You are on page 1of 10

Heat Transfer Enhancement in a

Rectangular (AR 5 3:1) Channel


With V-Shaped Dimples
C. Neil Jordan
Lesley M. Wright
e-mail: Lesley_Wright@Baylor.edu
Department of Mechanical Engineering,
Baylor University,
Waco, TX 76798-7356

An alternative to ribs for internal heat transfer enhancement of gas turbine airfoils is
dimpled depressions. Relative to ribs, dimples incur a reduced pressure drop, which can
increase the overall thermal performance of the channel. This experimental investigation
measures detailed Nusselt number ratio distributions obtained from an array of V-shaped
dimples (d/D 0.30). Although the V-shaped dimple array is derived from a traditional
hemispherical dimple array, the V-shaped dimples are arranged in an in-line pattern.
The resulting spacing of the V-shaped dimples is 3.2D in both the streamwise and spanwise directions. A single wide wall of a rectangular channel (AR 3:1) is lined with
V-shaped dimples. The channel Reynolds number ranges from 10,00040,000. Detailed
Nusselt number ratios are obtained using both a transient liquid crystal technique and a
newly developed transient temperature sensitive paint (TSP) technique. Therefore, the
TSP technique is not only validated against a baseline geometry (smooth channel), but it
is also validated against a more established technique. Measurements indicate that the
proposed V-shaped dimple design is a promising alternative to traditional ribs or hemispherical dimples. At lower Reynolds numbers, the V-shaped dimples display heat transfer and friction behavior similar to traditional dimples. However, as the Reynolds
number increases to 30,000 and 40,000, secondary flows developed in the V-shaped concavities further enhance the heat transfer from the dimpled surface (similar to angled
and V-shaped rib induced secondary flows). This additional enhancement is obtained
with only a marginal increase in the pressure drop. Therefore, as the Reynolds number
within the channel increases, the thermal performance also increases. While this trend
has been confirmed with both the transient TSP and liquid crystal techniques, TSP is
shown to have limited capabilities when acquiring highly resolved detailed heat transfer
coefficient distributions. [DOI: 10.1115/1.4006422]

Introduction
With gas turbines playing an increasingly important role around
the world, any improvements in the thermal efficiency of the
engines should be realized. Increasing the turbine inlet temperature of the working gas has shown to greatly improve the power
generation and the thermal efficiency of the engines. However,
increasing this turbine inlet temperature must be done cautiously
as the high gas temperatures can reduce the life of the turbine
components and thus the life of the engine.
Han et al. [1] have outlined advanced cooling technologies that
have been developed to combat the detrimental effects of the continuously rising turbine inlet temperatures. Both external and internal cooling strategies have been developed to ensure turbine
blades and vanes can withstand the extreme engine temperatures.
Thermal barrier coatings (TBC) and film cooling have become fixture techniques in protecting the outer surface of the airfoils. With
both TBC and film cooling, a layer of resistance is created
between the metallic airfoils and the hot mainstream gas. With
film cooling, the layer of resistance is created by the thin air film
created on the outer surface of the blades and vanes.
Within the hollow airfoils, several techniques are used in tandem to transfer heat from the airfoil walls to the coolant circulating through the internal cooling passages. Jet impingement is an
aggressive cooling technique that is typically employed in the
leading edge region of the airfoils. Jet impingement affords very
high heat transfer coefficients within the airfoils; however, a large
pressure drop is incurred with the jets. Pin-fin cooling is typically
Contributed by the International Gas Turbine Institute (IGTI) of ASME for publication in the JOURNAL OF TURBOMACHINERY. Manuscript received July 31, 2011; final
manuscript received August 20, 2011; published online October 30, 2012. Editor:
David Wisler.

Journal of Turbomachinery

used in the trailing edge region of the airfoils, where both high
heat transfer and structural support are required. Finally, rib turbulators are most commonly used throughout the midspan of the airfoils. These trip strips extend from the cooling passage to protrude
out of the viscous sublayer. Therefore, the ribs continually disturb
the turbulent boundary layer and increase the heat transfer from
the airfoil walls.
A general review of internal cooling technology was conducted
by Ligrani et al. [2]. This review indicated rib turbulators are capable of enhancing the heat transfer from 2 to 5 times that of a
smooth channel (depending on rib design, channel aspect ratio,
and coolant Reynolds number). This heat transfer enhancement
comes at the expense of increased pressure penalties as high as 71
times that of a smooth channel (although the majority of ribbed
channels yielded friction coefficients 1020 times greater than
smooth channels). Dimple concavities were shown to incur a significantly lower pressure penalty: topping out at only 4.5 times
that of a smooth channel. At this reduced pressure drop, dimples
are capable of enhancing heat transfer 2 to 3 times that of a
smooth channel. Kim et al. [3] compared jet impingement, rib turbulators, and hemispherical dimples at high Reynolds numbers
(26,000 < Re < 360,000), and they confirmed over this range of
Reynolds numbers dimples offer superior thermal performance
due to the significant reduction in pressure drop they provide.
Chyu et al. [4] showed rectangular channels with dimples can provide Nusselt number ratios approaching 2.5. In addition to the
dimples providing a reduced pressure drop, it was also indicated
dimples are a desirable alternative to ribs due to the decreased
weight associated with the removal of surface material compared
to the addition of material with ribs [4].
Early studies of internal cooling with dimples focused on staggered arrays of hemispherical dimples within the channel.

