Sie sind auf Seite 1von 13

Proceedings of ASME Turbo Expo 2015: Turbine Technical Conference and Exposition

GT2015
June 15 19, 2015, Montral, Canada

GT2015-42205
A PREDICTIVE MODEL FOR PRELIMINARY GAS TURBINE BLADE COOLING ANALYSIS
Nafiz H.K. Chowdhury, Hootan Zirakzadeh and Je-Chin Han
Turbine Heat Transfer Laboratory
Mechanical Engineering Department
Texas A&M University
College Station, TX 77843-3123
Email: jc-han@tamu.edu

allow higher TIT. On the other side, increasing the amount of


bleed air from the compressor to keep the TIT at the same level
results in lower efficiency of the compressor. Thus, a trade-off
study of the turbine blade heat transfer is essential.
As TIT in gas turbines have arisen, first stage vane and blade
cooling schemes have typically migrated toward more efficient
designs. However, with the introduction of combined cycles for
power generation, new types of efficient cooling technologies
are required to answer the potential demand. The best indicator
for a gas turbine cooling system design performance is the
temperature distribution around the blades. Since estimation of
the temperatures includes simultaneous analysis of external hot
gas flow, conduction through metal, and internal coolant flow,
therefore, this situation should treated as a conjugate heat
transfer (CHT) problem. Numerous heat transfer studies have
shown that CHT is a complex phenomenon which requires a
detailed 3D investigation. Generally, external heat transfer is
affected in the spanwise and chordwise directions by the nature
of oncoming flow field including inlet temperature profile, and
pressure distribution. Internal heat transfer depends on a coolant
flow characteristics which are highly influenced by the geometry
and shape of the cooling passage and rib turbulators.
In the last few decades, numerous studies have been
conducted on gas turbine heat transfer by many researchers
utilizing numerical or experimental approaches. Han et al. [1]
may have been the first who published their predictive computer
model in the open literature using energy and mass conservation
method to evaluate gas turbine blade cooling performance.
Later, Hylton et al. [2] employed a finite element solution of the
2D Laplacian heat conduction equation to calculate the heat
transfer coefficient for their vane surfaces using the measured
external temperature and internal cooling duct heat transfer
coefficient as boundary conditions. Additional input to their

ABSTRACT
The growing trend to achieve a higher Turbine Inlet
Temperature (TIT) in the modern gas turbine industry requires,
in return, a more efficient and advanced cooling system design.
Therefore, a complete study of heat transfer is necessary to
predict the thermal loadings in the turbine vane/blade. To
estimate the metal temperatures, it is important to simulate the
external hot gas flow condition, the conduction in the blade
material, and the internal coolant flow characteristics accurately
and simultaneously. As a result, proposing novel, quicker, and
more convenient ways to study the heat transfer behavior of gas
turbine blades is of absolute necessity. In the current work, a
predictive model for the gas turbine blade cooling analysis in the
form of a computer program has been developed to answer this
need. The program is capable of estimating distribution of
coolant mass flow rate, internal pressure and metal temperature
of a turbine blade based on external and internal boundary
conditions. The simultaneous solutions result from the coupled
equations of mass and energy balance. The model is validated by
showing its accuracy to predict the temperature distributions of
a NASA E3 blade with an uncertainty of less than +/-10%. Later,
this paper documents the overall analysis for a set of different
boundary conditions with the same blade model (E3) and
demonstrates the capability of the program to extend for other
cases as well.
INTRODUCTION
Gas turbine industries are always competing to reach higher
Turbine Inlet Temperature (TIT) in order to achieve higher
efficiency. However, this TIT range is restricted by the material
limitations such as the yielding temperature of the blade.
Therefore, it turns out to be a challenge to increase the blade
cooling efficiency with the same amount of bleed air that might

Copyright 2015 by ASME

program was 2D vane cross sectional geometry along with


material thermal conductivity, inlet total temperature and
average coolant temperature. However, their vane was a
simplified version without film cooling holes and used a simple
energy balance equation. Zecchi et al. [3] developed a quick
simulation tool where they implemented the conjugate heat
transfer to calculate metal temperature distributions. In their
work, they analyzed a specific case of the NASA C3X blade both
qualitatively and quantitatively. However, their study was on a
2D model of the blade that included only simple cooling ducts
without any film cooling holes.
Most recently, Mohammad Alizadeh et al. [4] predicted the
blade temperature distribution using a 1D coolant network code
along with CFD prediction by utilizing CHT. They also validated
their data using NASA C3X vane. In their work, they
investigated some parameters and claimed to find the most and
least critical parameters that influenced the blade temperature.
Again, they did not consider any film cooling in the design.
Luo and Razinsky [5] used their multiple analytical codes to
determine conjugate heat transfer effects on a realistic filmcooled turbine vane. They concluded that the V2F model could
capture the differences in the heat transfer rates and metal
temperature under different flow conditions very well. However,
they evaluated the model only by flat plate validation. Takahashi
et al. [6] also studied CHT analysis of a rotor blade where they
blended the 3D steady-state numerical analysis with 1D thermoflow calculation of internal cooling to estimate metal
temperature.
There are some studies [7-14] that have used 3D modeling
of vanes or blades where they employed 3D solvers to apply
CHT method for calculation of the metal temperature
distribution with more complex internal cooling passages. Rigby
and Lepicovsky [15] also used their 3D CHT CFD code called
Glenn-HT to analyze the internally cooled configurations and
compared the findings with the experimental results. Their code
can be modified with the advancement of cooling techniques if
necessary. Later, Downs and Landis [16] conducted a
comprehensive study on turbine cooling system designs where
they discussed the cooling parameters systematically.
Additionally, they identified some limitations that must be
overcome to advance the turbine cooling schemes in the future.
In order to achieve a more accurate solution of complex
CHT problems, generating a structured mesh is preferred by
many CFD experts utilizing commercial or in-house codes.
Creating such type of grids could be very burdensome and timeconsuming depending on the geometry. The aforementioned
point surfaces an urge to pursue a simple model that is easy to
use and can aid designers of the gas turbine blades to grasp an
understanding of the cooling system, quickly, for the preliminary
stage of design. On the other hand, the comprehensive analytical
study of gas turbine blade heat transfer utilizing a flexible tool
with inclusion of sophisticated details of internal and external
cooling is surprisingly limited in the open literature. Therefore,
in this scientific investigation, a step is taken toward providing
and validating such an interactive and simple model in the form

of a computer code to evaluate the cooling design performance


and extend its application and merits to gas turbine designers.
MODEL DESCRIPTION

