You are on page 1of 11

International Journal of Heat and Mass Transfer 138 (2019) 293–303

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


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

Fluid flow and heat transfer characteristics of heat sinks with laterally
perforated plate fins
Sakkarin Chingulpitak a,b, Ho Seon Ahn c, Lazarus Godson Asirvatham d, Somchai Wongwises b,e,⇑
a
The Joint Graduate School of Energy and Environment, King Mongkut’s University of Technology Thonburi, Bangmod, Bangkok 10140, Thailand
b
Fluid Mechanics, Thermal Engineering and Multiphase Flow Research Lab. (FUTURE), Department of Mechanical Engineering, Faculty of Engineering, King Mongkut’s University
of Technology Thonburi, Bangmod, Bangkok 10140, Thailand
c
Department of Mechanical Engineering, Incheon National University, Incheon, Republic of Korea
d
Department of Mechanical Engineering, Karunya Institute of Technology and Sciences, Coimbatore, Tamil Nadu, India
e
The Academy of Science, The Royal Society of Thailand, Sanam Suea Pa, Dusit, Bangkok 10300, Thailand

a r t i c l e i n f o a b s t r a c t

Article history: This study presents the thermal performance of laterally perforated plate-fin heat sinks (LAP-PFHSs) with
Received 23 September 2018 different numbers (Np) and diameters (Dp) of circular perforations. Based on the same plate fin dimension,
Received in revised form 23 March 2019 the available diameter of circular perforations is limited by their number. The largest diameters of per-
Accepted 5 April 2019
foration for Np = 14, 27 and 75 are 10 mm, 7 mm and 4 mm, respectively. The computational results of
solid fin heat sinks (SFHSs) and LAP-PFHSs are validated with the measured data and experimental data
from the available literature. The comparison results of LAP-PFHSs give the mean absolute error of about
Keywords:
3.6% and 5.3% for the pressure drop and thermal resistance, respectively. According to the numerical
Heat sink
Parallel
results, the LAP-PFHS with Dp = 3 mm and Np = 75 exhibits the highest heat transfer rate, about 11.6%
Perforation higher than that of the SFHS. Finally, the thermal performance factor, which is a consideration of the
Non-bypass flow Nusselt number and friction factor, is proposed to find the suitable design parameters. Under the same
Plate fin conditions, the optimized LAP-PFHS demonstrates 10.6% greater thermal performance and 28% lower vol-
ume of heat sink material than SFHS.
Ó 2019 Published by Elsevier Ltd.

1. Introduction studies on the effects of flow behavior and direction on the thermal
performance of conventional heat sinks. They collected literature
The recent trend in the development of electronic devices is that studied vortex generators added to the fins, vortex generators
moving to more compact or higher performance devices. Therefore, installed at the front of heat sinks and bypass flow. Kim and Kim
an electronic cooling system, comprising moving working fluid and [10] experimentally studied the thermal performance of square
a heat exchanger, is one of the most important things to ensure the pin-fin, plate-fin and cross-cut heat sinks. It was conducted under
extended lifetime of an electronic device. The heat sink is a simple the condition of parallel flow through heat sinks. The results
heat exchanger that can be separated into two types according to demonstrated that the thermal performance of single cross-cut fins
fin shape: plate fin and pin fin. Over the past few decades, some was better than that of multiple cross-cut fin. Finally, the authors
researchers have presented studies on the thermal performance proposed the friction factor and Nusselt number correlations for
of conventional plate-fin heat sinks to find the optimum plate-fin a single cross-cut heat sink. Chingulpitak et al. [11] experimentally
thickness [1–3]. Researchers have studied the integration of the and numerically investigated the optimal number and length of
volume of the heat sink as a constraint parameter [4–8]. cross-cuts. The results demonstrated that the optimal design
Moreover, some researchers have presented the heat transfer parameters were Nc = 6 and Lc = 1.5 mm. The cross-cut heat sink
improvement of plate-fin heat sinks by modification of fin geome- exhibited lower thermal resistance than the plate-fin heat sink
try. In 2015, Chingulpitak and Wonwises [9] summarized the by about 16.2%. Kim [12] presented a numerical model to find
the optimal design parameters of a branched plate-fin heat sink.
⇑ Corresponding author at: Fluid Mechanics, Thermal Engineering and The numerical model was developed using the volume averaging
Multiphase Flow Research Lab. (FUTURE), Department of Mechanical Engineering, theory (VAT). At the same base width, base length and fin height,
Faculty of Engineering, King Mongkut’s University of Technology Thonburi, the optimized branched-fin heat sink exhibited lower thermal
Bangmod, Bangkok 10140, Thailand. resistance than that of the optimized rectangular-fin heat sink by
E-mail address: somchai.won@kmutt.ac.th (S. Wongwises).

https://doi.org/10.1016/j.ijheatmasstransfer.2019.04.027
0017-9310/Ó 2019 Published by Elsevier Ltd.
294 S. Chingulpitak et al. / International Journal of Heat and Mass Transfer 138 (2019) 293–303

