light contrast).

    The bulk NiCoCrAlY alloy with the same nominal composition as the as-sprayed NiCoCrAlY coating

    was fabricated using spark plasma sintering (SPS). Briefly, the powder was packed in a column graphite

    die with an internal diameter of 28 mm, and densified at 1050°C using a SPS system (FCT-HP D25/4-

    SD) in vacuum (< 100 Pa ) at a load of 31 kN for 10 minutes. The as-sintered NiCoCrAlY alloy is fully

    dense (Fig.1b) and consists of a β-phase (dark contrast) and a γ-phase (light contrast).

    Rectangular (15 × 8 × 1.8 mm3) and disc (28 mm in diameter and 2 mm thick) samples were cut from

    the bond-coated superalloy strips and the bulk NiCoCrAlY alloy bar, respectively, using a SiC abrasive

    cutting blade in a precision cut-off machine (Accutom 5, Struers). The surfaces of the coatings and the

    alloys were progressively ground and polished to a 1 μm finish to remove the existing surface asperities .

    This step produced a similar initial root-mean-square roughness (~ 0.05 μm) on all samples. Some of

    the coatings were mechanically thinned to several thicknesses in order to assess the effect of the

  • 4

    coating thickness on rumpling development. A micrometer was used to monitor the thickness

    reduction during the thinning process and an optical microscope was used to double-check the final

    thickness of the cross-section of the coating edge. Subsequent oxidation experiments showed that

    coatings with thicknesses less than 50 μm tended to have TGO cracking and spallation after cyclic

    oxidation for a period of time (e.g. 50 1-hour cycles at 1150°C) and, therefore, were not used in this

    study. In addition, Vickers micro-hardness indentations were placed into the surfaces of several samples

    as markers so that the rumpling evolution of the identical surface areas with increasing number of

    thermal cycles could be tracked.

    2.2 Thermal treatment

    Cyclic oxidation was performed in laboratory air (1.01 × 105 Pa) between room temperature and

    1150ºC in a CMTM rapid cycle furnace. Each cycle consisted of 10 minute ramping, 1 hour “hot time”

    at 1150ºC and 10 minute fan-assisted air quenching. Some additional sets of cyclic oxidation

    experiments were performed with a shorter “hot time” (10 minutes) at 1150ºC in each cycle, while

    other cycling parameters remained unchanged. A few oxidation experiments were also carried out

    isothermally at 1150ºC. The use of different oxidation regimes is to assess the effect of thermal history

    on rumpling development. Apart from thermal exposure in air, several coating samples were cycled at

    1150ºC in vacuum where the partial pressure of oxygen was significantly reduced (~ 1 Pa). This was

    achieved by wrapping the samples in FeCrAlY foils (Goodfellow, UK) and then sealing them in quartz

    tubes in vacuum. Due to its great resistance to thermal shock, the quartz tube was intact throughout

    thermal exposure and therefore the vacuum was maintained.

    2.3 Characterisation methods

    The surface topography of the samples were characterised using a non-contact optical profilometer

    (MicroXAM, KLA-Tencor). The profilometer provided digital images in the form of surface height, z,

    as a function of lateral position, x and y. The rumpling magnitude is characterised by the root mean

    square roughness, Rq, given by:

    𝑅𝑞 = √1

    𝑛∑ (𝑍𝑖 − �̅�)2

    𝑛

    𝑖=1 (1)

  • 5

    where n is the number of total data points, Zi is the height of each point and �̅� is the mean height for

    the entire measured area. The images obtained were leveled to eliminate the effect of any macroscopic

    slope on the roughness calculation. Since the rumpling patterns show substantial periodicity, the

    wavelength, λ, of rumpling was characterised by the average distance between the adjacent undulation

    peaks or valleys, which is given by:

    𝜆 = 2𝜋

    ∆𝑅𝑞 (2)

    where ∆ is the root mean square slope of the surface, given by:

    ∆ = √1

    𝛿2𝑛∑ ⌈𝑍𝑖+1 − 𝑍𝑖⌉2𝑛𝑖=1 (3)

    and 𝛿 is the spacing between the points (0.02 μm). At least 10 locations were recorded for each sample.

    During thermal cycling, samples were removed from the rig at specific intervals for roughness

    measurements and then returned for further cycling.

    Following the surface roughness and stress measurements, the surface of the samples was carbon-

    coated and examined using a scanning electron microscope (SEM, Quanta 650, FEI). SEM images of

    the surface rumpling morphology were captured at the same tilt angle (60°) so the extent of rumpling

    could be directly compared. Finally, the samples were cross-sectioned using a SiC abrasive cutting blade

    and prepared in cross-section following conventional metallographic procedures. The microstructure

    of the cross-sections was then examined using a combination of SEM and electron backscatter

    diffraction (EBSD, NordlysNano, Oxford Instruments).

    3 Results

    3.1 Effect of thermal history and coating thickness on rumpling

    Fig.2a-c shows the surface morphologies of three NiCoCrAlY coatings after 50 1-hour cycles. No

    TGO spallation is observed throughout the oxidation. The initially planar coating surfaces rumple after

    50 1-hour cycles, but the extent of rumpling shows a dependence on the thickness of the coating: the

    rumpling becomes more pronounced as the thickness of the coating decreases from ~ 180 to ~ 60 μm.

    The observation is further illustrated by the profilometer images in Fig.3 in which the surface

  • 6

    topographies (Fig.3a-b) and typical height profiles of the line segments along the lateral direction

    (Fig.3d) , together with the corresponding Rq values, of a thin (~ 60 μm) and a thick (~ 180 μm)

    coating after 50 1-hour cycles are shown. The roughness of the thin coating is about twice that of the

    thick coating, yet the wavelengths (~ 60 - 70 μm, Table 2) are more or less the same. For comparison,

    the coating surfaces, irrespective of the coating thickness, show far less roughening after 50 hour

    isothermal oxidation (e.g. Fig.2d). This indicates that, for the same cumulative exposure time at 1150°C

    and at least for the oxidation time used in this study, the bond coat is more susceptible to rumpling

    under cyclic oxidation. The statistical rumpling amplitudes and wavelengths, associated with their

    standard deviations, of the samples under different experimental conditions are listed in Table 2.