C 2013 by ASME
Copyright V

JANUARY 2013, Vol. 135 / 011028-1

Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/ on 05/08/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

The benefit of dimples was quickly realized with such significant


reduction in pressure losses and moderate heat transfer enhancement. The use of dimples was also viewed favorably as the level
of heat transfer enhancement was shown to be independent of
Reynolds number (for ribs as the Reynolds number increases, the
Nusselt number ratios decrease) [5]. Moon et al. [6] confirmed the
Reynolds number independency, and from their study of detailed
heat transfer coefficient distributions, they determined the majority of heat transfer enhancement occurs outside the dimples (with
a reduction of heat transfer inside the concavities). From a computational study, Lin et al. [7] attributed the heat transfer enhancement to vortical structures formed within the concavities and
promoting turbulent mixing as they emerge from the dimples.
With knowledge that hemispherical dimples are promising
alternatives to rib turbulators, a group of researchers from the University of Utah considered various geometrical factors that can
influence the level of heat transfer enhancement within dimpled
channels [811]. Due to the vortical structures created by the dimples, the heat transfer enhancement increases as the ratio of air
temperature to wall temperature decreases [8]. Decreasing the
height of the channel was also shown to increase heat transfer
within the dimpled channels [9]. Within rectangular channels,
Nusselt numbers are also increased if the depth of the hemispherical dimples is increased (as the strength of the vortices formed
within the dimples increases) [10,11]. Griffith et al. [12] measured
similar levels of heat transfer enhancement within a rectangular
4:1 channel with two walls roughened with dimples. They were
also able to extend their study to quantify the effect of rotation on
heat transfer enhancement in dimpled channels.
While the majority of dimple studies have involved hemispherical dimples of different depths, several groups have considered dimples of different shapes. In addition to hemispherical
dimples, Chyu et al. [4] considered teardrop-shaped dimples. The
teardrop shape mitigates flow separation on the upstream half of
the dimple, and therefore, this region experiences less of a reduction in heat transfer (compared to the hemispherical dimples).
Moon and Lau [13] compared the thermal performance of hemispherical dimples to cylindrical dimples. This experimental study
indicated in a square channel the thermal performance of cylindrical dimples exceeds that of hemispherical dimples. Borisov
et al. [14] considered a banked dimple design that significantly
enhanced heat transfer within the channel; however, this
enhancement came at the expense of significantly increased friction coefficients within the channel. Zhou and Acharya [15] performed an experimental and numerical investigation of square,
triangular, circular, and teardrop shaped dimples. Similar to
Chyu et al. [4], they concluded that teardrop-shaped dimples provide the most heat transfer enhancement of the four dimple
shapes considered.
Although dimples have proven to provide improved thermal
performance compared to rib turbulators, they have not replaced
ribs as the staple internal cooling technology. The superior thermal performance is the result of reduced pressure losses while the
heat transfer enhancement is less than that expected in ribbed
channels. Therefore, it is necessary to optimize dimple geometries, so they offer an improved alternative to ribs. The current
experimental study proposes a V-shaped dimple design that
will induce secondary flow structures similar to those created in
ribbed channels. The thermal performance of the V-shaped dimples will be evaluated over a range of Reynolds numbers
(Re 10,00040,000). Both the heat transfer enhancement and
frictional losses are measured in a 3:1 rectangular channel with
one wall roughened with the V-shaped dimples. In addition to
evaluating the thermal performance of the proposed V-shaped
dimples, a newly developed transient temperature sensitive paint
is evaluated against the more traditional, transient liquid crystal
technique. Therefore, the objectives of this study are to evaluate
the viability of V-shaped dimples for use inside turbine airfoils
and to present a new TSP technique for the acquisition of detailed
heat transfer coefficient distributions.
011028-2 / Vol. 135, JANUARY 2013

Experimental Techniques
Detailed heat transfer coefficient distributions are obtained on a
dimple-roughened surface using two optical techniques. First, a
traditional, single color capturing, transient liquid crystal technique is used to obtain the detailed distributions. The results
obtained from the transient liquid crystal method are used to
assess the viability of a newly developed transient temperature
sensitive paint technique. The two transient methods are related as
they are both based on the one-dimensional, semi-infinite solid
assumption. In both cases, the edge of the dimples is a region of
concern with this assumption (as reduced material thickness
exists). However, throughout this study, no numerical correction
has been implemented around the dimples. The details of both experimental techniques are described below.

Transient Liquid Crystal Technique. Heat transfer coefficients are obtained on the dimpled surface by utilizing thermochromatic liquid crystals (TLC). With the traditional, single color
capturing technique, the time to reach a given surface temperature
(liquid crystal hue) is determined. With knowledge of the surface
temperature and the time required to reach this temperature, the
convective heat transfer coefficients can be calculated across the
surface using a one-dimensional, semi-infinite solid assumption
[16]. The one-dimensional, transient heat conduction equation is
given as
@ 2 T qs cp @T

@y2
ks @t

(1)

The boundary condition assumptions of the semi-infinite solid and


a purely convective surface, as well as the initial condition for the
transient experiment, are shown in Eq. (2).
t 0;
y 0;
y ! 1;

T Ti
@T
hTw  Tb
ks
@y
T Ti

(2)

Solving Eq. (1) at the surface (y 0) with the initial and boundary
conditions given in Eq. (2) yields the following solution [16]:
 2 
 p
Tw  Ti
h at
h at
erfc
(3)
1  exp
ks2
ks
Tb  Ti
Because the air temperature through the channel does not experience a true step change to initiate the transient test, the time dependency of the bulk fluid temperature must be taken into
account. The bulk air temperature is approximated as a series of
time steps taken through the course of the transient experiment.
Considering the time dependency of the air temperature, Eq. (3) is
modified with the incorporation of Duhamels theorem of superposition, as shown in Eq. (4).
2
0 q

13

 2 
N
h a t  sj
X
h
a
t

s
j
41  exp
A5
erfc@
Tw  Ti
ks
ks2
j1


 DTbj;j1
(4)
In Eq. (4), the initial temperature of the surface Ti is taken as a
constant value based on thermocouple measurements, and the
wall temperature Tw is known from the calibration of the narrowband liquid crystals (R25C1 from LCR-Hallcrest). The time t to
reach the given wall temperature Tw is determined from the continuous recording of the transient test using a Sony XCDSX90CR,
color CCD camera. The bulk air temperature is recorded through
the transient test, so the bulk temperature change (DTb) and time
Transactions of the ASME

Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/ on 05/08/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

step changes (s) are also known. With the material properties of
the dimpled surface known (ks and a), the only unknown in Eq.
(4) is the convective heat transfer coefficient. Therefore, Eq. (4)
can be solved iteratively to determine the heat transfer coefficient
distribution on the dimpled surface.
The dimpled surface is painted black and the narrow-band liquid crystals are sprayed on top of the black paint (the transient test
is recorded with the CCD camera mounted above the test section
and the surface is viewed through the smooth, top wall of the
channel). A rectangular copper bar is also painted with the black
and liquid crystal paints for calibration. Fluorescent lights are
arranged over the channel, and, with the lights positioned as they
are over the channel, the calibration piece is centered below the
camera and lights. The copper bar is heated, and the color change
of the TLC is recorded. In addition, the temperature change of the
copper bar is recorded with type-T thermocouples mounted inside
the copper. A relation is obtained between the temperature and the
hue of the TLC. From this calibration, the temperature corresponding to the color of maximum intensity can be determined.
For the current paint and lighting conditions, this temperature was
determined to be 25.54  C, and this temperature became the wall
temperature (a constant value) in Eq. (4).
For the actual transient test, the air is diverted away from the
rectangular channel using a three-way diverter valve. A 2 kW
pipe heater is used to heat the air to a desired temperature, and the
air is diverted away from the test section until this temperature is
reached. The transient test is initiated when the air is diverted to
the test section, and the response of the TLC on the dimpled test
section is recorded. Depending on the channel Reynolds number,
images are recorded at either 7.5 frames per second (Re 10,000
and 20,000) or 15 frames per second (Re 30,000 and 40,000).
At the lower Reynolds numbers, lower heat transfer coefficients
are measured, and the transient test requires more time for the
color change to occur. However, at high Reynolds numbers, the
color change can occur very quickly (less than 5 s with the highest
heat transfer coefficients), so the frame rate is increased to reduce
the uncertainty in determining the time to reach the desired wall
temperature. In all cases, the transient tests are completed in less
than 75 s to ensure the semi-infinite solid assumption is not
violated.
Transient Temperature Sensitive Paint Technique. The
temperature sensitive paint used in the current investigation is
comprised of photoluminescent molecules suspended in a polymer
binder [17]. The intensity of light emitted by these fluorescent
molecules is proportional to the temperature of the molecules
[18]. Therefore, using the TSP, it is possible to obtain a detailed
surface temperature distribution.
The theory for the transient TSP test is the same as for the transient TLC test. The experiment is based on a transient, onedimensional semi-infinite solid model. Therefore, a detailed heat
transfer coefficient distribution can be obtained from the TSP
using Eq. (4). The difference between the TSP and TLC experiments lies in what values are held constant in Eq. (4). As presented for the liquid crystal test, the wall temperature is a known,
constant value, and the time to reach this temperature varies from
pixel-to-pixel. For the TSP experiment, the surface intensity is
recorded as a function time. At a given instant in the time, the
detailed surface temperature distribution is calculated. At this
instant in time (t remains constant for every pixel), the heat transfer coefficients are calculated based on the spatial temperature distribution. Therefore, for the TSP test, the wall temperature varies
across the plate at a given time. This experimental approach is
similar to the approach described by Ekkad et al. [19] and Gao
et al. [20] for a transient IR experiment.
As a detailed temperature distribution of the surface is required
to solve Eq. (4) for the heat transfer coefficient distribution, it is
necessary to obtain a relationship of the intensity emitted by the
paint and the temperature of the paint. As the dimpled surface is
Journal of Turbomachinery

Fig. 1

TSP calibration curve

painted with a white base coat and TSP (UniCoat TSP from Innovative Scientific Solutions Inc. (ISSI)), a copper bar, instrumented
with type-T thermocouples is also painted. The copper bar is
placed under a 14-bit, monochrome CCD camera (with a 570 nm
optical filter) and 400 nm, LED array. As required during the transient test, a black and reference image is recorded prior to recording images for the TSP calibration. The black image removes
any background noise present with the optical components, and
the reference image is recorded to provide a reference condition
for normalizing both the calibration and transient test data [18].
The copper bar is then heated, and the intensity change of the TSP
is recorded along with the temperature change of the copper bar
from the thermocouple imbedded in the copper. Figure 1 shows
the relationship of the measured TSP intensity and copper temperature. As shown from the intensity and temperature ratios, as the
temperature of the paint increases, the intensity emitted by the
paint decreases (TSP is a temperature quenched paint). From
this relationship, a curve fit is used to relate the known temperature to the measured intensity. The relationship shown in Eq. (5)
can be used to determine the detailed temperature distribution on
the dimpled surface at any given time.






T
I  Ib 3
I  Ib 2
I  Ib 1
C3
C2
C1
C0
Tref
Iref  Ib
Iref  Ib
Iref  Ib
(5)
The transient TSP experiment is similar to the transient TLC
experiment. The air is heated to the desired temperature using the
pipe heater, and the air is diverted away from the test section until
the desired temperature is reached. Prior to diverting the air to the
channel, a black image is recorded (no LED illumination) and a
reference image is recorded (LED is on, no flow through the channel). Also, the intensity of the surface is recorded for the calculation of the initial temperature. The variation of the surface
temperature is less than 2% over the entire region of interest for
the dimpled channel. Finally, the transient test is initiated with the
hot air diverted to the channel. Images of the dimpled surface are
recorded at a rate of 5 Hz for 60 s. For the current test, the surface
temperature is calculated at 30 s, and this duration for each Reynolds number is used to calculate the detailed heat transfer coefficients. To solve for heat transfer coefficients at each pixel, Eq. (4)
is iteratively solved with a convergence criteria of 105. While
30 s data is presented throughout the paper, heat transfer coefficients were also calculated at other times (typically 15 and 45 s)
to ensure the calculated heat transfer coefficients are independent
of time.

