Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Received: 3 October 2017

| Accepted: 19 December 2017

DOI: 10.1111/jace.15409

ORIGINAL ARTICLE

3D characterization of porosity in an air plasma-sprayed


thermal barrier coating and its effect on thermal conductivity

Xun Zhang1,2 | Justyna Kulczyk-Malecka2,3 | James Carr1,2 | Ping Xiao2 |


Philip J. Withers1,2

1
Henry Moseley X-ray Imaging Facility,
School of Materials, The University of Abstract
Manchester, Manchester, UK Air plasma spray (APS) thermal barrier coatings (TBCs) contain pores, cracks
2
School of Materials, The University of and splat interfaces that are preferentially aligned normal to the heat flux direc-
Manchester, Manchester, UK
3
tion. These significantly reduce (by as much as 50%) the thermal conductivity
Surface Engineering Group, Manchester
Metropolitan University, Manchester, UK
over a fully dense coating. Here the microstructures of APS thermal barrier coat-
ings (TBCs) have been characterized in 3D using X-ray microcomputed tomogra-
Correspondence
phy (l-CT). Pores larger than 2.8 lm3 are resolved and their contribution to
Ping Xiao, School of Materials, The
University of Manchester, Manchester, thermal conductivity reduction is evaluated using image-based microstructurally
UK. realistic numerical models. However, the models overestimate (by almost 50%)
Email: p.xiao@manchester.ac.uk
the measured thermal conductivity of the as-deposited TBC sample. It is shown
Funding information that this discrepancy is due to very fine (micrometer and submicrometer) cracks
EPSRC, Grant/Award Number: EP/ which readily sinter after short-term exposure to temperatures representative of
F007906, EP/I02249X, EP/F028431;
TBC operating conditions causing the measured thermal conductivity to rise to
HEFCE; European Research Council,
Grant/Award Number: 695638 within 15% of that predicted. This suggests that under the realistic service condi-
tions, the 3D image-based models based on CT images provide a good indicator
of the likely long-term TBC performance. Virtual experiments showed that for the
retained pores, the larger pores which are flatter and more oriented in the plane of
the APS splats, contribute disproportionately to the beneficial reduction in thermal
conductivity. Our results demonstrate that X-ray imaging is a useful tool in estab-
lishing APS process conditions that give rise to a beneficial distribution of such
pores.

KEYWORDS
high temperature, image-based modeling, turbine blade, yttria-stabilized zirconia

1 | INTRODUCTION Among the many techniques to deposit TBCs, air plasma


spray (APS) has been widely used due to its relatively low
Thermal barrier coatings (TBCs) are widely used to pro- cost and versatility.1 During plasma spraying, the ceramic
vide thermal insulation and, in combination with a bond YSZ particles are melted and ejected toward the target as
coat, oxidation resistance for metallic turbine blades.1 The droplets. Upon impinging on the surface of the component,
most commonly used material for TBC applications is these droplets solidify rapidly and form splats.3 Conse-
yttria-stabilized zirconia (YSZ) which has a low thermal quently, APS TBCs have a splat-based microstructure with
conductivity (~2.5 W∕(mK)2). On top of this, the presence pores trapped in between the splats. In addition, the rapid
of pores has been found to further enhance the insulating solidification of the splat may induce cracking within the
performance of the coating.1 splat due to the quenching stress,4 whereas globular pores

J Am Ceram Soc. 2018;1–11. wileyonlinelibrary.com/journal/jace © 2017 The American Ceramic Society | 1


2
| ZHANG ET AL.