    Fig. 4a shows the surface morphology of a coating (~ 60 μm thick) after 50 10-minute cycles. While the

    coating surface rumples after 50 10-minute cycles, the extent of rumpling is less pronounced compared

    with that after 50 1-hour cycles (Fig.2c). This observation could be further illustrated by the

    profilometer images in Fig.3 where the surface roughness (Fig.3a and c) and profiles of line

    interceptions (Fig.3d) after 50 1-hour and 10-minute cycles are compared. In short, the observation

    indicates that for a given cycle number the exposure time at peak temperature plays an important role

    in rumpling growth. This suggests that for a given cycle number and coating thickness, a larger net

    TGO growth promotes rumpling development.

    3.2 Rumpling of bulk NiCoCrAlY alloys

    In contrast to the coating deposited on the superalloy substrate, the surface of the bulk NiCoCrAlY

    alloy shows less tendency to rumple after cyclic oxidation. This is further illustrated in Fig.4b in which

    the surface morphology of the alloy after 50 10-minute cycles is presented. The roughness, Rq, on the

    coating surface (Fig.4a and Table 2) is about 2.5 times that on the alloy surface (Fig.4b and Table 2).

    This indicates that while cyclic oxidation can produce rumpling on a bulk NiCoCrAlY as well, the

    rumpling magnitude is considerably smaller than that of the coating subject to the same thermal history,

    which suggests that the presence of the superalloy substrate is essential to rumpling development.

    Observations on rumpling of the bulk NiCoCrAlY alloys after 50 1-hour cycles are not shown since

    the samples have been through catastrophic TGO spallation and reoxidation, thus impeding the

  • 7

    application of the profilometer for the quantitative assessment of roughness. A possible reason for this

    phenomenon is that the TGO cannot undergo enough stress relaxation as surface rumpling of the bulk

    NiCoCrAlY alloys is small, the TGO stresses alleviate through spalling at a sufficiently high level of

    thickness [7]. Another possible reason could be embrittlement of the TGO/alloy interface and

    subsequent low interfacial strength caused by impurity segregation (e.g. Sulphur) during oxidation.

    However, future studies of interface properties are needed to confirm this.

    3.3 Evolution of rumpling with cyclic oxidation

    Fig.5a-e presents a series of profilometer images recorded from the identical regions of a coating

    surface (~ 60 μm thick) to illustrate the evolution of rumpling with increasing number of 10-minute

    cycles. The initially flat surface gradually roughens with increasing number of thermal cycles.

    Comparison of these images reveals that once rumpling is initiated, the areas above the average surface

    plane continue to bow up while the areas below the average surface plane continue to depress during

    the subsequent cyclic oxidation. Fig.5f shows the evolution of the height profile of a line segment (the

    white line in Fig.5a) across the area probed during cyclic oxidation. The height profiles clearly show that

    the coating surface progressively roughens with cycling oxidation (Rq, calculated based on the

    cross-sectional height profile using Eq (1), increases from 0.21 μm after 5 cycles to 0.87 μm after 50

    cycles).

    Compared with the coating samples, rumpling increment on bulk NiCoCrAlY alloy samples is less

    significant with increasing cycle number, as shown in Fig.6a-d where profilometer images recorded

    from identical regions are presented. The surface topography after 10 10-minute cycles shows a

    relatively small change compared with that after 50 10-minute cycles. This is further confirmed by the

    evolution of the height profile of a line segment (the white line in Fig.6a) with cyclic oxidation, as

    shown in Fig.6e where the roughness slightly increases from 0.17 μm after 10 cycles to 0.30 μm after 50

    cycles.

    3.4 Microstructure of coatings

    To date, numerous experiments and simulations conducted to study rumpling of bond coats have

    shown that the thermomechanical properties (e.g. coefficient of thermal expansion (CTE), creep and

  • 8

    yield strength) of the coatings greatly affect rumpling. These properties, in principle, are dictated by the

    intrinsic microstructure of the coating. This relationship between the microstructure and properties

    becomes even more complicated for a bond coat deposited on a superalloy substrate as the

    microstructure of the coating constantly changes with thermal exposure. Therefore, characterisation of

    the microstructure of the NiCoCrAlY coating and how it evolves with thermal cycling are of

    importance to understanding its rumpling development.

    Fig.7a-b shows the cross-sectional microstructure of a thick (~ 180 μm) and a thin coating (~ 60 μm)

    after 50 1-hour cycles. TGO (predominantly α-Al2O3) scales with a thickness of ~ 3.5 μm form on

    both coatings. The thickness of the TGO scale is fairly uniform over the coating surface, suggesting

    that rumpling could not be induced by oxide intrusions or variation of the oxide growth rate from

    place to place. Furthermore, the uniform TGO thickness over the bond coat surface suggests that the

    top surface of the TGO accurately follows the undulations of the bond coat, which in turn confirms

    that the Rq of the TGO is a good indicator of rumpling. Due to the selective oxidation of aluminium

    and the interdiffusion between the coating and the substrate, the β-phase near the TGO/coating

    interface and the coating/substrate interface decomposes to the γ-phase, which is also confirmed by the

    EBSD phase-contrast maps in Fig.7c and d. As a matter of fact, for the thin coating the aluminium

    depletion becomes so extensive that the β-phase could be hardly observed. Quantitative analyses (Fig.7e

    and f) based on cross-sectional images using ImageJ software show that the changes of the volume

    fractions of both phases follow a parabolic law, but the volume fraction of the β-phase in the thin

    coatings (~ 60 μm) decreases around 2.5 times faster than that in the thick coatings (~ 180 μm), which

    suggests that the thin coatings are less resistant to degradation during thermal cycling.