Nomenclature

cp specific heat (kJ/kg°C)  x, y, z Cartesian coordinates


Dh hydraulic diameter (m), Dh ¼ 2Hf W ch = Hf þ W ch
Dp perforation diameter (m) Greek letters
H height (m) D differential
k thermal conductivity (W/mK) l viscosity (kg/ms)
L length (m) q density (kg/m3)
m_ mass flow rate (kg/s) k thermal conductivity (W/m2°C)
Np total numbers of perforation / porosity
NR numbers of perforation row
P pressure (Pa) Subscripts
DP pressure drop (Pa)
a air
Q heat transfer rate (W) avg average value
Re Reynolds number b base of heat sink
Rth thermal resistance (K/W) ch channel
t time (s)
conv convection
tf fin thickness (m) f fin
T temperature (°C) in inlet
V velocity (m/s)
out outlet
U velocity vector p perforation
u, v, w velocity components (m/s) s solid fin
W width (m)

about 30%. Ahmed [13] numerically studied the effect of inserting plate-fin heat sink. Al-Sallami et al. [21] proposed the heat transfer
ribs into the flow passage of a plate-fin heat sink on the thermal and pressure drop of plate-fin heat sinks with a longitudinal notch,
performance. The ribbed plate-fin heat sink (RPFHS) was investi- rectangular perforation and multiple circular perforations. They
gated according to several parameters: the number, height and concluded that the notch and slot perforations resulted in better
position of ribs. The comparison results indicated that the RPFHS heat transfer and pressure drop than the circular perforations.
provided a Nusselt number ratio 1.55 times greater than the After that, Shaeri et al. [22] and Shaeri and Bonner [23,24] pre-
plate-fin heat sink under the same conditions. Subasi et al. [14] sented experimental and numerical studies of laterally perforated
showed numerical and experimental studies of aluminum heat plate-fin heat sinks (LAP-PFHSs). The heat transfer and pressure
sinks with hexagonal honeycomb fins. They presented numerical drop of heat sinks were presented with various numbers and
methods to find the optimal design parameters of the honeycomb dimensions of rectangular perforation.
fins. The variable parameters were fin thickness, fin height, attack Shaeri and Bonner [23] experimentally investigated the thermal
angle, and longitudinal pitch. The developed model was presented performance of laterally perforated-finned heat sinks (LA-PFHSs).
to determine optimum design parameters by considering at maxi- The air flow was assumed to be laminar flow and non-bypass flow.
mum the Nusselt number and at minimum the friction factor. The rectangular perforation on the plate-fin heat sink was imple-
Wang et al. [15] experimentally studied the effect of vortex gener- mented with three different perforation sizes, which were based
ators on the air-side performance of a plate-fin heat sink. The on five porosities. They proposed mass-based thermal resistance
results showed that the optimal vortex generator for heat sink as a performance indicator. It was calculated by multiplying the
was triangular attack fin. Moreover, the vortex generators were thermal resistance and mass of the heat sink. The results showed
also applied in the air-side performance improvement of the fin- that LA-PFHSs with the maximum porosity provided 45% lower
and-tube heat exchanger which was reviewed by Chimres and mass-based thermal resistance than conventional plate-fin heat
Wongwises [16]. In 2018, Chimres et al. [17,18] proposed numeri- sinks. Ismail et al. [25] performed a numerical study on the heat
cal investigations on the thermal performance of fin and tube heat transfer of plate fin with square, circular, triangular and hexagonal
exchangers with vortex generators. The comparison results perforations. The thermal performance of heat sinks was investi-
between semi-dimple and elliptical winglet vortex generators gated using a turbulent flow regime and bypass flow. They con-
and plain fins were presented. The results showed that the heat cluded that the plate-fin heat sink with hexagonal perforations
transfer coefficient of the vortex generator with semi-dimple was exhibited the highest fin effectiveness compared with heat sinks
similar to that of the vortex generator with elliptical winglet. How- with other perforation shapes. Dhanawade and Dhanawade [26]
ever, the elliptical winglet vortex generator provided the pressure and Dhanawade et al. [27,28] reported an experimental study on
drop higher than the semi-dimple vortex generator. They found the heat transfer of the plate-fin heat sinks with lateral square
that the semi-dimple vortex generator had a higher goodness fac- and circular perforations. The experimental study was conducted
tor than the plain fin by 15–20%. under the conditions of turbulent flow and bypass flow. At the
Regarding other fin designs, some researchers studied the ther- same perforation size, the results demonstrated that the fin effec-
mal performance of plate-fin heat sinks with perforation along the tiveness of the square perforated fin was higher than that of the
fin’s length [19–21] or lateral perforation [22–29]. Shaeri and circular perforated fin.
Yaghoubi [19,20] presented the numerical study of plate-fin heat Ibrahim et al. [29] carried out an experiment and simulation to
sinks with rectangular perforations along the fin’s length. The ther- study the effect of LAP-PFHSs on the heat transfer characteristic.
mal performance of heat sinks was investigated under the assump- The air flow characteristic on the fin tip was assumed to be bypass
tions of laminar and turbulence fluid flows and bypass flow. The flow. The laterally perforated plate fin was investigated with circu-
comparison results indicated that the Nusselt number of the con- lar, rectangular and triangular shapes. The highest temperature dif-
ventional plate-fin heat sink was higher than that of the perforated ference between the heat sink base and fin tip with circular,
S. Chingulpitak et al. / International Journal of Heat and Mass Transfer 138 (2019) 293–303 295