may arise around the entrained unmelted particles.5,6 As a use this image-based modeling approach to better under-
result, the overall porosity of APS TBCs typically lie within stand the process by which the thermal conductivity of a
15%~25%.1 Crucially, the plasma spray process produces TBC increases upon short-term thermal exposure and to
pores, cracks, and splat interfaces that are preferentially better understand how the size, distribution, morphology,
aligned perpendicular to the heat flux direction further and orientation of pores and cracks contribute to the ther-
enhancing the insulating effect. As a result, practical APS mal conductivity. This information will help to optimize
TBCs can have a thermal conductivity just 40% of the bulk the fabrication of APS TBCs to gain greatest advantage
conductivity.2 It is estimated that at 1500 K, contribution from the porosity thus introduced.
from radiative heat transfer is around 0.05~0.1 W∕(mK),
taking up <10% of the overall effective thermal conductiv-
ity.7 Therefore, even at a high operating temperature
2 | EXPERIMENTAL METHODS
(>1300°C),1 thermal conduction, which is sensitive to the
porosity within the coating, is still the dominant mechanism
2.1 | Coating deposition
of heat transfer through the TBCs. Porous 8 wt.% YSZ coatings having a thickness of
The correlation between porous structures and their ~450 lm were deposited by APS onto a Hastelloyâ X but-
effective thermal conductivity has been studied both analyti- ton having a diameter of 25.4 mm. The thickness of the
cally and experimentally.2,4,8-13 Specifically for APS TBCs, button was ~6.4 mm. The button surface was first grit
their highly complex and anisotropic microstructures have blasted using alumina grit 60 size. Prior to YSZ deposition,
made the microstructure-conductivity relationship difficult a NiCoCrAlY (AMPERITâ 410) overlay bond coat was
to determine.13 This requires a comprehensive 3 dimen- applied to the substrate, also by APS, to a thickness of
sional (3D) description of the TBC microstructure. Kulkarni ~150 lm using a Sulzer Metco Triplex Plasma Spray
et al4 have volume averaged defect properties in TBCs Guns. Finally, the YSZ topcoat was sprayed using a Sulzer
using anisotropic multiple small angle neutron scattering Metco Triplex Pro gun in Forschungszentrum J€ulich (IEK-
(MSANS). Deshpande et al14 used 2D image analysis to 1). A standard plasma gas composition of 46 standard liters
characterize the pores within the coatings. The validity of per minute (slpm) Ar and 4 slpm He was used19 with a
this image-based technique was authenticated by volume current of 420 A and a spraying distance of 120 mm.
averaged SANS data. To correlate the microstructural fea-
tures with the coating’s conductivity, predictive FE models
have been proposed based on the volumetrically averaged
2.2 | Thermal conductivity measurement
microstructural information determined from MSANS4 as The effective thermal conductivity (K) of the APS TBC
well as micrographs of the coatings.2,15 In a follow-up was determined at room temperature by laser flash analy-
study, Chi et al13 have probed the thermal conductivity-por- sis.20 In this analysis, a disk sample is heated on one side
osity relationship for various coatings at both ambient and using an unfocused laser pulse. In the meantime, the tem-
elevated temperatures. Their results suggest that closely perature on the other side of the sample is recorded. The
spaced interfaces in the coatings can have a strong effect on thermal diffusivity (a) along the thickness direction of the
their effective thermal conductivity although such interfaces sample can be calculated based on the temperature evolu-
only take up a small fraction of the total porosity. The sig- tion profile. According to Parker et al,20 the thermal diffu-
nificant and rapid increase in thermal conductivity upon sivity of the sample at temperature T can be calculated
thermal exposure has been ascribed to the closure of these using the time for the temperature to reach half of the
closely spaced interfaces.2,13 recorded maximum (t1∕2):
The 3D nature of the heat transfer pathways in APS
TBCs requires not only volumetrically averaged 1:38d2
aðTÞ ¼ (1)
microstructural information, but also knowledge of the spa- p2 t1=2 ðTÞ
tial distribution of the pores. Recent progress in X-ray
computed tomography (CT) has made possible the detailed where d is the thickness of the sample. However, Equa-
3D quantification of porous microstructures.16 Indeed, the tion 1 only holds for an infinitesimal pulse width. This is
subsurface microstructure of APS TBCs has recently been usually corrected for after measurement using the Cape-
studied with micrometer resolution using laboratory-based Lehman model.21 Consequently, the thermal conductivity
X-ray microscopes.17,18 of the sample can be estimated through the relation:
The aim of this paper is first to investigate whether 3D KðTÞ ¼ aðTÞ  qðTÞ  Cp ðTÞ (2)
X-ray l-CT images are sufficient to predict the thermal
conductivity performance of APS TBCs using image-based where q (T) and Cp (T) are the density and specific heat of
numerical models. The second step, once validated, is to the sample at T, respectively.
ZHANG ET AL. | 3