Different types of modern gas turbine cooling techniques


have been used in various engines over the last few decades as
shown in Figure 1.

Figure 1. Blade Cooling Terminology [17]


External cooling of the turbine blade is accomplished by
film cooling and applying thermal barrier coating (TBC) on its
external surface. Generally, film cooling involves the
introduction of secondary flows into the hot mainstream that are
injected out through the holes drilled into the internal cooling
passages. The aim is to form an insulating film layer that reduces
the heat load from the hot mainstream. Internal cooling can be
carried out by flowing the coolant in the shaped internal cooling
passages. Later, the cooling effect can be maximized by
introducing impingement cooling in the leading edge (LE)
region and pin-fin cooling near the trailing edge (TE) region.
However, rest of the passages is rib roughened to enhance the
heat transfer.
To begin with the present approach, the model is
approximated as a 1D network of finite element systems
connected in the coolant flow direction. According to this
convention, the choice of element numbers in the chord direction
(x) is restricted by the size and number of the internal coolant
passages where as in the span direction (y) is arbitrary. An
example of E3 blade [18] has been shown in Figure 2 to
demonstrate the model concept. This blade has eight internal
cooling ducts where first four (started from LE) are connected in
a single cooling circuit called forward loop and the rests are
connected using backward loop. The blade will necessarily be
divided at the centerline of any duct wall (as shown with red
dashed lines in Figure 2) so that each element is as wide as the
coolant duct associated plus half of each ducts side wall.
The choice of dimension of the element in the span direction
is fully independent and is a compromise between increased
accuracy and computation time. Figure 3 is representing eight
elements in the spanwise and eight elements in the chordwise

Copyright 2015 by ASME

direction that result in a total of 64 elements to perform the


computations on. At each element, mass is allowed to enter or
leave. Heat is transferred within an element resulting in an
increase in the internal energy of the transferred mass. For a
single element, conduction is only allowed through the wall
thickness (z-direction) where duct side walls are modeled as fin
surfaces. Therefore, there is no conduction from one element to
the neighboring elements. This conducted heat is carried away
through convection mechanism by the coolant. Finally, the
model is utilized to impose the continuity and energy balance
equations on every single discreet, 1D finite element. The overall
operation has been shown in Figure 4 for a single element.

Figure 2. Nodal subdivision in chord direction (x) [18]

Figure 4. Schematic of single element

Figure 3. 1D fluid network mapping in span direction (y)

The general process for the heat transfer mechanism in the


LE portion and the rest of the areas has been presented in the
Figure 5. Total heat, Q, is the sum of the heat load resulted from
suction side (qs) and pressure side (qp). However, this individual
heat (qs or qp) transfers by conduction through the blade wall (q1)
and duct side walls modeled as fin (q2).
The program executes in two steps. Initially, it calculates the
mass flow distribution and the film cooling effectiveness; then,
stores the results for each element. Later, it computes the
temperature distribution by solving the energy balance
equations.
Blade co-ordinates and the duct geometries are entered into
the program at the very beginning to evaluate the external
pressure distribution, hot gas temperature and velocity
distribution. It is to be mentioned that the program has the
capability to evaluate these distributions in both of the chordwise
and spanwise directions. In this manner, the results can be
considered as 2D.

Copyright 2015 by ASME

1
2

2
( + ) =

2 2
+ 2.5
+ 2.5
+

2
1 2 = 4 ( )(
)
2

(1-b)
(1-c)

If the channel is two-sided ribbed, the friction factor (2 ) is


calculated from Equation (1-d). To use this equation, is
obtained from Equation (1-b) for a four-sided ribbed channel,
then, is obtained from Equation (1-e) for a four-sided
smooth channel. Next, they are employed into Equation (1-d) to
find 2 which can be used in Equation (1-c) later to calculate
pressure drop for a two-sided ribbed channel.

= 2 + (2 )

64 ,
=
0.0460.2 ,

(a)

(1-d)
(1-e)

From these equations for ribbed channels, an increase of 5 to up


to 10 times is observed in the friction factor depending on the rib
geometry and flow Reynolds number.
In the turning region, the loss/pressure drop coefficient
which depends on the flow Reynolds number and channel aspect
ratio can be estimated from the following equation as discussed
in chapter 4 of Han et al [17].
=

= 1.23 27.07( ) + 17.86( )2


90
90

= 2.2~3.0

(1-f)

2
2

(2)

0.16
0.1175

where = 0.25 +
1.08

MASS FLOW RATE CALCULATION


Flow through the Mid Chord Region. To initiate the
iterative method, an estimate of the flow rate [from the boundary
condition] for a particular coolant passage is entered into the
code as input and the program is run. The pressure drop due to
friction can be computed for a smooth duct condition using
DarcyWeisbach equation. But in real design, ducts have
internal turbulence promoters i.e. repeated ribs which are
included to increase the frictional pressure drop to achieve higher
turbulent mixing. To predict the friction factor on a rib
roughened surface, it is necessary to consider some key factors
such as rib height, spacing and angle. A suitable correlation can
be chosen from chapter 4 in Han et al. [17] where the authors
collected a good amount of information based on those factors.
In this paper, a correlation is chosen [Equation (1-a), (1-b)] to
estimate the friction factor () for a four-sided ribbed channel.
Then, the calculated friction factor is utilized to calculate the
pressure drop using Equation (1-c) across the element shown in
Figure 4.