rectangular and triangular perforations was 51.29%, 45.57% and The thermal resistance (Rth) of the air flowing through the SFHS
42.28%, respectively. can be determined by
As mentioned above, the trend in thermal performance
improvement of the plate-fin heat sink is focused on the develop- ðT b;av g  T a;av g Þ
Rth ¼ ð2Þ
ment of newly designed fins. However, there are only a few articles Qa
on the thermal performance of plate-fin heat sinks with circular
perforations, which is especially based on force convection, turbu- where Tb,avg represents the average temperature at the heat sink
lent flow and non-bypass flow. For this reason, this article aims to base and Ta,avg represents the average air temperature. The Reynolds
study the thermal performance of LAP-PFHSs under a fixed area for number (Re) can be calculated by
perforation. The position of perforations is distributed equally over
the fin surface. Through numerical investigation, the air flow qV ch Dh
Re ¼ ð3Þ
through solid fin heat sinks (SFHSs) and LAP-PFHSs are considered l
based on parallel flow direction and non-bypass flow. The ratio of
the Nusselt number of perforated plate fins and solid fins to the where V means the velocity of the air in a channel and Dh means the
friction factor of perforated plate fins and solid fins hydraulic diameter of the channel. The hydraulic diameter of the
ðNu=Nus Þ=ðf =f s Þ is used to indicate the optimal LAP-PFHS with cir- channel can be calculated as follows:
cular perforation.
4Ach
Dh ¼ ð4Þ
Pch
2. Experimental apparatus and procedure
where Ach represents the cross-sectional area of the flow channel
The experimental apparatus is set up to study the heat transfer and P represents the wetted perimeter of the channel.
and flow characteristics of SFHSs. A schematic diagram of the
experimental apparatus is presented in Fig. 1. It consists of three
main sections, which are a wind tunnel, test section, and experi-
mental measurement. The air flow inside the wind tunnel is driven
Wch
by an axial fan controlled by an inverter. After that, the air flows tf
through the straightener and the test section. The hot wire
anemometer is used to measure the air velocity which is varied
between 1 and 4 m/s. The test section is made from aluminum
with a dimension of base width 27 mm, base length 75 mm, fin
thickness 1 mm, fin height 25 mm and channel width 3 mm. A
100-watt plate heater with a variable alternating current (AC)
transformer and 20 mm bakelite plate thickness are connected at
the heat sink bottom. At the heat sink base, the temperature is con-
trolled at 40 °C. The digital manometer is used to measure the
pressure drop across the SFHS. An accuracy of the digital manome- Hf
ter is ±0.5% of the full scale. The type T thermocouple, which has an
accuracy of ±0.2 °C, is used to measure the temperatures. More
details of this experimental setup are available from Chingulpitak
Hb
et al. [10]. The heat transfer rate from the SFHS to the air can be
calculated as follows:

_ a cp;a ðT a;av g;out  T a;av g;in Þ


Qa ¼ m ð1Þ

where m _ a means the mass flow rate of air, cp,a means the specific
heat of air and Ta,avg,in and Ta,avg,out mean the average air inlet and
Fig. 2. The geometrical design parameters of solid fin heat sink.
outlet temperatures, respectively.

1030
800
288 230
60 200 200
Insulator Acrylic plate
40 40 100 Fan
T
Straightener
A T T
45
45
35
25

35

98

A
27 Test section
Wind tunnel duct +
ΔP -
Section A-A Variac (Heat flux controller)
Unit: mm

Fig. 1. Schematic diagram of the experimental apparatus.