In this study, laser flash analysis was conducted on tomographic scan comprised 3201 projections equally
free-standing YSZ coating samples using a Netzsch LFA spaced over a rotation angle of 360°. A 109 optical lens
427 Laser Flash Apparatus. A ring-shaped trench with a was applied after the scintillator for magnification. A voxel
diameter of ~14 mm was made on the coating with a dia- size of 0.57 lm on the reconstructed virtual slices was
mond coated core hole saw. The free-standing disk YSZ achieved by the combination of optical lens and geometric
coating was obtained by immersing the button in magnification. Tomographic reconstruction was based on
hydrochloric acid to remove the substrate and bond coat. In the FDK (Feldkamp-Davis-Kress) algorithm23 using the
this way, the thermal conductivity in the through-thickness XMReconstructor software provided by Zeissâ.
direction (parallel to the spray direction) was measured. Image processing including filtering, segmentation, and
The effective thermal conductivity was first measured quantification was carried out using Avizo 9.0 software.
for the coating in the as-deposited state. Afterwards, the A nonlocal means filter was first applied to the recon-
sample was sintered at 1100°C under an Ar atmosphere structed virtual slices to reduce noise.24,25 A local thresh-
using the laser flash apparatus for 10 hours. The effective olding method (top hat) was used to segment the pores,
thermal conductivity of the sintered sample was measured which selects valleys by a series of morphological opera-
after the sample returned to ambient temperature. Prior to tions on the gray-scale image. Details of this method can
measurement, the sample surface was coated with a thin be found in.26
layer of graphite to improve the laser absorption. Due to Quantitative analysis of the microstructure was achieved
the relatively small sample thickness (~450 lm), the heat on the segmented 3D images. It was observed that the
diffusion time can be comparable to the pulse time. There- pores were essentially interconnected. To quantify the mor-
fore, the smallest pulse width (0.3 mseconds in this case) phology of the pores, each individual pore needed to be
has been used together with pulse correction based on the distinguished. This has been achieved using the “Separate
improved Cape-Lehman model.22 Five measurements have Object” module in Avizo.25 This method is based on the
been performed on 1 sample for consistency. The density watershed of the distance map of the binary image.27 As a
and specific heat of the sample were determined using the result, the connected pores were separated at their narrow-
Archimedes’ method and differential scanning calorimetry est points and the volume, shape, and orientation were
(DSC), respectively. quantified pore by pore.28 The pores could then be catego-
rized based on their 3D volumes as determined from image
analysis.
2.3 | Microstructural characterization
Scanning electron microscopy (SEM) was employed to
examine the pores inside the sample. The sample was first
2.4 | Image-based models
mounted in resin at room temperature. A cross section was To quantify the effect of pores on the effective thermal
made through the coating and substrate using a SiC abra- conductivity of the coating, image-based modeling was
sive cutting blade. This was followed by progressive grind- conducted utilizing the segmented 3D images of the YSZ
ing using SiC grinding paper from 400 to 1200 grit at 200 topcoat. The Avizo XLab Thermo Pack25 was used to sim-
interval. Finally, the cross section was polished to a ulate the heat flux through the porous TBC system.
0.25 lm finish using successively finer diamonds (6, 1 and Transient heat conduction in the YSZ solid in the TBC
0.25 lm). can be described using the partial differential equation:
X-ray l-CT was used to image the pore structure in 3D
@T
inside the coatings. The sample for l-CT was cut from the qCp  KYSZ r2 T ¼ 0; (3)
same button from which the free-standing YSZ disk had @t
been removed. A thin substrate-bondcoat-TBC slice where q is the density, Cp is the specific heat, KYSZ is
(600~800 lm) parallel to the thickness direction of the but- the thermal conductivity and T is the temperature of the
ton was first cut out. To achieve sufficient X-ray transmis- bulk YSZ. The thermal conductivity of the fully dense
sion, the substrate was thinned using SiC grinding paper to YSZ was taken as KYSZ = 2.5 W∕(mK).2 As the thermal
a thickness similar to that of the coating (~500 lm). As a conductivity of air (~0.026 W∕(mK) at room tempera-
result, a rectangular prism sample was made whose base ture29) is about 2 orders lower than that of YSZ, pores
had a diagonal length of ~1 mm. This sample was scanned have been assumed to be heat insulators in the simula-
using a Zeissâ Versa 520 X-ray microscope in the Henry tions. A steady state was assumed so that the transient
Moseley X-ray Imaging Facility. The X-ray tube was oper- heat conduction problem (Equation 3) reduced to solving
ated at 140 kV producing polychromatic X-rays. A lead- a time-invariant temperature field under the constraint
glass X-ray filter was used to harden the X-ray beam and of constant boundary temperatures as expressed in
achieve ~30% transmission for each projection. Each Equation (4).
4
| ZHANG ET AL.

