You are on page 1of 54

INL/JOU-21-61712-Revision-0

Thermal Barrier Coatings


Overview: Design,
Manufacturing and
Applications in High
Temperature Industries
April 2021

Kunal Mondal, Luis Nunez, Calvin Myer Downey, Isabella J Van Rooyen

INL is a U.S. Department of Energy National Laboratory operated by Battelle Energy Alliance, LLC
DISCLAIMER
This information was prepared as an account of work sponsored by an
agency of the U.S. Government. Neither the U.S. Government nor any
agency thereof, nor any of their employees, makes any warranty, expressed
or implied, or assumes any legal liability or responsibility for the accuracy,
completeness, or usefulness, of any information, apparatus, product, or
process disclosed, or represents that its use would not infringe privately
owned rights. References herein to any specific commercial product,
process, or service by trade name, trade mark, manufacturer, or otherwise,
does not necessarily constitute or imply its endorsement, recommendation,
or favoring by the U.S. Government or any agency thereof. The views and
opinions of authors expressed herein do not necessarily state or reflect
those of the U.S. Government or any agency thereof.
INL/JOU-21-61712-Revision-0

Thermal Barrier Coatings Overview: Design,


Manufacturing and Applications in High Temperature
Industries

Kunal Mondal, Luis Nunez, Calvin Myer Downey, Isabella J Van Rooyen

April 2021

Idaho National Laboratory


Idaho Falls, Idaho 83415

http://www.inl.gov

Prepared for the


U.S. Department of Energy
Under DOE Idaho Operations Office
Contract DE-AC07-05ID14517
Thermal Barrier Coatings Overview: Design, Manufacturing and
Applications in High-Temperature Industries

Kunal Mondal1, *, Luis Nunez2, Calvin M. Downey3, and Isabella J. van Rooyen4
1
Materials Science and Engineering Department, Idaho National Laboratory, Idaho Falls, ID,
83415
2
Irradiated Fuels and Materials Department, Idaho National Laboratory, Idaho Falls, ID, 83415
3
Experiment Design Department, Idaho National Laboratory, Idaho Falls, ID, 83415
4
Reactor System Design and Analysis Division, Idaho National Laboratory, Idaho Falls, ID,
83415

Abstract

Today’s competitive world economy is creating an indispensable demand for increased efficiency

of engineering components that operate in harsh environments (i.e., very high-temperature,

corrosive, or neutron irradiation environments), for applications in the energy, automotive,

aerospace, electronics, and power industries. Increased research is being done on thermal barrier

coatings (TBCs) for protecting such components, since the versatility of manufacturing techniques

and the scale of deployment result in increased life, economics, performance, and durability. This

review focuses on the advances that led to using TBCs for component life extension and, more

recently, as an integral part of advanced component design for high-temperature and other types

of harsh environments, such as those found in nuclear-related applications. Factors that led to state-

of-the-art advanced coating-fabrication techniques (e.g., electron-beam physical vapor deposition

[EB-PVD], plasma spray deposition, and electrophoretically deposited TBCs, as well as

functionally graded material [FGM] manufacturing) have also been emphasized in current coating

R&D. This review explores the current state of TBCs, the latest advances regarding their

1
fabrication and performance, associated challenges, and recommendations for their future use in

aerospace, nuclear, high-temperature, or otherwise harsh environments.

Keywords: Additive manufacturing, Advanced manufacturing processes, Thermal barrier

coatings, Gradient coating, High-temperature applications, Nuclear reactor components

* Corresponding author. Tel.: +1-208 5264960. E-mail address: Kunal.Mondal@inl.gov

2
1. Introduction

The global push for reliable green energy has created significant interest in increasing the
1,2
proportion of electricity generated by nuclear power . This has fostered new innovations in

power plant technologies, especially as regards the efficiency, safety, and lifespan of future plants
3
. It is vital that materials used in these systems perform at much higher temperatures and in harsher

corrosive and radiative environments than do materials in current reactor systems 4. Materials and

associated manufacturing techniques that are robust and reliable under such harsh environmental

conditions must be developed 5. Shielding both the reactor pressure vessel and power generation

components from harsh environments and high-temperature exposure is strategic for ensuring

sufficiently long operation lifetimes. As a result of high-quality research over the past few decades,

thermal barrier coating (TBC) systems were developed to safeguard critical components under

even the most demanding of operating conditions 6–8. TBCs are multilayer coatings that protect a

substrate (structure) from both heat and corrosion 9. TBCs not only restrict heat loss from the

environment into the substrate, thus ensuring greater thermal efficiency, but also help sustain lower

temperatures within the substrate, thereby allowing for increased operating temperatures 10–12.

Proven efficient over hundreds of hours of engine operations, TBCs for combustion engines

and power plants are available for use in future nuclear technologies 13. However, despite several

advances, more robust TBCs are still needed, whether to increase TBC service lifetimes under

present-day operating conditions or to operate in even higher temperatures and harsher

environments to attain more efficient energy conversion. As a result, development of new materials

and associated advanced manufacturing (AM) techniques, along with the characterization and

optimization of TBC properties, are now active areas of research.

3
Note that, in addition to assessing TBC performance in terms of service lifetimes and

durability, it is also important to consider their stability, reliability, and manufacturability.

Recently, AM techniques have introduced new methods of fabricating functionally graded

materials (FGMs), which change in composition and/or porosity as a function of the thickness of

the coating or layer. This area of research may lead to alleviating some of the challenges associated

with manufacturing advanced TBCs 14,15.

This review paper offers a view on the current state of TBCs, as well as the latest advances

regarding their fabrication and performance—including special consideration of FGMs.

Difficulties related to the stability and longevity of TBCs, along with their possible solution to

such challenges, are also reviewed, particularly in the context of high-temperature applications.

2. Design of Thermal Barrier Coatings

TBCs can provide insulation against long- and short-term problems pertaining to gas turbine,

internal combustion engine, power generation, and advanced nuclear components. By lowering

the thermal conductivity of the coating material, the component’s available temperature range is

thus increased, in turn increasing the overall efficiency of the system. TBCs can also slow material

degradation related to extreme temperature exposure and retain a thermal gradient between the

coating and the substrate interfaces. The basic classical TBC design features a substrate, a bond

coat, a thermally grown oxide (TGO) layer, and a topcoat (see Figure 1). In TBCs, the bond coat

acts as a pre-coating interface between the substrate and the topcoat to increase both the adhesion

between the layers and the structural integrity of the coating. The TGO layer is created via

diffusion of oxygen through the topcoat during manufacturing and operation. The topcoat,

4
otherwise referred to in the literature as the “TBC layer,” is generally a ceramic that provides

thermal protection to the substrate.

Figure 1: Schematic of a traditional TBC design.

2.1. Common Top Barrier Coat Materials

Effective thermal barrier materials should demonstrate such properties as low thermal

diffusivity, phase stability and thermal shock resistance during thermal cycling, strong adherence

to the substrate, and shielding against oxidation and corrosion for both the metallic bond coat and
16
the substrate . In general, titania, zirconia, alumina, porcelain, porcelainite, pyrochlores (A23+

B24+ O7), garnets (Y3AlxFe5-xO12), monazite (LaPO4), perovskites (ABO3), lanthanum magnesium

hexa-aluminate (LaMgAl1O19), etc. (even diamond), have all been measured to have low thermal

conductivity and can therefore be considered as potential TBC materials. 17.

5
Yttria partially stabilized zirconia (YSZ) is a ceramic material seen in a wide-range of

applications, owing to its exceptional overall properties (e.g., relatively low thermal conductivity,

high dielectric constant, excessive fracture toughness, and chemical inertness at high temperatures)
18
. Notably, YSZ has greater resistance to thermal shock than do other ceramic topcoats. The yttria

stabilizer in the YSZ coating helps preserve the tetragonal phase of zirconia at room temperature,

which suffers a monoclinic phase change under applied external stress 19. This phase change causes

a volumetric expansion and, thus, a buildup of compressive stresses in the vicinity of cracks,

thereby promoting transformation toughening by holding crack propagation 20,21. However, YSZ

barrier coatings are unreliable for long-term use in temperatures of over 1200°C, due to their

sinterability and the catastrophic phase transformation of the metastable tetragonal phase, which

escalate thermal conductivity and boost spallation in the TBCs 22,23.

Other ceramic topcoat layers such as zirconates have been the subject of ongoing academic

investigation. Pyrochlore La2Zr2O7 was found to be a potential alternative to YSZ thanks to its

lower Young’s modulus, comparable fracture toughness, and ~20% lower thermal conductivity at
24
elevated temperatures . The thermal expansion coefficient of the pyrochlore zirconate is

comparable to that of YSZ in the low-temperature range of 100–900°C, and only 10% less in the

high temperature range. Monazite, or LaPO4, is another potential topcoat material with high-

temperature stability, high thermal expansion, and low thermal conductivity 25. The LaTi2Al9O19

(LTA) system was studied in environments requiring material phase stability at over 1300°C 26.

LTA was found to possess excellent phase stability at over 1,600°C and had a thermal expansion

similar to that of YSZ over the same operational temperature range as well as low-temperature

range. LTA showed lower fracture toughness than YSZ; however, a double-layer YSZ/LTA could

balance this out 27. Doped SrZrO3 perovskite was also investigated as a ceramic topcoat material

6
for TBC applications. Both Yb2O3 and Gd2O3 dopants in the perovskite were observed to decrease

the thermal conductivity and Young’s modulus, while also maintaining a fracture toughness

similar to that of YSZ 28. In a study by Jian et al. 29, YSZ was co-doped with CeO2 and Gd2O3 to

address the phase instability problems inherent in conventional YSZ coatings. The results showed

non-transformable phase features, high fracture toughness, low/temperature-independent thermal

conductivity, and high sintering resistance at temperatures up to 1500°C 29. Certain garnets have

also garnered interest, including a Y3Al5O12 garnet that shows promise thanks to its good high-

temperature mechanical properties, low thermal conductivity, good phase stability, and low

oxygen diffusivity compared to traditional zirconia 30.

2.2. Common Bond Coat Materials

Two types of bond coats are commonly used in TBC applications. They are distinguished by

the coat application method used; namely, diffusion or overlay coatings. Diffusion coatings form

a diffused intermetallic layer between the substrate and bond coat to act as an oxygen diffusion

boundary. Aluminum coated onto a superalloy forms a NiAl layer—often with silicon and

chromium inclusions—along with oxide films in this layer 31. Current applications are more likely

to utilize platinum modified NiAl bond coats. Kim et al. 32 demonstrated that thermal cycle type

and service condition are important for TBC lifespans. It was suggested that, for lengthier cycle

times (i.e., >1 h), Pt modified NiAl bond coats function well; but for smaller cycle times of around

10 minutes, common MCrAlX bond coats perform better. MCrAlX bond coats are overlay

coatings, where M represents the base material—commonly mixed with Ni, Co, or Fe—and X

represents Y or Zr with a high reactivity to oxygen to form an oxide layer. However, in this coating

layer, the Al predominately acts as the oxide layer for oxidation protection, and the Cr addition

7
boosts the effective chemical activity of the Al. The X in the coating is generally yttrium, which

helps with thermal resistivity as well as thermal matching with the ceramic coating and grown

aluminum oxide. In the same way, the M base material is meant to increase the compatibility with

the substrate material. Versions of the MCrAlY overlay coating are used in most turbine blade

applications, with high Cr, 5–15% Al, and some Y addition locating on grain boundaries 33. This

overlay coating creates multiphase alloys within a ductile matrix. Overall, Cr, Ni, and Co provide

a bond coat system whose melting point is slightly higher than that of typical diffusion coatings
34
.

2.3. Double-Ceramic-Layer Coatings

Use of double-ceramic-layer (DCL) coatings has been proposed as a novel method to improve

TBC performance by adding primarily zirconate layers into the traditional YSZ topcoat. These

materials are designed to integrate into the ceramic topcoat and provide thermal stability and

improved oxidation resistance. Increasing the phase stability while decreasing the overall thermal

conductivity and thermal expansion mismatch can slow oxidation rates and, thus, potential layer

spallation in the TBC. Many studies have shown zirconate DCL coatings to be a viable option for

improving the overall lifetimes and performance of TBCs in various applications.

Zirconates such as La2Zr2O7 (LZ), Gd2Zr2O7 (GZ), and Nd2Zr2O7 (NZ) have lower

conductivity, higher-temperature phase transformations, and resistance to corrosion, but are

limited by fracture toughness and low thermal expansion coefficients. However, by double

layering the zirconate pyrochlore in a traditional YSZ layer, the thermal expansion mismatches are

minimized 35. In fact, it was found that a DCL topcoat between pyrochlore LZ or NZ with YSZ has

superior thermal properties and relieves thermal stresses between the pyrochlore and the bond coat

8
layer 35. Wang et al. 36
used finite element method analysis to explore the magnitude of thermal

residual stresses in YSZ/LZ TBCs. The finite element method simulation predicted that the

residual stresses from thermal expansion and shock would be lower for a double-layer YSZ/LZ

than for a single-layer TBC. A study on YSZ/LZ DCL coatings found that double-layered YSZ/LZ

TBCs showed better performance than graded YSZ/NZ and typical YSZ TBCs during thermal

cycling tests. Results showed an increase in the number of cycles: 1,380 for the typical YSZ, 3,390

for the graded YSZ/LZ, and 4,140 for the double-layer YSZ/LZ 37. A similar study on YSZ/GZ

showed better oxide layer growth results and preferred oxidation products in comparison with

YSZ, and the GZ was noted to provide a better oxygen diffusion boundary, due to its crystal

structure 38. Limiting the topcoat layer thickness remains important in YSZ-based DCL coatings,

as it was found that increased YSZ thickness leads to reduced durability and higher thermal

conductivity 39.

2.4. Porosity

Porosity is prevalent in many TBC applications to lower thermal conductivity and improve

thermal insulation. Since thermal diffusivity and specific heat capacity are both density-dependent,
40
porosity is a key factor in controlling these thermal properties . Depending on the fabrication

technique and parameters, porosity can vary in both size and location. High porosity values of up

to 25% have been achieved in zirconia TBCs. In an anodized aluminum study of TBCs, two kinds

of pores were typically revealed: nanopores (10–30 nm) and micropores (1~10 µm). These made

up about 80% of the overall porosity, forming on the crystalized silicon in the aluminum alloy

during the anodizing process. As the porosity of a coating increases, the Young’s modulus

decreases; therefore, controlled porosity can also play a crucial role in improving the bond between

9
the substrate, and coating allowing for strain-tolerant interfaces between the bond coat and

overcoming the coefficients of thermal expansion mismatch commonly observed, though porosity

above a certain value can degrade mechanical properties as a result of decreased cohesion 41.

