You are on page 1of 18

Case Studies in Thermal Engineering 52 (2023) 103727

Contents lists available at ScienceDirect

Case Studies in Thermal Engineering


journal homepage: www.elsevier.com/locate/csite

Effects of thermal barrier coating thickness and surface roughness


on cooling performance sensitivity of a turbine blade
Zequn Du a, b, Haiwang Li a, b, Ruquan You a, b, *, Yi Huang a, b
a
Research Institute of Aero-Engine, Beihang University, Beijing, China
b
National Key Laboratory of Science and Technology on Aero Engines Aero-thermodynamics, Beihang University, Beijing, China

A R T I C L E I N F O A B S T R A C T

Keywords: In this study, the conjugate heat transfer method is used to research the impact of the thermal
Thermal barrier coating thickness barrier coating thickness (0–300 μm) and surface roughness (Ra = 0–30 μm) on the cooling
Surface roughness characteristics of the high-pressure turbine blade. A physical thin-walled structure is applied to
Conjugate heat transfer simulate the TBC on the turbine blade surface. The results indicate that the temperature distri­
Cooling performance sensitivity bution on the coated metal surface is more uniform. With an increase in coating thickness, the
Turbine blades
heat flux of the blade surface decreases, reaching a maximum of 106 W/m2 at the leading edge.
Furthermore, leading-edge temperature is the most sensitive to changes in coating thickness, and
the pressure surface is the least sensitive. The overall cooling effectiveness is less sensitive to
changes in the roughness after coating. Increasing the surface roughness, the heat flux reduces at
the leading edge and increases at other places. The range of the overall cooling effectiveness rate
of the blade rises, and the sensitivity of the overall cooling effectiveness to the coating thickness
increases. When Ra reaches 20 μm, the insulation effect with different coating thickness remains
unaffected.

Nomenclature

Bi Biot number of metal


Bitbc Biot number of TBC
cp,dd6 Specific heat capacity of solid [J/kg⋅K]
dtbc Thickness of TBC [m]
d Thickness of mental [m]
Df Diameter of film hole [m]
Df t Diameter of outlet hole [m]
Di1 Diameter of impingement hole at the leading edge [m]
Di2 Diameter of impingement hole at the trailing edge [m]
Dp Diameter of pin fin [m]
h Heat transfer coefficient at the metal surface [W/m2 ⋅K]
he Heat transfer coefficient within the thermal boundary of high-temperature gas side [W/m2 ⋅K]
hc Heat transfer coefficient within the thermal boundary of the coolant side [W/m2 ⋅K]

* Corresponding author. Research Institute of Aero-Engine, Beihang University, Beijing, China.


E-mail address: youruquan10353@buaa.edu.cn (R. You).

https://doi.org/10.1016/j.csite.2023.103727
Received 3 June 2023; Received in revised form 21 October 2023; Accepted 3 November 2023
Available online 4 November 2023
2214-157X/© 2023 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
Z. Du et al. Case Studies in Thermal Engineering 52 (2023) 103727

h′ Heat transfer coefficient at TBC surface [W/m2 ⋅K]


Δh Heat transfer coefficient change [W/m2 ⋅K]
ks Equivalent sand grain roughness height [m]
Pin Total pressure of inlet mainstream [Pa]
Pout Static pressure of outlet mainstream [Pa]
q Heat flux at the interface of fluid and solid [W /m2 ]
Ra Surface roughness [m]
Tin Total temperature of inlet mainstream [K]
T∞ Temperature of inlet mainstream [K]
Tc Temperature of inlet cooling airflow [K]
Te,conv Driving temperature of mainstream [K]
Ti,conv Driving temperature of cooling airflow [K]
Tw Metal wall temperature without TBC [K]
T′w Metal wall temperature with TBC [K]
T′ Temperature of TBC surface [K]
ΔT Insulation effect of TBC [K]

Greek Symbols
λtbc Thermal conductivity of TBC [W/m⋅K]
λ Thermal conductivity of metal [W/m⋅K]
φ Overall cooling effectiveness without TBC
φ′ Overall cooling effectiveness with TBC
φ′smooth Overall cooling effectiveness with TBC on smooth surface
φ′rough Overall cooling effectiveness with TBC on rough surface
Δφ Overall cooling effectiveness change rate
τ Coating effectiveness
α Dimensionless driving temperature of the internal heat transfer process
η Dimensionless driving temperature of the external heat transfer process

Abbreviation
TBC Thermal barrier coating
CFD Computational fluid dynamics
CHT Conjugate heat transfer

1. Introduction
The thrust-to-weight ratio is an important index to measure engine performance. To optimize the engine performance with a
smaller size and lighter weight, the turbine inlet temperature is well above the melting point of Ni-base superalloy. In this case, the
cooling technologies cannot keep the blades working effectively. Therefore, the blade surface must be coated with thermal barrier
coating (TBC). The application of TBC can increase the thermal resistance of the blade and provide thermal insulation. TBC consists of a
ceramic layer and a bonding layer, providing superior resistance to high temperatures, corrosion, and abrasion. Therefore, more
studies have focused on TBCs and their applications [1–3].
The insulation effect of TBC has been studied extensively. Studies showed that applying TBC with a thickness of 0.25 mm could
reduce substrate surface temperature by up to 127K. This effect was equivalent to the sum of the development of superalloys to
improve the temperature capacity over the past 30 years [4–7]. Rossette et al. [8] conducted conjugate heat transfer (CHT) calculations
for high-pressure turbine blades without film holes. They presented that when the TBC thickness was 0.4 mm, the surface temperature
could be reduced by up to 200 K, resulting in a 36 % reduction in cooling air usage and an increase in turbine efficiency. Liu et al. [9]
used a one-dimensional heat transfer model to derive the heat insulation effect with dimensionless parameters. They concluded that
the thermal insulation properties might be enhanced by increasing TBC thickness or reducing the TBC’s thermal conductivity. Chen
et al. [10] employed numerical simulations to explore the impact of coating thickness on the cooling effectiveness of a plate with film
holes. The results showed that using a thicker TBC led to a more uniform distribution of cooling effectiveness.
Although TBC theoretically has a good thermal insulation effect, it doesn’t perform as well as expected in practical applications.
The structure’s thermal expansion coefficient is different from the substrate’s. Influenced by the film cooling, the temperature
downstream of the film holes is low. The surface temperature is not uniform, and there are significant thermal stresses. These stresses
may cause the coating to fall off. During the coating process, the film holes are blocked. This changes the size and angle of the film
holes, which affects the overall cooling effectiveness of the blade. Meanwhile, working in a severe environment, dust deposits on the
coating surface, and sand grains wear down the blade [11–15]. The impact of film hole blockage caused by coating on the adiabatic
cooling effectiveness of plates under various blowing ratios was experimentally investigated by Bunker et al. [16]. Sundaram et al. [17]