Q Tout  Tin
¼ Ka (4)
S t
where Q is the total heat flux through the input face, S is
the area of the input surface, Tout and Tin are the imposed
temperatures at the outer and inner surface, respectively, t
is the thickness of the region through which heat flows,
and Ka is the apparent thermal conductivity of the porous
topcoat in the coating’s thickness direction.
A finite volume method has been used with each voxel
serving as an isotropic cubic volume element.25 Once dis-
cretized, the steady-state temperature field was found by
solving the Fourier equation systems using a fully implicit
method. A conjugate gradient iteration with an ILU precon-
ditioner was performed until the relative decrease in norm F I G U R E 1 Thermal conductivity as a function of the (square
of the residual was less than a preset criterion (105 in this based) ROI size determined at 3 randomly picked positions. All ROI
span the coating thickness
study).25
Simulations were conducted using the 3D micrographs
of the real coating sample (Figure 4). As has been pointed a virtual l-CT cross section. The 3 distinct layers within
out in,2 the model domain needs to be large enough, and the APS TBC system are clearly distinguished. Pores are
contain sufficient microstructural details, to represent the observed in the topcoat and bond coat. Both globular and
average features of the coating. On the other hand, a large interlamellar cracks within the topcoat and bond coat are
model domain means a much higher computational cost. observed on the virtual l-CT cross section. In the bond
Considering the voxel size of the reconstructed virtual slice coat, interlamellar cracks predominate, whereas in the top-
(0.57 lm), a region of interest (ROI) with a size of coat the globular pores are more pronounced. The SEM
100 9 100 9 100 lm3 contains around 5 400 000 volume image suggests that in addition to the wide interlamellar
elements. Therefore, a domain size check was first con- cracks (yellow rectangle in Figure 2B), fine (submicrome-
ducted by incrementally changing the size of the region of ter opening) interlamellar cracks are present (eg, the insert
interest (ROI) for the simulation until the calculated ther- in Figure 2B). Their fine scale means that are not resolved
mal conductivity converged. Further increase in the domain in the l-CT cross section. Moreover, although some of the
had little benefits on the robustness of the calculated ther- intersplat cracks can be detected in the virtual l-CT cross
mal conductivity, but would significantly increase the com- section, they have not been clearly resolved. Instead, they
putational cost. appear as blurred and grayish phases in the slice due to the
Figure 1 shows the predicted thermal conductivity val- limited resolution and partial volume effect.30
ues as a function of the size of the ROI picked randomly
at 3 positions within the sample. All the ROIs span the full
3.2 | Quantification of defects in the YSZ
coating thickness. Figure 1 suggests that a ROI size of at
topcoat
least 200 lm is necessary to produce a representative volu-
metrically averaged thermal conductivity value. The fluctu- To quantify the different types of defects (ie, globular
ations observed at a small ROI size are mainly due to the pores and interlamellar cracks), they must first be seg-
uneven distribution of the microstructural constituents, for mented from the CT volume images as shown in Figure 3.
example, pore clusters or local pore-free zones larger than The top-hat method gives a more or less satisfactory
the ROI. Therefore, all simulations have been run with a segmentation result (see Figure 3C) for the pores in the
ROI size of around 200 9 200 9 450 lm3. YSZ topcoat. Comparing Figure 3B,C, it can be seen that
the large and globular pores in the YSZ are precisely
labeled. However, when it comes to the fine and planar
3 | RESULTS interlamellar cracks (yellow arrow in Figure 3A,B), seg-
mentation proves difficult. In some situations, the cracks
3.1 | Microstructural features in the could not be segmented from the YSZ as their widths are
as-deposited sample much smaller than the voxel size of the CT image (eg,
To validate the l-CT results, the polished cross section of insert in Figure 2B). There are other situations where over-
the as-deposited coating was imaged using scanning elec- segmentation occurred when the cracks were detected and
tron microscope (SEM) and shown in Figure 2B alongside segmented, whereas their opening has been less than the
ZHANG ET AL. | 5

F I G U R E 2 Comparison of (A) a
virtual l-CT cross section and (B) a
destructive cross section imaged by SEM
(BSE) in the as-deposited condition

F I G U R E 3 A typical virtual cross section (A) before; (B) after nonlocal means filtering and (C) after segmentation. The filtering has helped
remove noise in the image and differentiate the layers but has also smeared out small features

size of 1 voxel.31 In spite of these difficulties, the porosity it can be observed that the larger pores (ie, >10 000 lm3),
of the YSZ topcoat has been determined from the 3D which constitute the majority of the porosity detected
images as 16.1  0.7%. By comparison, image analysis on (according to Figure 4A), have complex surface morpholo-
the cross section SEM micrographs yields a value of gies. To quantitatively describe the shape of the pores, their
17.6  1.3%. The discrepancy is primarily due to the fine sphericity (Φ) has been determined from the 3D images.
interlamella cracks not being resolved by X-ray l-CT. The sphericity of a 3D particle is defined as:
Moreover, sample preparation, that is grinding and polish-
p1=3 ð6VÞ2=3
ing may have introduced damage to the cross sections for U¼ (5)
SEM investigation, leading to a higher porosity level. A
In addition to obtaining the total porosity, the pores cap- where V and A are the volume and surface area of the par-
tured by X-ray l-CT have been categorized based on their ticle, respectively, where a sphere gives Φ = 1.
3D volumes. Figure 4 shows 3D images of the pores on Figure 5 plots the sphericity and the alignment of the
each volume range and the contribution of each type pores as a function of their size. Figure 5A confirms the
of pores to the total porosity by the cumulative porosity observation in Figure 4 that the pores become less spherical
profile. as their sizes increase. The sphericity of the pores is seen to
Figure 4B-G show a transition with increasing size in decrease linearly with the logarithm of their volume. Like-
the morphology of the pores from randomly distributed wise the pores become more aligned with the plane of the
spheres to more torturous and oriented clusters. In addition, coating (plane of the splats) as their size increases. The
6
| ZHANG ET AL.

F I G U R E 4 3D images of typical pores


categorized based on their 3D volume and
contribution of each type of pores to the
total porosity: (A) cumulative porosity
determined from segmented 3D X-ray
image and 3D view of pores in the size
range (B) 0-500 lm3; (C) 500-5000 lm3;
(D) 5000-20 000 lm3; (E) 20 000-
100 000 lm3; (F) 100 000-200 000 lm3,
and (G) >200 000 lm3