1 2 =

Figure 5. Heat Balance in (a) LE and (b) other portion

( )0.35 ( )
10

Flow through the Trailing Edge Region. As pressure drop


in the pin-fin passage is quite higher, it can be evaluated using a
different correlation [19] showed in Equation (2).

(b)

Pumping Effect. The change in internal pressure due to


pumping effect is also considered in this model using the
Equation (3). This pumping effect may serve to either increase
or decrease the static pressure based on the direction of the
coolant flow.
1

1 1
1
2

2 2 1 2 2 2
=

30 2 144

(3)

Film Injection Effect. The mass injection rate into the


surrounding hot gas for film cooling is calculated utilizing
Equation (4) that is used for regular orifice flow calculation.
1

=
.
( )
1

(4)

The internal static pressure is used as the driving force for the
injection into the surrounding external boundary layer. The
program is capable of calculating the injection rate into either
suction side (WSinj)/pressure side (WPinj), neither or both based
on the design of a blade, showed in Figure 4. Injection is made
through regularly spaced film cooling holes of equal diameter.

(1-a)

Copyright 2015 by ASME

The injected mass flow rate is subtracted from the mass flow rate
in the duct and the static pressure is recomputed. The computed
pressure and mass flow rate become the inlet conditions to the
adjacent element and the procedure is repeated until the tip is
reached. If at any element in the duct the internal pressure is less
than the external pressure, then the program stops and generates
the appropriate error message. Again, if the duct is multi-pass,
then the procedure is repeated (with film cooling injection, if
required) until the correct number of passes has been calculated
for and a satisfactory convergence is obtained. This process is
presented in Figure. 3.
Flow through the Leading Edge Region. Flow through the
passage adjacent to the LE is handled in a similar manner except
for the procedure taken to calculate the mass injection rate. The
coolant is impinged into the LE inner wall that is computed using
the equation for orifice flow [Equation (4)]. Figure 3 illustrates
the flow arrangement in LE region. Again, it is not possible to
calculate the pressure inside the LE duct in the same manner as
before. Since there is no available method to determine this
pressure, therefore, it is calculated with an iterative approach.
Initially, the static pressure inside the first element of the LE duct
(element 25 in forward loop, Figure 3) is calculated as the
average of external main stream flow and adjacent duct element
(element 17 in forward loop, Figure 3) pressure and then
continues until the last element of the same duct. A minimum
target residual value of mass that remains in the last element
(element 32 in forward loop, Figure 3) is set by user for
convergence check.

estimate the effectiveness quite well. Both of these correlations


predict the cooling effectiveness of shaped-hole geometries very
well and, therefore, either of them is applicable to the current
problem at hand.
=
=

where =

(+1.909)

(5-b)

0.1

1.85
0.6 0.1

(5-c)

+ 1 . 3

4

1 = 0.1721, 2 = 0.2664, 3 = 0.8749,

=2

For LE film cooling effectiveness, it is difficult to find any


complete correlation. However, it can be estimated based on the
previously mentioned correlations, then, the predictions can be
justified with the experimental results shown by Takeishi et al.
[24], Li et al. [25], as shown in chapter 3 of Han et al. [17].
The performance of film cooling can be different in
rotational condition. Many studies indicate the reduction of film
cooling effectiveness on both pressure and suction sides due to
upstream unsteady flow. Again, film tends to be pushed towards
the pressure surface near tip region due to rotating centrifugal
force that could also affect the results from the stationary case.
Thus, a different correlation can be utilized for rotating condition
as discussed in chapter 3 of Han et al. [17].
The average film cooling effectiveness along with corrected
hot gas temperature is used to compute the film temperature
[Equation (6)]. Initially, the hot gas temperature ( ) is
calculated based on the span and chord location. Then, it is
corrected, termed as recovery temperature(r ), so that the
aerodynamic heating that is resulted from the kinetic energy of a
high velocity stream moving over a stationary body can be
considered.
= (1 ) +

(6)

2
2
1
1
[ = , = ]
2
3
where = +

(5-a)

2 0.25

where =

FILM COOLING EFFECTIVENESS


Film cooling helps the boundary layer to cool down to a
certain limit and reduces heat transfer. Film cooling holes can be
different in shape such as cylindrical or shaped and result in
different order of effectiveness.
Initially, the cylindrical holes are transformed in terms of a
2D equivalent slot, then, effectiveness can be calculated based
on correlations in Equation (5-a) provided by Goldstein [20].
These equations can predict the experimental results quite well
for a certain range of [10~100]. But for the rest of the range
[0~10] a constant value of [= 0.45] is selected for the present
model based on the experimental evidence [21].
= (1 + 0.249)0.8

+ 1 2 3

CALCULATION OF CONVECTIVE HEAT


TRANSFER COEFFICIENTS
The program also calculates the external heat transfer
coefficient using the conventional laminar-turbulent well
established experimental correlations, described in the following
paragraphs. These correlations can be supplied to the program as
a function of span and chord length co-ordinates.

S = width of equivalent 2D slot (= 6)


Re2= injected Reynolds number based on equivalent 2D
slot (= 2 2 2 )
Shaped-holes film cooling effectiveness calculation is
different from the cylindrical ones. Equation (5-b), proposed by
Colban et al. [22], can be used to predict the film cooling
effectiveness. Later, Chen et al. [23] proposed an improved
correlation [Equation (5-c)] for the shaped-holes that can

External Heat Transfer Coefficient Calculation. The


external heat transfer coefficients are calculated based on two
physical models.