    3.5 Rumpling of coatings after thermal cycling in vacuum

    Fig.8 shows the surface and cross-sectional morphologies of a coating (~ 60 μm thick) after 50 1-hour

    cycles in vacuum. The coating surface rumples (Fig.8a and d), yet the surface roughness (Rq ≈ 1.09 μm,

    Fig.8a and c, Table 2) is lower than that after cycling in air (Rq ≈ 1.26 μm, Fig.2c and 3a, Table 2). The

    thickness of the TGO scale is about 1.6 μm (Fig.8b), which is approximately half the thickness of the

    TGO scale generated in air for the same exposure time (~ 3.5 μm). No β-phase is observed on the

  • 9

    cross-section of the coating after 50 1-hour cycles (Fig.8d). Fig.8e shows an inverse pole figure EBSD

    map of the coating after 50 1-hour cycles where the crystallographic directions of the γ-grains parallel

    to the RD direction (coating surface normal) are colour-coded. Only the γ-phase is indexed throughout

    the entire coating. No correlation has been found between the concave/convex areas and the

    crystallographic orientations of the metal grains underneath the TGO, which suggests that rumpling is

    not associated with the crystallographic anisotropy of the γ-grains, whether through a dependence on

    elastic modulus, yield strength, creep strength or diffusivity. These microstructural features (e.g. grain

    size, grain morphology and phase composition) are extremely similar to that of the thin coating (with

    the same thickness) after 50 1-hour cycles in air (Fig.7b and d). This suggests that, under the same

    thermomechanical history (e.g. 50 1-hour cycles in air and vacuum), the difference in rumpling

    magnitude between cycling in air and in vacuum is not associated with the microstructural difference of

    the coatings. Therefore, the lower surface roughness after cycling in vacuum suggests that the TGO

    growth is an important factor for rumpling development.

    4 Discussion

    4.1 Fundamental rumpling mechanism

    In terms of the geometric profile, surface rumpling indicates overall lateral lengthening of the TGO

    relative to that on a flat surface. The lengthening of the TGO could be associated with a couple of

    mechanisms. One is the lateral growth of the TGO; the other is the out-of-plane distortion of the

    coating surface, resulted from stress-driven plastic deformation of the coating. Unless the TGO

    delaminates from the bond coat, TGO lengthening must occur along with plastic deformation of the

    bond coat. Both mechanisms change the total strain energy of the system. The observations in this

    investigation are used to evaluate the relative contribution of these mechanisms on the observed

    rumpling behaviour.

    The observations above have shown that, for a given cycle number and bond coat thickness, a longer

    dwell time at the peak temperature produces a higher surface roughness and cycling in air produces a

    higher surface roughness than that in vacuum. This indicates that the TGO growth is important to

    rumpling development. However, the facts that isothermal oxidation produces limited surface

  • 10

    roughness on NiCoCrAlY coatings (Fig.2d) and the rumpling magnitude (Fig.4b) or increment rate

    (Fig.6) of the NiCoCrAlY alloy during cyclic oxidation is relatively low during thermal cycling suggest

    that the lateral TGO growth stress alone or thermal ratcheting of TGO (which involves both TGO

    growth stress and TGO/alloy thermal mismatch stress) on an unsupported alloy alone is not sufficient

    for accounting for the substantial rumpling development on the NiCoCrAlY coatings during cyclic

    oxidation. This in turn indicates that the thermal mismatch between the coating and superalloy

    substrate plays a crucial role in rumpling development (see Table 3 for CTEs of the NiCoCrAlY

    coating and Hastelloy X substrate). However, the coating/substrate thermal mismatch alone could not

    explain some observations in this study either. One implication of this mechanism is that, as long as the

    exposure time at the peak temperature is sufficient for fully relaxation of the stress in the coating,

    rumpling would only depend on the cycle number but not the dwell time at the peak temperature. On

    the other hand, it is believed that full relaxation of the stress in the NiCoCrAlY coating should occur

    within 10 minutes as previous stress measurements at 1150 °C have shown that the stress in this

    coating is close to 0 after holding at this temperature for 10 minutes [28]. This is inconsistent with the

    observations in this study, which show that the rumpling magnitude increases with increasing dwell

    time at the peak temperature for a given cycle number. Furthermore, if the coating/substrate thermal

    mismatch is the only single driving force for rumpling, thermal cycling in vacuum would be expected to

    produce a larger, or, at least, an equivalent, rumpling magnitude than cycling in air (for the same coating

    thickness, dwell time at peak temperature and cycle number). This is because the TGO scale in vacuum

    is thinner than that in air and therefore has a lower bending stiffness, which makes it easier to deform

    to accommodate the distortion of the coating surface. However, this is also contradictory to the

    experimental observations in this study. Therefore, it is suggested that rumpling of the NiCoCrAlY

    coating is driven by both TGO lateral growth and coating/substrate thermal mismatch.

    4.2 Effect of coating thickness on rumpling development

    The observations in this work also demonstrate that the rumpling magnitude of the NiCoCrAlY bond

    coat increases with decreasing coating thickness. This dependence of the extent of rumpling on the

    coating thickness is in agreement with the work of Deb et al. However, previous experimental [13] and

    modeling work [29] has suggested that maximum rumpling occurs at intermediate bond coat thickness.

  • 11

    The reason for the discrepancy between these observations is explored below.

    The thermal misfit stress ( ) generated in the bond coat upon cooling is given by (assume no creep

    relaxation):

    =𝐸 (𝛼 − 𝛼𝑆)∆𝑇

    (1 − 𝑣 ) + (1 − 𝑣𝑆)ℎ 𝐸 /(ℎ𝑆𝐸𝑆) (4)

    where E is the elastic modulus; α is the CTE; ΔT is the temperature drop; v is Poisson’s ratio and h is

    the thickness. The subscripts “BC” and “S” refer to the bond coat and substrate, respectively. As the

    thickness of the substrate is about 10 times or more thicker than that of the coating, the stress in the

    coating is expected to be nearly independent of the coating thickness according to Eq.(4). For instance,

    using the CTE and elastic modulus data given in Table 3 and volume fractions in Fig.7e and f (the

    elastic modulus and CTE of the coating are functions of the volume fractions of the γ-phase and β-

    phase), the calculated residual stress ranges from 629.2 to 654.8 MPa in the thick (~ 180 μm) coating

    and from 659.7 to 683.6 MPa in the thin (~ 60 μm) coating, respectively, throughout 50 1-hours cycles.