Experimental Facilities
The experimental facility consists of a one-pass, rectangular
channel (AR 3:1) modified from that of Wright and Gohardani
JANUARY 2013, Vol. 135 / 011028-3

Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/ on 05/08/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 2

Overview of experimental setup

[21]. The test section consists of a rectangular, smooth entrance


section. As shown in Fig. 2, the lengths of both the smooth entrance and the dimpled channel are greater than ten hydraulic
diameters. With the addition of flow straightening screens in the
entrance section of the channel, the coolant should be hydrodynamically developed upon entering the dimpled roughened test
section. The cross-section is maintained constant through both the
entrance section and the dimpled channel providing the aspect
ratio 3:1. All walls of both the entrance section and the dimpled
section are fabricated from Plexiglas that is at least 1.27 cm thick
(1.91 cm for the entrance and 1.27 cm dimpled section). The combination of the material thickness, material thermal conductivity,
and time of the transient test ensure the 1D semi-infinite solid
assumption is maintained for all tests.
A single wide wall of the dimpled test section is lined with
three rows of 22 inline V-shaped dimples. Figure 3 illustrates
the geometry of the test section. Note that the optical region of interest is located in the latter half of the channel. This ensures the
turbulent flow is fully developed in the region where the measurements are recoded. Two type-T thermocouples are placed in the
center of the channel (immediately upstream and downstream of
the region of interest) to measure the bulk temperature of the cooling air. The region of interest is located where the flow is known
to be thermally fully developed and free of end effects (8.5  x/
Dh  10.1). At this location, the change in the streamwise temperature between the two thermocouples is less than 1  C; therefore,
the two temperature measurements are averaged and assumed to
be the bulk temperature for the entire region of interest. A single
static pressure tap is located near the exit of the entrance channel.
This measured pressure, along with the exit pressure at the outlet
of the channel (atmospheric pressure), provides a basis for the calculation of the frictional losses created by the V-shaped dimples.
To investigate the effect of coolant Reynolds number on the
heat transfer enhancement within the rectangular channel, the
mass flow rate of the coolant is varied. The mass flow rate of the
air is measured with an ASME square edge flow meter placed
upstream of the pipe heater. The flow is rate is varied to achieve
the desired coolant Reynolds numbers of 10,000, 20,000, 30,000,
and 40,000, based on the hydraulic diameter (Dh 4.0 cm) of the
channel.

Fig. 3

011028-4 / Vol. 135, JANUARY 2013

Fig. 4 V-shaped dimple details

V-Shaped Dimple Configuration. The V-shaped dimple array


is derived from a traditional staggered hemispherical dimple
array. Figure 4 shows the derived V-shaped dimple array along
with the original array of hemispherical dimples (outlined in
white). The baseline hemispherical array is based on that from
Griffith et al. [12]. The dimple diameter is 0.635 cm, and the depth
of the dimples is d/D 0.3. Both the spanwise and streamwise
spacing of the hemispherical dimples was Sxd/D Syd/D 0.8.
A V-shaped dimple is created by connecting three hemispherical dimples. Therefore, the staggered array of hemispherical dimples leads to an inline array of V-shaped dimples. The depth of
the V-shaped dimples remains d/D 0.3; however, the spacing of
the dimples becomes Sx/D Sy/D 3.2. As with the work from
Griffith et al. [12], the V-shaped dimples were created with a
0.635 cm (0.25 in.) ball mill.

Data Reduction
To evaluate the benefit of the proposed V-shaped concavities, it
is necessary to consider both the heat transfer enhancement
afforded by the dimple geometry, as well as the frictional losses
incurred by the dimples.
Heat Transfer Enhancement. Detailed heat transfer coefficient distributions are obtained using both the transient liquid
crystal and transient TSP tests described above. Equation (6)
shows how the measured heat transfer coefficients can be nondimensionalized to form detailed Nusselt number distributions.
Nu

hDh
kf

(6)

To evaluate the heat transfer performance of the proposed


V-shaped dimples against other heat transfer enhancement

Dimple roughened channel details

Transactions of the ASME

Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/ on 05/08/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

mechanisms, the Nusselt number is normalized with the Nusselt


number expected for fully developed, turbulent flow in a smooth
tube. As shown with the Nusselt number ratio in Eq. (7), the calculated Nusselt numbers are compared to those predicted by the
DittusBoelter correlation, and the heat transfer enhancement
resulting from the V-shaped dimples is presented in terms of this
Nusselt number ratio.
Nu
hDh
1


Nuo
kf 0:023Re0:8 Pr0:4

(7)

As two different measurement techniques are considered for the


acquisition of detailed heat transfer coefficient distributions, it is
necessary to consider the experimental uncertainty of both experimental methods. While both the TLC and TSP techniques are
based on a transient test, the experimental uncertainty resulting
from these tests is dominated by different factors. With the transient liquid crystal test, the surface temperature of the dimpled
surface is known, and the time required to reach this temperature
is measured (from the frame rate measurement of the camera). In
regions of very high heat transfer coefficients, the transient test
occurs very quickly. However, in regions of low heat transfer
coefficients, the time required to achieve the color change is significantly longer. Therefore, the time measurement dominates the
accuracy of the TLC calculation (with the uncertainty increasing
as the heat transfer coefficient increases). With the TSP tests, the
time is held constant, and a sufficiently large time is selected to
minimize error associated with this measurement. However, at a
given time, the temperature across the surface varies. In regions of
low heat transfer, the wall temperature has a limited increase over
the initial temperature of the surface. In contrast to the TLC
experiments, the uncertainty with the TSP method increases as the
heat transfer coefficients decrease. Coupling the method proposed
by Camci et al. [22] with that from Kline and McClintock [23]
indicates the maximum experimental uncertainty in the calculated
Nusselt number ratios (on the dimpled surface) for the TLC
experiments is approximately 12%. The calculated uncertainty
with the TSP experiments is significantly higher, and can exceed
25% in regions of relatively low heat transfer enhancement.
Frictional Losses. While it is necessary to know how the dimples increase the convective heat transfer within the rectangular
channel, it is also necessary to know the cost of the heat transfer
enhancement. This cost is evaluated in terms of the pressure penalty incurred through the dimple roughened channel. The gage
pressure at the inlet of the roughened test section is measured
using a 1 psid pressure scanner from Scanivalve Corp. With the
pressure differential from the inlet to the outlet of the dimple
roughened section, the friction factor for the channel can be calculated from Eq. (8).
f