Copyright 2015 by ASME

The first model assumes that at the LE of the blade the local
heat transfer coefficient may be estimated by flow over a simple
cylinder [26] and can be calculated from Equation (7).
3

= 1.14 ( )0.5 0.4 1

90
[0 < < 80 ]

through a smooth duct using Equation (10). This equation is only


applicable for turbulent flow inside a duct.

= 0.023 45 1/3

To enhance the internal blade cooling efficiency, modern


design of turbine blade includes turbulence promoters called rib
turbulators [Figure 1]. To reflect this effect, heat transfer
coefficient is computed using Equations (11-a) and (11-b)
proposed by Han and Park [28] based on the rib configuration.
This enhancement in heat transfer is about 2~4 times depending
on the rib geometry and flow Reynolds number.

(7)

As upstream turbulence is responsible for heat transfer


augmentation, then, the higher value of external HTC should be
predicted. If Reynolds number (ReD), turbulence intensity (Tu)
and energy scale (Lu) are known, then the average HTC can be
calculated using Equation (8) [27].
=

1
2


0.95 + 0.038 12

1
3

0.1 / + 0.35
= 2.24
( )

90
10

1
2
( + , ) = ( + ) +

( )1/2

(8)

Then, an average of local values from Equation 7 is


compared with the value obtained from Equation 8. The resulted
augmentation ratio is multiplied back to HTCs from Equation 7
and they are updated in the program accordingly.
In the second model, the convective heat transfer coefficient
is calculated for the rest of the curve length along the
suction/pressure side by being simulated as external flow over a
flat plate. The HTC is estimated using the so called laminar and
turbulent correlations, presented in Equations (9-a) and (9-b)
[26]. The appropriate equation is chosen based on the critical
Reynolds number (Rec = 5 x 105). In the real engine condition,
the oncoming flow is highly turbulent and the blade surface
offers certain roughness that may lead to the early transition.
Thus, the critical Reynolds number may vary for particular
applications to compensate that early transition.
12

1/3 []

= 0.0296 4 5 1/3 []

= 0.332

where, =

(11-a)
(11-b)

Again, if there is any flow turning at the hub or tip, then, it


improves the heat transfer in that specific region. So a suitable
enhancement factor could be used to calculate the internal HTC
using Equation 9-c.
The LE portion of a turbine blade is the most crucial part as
it sees the high inlet temperature. Thus, special treatment has
been taken under consideration. Jet impingement cooling
scheme is employed inside the LE where coolant stream is
directed to hit the surface at spatial intervals [Figure 1]. This heat
transfer coefficient is computed using Equation (12) [29].

(12)
0.5
0.6
0.5
1.2

= 0.63 ( )0.7 1.27

(9-a)

Due to external flow separation and the resultant increase in


the external convective heat transfer coefficient, it is essential to
achieve higher value of internal convective heat transfer
coefficient on the TE. This is the reason pin-fin arrangement is
employed to obtain higher value of convective heat transfer.
Average array heat transfer coefficient for the case of staggered
pins can be calculated using Equation (13) suggested by Metzger
et al. [30]

(9-b)

, and is the local mainstream velocity

along the blade.


For the rotational condition, the literature survey clearly
[17] indicates that the blade heat transfer can be affected by the
unsteadiness in the flow caused by the upstream airfoil. Many
studies have shown that unsteady wake from vane TE along with
free stream turbulence or even shock phenomenon not only
promotes early transition, but also increases heat transfer.
Therefore, these effects may be considered to accurately predict
the external heat transfer. Suitable information can be found in
chapter 2 of Han et al. [17] and augmented HTC can be
calculated in the following manner.

(10)

0.34

0.685

= 0.135

(13)

All the above correlations are evaluated for the stationary


state. But for rotational condition, results may vary as flow in a
stationary channel is substantially different from flow in a
rotating channel. Coriolis and rotational buoyancy forces
resulted from the rotation of a rotor blade can significantly affect
the internal heat transfer. Again, heat transfer coefficients can be
different on the leading and trailing walls of the internal duct. In
this case, appropriate correlations are presented and discussed in
chapter 5 of Han et al. [17].

(9-c)

Internal Heat Transfer Coefficient Calculation. The


internal convective heat transfer coefficient of the model is
computed for three different possibilities based on particular
blade designs. The internal convective heat transfer coefficient
for the central region of the blade element will be estimated using
conventional Dittus-Boelter equation for forced convection

CALCULATION OF THERMO-FLUID PROPERTIES


Thermal properties of the fluid, density, viscosity, and
thermal conductivity, were considered to be temperature
dependent and are calculated from the following equations:

Copyright 2015 by ASME

Dynamic viscosity is obtained from Sutherlands law:


=

2/2 = [ tanh() ]/
Where, = , =
P = perimeter and As = fin area

(14-a)

Increase in internal energy in a single element:

0 + 2
(14-b)
() = 0

+ 0
where 0 =0.00001827 Pa-s, 0 = 291.15 , = 120

() = 3.225282 1015 4 + 1.96488


1011 3 5.15701 108 2
+ 1.0334 104 6.40447
104
[175 < < 1900]

Heat balance:

/ = [( )( )]/

/ = [(

) ( )]/

(14-c)

) ( )]/

Internal convection on pressure/suction side:

1/1 = [( ) ]/

Fin conduction through side walls of a duct:

1
2

(2 + 1 )

(20)
(21)
(22)
(23)

From the analysis, it is apparent that the thermal circuit


includes five different resistances as shown in the Figure 6. Each
resistance is considered as a unique input for solving these set of
equations where it can be set as zero if necessary.