Pore/crack orientation and size play a role in the thermo-mechanical properties obtained

through coatings. Horizontal cracks and pores can improve strain tolerance and help lower thermal

conductivity. Electron-beam physical vapor deposition (EB-PVD) produced YSZ topcoats with

columnar microstructures, while atmospheric plasma spraying (APS) produced coatings with

microcracks and pores horizontal to the surface. Both proved more effective at reducing overall

thermal conductivity: 1.5–1.9 and 0.8–1.1 Wm-1K-1, respectively [10]. Studies of other porous

architecture methods using suspension plasma sprayed (SPS) coatings were also conducted, and it

was observed through experiments that control of the process parameters, and thus the pore

architecture, via low-density coatings directly correlated to a reduced thermal conductivity of 0.5–

0.9 Wm-1K-1, much lower than that achieved via the APS and EB-PVD methods. Total porosity

values of 28.2–43.9% were achieved, with a larger fraction of pores ≥ 1 µm in the samples

containing high total porosity 42. It was observed that 90% of the nanopores seemed smaller than

the 360-nm powder used in the study; thus, pore-size control was thought to correlate to the powder

size. The nano-submicron pores had a greater influence on thermal properties than did the larger

pores, and the increased presence of nanopores was an effective way to improve the thermal

insulation properties of SPS coatings. In a study of SPS coatings with columnar grain orientations,

particle size was seen to affect pore size, and SPS coatings have also been compared to their APS

counterparts with multilayered columnar-oriented porous structures [43]. In APS-prepared TBCs,

small pores were not exhibited, due to the larger particle size of the powder in comparison with

the SPS-fabricated ones, while the total porosity remained lower in the APS coatings 43.

10
While pores can help improve the thermal performance of a coating, it also presents issues. For

example, in engine applications, open pores can lead to decreased volumetric efficiency, allowing

gas to become entrapped in the coating. To address the problem of open pores, many different

methods have been studied for creating dense outer surfaces. These include thermal shock heating,

laser glazing, sol-gel and aluminum phosphate sealings, and segmentation cracking introduced by
44,45
thick spray coatings . In a study of 8Y2O3-ZrO2, coatings plasma sprayed onto 4142-steel

substrates had their open pores significantly reduced via the surface modification methods of laser

glazing, sol-gel sealing, and segmentation cracking. While very dense surfaces were seen when

laser glazing was applied at a penetration depth of 80–120 um, the overall porosity (determined

via image analysis) significantly decreased, whereas significantly improved wear and erosion

properties were noted of the as-sprayed reference coatings 45. New fabrication methods that afford

control of pore size, orientation, and location, can improve the thermal insulation and mechanical

properties of TBCs.

2.5. Functionally Graded Barrier Coatings


46
One of the most intriguing innovations regarding TBCs is the use of FGMs . FGMs are

materials that have spatially dependent and carefully controlled microstructures and compositions,

and have been used in various high-temperature applications 47. Figure 2 shows an example of a

functionally graded TBC with a constant 20% gradient in the interface composition between the

bond coat and topcoat. Originally, FGMs were fundamentally designed to reduce thermal stresses

and the delamination of metal-ceramic interfaces for heat resistance applications; however, FGMs
48
also indicate exceptional opportunities for multifaceted applications in various industries .

Compositionally graded FGMs utilize the gradual changes in material layer composition to

11
alleviate concerns over failures caused by material property mismatching 49. In TBC applications,

a functionally graded coating (FGC) is where the intermetallic bond coat layer is graded between

the metallic substrate layer and ceramic topcoat. This reduces the discontinuity of the thermal

expansion coefficient between the layers and alleviates much of the thermal stresses in the system
50
. In a similar manner, the thermal conductivity and oxygen diffusion through the graded layers

can be controlled spatially. Research has also been conducted into grading porosity in FGMs, a

process that can reduce the thermal conductivity through the bond coat and metal substrate layers

(thus improving thermal shock resistance), reduce residual thermal stresses, and slow the oxide

layer growth rate. In the high-temperature regime of 900–1000°C, the oxide layer growth rate in

compositionally graded TBCs was found to be slower than in traditional duplex TBCs.

Additionally, the nature of the residual stresses in traditional TBCs were found to be tensile and

constant, while that in the compositionally graded TBCs were compressive and increased with

both time and temperature. In a study of porous FGMs that featured a conventional NiCoCrAlY,

TBCs were fabricated with APS porosity (increasing towards the surface of the topcoat) to address

the residual stresses that develop between the topcoat and bond coat. The results demonstrated

improved thermal conductivity and thermal shock resistance, along with the potential for increased

inlet temperatures in gas turbines 44.

In high-temperature applications, FGCs could improve component performance by optimizing

the thermal conductivity across the coating without compromising the structural integrity of the

coating layers. More research is needed to quantify FGC performance in relation to the various

harms that TBCs experience in specific harsh environments. Additionally, FGMs could enable the

use of materials normally made incompatible by property mismatching. In fact, by tailoring the

bond coat material and relative porosity of the FGC while also controlling the desired

12
microstructure, it may be possible to create environment-specific thermally compatible topcoat

layers on a wide variety of metal substrate materials.

Figure 2: Schematic of a functionally graded TBC.

3. TBC Fabrication Techniques

Several types of available coating methods have been used for various applications in the

automotive and power generation industries. These include diffusion coating processes, thermal

spray processes, Ni-dispersion coating, electric arc wire spray coating, physical vapor deposition

(PVD) processes, and magnetron sputtering systems 51. Plasma spray techniques have been widely

used to yield ceramic coatings, due to their flexibility and ease of application. They have the

capability of depositing a wide range of coatings onto various substrates 52. Issues associated with

these techniques can lead to reduced TBC lifetimes, and they include high porosity, low bond

strength with substrate, internal residual stresses, and oxidation during fabrication itself. Newly

developed chemical vapor deposition (CVD) and laser-induced CVD (LCVD) processes have

demonstrated better microstructure controllability, but low deposition rates make large-scale

application to TBCs unrealistic in comparison with traditional or plasma spray techniques.

13
The two common methods for fabricating TBCs are EB-PVD and plasma spray deposition 8.

Among many other factors, thermal conductivity, strain tolerance, and reproducibility must be

considered when promoting the widespread use of TBC deposition methods 53.

3.1. Electron-Beam Physical Vapor Deposition

In EB-PVD, a target anode is bombarded by a highly energized electron beam produced by a

charged tungsten filament within a high-vacuum environment 54. The electron beam triggers atoms

from the target to convert into the gaseous phase. These high energized atoms then precipitate into

a thin, solid coating form of the anode material, covering everything in the vacuum. This is also

known as an “electron evaporation system,” since the incident electron beam evaporates the

coating (source) material and deposits on the substrate. Coating materials used for EB-PVD

contain but are not restricted to ceramic, titanium, and zirconium, with Aluminum Titanium

Nitride (TiAlN) 55 YSZ being the standard TBC material for turbine applications fabricated in this
56
manner . Coating thickness can range from 1 nm to a few microns, and can enhance the

substrate’s thermal and optical properties 57. The major applications of EB-PVD coatings lie in

employing TBCs on automotive, aerospace, power generation, and marine high-temperature

components 58. They also involve employing thin films on medical equipment technology such as
59
biomedical devices and implants . For example, the hydroxyapatite (HAp) coating commonly

used for dental and orthopedic prostheses was also produced via the EB-PVD technique. The

metallization of semiconductor constituents—more commonly used in micro-electromechanical

systems, radiofrequency (RF) power amplifiers, light-emitting diodes, laser recorders, and

consumer electronics—is accomplished through EB-PVD.

14
3.2. Plasma Spray Deposition

The foremost processes for fabricating YSZ topcoats are APS 60 and EB-PVD 61. Plasma spray

deposition in controlled surroundings was established in the late 1960s to decrease the detrimental

effects (e.g., oxidation) and undesired contaminations found in the coatings as a result of the in-
51
flight heated particles interacting with the surrounding environment . The plasma-sprayed

coating can be seen in Figure 3 (a–d)—as observed via environmental scanning electronic

microscopy (SEM)—showing a compact and homogeneous microstructure 62.

Figure 3: Environmental SEM fractography images of plasma-sprayed yttria-stabilized zirconia


ceramic coatings on a metallic substrate: Coatings of (a) YSZ1, (b) YSZ2, (c) YSZ3, and (d)
YSZ4, where 1, 2, 3, and 4 represent different sets of coatings. Reprinted with permission from
ref. 62. ©2014 American Chemical Society.

Plasma spray deposition can be classified into three major categories: suspension/solution

precursor APS, low-pressure/vacuum plasma spraying, and suspension/solution precursor plasma

spraying (SPPS). Lately, new thermal spray methods such as suspension plasma spraying and

plasma spray-PVD have been largely explored in terms of TBC topcoat deposition.

15
In the case of APS, a thermal plasma is produced via an RF discharge or through a direct

current (DC) arc. A flame temperature of 8000–14,000 K in the jet core, in addition to particle

velocities of 20–500 m/s, may be possible depending on the particle size distribution 52. The high

temperatures harvest a high volume of particle melt that, along with the high particle velocities,

yields outstanding deposition densities, low-porosity coatings, and better bonding strengths in

comparison to thermal spraying processes 51.

In terms of application, the APS technique is restricted to depositing small particles sized 10–
63
100 µm, due to the feedstock powder’s unsuitable flowability for plasma spraying . To allow

usage of nanoscopic powders in the feedstocks, a wide range of diverse solutions have been

industrialized to substitute for the traditional injection of powder. Most notable among them are
64
SPS and SPPS . The main difference between these two methods is the precipitation of the

deposited particles in-flight. These methods increase the flexibility of plasma deposition

technologies by using smaller particle sizes for the feedstock materials and permitting deposited

coatings with diverse microstructures. Therefore, creating TBCs for high-temperature applications

is an area in which SPS and SPPS have been utilized. This is mostly due to the fine porosity, strain-

tolerant columnar structures or vertical cracks in the SPS- and SPPS-deposited coatings, resulting

in a thermal conductivity lower than that of EB-PVD or traditional APS coatings.

Use of low and very low pressures in plasma deposition fosters the growth of high-quality

thermal-sprayed coatings 65. The extent of the pressures used may differ, usually in the range of

4,000–40,000 Pa for low-pressure plasma spraying, and as low as 100 Pa for very low-pressure

plasma spraying. When the pressure value is lower than 100 Pa, the process is better known as

“vacuum plasma spraying” 66. This very low-pressure technique produces plasma-sprayed coatings

16
with low porosity values of ~1% and generates columnar structures similar to those created via

PVD—a fact of significant interest in regard to various TBC applications 16.

Despite being a popular spray coating method, plasma spraying yields ceramic coatings with

porous microstructures that are extremely prone to degradation in harsh environments and high

temperatures. However, CVD can deliver dense, highly pure coatings with outstanding

morphologies and better conformal coverage, albeit at a relatively low deposition rate. Yet, one

recently established LCVD method is both efficient and able to provide a superior deposition rate
67
, making it comparable to plasma spraying and EB-PVD. The microstructure achieved by LCVD

is porous, but the presence of nanoscale pores in TBCs is desirable for reducing thermal

conductivity.

3.3. Electrophoretic Deposition


68
Lately, electrochemical processes have been utilized to fabricate metal/ceramic coatings .
69
These methods can be categorized as electrophoretic deposition processes and electrolytic

depositions 70. Electrophoretic deposition processes are techniques in which charged particles are

deposited from a suspension source into an oppositely charged electrode, and electrolytic

depositions involve coating metal oxides onto a cathode that consume an aqueous solution of the

metallic salt. Electrolytic deposition uses soluble salts of the targeted metals and deposit an oxide

layer, whereas electrophoretic deposition uses suspended solid metal oxide particles in an
22
electrolyte solution . Realizing the ceramic coatings via electrochemical means is a modern

coating fabrication technique. The electrophoretic technique is a low-temperature process that

could deliver efficient coatings with improved sinterability and a homogeneous microstructure

throughout 71.

17
For example, Hashaikeh et al. demonstrated an electrochemical method of fabricating a
71
NiCoCrAlY/MgO/YSZ multilayered TBC for nickel superalloy substrates . The NiCoCrAlY

alloy particles and MgO thin film layers were achieved using the electrophoretic deposition and

electrolytic deposition methods, respectively, when the YSZ layer was coated electrophoretically.

The middle MgO layer was initiated to acclimatize the change in thermal expansion coefficient

between the YSZ ceramic layer and the NiCoCrAlY metal alloy layer. Hu et al. 72 demonstrated

the DC electrophoretic deposition method for depositing gadolinium‐doped ceria (GDC) as a

barrier layer onto YSZ for solid oxide fuel cells. However, at a sufficient current density using this

deposition technique, the produced H2 grows in the form of gas bubbles that damage the green

density of the coating and the adhesion between the deposit and substrate. Recently, a thick GDC

layer was successfully realized via high-frequency alternating current electrophoretic deposition

(AC-EPD) with an asymmetric square wave 73. The high frequency confirms that the total current

is controlled by a charging current, thus preventing gas bubble formation. The deposit rate

enhances with the increased voltage ratio and forward width percentage. In the AC-EPD technique,

the deposition rate of GDC particles is conjointly controlled by the transport and desorption

processes. Figure 4 shows a surface and cross-section morphology of the GDC coating, as obtained

via AC-EPD at 500 Hz, a voltage ratio of −100/+80, and a duty cycle of 50% with high-temperature

sintering at 1250°C for 4 hours. The attained GDC coating is almost fully dense, with a just few

isolated pinholes. This suggests that it is capable of performing as the barrier layer 73.

18
Figure 4: Surface and cross-sectional morphology of the GDC layer. Reprinted with permission

from ref. 73. ©2020 American Chemical Society.