296 S. Chingulpitak et al. / International Journal of Heat and Mass Transfer 138 (2019) 293–303

The air flow behavior through the plate-fin heat sink at low
frontal velocity is considered turbulent flow. In the past decade,
Dp
most research studies were on the flow characteristic of air flowing
Hf Sy NR through plate-fin heat sinks using the k–e and k–x turbulence
models. The standard k–x model is suitable for predicting the flow
Sx
at the near-wall region. The k–e model is used to determine far-
field free-stream independence. After that, Menter [32] proposed
tb Lb = Lf the shear stress transport (SST) k-x turbulent model, which was
developed by combining the standard k–x model and the k–e
Fig. 3. The schematic view of laterally perforated plate-fin heat sink with circular
model. Therefore, it can be widely used for predicting the flow
perforations.
characteristic of aerodynamic applications. In the present model,
the SST k-x turbulent model is used to predict the air flow charac-
3. Theory teristic along the flow passage, which is in agreement with the
numerical studies of Zhou et al. [33] and Kanargi et al. [34].
3.1. Mathematical model and numerical method
3.2. Computational domain and simulation settings
The heat transfer of the plate-fin heat sinks and three-
dimensional turbulent flow of the air are presented. The steady, To meet the aims of this study, the comparison criteria of plate-
incompressible and three-dimensional flows are considered to be fin heat sinks is considered under the same fin dimensions. Thus,
the assumptions of air flowing through the plate-fin heat sink. the area for circular perforation is 25  75 mm according to the
The mathematical model does not include the buoyancy effect or dimensions of a solid fin. The geometrical parameters of the SFHS
radiation heat transfer. For the steady turbulent flow, the govern- and LAP-PFHS are shown in Fig. 2 and Fig. 3, respectively. To save
ing equations of continuity, momentum and energy [30,31] are computer resources, the SFHS and LAP-PFHS are considered a por-
as follows: tion of the body. The computational domains of the heat sink and
Continuity equation: air are presented in Fig. 4. The height, length and width of the heat
@u @ v @w sink base are 0.5 mm, 75 mm and 4 mm, respectively. The other
þ þ ¼0 ð5Þ design parameters of the SFHS and LAP-PFHS are shown in Table 1.
@x @y @z
The computational domain of air is considered to be three-
Momentum equation: dimensional fluid flow and steady-state. The top and bottom walls
Du @P are assumed to feature adiabatic and no-slip conditions. The side
x  momentum q ¼ þ div ðl grand uÞ ð6aÞ walls are considered symmetrical. The air flow is considered as
@t @x
non-bypass flow. It is given to be a constant temperature of
Dv @P 30 °C, ambient pressure and uniform flow at the inlet boundary
y  momentum q ¼ þ div ðl grand v Þ ð6bÞ
@t @y condition. At the outlet boundary of the air domain, the relative
pressure is set as 0.0 Pa. The thermal contact resistance between
Dw @P the fin and base is neglected as a consideration of continuity mate-
z  momentum q ¼ þ div ðl grand wÞ ð6cÞ
@t @z rial. The interface area of the heat sink and air is considered to have
no thermal contact resistance and a no-slip wall. The thermal con-
Energy equation:
ductivity of the air and aluminum heat sink is 0.0265 W/mK, and
Di 237 W/mK, respectively. The specific heat, density, and viscosity of
q ¼ p div U þ div ðk grand T Þ þ U ð7Þ
Dt the air is 1.005 kJ/kgK, 1.149 kg/m3 and 1.629  105 kg/ms,

Fig. 4. The computational domains of air and laterally perforated plate-fin heat sink.
S. Chingulpitak et al. / International Journal of Heat and Mass Transfer 138 (2019) 293–303 297

Table 1
Design parameters of laterally perforated plate-fin heat sink for numerical investigation. Unit: mm.

Base Fin Perforation


Height of base, Hb 0.5 Length of fin, Lf 75 Number of perforation, Np 14, 27, 75
Width of base, Wb 4 Height of fin, Hf 25 Number of perforation row, NR 2, 3, 5
Length of base, Lb 75 Thickness of fin, tf 1 Diameter of perforation, Dp 1, 2, 3, 4, 5, 6, 7, 8, 9, 10

Fig. 5. Grid independence for solid fin and laterally perforated plate-fin for
Fig. 6. Comparison between numerical results of solid fin heat sink and measured
Re = 3529: (a) Pressure drop (b) Heat transfer rate.
data: (a) Pressure drop and (b) Thermal resistance.

respectively. The governing equations under considered boundary the computation domains varies between 9,394,572 and
conditions are solved by using the commercial computational fluid 12,193,439.
dynamics software ANSYS CFX 17.0. The high-resolution transient
scheme is employed to discretize the governing equations. The 3.3. Validation and optimization
iteration of numerical solutions is repeated until the residual root
mean square (RMS) errors of all parameters are less than 105. For The validation of numerical results is carried out by comparison
all domains, the mesh is generated with a nonuniform structured with the measured data for the SFHS and the experimental results
mesh. The grid independence is considered carefully to find opti- of Shaeri and Bonner [24] for the LAP-PFHS. The heat transfer coef-
mum mesh elements that give accurate simulation results with a ficient for convection and Nusselt number are calculated as fol-
proper number of elements, as shown in Fig. 5. For example, the lows, respectively,
number of elements for LAP-PFHS with Dp = 10 mm is investigated
Qa
at 2,106,697, 9,394,572 and 28,833,837 elements. The simulation h¼  ð8Þ
results show that a number of elements beyond 9,394,572 leads Aconv T wall;av g  T a;av g
to a less than 0.26% variation in the pressure drop and 0.43% for
the heat transfer rate. However, it cannot be used for all heat sink hDch
Nu ¼ ð9Þ
models. Hence, in this numerical study, the number of elements for k
298 S. Chingulpitak et al. / International Journal of Heat and Mass Transfer 138 (2019) 293–303

The porosity (/) of the LAP-PFHS is calculated by the ratio of


void volume of the perforated fin to volume of the solid fin, as
follows:

Np pD2p
/¼ ð11Þ
4Hf Lf
The indicator parameter is used to determine the optimum size
of LAP-PFHS, as shown in Eq. (12). The thermal performance factor
(g) can be calculated from

ðNu=Nus Þ
g¼ ð12Þ
ðf =f s Þ

where Nus and fs are the Nusselt number and friction factor of the
SFHS.