2
Z. Du et al. Case Studies in Thermal Engineering 52 (2023) 103727

conducted a large-scale turbine vane cascade experiment to explore the impact of TBC spallation. The study showed that TBC spall­
ation decreased the efficiency at the leading edge by 10 %. Wear and peeling of coatings change the surface roughness of blades and
affect the flow heat transfer characteristics [18–22]. Stripf et al. [23] and So et al. [24,25] experimentally evaluated the impact of
roughness on the heat transfer distribution of the turbine blade. They concluded that compared with a smooth surface, the change in
roughness would lead to a forward transition point and a 50 % increase in the heat transfer coefficient within the turbulent boundary
layer. Prapa et al. [26] employed numerical calculations of coupled heat transfer to study the comprehensive impact of roughness and
TBCs on cooling effectiveness. The results showed that increasing surface roughness enhanced the heat transfer, which was harmful to
blade cooling but improved the thermal insulation of TBC. Without TBC, the overall cooling effectiveness variation range with
roughness was − 8 %–4 %. Conversely, the variation range was 0–30 % with TBC. Rutledge et al. [27] experimentally investigated the
effect of roughness on the cooling characteristics at the suction surface of the blade. Their results indicated that compared with smooth
surfaces, rough surfaces could increase surface heat flux by 30 %–70 %.
Over the years, researchers have conducted numerous experimental and numerical studies on the impacts of roughness and TBCs on
the cooling characteristics of turbine blades. However, there are complex heat transfer processes in the blade, and the above pa­
rameters cannot be changed separately during the operation of real blades. The coating thickness and surface roughness will interact
with each other, leading to the nonlinear effects. Therefore, it is necessary to closely combine boundary conditions with real working
conditions to further investigate the combined effect of roughness and coating thickness. Meanwhile, most of the previous numerical
calculations used a simplified method to imitate TBC on the blade [28–32]. A thin wall with virtual thickness outside the blade was set
in CFX to simulate the equivalent thermal resistance. Only the flow and temperature parameters on the substrate’s surface can be
obtained through this method. It is challenging to get the TBC’s thermal insulation and heat transfer characteristics. The effect of
coatings on the flow in the cascade passage is ignored. After coating, the pressure inside the passage changes, and the mass flow
distribution of the film changes accordingly. Therefore, it has great limitations. In this study, a thin wall of true thickness was proposed
to simulate TBC structures on turbine blade surfaces. It is convenient to obtain the heat transfer on the TBC surface and heat conduction
inside the TBC in the heat transfer processes. The results are more detailed than those of previous research, which provides the
possibility to comprehensively analyze the complex cooling processes of turbine blades. And the simulation method used in this paper
effectively takes into account the influence of coating on the flow, and the calculation results are more accurate, which is of stronger
value for the optimal design of blades in engineering.
In this paper, the theoretical derivation of a one-dimensional heat transfer model with TBC was carried out, and the influence
mechanism of TBC thickness and surface roughness on the heat transfer process was investigated. Computational fluid dynamics (CFD)
and conjugate heat transfer (CHT) methods were used to numerically calculate a simplified model of a high-pressure turbine blade with
multiple exhaust film holes. The influence of TBC thickness and surface roughness on the cooling characteristics of blades was
investigated, and the difference between the calculated results and the one-dimensional deduced results were compared with the
effects of lateral thermal conduction and various cooling structures. The sensitivity of the overall cooling effectiveness and insulating
effect of TBC under the combined effects of coating thickness and roughness was studied. It is of significant engineering value.

2. Theoretical and numerical models


2.1. A one-dimensional theoretical heat transfer model with TBC
The thickness of TBC in turbine blades is very small, and the thickness direction is substantially smaller than the other directions.

Fig. 1. One-dimensional heat transfer model.

3
Z. Du et al. Case Studies in Thermal Engineering 52 (2023) 103727

Therefore, the temperature of turbine blades with TBC can be approximately simulated by a one-dimensional model. Fig. 1 shows the
one-dimensional theoretical heat transfer model. The heat transfer process includes thermal convection in the mainstream side and the
internal cooling channel, as well as heat conduction of the TBC and blade. According to Fourier’s law of heat conduction, the rela­
tionship of the whole heat transfer process can be established as Eq. (1). φ′ can be obtained:
T∞ − Tc
q= 1 (1)
h′
+ dλtbc
tbc
+ dλ + h1c

T∞ − Tw′ 1 + Bitbc
φ′ = = (α − η) + η (2)
T∞ − Tc 1 + Bitbc + Bi + hhec

where h′ is the overall heat transfer coefficient between high-temperature gas and TBC surface. Bi and Bitbc are the Biot number of blade
and TBC respectively. In this paper, the heat transfer process is considered to occur in the thermal boundary layer. Two dimensionless
driving temperatures α and η are defined for the internal and external heat transfer processes, as shown in Eqs. (3) and (4).
T∞ − Ti,conv
α= (3)
T∞ − Tc

T∞ − Te,conv
η= (4)
T∞ − Tc

where Te,conv and Ti,conv are the driving temperatures of mainstream and coolant sides within the thermal boundary layer. T∞ , Tc , and T′w
are the temperatures of the inlet mainstream, inlet cooling airflow, and the substrate after coated, respectively.
According to Eq. (2), increasing the TBC thickness can effectively improve the overall cooling effectiveness of the blade. For a one-
dimensional heat transfer model without film cooling, increasing the surface roughness will reduce the overall cooling effectiveness.
However, there are complex cooling structures in blades, and the cooling effects are influenced by the combination of various cooling
methods, such as film cooling, TBC heat insulation, etc. Therefore, the influence of TBC thickness and surface roughness is no longer
simple. TBC has a certain lateral thermal conductivity, while the thermal conductivity of the substrate is relatively large. The lateral
thermal conductivity effect is more obvious. The one-dimensional steady-state heat transfer formula based on Fourier heat conduction
law is no longer applicable. Therefore, this paper proposed the three-dimensional turbine blade model and compared the difference
between the calculated results and the one-dimensional deduced results.