3.4. Additive Manufacturing

Additive manufacturing is well-suited for FGM fabrication. FGMs of various alloys have been

studied using manufacturing techniques such as directed energy deposition (DED), powder bed
15
fusion, stereolithography, material jetting, and fused deposition modeling . Specifically, DED

techniques such as laser engineered net shaping (LENS) and laser direct metal deposition allow

for fine control of process parameters, most importantly the feed powder composition via the use
74
of two or more feeder systems . The LENS method is especially attractive due to its fast

processing speed, parameter control, and ability to produce spatially dependent components and

FGMs with complex shapes and microstructures. AM of metal FGMs has recently gained

attention—especially in terms of LENS, for which the ability to gradually vary the mixing of

multiple powders and change the chemical composition and processing parameters at each location

benefit from the FGM design. For example, Gualtieri et al. found LENS well-suited for fabricating

ceramic vanadium carbide in a compositional gradient using 304L stainless steel (SS) 75. In a study

pertaining to a functionally graded YSZ coating on a 316L SS substrate, the LENS method
76
produced a good interfacial microstructures . LENS can also be used to control resulted in

structural features such as porosity, and FGM porous components have been fabricated via LENS

to improve the strength-to-weight ratio and mechanical properties, sparking a reduction in laser

power that correlates to a decrease in the Young’s modulus and a higher porosity content 77.

Other stereolithography methods such as Digital Light Processing (DLP) can be used to create

high tolerance “green” FGM parts that require specific post-debinding processes to make them solid.

19
Fine lattice structures ceramics predesigned with varying porosities were fabricated through DLP,

and computed tomography (CT) scanning and microscopy revealed the total porosity to be 50–80%,
78
depending on the debinding and sintering processes implemented . Thermal conductivity was

initially studied in a similar ultraviolet (UV) process (i.e., binder jet printing), with increased pore-related

thermal conductivity in copper parts proven via laser flash methods. Then, a model was developed for

predicting the thermal conductivity and upper limit of the grain boundary thermal resistance. The

observed thermal conductivity was lower than those predicted for both powder metallurgy and

structural components, and the author attributes this to less uniform porosity distributions, as well as

grain variation due to inconsistencies in the powder and processing conditions—key factors

requiring further investigation to realize optimum TBC applications in AM. Along with those seen

in common spray deposition methods, new challenges arise when using high-energy input methods

such as LENS and powder bed fusion to develop FGMs. Such challenges include cracking,

intermetallic phases, unmelted particles, cooling rates and deposition parameter control 14. Studies

remain to be conducted on the performance and design of AM-produced FGMs for use in extreme

environments, especially in cases of fusion-based DED AM, which may present issues regarding the

coalescence of dissimilar metals and the formation of undesired phases in the microstructure. These

challenges can be addressed through improved spatial design of microstructures and compositions

for specific coatings intended for industrial applications.

4. TBC Characterization

The effect of harsh operating conditions (e.g., corrosion, foreign object damage, etc.) on TBCs

causes erosion and delamination, thus reducing the overall functionality and strength of the TBCs
11,79
. Consequently, various advanced nondestructive test (NDT) methods for TBC evaluation and

20
failure detection during service are commonly used in TBC applications 80,81. Among such NDT

methods, characterization techniques such as ultrasonic waves testing, eddy current testing, x-ray
82
imaging, rare-earth luminescence examination, etc., are particularly significant . Some NDT

technologies can also measure TBC thickness and monitor their condition 83,84.

However, these characterization methods have their downsides; for example, in evaluating

complex coating components, the eddy current test signal is vulnerable when nonmetallic materials

are involved, as it requires the lift-off operation step, which creates large noise in the signal 85. The

ultrasound acoustic waves are significantly limited by the edge effects of the coatings, and the
86
constraints imposed by liquid coolants make this technique unpopular . The rare-earth

luminescence process requires additional doping of rare-earth elements into the coating; thus, it

may not be helpful in quantitatively characterizing the morphology of barrier coatings 87,88. On the

other hand, x-rays are detrimental to the human body, due to their high radiant energy 89,90. Hence,

these characterization techniques may be unsuitable for use as real-time NDT techniques to

characterize the interface morphology of TBCs.

Recently, nondestructive interface morphology characterization of TBCs via terahertz time-

domain spectroscopy (THz-TDS) has garnered huge interest, since the frequency range of terahertz

waves is 0.1–10 THz, which lies between that of infrared waves and microwaves and is therefore

appropriate for analyzing TBCs 91. Also, many nonmetallic dielectric coating materials—which

are opaque in the range of infrared and visible light—can be accessed by THz waves. Additionally,

THz-TDS characterization affords noncontact, nonionizing, nondestructive, real-time assessments


92
. The thickness of the TBC, the evolution and growth of the TGO layer, and stress-induced air-

filled pores have also been determined via this technique 93,94, which has also been effectively used

21
95
to characterize materials with composite structures (e.g., polymer-coated steel , glass-fiber-

reinforced plastics 96,97, integrated circuit packages 98, and pharmaceutical drugs 99,100).

TBC erosion has been considered the secondary cause of failure; however, this has become

more of an issue and received more attention in recent research 101. Ye et al. 80 reported a method

of characterizing TBC erosion using a THz inspection. The coating thickness—as determined

through SEM—was 405.68 ± 11.33 μm, as shown in Figure 5. In this study, a THz-TDS

characterization of the reflection mode was employed to analyze the thickness and interface

morphology of TBCs post-erosion. Laser-based NDT characterization techniques employing

cheaper, nanosecond optical pulses or continuous-wave laser beams also exhibited the potential

for measuring thermal transport and ultrasonic wave propagation at the micron scale. The transient

grating spectroscopy method uses laser pulses from a nanosecond laser source and launches

surface acoustic waves during surface probing, thus delivering information about material

properties such as the elastic modulus, thermal diffusivity, etc. This technique has also been widely

used in layered thin film materials 102 and nanostructured acoustic metamaterials 103, and could be

useful in TBC characterization. Lastly, laser-based thermoreflectance techniques (e.g., spatial

domain thermoreflectance) utilizing continuous-wave lasers have also proven very effective at

characterizing mesoscale thermal transport properties. This continuous-wave lasers based

technique has been widely used to study thermal conductivity in ion-irradiated nuclear fuels and

can potentially be used for characterizing TBCs 104.

22
Figure 5: SEM interface morphology of TBCs pre-erosion. Reprinted with permission from ref. 80.
©2019 by the authors and licensee MDPI, Basel, Switzerland.

In-situ characterizations of TBC failure mechanisms have also been proposed, such as those
105
done by Zhu et al. . In that study, key TBC durability failure mechanisms such as the elastic

modulus and fracture toughness were inspected using a method comprised of high-temperature

three-point bending and digital image correlation for APS-fabricated TBCs. Zhu et al. 105 proposed

that the topcoat and substrate elastic moduli and surface interfacial fracture toughness can be

determined at high temperatures by monitoring crack initiation and propagation through the

surface and interfacial layers during bending testing. It was demonstrated that, when the

temperature increases from 30 to 800°C, the TBC elastic modulus and fracture toughness decline

from 20.3 to 13.1 GPa and from 1.31 to 1.16 MPa m1/2, respectively, while the interfacial fracture

toughness rises from 83.7 to 156.3 J/m2. Elastic modulus and fracture toughness results were

validated using data available in the literature, in addition to tests such as high-temperature

indentation. A decrease in the topcoat fracture toughness and elastic modulus with increased

temperature was observed, while the interfacial fracture toughness increased 105.

23
Microstructural characterization via classic destructive methods is still widely conducted to

assess TBC properties. Using mercury porisemetry and SEM image analyses, Portinhaa et al.

successfully characterized APS-fabricated TBCs with porosity gradients. Image analysis is an

effective way to determine the microstructures and defects present in coating structures. In a study

by Nouri et al., microstructural characterization of the APS-fabricated TBC on both a single-layer,

Co-based topcoat as well as a two-layer YSZ topcoat on a γ-TiAl intermetallic alloy was performed

using high-temperature isothermal oxidation tests (i.e., x-ray diffraction, SEM, and energy

dispersive spectroscopy), showing that the two-layer TBC performed better under high-

temperature oxidation 106. The laser flash method has been used to measure the thermal properties

of porous architected YSZ coatings fabricated via SPS, revealing the effect of process parameters

on porosity, the fact that increased porosity is an effective mechanism for reducing heat transfer,

and the larger effect of smaller pores (<1 µm) on overall thermal properties as compared to that of

larger pores 42.

5. TBC Failure Mechanisms

Research and industrial experience reveal various conditions that can result in the structural

harm and performance degradation of TBCs. Damage in TBCs is mostly the result of external

mechanical damage and compaction, thermal shock and gradients, sintering, calcium magnesium

alumina silicate (CMAS)- and environment-induced erosion, and corrosion and oxidation. Figure

6 shows a pictorial diagram of various types of damage that can occur to TBCs in high-temperature

or otherwise harsh conditions. This damage stems from a variety of factors and affects the TBC

structure in a variety of observable ways. Of all the various damage-causing phenomena, those

that most impact a TBC in its operating temperature regime is thermal expansion mismatching,

24
diffusion between layers, and oxidation in the bond coat. Certain experimental strategies are being

employed to improve TBC performance by increasing their resistance to these common failure

mechanisms.

Figure 6: Schematic illustrating the types of harm that occur to TBCs in high-temperature or
otherwise harsh conditions.

Thermal property mismatching is a primary driver of material failure, as differences in thermal

expansion during thermal shock or gradients can result in material fracture. Thermal stresses and
107
any remaining residual stresses can have a similar effect as well . This commonly results in

spallation of the ceramic topcoat from the bond coating, due to variations in linear and volumetric

expansion between the two materials, thus leading to formation of cracks and voids in the material
108
. Though the TBCs are designed with this specifically in mind, the problem can be exacerbated,

depending on anisotropy between the layers that results from the specific material’s processing

and manufacturing history. Additionally, the creation of unwanted phases in the bond coats and

ceramic topcoat can add to the property mismatching in the material. These phase changes can

also impact other degradation mechanisms such as excessive oxidation.

25
The bond coat in a TBC will preferentially form an oxide layer, as oxygen readily diffuses

through the zirconia layer due to its porosity. The oxide layer between the bond coat and the

ceramic topcoat could also cause delamination and spallation of the layers once the oxide reaches

a critical thickness. This also occurs due to property mismatching between the oxide and

surrounding layers, with thermal stresses and crack tip formation usually occurring on the oxide

layer interface peaks 109. Note that the shock resistance of the coating decreases as a function of

the oxide layer thickness, as well as due to the bond coat layer’s decreased ductility, which

normally relieves some of the thermal stresses. Another type of oxidation failure occurs when

stacked laminar sections of oxides form in the coating. If this phenomenon occurs, the coatings

are more likely to experience cracking perpendicular to the coating layers, as well as crack growth

through the layers. It has been shown that the oxide growth rate in TBCs can be slowed by

increasing the thickness of the top barrier coat, thereby increasing the overall oxidation life.
110
Izadinia et al. found spallation due to lateral cracking at the oxide layer interface to be the

dominant failure mechanism in both typical and thick TBCs under isothermal oxidation conditions.

Note, however, that increased coating thickness decreases the overall bond strength of the layers,

since thicker coatings have increased tensile edge stresses that increase the possibility of

delamination. The solid-state diffusion of elements between layers can create complications in the

oxidation resistance of the TBC. In MCrAlY (e.g., NiCrAlY and CoCrAlY) bond coating, the

aluminum-created alumina layer serves as a vital oxygen diffusion barrier in the coating. If a

substantial amount of Al diffuses to the substrate layer, there will be less aluminum to react with

the oxygen, and the lifetime of the oxide layer will decrease. The typical morphologies (i.e.,

CoCrAlY and YSZ) of MCrAlY coatings are shown in Figure 7 111.

26
Figure 7: Morphologies of MCrAlY powders: (a) CoCrAlY and (b) YSZ. Reprinted with
permission from ref. 111. ©2014 American Chemical Society.

A handful of methods are suggested in the literature to improve TBC resistance to typical

failure mechanisms. The primary method is to enhance oxidation and thermal shock resistance by

reducing the thermal expansion coefficient mismatching in the material. This can also be

accomplished by decreasing the oxygen permeation and thermal conductivity while increasing the
108
overall phase stability of the topcoat at high temperatures . Use of DCLs and FGMs is an

attractive way to improve these properties.

6. TBC Applications

TBCs play a precarious role in the development of insulation capabilities for a wide spectrum

of constituents in numerous industries such as those involving aero engines, gas turbines, and parts

for combustion/nuclear power plants that generate extremely high-temperature and otherwise

harsh environments. TBCs are also considered by their uniquely low thermal conductivity and

ability to withstand a large temperature gradient upon exposure to heat flow. TBCs provide a wide

range of benefits such as increased thermal conductivity, increased engine power efficiency,

decreased fuel consumption, increased exhaust gas temperature, high thermomechanical stability,

increased lifespan of parts through decreased fatigue and stress on components, and shielding of

27
metallic structural components from extreme temperatures and harsh environments. Moreover,

when a bond coat is used to afford TBCs better adhesion, even more advantages become apparent.

Among the various types of TBC, YSZ is the most successful ceramic top layer, owing to its

extremely low thermal conductivity and good phase stability, as well as the fact that it is widely

manufacturable, possesses a good combination of low thermal conductivity and thermal expansion,

and is stable at up to 1200oC when combined with a metallic interlayer 112


. Applying a 0.1–0.2-

mm-thick YSZ layer to the surface of a high-temperature component can reduce the exterior
113
temperature to 170°C . Such increased efficiency in high-temperature components such as

turbine can increase the thrust by 1% and efficiency by the order of 1% with associated fuel

economies 114. In diesel engine applications, YSZ TBCs have been studied extensively in terms of

various features such as the effects of coating porosity on the combustion cycle, the effects of

thermal insulation on volumetric efficiency, and combustion wall temperature 115. TBCs for diesel

engine applications were recently studied in terms of a novel architecture combining YSZ and

gadolinium zirconate structures, with SPS columnar topcoats on both ceramic and metallic bond

coats under thermal cycling, resulting in extended service lifetimes as compared to APS coatings
116–118
.

6.1. Aerospace TBC Applications

Green energy demands and growing fuel costs are forcing aerospace industries to develop more

efficient turbine engines, rockets, scramjet engines, re-entry space vehicles, and other such

aerospace components. As combustion temperatures rise in commercial jet engines, efficiency

escalates while nitrogen oxides and carbon dioxide emissions decrease as the engines consume

fuel more efficiently. Thus, turbine blades must be able to function for long operating periods in

28
temperatures well above their melting point. Development of materials and coatings that can

withstand these intense temperatures has for many years been the focus of aerospace industry.