    The difference in bond coat stress due to thickness variation for a given cycle number is less than 5%,

    which is believed to be not the main reason for the appreciable increase in rumpling as the thickness of

    the coating decreases.

    In the modeling conducted by Balint et al. [29], the thermomechanical properties of the coating are

    taken to be independent of time and thickness over the entire thermal cycling process. However, in the

    case of the NiCoCrAlY/Hastelloy X system used in this study, the coatings, especially the thin ones, are

    subject to extensive aluminium depletion and a substantial β to γ phase transformation after 50 1-hour

    cycles (Fig.7). This degradation could lead to considerable changes of the thermomechanical properties

    of the coatings and compromise the prediction of the simulations using time and thickness

    independent material properties. Specifically, the degradation of the coatings reduces their creep

    strength at high temperature. Creep data from previous experimental tests and simulations have shown

    that the creep strain rate of the polycrystalline β-NiAl is about two or more orders of magnitude

    smaller than that of the polycrystalline γ-Ni [30-32]. According to a number of modelling results [6,

    33-35], the decrease of the creep strength of the coating would promote rumpling growth. Indeed,

  • 12

    experimental studies on coatings with high creep strength (e.g. Pt-diffused γ/γ’ bond coats [36-40] and

    γ/γ’ EQ (Equilibrium) bond coats [20]) have shown negligible rumpling even after extensive thermal

    cycling at sufficiently high temperatures. The decrease of the creep resistance becomes more

    pronounced as the thickness of the coating decreases since a thinner coating undergoes a more

    extensive aluminium depletion and subsequent β to γ transformation during thermal cycling (Fig.7e and

    f). The observations in this work suggests that the relatively rapid loss of the strength of the thin

    coating, caused by the depletion of aluminium, makes it more susceptible to rumpling growth and,

    therefore, play an important role in rumpling development.

    4.3 Comparison of rumpling between the NiCoCrAlY coating and β-(Ni,Pt)Al coating

    This work provides an opportunity to compare the rumpling behaviour of the NiCoCrAlY coating

    with that of the β-(Ni,Pt)Al coating which has been extensively studied in the literature. The Rq of the

    NiCoCrAlY coating after 50 1-hour cycles at 1150 °C ranges from 0.71 to 1.26 μm, increasing with

    decreasing coating thickness. The β-(Ni,Pt)Al coating, however, has a Rq ~ 2.3 μm after exposure under

    the same thermal history [41], which is 2 to 3 times that of the NiCoCrAlY coating. The characteristic

    wavelengths of the rumpling patterns on the two types of coatings, however, are more or less the same

    (~ 70 μm) [41]. To understand the reason for this difference between the rumpling magnitudes of the

    two coatings, it is necessary to compare the respective thermophysical properties of the two coating

    systems first.

    Table 3 lists the CTE, phase composition, elastic modulus and TGO growth stress (TGO yield strength)

    of the β-(Ni,Pt)Al/René N5 and NiCoCrAlY/Hastelloy X coating systems given in the literature [42-

    47]. For the NiCoCrAlY coating, as it has been through a substantial β to γ phase transformation

    during cyclic oxidation at 1150 °C, the CTE of the coating depends on its specific phase composition

    [48]. The same argument also applies to the variation of the elastic modulus of the NiCoCrAlY coating.

    The β-(Ni,Pt)Al bond coat deposited on the René N5 superalloy substrate, however, shows few phase

    transformations from β to γ even after 100 1-hour cycles at 1150 °C [12], and therefore the CTE of the

    coating can be reasonably seen as a constant. On the other hand, previous stress relaxation tests have

    shown that the β-(Ni,Pt)Al coating shows a slower creep rate than that of the NiCoCrAlY coating [49-

  • 13

    51]. This suggests that the β-(Ni,Pt)Al is more creep resistant than the two-phase NiCoCrAlY. The

    TGO growth stress on the NiCoCrAlY coatings is calculated based on the combination of the room

    temperature TGO stress and the thermal-elastic parameters of the TGO and Hastelloy X substrate [52].

    As the magnitude of the TGO growth stress is a result of a dynamic competition between the stress

    generated by the lateral growth strain and any concurrent stress relaxation processes, the TGO growth

    stress is also a good approximation for the yield strength of the TGO at the growing temperature.

    The comparison of the thermophysical properties in Table 3 shows that compared with the β-(Ni,Pt)Al

    coating deposited on the René N5 superalloy substrate, the NiCoCrAlY coating deposited on the

    Hastelloy X substrate shows a higher CTE mismatch with the substrate, a higher elastic modulus, a

    lower creep strength and a higher TGO yield strength when thermally cycled between ambient and

    1150 °C. All these features would lead to a higher rumpling magnitude of the NiCoCrAlY coating for a

    given thermal history based on the predictions of the simulations reported in the literature [17, 33, 34].

    The experimental results in this work, however, show an opposite trend with the prediction of the

    simulations. The discrepancy, therefore, should be accounted for by some other factors.

    In searching for the reasons for the difference in the rumpling magnitudes between the NiCoCrAlY

    coating and β-(Ni,Pt)Al coating, it is helpful to recall other key parameters essential to rumpling

    development. In most simulations conducted in the literature, the lateral growth strain of the TGO is a

    key parameter governing rumpling growth and rumpling is shown to increase with increasing the strain

    incorporated in the models [6].On the other hand, previous experimental studies have shown that the

    addition of minor reactive elements (e.g. yttrium and hafnium) into FeCrAl alloys or β-(Ni,Pt)Al bond

    coats largely reduces surface undulation of the growing oxide [15, 53]. This effect has been ascribed to

    a smaller lateral growth strain in the TGO due to the segregation of reactive elements in the growing

    TGO grain boundaries [53-57]. Hence, a likely explanation of the smaller rumpling magnitude of the

    NiCoCrAlY coating is a smaller lateral growth strain of the TGO resulting from segregation of yttrium

    at the growing TGO grain boundaries.