2.2. Multi-field coupling calculation model


2.2.1. Blade model
Fig. 2 shows the blade model. The original cascade is a circumferential layout with 66 rotor blades, and the three-channel structure
is used for internal cooling. The cooling channel includes impingement holes, 90◦ straight ribs, and pin fins. Table 1 displays the
geometrical features of the film holes. There are 14 rows with 326 film holes. The leading edge and pressure surface each have six
exhaust film holes, and the suction surface and the tip each have one exhaust film hole. The spanwise angle of film holes is 60◦ . There
are four outlet holes at the tip of the blade each with a diameter (Dft ) of 0.6 mm. The diameters of impingement holes at the leading
edge and the trailing edge are 1.2 mm and 1 mm, respectively. The TBC thickness was set to 0, 50, 100, 150, 200, 250, and 300 μm.
Turbine blades can operate effectively in high-temperature and high-pressure environments through composite cooling methods of
external film cooling, TBC insulation, internal impingement cooling, and reinforcement heat transfer by pin fin and rib.

Fig. 2. (a) Blade model, (b) internal cooling passage, (c) positions of film holes.

4
Z. Du et al. Case Studies in Thermal Engineering 52 (2023) 103727

Table 1
Geometric parameters of film holes.

Row Number of Holes Position Hole Diameter Df (mm) Flow Angle (◦ )

1 27 Suction surface 0.33 60


2 28 Leading edge 0.32 60
3 23 Leading edge 0.30 60
4 32 Leading edge 0.40 60
5 32 Leading edge 0.40 90
6 31 Leading edge 0.40 90
7 28 Leading edge 0.33 90
8 9 Pressure surface 0.38 60
9 23 Pressure surface 0.58 60
10 18 Pressure surface 0.33 60
11 18 Pressure surface 0.33/0.46 60
12 20 Pressure surface 0.38/0.48 60
13 15 Pressure surface 0.38 60
14 22 Blade tip 0.33 60

2.2.2. Numerical computation method


The numerical calculation method of coupled heat transfer was adopted, and the mesh was divided by Fluent Meshing. Considering
the complexity of the model, the fluid and solid domains were divided into unstructured grids, as shown in Fig. 3. To guarantee that the
mesh at the interface had common nodes and prevent the interpolation error in the calculation process, the “Join” function was
employed in the mesh construction of the fluid and solid domains. Due to the complexity of the mixing process of cold and hot fluids
near the outlets of film holes, mesh encryption was carried out at small structures such as film holes, tip gaps, and TBC. A boundary
layer mesh was added to the end walls of the main channel and blade surface to improve the solving accuracy in the viscous boundary
layer. The mesh height of the first layer was 0.0015 mm, and the growth rate was 1.2. There were 15 layers of boundary mesh.
Uncoated blades were selected for a grid independence test, and five types of meshes were adopted, which were 23 million, 31
million, 38 million, 53 million, and 82 million. Fig. 4 shows that the temperature changes along the blade surface at midspan under
different meshes. When the mesh number is lower than 38 million, the surface temperature changes with mesh numbers. The local
temperature difference is up to 40 K. Increasing the mesh number, the temperature along the blade surface is almost constant.
Considering the calculation time and accuracy, the final mesh number was 38 million.
The numerical calculations need to be consistent with the continuity, momentum and energy equations. The equations are shown
below.
∂ρ
+∇•(ρu)= 0 (5)
∂t

Fig. 3. (a) Solid domain, (b) fluid domain, (c) mesh at tip clearance, (d) mesh at midspan.

5
Z. Du et al. Case Studies in Thermal Engineering 52 (2023) 103727

Fig. 4. Temperature distributions of five meshes at midspan.

∂ρu
+∇ • {ρu ⊗ u}= ∇ • {τ − ρu′ ⊗ u′} + SM (6)
∂t

∂ρHtotal ∂ρ
+∇ • {ρuHtotal + ρu′H − k∇τ} = + SE (7)
∂t ∂t

where u′ is the pulsation velocity. And SM and SE are the source terms.
The steady-state solution was performed using CFX as the solver. Shear stress transport (SST) and γ-θ transition models were the
selected turbulence models. Ho [33] compared the numerical calculation results of turbine blades with experimental data for four
turbulence models (SA, RNG, k-ε, and SST). He presented that the SST model was the most suitable for the fluid-thermo-solid coupling
calculation of blades with film cooling. The flow conservation equation used a high-resolution discretization scheme with second-order
accuracy. To make the calculation easier to converge, the time factor was adjusted appropriately. The criterion of the convergence
calculation is that the residuals of the momentum and energy equations were less than 10− 4 . Meanwhile, the average temperature of
the coating surface was monitored. When the temperature stabilized, the calculation was considered to be convergent.

2.2.3. Boundary conditions


The fluid domain was chosen as an ideal gas. The heat capacity and thermal conductivity were calculated according to molecular
theory. The viscosity coefficient followed Sutherland’s law. Using mass flow rate and total temperature, the cool inlet was regulated.
The total temperature, total pressure, and velocity direction were given to the main inlet. The main flow outlet was characterized by
static pressure. The variation rules of each parameter along the radial direction were shown in Fig. 5. The turbulence of the mainstream
and cold air was set as 5 %, and the boundary conditions in the fluid domain were listed in Table 2.
The blade and TBC were elements of the solid domain. Considering the thinness of the bonding layer, TBC was simplified to a
ceramic layer structure. The material was zirconia (ZrO2 ), and the physical property parameters were constant. The specific heat
capacity and thermal conductivity were 502 J/(kg⋅K) and 1.02 W/(m⋅K), respectively. DD6 was selected as the blade material, and the
variation of specific heat capacity and thermal conductivity with temperature followed Eqs. (8) and (9).

cp,dd6 = − 3.4965×10− 7 T 2 +0.34689×T+227.7[J / (kg⋅K)] (8)

Fig. 5. Distribution along the radial direction of (a) total temperature (T) of inlet mainstream, (b) total pressure of inlet mainstream (Pin), (c) static pressure of outlet
mainstream (Pout).

6
Z. Du et al. Case Studies in Thermal Engineering 52 (2023) 103727

Table 2
Boundary conditions of the fluid domain.

Parameters Value Units

Total temperature of inlet mainstream Tin K


Total pressure of inlet mainstream Pin MPa
Mass flow of inlet cooling airflow 0.08 kg/s
Total temperature of inlet cooling airflow 868 K
Static pressure of outlet mainstream Pout MPa
Rotation speed 13181 rpm

λ= − 7.0907×10− 9 T 3 +2.1458×10− 5 T 2 +1.5867×10− 3 T+4.62[W / (m ⋅ K)] (9)


The roughness values for the turbine blades at various places were reported by Bogard et al. [13]. The surface roughness varied
from 8.3 to 46.2 μm. The corresponding relationship between the equivalent sand grain roughness height (ks ) and surface roughness
parameter (Ra) was established as ks ≈ 4Ra. In this study, Ra was set to 0 (smooth), 8 μm, 20 μm, and 30 μm, while ks was set to
0 (smooth), 32 μm, 80 μm, and 120 μm to investigate the effects of roughness on cooling characteristics. The walls on both sides of the
fluid were rotationally periodic. The fluid-solid coupling surface was set as the interface for heat exchange. The upper wall in the fluid
domain was set to the counter-rotating wall, which was equivalent to the speed of the turbine blade in the opposite direction to
simulate the turbine engine in real working conditions. The rest of the walls were set up as adiabatic non-slip walls.