dependence of both the pores’ shape and orientation on their splat interfaces and fine intersplat cracks are prone to sin-
volume has been a result of the splat-based microstructure. tering considering the service temperature of the TBCs.
As noted previously,32 the larger pores tend to be those aris- This may explain why it has been reported that the thermal
ing from air trapped in between the splats: as a result they conductivity of an as-deposited TBC increases significantly
tend to follow the shape and orientation of the splats. (by as much as ~46%) and rapidly (after 10 hours) upon
thermal exposure (1100°C) due to the closure of closely
spaced splat interfaces despite little change in the total vol-
3.3 | Thermal conductivity ume fraction of porosity.13
The measured thermal conductivity has been compared To investigate the effect of sintering on the thermal con-
with the model predictions and the results are summarized ductivity, we have sintered the free-standing TBC sample
in Figure 6. at 1100°C for 10 hours. After sintering the thermal conduc-
The effective through-thickness thermal conductivity tivity of the sintered sample has increased (from
measured on the as-deposited free-standing coating is 1.20  0.12 W∕(mK) to 1.55  0.18 W∕(mK)), which
1.20  0.12 W∕(mK; 48  5% of the bulk value). By is close to the predictions from the l-CT image-based sim-
comparison, the image-based model predicts a conductivity ulations of the as-deposited sample.
of 1.76  0.05 W∕(mK; 30  1% reduction from the
bulk value). This suggests that the pores detected by X-ray
l-CT account for only ~60% of the total reduction in ther-
4 | DISCUSSION
mal conductivity. The splat boundaries and fine inter-
4.1 | Investigation into the role of the fine
lamella cracks (on the scale of submicrometer), which have
cracks
not been resolved in the l-CT slices are probably the main
reason for the difference between the measured and pre- It is reported above that the 3D X-ray l-CT image-based
dicted thermal conductivity2 and this is investigated below. model overestimates (by almost 50%) the thermal conduc-
It should be noted that, effective as they are in reducing tivity of the as-deposited TBC sample and it is postulated
the thermal conductivity of the as-deposited coating, the that this is due to very fine cracks. The SEM image in
ZHANG ET AL. | 7

wide range of scales (from ~100 lm down to nanome-


ters).33-35 Such fine cracks cannot be clearly resolved in
the l-CT virtual slices even though there is some contrast
suggesting the presence of such features. Under the current
scanning setup, a voxel size of ~0.57 lm has been
employed, which alongside image filtering to remove noise
in the images and facilitate segmentation, leads to the loss
of detailed features (eg, interfaces in Figure 3A). As a
result, only pores and cracks with a volume larger than 15
voxels were considered as defects (pores and cracks). This
corresponds to a volume of around 2.8 lm3.
To establish whether the fine defects and interfaces
could explain the ~50% overestimation in the thermal con-
ductivity of the as-deposited TBC sample, we have
employed the contact-zone model as defined in36 to quan-
tify the contribution to the thermal conductivity reduction
from these features. The idea being to treat the intersplat
gaps as a semi-insulating medium with a degraded thermal
conductivity value. In this case, high-resolution SEM
micrographs (pixel size  0.1 lm) were used as input for
the image-based models. Features smaller than the X-ray
l-CT resolution limit, that is, intersplat gaps thinner than
1 lm and pores smaller than ~1.7 lm (this corresponds to
a sphere smaller than 2.8 lm3 in 3D) were taken as contact
zones to which a degraded thermal conductivity has been
assigned (Figure 7).
As explained in,36 it is not possible to determine the
effective thermal conductivity of the contact zones experi-
mentally. Instead, a series of reduced thermal conductivity
F I G U R E 5 Shape and orientation of the pores as a function of
values have been used for the contact zones. The calculated
their volume determined from the 3D images: (A) sphericity and (B)
alignment of the pore’s long axis with the normal to the coating. The
effective thermal conductivity is plotted against that of the
horizontal error bar represents the scatter in pore size and the vertical contact zone in Figure 8. When the contact zones are fully
one represents the scatter in sphericity and orientation of the pores insulating (KCZ/KBulk < 104), the effective thermal con-
with that range of size ductivity of the sample is only 0.29 W∕(mK), which is
<25% of the measured value. In fact, the simulated thermal
conductivity matches the experiment measurements when
Figure 2B confirms the existence of cracks both on the KCZ is about half of KBulk. This could be because specimen
micrometer and submicrometer scales. Other researchers preparation introduces some of the fine cracks seen in the
have also reported cracks/pores in APS TBCs across a micrograph, or that the intersplat gaps are only 50%

F I G U R E 6 Comparison of the
measured reduction in effective through-
thickness thermal conductivity (expressed
as the % of the bulk conductivity) with
model predictions. For the model
predictions on the as-deposited sample, the
contribution of each type of pores has been
isolated and highlighted using the sample
color code shown in Figure 4
8
| ZHANG ET AL.