One such coating material now being explored for use in TBCs is rare-earth zirconates 11.

Two fabrication processes are generally used to apply TBCs onto aerospace components.

Plasma-sprayed coatings came first, and these remain in service today. Recently, however, a

second type of TBC fabrication process, PVD, was successfully brought into commercial service.

The advantages of both types of TBC deposition process for commercial aircraft gas turbine

applications—especially regarding PVD coatings—have been pivotal in the acceptance of recent

technologies.

TBC applications for aeroengine have dramatically increased over the past few decades. These

TBCs are extensively used on hot stator and rotor components such as fuel vaporizers, combustion

chambers, vanes, and blades to extend component service lifetimes, thus improving the durability

of aero engines and decreasing overall operating costs. In the U.S., various TBCs were

efficaciously applied on the hot stator and rotor sections of various aero engines by both the Pratt
119
& Whitney Group and General Electric . Table 1 outlines the different TBC applications in

aerospace industries.

Table 1: TBC applications on commercial aero-engine parts 119.


Bond coat Topcoat Pratt & Whitney General Electric
Group
APS MCrAlY APS YSZ not applied CF6-80 Vane 2
(aluminized)
Pt-Al EB-PVD YSZ not applied CF6-80 Blade 1
aluminized EB-PVD YSZ not applied CFM56-7 Vane 1
vacuum plasma- APS YSZ V2500 Vane 1 CF6-50 Vane 2
sprayed MCrAlY

EB-PVD MCrAlY EB-PVD YSZ PW2000 Blade 1 not applied

29
The TBCs reduce the temperature of the blade alloy and protect against oxidation and hot
120
corrosion from high-temperature gases . TBCs are usually used to shield nickel-based

superalloys against both melting and thermal cycling in turbines. Together with cold air flow,

TBCs help to expand the allowable gas temperature above that of the superalloy’s melting point,

thus significantly improving turbine performance, life expectancy, and efficiency. Besides the

improved thermal behavior, the strength and stiffness of the aerospace sandwich panels were

remarkably improved by introducing the TBC layers 121.

Radiation exposure for aerospace components can be reduced by consuming suitable optical

reflective coatings. Tactical placement of oriented cracks and tunable ordered pores within the

topcoat, along with the suppression of sintering at operating temperatures, can also decrease both
122
phonon conduction and radiative heat transfer . Although the search for suitable new TBC

ceramics for aerospace industries remains ongoing, success will eventually come in the form of a

more complete appreciation of all the promising characteristics that have made YSZ TBCs so

effective to date, then combining these characteristics into a ceramic with lower thermal

conductivity and a higher use temperature.

6.2. Electronics TBC applications

Due to the complex effects of interface coupling and the vast differences in the microstructures

and adhesion properties of different substrates, failure of thin films and coatings is always

unpredictable. TBCs are also being considered for an ever-increasing number of applications, such

as in communication tools, electronic devices, and high-tech displays.

Zirconia-based TBCs, along with promoting surface biocompatibility and bioactivity, improve

thermal insulation. The low thermal conductivity of such coatings ensures that the temperature of

30
the metallic substrate is maintained at its original value 123. The coatings have also exhibited strain-

tolerant properties that afford protection to biomedical devices and electronics components 124.

6.3. Automobile TBC Applications

To reduce greenhouse gas emissions, the automobile industry has been applying TBCs on

combustion engine components such as pistons, cylinder heads, chamber walls, valves, and ports

to enhance thermal insulation and engine efficiency, especially in low-heat-rejection engine

applications. The efficiency of low-heat-rejection engines is improved by coatings that can reduce

thermal swing—namely, the temperature difference between the cylinder wall and the gas—by

increasing wall temperatures as a result of affording them insulation, thus reducing heat loss from

the cylinder and thereby preventing detrimental effects such as intake air heating and poor

volumetric efficiency 40,125.

Requirements for effective coating materials in combustion applications include low thermal

conductivity, corrosive environment resistance, coefficient of thermal expansion, and thermal

shock resistance. Common TBC materials such as zirconia and silicon nitride have proven

effective at increasing operating temperatures and providing thermal insulation 126. The adoption

of TBCs has been hindered because application thereof can reduce volumetric efficiency because

of the higher wall and gas temperatures experienced due to the increased thermal insulation. YSZ

TBCs were applied by Sivakumar et al. to pistons in diesel engines, 8 mol% Y2O3 and fully

stabilized ZrO2 were applied to pistons via plasma spray at a layer thickness of 100 µm, and

experimental analysis of heat loss, thermal efficiency, brake-specific fuel consumption, and

emissions in four-stroke diesel engines were performed. Significant improvements were noted,

such as a specific fuel consumption reduction of 3.38–28.59% and a hydrocarbon emissions

31
126
reduction of 35.27% . In a study by Caputo et al., the effect of anodized aluminum, a new

material whose thermal conductivity and heat capacity are lower than those of YSZ, was

investigated via computational fluid dynamics and experiments on aluminum pistons. It was found

that aluminum pistons applied with for its fuel consumption and heat transfer reduction, finding

about a 1% and 6% reduction, respectively. A decrease of about 2% in experimental engine

efficiency was observed when compared to numerical simulations, and the primary cause of the

inefficiency was thought to be the surface roughness of the TBC: namely, a roughness of Ra = 8

µm as compared to the Ra = 3.2 µm of the uncoated piston. The surface roughness of the new

multilayered zirconia TBC on pistons was also shown to affect heat loss when compared to

baseline pistons and zirconia spray coated pistons in numerical simulation of heat loss 127. In this

multilayered TBC system, as investigated by Uchida et al., a 200-µm lower- layer coating of high

porosity (~60%) was implemented to reduce thermal conductivity and increase surface

temperature, while a denser top layer of 50–80 µm was designed to reduce porosity—specifically,

open pores—as well as surface roughness, which can hinder thermal performance. Uchida et al.

speculated that open surface pores could allow gas flow into the pores and increase heat transfer,

while the high-porosity layer led to premature failure compared to the baseline pistons. Thermal

swing heat loss reduction by applying a porous TBC has also been seen in technologies such as

Thermo-Swing Wall Insulation Technology, utilizing silica-reinforced porous anodized aluminum

coatings applied to the piston 40. Silica-reinforced porous anodized aluminum coatings applied at

a thicknesses of 20–300 µm to the top surface of pistons in turbocharged engines demonstrated

improved fuel efficiency, increased exhaust gas temperatures, reduced cooling heat loss, and

coating thickness optimization based on pump performance 40. While improved thermal insulation

via multilayered TBCs has been proven to enhance engine efficiency, challenges such as

32
deterioration of lubrication properties at high temperatures, increased carbon monoxide emissions,

and decreased volumetric efficiency have yet to be overcome 128.

6.4. Nuclear TBC Applications

Nuclear applications bring about greater complications for TBCs than do traditional

environments. Since radiation-induced kinetics can cause even greater coating damage, radiation

damage resistance is a major consideration. Corrosion kinetics experienced in next-generation

reactors are mostly due to molten metal, molten salt, or gas, and occur at different temperature

regimes than found in current pressurized-water reactors and turbine environments. The neutronics

properties of coating materials are of great importance, as unwanted neutron poisons, moderators,

and activated materials may impact component performance. Material transmutation should also

be considered.

6.4.1. Nuclear Thermal Propulsion

In carbon-fuel-based nuclear thermal propulsion (NTP) applications, hot corrosion of the

graphite structural fuel elements negatively impacts the neutronics and performance, due to the H2

propellants’ reactivity with the carbon at high temperatures. Certain carbides have been developed

to alleviate this effect, but thermal expansion mismatching creates the need for interface layers of

Mo, Mo2C, and Nb that insulate before the outer ZrC layer. In this technology, Mo2C acts a

diffusion barrier to minimize the amount of carbon reacting with the Mo. The biggest challenges

in the NTP research field relate to minimizing three things: high-temperature hydrogen corrosion,

fracture caused by temperature gradients, and radiation damage that would impede reactor

performance during operation 129. In general, reduced hydrogen ingress into carbon fuel elements,

33
along with reduced property mismatching in the materials could improve component lifetime and

overall performance. Another fuel option being studied for use in NTP systems is W/UO2 and

Mo/UO2 ceramic metal (CERMET) fuel, which must be able to operate in excess of 2700 K while
130
remaining compatible with the H2 propellant . Thermal and chemical extremes can induce

undesirable effects in CERMET fuel, such as cracking of fuel pellets, swelling, and migration of

uranium into the surrounding fuel matrix. Certain CERMET designs that employ an aluminum

cladding around the bulk fuel in addition to interfacial barrier coatings between the fuel and

cladding have been developed to reduce the irradiation-induced diffusion between U and Al, which

creates a diffusion layer identified as a cause of structural failure in these fuels 131. It was shown

that a barrier layer interface such as ZrN can reduce the creation of diffusion layers in the materials.

Though a classical TBC with thermal conductivity reduction is undesired for NTP applications,

techniques applied to TBCs could be utilized to optimize aspects of NTP performance, such as via

diffusion mitigation and the reduced effect of thermal gradients in the material.

6.4.2. Nuclear Waste Management

Waster vitrification is a recommended nuclear waste management solution that places waste

into a borosilicate glass matrix to be stored for tens of thousands of years 132. The melter pots used

in the waster vitrification process experience high-temperature corrosive environments of

approximately 1273 K and must be protected from potential material failure during this process.

These melter pots often use nickel-based alloys that show naturally good mechanical properties

and resistance to corrosion and stress corrosion cracking 133. However, the operation lifetimes of

these alloys could be increased via diffusion barrier coatings. From a research perspective, graded

34
Ni-YSZ coatings are attractive options for diffusion barrier coatings in harsh nuclear environments
134
.

6.4.3. Advanced Nuclear Power Plants

6.4.3.1. Liquid Metal Fission Reactors

Liquid metals have been proposed as an alternative coolant to water, fostering lower-pressure,

high-temperature, fast neutron flux systems in advanced nuclear reactors. Structural materials in

contact with the primary coolant in sodium-cooled fast reactors will be exposed to operating

temperatures of over 600°C, high radiation fluence, cyclic thermal stresses, and corrosive

environments that readily corrode oxide passivation layers and dissolve the metal. Diffusion

coatings of NiAl have been studied as potential barrier coatings on both Ni- and Fe-based alloys

for applications in high-temperature sodium environments. The corrosion performance of the

nickel aluminide diffusion coatings was attributed to a NaAlO2 film proven stable in the sodium
135
environment . A lead-cooled fast reactor (LFR) is another design that utilizes a liquid metal

coolant and affords intrinsic safety features due to the lead or lead-based coolant. However, lead,

lead-bismuth, and lead-lithium coolants cause extreme corrosion and dissolution in many

structural metals, partly due to insufficient oxygen to grow a passivation layer. Alumina coatings

are particularly appealing as a solution to corrosion problems in LFRs because alumina is very

insoluble in lead over a wide range of potential reactor conditions. It was shown that alumina-

coated austenitic steels prepared via pulse laser deposition show no corrosion at 1,000 h under ion
136
irradiation, and retain their structural integrity under steady-state LFR conditions . Similar

alumina coatings produced via sol-gel methods with yttrium additions were studied for use in lead-

35
bismuth eutectic reactors. Accelerated corrosion tests conducted at 650°C under dynamic

conditions resulted in no erosion or lead-bismuth permeation in the coating 137.

Many liquid-metal-cooled fast reactor designs employ a core catcher, a passive safety device

for collecting molten fuel and structural materials in the event of a hypothetical core meltdown.

The purpose of the core catcher is to store the melted down fluid in a subcritical, coolable

geometry. Sacrificial ceramic coatings are being considered for decreasing the initial thermal

shock of the catcher structural material; the cooling rate will be low and thus allowing the system

to reach at an equilibrium as the coating is dissolved into the melt—with reduced thermal gradients

and system shock. Both alumina and YSZ produced via plasma spray processes are viable

candidate materials for sacrificial coatings in core catchers. It was shown that YSZ completely

leaches into the sodium melt, leaving behind only the substrate and bond coat layer. This

compatibility somewhat depends on the porosity percentage in the topcoat. In comparison, the

alumina coating showed almost no degradation or thickness reduction under the same conditions,

with only minor sodium interaction zones. However, at the high pressures likely to be experienced

in the core catcher, the alumina coating manifested delamination, cracking, and subsequent

detachment 138.

6.4.3.2. Molten-Salt Fission Reactors

Molten-salt reactors are a popular design for next-generation nuclear power, allowing for the

use of liquid fuel and high operating temperatures. In the development of fluoride-salt-cooled high-

temperature reactors, the coolant—commonly 2LiF-BeF2 (FLiBe)—continuously generates

tritium under neutron flux via multiple reactions, primarily Li-6 neutron capture. At high

temperatures, tritium readily diffuses into metal structures such as the reactor pressure vessel,

36
piping, and heat exchangers, eventually creating transport pathways through the reactor and

opening up the possibility of environmental release. Zheng et al. 139 found that alumina coatings

on 316 SS tubing significantly reduced the permeability of tritium. Tritium permeation barriers

deposited via plasma thermal spray and consisting of a NiCr topcoat, a NiCr + alumina bond coat,

and a 316 SS substrate were exposed to neutron irradiation to study their tritium permeation at

700°C, proving successful at decreasing the tritium’s permeability. Moderated fluoride-salt-cooled

high-temperature reactors and other molten-salt designs often use reactor-grade graphite, which

can be permeated by molten salt and other gases that affect the material’s microstructure and

increase the diffusion and retention of fission products and tritium 140. Additionally, depending on

the redox potentials in the molten salt, graphite can carburize metals—resulting in carbide
141
deposition at alloy surfaces—then corrode the material . In research pertaining to the Molten

Salt Reactor Experiment, 50-µm-thick CVD metal coatings of Mo, Nb and pyrolytic graphite were

under consideration, given corrosion and graphite compatibility, but early developed coatings were

deemed unsuitable due to in-core neutronics issues and spallation 142. Recent advances in coatings

and manufacturing have led to the consideration of materials such as glassy carbons and SiC for

these molten-salt applications. Glassy carbon consists of randomly oriented nanometer grains and

is prepared using various heat treatment processes on polymeric resins. A layer of glassy carbon

on typical graphite decreases the salt permeation, especially in tandem with a pyrolytic graphite
143
interlayer . Use of SiC as a graphite sealant coating has also proven feasible for reducing

permeation and diffusion through graphite. These coatings have been fabricated using CVD,

impregnation and pyrolysis, and siliconization. Research has shown that, when using an

impregnation and pyrolysis technique, the resulting SiC coatings exhibit a void-filling

phenomenon that reduces salt permeation. Double-layer coatings of glassy carbon and SiC have

37
been proposed and studied in regard to manufacturing a successful SiC coating using chemical
144
vapor reaction . More research is needed to characterize the performance of barrier coatings

under high neutron fluence and a wide range of corrosive and high-temperature environments.