    In order to examine if the argument above is true or not, it would be helpful to see if there is

    segregation of yttrium at the TGO grain boundaries. To achieve this, the chemistry of the TGO grain

  • 14

    boundaries was analysed using energy dispersive spectroscopy (EDS, X-MaxN 80T, Oxford Instruments)

    equipped on a transmission electron microscope (TEM, TecnaiTM G2, FEI). The TEM sample was

    prepared by a focused ion beam (FIB, Quanta 3D, FEI) cutting through the TGO surface followed by

    an in-situ lift-out technique. Fig.9 shows a bright-field image (Fig.9a) and the corresponding STEM

    annular dark-field image (Fig.9b) of the TGO on a NiCoCrAlY coating (~ 180 μm thick) after 2 1-hour

    cycles. The TGO is about 0.8 μm thick, with its microstructure featured by the outer portion composed

    of equiaxed grains and the inner portion composed of columnar grains. An EDS linescan profile (the

    inset in Fig.9b) across a couple of TGO grain boundaries shows that the signals from yttrium bump up

    when the linescan intersects with either of the grain boundaries, suggesting the presence of yttrium

    segregation at the TGO grain boundaries.

    It should be finally noted that Jackson et al [58, 59], however, recently have shown that rumpling of Hf-

    doped B2 bond coats is comparable to the Hf-free β-(Ni,Pt)Al coating during thermal cycling. The

    reasons for this are not completely understood, but could be probably explained from two aspects. First,

    apart from the thermomechanical properties of the coating system, rumpling is also affected by other

    parameters (e.g. initial roughness and wavelength [6, 11, 17]). On the other hand, the sustenance of the

    dynamic segregation of reactive elements at the TGO grain boundaries and its subsequent effect on

    rumpling is dictated by the reservoir of the elements in the underlying coating and its depletion during

    thermal exposure [60]. After exposed at 1150 °C or above for sufficiently long time, the reactive

    elements in the bond coat could be depleted to an extent that its concentration is so low that its

    beneficial effect on rumpling becomes insignificant during subsequent thermal cycling.

    5 Summary

    The surface of a NiCoCrAlY coating deposited on a Ni-based superalloy progressively roughens during

    cyclic oxidation. The rumpling magnitude depends on thermal history, coating thickness and oxidation

    atmosphere. Compared with the coating, the bulk NiCoCrAlY alloy with the same composition shows

    a smaller tendency to rumple after thermal cycling. The coatings, especially the thin ones, experience

    substantial degradation (e.g. β to γ phase transformation) induced by oxidation and coating/substrate

    interdiffusion. The observations together suggest that rumpling of the NiCoCrAlY coating is driven by

  • 15

    a combination of the lateral growth of the thermally grown oxide and coating/substrate thermal

    mismatch. It is suggested that the chemical degradation of the NiCoCrAlY coatings, associated with

    their microstructural/mechanical degradation, promotes rumpling growth. This effect is more

    pronounced in a thinner coating as it is through a more substantial degradation for a given thermal

    history. Compared with the β-(Ni,Pt)Al coating deposited on the René N5 superalloy substrate, the

    NiCoCrAlY deposited on the Hastelloy X superalloy shows a rumpling magnitude 2-3 times lower,

    which is due to the smaller TGO growth strain resulted from segregation of yttrium at the growing

    TGO grain boundaries.

    Acknowledgment

    The authors would like to acknowledge Nicholas Curry, Nicolaie Markocsan, Per Nylen from

    Production Technology Centre, University West, Sweden for supply of the coating samples. The

    authors are also grateful to Mr. David Gordon for the help in vacuum sealing and Professor Brain

    Ralph for proof-reading the manuscript.

    References 1. D.R. Clarke and C.G. Levi, Materials design for the next generation thermal barrier coatings. Annual Review

    of Materials Research, 2003. 33. 383-417.

    2. D.R. Clarke, M. Oechsner, and N.P. Padture, Thermal-barrier coatings for more efficient gas-turbine engines.

    Mrs Bulletin, 2012. 37(10). 891-902.

    3. N.P. Padture, M. Gell, and E.H. Jordan, Thermal barrier coatings for gas-turbine engine applications. Science,

    2002. 296(5566). 280-284.

    4. A.G. Evans, M.Y. He, and J.W. Hutchinson, Mechanics-based scaling laws for the durability of thermal

    barrier coatings. Progress in Materials Science, 2001. 46(3-4). 249-271.

    5. A.G. Evans, D.R. Mumm, J.W. Hutchinson, G.H. Meier, and F.S. Pettit, Mechanisms controlling the durability

    of thermal barrier coatings. Progress in Materials Science, 2001. 46(5). 505-553.

    6. M.Y. He, A.G. Evans, and J.W. Hutchinson, The ratcheting of compressed thermally grown thin films on

    ductile substrates. Acta Materialia, 2000. 48(10). 2593-2601.

    7. B. Heeg, V.K. Tolpygo, and D.R. Clarke, Damage Evolution in Thermal Barrier Coatings with Thermal

    Cycling. Journal of the American Ceramic Society, 2011. 94. S112-S119.

    8. V.K. Tolpygo, D.R. Clarke, and K.S. Murphy, Evaluation of interface degradation during cyclic oxidation of

    EB-PVD thermal barrier coatings and correlation with TGO luminescence. Surface & Coatings Technology,

    2004. 188. 62-70.

    9. V. Tolpygo and D.R. Clarke, Morphological evolution of thermal barrier coatings induced by cyclic oxidation.

    Surface & Coatings Technology, 2003. 163. 81-86.

    10. S. Dryepondt and D.R. Clarke, Effect of superimposed uniaxial stress on rumpling of platinum-modified nickel

    aluminide coatings. Acta Materialia, 2009. 57(7). 2321-2327.