2.2.4. Thermal parameters


In this paper, the separate and combined effects of TBC thickness and surface roughness on blade cooling characteristics were
evaluated by defining heat transfer parameters including overall cooling effectiveness, coating effectiveness, and heat transfer coef­
ficient. According to the CHT method, overall cooling efficiency represents the combined effect of the heat conduction of TBC and
blade, and the thermal convection within the thermal boundary layer. Coating effectiveness defines the thermal insulation perfor­
mance of TBC. The heat exchange between the surface of the blade and the mainstream is measured using heat transfer coefficient.
(1) Overall cooling effectiveness φ:
T∞ − Tw
φ= (10)
T∞ − Tc

where T∞ , Tc , and Tw are the temperatures of the inlet mainstream, inlet cooling airflow, and metal surface, respectively.
For coated blades, Eq. (10) becomes:

T∞ − Tw′
φ′ = (11)
T∞ − Tc

where T′w represents the metal temperature of the coated blade.

(2) Coating effectiveness τ:

T∞ − T′
τ= (12)
T∞ − Tc

where T′ represents the temperature of TBC surface.


(3) Heat transfer coefficient h:
q
h= (13)
T∞ − Tw

where q represents the heat flux on the metal blade’s surface. Heat is transmitted from the mainstream to the metal surface when q is
positive. A negative value, however, denotes heat being transported away from the metal surface.
For coated blades, Eq. (13) becomes
q
h′ = (14)
T∞ − T′

3. Results and discussions


3.1. Effects of TBC thickness
In this section, the influence of TBC thickness variation (d = 0–300 μm) on blade cooling performance was investigated for a smooth
surface.
The surface temperature distribution of the metal substrate with various TBC thickness is shown in Fig. 6. It is similar at different
thickness. The hot zones are located along the central leading edge, the tip, and the first and second rows of holes at the suction surface.

7
Z. Du et al. Case Studies in Thermal Engineering 52 (2023) 103727

The area enclosed by the film holes has a low temperature. Along the height direction, the temperature of the metal surface increases
gradually. Because the temperature at the inlet of the mainstream tends to be high in the middle and low at both ends. The inlet
temperature at the middle of the leading edge is relatively high, and it is directly impacted by high-temperature gas. Affected by the
pressure gradient, the high-temperature gas leaks into the suction surface side from the pressure surface through the tip gap. The total
pressure at the outlet hole at the tip of the blade is not large, and the flow rate is relatively low. Moreover, the cold air conducts
convective heat transfer in the inner cold channel, increasing the coolant’s temperature and decreasing the cooling capacity. There
were a few film holes at the suction surface, which limited the cooling capacity of the film. As the TBC’s thickness increases, the surface
temperature of the substrate reduces, and the high-temperature region shrinks. The temperature decreases faster and wider, especially
in the high-temperature region covered by film. The temperature at the tip also decreases, but the amplitude and range of the reduction
are relatively small. In the middle of the pressure surface, the cold air flow is large, and the cooling effect is better. Therefore, the
surface temperature there is lower and it is less sensitive to TBC thickness. When the TBC thickness increases from 0 to 300 μm, the
average temperature of the substrate decreases from 1158K to 1044 K. The temperature at A, B, and C decreases from 964K to 919K,
1428K–1384K, and 1292K–1163K, respectively. TBC increases the thermal conductivity resistance so that the heat transfer is greatly
reduced. This results in a lower temperature difference and a more uniform temperature distribution on the blade surface.
Fig. 7 shows the temperature distribution of the TBC surface under various coating thickness. On the whole, the rule of temperature
distribution is basically the same with different coating thickness. The high-temperature zone is mainly located in the leading edge and
tip of the blade. Due to the high total pressure of coolant at the root, the flow rate is large, and the cooling capacity is strong. Complex
vortexes at the adjacent blade pressure surface enter the middle of the suction surface, so there is a low-temperature zone at the root of
the pressure surface and the middle of the suction surface root. As the coating thickness increases, the TBC surface temperature steadily
rises and the high-temperature area expands, especially at the leading edge. A stagnation occurs at the leading edge where the heat flux
is large, and the heat transfer is intense, as can be seen in Fig. 10. As TBC thickness rises, the heat flux transferred from the TBC to the
substrate gradually decreases, and more heat is isolated on the coating surface, raising the temperature at the leading edge. When the
coolant flows through the film hole, there is convective heat transfer in the hole, raising the temperature of the coolant. According to
Eq. (2), the insulating effect increases with TBC thickness. The temperature of the internal cooling channel decreases, which weakens
the heat transfer to the coolant and lowers the coolant temperature. Therefore, the film coverage effect near the film hole exit is better
and the coolant and local temperature are lower. When the coating thickness increases from 50 μm to 300 μm, the average temperature
of the TBC increases from 1175 K to 1230 K. The temperature at D decreases from 964 K to 924 K, and the temperature at E increases
from 1289 K to 1438 K. Compared with the substrate, TBC has a lower thermal conductivity, so the lateral heat conduction of the
surface is weaker, resulting in a lower temperature in the area covered by the film and a significantly higher temperature in the area
without film. The temperature distribution on the TBC surface varies widely. However, the low thermal conductivity of the TBC in­
hibits the inward transfer of its surface temperature difference. The effect of film cooling on the metal temperature is weakened, and
the temperature is more uniform.
Fig. 8 shows the structural temperature distributions at midspan with different thickness. There is a single row of film holes on the
suction surface, and the cooling capacity is limited, while there are many film holes on the pressure surface. As a result, the suction
surface has a higher temperature than the pressure surface. The temperature at the leading edge is also high. As the thickness increases,
the thermal resistance increases, hindering the heat transfer from the mainstream to the blade. The coating temperature increases, and
the thermal insulation performance improves. The metal temperature decreases significantly, and the temperature gradient decreases
accordingly. The temperature of the cooling channel is greatly reduced, which weakens the convective heat transfer from the inner
wall to the cold air and improves the cooling performance of the cold air.

Fig. 6. Surface temperature distributions of blades with different thickness.

8
Z. Du et al. Case Studies in Thermal Engineering 52 (2023) 103727

Fig. 7. Surface temperature distributions of TBCs with different thickness.

Fig. 8. Structural temperature distributions at midspan with different thickness.