6.4.3.3 Thermonuclear Fusion Reactors

Thermonuclear fusion reactor systems utilize the fusion of hydrogen isotopes to create energy
145
generally within a magnetic confinement . Permeation of hydrogen isotopes through flow

channels and plasma-facing components results in fuel loss and potential contamination. Thus,

application of a tritium permeation resistance coating on structural materials is an effective way to

alleviate tritium transport and safety risks. Coatings used in blanket channels must perform well

as permeation and diffusion barriers in high-temperature corrosive environments, have good

structural integrity, and be compatible with breeding blanket materials such as Pb-Li. Composite

Y2O3/Cr2O3 coatings—in which the Cr2O3 acts as an interlayer between the substrate and the

Y2O3—applied via CVD on the 316L SS substrate have been analyzed regarding tritium

permeation. Use of a double-layer permeation barrier coating showed better deuterium permeation

performance than that of a single Y2O3 layer of the same thickness; this finding holds true for other
146
hydrogen isotopes, as well . It was found that a Cr2O3/Al2O3 coating fabricated using PVD

varied process parameters showed exceptional hardness, good corrosion and irradiation resistance,

and excellent permeation performance at 823–973 K. The deposition of nitrides such as AlN, TiN,

and BN can act as a barrier for hydrogen permeation 147. Bazanella et al. showed that a 1.7-micron

double-layer AlN/TiN coating on a 316L SS substrate reduces deuterium permeability by 2–3


148
orders of magnitude as compared to the naked substrate . Though there are many options in

research for permeation barriers for fusion applications, most solutions involve thin film ceramic

38
layers that delaminate and increase in permeability as high radiation fluence damages the material
118
.

One challenge in designing TBCs for high-temperature nuclear components is the spallation

of top coating under harsh thermal cycling conditions. FGM ceramic coatings can be fabricated to
149
inhibit spallation and increase TBC thermomechanical performance . Though not currently

applied in industry, areas such as protective coatings for nuclear propulsion fuel elements,

diffusion barrier coatings for high-flux nuclear reactors, and uranium fuels could benefit from

these coatings.

7. Discussions and Prospects

The potential of TBC materials highly depends on the application and fabrication methods

involved. EB-PVD and plasma-spray deposition are two established fabrication methods, while

newer coating methods such as electrophoretic fabrication and advanced thermal spray techniques

(e.g., plasma-spray PVD) are under development and show appealing advanced properties. Three

major factors are typically considered in developing efficient TBC deposition methods: thermal

conductivity, strain tolerance, and reproducibility.

New advanced reactor designs drive the need for TBCs that can withstand harsher conditions

such as very high temperatures under neutron irradiation and highly corrosive environments.

Environmental diversity and performance requirements further contribute to the increased interest in

TBCs. To develop TBCs with suitable lifetimes under the demands of such conditions, advances in

new materials, manufacturing techniques, and TBC design should be considered. With evolving

performance requirements, new specialized coating functions are needed to achieve the desired

design performance in each application. New material concepts for TBCs should include optimized

39
thermal conductivity and exhibit better temperature stability than existing industrial options.

Furthermore, multifunctional layer systems using different application-focused materials are being

considered.

Fabrication of porous and compositional gradient FGMs has been employed to overcome the

mismatching in thermal expansion coefficients and stress concentrations within TBCs. In general,

the opportunity for advancements in barrier coatings is quite high, and a variety of new

manufacturing processes and materials can enhance the success rate and potential options for future

applications. These new manufacturing methods and science-based material approaches can lead to

the fabrication of better, more durable, more efficient coatings. TBCs can serve as an insulation

solution for both long- and short-term problems in the gas turbine, internal combustion, power

generation, and nuclear industries.

The global EB-PVD coating market was valued at $1.8 billion in 2019, and is projected to

reach $2.8 billion by 2027, growing at a rate of 5.9% increase each year from 2020 to 2027.

Moreover, increased automotive production and the implementation of solar energy and clean

power generation (e.g., nuclear and wind) are likely to expand the global EB-PVD market. Still,

the convenience of alternative options (e.g., electroplating) or even other PVD techniques (e.g.,

CVD, plasma-enhanced CVD, and magnetron sputtering) will likely reduce the reliance on EB-

PVD coatings, in turn hindering the growth of the global EB-PVD coating market.

8. Conclusions

Use of TBCs on components such as combustors, high-pressure turbine blades, and other such

structures is aggressively increasing in both commercial and military applications. This trend will

surely continue due to the TBCs’ insulating capability, which enables higher operating

40
temperatures and reduced cooling system costs, thus improving overall component efficiency.

Power generation sectors (e.g., fossil fuel combustion and nuclear reactors) are also progressively

increasing their usage of TBCs, and sectors such as nuclear are quickly adopting this technique for

use in next-generation power plants. TBCs are also scheduled to be used on modules fabricated

with ceramic matrix composites and metal-metal and metal-ceramic frameworks. Of the numerous

organic and inorganic compounds that have the potential to be used as TBC materials, zirconia in

the form of YSZ has been the primary focus of consideration, owing to its low thermal

conductivity, exceptional toughness, and high chemical inertness; however, other advanced

materials are also being researched. The plasma spraying technique is the most widely accepted

coating method for fabricating ceramics and TBCs, thanks to its flexibility for various substrate-

coating arrangements. Recently developed coating techniques such as the LCVD process offer

much higher deposition rates and better microstructures with nanopores; hence, they can produce

better TBCs. The stability of TBCs under high operating temperatures and harsh environmental

conditions has been a big concern. With the proper tuning of TBC and FGM manufacturing

parameters, a suitable TBC—graded in porosity and functional composition—can be developed,

thus improving thermal resistance and overall coating performance for extreme environments.

Certain AM techniques (e.g., LENS and photolithography) with carefully controlled process

parameters could produce enhanced performance in TBCs over a wide range of applications.

Author Contributions: Conceptualization, K.M., L.N., C.M.D., and I.J.V.R.; resources, I.J.V.R.;

writing—original draft preparation, K.M.; writing—review and editing, K.M., L.N., C.M.D., and

I.J.V.R.; supervision and project administration, I.J.V.R.; funding acquisition, I.J.V.R. All authors

have read and agreed to the published version of the manuscript.

41
Funding: This research was funded by the INL Laboratory Directed Research & Development

(LDRD) Program under DOE Idaho Operations Office Contract DE-AC07-05ID14517. The APC

was funded by the same project grant.

Acknowledgments: The authors gratefully acknowledge the U.S. Department of Energy for their

funding and wish to thank Amey R. Khanolkar and Subhashish Meher for their useful discussions.

Conflicts of Interest: The authors declare no conflict of interest. The funders had no role in the

design of the study; in the collection, analyses, or interpretation of data; in the writing of the

manuscript; or in the decision to publish the results.

References

(1) Brook, B. W.; Alonso, A.; Meneley, D. A.; Misak, J.; Blees, T.; van Erp, J. B. Why Nuclear Energy Is
Sustainable and Has to Be Part of the Energy Mix. Sustain. Mater. Technol. 2014, 1–2, 8–16.
https://doi.org/10.1016/j.susmat.2014.11.001.
(2) Sadekin, S.; Zaman, S.; Mahfuz, M.; Sarkar, R. Nuclear Power as Foundation of a Clean Energy
Future: A Review. Energy Procedia 2019, 160, 513–518.
https://doi.org/10.1016/j.egypro.2019.02.200.
(3) Eriksson, O. Nuclear Power and Resource Efficiency—A Proposal for a Revised Primary Energy
Factor. Sustainability 2017, 9 (6), 1063. https://doi.org/10.3390/su9061063.
(4) Jayakumar, T.; Mathew, M. D.; Laha, K. High Temperature Materials for Nuclear Fast Fission and
Fusion Reactors and Advanced Fossil Power Plants. Procedia Eng. 2013, 55, 259–270.
https://doi.org/10.1016/j.proeng.2013.03.252.
(5) Sengupta, P.; Manna, I. Advanced High-Temperature Structural Materials for Aerospace and
Power Sectors: A Critical Review. Trans. Indian Inst. Met. 2019, 72 (8), 2043–2059.
https://doi.org/10.1007/s12666-019-01598-z.
(6) Chattarki, Anoop. R.; Basavakumar, K. G. Thermal Barrier Coating on IC Engines; A Review. In
Recent Trends in Mechanical Engineering; Narasimham, G. S. V. L., Babu, A. V., Reddy, S. S.,
Dhanasekaran, R., Eds.; Lecture Notes in Mechanical Engineering; Springer Singapore:
Singapore, 2021; pp 47–68. https://doi.org/10.1007/978-981-15-7557-0_5.
(7) Cao, X. Q.; Vassen, R.; Stoever, D. Ceramic Materials for Thermal Barrier Coatings. J. Eur. Ceram.
Soc. 2004, 24 (1), 1–10. https://doi.org/10.1016/S0955-2219(03)00129-8.
(8) Miller, R. A. Current Status of Thermal Barrier Coatings — An Overview. Surf. Coat. Technol. 1987,
30 (1), 1–11. https://doi.org/10.1016/0257-8972(87)90003-X.
(9) Marino, K. A.; Hinnemann, B.; Carter, E. A. Atomic-Scale Insight and Design Principles for Turbine
Engine Thermal Barrier Coatings from Theory. Proc. Natl. Acad. Sci. 2011, 108 (14), 5480–5487.
https://doi.org/10.1073/pnas.1102426108.

42
(10) Gleeson, B. Thermal Barrier Coatings for Aeroengine Applications. J. Propuls. Power 2006, 22 (2),
375–383. https://doi.org/10.2514/1.20734.
(11) Padture, N. P. Thermal Barrier Coatings for Gas-Turbine Engine Applications. Science 2002, 296
(5566), 280–284. https://doi.org/10.1126/science.1068609.
(12) Clarke, D. R.; Levi, C. G. Materials Design for the Next Generation Thermal Barrier Coatings. Annu.
Rev. Mater. Res. 2003, 33 (1), 383–417.
https://doi.org/10.1146/annurev.matsci.33.011403.113718.
(13) Gupta, M.; Markocsan, N.; Li, X.-H.; Peng, R. L. Improving the Lifetime of Suspension Plasma
Sprayed Thermal Barrier Coatings. Surf. Coat. Technol. 2017, 332, 550–559.
https://doi.org/10.1016/j.surfcoat.2017.07.078.
(14) Yan, L.; Chen, Y.; Liou, F. Additive Manufacturing of Functionally Graded Metallic Materials Using
Laser Metal Deposition. Addit. Manuf. 2020, 31, 100901.
https://doi.org/10.1016/j.addma.2019.100901.
(15) Zhang, C.; Chen, F.; Huang, Z.; Jia, M.; Chen, G.; Ye, Y.; Lin, Y.; Liu, W.; Chen, B.; Shen, Q.; Zhang,
L.; Lavernia, E. J. Additive Manufacturing of Functionally Graded Materials: A Review. Mater. Sci.
Eng. A 2019, 764, 138209. https://doi.org/10.1016/j.msea.2019.138209.
(16) Young, E. J.; Mateeva, E.; Moore, J. J.; Mishra, B.; Loch, M. Low Pressure Plasma Spray Coatings.
Thin Solid Films 2000, 377–378, 788–792. https://doi.org/10.1016/S0040-6090(00)01452-8.
(17) Saini, A. K.; Das, D.; Pathak, M. K. Thermal Barrier Coatings -Applications, Stability and Longevity
Aspects. Procedia Eng. 2012, 38, 3173–3179. https://doi.org/10.1016/j.proeng.2012.06.368.
(18) Song, X.; Xie, M.; Zhou, F.; Jia, G.; Hao, X.; An, S. High-Temperature Thermal Properties of Yttria
Fully Stabilized Zirconia Ceramics. J. Rare Earths 2011, 29 (2), 155–159.
https://doi.org/10.1016/S1002-0721(10)60422-X.
(19) Ren, X.; Pan, W. Mechanical Properties of High-Temperature-Degraded Yttria-Stabilized Zirconia.
Acta Mater. 2014, 69, 397–406. https://doi.org/10.1016/j.actamat.2014.01.017.
(20) Smith, M. F.; Hall, A. C.; Fleetwood, J. D.; Meyer, P. Very Low Pressure Plasma Spray—A Review of
an Emerging Technology in the Thermal Spray Community. Coatings 2011, 1 (2), 117–132.
https://doi.org/10.3390/coatings1020117.
(21) Hashaikeh, R.; Szpunar, J. A. Electrolytic Processing of MgO Coatings. J. Phys. Conf. Ser. 2009, 165,
012008. https://doi.org/10.1088/1742-6596/165/1/012008.
(22) Zhitomirsky, I. Cathodic Electrodeposition of Ceramic and Organoceramic Materials. Fundamental
Aspects. Adv. Colloid Interface Sci. 2002, 97 (1–3), 279–317. https://doi.org/10.1016/S0001-
8686(01)00068-9.
(23) Boccaccini, A. R.; Zhitomirsky, I. Application of Electrophoretic and Electrolytic Deposition
Techniques in Ceramics Processing. Curr. Opin. Solid State Mater. Sci. 2002, 6 (3), 251–260.
https://doi.org/10.1016/S1359-0286(02)00080-3.
(24) Zhang, J.; Guo, X.; Jung, Y.-G.; Li, L.; Knapp, J. Lanthanum Zirconate Based Thermal Barrier
Coatings: A Review. Surf. Coat. Technol. 2017, 323, 18–29.
https://doi.org/10.1016/j.surfcoat.2016.10.019.
(25) Sudre, O.; Cheung, J.; Marshall, D.; Morgan, P.; Levi, C. G. Thermal Insulation Coatings of Lapo4. In
Ceramic Engineering and Science Proceedings; Singh, M., Jessen, T., Eds.; John Wiley & Sons,
Inc.: Hoboken, NJ, USA, 2001; Vol. 22, pp 367–374.
https://doi.org/10.1002/9780470294703.ch44.
(26) Xie, X.; Guo, H.; Gong, S.; Xu, H. Thermal Cycling Behavior and Failure Mechanism of
LaTi2Al9O19/YSZ Thermal Barrier Coatings Exposed to Gas Flame. Surf. Coat. Technol. 2011, 205
(17–18), 4291–4298. https://doi.org/10.1016/j.surfcoat.2011.03.047.
(27) Ghosh, S. Thermal Barrier Ceramic Coatings — A Review. In Advanced Ceramic Processing;
Mohamed, A. M. A., Ed.; InTech, 2015. https://doi.org/10.5772/61346.