F I G U R E 7 (A) High magnification


SEM micrograph of the coating and (B) the
same image showing the segmented pores
(blue) and contact zones (red)

effective in providing insulation in the real TBCs possibly pore (>2.8 lm3) inside the sample before and after sinter-
because they are not fully open. ing. Both the 2D slice and 3D image suggest that no
detectable changes can be identified from the X-ray images
of the sample after the sintering treatment. Based on this,
4.2 | Investigation of the effect of sintering
the quantification on the cumulative porosity, pore mor-
on microstructure
phology and orientation (as shown in Figures 4 and 5) will
It was postulated above that the increase in the measured remain unchanged after sintering. Indeed, the porosity
thermal conductivity after sintering is due to the closure of determined from the X-ray images of the as-sintered sam-
closely spaced splat interfaces. To investigate this further ple is 16.0  0.5%.
the same ROI on the sample was reimaged by X-ray l-CT The cross-sectional SEM images (Figure 9E,F) also sug-
after sintering at 1100°C for 10 hours in air. A pair of cor- gest that the large pores (>10 lm in diameter) have chan-
responding virtual cross sections are shown in Figure 9. ged little during sintering. However, the thin intersplat
As evident in Figure 9, upon sintering a significant ther- cracks (inserts in Figure 9E,F) seem to have narrowed after
mally grown oxide (TGO) has developed at the topcoat/ sintering due to diffusion and grain growth across the inter-
bond coat interface. In addition cracks have initiated at splat boundaries at sintering temperature (1100°C in this
some locations within the topcoat close to the topcoat/bond study).37
coat interface. However, these changes have little effect on In view of the fact that the thermal conductivity pre-
the thermal conductivity measurements as these were made dicted by the X-ray CT image-based model
on the free-standing as-sintered coating. Other than these, (1.74  0.03 W∕(mK) for the as-sintered sample) now
no detectable changes could be observed on the virtual comes into close agreement with the measured as-sintered
cross sections of the YSZ topcoat. The same image pro- conductivity (model prediction is about 12% higher than
cessing routine was used on the X-ray images of the as-sin- the measured value), this provides conclusive evidence that
tered sample. Figure 9C,D show an example of 1 large it is the fine scale cracks that are responsible for the
increase in thermal conductivity during the early stages of
thermal exposure. Further it validates that X-ray CT imag-
ing is able to capture the larger scale pores and cracks that
are responsible for the long-term thermal conductivity of
the coating and that image-based modeling can reliably
predict long-term thermal conductivity.

4.3 | The contribution to the thermal


conductivity of different types of pores
Given that the image-based model is in close accordance
with the as-sintered (ie, likely in-service) thermal conduc-
tivity of the coating, it can be used to assess the contribu-
tion to the conductivity of the different types of pores
found in the coating. The contribution to the total porosity
F I G U R E 8 Simulated TBC thermal conductivity as a function of
the contact zone conductivity, KCZ, normalized by that of bulk YSZ
and reduction in thermal conductivity for each type of pore
(see Figure 4) has been plotted in Figure 10. It is evident
ZHANG ET AL. | 9

F I G U R E 9 The same X-ray l-CT


virtual cross section viewed (A) as-
deposited and (B) after sintering at 1100°C
for 10; 1 large pore inside the YSZ topcoat
viewed in 3D (C) as-deposited and (D)
after sintering along with SEM images (E)
as deposited and (F) after sintering

that pores having volumes in the range


20 000~100 000 lm3 (size 4 in Figure 10) constitute most
of the pore space and contribute most to the reduction in
conductivity over bulk YSZ. Although the relatively small
pores (<20 000 lm3, ie, size 1, 2 and 3 in Figure 10) are
large in number, as visualized in Figure 4, they account for
<25% of the total porosity and 5.3% of the reduction in
thermal conductivity. It is noteworthy that although the
total volume of pores lager than 200 000 lm3 (size 6 in
Figure 10) is less than that of pores on the range of
5000~20 000 lm3 (size 3 in Figure 10), they contribute
more to the reduction in thermal conductivity. F I G U R E 1 0 A comparison of the predicted contribution of each
To compare the effectiveness of each type or pore in individual pore type to the total porosity and thermal conductivity
reducing the conductivity, the volume normalized conduc- reduction predicted by image-based simulations. The color code for
tivity reduction has been plotted in Figure 11. Figure 11 the columns are the same as in Figures 4 and 6
suggests that the effectiveness of pores in reducing the
10
| ZHANG ET AL.