Fig. 9 shows the distribution of overall cooling effectiveness at midspan with different coating thickness. l/L is the non-dimensional
streamwise position, which is equal to 0 at the stagnation point, negative at the pressure surface, and positive at the suction surface.
The distribution of overall cooling effectiveness (φ′) along the chord length is similar for different coating thickness. With an increase in
coating thickness, φ′ gradually increases. The average φ′ increases by 25.4 %, and the local φ′ increases to a maximum of 79.4 %. When
the thickness of TBC increases to 200 μm, φ′ increases slowly, and the insulation efficiency per unit thickness of the TBC decreases from
0.83 K/μm to 0.57 K/μm. The overall cooling effectiveness at the leading edge is more sensitive to the presence or absence of the TBC,
the suction surface is the second most sensitive, and the pressure surface is the least sensitive. The TBC thickness has less impact on φ′
near the film holes. This is mainly because the wall temperature is low whereas the local cooling effectiveness is high, and the tem­
perature difference is small. Therefore, the insulation effect of the TBC is not obvious.
Fig. 10 shows the distribution of heat flux at midspan. Because of the low thermal conductivity of TBC, the thermal conduction
resistance is enhanced. After coating, the heat flux decreases. It is more challenging for the mainstream to transfer heat to the blade.

9
Z. Du et al. Case Studies in Thermal Engineering 52 (2023) 103727

Fig. 9. Distributions of overall cooling effectiveness at midspan with different thickness.

Meanwhile, the heat flux close to the film holes and the position with high cooling effectiveness is rarely affected by the coating
thickness. The TBC thickness has the greatest impact on the heat flux at the leading edge. As TBC thickness increases, the change rate of
heat flux decreases gradually. Eq. (15) can be obtained based on Fourier’s law of heat conduction. Due to the extremely thin thickness
of TBC, its effect on the flow in the blade cascade is assumed almost negligible. Therefore, h1e + dλ + h1c can be approximated as constant.
For a particular position, q can be seen as an inverse proportion function of dtbc . As can be seen from Eq. (15), the larger Te,conv is, the
more sensitive q is to the change of dtbc . Stagnation occurs at the leading edge and heat transfer is enhanced. Te,conv is large there, so dtbc
has a significant impact on q. The changing rate of q slows down as dtbc rises and q’ gradually declines.
Te,conv − Ti,conv
q= (15)
1
+ dλtbc
he
tbc
+ dλ + h1c

Where he is the heat transfer coefficient in the thermal boundary layer of the high-temperature gas side. It is worth noting that Fourier’s
law of heat conduction applied here is only used for qualitative analysis, not for quantitative analysis, limited by the transverse heat
conduction in the blade.
Figs. 11 and 12 indicate how the TBC thickness affects the metal substrate’s average temperature and thermal insulation effect,
respectively. With the increase of TBC thickness, the temperature at the leading edge drops the most, and the thermal insulation effect
is the best. The average temperature of the pressure surface changes in proportion to that of the suction surface, and a similar trend is
observed with the thermal insulation effect. Without TBC, the average temperature of the leading edge is the highest, while the
pressure surface is the lowest. When TBC thickness reaches 50 μm, the leading edge’s average temperature decreases by 122.69 K,
lower than that of the suction surface. As shown in Fig. 10, the heat flux at the leading edge decreases, reaching a maximum of

Fig. 10. Distributions of heat flux at midspan with different thickness.

10
Z. Du et al. Case Studies in Thermal Engineering 52 (2023) 103727

Fig. 11. The average temperature of the metal substrate varies with different thickness.

Fig. 12. The insulation effect of the TBC varies with different thickness.

approximately 106 W/m2 . This makes the average temperature of the leading edge more sensitive to changes in coating thickness.
When the TBC thickness is 150 μm, the thermal insulation effect of the TBC at the leading edge reaches 169.94 K. When the TBC
thickness is 300 μm, the thermal insulation effect reaches 258.4 K at the leading edge, while that of the pressure surface and suction
surface is relatively low at approximately 180 K. The growth rate of the insulation effect slows down as the coating thickness increases.

Fig. 13. Surface temperature distribution of blade with different roughness (a) Ra = 0 μm(b) Ra = 8 μm(c) Ra = 20 μm(d) Ra = 30 μm.

11
Z. Du et al. Case Studies in Thermal Engineering 52 (2023) 103727

3.2. Effects of surface roughness


In this section, the influence of surface roughness variation (Ra = 0–30 μm) on the cooling performance was investigated when the
TBC thickness was 100 μm.
Fig. 13 shows the surface temperature distribution of the substrate with different roughness. The metal surface temperature shows a
general increasing trend as the roughness increases, and the temperature gradient rises along the radial direction. In contrast, the
temperature and gradient decreased in the radial direction at the leading edge. Because the increase in surface roughness strengthens
the overall heat transfer from the mainstream to the surface. This increases the heat flux on the blade surface, causing the metal surface
temperature to rise. When the coolant moves through the cooling channel, convective heat transfer takes place, and the temperature
rises along the radial direction. So, the cooling capacity decreases, and the temperature gradient increases. The opposite applies at the
leading edge. Increasing roughness causes a decrease in heat flux and surface temperature. The temperature gradient of the cold air in
the radial direction decreases, which decreases that at the leading edge. More detailed reasons for the decrease in temperature at the
leading edge were explained later.
The temperature distribution of the TBC surface for various roughness is shown in Fig. 14. Compared with the metal temperature,
the TBC surface temperature is far more sensitive to changes in roughness. When the surface changes from smooth to Ra = 30 μm, the
average temperature of the substrate rises from 1099.46 K to 1158.85 K, and that of TBC rises from 1176.58 K to 1255.03 K. Once the
coolant exits from the film hole, the film with low temperature develops near the wall. The mainstream is isolated from the surface, and
the local cooling effectiveness is high. The change of TBC surface temperature with roughness is relatively low. Roughness promotes
the mixing of mainstream and coolant, which reduces the film’s effective coverage area. As the roughness increases, the surface
temperature of TBC rises in the area lacking film coverage at the pressure and suction surfaces. The heat flux decreases in the middle
and lower part of the leading edge, leading to a drop in the TBC’s surface temperature. Therefore, it performs more uniformly.
It is worth noting that as roughness increases, the wall temperature decreases at the leading edge and the gradient gradually
decreases along the radial direction, as shown in Fig. 13 and Fig. 14. To analyze the reason for this, the heat transfer coefficient (he )
distribution of the TBC surface is investigated with roughness at the leading edge, as plotted in Fig. 15. Based on the assumption that
the heat transfer at the gas side occurs within the thermal boundary layer, he is defined as follow.
q
he = (16)
Te,conv − T′