Figure 6. Thermal circuit


The simultaneous operation of the above equations results
in calculation of internal and external temperatures on pressure
and suction side for each element. To initiate the iteration
process, an assumption is made where the average coolant
temperature ( ) is equal to the inlet temperature of the first
element of the first duct in a single loop as shown in Figure 3.
The end results are written in a specific file and at the same time
it can be analyzed in a contour format. Overall operation has
been presented in the form of a flow chart in Figure 7.

(15)
(16)

Conduction through wall on pressure/suction side:


/ = [(

/ = 1/1 + 2/2
= +

Average coolant temperature:

CALCULATION OF BLADE METAL TEMPERATURE


To calculate the temperature of a single element, the
estimation of the internal and external convective heat transfer
coefficients are required at that specific surface location of the
blade. These calculations are made based on the following five
mechanisms of heat transfer for a general element.
1. External convection from external gas flow to the
suction and pressure sides.
2. Radiation from the hot gas.
3. Conduction through the suction and pressure side of the
blade material.
4. Internal convection from the suction and pressure sides
to the coolant.
5. Increase in the internal energy of the coolant as it
collects the heat.
Radiation in this problem is negligible in comparison to the
convection and conduction orders. Therefore, by neglecting its
effect, the model can be simplified without losing much
accuracy. This assumption had also been justified in the
literature before [1]. As discussed earlier, heat conduction is one
dimensional in nature, so conduction in the span and chord
direction has been neglected. As a result, heat is only conducted
from the pressure or suction side through the ducts side wall
boundaries which are treated as fins. Then, convection takes
place through the inner surface (Ai) and fin areas. It is essential
to mention that the external area (A0) is greater than internal area
(Ai). Based on these assumptions, corresponding relations can be
extracted in the equation forms [Equations (15-23)]. The same
relations are applicable for both pressure and suction sides.
Therefore, the subscripts in the Equations (15-21) are presented
to stand for either pressure/suction side.
External convection on blade pressure/suction side:
Conduction through TBC layer on pressure/suction side:

= (2 1 )

(19)

Figure 7. Flow chart

(17)

RESULTS AND DISCUSSION


This section is divided into two parts. First, the validity of
the model is assessed by comparing the calculated temperatures

(18)

Copyright 2015 by ASME

to the ones available from the open literature for an E3 blade


which serves as the main purpose of this paper. In the second
part, the model is further applied on the same blade profile but
with new boundary conditions to reflect the flexibility of the
code.
PART I
For the model validation, an E3 rotor blade profile, studied
earlier by Halila et al. [18], is selected. Then, a comparison is
made to evaluate the model performance. The E3 blade has 8
cooling ducts including impingement, showerhead and pin-fin
cooling schemes along with film cooling holes at several
locations. A breakdown of the cooling arrangement is shown in
the Figure 8.

Figure 2. Later, the 2D model of the blade section is sub divided


in a total 64 elements where 8 sections are in the chordwise
direction and 8 sections are in the spanwise direction (Figure 3).
Figures 9 shows that Rotor Inlet Temperature (RIT) varies
in the spanwise direction while mach number profile varies in
both the spanwise and chordwise directions.
All these information are directly reproduced from the
reference report [18] except the temperature profile. Instead of
temperature profile, energy extraction from each stage has been
utilized, thus, a profile is produced in terms of inlet temperature.
Later, a linear drop of 200oC in the chord direction is considered
for the present model to meet the boundary condition.
The external HTC distribution (Figure 10) used for
validation has been taken from the report [18]. Reynolds number
inside the internal passages is calculated based on the distributed
mass flow rate in the backward and forward loops as shown in
Figure 11-a and 11-b. The plots are exhibited elementwise for
each passage representing the flow loop that was shown earlier
in Figure 3. Since the second duct is losing its coolant mass flow
rate by injection into duct 1, an obvious linear drop is observed
in Figure 11-a for the forward loop. An identical trend is also
observed in Figure 11-b for the backward loop for ducts 5 and 7.
Since duct 7 is supplying coolant mass to the pin-fin passage and
duct 5 has film cooling holes on the pressure side, then, a linear
drop occurs in Reynolds numbers.

Figure 8. Detailed view of the reference case [18]


There are two internal multi-pass coolant loops, termed as
forward loop (1.63% of W25 mass flow rate) containing the LE
impingement passage and backward loop (1.67% of W25 mass
flow rate) containing the TE pin-fin passage. Forward loop
includes 12 LE impingement cooling holes, 3 rows of LE holes
(10/row), 23 gill holes on the pressure side, and diffused type
suction side gill holes. Backward loop includes 15 pressure side
holes and the pin-fin passage. However, passages are rib
roughened in both loops. The inlet boundary conditions and
necessary information are supplied in Table 1.

(a)

Table 1: Boundary conditions [18]


Number of Blades
76
Pressure ratio
2.25
Mainstream Inlet Pressure (Total)
2.52 MPa
Rotor Inlet Total Temperature
1396 C
Coolant Inlet Pressure (Total)
2.61 MPa
Coolant Inlet Total Temperature
628 C
Inlet Mach Number
0.34
Exit Mach Number
0.84
Speed
13287 rpm
The blade is subdeivided into eight different nodal points, as
mentioned earlier to compare with the reference case along the
pressure and suction sides in the chordwise direction similar to

(b)

Figure 9. (a) Spanwise non-dimensional RIT and inlet Mach


number profile [18], (b) Surface Mach number distribution
[18]

Copyright 2015 by ASME

Later, internal HTC is calculated for different passages and


shown in the similar fashion in Figures 12-a and 12-b.
Next, the results of temperature distributions are generated
using all the details and compared with the reported data [18] as
depicted in Figure 13 and 14 for the mid span region. According
to Table 2, the results demonstrate a close estimation of the metal
temperature prediction with a maximum relative error of
10.64%. Particularly, Suction side prediction is more accurate
than the pressure side one. This error is simply the ratio of
difference between the calculated and the known value over the
known value.