2. It was shown that fine cracks, mostly intersplat cracks


identified in SEM micrographs, are more than sufficient
to explain the remaining reduction in conductivity. Indeed
a good match between experiment and model is reached
when the conductivity of the interspat gap is about half of
the bulk value for the as-deposited sample.
3. The narrowing of overlapping splats via diffusion and
grain growth across intersplat boundaries during sintering
treatment causes the measured thermal conductivity of the
coating to rise from 1.20  0.12 W∕(mK) to 1.55 
0.18 W∕(mK). At the same time the coarser defects (ie,
defects larger than 2.8 lm3) are found to be stable during
this sintering treatment. As a result, the measured conduc-
tivity of the as-sintered coating comes to within 15% of
that predicted on the basis of the coarser defects. This
F I G U R E 1 1 Volume averaged reduction in thermal conductivity
implies that the long-term thermal conductivity of an APS
for pores with different volume
TBC topcoat can be predicted by X-ray l-CT image-
based models developed from an as-deposited sample.
conductivity is linearly related to their volume. The reduc- 4. The shapes of the pores within the YSZ topcoat are
tion per unit volume of a 250 000 lm3 pore is more than strongly related to their size. Small pores are roughly
twice as effective as a 1500 lm3 pore. A numerical study spherical, whereas large pores are more tortuous and
looking at different sizes and shapes of pores confirms that much flatter and tend to lie within the plane of the
this is probably because the larger pores are flatter and tend splats. As a result the larger pores are more effective in
to lie within the plane of the coating (perpendicular to the reducing the thermal conductivity.
heat flux).
The fact that the X-ray l-CT can detect the pores that
are stable under realistic in-service temperatures suggests
5 | CONCLUSIONS that 3D X-ray imaging is a good way of helping to opti-
mize APS process control in terms of the optimization of
In this study, the pore microstructure-conductivity relation- retained porosity for improved thermal conductivity.
ship has been explored for a TBC deposited by APS.
Detailed microstructural characterization has been carried
ACKNOWLEDGMENTS
out by l-CT and scanning electron microscopy in the as-
deposited condition, as well as after sintering at tempera- The authors acknowledge the provision of beam time at
tures representative of in-service conditions. Considering Henry Moseley Manchester X-ray Imaging Facility
that thermal conduction through the YSZ solid is still the (HMXIF) at the University of Manchester, UK, established
dominant heat transfer mechanism even at the service tem- with funding from EPSRC through grants EP/F007906, EP/
perature of the TBCs, 3D image-based microstructurally I02249X and EP/F028431 and HEFCE funding of the Mul-
realistic models have been developed to explore the effect tiscale Characterisation Facility. PJW is grateful for a Euro-
of the porosity on the reduction in conductivity of the coat- pean Research Council for funding COREL-CT under grant
ing. The following concluding remarks can be drawn: No 695638. The authors acknowledge Prof. Robert Vaßen
and Dr. Denise Koch for providing the APS TBC samples.
1. The effective thermal conductivity of the as-deposited
TBC topcoat can be as low as 50% of a fully dense
ORCID
YSZ due to the existence of pores and splat interfaces.
Among these, X-ray l-CT was used to image the pores Ping Xiao http://orcid.org/0000-0002-6063-3681
larger than 2.8 lm3 in volume. Comparison of the
X-ray virtual slices taken before and after sintering
suggests that the X-ray images are robust in resolving REFERENCES
such large defects in the APS TBC. X-ray l-CT image-
1. Padture NP, Gell M, Jordan EH. Materials science - Thermal bar-
based modeling suggested these contribute ~60% of the rier coatings for gas-turbine engine applications. Science.
total reduction in thermal conductivity. 2002;296:280-284.
ZHANG ET AL. | 11