Pin  Pout


L 1
qf V 2
4
Dh 2

(8)

As with the heat transfer enhancement, it is beneficial to compare


the measured friction factor within the dimpled channel to that in
a smooth tube. Therefore, the friction factor can be compared to
that predicted by the Blasius equation for fully developed
flow through a smooth tube. The friction factor ratio is shown in
Eq. (9).
f
f

fo 0:079Re0:25

(9)

According to the method described by Kline and McClintock


[23], the experimental uncertainty in the friction factor ratio is
less than 15% for Reynolds numbers of 30,000 and 40,000. HowJournal of Turbomachinery

ever, at the lower Reynolds numbers, the inlet pressure to the


channel decreases, and the experimental uncertainty in the friction
factor ratio can exceed 30%.
Thermal Performance. Based on the heat transfer enhancement (Eq. (7)) and the frictional losses (Eq. (10)), the overall thermal performance of the proposed V-shaped dimple geometry can
be considered. Based on the constant pumping power condition
used by Han et al. [24], the thermal performance of the dimple geometry can be calculated.


Nu
Nu
(10)
g  01=3
f
f0
Combining the experimental uncertainties from the transient liquid tests with those from the pressure drop, it is possible to estimate the experimental uncertainty of the overall thermal
performance. Over the range of Reynolds numbers considered, the
average experimental uncertainty is estimated to be 11.5% of the
measured value.

Results and Discussion


The present study yields interesting trends associated with
increasing Reynolds number in a channel roughened with
V-shaped dimples. The following includes the presentation of
results obtained using the TLC technique, a brief discussion of
flow characteristics over the range of Reynolds numbers considered, a comparison of the results obtained using the TLC technique to that of the TSP technique, and a discussion of the thermal
performance of V-shaped dimples.
Heat Transfer Enhancement. A comparison of detailed
Nusselt number ratio distributions obtained using both transient
techniques is provided in Fig. 5. In each image, the air flows from
left to right, and each dimple print is overlaid on the distributions.
For both techniques at a specific Reynolds number, the measured
Nusselt number distributions are uniform in both the streamwise
and spanwise directions. As the Nusselt numbers are not changing
in the streamwise direction, the turbulent flow can be considered
fully developed across the width of the channels. For the lowest
Reynolds numbers of 10,000 and 20,000, the heat transfer
enhancement is very similar to that of dimples. Inside the dimple,
the flow separates causing very low Nusselt number ratios (less
than one inside the dimples). Most of the heat transfer enhancement occurs outside of the dimple where the flow reattaches to the
surface (similar to the trends reported for traditional hemispherical
dimples). However, at Reynolds numbers greater than 30,000,
overall Nusselt number ratios increase significantly, as shown by
both experimental techniques (Nusselt number ratios for traditional hemispherical dimples are relatively invariant to Reynolds
number).
To better visualize the changes in heat transfer enhancement,
detailed Nussselt number ratio distributions for the center dimple
obtained using the TLC technique at Reynolds numbers of 10,000
and 40,000 are shown in Fig. 6 (note the scale has been increased
from Fig. 5). Again, at low Reynolds numbers most of the heat
transfer enhancement occurs outside of the dimple as the coolant
reattaches to the surface downstream of the dimple. While at high
Reynolds numbers, there is increased heat transfer inside the dimple. Not only are elevated heat transfer coefficients present in the
apex of the V (where the coolant is impinging within the concavity), but the heat transfer coefficients are increasing inside the legs
of the V, moving in the spanwise direction across the dimple.
The increased Nusselt number ratios are thought to be the result
of the increased momentum associated with the flow at high
JANUARY 2013, Vol. 135 / 011028-5

Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/ on 05/08/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 6 Detailed Nusselt number ratios from TLC

Fig. 7
flow

Fig. 5 Detailed Nusselt number ratio distributions from TLC


((a)(d)) and TSP ((e)(h))

Reynolds numbers. Higher momentum induces strong counterrotating vortices within the secondary flow inside the dimple;
such vortices produce high heat transfer coefficients along the
trailing edge and downstream of the dimple. Three pairs of counter-rotating vortices are formed inside the channel as a result of
the corresponding three rows of dimples. Similar vortical structures are thought to increase heat transfer along the legs of Vshaped ribs [25]. Along with increasing heat transfer enhancement
downstream of each dimple, the dimple-induced secondary flows
increase Nusselt number ratios inside the dimple as well as laterally, between each row of dimples where the fluid is driven by the
angled legs of the dimples. The counter-rotating vortices, in
effect, pull the flow into the dimple, and cause the flow to reattach
inside of the downstream edge of the dimple. Such rotation also
causes the flow to reverse as it enters and flows through each leg
of the dimple; this recirculation causes a small area of high heat
transfer on the leading edge of the dimples. A conceptual view of
the dimple induced secondary flow is shown in Fig. 7.
For the TLC technique, the streamwise centerline (z/W 0)
Nusselt number ratios for each case are compared in Fig. 8. The
highest Nusselt number ratios occur at the trailing edge of each
dimple. At the relatively high Reynolds numbers (30,000 and
011028-6 / Vol. 135, JANUARY 2013