100
(%) =

The major part of uncertainty in this work is due to lack of


complete geometrical information from the reference report such
as the size of the film cooling holes or coolant duct dimensions.

Figure 10. HTC distribution along the blade surface [18]

(a)

(a)

(b)

(b)

Figure 12. (a) Calculated internal HTC for the forward


loop, MFR = 1.63%W25, (b) Calculated internal HTC for
the forward loop, MFR = 1.67%W25

Figure 11. (a) Calculated internal Reynolds number for the


forward loop, MFR = 1.63%W25, (b) Calculated internal
Reynolds number for the backward loop, MFR = 1.67%W25

Copyright 2015 by ASME

so on. However, for brevity, not all but some of these factors are
altered; then, the new results are obtained, and visualized in the
form of contour plots. Consequently, a higher RIT (1700 oC), a
spanwise variable HTC distribution with transition model, and a
layer of TBC were applied while keeping the other boundary
conditions and coolant MFR at the same values as were in the
previous part.
In contrast to PART I, the HTC profile is not taken from the
available literature [18] in this part and it is calculated utilizing
(9-a) and (9-b), instead.
Since in real gas turbine engines HTC is different from tip
to hub due to phenomena such as tip leakage or end wall vortices,
then, the blade is sub-divided into three various regions to count
for these effects. These three regions are near-hub-zone (0-20%
span), mid-span (20-80% span) and near-tip-zone (80-100%
span). Different velocity distributions specific to each of the
aforestated regions are replicated from the report [18] and
implemented into the HTC correlations.
As from literature [17], it can be observed that boundary
layer transition starts on suction side within 15-25% surface
distance from the LE, then a bit different value of critical
Reynolds number (3x105) has been set for flow shift from
laminar to turbulent. However, this behavior is highly case
specific and depends on the blade profile.
Also, high free stream turbulence can augment the external
heat transfer near the stagnation region. Thus, a more realistic
estimation of HTC is made on this region using Equation (8)
where turbulence parameters for aero-combustor simulator are
considered from the detailed study of turbulence characteristics
by Chowdhury et al. [31]. By comparing the results with
Equation (7), an increase of about 40% in heat transfer values
was suggested. Based on all these information, Figure 15 has
been generated where it is showing three different HTC
distributions at various span regions.
Thermal barrier coating (TBC) is also considered in the
calculation, as the RIT is high enough. The presence of TBC
layer can moderately lessen the heat load by offering an
insulating effect between the hot main stream flow and the blade
external surface. A certain thickness of 0.25 mm TBC layer [16]
is considered for this study with the thermal conductivity of 1.3
W/m/K.
The temperature contours, in Figure 16, shows the computed
results on the suction side and pressure side of the blade for an
88 i.e. a total of 64 elements on a 2D plane. Of course, the
number of elements can be increased easily if designers desire to
have a more sophisticated plot. It can be seen from the contours
that a hot spot zone is clearly detected on the upper left corner
region for the both suction side and pressure side. This
observation can be attributed to the injection of coolant through
impingement holes to the mainstream hot gas flow as the coolant
moves upward towards the tip region inside the forward loop
duct. Now, the true value of the code reveals itself to an expert
of gas turbine cooling systems. Failure of many gas turbine
engines starts from the hot spots of the blades. Utilizing this
program, these hot spots are easily surfaced and targeted to be
modified. For instance, after the very high temperature locations

Figure 13. Results comparison with the reference case for


external surface temperature (Two), RIT = 1396oC, MFR =
3.3%W25, no TBC layer

Figure 14. Results comparison with the reference case for


internal surface temperature (Twi), RIT = 1396oC, MFR =
3.3%W25, no TBC layer
Table 2: Error (%) chart for the calculated results
PS
External
Internal
SS
External
Internal
Node Temp,Error (%) Temp,Error (%) Node Temp,Error (%) Temp,Error (%)
-1

9.87

0.37

8.88

1.92

-2

0.09

1.15

6.33

3.93

-3

8.18

7.12

1.36

7.58

-4

10.33

8.92

0.75

10.64

-5

4.54

5.13

0.73

2.28

-6

1.47

5.26

4.91

1.36

-7

6.45

5.89

0.67

1.23

-8

3.40

0.001

2.32

4.54

PART II
The goal of this part is to show the interactive features of the
code and how easily a user can change the input design
parameters such as TBC thickness, HTC distribution, TIT,
coolant MFR, film cooling hole and internal duct geometry, and

10

Copyright 2015 by ASME

have been spotted by this code, a designer can move forward to


heal these spots by modifying the current design through
adding/removing extra features such as additional cooling holes
while trying to keep the compressors efficiency as before. This
obviously leaves a great opportunity for the whole gas turbine
industry to make more energy efficient and safer engines.

a) Suction side

Figure 15. Variable HTC distributions along the blade surface


Higher HTC on suction side, as seen in Figure 15, results in
capturing higher temperatures on this surface. Again, one reason
for this higher heat load could be the existence of only one single
row of film cooling at the suction side. Addition of extra rows
with increased MFR on the suction side could aid the surface to
cool down furthermore.
Conduction in the blade, for each separate element, is
modeled to be 1D and the problem is simplified by neglecting
conduction effects in the span and chord directions. But it is
apparent that there is a noticeable difference between
neighboring elements in both directions and it will obviously
affect the blade surface temperature. Therefore, an assessment of
uncertainties is essential to ensure the results of the aforestated
1D assumption is reliable.
The blade heat conduction in three different directions can
be calculated as following.