It can be seen that he decreases significantly with the increase of Ra. Due to the dense arrangement of film holes at the leading edge
and the fast coolant flow rate, the jet separates from the wall. The film interacts with each other, and the outflow is almost invisible.
The increase in roughness enhances the turbulent pulsation near the wall, and strengthens the spanwise mixing and diffusion between
the coolant and the mainstream. This helps to increase the quantity of the coolant near the wall, distributing it more uniformly. It
prevents heat transfer from the hot gas to the wall and reduces the heat flux.
To further investigate the impact of roughness on the cooling properties of coated blades, the overall cooling effectiveness change
φ′smooth − φ′rough
rate (Δφ) is defined and calculated as Δφ = φ′smooth
. A positive value indicates that the roughness strengthens the heat transfer
between the blade and high-temperature gas and reduces the overall cooling effectiveness, while a negative value indicates that the
overall cooling effectiveness is improved. Fig. 16 shows the distribution of Δφ on the blade’s surface at various roughness. The result
shows that negative Δφ mainly exists at the leading edge. A large value of Δφ is located at the bottom of the pressure surface, middle
and upper parts of the suction surface and trailing edge. With the increase in roughness, the metal temperature difference gradually
grows, and Δφ significantly increases Overall, the pressure surface’s cooling effectiveness variation range is narrower than the suction
surface’s. When Ra increases from 8 to 30 μm, the maximum of Δφ changes from 0.16 to 0.28 at the pressure surface, and the cooling
effectiveness decreases by 14 %. The maximum of Δφ changes from − 0.08 to − 0.2 at the leading edge, and the cooling effectiveness
increases by 11 %. The maximum of Δφ changes from 0.24 to 0.32 at the suction surface, and the cooling effectiveness decreases by 11
%.
Fig. 17 compares coated blades with uncoated blades for changes in the heat transfer coefficient under various roughness. Δh = h −
h′ was used to define the heat transfer coefficient change (Δh). Positive Δh indicates that TBC has a good heat insulation effect and
reduces heat transport to the blade. The result shows that Δh is negative at the leading edge, near the film hole outlet, and the trailing
edge of the top of the pressure surface. TBC prevents heat transfer from the blade to the coolant. With the increase in surface roughness,
Δh increases, and the negative range dramatically shrinks at the leading edge. Additionally, the overall insulation effect of the TBC is
enhanced. However, the negative value range of Δh near the outlet of film holes expands. This is mainly because roughness enhances
the turbulent pulsation near the wall, and strengthens the heat transfer between the blade and mainstream. The surface temperature
increases, making the TBC’s heat insulation effect more noticeable. In contrast, near the film hole outlet, roughness increases the
turbulence intensity, and exacerbates the mixing of the mainstream and cold air.
The structural temperature distributions at midspan with different roughness are shown in Fig. 18. The pressure surface’s tem­
perature near the trailing edge is low, because the film holes are densely arranged there and the cold air flow is large. There are few film
holes on the suction surface, so the overall temperature is high. With the increase in roughness, the temperature of the local low-
temperature region rises at the pressure surface. The temperature distribution performs more uniformly. The suction surface tem­
perature increases significantly, and the range of the high-temperature area expands greatly.
Figs. 19 and 20 show the distribution of the overall cooling effectiveness (φ′), coating effectiveness (τ), and heat flux at midspan

12
Z. Du et al. Case Studies in Thermal Engineering 52 (2023) 103727

Fig. 14. Surface temperature distribution of TBCs with different roughness (a) Ra = 0 μm(b) Ra = 8 μm(c) Ra = 20 μm(d) Ra = 30 μm.

Fig. 15. Heat transfer coefficient (he ) distribution with different roughness (a) Ra = 0 μm(b) Ra = 8 μm(c) Ra = 20 μm(d) Ra = 30 μm.

Fig. 16. Overall cooling effectiveness changes with different roughness (a) Ra = 8 μm (b) Ra = 20 μm (c) Ra = 30 μm.

Fig. 17. Surface temperature distribution of blade with different roughness.

13
Z. Du et al. Case Studies in Thermal Engineering 52 (2023) 103727

Fig. 18. Structural temperature distributions at midspan with different roughness.

Fig. 19. Distribution of overall cooling effectiveness and coating effectiveness at midspan with different roughness.

Fig. 20. Distributions of heat flux at midspan with different roughness.

with different roughness. Generally, due to the low thermal conductivity of TBC, the surface temperature of substrate is typically lower
than that of TBC. Thus, overall cooling effectiveness is higher than coating effectiveness. But it is not the same for all regions. The
transverse heat conduction effect of TBC surface is weak, resulting in a low temperature in the area covered with film. However, the
blade’s thermal conductivity is rather high, and there is transverse heat conduction. Therefore, there are areas where the TBC tem­
perature is lower than the blade temperature, and the local φ′ is lower than τ. The effect of roughness is relatively small around the film

14
Z. Du et al. Case Studies in Thermal Engineering 52 (2023) 103727

hole. With the increase in roughness, both φ′ and τ decrease, the change rate gradually decreases, and the change range of τ is higher
than φ′. Because increasing roughness directly enhances the heat exchange on the mainstream side and raises the temperature of TBC.
The temperature of the substrate is indirectly increased by the heat conduction of TBC. However, the increase in the TBC temperature is
higher than that of the blade surface. With the change in roughness, the heat flux decreases gradually at the leading edge, while the
pressure surface and suction surface are exactly opposite. When Ra reaches 20 μm, the heat flux remains unaffected with different
roughness.
Figs. 21 and 22 show the variation of average metal substrate temperature and TBC insulation effect with roughness, respectively.
The results show that the average temperature of the substrate is higher at the suction surface than that of the leading edge for each
roughness. The thermal insulation effect is greater at the leading edge than at the suction surface. And the pressure surface is the least
affected. As roughness rises, both the metal surface temperature and the insulation effect of TBC increase, but the growth rate decreases
gradually. The average temperature change range of the metal is greater at the suction surface than at the pressure surface, and it is
smallest at the leading edge. The insulation effect at the leading edge has the largest variation range, and the variation ranges of the
suction and pressure surfaces are similar. When the roughness increases from 0 to 30 μm, the average temperature at the leading edge
only increases by 25.6 K, and the heat insulation effect increases by 36.3 %. In contrast, the suction surface with fewer film holes has
poor cooling effectiveness. The average temperature of the metal surface is more sensitive to the roughness change, and increases by
73.99 K.