4. Results and discussion

The prediction results of thermal resistance and pressure drop


according to the present model are validated with the measured
data and experimental results of Shaeri and Bonner [24]. The com-
parison results between the numerical results and measured data
for the SFHS are presented in Fig. 6. The mean absolute errors
(MAEs) of the comparison results are 6.2% and 4.9% for the pressure
drop and thermal resistance, respectively. Fig. 7 shows the compar-
ison results between the predicted results and the experimental
results of Shaeri and Bonner [24]. According to their study, the
dimensions of the plate-fin heat sink are Lb = Lf = 203 mm,
Hf = 22.86 mm, Wch = 2.18 mm and tf = 0.96 mm. The comparison
results show that the MAEs of pressure drop and thermal resis-
tance for LAP-PFHSs with square perforations are 3.6% and 5.3%,
respectively. Moreover, it shows that the tendency of predicted
results is consistent with their experimental results. Thus, it can
be summarized that the present model is reasonably accurate to
predict the pressure and temperature of other plate-fin heat sinks.
Consequently, the optimum size of circular perforation is investi-
gated under the same perforation area. The convective heat trans-
Fig. 7. Comparison between numerical results of laterally perforated plate-fin heat fer area, distance between perforations and porosities of all case
sink and experimental data of Shaeri and Bonner [24]: (a) Pressure drop and (b)
studies are presented in Table 2.
Thermal resistance.

4.1. Flow behavior


where Aconv represents the total convective heat transfer area.
Twall,avg and Ta,avg represent the average temperatures of the wall The air flow behaviors inside LAP-PFHS for Np = 27 and
surface and air, respectively. Dp = 5 mm are demonstrated with velocity vector. From the simu-
The friction factor in the flow passage of a plate-fin heat sink is lation results, the flow visualizations are presented with 3 different
calculated by is calculated from the pressure drop of heat sink, as distances of the x-z plane on the negative direction of the y-axis, as
follows: shown in Fig. 8. It is evident that the air flow behavior can be sep-
arated into recirculation and mixing zones. The recirculation zone
2DPDch
f ¼ ð10Þ is at the upstream side of the circular perforation for all sections, as
qV 2ch Lch shown in Fig. 8(a)–(c). It is the cause of increasing the width of the

Table 2
Details of the convective heat transfer area, position, and porosity of laterally perforated plate-fin heat sink.

Dp (mm) Np = 14 (NR = 2) Np = 27 (NR = 3) Np = 75 (NR = 5)


2 2
Aconv (mm ) Sx/Dp Sy/Dp / Aconv (mm ) Sx/Dp Sy/Dp / Aconv (mm2) Sx/Dp Sy/Dp /
1 4047 9.5 8.7 0.01 4067 7.6 6.5 0.01 4143 4.8 4.3 0.03
2 4025 4.8 4.5 0.02 4025 3.9 3.4 0.05 4025 2.4 2.3 0.13
3 3959 3.3 3.1 0.05 3898 2.6 2.3 0.10 3672 1.6 1.6 0.28
4 3849 2.5 2.4 0.09 3686 2.0 1.8 0.18 3083 1.2 1.2 0.50
5 3695 2.0 2.0 0.15 3389 1.6 1.5 0.28 – – – –
6 3497 1.7 1.7 0.21 3007 1.4 1.3 0.41 – – – –
7 3255 1.5 1.5 0.29 2541 1.2 1.1 0.55 – – – –
8 2969 1.3 1.4 0.38 – – – – – – – –
9 2640 1.2 1.3 0.48 – – – – – – – –
10 2266 1.1 1.2 0.59 – – – – – – – –
S. Chingulpitak et al. / International Journal of Heat and Mass Transfer 138 (2019) 293–303 299

Fig. 8. Flow behavior of air inside a circular perforation for Dp = 5.0 mm and Re = 2646: (a) at the perforation center (b) at Dp/6 section and (c) at Dp/3 section.