Conceptual view of V-shaped dimple induced secondary

40,000), note the increased slope of the Nusselt number ratios


along the centerline as the flow leaves the dimple. These increased
Nusselt numbers along the centerline are the result of the coolant
impingement occurring inside the dimple. Figure 8 also shows the
Nusselt numbers reduce to half of those predicted inside a smooth
channel. At all Reynolds numbers, within the upstream half of the
dimple, the separation that forms and traps a pocket of air within
the dimple results in decreased heat transfer within this region.
However, this reduced heat transfer is balanced by the enhanced
heat transfer between the dimples. For all Reynolds numbers, the
Nusselt numbers on the flat surface are approximately two times
those of a traditional smooth channel.
Spanwise Nusselt number ratio distributions for the center
dimple at x 9.4Dh are given in Fig. 9. These spanwise distributions correspond to the downstream edge of the V-shaped dimple.
As expected, the Nusselt number ratios are symmetrical about the
centerline (z/W 0) of the dimple, and on each half of the dimple,
two local maxima are present at each Reynolds number. The
absolute maximum occurring at z/W 60.04 corresponds to the
downstream edge of the leg of the V. As this is the true edge of
the V-shaped concavity, it corresponds to location of the
maximum Nusselt number ratios. The secondary maximum occurs
at z/W 60.08. At this location (at the outermost position along
the dimple), the magnitude of the Nusselt number ratio is strongly
affected by Reynolds number. With the intense vortices
created by the V-shaped dimples at elevated Reynolds numbers,
Transactions of the ASME

Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/ on 05/08/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 8

Centerline Nusselt number distributions

Fig. 9 Spanwise Nusselt number ratio distributions at trailing


edge of V-shaped dimple

enhanced heat transfer coefficients result along the leg of the


dimple.
The Reynolds number dependency of heat transfer enhancement or degradation inside of the dimple is more easily seen in
Fig. 10; here, spanwise Nusselt number ratio distributions are
shown at the leading edge (x/Dh 9.15), the center (x/Dh 9.28),
and the trailing edge (x/Dh 9.40) of the center dimple. At the
leading edge of the dimple (x/Dh 9.15), the level of heat transfer
enhancement is independent of Reynolds number. Here the Nusselt numbers are dominated by the flow patterns apparent between
two dimples, and in all cases the level of enhancement is
approaching two times that of a smooth channel. The effect of
Reynolds number is most obviously observed in the center of the
dimple (x/Dh 9.28). At the lower Reynolds numbers of 10,000
and 20,000, the Nusselt number ratios are significantly less than
unity. This indicates the dominant secondary flow feature inside
the dimple is the traditional separation occurring at the leading
edge of the dimple. However, increasing the Reynolds number to
30,000 and 40,000, the Nusselt numbers inside the dimples begin
to increase significantly. Regions of high heat transfer develop at
Journal of Turbomachinery

Fig. 10 Effect of streamwise location on the spanwise Nusselt


number ratio

z/W 0 (centerline) and 0.075 (outer edge of dimple). These


regions of increased heat transfer support the theory of counterrotating vortices forming in the legs of the V-shaped dimple.
Experimental Technique Comparison. The area averaged
Nusselt number ratios for both experimental techniques are presented in Fig. 11. Included in the figure is a baseline case where
the experimental setup is validated (for both the TSP and TLC
techniques) with a smooth channel. As expected, each technique
yields unity Nusselt number ratios for the smooth channel.
For the V-shaped dimples, each technique yields increasing
Nusselt number ratios with increasing Reynolds number; however, the TSP technique yields lower overall heat transfer
JANUARY 2013, Vol. 135 / 011028-7

Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/ on 05/08/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Table 1

Geometrical comparison with previous studies

Reference

Dimple shape Aspect ratio Sx/d d/Dh

d/D

Dimples Present study V-shaped


3
10.7 0.047
0.30
5
Hemispherical
4
4 0.157
0.25
5
Teardrop
4
4 0.157
0.25
6
Hemispherical
3.78
5.56 0.083
0.19
Reference Rib geometry Aspect ratio P/e e/Dh Rib angle
Ribs

Fig. 11 Comparison of transient TLC and TSP techniques

Fig. 12 Average Nusselt number ratio comparison with previous studies

enhancement than that of the TLC technique. Recall that the TSP
technique measures an intensity ratio to determine wall temperature. By the nature of this transient technique, the intensity ratio
change, at any time within the bounds of the semi-infinite solid
assumption, is minimal (especially at low Reynolds numbers).
Thus, highly resolved, detailed distributions are difficult to obtain.
As a result, TSP is currently best suited for observing general
trends associated with varying flow conditions, as well as obtaining regionally averaged Nusselt number ratios.
Due to the nature of the TSP, the detailed distributions shown
in Fig. 5 are less resolved (spatially) than those from the TLC.
While both techniques capture the same heat transfer enhancement trends, the traditional TLC data reflects both the extreme
high and extreme low heat transfer coefficients. While the transient TSP has clear drawbacks, with the inclusion of error bars in
Fig. 11, it is obvious the data from the two techniques do overlap.
It is possible, however, that a more sensitive TSP and more
powerful camera-LED array combination could improve the quality of the detailed distributions produced with the TSP technique.
Also, due to the temperature quenching nature of the paint, as the
temperature of the surface increases, the emission intensity of the
paint decreases. Decreasing the magnitude of the measured intensity is a drawback for the current TSP tests. Regardless of the
011028-8 / Vol. 135, JANUARY 2013