3.3. Comprehensive effect


The combined influence of TBC thickness and roughness on the overall cooling effectiveness and insulation effect of TBC was
analyzed. The results are shown in Figs. 23 and 24.
When the surface is uncoated, the metal’s surface temperature is directly affected by the mainstream. Surface roughness has a large
effect on the overall cooling effectiveness, as shown in Fig. 23. The influence range of roughness variation on φ is − 65.92 %, and the
sensitivity of φ to roughness is about − 2.2 (%/μm). When the surface changes from smooth to roughness of 8 μm and 30 μm, φ de­
creases by 42.83 % and 65.92 %, respectively. However, for coated blades, roughness affects the TBC’s surface temperature distri­
bution. And the metal temperature is indirectly affected by the thermal conductivity of TBC with high thermal resistance. Therefore,
the effect of roughness on φ′ is significantly reduced after coating. With the increase in TBC thickness, the sensitivity of φ′ to the
roughness is significantly reduced. When the surface changes from smooth to rough (Ra = 30 μm), φ′ of the blades with TBC thickness
of 100 μm and 300 μm decreases by 16.21 % and 7.90 %, respectively. When the TBC thickness is 300 μm, the influence range of
roughness change on φ′ is − 7.9 %, and the sensitivity of φ′ to roughness is about − 0.26 (%/μm).
When the surface changes from smooth to rough, the turbulent pulsation near the wall significantly increases, and the heat transfer
from the mainstream to the blade is enhanced. Therefore, when Ra increases from smooth to 8 μm, the overall cooling effectiveness
decreases significantly, and the attenuation rate slows down as the surface roughness continues to increase. The greater the surface
roughness is, the more sensitive the overall cooling effectiveness is to the change in the TBC thickness. Because the larger the roughness
is, the stronger the heat transfer from the mainstream to the walls. This leads to an increase in the difference between TBC surface and
the cold air, making the thermal insulation of TBC more and more important. When coating thickness increases from 0 to 300 μm, the
overall cooling effectiveness of the smooth surface increases by 31.15 %, and the rough surface with Ra = 30 μm increases by one time.
For smooth surface, the effect of the TBC thickness on the overall cooling effectiveness is 16 %. The sensitivity of the overall cooling
effectiveness to the change of unit coating thickness is about 5.3 (%/100 μm). When Ra = 30 μm, the effect of the TBC thickness on the
overall cooling effectiveness is 30 %, and the sensitivity of the overall cooling effectiveness to the change of unit coating thickness is
about 10 (%/100 μm).

Fig. 21. Average temperature of metal substrate varies with different roughness.

15
Z. Du et al. Case Studies in Thermal Engineering 52 (2023) 103727

Fig. 22. Insulation effect of TBC varies with different roughness.

Fig. 23. Effect of TBC thickness and surface roughness on overall cooling effectiveness.

Fig. 24. Effect of TBC thickness and surface roughness on insulation effect of TBC.

16
Z. Du et al. Case Studies in Thermal Engineering 52 (2023) 103727

As shown in Fig. 24, TBC’s thermal insulation effect increases by 66.83 % when Ra increases from 0 to 8 μm. As the roughness
continues to increase, TBC’s thermal insulation effect gradually slows down. When Ra reaches 20 μm, the insulation effect with
different coating thickness remains unaffected.

4. Conclusion
In this study, a thin wall of true thickness was constructed to simulate TBC on the surface of turbine blades. CFD and CHT methods
were employed to numerically calculate the simplified model of a high-pressure turbine blade with multiple exhaust film holes. The
effects of the TBC thickness (d = 0–300 μm) and surface roughness (Ra = 0–30 μm) on the cooling performance sensitivity of a turbine
blade were studied, and the following conclusions were obtained.
(1) Applying TBC on the blade surface reduces the heat flux. The effect of film cooling on the metal temperature is weakened, and
the temperature is more uniform. There is a large variation in the surface temperature distribution on the TBC. The temperature
is low in the area covered by the film, and high in the area without film.
(2) As the TBC thickness increases, the metal temperature decreases, while the TBC temperature rises. The change rate of the
temperature slows down. The heat flux decreases at the leading edge, reaching a maximum of approximately 106 W/ m2 .
(3) The leading-edge temperature is the most sensitive to changes in the TBC thickness, and the pressure surface is the least sen­
sitive. When the TBC thickness increases from 0 to 300 μm, the average temperature of the substrate decreases by 114 K, and the
thermal insulation effect increases to 186 K.
(4) The temperature of TBC is much more sensitive to roughness than that of the substrate. Roughness has less impact on the
temperature close to the film hole’s exit. The sensitivity of metal temperature to roughness is the highest at the suction surface,
and the least at the leading edge.
(5) With an increase in roughness, the temperature gradient and heat flux at the leading edge both drop, and overall cooling
effectiveness rises. At the pressure surface and suction surface, the heat flux increases, and the cooling effectiveness decreases
significantly.
(6) When the surface is uncoated, the overall cooling effectiveness is greatly affected by the roughness. With an increase in TBC
thickness, the sensitivity of the overall cooling effectiveness to the roughness reduces. When the surface changes from smooth to
rough (Ra = 30 μm), the overall cooling effectiveness of the blades with TBC thickness of 100 μm and 300 μm decreases by 16.21
% and 7.90 %, respectively.
(7) The larger the roughness is, the more sensitive the overall cooling effectiveness is to the change in the TBC thickness. When Ra
reaches 20 μm, the insulation effect with different coating thickness remains unaffected.

Author statement
Zequn Du: Conceptualization, Methodology, Software, Validation, Formal analysis, Investigation, Resources, Data Curation,
Writing-Original Draft, Writing-Review & Editing, Visualization. Haiwang Li: Supervision, Funding acquisition. Ruquan You:Project
administration Yi Huang: Project administration.

Declaration of competing interest


The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Data availability

No data was used for the research described in the article.

Acknowledgments
The present work was financially supported by the Fundamental Research Funds for the Central Universities [No. YWF-22-BJ-J-
822], and National Science and Technology Major Project [2017-III-0003-0027].

References
[1] D.R. Clarke, M. Oechsner, N.P.J.M.b. Padture, Thermal-barrier coatings for more efficient gas-turbine engines 37 (2012) 891–898.
[2] Z. Meng, Y. Liu, Y. Li, X.J.A.T.E. He, The performance evaluation for thermal protection of turbine vane with film cooling and thermal barrier coating, Appl.
Thermal Eng. 210 (2022), 118405.
[3] Y. Şöhret, E. Açıkkalp, A. Hepbasli, T.H.J.E. Karakoc, Advanced exergy analysis of an aircraft gas turbine engine: splitting exergy destructions into parts, Energy
90 (2015) 1219–1228.
[4] R.A. Miller, C.E. Lowell, Failure mechanisms of thermal barrier coatings exposed to elevated temperatures, Thin Solid Films 95 (1982) 265–273.
[5] K.D. Sheffler, D.K. Gupta, Current Status and Future Trends in Turbine Application of Thermal Barrier Coatings, 1988.
[6] S.M. Meier, D. Gupta, The Evolution of Thermal Barrier Coatings in Gas Turbine Engine Applications, 1994.
[7] N.P. Padture, M. Gell, E.H.J.S. Jordan, Thermal barrier coatings for gas-turbine engine applications 296 (2002) 280–284.
[8] A. Hernández Rossette, Z. Mazur C, R. Muñ oz, E. Bautista, Start-up and shut-down simulation of a thermal barrier coated gas turbine bucket using CFD code, in:
ASME Power Conference, vol. 48329, 2008, pp. 75–83.