b) Pressure Side
Figure 16. Temperature contour for E3 blade, RIT =
1700oC, MFR = 3.3%W25, with TBC layer (0.25 mm)
CONCLUSIONS
The present work is intended to develop an analytical model
that can serve as a quick tool for preliminary design of a cooled
vane/blade. This paper is demonstrated in a few steps. First of
all, the model is compared with a reference blade condition that
is available in the open domain. The temperature results exhibit
reasonable agreement with a maximum uncertainty of about
11%. This close prediction ensures the capability of this program
to evaluate the temperature distributions with higher accuracy if
complete geometrical details, kept confidential by many
companies, are provided to the code.
Next, the flexibility and user-friendliness of the code
is illustrated by applying a new set of values for input parameters
such as TBC thickness, HTC distribution, and RIT. The
predicted temperature distributions are depicted in the form of a
contour plot as a visualizing tool for designers.
Moreover, this model is quite flexible to any type of
modification that is required for new cooling designs and even
totally new blade profiles. Therefore, this program gives an
opportunity to the designers to justify the performance of the
cooling system and to modify its design as needed in an optimum

( ) [ ()]

( ) [ ()]
=

( ) [ ()]
=

After the calculation of conductions for different directions,


the ratio of (qc + qs) / qd is found to be less than 10% which
demonstrates the validity of the simplifying assumption and
proves the reliability of the heat conduction model.

11

Copyright 2015 by ASME

way. In other words, this novel program is a chance for the


industry to reduce costs in the preliminary stage of gas turbine
design.

Acronyms
1D
2D
3D
2side
E3
LE
RIT
TE
TIT
Subscripts
1p
2p
1s
2s
1
2
25
c
cf
f
h
i

inj
l
o
P/p
r
S/s
TBC
TBCP
TBCS
wo/wi
WIP/WIS
WOP/WOS

ACKNOWLEDGEMENT
The authors gratefully acknowledge the support for this
project from the Samsung Techwin R&D Division.
NOMENCLATURE
Symbols
A
cp
CD
d
D
DR
e
e+
f
G
h /h
H
k
K
l
Lu
M
Ma
n
N
P
Pr
PS
Q/q
R/r
Re
S
St
T
T
Tu
V
W
X/x

area [m2]
specific heat capacity [J/kg/K]
discharge coefficient [-]
impingement/film cooling hole diameter [m]
diameter [m]
density ratio [-]
roughness height [m]
roughness Reynolds number [-]
friction factor [-]
heat transfer roughness function [-]
(average)/heat transfer coefficient [W/m2/K]
channel height [m]
thermal conductivity [W/m/K]
pressure drop/loss coefficient
jet to target plate separation distance [m]
energy scale [m]
blowing ratio [-]
Mach number [-]
rotational speed [rpm]
row numbers
pitch distance between ribs [m]
Prandtl number [-]
static pressure [Pa]
heat load [Watt]
roughness function [-]/radius of the element [m]
Reynolds number [-]
distance between impingement holes/pins
Stanton number, (St = h/VCp) [-]
temperature [oC]
temperature difference [oC]
turbulence level [%]
approach//local mainstream velocity [m/s]
mass flow rate [kg/s]/channel width [m]
distance from injection hole center [m]/
chorwise direction
Y
spanwise direction
Greek Symbols

rib angle [degree]


thickness [m]

laterally averaged effectiveness

angular displacement from stagnation line


coolant viscosity at the injection point [pa-s]
2
mainstream viscosity [pa-s]

density [kg/m3]

2D film cooling parameter

one- dimensional
two dimensional
three dimensional
two-sided ribbed channel
energy efficient engine
leading edge
rotor inlet temperature
trailing edge
turbine inlet temperature
normal direction to pressure side surface
normal direction to fin surface (pressure side)
normal direction to suction side surface
normal direction to fin surface (suction side)
at inlet of element
at exit of element
core compressor inlet plane (to combustor)
cross section/coolant
coolant flow
film
hydraulic
internal
main stream
injection
leading edge
external
pressure side/pin
rough/recovery
suction side
thermal barrier coating
thermal barrier coating on pressure side
thermal barrier coating on suction side
external/internal wall
internal wall on pressure/suction side
external wall on pressure/suction side

REFERENCES
[1] Han, J.C., Ortman, D.W., and Lee, C.P., A Computer
Model for Gas Turbine Blade Cooling Analysis, ASME Paper
No. 82-JPGC-GT-6, presented at the ASME Joint Power
Generation Conference, Denver, Colorado, October 1982.
[2] Hylton L. D., Mihelc M. S., Turner E. R., Nealy D. A.,
York R. E., 1983, Analytical and Experimental Evaluation of
the Heat Transfer Distribution Over the surfaces of turbine
vanes, NASA CR 168015.
[3] Zecchi, S., Arcangeli, L., Facchini, B., and Coutandin, D.,
2004, Features of a Cooling System Simulation Tool Used in
Industrial Preliminary Design Stage, ASME Paper No.
GT2004-53547.
[4] Alizadeh, M., Izadi A., and Fathi, A., 2014, Sensitivity
Analysis on Turbine Blade Temperature Distribution Using
Conjugate Heat Transfer Simulation, ASME J. Turbomach.,
vol. 136, pp. 542550.