43
(28) Ma, W.; Mack, D.; Malzbender, J.; Vaßen, R.; Stöver, D. Yb2O3 and Gd2O3 Doped Strontium
Zirconate for Thermal Barrier Coatings. J. Eur. Ceram. Soc. 2008, 28 (16), 3071–3081.
https://doi.org/10.1016/j.jeurceramsoc.2008.05.013.
(29) Jiang, K.; Liu, S.; Wang, X. Low–Thermal–Conductivity and High–Toughness CeO2–Gd2O3 Co–
Stabilized Zirconia Ceramic for Potential Thermal Barrier Coating Applications. J. Eur. Ceram.
Soc. 2018, 38 (11), 3986–3993. https://doi.org/10.1016/j.jeurceramsoc.2018.04.065.
(30) Padture, N. P.; Klemens, P. G. Low Thermal Conductivity in Garnets. J. Am. Ceram. Soc. 2005, 80
(4), 1018–1020. https://doi.org/10.1111/j.1151-2916.1997.tb02937.x.
(31) Fan, X.; Zou, B.; Gu, L.; Wang, C.; Wang, Y.; Huang, W.; Zhu, L.; Cao, X. Investigation of the Bond
Coats for Thermal Barrier Coatings on Mg Alloy. Appl. Surf. Sci. 2013, 265, 264–273.
https://doi.org/10.1016/j.apsusc.2012.10.192.
(32) Kim, G. M.; Yanar, N. M.; Hewitt, E. N.; Pettit, F. S.; Meier, G. H. The Effect of the Type of Thermal
Exposure on the Durability of Thermal Barrier Coatings. Scr. Mater. 2002, 46 (7), 489–495.
https://doi.org/10.1016/S1359-6462(02)00016-7.
(33) Bai, M.; Song, B.; Reddy, L.; Hussain, T. Preparation of MCrAlY–Al2O3 Composite Coatings with
Enhanced Oxidation Resistance through a Novel Powder Manufacturing Process. J. Therm. Spray
Technol. 2019, 28 (3), 433–443. https://doi.org/10.1007/s11666-019-00830-y.
(34) Nowak, W. J.; Kubaszek, T.; Góral, M.; Wierzba, B. Durability of Underaluminized Thermal Barrier
Coatings during Exposure at High Temperature. Surf. Coat. Technol. 2020, 382, 125236.
https://doi.org/10.1016/j.surfcoat.2019.125236.
(35) Guo, X.; Lu, Z.; Park, H.-Y.; Li, L.; Knapp, J.; Jung, Y.-G.; Zhang, J. Thermal Properties of La 2 Zr 2 O 7
Double-Layer Thermal Barrier Coatings. Adv. Appl. Ceram. 2019, 118 (3), 91–97.
https://doi.org/10.1080/17436753.2018.1510820.
(36) Wang, L.; Wang, Y.; Sun, X. G.; He, J. Q.; Pan, Z. Y.; Wang, C. H. Finite Element Simulation of
Residual Stress of Double-Ceramic-Layer La2Zr2O7/8YSZ Thermal Barrier Coatings Using Birth
and Death Element Technique. Comput. Mater. Sci. 2012, 53 (1), 117–127.
https://doi.org/10.1016/j.commatsci.2011.09.028.
(37) Bobzin, K.; Lugscheider, E.; Bagcivan, N. Thermal Cycling Behavior of Yttria Stabilized Zirconia and
Lanthanum Zirconate as Graded and Bilayer EB-PVD Thermal Barrier Coatings. High Temp.
Mater. Process. Int. Q. High-Technol. Plasma Process. 2006, 10 (1), 103–116.
https://doi.org/10.1615/HighTempMatProc.v10.i1.80.
(38) Doleker, K. M.; Karaoglanli, A. C.; Ozgurluk, Y.; Kobayashi, A. Performance of Single YSZ,
Gd2Zr2O7 and Double-Layered YSZ/Gd2Zr2O7 Thermal Barrier Coatings in Isothermal Oxidation
Test Conditions. Vacuum 2020, 177, 109401. https://doi.org/10.1016/j.vacuum.2020.109401.
(39) Mahade, S.; Curry, N.; Jonnalagadda, K. P.; Peng, R. L.; Markocsan, N.; Nylén, P. Influence of YSZ
Layer Thickness on the Durability of Gadolinium Zirconate/YSZ Double-Layered Thermal Barrier
Coatings Produced by Suspension Plasma Spray. Surf. Coat. Technol. 2019, 357, 456–465.
https://doi.org/10.1016/j.surfcoat.2018.10.046.
(40) Kawaguchi, A.; Iguma, H.; Yamashita, H.; Takada, N.; Nishikawa, N.; Yamashita, C.; Wakisaka, Y.;
Fukui, K. Thermo-Swing Wall Insulation Technology; - A Novel Heat Loss Reduction Approach on
Engine Combustion Chamber -; 2016; pp 2016-01–2333. https://doi.org/10.4271/2016-01-2333.
(41) Kováčik, J. Correlation between Young’s Modulus and Porosity in Porous Materials. J. Mater. Sci.
Lett. 1999, 18 (13), 1007–1010. https://doi.org/10.1023/A:1006669914946.
(42) Zhao, Y.; Wen, J.; Peyraut, F.; Planche, M.-P.; Misra, S.; Lenoir, B.; Ilavsky, J.; Liao, H.; Montavon,
G. Porous Architecture and Thermal Properties of Thermal Barrier Coatings Deposited by
Suspension Plasma Spray. Surf. Coat. Technol. 2020, 386, 125462.
https://doi.org/10.1016/j.surfcoat.2020.125462.

44
(43) Aranke, O.; Gupta, M.; Markocsan, N.; Li, X.-H.; Kjellman, B. Microstructural Evolution and
Sintering of Suspension Plasma-Sprayed Columnar Thermal Barrier Coatings. J. Therm. Spray
Technol. 2019, 28 (1–2), 198–211. https://doi.org/10.1007/s11666-018-0778-z.
(44) Portinha, A.; Teixeira, V.; Carneiro, J.; Martins, J.; Costa, M. F.; Vassen, R.; Stoever, D.
Characterization of Thermal Barrier Coatings with a Gradient in Porosity. Surf. Coat. Technol.
2005, 195 (2–3), 245–251. https://doi.org/10.1016/j.surfcoat.2004.07.094.
(45) Ahmaniemi, S.; Tuominen, J.; Vippola, M.; Vuoristo, P.; Mäntylä, T.; Cernuschi, F.; Gualco, C.;
Bonadei, A.; Di Maggio, R. Characterization of Modified Thick Thermal Barrier Coatings. J.
Therm. Spray Technol. 2004, 13 (3), 361–369. https://doi.org/10.1361/10599630420371.
(46) Ait Ferhat, Y.; Boulenouar, A. Computation of SIFs for Cracks in FGMs and TBC under Mechanical
and Thermal Loadings. Int. J. Interact. Des. Manuf. IJIDeM 2020, 14 (4), 1347–1356.
https://doi.org/10.1007/s12008-020-00675-8.
(47) Functionally Graded Materials; Miyamoto, Y., Kaysser, W. A., Rabin, B. H., Kawasaki, A., Ford, R.
G., Eds.; Ford, R. G., Series Ed.; Materials Technology Series; Springer US: Boston, MA, 1999; Vol.
5. https://doi.org/10.1007/978-1-4615-5301-4.
(48) Parihar, R. S.; Setti, S. G.; Sahu, R. K. Recent Advances in the Manufacturing Processes of
Functionally Graded Materials: A Review. Sci. Eng. Compos. Mater. 2018, 25 (2), 309–336.
https://doi.org/10.1515/secm-2015-0395.
(49) Sahasrabudhe, H.; Harrison, R.; Carpenter, C.; Bandyopadhyay, A. Stainless Steel to Titanium
Bimetallic Structure Using LENSTM. Addit. Manuf. 2015, 5, 1–8.
https://doi.org/10.1016/j.addma.2014.10.002.
(50) Rangaraj, S.; Kokini, K. Interface Thermal Fracture in Functionally Graded Zirconia–Mullite–Bond
Coat Alloy Thermal Barrier Coatings. Acta Mater. 2003, 51 (1), 251–267.
https://doi.org/10.1016/S1359-6454(02)00396-8.
(51) Tejero-Martin, D.; Rezvani Rad, M.; McDonald, A.; Hussain, T. Beyond Traditional Coatings: A
Review on Thermal-Sprayed Functional and Smart Coatings. J. Therm. Spray Technol. 2019, 28
(4), 598–644. https://doi.org/10.1007/s11666-019-00857-1.
(52) Fauchais, P.; Vardelle, A.; Dussoubs, B. <I>Quo Vadis</I> Thermal Spraying? J. Therm. Spray
Technol. 2001, 10 (1), 44–66. https://doi.org/10.1361/105996301770349510.
(53) Sampath, S.; Schulz, U.; Jarligo, M. O.; Kuroda, S. Processing Science of Advanced Thermal-Barrier
Systems. MRS Bull. 2012, 37 (10), 903–910. https://doi.org/10.1557/mrs.2012.233.
(54) Electron Beam Processes. In Advanced Thermally Assisted Surface Engineering Processes; Springer
US: Boston, MA, 2004; pp 135–147. https://doi.org/10.1007/1-4020-7764-5_4.
(55) Miguel-Pérez, V.; Martínez-Amesti, A.; Nó, M. L.; Calvo-Angós, J.; Arriortua, M. I. EB-PVD
Deposition of Spinel Coatings on Metallic Materials and Silicon Wafers. Int. J. Hydrog. Energy
2014, 39 (28), 15735–15745. https://doi.org/10.1016/j.ijhydene.2014.07.115.
(56) Xu, Z.; Wang, Z.; Huang, G.; Mu, R.; He, L. Morphology, Bond Strength and Thermal Cycling
Behavior of (Ni, Pt)Al/YSZ EB-PVD Thermal Barrier Coatings. J. Alloys Compd. 2015, 651, 445–
453. https://doi.org/10.1016/j.jallcom.2015.08.113.
(57) Soltani, R.; Samadi, H.; Garcia, E.; Coyle, T. W. Development of Alternative Thermal Barrier
Coatings for Diesel Engines; 2005; pp 2005-01–0650. https://doi.org/10.4271/2005-01-0650.
(58) Schulz, U.; Leyens, C.; Fritscher, K.; Peters, M.; Saruhan-Brings, B.; Lavigne, O.; Dorvaux, J.-M.;
Poulain, M.; Mévrel, R.; Caliez, M. Some Recent Trends in Research and Technology of Advanced
Thermal Barrier Coatings. Aerosp. Sci. Technol. 2003, 7 (1), 73–80.
https://doi.org/10.1016/S1270-9638(02)00003-2.
(59) Kaliaraj, G. S.; Bavanilathamuthiah, M.; Kirubaharan, K.; Ramachandran, D.; Dharini, T.;
Viswanathan, K.; Vishwakarma, V. Bio-Inspired YSZ Coated Titanium by EB-PVD for Biomedical

45
Applications. Surf. Coat. Technol. 2016, 307, 227–235.
https://doi.org/10.1016/j.surfcoat.2016.08.039.
(60) Koch, D.; Mauer, G.; Vaßen, R. Manufacturing of Composite Coatings by Atmospheric Plasma
Spraying Using Different Feed-Stock Materials as YSZ and MoSi2. J. Therm. Spray Technol. 2017,
26 (4), 708–716. https://doi.org/10.1007/s11666-017-0537-6.
(61) Yanar, N. M.; Helminiak, M.; Meier, G. H.; Pettit, F. S. Comparison of the Failures during Cyclic
Oxidation of Yttria-Stabilized (7 to 8 Weight Percent) Zirconia Thermal Barrier Coatings
Fabricated via Electron Beam Physical Vapor Deposition and Air Plasma Spray. Metall. Mater.
Trans. A 2011, 42 (4), 905–921. https://doi.org/10.1007/s11661-010-0436-7.
(62) Campo, L. del; De Sousa Meneses, D.; Wittmann-Ténèze, K.; Bacciochini, A.; Denoirjean, A.;
Echegut, P. Effect of Porosity on the Infrared Radiative Properties of Plasma-Sprayed Yttria-
Stabilized Zirconia Ceramic Thermal Barrier Coatings. J. Phys. Chem. C 2014, 118 (25), 13590–
13597. https://doi.org/10.1021/jp5014362.
(63) Fauchais, P.; Vardelle, M.; Vardelle, A.; Goutier, S. What Do We Know, What Are the Current
Limitations of Suspension Plasma Spraying? J. Therm. Spray Technol. 2015, 24 (7), 1120–1129.
https://doi.org/10.1007/s11666-015-0286-3.
(64) Kassner, H.; Siegert, R.; Hathiramani, D.; Vassen, R.; Stoever, D. Application of Suspension Plasma
Spraying (SPS) for Manufacture of Ceramic Coatings. J. Therm. Spray Technol. 2008, 17 (1), 115–
123. https://doi.org/10.1007/s11666-007-9144-2.
(65) Cai, W.; Liu, H.; Sickinger, A.; Muehlberger, E.; Bailey, D.; Lavernia, E. J. Low-Pressure Plasma
Deposition of Tungsten. J. Therm. Spray Technol. 1994, 3 (2), 135–141.
https://doi.org/10.1007/BF02648270.
(66) Salimijazi, H. R.; Coyle, T. W.; Mostaghimi, J. Vacuum Plasma Spraying: A New Concept for
Manufacturing Ti-6Al-4V Structures. JOM 2006, 58 (9), 50–56. https://doi.org/10.1007/s11837-
006-0083-z.
(67) Besling, W. F. A.; Goossens, A.; Meester, B.; Schoonman, J. Laser-Induced Chemical Vapor
Deposition of Nanostructured Silicon Carbonitride Thin Films. J. Appl. Phys. 1998, 83 (1), 544–
553. https://doi.org/10.1063/1.366669.
(68) Mbugua, N. S.; Kang, M.; Zhang, Y.; Ndiithi, N. J.; V. Bertrand, G.; Yao, L. Electrochemical
Deposition of Ni, NiCo Alloy and NiCo–Ceramic Composite Coatings—A Critical Review.
Materials 2020, 13 (16), 3475. https://doi.org/10.3390/ma13163475.
(69) Mišković-Stanković, V. B. Electrophoretic Deposition of Ceramic Coatings on Metal Surfaces. In
Electrodeposition and Surface Finishing: Fundamentals and Applications; Djokić, S. S., Ed.;
Springer New York: New York, NY, 2014; pp 133–216. https://doi.org/10.1007/978-1-4939-
0289-7_3.
(70) Meletis, E. I.; Nie, X.; Wang, F. L.; Jiang, J. C. Electrolytic Plasma Processing for Cleaning and
Metal-Coating of Steel Surfaces. Surf. Coat. Technol. 2002, 150 (2), 246–256.
https://doi.org/10.1016/S0257-8972(01)01521-3.
(71) Hashaikeh, R.; Szpunar, J. A. Electrophoretic Fabrication of Thermal Barrier Coatings. J. Coat.
Technol. Res. 2011, 8 (2), 161–169. https://doi.org/10.1007/s11998-010-9277-y.
(72) Hu, S.; Li, W.; Yao, M.; Li, T.; Liu, X. Electrophoretic Deposition of Gadolinium-Doped Ceria as a
Barrier Layer on Yttrium-Stabilized Zirconia Electrolyte for Solid Oxide Fuel Cells. Fuel Cells 2017,
17 (6), 869–874. https://doi.org/10.1002/fuce.201700122.
(73) Hu, S.; Finklea, H.; Li, W.; Li, W.; Qi, H.; Zhang, N.; Liu, X. Alternating Current Electrophoretic
Deposition of Gadolinium Doped Ceria onto Yttrium Stabilized Zirconia: A Study of the
Mechanism. ACS Appl. Mater. Interfaces 2020, 12 (9), 11126–11134.
https://doi.org/10.1021/acsami.9b17504.