17
Z. Du et al. Case Studies in Thermal Engineering 52 (2023) 103727

[9] Z. Liu, W. Zhu, L. Yang, Y. Zhou, Numerical prediction of thermal insulation performance and stress distribution of thermal barrier coatings coated on a turbine
vane, Int. J. Thermal Sci. 158 (2020), 106552.
[10] X. Chen, J. Li, Y. Long, Y. Wang, S. Weng, S. Yavuzkurt, A conjugate heat transfer and thermal stress analysis of film-cooled superalloy with thermal barrier
coating, in: Turbo Expo: Power for Land, Sea, and Air, vol. 84171, American Society of Mechanical Engineers, 2020. V07BT12A074.
[11] F. Nemdili, A. Azzi, B.A. Jubran, F. Flow, Numerical investigation of the influence of a hole imperfection on film cooling effectiveness, Numeric. Investig.
Influence Hole Imperfect. Film Cool. Effect. 21 (2011) 46–60.
[12] R.A. Wenglarz, Deposition, erosion and corrosion protection for coal-fired gas turbines, in: Turbo Expo: Power for Land, Sea, and Air, vol. 79436, Citeseer, 1985.
V002T003A004.
[13] D.G. Bogard, D.L. Schmidt, M. Tabbita, Characterization and Laboratory Simulation of Turbine Airfoil Surface Roughness and Associated Heat Transfer,
American Society of Mechanical Engineers, 1996.
[14] K. Tian, J. Wang, C. Liu, L. Yang, B.J.E. Sundén, Effect of blockage configuration on film cooling with and without mist injection 153 (2018) 661–670.
[15] X. Chen, Y. Wang, Y. Long, S. Weng, Effect of partial blockage on flow and heat transfer of film cooling with cylindrical and fan-shaped holes, Int. J. Thermal Sci.
164 (2021), 106866.
[16] R.S. Bunker, Effect of partial coating blockage on film cooling effectiveness, in: Turbo Expo: Power for Land, Sea, and Air, vol. 78569, American Society of
Mechanical Engineers, 2000. V003T001A051.
[17] N. Sundaram, K. Thole, Effects of Surface Deposition, Hole Blockage, and Thermal Barrier Coating Spallation on Vane Endwall Film Cooling, 2007.
[18] F. Tarada, M. Suzuki, External heat transfer enhancement to turbine blading due to surface roughness, in: Turbo Expo: Power for Land, Sea, and Air, vol. 78897,
American Society of Mechanical Engineers, 1993. V002T008A006.
[19] R.P. Taylor, Surface Roughness Measurements on Gas Turbine Blades, 1990.
[20] J.P. Bons, R.P. Taylor, S.T. McClain, R.B. Rivir, The many faces of turbine surface roughness, in: Turbo Expo: Power for Land, Sea, and Air, vol. 78521, American
Society of Mechanical Engineers, 2001. V003T001A042.
[21] J.P. Bons, A Review of Surface Roughness Effects in Gas Turbines, 2010.
[22] M. Morini, M. Pinelli, P.R. Spina, M. Venturini, Numerical Analysis of the Effects of Nonuniform Surface Roughness on Compressor Stage Performance, 2011.
[23] M. Stripf, A. Schulz, S.J.J.T. Wittig, Surface roughness effects on external heat transfer of a HP turbine vane 127 (2005) 200–208.
[24] R.P. Somawardhana, D.G. Bogard, Effects of surface roughness and near hole obstructions on film cooling effectiveness, in: Turbo Expo: Power for Land, Sea, and
Air, vol. 47934, 2007, pp. 799–809.
[25] R.P. Somawardhana, D.G. Bogard, Effects of Obstructions and Surface Roughness on Film Cooling Effectiveness with and without a Transverse Trench, 2009.
[26] P. Prapamonthon, B. Yin, G. Yang, M. Zhang, E. Applications, Separate and combined effects of surface roughness and thermal barrier coating on vane cooling
performance, J. Therm. Sci. Eng. Appl. 12 (2020), 051017.
[27] J.L. Rutledge, D. Robertson, D.G. Bogard, Degradation of film cooling performance on a turbine vane suction side due to surface roughness, in: Turbo Expo:
Power for Land, Sea, and Air, vol. 47268, 2005, pp. 879–887.
[28] P. Prapamonthon, S. Yooyen, S. Sleesongsom, CHT/CFD Analysis of Thermal Sensitivity of a Transonic Film-Cooled Guide Vane, Comput. Model. Eng. Sci. 119
(2019) 593–615.
[29] P. Prapamonthon, H. Xu, W. Yang, J.J.E. Wang, Numerical study of the effects of thermal barrier coating and turbulence intensity on cooling performances of a
nozzle guide vane, 10, 2017, p. 362.
[30] P. Prapamonthon, S. Yooyen, S. Sleesongsom, D. Dipasquale, H. Xu, J. Wang, Z.J.E. Ke, Investigation of cooling performances of a non-film-cooled turbine vane
coated with a thermal barrier coating using conjugate heat transfer, 11, 2018, p. 1000.
[31] T. Takahashi, K. Watanabe, T. Fujii, T. Fujioka, Numerical analysis of temperature distribution of a film-cooled and TBC coated blade, in: Turbo Expo: Power for
Land, Sea, and Air, vol. 43147, 2008, pp. 691–702.
[32] S. Khumhaeng, T. Suksa, N. Laohalertchai, B. Chaiprasit, T. Chotroongruang, P. Prapamonthon, Numerical prediction of thermomechanical sensitivity of the
first stage nozzle guide vane with film cooling, in: ASME Turbo Expo 2022: Turbomachinery Technical Conference and Exposition, Heat Transfer — General
Interest/Additive Manufacturing Impacts on Heat Transfer; Internal Air Systems, 6B, Internal Cooling, 2022.
[33] K.-S. Ho, J.S. Liu, T. Elliott, B. Aguilar, Conjugate heat transfer analysis for gas turbine film-cooled blade, in: Turbo Expo: Power for Land, Sea, and Air, vol.
49781, American Society of Mechanical Engineers, 2016. V05AT10A003.

18

You might also like