21
26

V-shaped
Angled

3
3

7.86 0.08
8 0.08

45 deg
45 deg

lower quality images obtained with the TSP technique, the use of
TSP has multiple advantages; most notable of such advantages is
ability for TSP to accurately measure temperatures over a much
wider range than that of TLC. For transient tests such as this, it is
also beneficial to hold time constant (as opposed to temperature
for TLC). This allows for a detailed temperature distribution at
any time.
Overall Thermal Performance. The average heat transfer
enhancement (TLC) for each case is compared in Fig. 12. Also
included on the plot are corresponding data for existing heat transfer enhancement configurations. Geometrical details of each comparison are specified in Table 1. The comparison data of Wright
and Gohardani [21,26] is calculated using a similar channel with
one wide wall roughened with V-shaped or angled rib turbulators.
The results obtained by Moon et al. [6] in a rectangular channel
where one wide wall is roughened with hemispherical dimples are
slightly lower than that of Chyu et al. [5] where both wide walls
of a rectangular channel are roughened with either hemispherical
or teardrop shaped dimples. The addition of dimples on two wide
walls of the channel increases the turbulent mixing inside the
channel. This results in relatively higher heat transfer enhancement within the roughened passage. V-shaped dimples appear to
increase heat transfer enhancement with increasing Reynolds
numbers. Such heat transfer enhancement trends are vastly different from hemispherical dimples, which are invariant to Reynolds
number, and rib turbulators, which decrease with increasing Reynolds number. Generally the V-shaped dimples outperform the traditional hemispherical dimple configurations; in fact, overall heat
transfer enhancement of the V-shaped dimples approaches that of
traditional rib turbulators with increasing Reynolds number.
No cooling scheme can be limited only to comparisons of heat
transfer enhancement. Recall that any form of heat transfer
enhancement comes at the expense of increased pressure loss
through the channel. A comparison of friction factor ratios for
each case is provided in Fig. 13. While V-shaped dimple friction
factor ratios are slightly higher than that of traditional dimples,
they are significantly lower than that of rib turbulators. Comparison of thermal performance to the relative height and spacing of
the ribs and the relative depth and spacing of the V-shaped dimples reveals that the V-shaped dimples have favorable thermal
performance for a larger spacing and smaller depth. This contributes to the significantly lower pressure drop incurred by the dimples. For this experiment, at higher Reynolds numbers, there are
slightly higher friction factor ratios. Again, for the V-shaped dimple configuration, increasing Reynolds number increases the
strength of the counter rotating vortices within the channel.
Increased vortical strength increases overall heat transfer; however, it also increases pressure drop through the test section. From
strict observation of only the overall Nusselt number ratios and
the friction factor ratios, it is unclear how this particular cooling
scheme compares to other cooling schemes.
As a result, frictional losses must be combined with overall
heat transfer enhancement, in the form of thermal performance, to
fully evaluate the V-shaped dimple cooling scheme. A comparison of the thermal performance for this cooling scheme and other
Transactions of the ASME

Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/ on 05/08/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

is the increasing heat transfer enhancement with increasing Reynolds number. As the Reynolds number increases, the heat transfer
enhancement rises due to dimple-induced counter-rotating vortices formed within the legs of the dimples. The extent of the rising
Nusselt number ratios should be further investigated at increased
Reynolds numbers (beyond 40,000). Previous studies show that a
channel utilizing two walls roughed with dimples will produce
higher heat transfer enhancement than that of one roughened wall;
as a result, future studies should examine the effect of roughening
both wide walls of the channel with V-shaped dimples.
Two transient techniquesthermochromatic liquid crystals and
temperature sensitive paintwere used to generate detailed Nusselt Number ratio distributions. The temperature sensitive paint
technique yields less resolved detailed distributions than that of its
liquid crystal counterpart. Although the magnitude of the TSP
varies from the TLC, the TSP does an adequate job of identifying
the areas of both high and low heat transfer enhancement. As
more researchers consider the use of temperature sensitive paints,
the technique is likely to become a more suitable alternative to
transient liquid crystal and IR tests.

Acknowledgment
Fig. 13

Friction factor ratio comparison with previous studies

The authors appreciate the work of Baylor University students


Evan Martin, Sam Strickling, Shane Wallace, and Doug Wise.
Their creativity and effort have been vital to the completion of
this experimental study.

Nomenclature

Fig. 14 Thermal performance comparison with previous


studies

traditional cooling schemes is provided in Fig. 14. At high Reynolds numbers, the thermal performance of the V-shaped configuration exceeds that of traditional rib turbulators and is comparable
to hemispherical dimple cooling schemes. It is interesting to note
that even as friction factors increase with increasing Reynolds
number, the thermal performance also increases.

Conclusions
The present experimental investigation primarily considers the
effect of increasing Reynolds number on overall heat transfer
enhancement and thermal performance of a V-shaped dimple
cooling scheme in a rectangular channel (AR 3:1). Based on
both the heat transfer enhancement and friction losses measured
in the rectangular channel, the proposed V-shaped dimples offer a
potential alternative to both rib turbulators and hemispherical
dimples. The primary advantage offered by the V-shaped dimples
Journal of Turbomachinery

Ac
AR
C
cp
D
Dh
e
f
fo
h
I
Ib
Iref
j
kf
ks
L
Le
N
Nu
Nu
Nuo
P
Pin
Pout
Pr
Re
Sx
Sxd
Sy
Syd
t
T
Tb
Ti
Tref
Tw
V
W

cross-sectional area of rectangular channel ( wh)


channel aspect ratio ( W/h)
TSP calibration coefficients (C3, C2, C1, C0)
specific heat of test section material
dimple diameter
hydraulic diameter of rectangular channel ( 4Ac/P)
rib height
friction factor
friction factor of fully developed flow in a smooth tube
channel height or heat transfer coefficient
TSP emission intensity
TSP emission intensity of black image
TSP emission intensity of reference image
time step for bulk temperature change
thermal conductivity of air
thermal conductivity of test section material
channel length
smooth entry length of channel
total number of time steps for bulk temperature change
Nusselt number
average Nusselt number over entire region of interest
Nusselt number of fully developed flow in a smooth tube
perimeter of rectangular channel ( 2[w h]) or streamwise spacing of rib turbulators
inlet air pressure
outlet air pressure (atmospheric pressure)
Prandtl number of air
Reynolds number coolant flow through channel
streamwise spacing of V-shaped dimples
streamwise spacing of traditional hemispherical dimples
spanwise spacing of V-shaped dimples
spanwise spacing of traditional hemispherical dimples
time
temperature
bulk temperature of cooling air
initial temperature of heat transfer surface
reference temperature for TSP test and calibration
wall temperature
coolant velocity
width of channel
JANUARY 2013, Vol. 135 / 011028-9

Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/ on 05/08/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

x
y
z
a
d
g
qs
qf
s

streamwise coordinate within channel


coordinate direction through the channel wall
spanwise coordinate within channel
thermal diffusivity of test section material
dimple depth
overall thermal performance
density of test section material
density of air
time step change of bulk fluid temperature