46
(74) Carroll, B. E.; Otis, R. A.; Borgonia, J. P.; Suh, J.; Dillon, R. P.; Shapiro, A. A.; Hofmann, D. C.; Liu, Z.-
K.; Beese, A. M. Functionally Graded Material of 304L Stainless Steel and Inconel 625 Fabricated
by Directed Energy Deposition: Characterization and Thermodynamic Modeling. Acta Mater.
2016, 108, 46–54. https://doi.org/10.1016/j.actamat.2016.02.019.
(75) Gualtieri, T.; Bandyopadhyay, A. Additive Manufacturing of Compositionally Gradient Metal-
Ceramic Structures: Stainless Steel to Vanadium Carbide. Mater. Des. 2018, 139, 419–428.
https://doi.org/10.1016/j.matdes.2017.11.007.
(76) Balla, V. K.; Bandyopadhyay, P. P.; Bose, S.; Bandyopadhyay, A. Compositionally Graded Yttria-
Stabilized Zirconia Coating on Stainless Steel Using Laser Engineered Net Shaping (LENSTM). Scr.
Mater. 2007, 57 (9), 861–864. https://doi.org/10.1016/j.scriptamat.2007.06.055.
(77) Lin, W.-S.; Starr, T. L.; Harris, B. T.; Zandinejad, A.; Morton, D. Additive Manufacturing Technology
(Direct Metal Laser Sintering) as a Novel Approach to Fabricate Functionally Graded Titanium
Implants: Preliminary Investigation of Fabrication Parameters. Int. J. Oral Maxillofac. Implants
2013, 28 (6), 1490–1495. https://doi.org/10.11607/jomi.3164.
(78) Guo, J.; Zeng, Y.; Li, P.; Chen, J. Fine Lattice Structural Titanium Dioxide Ceramic Produced by DLP
3D Printing. Ceram. Int. 2019, 45 (17), 23007–23012.
https://doi.org/10.1016/j.ceramint.2019.07.346.
(79) Torkashvand, K.; Poursaeidi, E.; Mohammadi, M. Effect of TGO Thickness on the Thermal Barrier
Coatings Life under Thermal Shock and Thermal Cycle Loading. Ceram. Int. 2018, 44 (8), 9283–
9293. https://doi.org/10.1016/j.ceramint.2018.02.140.
(80) Ye, D.; Wang, W.; Huang, J.; Lu, X.; Zhou, H. Nondestructive Interface Morphology
Characterization of Thermal Barrier Coatings Using Terahertz Time-Domain Spectroscopy.
Coatings 2019, 9 (2), 89. https://doi.org/10.3390/coatings9020089.
(81) Ogawa, K.; Minkov, D.; Shoji, T.; Sato, M.; Hashimoto, H. NDE of Degradation of Thermal Barrier
Coating by Means of Impedance Spectroscopy. NDT E Int. 1999, 32 (3), 177–185.
https://doi.org/10.1016/S0963-8695(98)00069-3.
(82) Fukuchi, T.; Ozeki, T.; Okada, M.; Fujii, T. Nondestructive Inspection of Thermal Barrier Coating of
Gas Turbine High Temperature Components: NONDESTRUCTIVE INSPECTION OF THERMAL
BARRIER COATING. IEEJ Trans. Electr. Electron. Eng. 2016, 11 (4), 391–400.
https://doi.org/10.1002/tee.22255.
(83) Sayar, M.; Seo, D.; Ogawa, K. Non-Destructive Microwave Detection of Layer Thickness in
Degraded Thermal Barrier Coatings Using K- and W-Band Frequency Range. NDT E Int. 2009, 42
(5), 398–403. https://doi.org/10.1016/j.ndteint.2009.01.003.
(84) Moskal, G.; Swadźba, R.; Witala, B. Non-Destructive Measurement of Top Coat Thickness in TBC
Systems by 3D Optical Topometry. Nondestruct. Test. Eval. 2015, 30 (1), 39–48.
https://doi.org/10.1080/10589759.2014.984297.
(85) Li, Y.; Yan, B.; Li, W.; Li, D. Thickness Assessment of Thermal Barrier Coatings of Aeroengine
Blades Via Dual-Frequency Eddy Current Evaluation. IEEE Magn. Lett. 2016, 7, 1–5.
https://doi.org/10.1109/LMAG.2016.2590465.
(86) Chen, H.-L. R.; Zhang, B.; Alvin, M. A.; Lin, Y. Ultrasonic Detection of Delamination and Material
Characterization of Thermal Barrier Coatings. J. Therm. Spray Technol. 2012, 21 (6), 1184–1194.
https://doi.org/10.1007/s11666-012-9811-9.
(87) Eldridge, J. I.; Singh, J.; Wolfe, D. E. Erosion-Indicating Thermal Barrier Coatings Using
Luminescent Sublayers. J. Am. Ceram. Soc. 2006, 89 (10), 3252–3254.
https://doi.org/10.1111/j.1551-2916.2006.01210.x.
(88) Eldridge, J. I.; Bencic, T. J. Monitoring Delamination of Plasma-Sprayed Thermal Barrier Coatings
by Reflectance-Enhanced Luminescence. Surf. Coat. Technol. 2006, 201 (7), 3926–3930.
https://doi.org/10.1016/j.surfcoat.2006.08.008.

47
(89) Taqi, A. H.; Faraj, K. A.; Zaynal, S. A. The Effect of Long-Term X-Ray Exposure on Human
Lymphocyte. J. Biomed. Phys. Eng. 2019, 9 (1), 127–132.
(90) Davies, G.; Linfield, E. Bridging the Terahertz Gap. Phys. World 2004, 17 (4), 37–41.
https://doi.org/10.1088/2058-7058/17/4/34.
(91) Wright, O. B.; Hyoguchi, T.; Kawashima, K. Laser Picosecond Acoustics in Thin Films: Effect of
Elastic Boundary Conditions on Pulse Generation. Jpn. J. Appl. Phys. 1991, 30 (Part 2, No. 1B),
L131–L133. https://doi.org/10.1143/JJAP.30.L131.
(92) Tao, Y. H.; Fitzgerald, A. J.; Wallace, V. P. Non-Contact, Non-Destructive Testing in Various
Industrial Sectors with Terahertz Technology. Sensors 2020, 20 (3), 712.
https://doi.org/10.3390/s20030712.
(93) Fukuchi, T.; Fuse, N.; Fujii, T.; Okada, M.; Fukunaga, K.; Mizuno, M. Measurement of Topcoat
Thickness of Thermal Barrier Coating for Gas Turbines Using Terahertz Waves. Electr. Eng. Jpn.
2013, 183 (4), 1–9. https://doi.org/10.1002/eej.22385.
(94) Chen, C.-C.; Lee, D.-J.; Pollock, T.; Whitaker, J. F. Terahertz Characterization of Interfacial Oxide
Layers and Voids for Health Monitoring of Ceramic Coatings. In 2009 34th International
Conference on Infrared, Millimeter, and Terahertz Waves; IEEE: Busan, South Korea, 2009; pp 1–
2. https://doi.org/10.1109/ICIMW.2009.5325767.
(95) Dong, J.; Locquet, A.; Citrin, D. S. Terahertz Quantitative Nondestructive Evaluation of Failure
Modes in Polymer-Coated Steel. IEEE J. Sel. Top. Quantum Electron. 2017, 23 (4), 1–7.
https://doi.org/10.1109/JSTQE.2016.2611592.
(96) Kim, D.-H.; Ryu, C.-H.; Park, S.-H.; Kim, H.-S. Nondestructive Evaluation of Hidden Damages in
Glass Fiber Reinforced Plastic by Using the Terahertz Spectroscopy. Int. J. Precis. Eng. Manuf.-
Green Technol. 2017, 4 (2), 211–219. https://doi.org/10.1007/s40684-017-0026-x.
(97) Ryu, C.-H.; Park, S.-H.; Kim, D.-H.; Jhang, K.-Y.; Kim, H.-S. Nondestructive Evaluation of Hidden
Multi-Delamination in a Glass-Fiber-Reinforced Plastic Composite Using Terahertz Spectroscopy.
Compos. Struct. 2016, 156, 338–347. https://doi.org/10.1016/j.compstruct.2015.09.055.
(98) Park, S.-H.; Jang, J.-W.; Kim, H.-S. Non-Destructive Evaluation of the Hidden Voids in Integrated
Circuit Packages Using Terahertz Time-Domain Spectroscopy. J. Micromechanics
Microengineering 2015, 25 (9), 095007. https://doi.org/10.1088/0960-1317/25/9/095007.
(99) Markl, D.; Strobel, A.; Schlossnikl, R.; Bøtker, J.; Bawuah, P.; Ridgway, C.; Rantanen, J.; Rades, T.;
Gane, P.; Peiponen, K.-E.; Zeitler, J. A. Characterisation of Pore Structures of Pharmaceutical
Tablets: A Review. Int. J. Pharm. 2018, 538 (1–2), 188–214.
https://doi.org/10.1016/j.ijpharm.2018.01.017.
(100) Lin, H.; Dong, Y.; Shen, Y.; Axel Zeitler, J. Quantifying Pharmaceutical Film Coating with Optical
Coherence Tomography and Terahertz Pulsed Imaging: An Evaluation. J. Pharm. Sci. 2015, 104
(10), 3377–3385. https://doi.org/10.1002/jps.24535.
(101) Wellman, R. G.; Nicholls, J. R. A Review of the Erosion of Thermal Barrier Coatings. J. Phys. Appl.
Phys. 2007, 40 (16), R293–R305. https://doi.org/10.1088/0022-3727/40/16/R01.
(102) Rogers, J. A.; Fuchs, M.; Banet, M. J.; Hanselman, J. B.; Logan, R.; Nelson, K. A. Optical System for
Rapid Materials Characterization with the Transient Grating Technique: Application to
Nondestructive Evaluation of Thin Films Used in Microelectronics. Appl. Phys. Lett. 1997, 71 (2),
225–227. https://doi.org/10.1063/1.119506.
(103) Khanolkar, A.; Wallen, S.; Abi Ghanem, M.; Jenks, J.; Vogel, N.; Boechler, N. A Self-Assembled
Metamaterial for Lamb Waves. Appl. Phys. Lett. 2015, 107 (7), 071903.
https://doi.org/10.1063/1.4928564.
(104) Hurley, D. H.; Schley, R. S.; Khafizov, M.; Wendt, B. L. Local Measurement of Thermal Conductivity
and Diffusivity. Rev. Sci. Instrum. 2015, 86 (12), 123901. https://doi.org/10.1063/1.4936213.