    11. V.K. Tolpygo and D.R. Clarke, Surface rumpling of a (Ni, Pt)Al bond coat induced by cyclic oxidation. Acta

    Materialia, 2000. 48(13). 3283-3293.

    12. V.K. Tolpygo and D.R. Clarke, On the rumpling mechanism in nickel-aluminide coatings - Part I: an

    experimental assessment. Acta Materialia, 2004. 52(17). 5115-5127.

    13. V.K. Tolpygo and D.R. Clarke, Rumpling induced by thermal cycling of an overlay coating: the effect of

    coating thickness. Acta Materialia, 2004. 52(3). 615-621.

    14. V.K. Tolpygo and D.R. Clarke, Rumpling of CVD (Ni,Pt)Al diffusion coatings under intermediate temperature

  • 16

    cycling. Surface & Coatings Technology, 2009. 203(20-21). 3278-3285.

    15. V.K. Tolpygo, K.S. Murphy, and D.R. Clarke, Effect of Hf, Y and C in the underlying superalloy on the

    rumpling of diffusion aluminide coatings. Acta Materialia, 2008. 56(3). 489-499.

    16. P. Deb, D.H. Boone, and T.F. Manley, Surface Instability of Platinum Modified Aluminide Coatings during

    1100-Degrees-C Cyclic Testing. Journal of Vacuum Science & Technology a-Vacuum Surfaces and Films,

    1987. 5(6). 3366-3372.

    17. D.S. Balint and J.W. Hutchinson, An analytical model of rumpling in thermal barrier coatings. Journal of the

    Mechanics and Physics of Solids, 2005. 53(4). 949-973.

    18. D.S. Balint, S.S. Kim, Y.F. Liu, R. Kitazawa, Y. Kagawa, and A.G. Evans, Anisotropic TGO rumpling in EB-

    PVD thermal barrier coatings under in-phase thermomechanical loading. Acta Materialia, 2011. 59(6). 2544-

    2555.

    19. C. Mercer, D. Hovis, A.H. Heuer, T. Tomimatsu, Y. Kagawa, and A.G. Evans, Influence of thermal cycle on

    surface evolution and oxide formation in a superalloy system with a NiCoCrAlY bond coat. Surface &

    Coatings Technology, 2008. 202(20). 4915-4921.

    20. C. Mercer, K. Kawagishi, T. Tomimatsu, D. Hovis, and T.M. Pollock, A comparative investigation of oxide

    formation on EQ (Equilibrium) and NiCoCrAlY bond coats under stepped thermal cycling. Surface & Coatings

    Technology, 2011. 205(8-9). 3066-3072.

    21. R. Panat and K.J. Hsia, Experimental investigation of the bond-coat rumpling instability under isothermal and

    cyclic thermal histories in thermal barrier systems. Proceedings of the Royal Society a-Mathematical Physical

    and Engineering Sciences, 2004. 460(2047). 1957-1979.

    22. R.C. Pennefather and D.H. Boone, Mechanical degradation of coating systems in high-temperature cyclic

    oxidation. Surface & Coatings Technology, 1995. 76(1-3). 47-52.

    23. Y.H. Sohn, J.H. Kim, E.H. Jordan, and M. Gell, Thermal cycling of EB-PVD/MCrAlY thermal barrier

    coatings: 1. Microstructural development and spallation mechanisms. Surface & Coatings Technology, 2001.

    146. 70-78.

    24. C. Mercer, S. Faulhaber, N. Yao, K. McIlwrath, and O. Fabrichnaya, Investigation of the chemical composition

    of the thermally grown oxide layer in thermal barrier systems with NiCoCrAlY bond coats. Surface & Coatings

    Technology, 2006. 201(3-4). 1495-1502.

    25. D.R. Mumm and A.G. Evans, On the role of imperfections in the failure of a thermal barrier coating made by

    electron beam deposition. Acta Materialia, 2000. 48(8). 1815-1827.

    26. G.M. Kim, N.M. Yanar, E.N. Hewitt, F.S. Pettit, and G.H. Meier, The effect of the type of thermal exposure on

    the durability of thermal barrier coatings. Scripta Materialia, 2002. 46(7). 489-495.

    27. N.M. Yanar, G.H. Meier, and F.S. Pettit, The influence of platinum on the failure of EBPVD YSZ TBCs on

    NiCoCrAlY bond coats. Scripta Materialia, 2002. 46(4). 325-330.

    28. Y. Chen, X. Zhao, Y. Dang, P. Xiao, N. Curry, N. Markocsan, and P. Nylen, Characterization and

    understanding of residual stresses in a NiCoCrAlY bond coat for thermal barrier coating application. Acta

    Materialia, 2015. 94. 1-14.

    29. D.S. Balint, T. Xu, J.W. Hutchinson, and A.G. Evans, Influence of bond coat thickness on the cyclic rumpling

    of thermally grown oxides. Acta Materialia, 2006. 54(7). 1815-1820.

    30. U. Hermosilla, M.S.A. Karunaratne, I.A. Jones, T.H. Hyde, and R.C. Thomson, MCrAlY creep behaviour

    modelling by means of finite-element unit cells and self-consistent constitutive equations. Proceedings of the

    Institution of Mechanical Engineers, Part L: Journal of Materials Design and Applications, 2009. 223(1). 41-

    51.

    31. R.D. Noebe, R.R. Bowman, and M.V. Nathal, Physical and Mechanical-Properties of the B2 Compound Nial.

    International Materials Reviews, 1993. 38(4). 193-232.

    32. H.J. Frost and M.F. Ashby, Deformation-mechanism maps: the plasticity and creep of metals and ceramics.

    1982.

    33. A.M. Karlsson, J.W. Hutchinson, and A.G. Evans, A fundamental model of cyclic instabilities in thermal

    barrier systems. Journal of the Mechanics and Physics of Solids, 2002. 50(8). 1565-1589.

    34. A.M. Karlsson, J.W. Hutchinson, and A.G. Evans, The displacement of the thermally grown oxide in thermal

    barrier systems upon temperature cycling. Materials Science and Engineering a-Structural Materials Properties

    Microstructure and Processing, 2003. 351(1-2). 244-257.