flow channel, which results in a separated flow. Moreover, the size flows, striking the edge of the circular perforation and returning
of the recirculation zone is smaller when the distance between the to the main flow region, which results in an increase of turbulent
horizontal plane and perforation center is greater. This is because flow.
the available flow distance inside the circular perforation at the Fig. 9 proposed the air flow behavior between two adjacent per-
upper and lower edge regions (section C-C) is shorter than in the forations with three different distances between perforations (Sx)
other regions (sections A-A and B-B). At the mixing zone, the air under the same perforation diameter (Dp = 4 mm) and Reynolds
300 S. Chingulpitak et al. / International Journal of Heat and Mass Transfer 138 (2019) 293–303

Fig. 9. Flow behavior of air inside two adjacent perforations with three different distance between perforations (Sx) for Dp = 4.0 mm and Re = 4413: (a) Sx/Dp = 2.5 (b) Sx/
Dp = 2.0 and (c) Sx/Dp = 1.2.

number (Re = 4413). The air flow behavior is presented with the larger recirculation region, as described by Kindere and Ganap-
velocity vector on an x-z plane. The simulation results show that athisubramani [35].
the air flow behavior in section A-A is similar to that in section
B-B, as shown in Fig. 9(a) and (b), respectively. The air strikes the 4.2. Effect of perforation diameter
downstream side of the perforation, which results in the mixing
of air flow. Then, the air flow reattaches to the fin surface between The LAP-PFHS is investigated with 10 different perforation
the two perforations. However, the reattachment flow is not found diameters. In all case studies, the lateral perforations are consid-
in the C-C section, as shown in Fig. 9(c). Moreover, it can be ered under a given fin area of 1875 mm2 (75  25 mm). Fig. 10
observed that the recirculation zone in section C-C is larger than shows the pressure drop of the LAP-PFHS with various perforation
that of other sections. This is because the shear layer of the short diameters and Reynolds numbers. Fig. 10(a), (b) and (c) present the
rib provides turbulence more than that of long rib and leads to a pressure drop of the LAP-PFHS at Np = 14, 27 and 75, respectively.
S. Chingulpitak et al. / International Journal of Heat and Mass Transfer 138 (2019) 293–303 301

Fig. 10. The effect of perforation diameters on pressure drop: (a) Np = 14 (b) Np = 27
(c) Np = 75.

Fig. 11. The effect of perforation diameters on heat transfer rate: (a) Np = 14 (b)
At the same as Np and Re, it can be observed that the total pressure
Np = 27 (c) Np = 75.
drop increases with increasing perforation diameter. The total
pressure drop continuously increases, although the frictional pres- can be observed that the maximum pressure drops for each perfo-
sure drop is decreased due to the decrease of the surface area of the ration number are obtained at the largest perforation diameter.
perforated plate fin. This is because the effect of pressure drag due Meanwhile, the pressure drop is not affected by the circular perfo-
to the perforation on total pressure drop is dominant, as explained ration at Dp = 1 mm. This is because the air flow over the perforated
by Gatski and Grosch [36]. The results indicate that the pressure plate fin with Dp = 1 mm acts like the flow over the solid fin. Thus,
drop across the LAP-PFHS increases about 96.3% when the diame- it can be concluded that the effect of perforation diameter on the
ter of perforation is changed from 1 mm to 10 mm. Moreover, it total pressure drop is achieved at Dp > 1 mm.
302 S. Chingulpitak et al. / International Journal of Heat and Mass Transfer 138 (2019) 293–303

Fig. 12. The effect of numbers of circular perforations on pressure drop. Fig. 14. The thermal performance factor with various diameters and numbers of
circular perforations.