2. Wang Z, Kulkarni A, Deshpande S, Nakamura T, Herman H. 21. Cape JA, Lehman GW. Temperature and finite pulse-time effects
Effects of pores and interfaces on effective properties of plasma in flash method for measuring thermal diffusivity. J Appl Phys.
sprayed zirconia coatings. Acta Mater. 2003;51:5319-5334. 1963;34:1909-1913.
3. Sampath S, Schulz U, Jarligo MO, Kuroda S. Processing science 22. Blumm J, Opfermann J. Improvement of the mathematical model-
of advanced thermal-barrier systems. MRS Bull. 2012;37:903-910. ing of flash measurements. High Temp High Press. 2002;34:515-
4. Kulkarni A, Wang Z, Nakamura T, et al. Comprehensive microstruc- 521.
tural characterization and predictive property modeling of plasma- 23. Feldkamp LA, Davis LC, Kress JW. Practical cone-beam algo-
sprayed zirconia coatings. Acta Mater. 2003;51:2457-2475. rithm. J Opt Soc Am A Opt Image Sci Vis. 1984;1:612-619.
5. Mcpherson R. A review of microstructure and properties of 24. Liu YL, Wang J, Chen X, Guo YW, Peng QS. A robust and fast
plasma sprayed ceramic coatings. Surf Coat Technol. non-local means algorithm for image denoising. J Comput Sci
1989;39:173-181. Technol. 2008;23:270-279.
6. Kulkarni A, Vaidya A, Goland A, Sampath S, Herman H. Pro- 25. Avizo User Guide, Visualization sciences group SAS. 2013.
cessing effects on porosity-property correlations in plasma 26. Serra J. Introduction to mathematical morphology. Comput Vis
sprayed yttria-stabilized zirconia coatings. Mater Sci Eng. Graph Image Process. 1986;35:283-305.
2003;359:100-111. 27. Kimmel R, Kiryati N, Bruckstein AM. Sub-pixel distance maps
7. Golosnoy IO, Cipitria A, Clyne TW. Heat transfer through and weighted distance transforms. J Math Imaging Vis.
plasma-sprayed thermal barrier coatings in gas turbines: a review 1996;6:223-233.
of recent work. J Therm Spray Technol. 2009;18:809-821. 28. Brabant L, Vlassenbroeck J, De Witte Y, et al. Three-dimensional
8. Cernuschi F, Bison P, Moscatelli A. Microstructural characteriza- analysis of high-resolution X-ray computed tomography data with
tion of porous thermal barrier coatings by laser flash technique. Morpho+. Microsc Microanal. 2011;17:252-263.
Acta Mater. 2009;57:3460-3471. 29. Lima JAP, Marin E, da Silva MG, et al. Characterization of the
9. Cernuschi F, Ahmaniemi S, Vuoristo P, Mantyla T. Modelling of thermal properties of gases using a thermal wave interferometer.
thermal conductivity of porous materials: application to thick Meas Sci Technol. 2001;12:1949-1955.
thermal barrier coatings. J Eur Ceram Soc. 2004;24:2657-2667. 30. Erlandsson K, Buvat I, Pretorius PH, Thomas BA, Hutton BF. A
10. Nait-Ali B, Haberko K, Vesteghem H, Absi J, Smith DS. Ther- review of partial volume correction techniques for emission
mal conductivity of highly porous zirconia. J Eur Ceram Soc. tomography and their applications in neurology, cardiology and
2006;26:3567-3574. oncology. Phys Med Biol. 2012;57:R119-R159.
11. Ahmaniemi S, Vuoristo P, Mantyla T, Cernuschi F, Lorenzoni L. 31. Withers PJ, Preuss M. Fatigue and damage in structural materials
Modified thick thermal barrier coatings: thermophysical character- studied by X-ray tomography. Annu Rev Mater Res. 2012;42:81-
ization. J Eur Ceram Soc. 2004;24:2669-2679. 103.
12. Zhu DM, Miller RA. Thermal conductivity and elastic modulus 32. Mcpherson R. The relationship between the mechanism of forma-
evolution of thermal barrier coatings under high heat flux condi- tion, microstructure and properties of plasma-sprayed coatings.
tions. J Therm Spray Technol. 2000;9:175-180. Thin Solid Films. 1981;83:297-310.
13. Chi WG, Sampath S, Wang H. Microstructure-thermal conductiv- 33. Funke C, Mailand JC, Siebert B, Vassen R, Stover D. Characteri-
ity relationships for plasma-sprayed yttria-stabilized zirconia coat- zation of ZrO2-7 wt% Y2O3 thermal barrier coatings with differ-
ings. J Am Ceram Soc. 2008;91:2636-2645. ent porosities and FEM analysis of stress redistribution during
14. Deshpande S, Kulkarni A, Sampath S, Herman H. Application of thermal cycling of TBCs. Surf Coat Technol. 1997;94:106-111.
image analysis for characterization of porosity in thermal spray 34. Vassen R, Cao XQ, Tietz F, Basu D, Stover D. Zirconates as
coatings and correlation with small angle neutron scattering. Surf new materials for thermal barrier coatings. J Am Ceram Soc.
Coat Technol. 2004;187:6-16. 2000;83:2023-2028.
15. Arai M, Ochiai H, Suidzu T. A novel low-thermal-conductivity 35. Guo HB, Vassen R, Stover D. Atmospheric plasma sprayed thick
plasma-sprayed thermal barrier coating controlled by large pores. thermal barrier coatings with high segmentation crack density.
Surf Coat Technol. 2016;285:120-127. Surf Coat Technol. 2004;186:353-363.
16. Maire E, Withers PJ. Quantitative X-ray tomography. Int Mater 36. Zivelonghi A. Thermomechanical behaviour of two heterogeneous
Rev. 2014;59:1-43. tungsten materials via 2D and 3D image-based FEM. 2011.
17. Zhao Y, Shinmi A, Zhao X, et al. Investigation of interfacial 37. Thompson JA, Clyne TW. The effect of heat treatment on the
properties of atmospheric plasma sprayed thermal barrier coatings stiffness of zirconia top coats in plasma-sprayed TBCs. Acta
with four-point bending and computed tomography technique. Mater. 2001;49:1565-1575.
Surf Coat Technol. 2012;206:4922-4929.
18. Ahmadian S, Browning A, Jordan EH. Three-dimensional X-ray
micro-computed tomography of cracks in a furnace cycled air
plasma sprayed thermal barrier coating. Scripta Mater.
How to cite this article: Zhang X, Kulczyk-
2015;97:13-16. Malecka J, Carr J, Xiao P, Withers PJ. 3D
19. Mauer G, Du LN, Vassen R. Atmospheric plasma spraying of characterization of porosity in an air plasma-sprayed
single phase lanthanum zirconate thermal barrier coatings with thermal barrier coating and its effect on thermal
optimized porosity. Coatings. 2016;6:49. conductivity. J Am Ceram Soc. 2018;00:1–11.
20. Parker WJ, Jenkins RJ, Abbott GL, Butler CP. Flash method of https://doi.org/10.1111/jace.15409
determining thermal diffusivity, heat capacity, and thermal con-
ductivity. J Appl Phys. 1961;32:1679-1684.

You might also like