References
[1] Han, J. C., Dutta, S., and Ekkad, S., 2000, Gas Turbine Heat Transfer and Cooling Technology, Taylor and Francis, New York.
[2] Ligrani, P. M., Oliveira, O., and Blaskovich, T., 2003, Comparison of Heat
Transfer Augmentation Techniques, AIAA J., 41, pp. 337361.
[3] Kim, Y. W., Arellano, L., Vardakas, M., Moon, H. K., and Smith, K. O., 2003,
Comparison of Trip-Strip/Impingement/Dimple Cooling Concepts at High
Reynolds Numbers, ASME Paper No. GT2003-38935.
[4] Chyu, M. K., Yu, Y., and Ding, H., 1999, Heat Transfer Enhancement in Rectangular Channels With Concavities, J. Enhanced Heat Transfer, 6(6), pp.
429439.
[5] Chyu, M. K., Yu, Y., Ding, H., Downs, J. P., and Soechting, F. O., 1997,
Concavity Enhanced Heat Transfer in an Internal Cooling Passage, ASME
Paper No. 97-GT-437.
[6] Moon, H. K., OConnell, T., and Glezer, B., 2000, Channel Height Effect on
Heat Transfer and Friction in a Dimpled Passage, ASME J. Eng. Gas Turbines
Power, 122, pp. 307313.
[7] Lin, Y. L., Shih, T. I-P., and Chyu, M. K., 1999, Computations of Flow and
Heat Transfer in a Channel with Rows of Hemispherical Cavities, ASME Paper No. 99-GT-263.
[8] Mahmood, G. I., Hill, H. L., Nelson, D. L., Ligrani, P. M., Moon, H. K., and
Glezer, B., 2001, Local Heat Transfer and Flow Structure on and Above a
Dimpled Surface in a Channel, ASME J. Turbomach., 123, pp. 115123.
[9] Mahmood, G. I., and Ligrani, P. M., 2002, Heat Transfer in a Dimpled Channel: Combined Influences of Aspect Ratio, Temperature Ratio, Reynolds Number, and Flow Structure, Int. J. Heat Mass Transfer, 45, pp. 20112020.
[10] Burgess, N. K., Oliveria, M. M., and Ligrani, P. M., 2003, Nusselt Number
Behavior on Deep Dimpled Surfaces Within a Channel, ASME J. Heat Transfer, 125, pp. 1118.
[11] Ligrani, P. M., Burgess, N. K., and Won, S. Y., 2004, Nusselt Numbers and
Flow Structure On and Above a Shallow Dimpled Surface Within a Channel

011028-10 / Vol. 135, JANUARY 2013

[12]

[13]

[14]

[15]

[16]

[17]
[18]

[19]

[20]

[21]

[22]

[23]
[24]

[25]
[26]

Including Effects of Inlet Turbulence Intensity Level, ASME Paper No.


GT2004-54231.
Griffith, T. S., Al-Hadhrami, L., and Han, J. C., 2003, Heat Transfer in Rotating Rectangular Cooling Channels (AR 4) With Dimples, ASME J. Turbomach., 125, pp. 555564.
Moon, S. W., and Lau, S. C., 2002, Turbulent Heat Transfer Measurements on
a Wall With Concave and Cylindrical Dimples in a Square Channel, ASME
Paper No. GT2002-30208.
Borisov, I., Khalatov, A., Kobzar, S., and Glezer, B., 2004, Comparison of
Thermo-Hydraulic Characteristics for Two Types of Dimpled Surfaces,
ASME Paper No. GT2004-54204.
Zhou, F., and Acharya, S., 2009, Experimental and Computational Study of
Heat / Mass Transfer and Flow Structure for Four Dimple Shapes in a Square
Internal Passage, ASME Paper No. GT2009-60240.
Ekkad, S. V., and Han, J. C., 2000, A Transient Liquid Crystal Thermography
Technique for Gas Turbine Heat Transfer Measurements, Meas. Sci. Technol.,
11, pp. 957968.
Liu, T., and Sullivan, J. P., 2005, Pressure and Temperature Sensitive Paints,
Springer, Berlin.
Wright, L. M., Gao, Z., Varvel, T. A., and Han, J. C., 2005, Assessment of
Steady State PSP, TSP, and IR Measurement Techniques for Flat Plate Film
Cooling, ASME Paper No. HT2005-72363.
Ekkad, S. V., Ou, S., and Rivir, R. B., 2004 A Transient Infrared Thermography Method for Simultaneous Film Cooling Effectiveness and Heat Transfer
Coefficient Measurements From a Single Test, ASME J. Turbomach., 126, pp.
597603.
Gao, Z., Wright, L. M., and Han, J. C., 2005, Assessment of Steady State PSP
and Transient IR Measurement Techniques for Leading Edge Film Cooling,
ASME Paper No. IMECE2005-80146.
Wright, L. M., and Gohardani, A. S., 2009, Effect of Coolant Ejection in Rectangular and Trapezoidal Trailing Edge Cooling Passages, AIAA J. Thermophys. Heat Transfer, 23(2), pp. 316326.
Camci, C., Kim, K., and Hippensteele, S. A., 1993, Evaluation of a Hue Capturing Based Transient Liquid Crystal Method for High-Resolution Mapping of
Convective Heat Transfer on Curved Surfaces, ASME J. Heat Transfer,
115(2), pp. 311318.
Kline, S. J., and McClintock, F. A., 1953, Describing Uncertainties in SingleSample Experiments, Mech. Eng. (Am. Soc. Mech. Eng.), 75, pp. 38.
Han, J. C., Park, J. S., and Lei, C. K., 1985, Heat Transfer Enhancement in
Channels With Turbulence Promoters, ASME J. Eng. Gas Turbines Power,
107, pp. 628635.
Tanda, G., 2003, Heat Transfer in Rectangular Channels With Transverse and
V-Shaped Broken Ribs, Int. J. Heat Mass Transfer, 47, pp. 229243.
Wright, L. M., and Gohardani, A. S., 2008, Effect of Turbulator Width and
Spacing on the Thermal Performance of Angled Ribs in a Rectangular Channel
(AR 3:1), ASME Paper No. IMECE2008-66842.

Transactions of the ASME

Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/ on 05/08/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

You might also like