    35. T. Xu, M.Y. He, and A.G. Evans, A numerical assessment of the durability of thermal barrier systems that fail

    by ratcheting of the thermally grown oxide. Acta Materialia, 2003. 51(13). 3807-3820.

    36. B. Gleeson, Thermal barrier coatings for aeroengine applications. Journal of Propulsion and Power, 2006.

    22(2). 375-383.

    37. A. Selcuk and A. Atkinson, The evolution of residual stress in the thermally grown oxide on Pt diffusion bond

    coats in TBCs. Acta Materialia, 2003. 51(2). 535-549.

    38. X. Wang, G. Lee, and A. Atkinson, Investigation of TBCs on turbine blades by photoluminescence

    piezospectroscopy. Acta Materialia, 2009. 57(1). 182-195.

    39. L.T. Wu, R.T. Wu, X. Zhao, and P. Xiao, Microstructure parameters affecting interfacial adhesion of thermal

    barrier coatings by the EB-PVD method. Materials Science and Engineering a-Structural Materials Properties

  • 17

    Microstructure and Processing, 2014. 594. 193-202.

    40. D.J. Jorgensen, A. Suzuki, D.M. Lipkin, and T.M. Pollock, Bond coatings with high rumpling resistance:

    Design and characterization. Surface and Coatings Technology, 2016. 300. 25-34.

    41. V.K. Tolpygo and D.R. Clarke, On the rumpling mechanism in nickel-aluminide coatings - Part II:

    characterization of surface undulations and bond coat swelling. Acta Materialia, 2004. 52(17). 5129-5141.

    42. HASTELLOY X TECHNICAL DATA. http://www.hightempmetals.com/techdata/hitempHastXdata.php, 2015.

    43. P.Y. Hou, A.P. Paulikas, and B.W. Veal, Growth strains in thermally grown Al2O3 scales studied using

    synchrotron radiation. Jom, 2009. 61(7). 51-55.

    44. B.W. Veal, A.P. Paulikas, B. Gleeson, and P.Y. Hou, Creep in alpha-Al(2)O3 thermally grown on beta-NiAl and

    NiAlPt alloys. Surface & Coatings Technology, 2007. 202(4-7). 608-612.

    45. B.W. Veal, A.P. Paulikas, and P.Y. Hou, Creep in protective alpha-Al2O3 thermally grown on beta-NiAl.

    Applied Physics Letters, 2007. 90(12).

    46. A.H. Heuer, A. Reddy, D.B. Hovis, B. Veal, A. Paulikas, A. Vlad, and M. Ruhle, The effect of surface

    orientation on oxidation-induced growth strains in single crystal NiAl: An in situ synchrotron study. Scripta

    Materialia, 2006. 54(11). 1907-1912.

    47. B.S. Kang, C. Feng, J.M. Tannenbaum, and M.A. Alvin, A Load-based Micro-indentation Technique for

    Mechanical Property and NDE Evaluation. Proceedings of the SEM Annual Conference, June 1-4, 2009

    Albuquerque New Mexico USA, 2009.

    48. J.A. Haynes, B.A. Pint, W.D. Porter, and I.G. Wright, Comparison of thermal expansion and oxidation

    behavior of various high-temperature coating materials and superalloys. Materials at High Temperatures,

    2004. 21(2). 87-94.

    49. D. Pan, M.W. Chen, P.K. Wright, and K.J. Hemker, Evolution of a diffusion aluminide bond coat for thermal

    barrier coatings during thermal cycling. Acta Materialia, 2003. 51(8). 2205-2217.

    50. W.J. Brindley and J.D. Whittenberger, Stress-Relaxation of Low-Pressure Plasma-Sprayed Nicraly Alloys.

    Materials Science and Engineering a-Structural Materials Properties Microstructure and Processing, 1993.

    163(1). 33-41.

    51. K.J. Hemker, B.G. Mendis, and C. Eberl, Characterizing the microstructure and mechanical behavior of a two-

    phase NiCoCrAlY bond coat for thermal barrier systems. Materials Science and Engineering a-Structural

    Materials Properties Microstructure and Processing, 2008. 483-84. 727-730.

    52. Y. Chen and P. Xiao, To be submitted for publication.

    53. V.K. Tolpygo and D.R. Clarke, Wrinkling of alpha-alumina films grown by thermal oxidation - I. Quantitative

    studies on single crystals of Fe-Cr-Al alloy. Acta Materialia, 1998. 46(14). 5153-5166.

    54. D.R. Clarke, The lateral growth strain accompanying the formation of a thermally grown oxide. Acta

    Materialia, 2003. 51(5). 1393-1407.

    55. B.A. Pint, Experimental observations in support of the dynamic-segregation theory to explain the reactive-

    element effect. Oxidation of Metals, 1996. 45(1-2). 1-37.

    56. F.H. Stott, G.C. Wood, and J. Stringer, The Influence of Alloying Elements on the Development and

    Maintenance of Protective Scales. Oxidation of Metals, 1995. 44(1-2). 113-145.

    57. B.A. Pint, Optimization of reactive-element additions to improve oxidation performance of alumina-forming

    alloys. Journal of the American Ceramic Society, 2003. 86(4). 686-695.

    58. R.W. Jackson, D.M. Lipkin, and T.M. Pollock, The oxidation and rumpling behavior of overlay B2 bond coats

    containing Pt, Pd, Cr and Hf. Surface & Coatings Technology, 2013. 221. 13-21.

    59. R.W. Jackson, D.M. Lipkin, and T.M. Pollock, Thermal barrier coating adherence to Hf-modified B2 NiAl

    bond coatings. Acta Materialia, 2014. 80. 39-47.