total pressure drop of the LAP-PFHS for Np = 75 is higher than that


of other perforation numbers. In the case of Dp = 4.0 mm, the total
pressure drop rapidly increases by about 50–63% when the number
of perforations is changed between 27 and 75, as shown in Fig. 12.
This can be explained by the air flow behavior inside two adjacent
perforations, as shown in Fig. 9. That is, the increase of perforation
number results in the increase of pressure drag along the flow
channel. Fig. 13 presents the numerical results of the heat transfer
rate of LAP-PFHSs with various perforation numbers and diame-
ters, as well as Reynold numbers. For the perforation diameter of
1 mm, the heat transfer rate dramatically decreases when the
number of perforations increases. The results demonstrate that
LAP-PFHSs provide a lower heat transfer rate than SFHSs in the
range of 0.9–1.7%. For Dp = 2.0 mm and 3.0 mm, the present data
show that the heat transfer rate of LAP-PFHSs increases when the
number of perforations increases. However, the maximum heat
transfer rate is obtained at Np = 27 and Dp = 4 mm. In the range of
Fig. 13. The effect of numbers of circular perforations on heat transfer rate. increased heat transfer rate, a high number of perforations pro-
vides a higher heat transfer rate than a low number of perforations.
As shown in Fig. 11, the relationships between the heat transfer This is because the increase of perforation number and the distri-
rate and perforation diameters are investigated with 5 different bution of perforation row (NR = 2, 3 and 5) lead to the increase of
Reynolds numbers. It can be observed that the effect of perforation turbulent flow along the flow channel. After that, the heat transfer
diameter that creates more turbulent flow on the heat transfer rate rate dramatically decreases when the number of perforations
is obtained at Dp > 1 mm. For Dp = 1 mm, the heat transfer rate of increases. At Re = 4413, it is clearly seen that the heat transfer rate
the LAP-PFHS is lower than that of the SFHS, although the convec- of Np = 75 is lower than that of Np = 14 and 27. This is because the
tive heat transfer area is greater. This is because the air inside the reduction of the heat transfer area due to the increase of perfora-
channel cannot flow into the small circular perforation. Thus, the tion number is dominant.
added heat transfer area due to the fin thickness and the perfora-
tion circumference does not result in heat transfer enhancement. 4.4. Optimization
At Np = 14 and Re = 2646, as shown in Fig. 11(a), the heat transfer
rate of the LAP-PFHS increases by about 9.8% when the perforation In this study, the LAP-PFHS is studied with 21 different geome-
diameter is changed from 1 mm to 6 mm. The maximum heat tries. Based on the typical size of the solid fin, the dimensions of
transfer rates for Np = 27 and 75 are achieved at Dp = 5 and 3 mm, the laterally perforated fin are given as tf = 1 mm, Lf = 75 mm and
as shown in Fig. 11(b) and (c), respectively. However, it is notewor- Hf = 25 mm. In order to obtain the optimal LAP-PFHS, the thermal
thy that the heat transfer rate decreases when the perforation performance factor (g) is presented, as shown in Fig. 14. It is
diameter is higher. This is because the increase of the perforation defined as the ratio of the Nusselt number of the perforated plate
diameter results in the decrease of the heat transfer area, which fin to that of the solid fin to the friction factor of the perforated
decreases the heat transfer rate. plate fin to that of the solid fin, as shown in Eq. (12). For all condi-
tions, the numerical results show that LAP-PFHSs with Dp = 1 and
4.3. Effect of perforation number 2 mm provide lower thermal performance than SFHSs. Moreover,
it can be observed that the thermal performance (g) of LAP-
The effects of perforation number on pressure drop and heat PFHSs decreases when the Reynolds number increases. This is
transfer rate are presented in Figs. 12 and 13, respectively. The because the total pressure drop is the dominant effect in the ther-
LAP-PFHSs are investigated with four different perforation diame- mal performance (g) of LAP-PFHSs. At Re = 2646, the optimal diam-
ters of 1, 2, 3 and 4 mm. At a given the perforation diameter, the eter and number of circular perforation are obtained at Dp = 5 mm
S. Chingulpitak et al. / International Journal of Heat and Mass Transfer 138 (2019) 293–303 303