12

Copyright 2015 by ASME

[5] Luo, J., and Razinsky, E. H., 2007, Conjugate Heat Transfer
Analysis of a Cooled Turbine Vane Using the v2-f Turbulence
Model, ASME J. Turbomach., 129, pp. 773781.
[6] Takahashi, T., Watanabe, K., and Sakai, T., 2005,
Conjugate Heat Transfer Analysis of a Rotor Blade with RibRoughened Internal Cooling Passages, ASME Paper No.
GT2005-68227
[7] Carcasci, C., Facchini, B., 1996, A numerical procedure
to design internal cooling of gas turbine stator blades, Revue
Gnrale de Thermique, Volume 35, Issue 412, Pages 257268.
[8] Andreini, A., Bonini, A., Carcasci, C., Facchini, B.,
Innocenti, L. and Ciani, A., 2012, Conjugate heat transfer
calculations on GT rotor blade for industrial applications. Part I:
equivalent internal uid network setup. ASME Paper No.
GT2012-69846.
[9] Brillert, D., Dohmen H. J., Benra, F. K., Schneider O. and
Mirzamoghadam, A. V., 2003, Application of Conjugate CFD
to the Internal Cooling Air Flow System of Gas Turbines,
ASME Paper No. GT2003-38471.
[10] Yamane, T., Yoshida, T., Enomoto, S., Takaki, R. and
Yamamoto, K., 2004, Conjugate Simulation of Flow and Heat
Conduction with a New Method for Faster Calculation, ASME
Paper No. GT2004-53680.
[11] Sipatov, A., Gomzikov, L., Latyshev, V., and Gladysheva,
N., 2009, Three Dimensional Heat Transfer Analysis of High
Pressure Turbine, ASME Paper No. GT2009-59163.
[12] Mangani, L., Cerutti, M., Maritano, M., and Spel, M.,
2010, Conjugate Heat Transfer Analysis of NASA C3X Film
Cooled Vane with an Object-Oriented CFD Code, ASME Paper
No. GT2010-23458.
[13] Chi, Z., Ren, J. and Jia, H., 2013 Coupled AeroThermodynamics Optimization for the Cooling System of A
Turbine Vane, ASME Paper No. GT2013-94528.
[14] Heidmann, J. D., Kassab A. J., Divo E. A., Rodriguez F. and
Steinthorsson, E., 2003, Conjugate Heat Transfer Effects on a
Realistic Film-Cooled Turbine Vane, ASME Paper No.
GT2003-38553.
[15] Rigby D. L. and Lepicovsky, J., 2001 Conjugate Heat
Transfer Analysis of Internally Cooled Congurations, ASME
Paper No. GT2001-0405.
[16] Downs, J. P., and Landis, K. K., 2009, Turbine Cooling
Systems Design Past, Present and Future, ASME Paper No.
GT2009-59991.
[17] Han, J. C., Dutta, S., and Ekkad, S., 2012, Gas Turbine
Heat Transfer and Cooling Technology, 2nd edition, Taylor &
Francis Group, CRC Press.
[18] Halila, E. E., Lenahan, D.T. and Thomas, T.T., 1982, High
Pressure Turbine Test Hardware Detailed Design Report,
Unclassified NASA Technical Report.
[19] Jacob, M., 1938, Heat Transfer and Flow Resistance in
Cross Flow of Gases Over Tube Banks, Trans. ASME, 59, pp.
384386.
[20] Goldstein, R.J., 1971, "Film Cooling, Advances in Heat
Transfer, Vol.7, Academic Press, New York and London.
[21] Jabbari, M.Y., and Goldstein, R.J., "Adiabatic Wall
Temperature and Heat Transfer Downstream of Injection

through Two Rows of Holes," ASME Journal of Engineering for


Power, Vol.100, pp.303-307, 1978.
[22] Colban, W. F., Thole, K. A., and Bogard, D., 2010, "A FilmCooling Correlation for Shaped Holes on a Flat-Plate Surface,"
Journal of Turbomach., 133, pp. 011002.
[23] Chen, A. F., Li, S., J., and Han, J. C., 2014, Film Cooling
With Forward and Backward Injection for Cylindrical and FanShaped Holes Using PSP Measurement Technique, ASME
Paper No. GT2014-26232
[24] Takeishi, K., Aoki, S., Sato, T. and Tsukagoshi, K., 1992,
Film Cooling on a Gas Turbine Rotor Blade, Journal of
Turbomach., Vol. 114, pp. 828-834.
[25] Li, S. J., Yang, S. F., and Han, J. C., 2014, "Effect of
Coolant Density on Leading Edge Showerhead Film Cooling
Using the Pressure Sensitive Paint Measurement Technique,"
Journal of Turbomach., Vol. 136, pp. 051011.
[26] Kreith, F., Bohn, M.S., Principles of Heat Transfer,, 5th
edition, PWS Publishing Company, 1998, pp. 292 and 452.
[27] Ames, F. E., 1997, The Influence of Large Scale, High
Intensity Turbulence on Vane Heat Transfer, ASME J.
Turbomachinery, 119, pp. 2330.
[28] Han, J. C., Park, J. S., 1988, Developing Heat Transfer in
Rectangular Channels with Rib Turbulators, Journal of Heat
and Mass Transfer, Vol. 31, No. 1, pp. 183-195.
[29] Chupp, R. E., Helms, H.E., Mcfadden, P.W. and Brown, T.
R., 1969, Evaluation of Internal Heat Transfer Coefficients for
Impingement Cooled Turbine Airfoils, AIAA Journal of
Aircraft, Vol. 6, pp. 203-208.
[30] Metzger, D. E., Shepard, W. B., and Haley, S. W., 1986,
Row Resolved Heat Transfer Variations in Pin Fin Arrays
Including Effects of Non-Uniform Arrays and Flow
Convergence, ASME Paper No. 86-GT-132.
[31] Chowdhury, N., Ames, F. E., 2013, The Response of High
Intensity Turbulence In The Presence of Large Stagnation
Regions, ASME Paper No. GT-2013-95055.

13

Copyright 2015 by ASME

Das könnte Ihnen auch gefallen