48
(105) Zhu, W.; Wu, Q.; Yang, L.; Zhou, Y. C. In Situ Characterization of High Temperature Elastic
Modulus and Fracture Toughness in Air Plasma Sprayed Thermal Barrier Coatings under Bending
by Using Digital Image Correlation. Ceram. Int. 2020, 46 (11), 18526–18533.
https://doi.org/10.1016/j.ceramint.2020.04.158.
(106) Nouri, S.; Sahmani, S.; Asayesh, M.; Aghdam, M. M. Microstructural Characterization of YSZ-
CoNiCrAlY Two-Layered Thermal Barrier Coating Formed on γ-TiAl Intermetallic Alloy via APS
Process. Intermetallics 2020, 118, 106704. https://doi.org/10.1016/j.intermet.2020.106704.
(107) Ali, M. S.; Song, S.; Xiao, P. Degradation of Thermal Barrier Coatings Due to Thermal Cycling up to
1150°C. J. Mater. Sci. 2002, 37 (10), 2097–2102. https://doi.org/10.1023/A:1015245920054.
(108) Shi, J.; Zhang, T.; Sun, B.; Wang, B.; Zhang, X.; Song, L. Isothermal Oxidation and TGO Growth
Behavior of NiCoCrAlY-YSZ Thermal Barrier Coatings on a Ni-Based Superalloy. J. Alloys Compd.
2020, 844, 156093. https://doi.org/10.1016/j.jallcom.2020.156093.
(109) Keyvani, A.; Saremi, M.; Sohi, M. H. Oxidation Resistance of YSZ-Alumina Composites Compared
to Normal YSZ TBC Coatings at 1100°C. J. Alloys Compd. 2011, 509 (33), 8370–8377.
https://doi.org/10.1016/j.jallcom.2011.05.029.
(110) Izadinia, M.; Soltani, R.; Sohi, M. H. Effect of Segmented Cracks on TGO Growth and Life of Thick
Thermal Barrier Coating under Isothermal Oxidation Conditions. Ceram. Int. 2020, 46 (6), 7475–
7481. https://doi.org/10.1016/j.ceramint.2019.11.245.
(111) Cai, J.; Lv, P.; Guan, Q.; Xu, X.; Lu, J.; Wang, Z.; Han, Z. Thermal Cycling Behavior of Thermal
Barrier Coatings with MCrAlY Bond Coat Irradiated by High-Current Pulsed Electron Beam. ACS
Appl. Mater. Interfaces 2016, 8 (47), 32541–32556. https://doi.org/10.1021/acsami.6b11129.
(112) Gell, M.; Wang, J.; Kumar, R.; Roth, J.; Jiang, C.; Jordan, E. H. Higher Temperature Thermal Barrier
Coatings with the Combined Use of Yttrium Aluminum Garnet and the Solution Precursor
Plasma Spray Process. J. Therm. Spray Technol. 2018, 27 (4), 543–555.
https://doi.org/10.1007/s11666-018-0701-7.
(113) Rao, H.; Oleksak, R. P.; Favara, K.; Harooni, A.; Dutta, B.; Maurice, D. Behavior of Yttria-Stabilized
Zirconia (YSZ) during Laser Direct Energy Deposition of an Inconel 625-YSZ Cermet. Addit.
Manuf. 2020, 31, 100932. https://doi.org/10.1016/j.addma.2019.100932.
(114) Yonushonis, T. M. Overview of Thermal Barrier Coatings in Diesel Engines. J. Therm. Spray
Technol. 1997, 6 (1), 50–56. https://doi.org/10.1007/BF02646312.
(115) Uczak de Goes, W.; Markocsan, N.; Gupta, M.; Vaßen, R.; Matsushita, T.; Illkova, K. Thermal
Barrier Coatings with Novel Architectures for Diesel Engine Applications. Surf. Coat. Technol.
2020, 396, 125950. https://doi.org/10.1016/j.surfcoat.2020.125950.
(116) Benamati, G.; Chabrol, C.; Perujo, A.; Rigal, E.; Glasbrenner, H. Development of Tritium
Permeation Barriers on Al Base in Europe. J. Nucl. Mater. 1999, 271–272, 391–395.
https://doi.org/10.1016/S0022-3115(98)00792-2.
(117) Hollenberg, G. W.; Simonen, E. P.; Kalinin, G.; Terlain, A. Tritium/Hydrogen Barrier Development.
Fusion Eng. Des. 1995, 28, 190–208. https://doi.org/10.1016/0920-3796(95)90039-X.
(118) Causey, R. A.; Karnesky, R. A.; San Marchi, C. Tritium Barriers and Tritium Diffusion in Fusion
Reactors. In Comprehensive Nuclear Materials; Elsevier, 2012; pp 511–549.
https://doi.org/10.1016/B978-0-08-056033-5.00116-6.
(119) Liu, Q.; Huang, S.; He, A. Composite Ceramics Thermal Barrier Coatings of Yttria Stabilized
Zirconia for Aero-Engines. J. Mater. Sci. Technol. 2019, 35 (12), 2814–2823.
https://doi.org/10.1016/j.jmst.2019.08.003.
(120) Li, B.; Fan, X.; Li, D.; Jiang, P. Design of Thermal Barrier Coatings Thickness for Gas Turbine Blade
Based on Finite Element Analysis. Math. Probl. Eng. 2017, 2017, 1–13.
https://doi.org/10.1155/2017/2147830.

49
(121) Nguyen, C. H.; Chandrashekhara, K.; Birman, V. Multifunctional Thermal Barrier Coating in
Aerospace Sandwich Panels. Mech. Res. Commun. 2012, 39 (1), 35–43.
https://doi.org/10.1016/j.mechrescom.2011.10.003.
(122) Schlichting, K. W.; Padture, N. P.; Klemens, P. G. Thermal Conductivity of Dense and Porous Yttria-
Stabilized Zirconia. J. Mater. Sci. 2001, 36 (12), 3003–3010.
https://doi.org/10.1023/A:1017970924312.
(123) Song, X.; Liu, Z.; Kong, M.; Lin, C.; Huang, L.; Zheng, X.; Zeng, Y. Thermal Stability of Yttria-
Stabilized Zirconia (YSZ) and YSZ-Al2O3 Coatings. Ceram. Int. 2017, 43 (16), 14321–14325.
https://doi.org/10.1016/j.ceramint.2017.07.186.
(124) Nourani-Vatani, M.; Ganjali, M.; Solati-Hashtjin, M.; Zarrintaj, P.; Reza Saeb, M. Zirconium-Based
Hybrid Coatings: A Versatile Strategy for Biomedical Engineering Applications. Mater. Today
Proc. 2018, 5 (7), 15524–15531. https://doi.org/10.1016/j.matpr.2018.04.159.
(125) Caputo, S.; Millo, F.; Boccardo, G.; Piano, A.; Cifali, G.; Pesce, F. C. Numerical and Experimental
Investigation of a Piston Thermal Barrier Coating for an Automotive Diesel Engine Application.
Appl. Therm. Eng. 2019, 162, 114233. https://doi.org/10.1016/j.applthermaleng.2019.114233.
(126) Sivakumar, G.; Senthil Kumar, S. Investigation on Effect of Yttria Stabilized Zirconia Coated Piston
Crown on Performance and Emission Characteristics of a Diesel Engine. Alex. Eng. J. 2014, 53
(4), 787–794. https://doi.org/10.1016/j.aej.2014.08.003.
(127) Uchida, N.; Osada, H. A New Piston Insulation Concept for Heavy-Duty Diesel Engines to Reduce
Heat Loss from the Wall. SAE Int. J. Engines 2017, 10 (5), 2565–2574.
https://doi.org/10.4271/2017-24-0161.
(128) L. Salvalaio, C.; P. Silva, F.; S. Pinho, A.; Pohlmann, M. Qualitative Evaluation of Physical Effort in
Bass Drum Pedal Drive by Thermography. Sci. Technol. 2012, 1 (1), 1–6.
https://doi.org/10.5923/j.scit.20110101.01.
(129) Burns, D.; Johnson, S. Nuclear Thermal Propulsion Reactor Materials. In Nuclear Materials
[Working Title]; IntechOpen, 2020. https://doi.org/10.5772/intechopen.91016.
(130) Cureton, W. F.; Zillinger, J.; Rosales, J.; Wilkerson, R. P.; Lang, M.; Barnes, M. Microstructural
Evolution of Mo-UO2 Cermets under High Temperature Hydrogen Environments. J. Nucl. Mater.
2020, 538, 152297. https://doi.org/10.1016/j.jnucmat.2020.152297.
(131) Schmid, W.; Dirndorfer, S.; Juranowitsch, H.; Kress, M.; Petry, W. Adhesion Strength of Sputter
Deposited Diffusion Barrier Layer Coatings for the Use in U–Mo Nuclear Fuels. Nucl. Eng. Des.
2014, 276, 115–123. https://doi.org/10.1016/j.nucengdes.2014.05.025.
(132) Dharini, T.; Kuppusami, P.; Panda, P.; Ramaseshan, R.; Kirubaharan, A. M. K. Nanomechanical
Behaviour of Ni – YSZ Nanocomposite Coatings on Superalloy 690 as Diffusion Barrier Coatings
for Nuclear Applications. Ceram. Int. 2020, 46 (15), 24183–24193.
https://doi.org/10.1016/j.ceramint.2020.06.198.
(133) Samantaroy, P. K.; Girija, S.; Mudali, U. K. Effect of Heat Treatment on Corrosion Behavior of Alloy
690 and Alloy 693 in Simulated Nuclear High-Level Waste Medium. Corrosion 2012, 68 (4),
046001-1-046001–046013. https://doi.org/10.5006/0010-9312-68-4-7.
(134) Sengupta, P.; Rogalla, D.; Becker, H. W.; Dey, G. K.; Chakraborty, S. Development of Graded Ni–
YSZ Composite Coating on Alloy 690 by Pulsed Laser Deposition Technique to Reduce Hazardous
Metallic Nuclear Waste Inventory. J. Hazard. Mater. 2011, 192 (1), 208–221.
https://doi.org/10.1016/j.jhazmat.2011.05.006.
(135) Johnson, R. N. Coatings for Fast Breeder Reactor Components. Thin Solid Films 1984, 118 (1), 31–
48. https://doi.org/10.1016/0040-6090(84)90104-4.
(136) Ferré, F. G.; Mairov, A.; Iadicicco, D.; Vanazzi, M.; Bassini, S.; Utili, M.; Tarantino, M.; Bragaglia,
M.; Lamastra, F. R.; Nanni, F.; Ceseracciu, L.; Serruys, Y.; Trocellier, P.; Beck, L.; Sridharan, K.;

50
Beghi, M. G.; Di Fonzo, F. Corrosion and Radiation Resistant Nanoceramic Coatings for Lead Fast
Reactors. Corros. Sci. 2017, 124, 80–92. https://doi.org/10.1016/j.corsci.2017.05.011.
(137) Kasada, R.; Dou, P. Sol–Gel Composite Coatings as Anti-Corrosion Barrier for Structural Materials
of Lead–Bismuth Eutectic Cooled Fast Reactor. J. Nucl. Mater. 2013, 440 (1–3), 647–653.
https://doi.org/10.1016/j.jnucmat.2013.06.014.
(138) Shukla, P. K.; Hemanth Rao, E.; Vetrivendan, E.; Anne, R. S.; Das, S. K.; Malarvizhi, B.; Ponraju, D.;
Nashine, B. K.; Kamachi Mudali, U.; Selvaraj, P. Evaluation of Plasma Sprayed Sacrificial Thermal
Barrier Coatings for Core Catcher of Future Sodium Cooled Fast Reactors. Ann. Nucl. Energy
2017, 107, 31–36. https://doi.org/10.1016/j.anucene.2017.04.015.
(139) Zheng, G.; Carpenter, D.; Dolan, K.; Hu, L. Experimental Investigation of Alumina Coating as
Tritium Permeation Barrier for Molten Salt Nuclear Reactors. Nucl. Eng. Des. 2019, 353, 110232.
https://doi.org/10.1016/j.nucengdes.2019.110232.
(140) Lee, J. J.; Arregui-Mena, J. D.; Contescu, C. I.; Burchell, T. D.; Katoh, Y.; Loyalka, S. K. Protection of
Graphite from Salt and Gas Permeation in Molten Salt Reactors. J. Nucl. Mater. 2020, 534,
152119. https://doi.org/10.1016/j.jnucmat.2020.152119.
(141) Olson, L. C.; Ambrosek, J. W.; Sridharan, K.; Anderson, M. H.; Allen, T. R. Materials Corrosion in
Molten LiF–NaF–KF Salt. J. Fluor. Chem. 2009, 130 (1), 67–73.
https://doi.org/10.1016/j.jfluchem.2008.05.008.
(142) McNeese, L. Molten-Salt Reactor Program. Semiannual Progress Report for Period Ending
February 29, 1976; Oak Ridge National Lab., Tenn.(USA), 1976.
(143) Bernardet, V.; Gomes, S.; Delpeux, S.; Dubois, M.; Guérin, K.; Avignant, D.; Renaudin, G.; Duclaux,
L. Protection of Nuclear Graphite toward Fluoride Molten Salt by Glassy Carbon Deposit. J. Nucl.
Mater. 2009, 384 (3), 292–302. https://doi.org/10.1016/j.jnucmat.2008.11.032.
(144) He, X.; Song, J.; Tan, J.; Zhang, B.; Xia, H.; He, Z.; Zhou, X.; Zhao, M.; Liu, X.; Xu, L.; Bai, S. SiC
Coating: An Alternative for the Protection of Nuclear Graphite from Liquid Fluoride Salt. J. Nucl.
Mater. 2014, 448 (1–3), 1–3. https://doi.org/10.1016/j.jnucmat.2014.01.034.
(145) Luo, L.-M.; Liu, Y.-L.; Liu, D.-G.; Zheng, L.; Wu, Y.-C. Preparation Technologies and Performance
Studies of Tritium Permeation Barriers for Future Nuclear Fusion Reactors. Surf. Coat. Technol.
2020, 403, 126301. https://doi.org/10.1016/j.surfcoat.2020.126301.
(146) Wu, Y.; He, D.; Li, S.; Liu, X.; Wang, S.; Jiang, L. Deuterium Permeation Properties of Y2O3/Cr2O3
Composite Coating Prepared by MOCVD on 316L Stainless Steel. Int. J. Hydrog. Energy 2016, 41
(18), 7425–7430. https://doi.org/10.1016/j.ijhydene.2016.03.132.
(147) Wang, L.; Wu, Y. Y.; Luo, X. F.; Ning, Z. E.; Wang, J. H.; Yang, J. J.; Feng, Y. J.; Liao, J. L.; Yang, Y. Y.;
Feng, K. M.; Liu, N.; Gong, M. Effects of Ar/O2 Ratio on Preparation and Properties of Multilayer
Cr2O3/α-Al2O3 Tritium Permeation Barrier. Surf. Coat. Technol. 2018, 339, 132–138.
https://doi.org/10.1016/j.surfcoat.2018.02.021.
(148) Bazzanella, N.; Checchetto, R.; Miotello, A.; Patton, B.; Kale, A. N.; Kothari, D. C. High
Temperature Efficient Deuterium Permeation and Oxidation (Al,Ti)N Barriers Deposited on
Stainless Steel. Appl. Phys. Lett. 2002, 81 (20), 3762–3764. https://doi.org/10.1063/1.1521576.
(149) Leushake, U.; Krell, T.; Schulz, U. Graded Thermal Barrier Coating Systems for gas turbine
applications. Mater. Werkst. 1997, 28 (8), 391–394.
https://doi.org/10.1002/mawe.19970280817.

51

You might also like