    60. B.A. Pint and K.L. More, Characterization of alumina interfaces in TBC systems. Journal of Materials

    Science, 2009. 44(7). 1676-1686.

    http://www.hightempmetals.com/techdata/hitempHastXdata.php

  • 18

    Fig.1

    Fig.1 Microstructure of the NiCoCrAlY bond coat and bulk NiCoCrAlY alloy: (a) optical image of the

    cross-section of the as-deposited bond coat; the inset at the top right corner of Fig.1a is a high-

    magnification backscattered electron (BSE) image showing that the bond coat consists of β-phase (grey

    contrast), γ-phase (light contrast) and interfacial pores (dark contrast) between the NiCoCrAlY particles;

    (b) BSE image of the as-sintered bulk NiCoCrAlY alloy; the dark contrast is the β-phase and the light

    contrast γ-phase

  • 19

    Fig.2

    Fig.2 Surface morphology of (a) a thick coating (~ 180 μm), (b) a coating with an intermediate

    thickness (~ 120 μm) and (c) a thin coating (~ 60 μm) after 50 1-hour cycles; (d) surface morphology

    of a coating (~ 60 μm thick) after 50 hours isothermal oxidation

  • 20

    Fig.3

    Fig.3 Profilometer images of three coatings after cyclic oxidation: (a) a thin coating (~ 60 μm) after 50

    1-hour cycles; (b) a thick coating (~ 180 μm) after 50 1-hour cycles and (c) a thin coating (~ 60 μm)

    after 50 10-minute cycles. (d) height profiles of the white lines in a-c.

  • 21

    Fig.4

    Fig.4 Surface morphology of (a) a NiCoCrAlY coating (~ 60 μm thick) and (b) a bulk NiCoCrAlY alloy

    after 50 10-minute cycles

  • 22

    Fig.5

    Fig.5 Profilometer images of a NiCoCrAlY coating surface after (a) 5 10-minute cycles, (b) 10 10-

    minute cycles, (c) 20 10-minute cycles, (d) 35 10-minute cycles and (e) 50 10-minute cycles recorded at

    identical regions, as illustrated by the indentation marks. (f) The evolution of the surface profile and

    corresponding Rq of the line interception shown in a.

  • 23

    Fig.6

    Fig.6 Profilometer images of a bulk NiCoCrAlY alloy surface after (a) 10 10-minute cycles, (b) 20 10-

    minute cycles, (c) 35 10-minute cycles and (d) 50 10-minute cycles and recorded at identical regions, as

    illustrated by the indentation mark. (e) The evolution of the surface profile and corresponding Rq of

    the line interception shown in d.

  • 24

    Fig.7

    Fig.7 BSE images of the cross sections of (a) a thick coating (~ 180 μm) and (b) a thin coating (~ 60

    μm) after 50 1-hour cycles; (c-d) EBSD phase-contrast maps of the thick and the thin coating after 50

    1-hour cycles, respectively. The γ-phase is coded with red and the β-phase blue; (e-f) evolution of the

    volume fractions of the γ-phase and the β-phase in the thick (~ 180 μm) and thin (~ 60 μm) coatings,

    respectively, with thermal cycling. The volume fractions are estimated from the area ratios of the two

    phases by processing cross-sectional images of the coatings using ImageJ software

  • 25

    Fig.8

    Fig.8 Surface and cross-sectional morphology of a coating (~ 60 μm thick) after 50 1-hour cycles in

    vacuum: (a) surface morphology, (b) fractured cross-section of the TGO at the edge of surface, (c)

    profilometer image, (d) general view of the cross section and (e) an EBSD map of the coating where

    the crystallographic directions of the γ-grains parallel to RD are colour-coded

  • 26

    Fig.9

    Fig.9 TEM analyses of the TGO after 2 1-hour cycles. (a) bright-filed image, (b) STEM annular dark-

    field image. The inset in (b) shows an EDS linescan profile across a couple of TGO grain boundaries.

    The signal from yttrium bumps up when the linescan intersect with the grain boundaries.

  • 27

    Table 1 Chemical compositions (wt. %) of Hastelloy® X superalloy and NiCoCrAlY coating

    Ni Cr Co Al Y Fe Mo W C Mn Si B

    Hastelloy® X Bal 22 1.5 / / 18 9 0.6 0.1 1* 1* 0.008* NiCoCrAlY Bal 17 23 12.5 0.6 / / / / / / /

    * Maximum

    Table 2 Roughness (Rq ) and wavelength (λ) of coatings after thermal cycling

    Cycle regime Sample configuration Atmosphere exposed Rq (μm) λ (μm)

    50 × 1-hour Coating (~ 180 μm thick) Air 0.71 ± 0.07 64.39 ± 3.52 50 × 1-hour Coating (~ 120 μm thick) Air 0.99 ± 0.09 66.46 ± 3.78 50 × 1-hour Coating (~ 60 μm thick) Air 1.26 ± 0.11 65.13 ± 4.56 50 × 10-minute Coating (~ 60 μm thick) Air 0.85 ± 0.08 67.16 ± 3.89 50 × 10-minute Bulk alloy Air 0.35 ± 0.06 69.23 ± 5.21 50 × 1-hour Coating (~ 60 μm thick) Vacuum 1.09 ± 0.11 66.17 ± 4.16

    Table 3 Thermophysical properties of the β-(Ni,Pt)Al/ René N5 and NiCoCrAlY/Hastelloy X coating systems. α is the mean CTE over the temperature range between room temperature and 1150 °C. E is

    the elastic modulus at room temperature. σ𝑇𝐺𝑂𝐺 is the TGO growth stress at 1150 °C, which is also

    taken as the TGO yield strength at 1150 °C. The mean CTE of the Hastelloy X is obtained by fitting the mean CTEs given in Ref. [42] using a quadratic polynomial and extrapolating it to 1150 °C.

    α ( ×10-6/°C) Phase

    composition E (GPa)

    𝛔𝑻𝑮𝑶𝑮 /TGO yield strength

    (GPa)

    NiCoCrAlY 19.3 ~ 20.3

    [48] γ+β

    155 ~ 200 [51]

    ~ 1 [52]

    β-(Ni,Pt)Al 16.0 [48] β 117 [49] ~ 0 - 0.4 [43-46] Hastelloy

    X 17.1 [42] γ 196 [42] /

    René N5 17.2 [48] γ/γ’ 145 [47] /