and Np = 27, respectively. The optimized LAP-PFHS provides about [7] C.J. Shih, G.C. Liu, Optimal design methodology of plate-fin heat sink for
electronic cooling system using entropy generation strategy, IEEE Trans.
10.6% higher thermal performance than the SFHS under the same
Compon. Packag. Manuf. Technol. 27 (3) (2004) 551–569.
conditions. Meanwhile, it can reduce the volume of heat sink mate- [8] C.T. Chen, C.K. Wu, C. Hwang, Optimal design and control of CPU heat sink
rial to 28%, which is presented in terms of porosity (/), as shown in processes, IEEE Trans. Compon. Packag. Manuf. Technol. 31 (1) (2008) 184–
Table 2. The maximum reduction of the volume of heat sink mate- 195.
[9] S. Chingulpitak, S. Wonwises, A review of the effect of flow directions and
rial by about 59% is achieved at Dp = 10 mm and Np = 14. It provides behaviors on the thermal performance of conventional heat sinks, Int. J. Heat
higher thermal performance than SFHS by about 8%. Mass Transf. 81 (2015) 10–18.
[10] T.Y. Kim, S.J. Kim, Fluid flow and heat transfer characteristics of cross-cut heat
sinks, Int. J. Heat Mass Transf. 52 (2009) 5358–5370.
5. Conclusions [11] S. Chingulpitak, N. Chimres, K. Nilpueng, S. Wongwises, Experimental and
numerical investigations of heat transfer and flow characteristics of cross-cut
heat sinks, Int. J. Heat Mass Transf. 102 (2016) 142–153.
This paper presents the thermal performance of a laterally per-
[12] D.-K. Kim, Thermal optimization of branched-fin heat sinks subject to a
forated plate-fin heat sink (LAP-PFHS) under a given area for perfo- parallel flow, Int. J. Heat Mass Transf. 77 (2014) 278–287.
ration. The main aim is to find the optimal design parameters of [13] H.E. Ahmed, Optimization of thermal design of ribbed flat-plate fin heat sink,
Appl. Therm. Eng. 102 (2016) 1422–1432.
LAP-PFHSs with circular perforations. The effects of perforation
[14] A. Subasi, B. Sahin, I. Kaymaz, Multi-objective optimization of a honeycomb
diameter (Dp) and perforation number (Np) on pressure drop, heat heat sink using Response Surface Method, Int. J. Heat Mass Transf. 101 (2016)
transfer rate, and thermal performance are proposed. The solid fin 295–302.
heat sink (SFHS) is presented to compare with the LAP-PFHSs. The [15] K.-S. Yang, S.-L. Li, I.Y. Chen, K.-H. Chien, R. Huc, C.-C. Wang, An experimental
investigation of air cooling thermal module using various enhancements at
present numerical model is validated by comparing the results low Reynolds number region, Int. J. Heat Mass Transf. 53 (2010) 5675–5681.
with the experimental data from available literature. The mean [16] N. Chimres, S. Wongwises, A critical review of the prominent method of heat
absolute error of the comparison results between LAP-PFHS with transfer enhancement for the fin-and-tube heat exchanger by interrupted fin
surface: the vortex generators approach, Int. J. Air-Cond. Refrig. 26 (3) (2018)
square perforation and experimental data is up to 5.3% for 1830001.
thermal resistance and 3.6% for the pressure drop. For Np = 14, [17] N. Chimres, C.-C. Wang, S. Wongwises, Optimal design of the semi-dimple
the highest heat transfer rate is obtained between Dp = 5–6 mm. vortex generator in the fin and tube heat exchanger, Int. J. Heat Mass Transf.
120 (2018) 1173–1186.
For Np = 27 and 75, the maximum heat transfer rate is achieved [18] N. Chimres, C.-C. Wang, S. Wongwises, Effect of elliptical winglet on the air-
at Dp = 4–5 mm and 3 mm, respectively. Moreover, the present side performance of fin-and-tube heat exchanger, Int. J. Heat Mass Transf. 123
results show that LAP-PFHSs with Dp = 1 mm provide a lower heat (2018) 583–599.
[19] M.R. Shaeri, M. Yaghoubi, Numerical analysis of turbulent convection heat
transfer rate than SFHSs for all conditions. In addition, the thermal
transfer from an array of perforated fins, Int. J. Heat Fluid Flow 30 (2009) 218–
performance factor (g) is proposed to find the optimum LAP-PFHS. 228.
From the results, it can be concluded that the diameter and [20] M.R. Shaeri, M. Yaghoubi, Thermal enhancement from heat sinks by using
perforated fins, Energy Convers. Manage. 50 (2009) 1264–1270.
number of perforations have a significant effect on the thermal
[21] W. Al-Sallami, A. Al-Damook, H.M. Thompson, A numerical investigation of the
performance of LAP-PFHSs. The best LAP-PFHS performance is thermal-hydraulic characteristics of perforated plate fin heat sinks, Int. J.
obtained at Dp = 5 mm and Np = 27. Therm. Sci. 121 (2017) 266–277.
[22] M.R. Shaeri, M. Yaghoubi, K. Jafarpur, Heat transfer analysis of lateral
perforated fin heat sinks, Appl. Energy 86 (2009) 2019–2029.
Conflict of interest [23] M.R. Shaeri, R. Bonner, Laminar forced convection heat transfer from laterally
perforated-finned heat sinks, Appl. Therm. Eng. 116 (2017) 406–418.
[24] M.R. Shaeri, R. Bonner, Heat transfer and pressure drop in laterally perforated-
We and our institution don’t have any conflict of interest and finned heat sinks across different flow regimes, Int. Commun. Heat Mass. 87
don’t have any financial or other relationship with other people (2017) 220–227.
or organizations that may inappropriately influence this work. [25] Md.F. Ismail, M.N. Hasan, S.C. Saha, Numerical study of turbulent fluid flow
and heat transfer in lateral perforated extended surfaces, Energy 64 (2014)
632–639.
Acknowledgments [26] K.H. Dhanawade, H.S. Dhanawade, Enhancement of forced convection heat
transfer from fin arrays with circular perforation, in: IEEE, Frontiers in
The first and fourth authors acknowledge the support provided Automobile and Mechanical Engineering (FAME), 2010, pp. 192–196.
[27] K.H. Dhanawade, V.K. Sunnapwar, H.S. Dhanawade, Thermal analysis of square
by the ‘‘Research Chair Grant” National Science and Technology and circular perforated fin arrays by forced convection, Int. J. Curr. Eng.
Development Agency (NSTDA) and King Mongkut’s University of Technol. 2 (2) (2014) 109–114.
Technology Thonburi through the ‘‘KMUTT 55th Anniversary Com- [28] K.H. Dhanawade, V.K. Sunnapwar, H.S. Dhanawade, Optimization of design
parameters for lateral circular perforated fin arrays under forced convection,
memorative Fund”. The second author would like to thank interna- Heat Transf. Asian Res. 45 (2016) 30–45.
tional co-operation in research provided by the National Research [29] T.K. Ibrahim, M.N. Mohammed, M.K. Mohammed, G. Najafi, N.A.C. Sidik, F.
Council of Thailand (NRCT) and the National Research Foundation Basrawi, A.N. Abdalla, S.S. Hoseini, Experimental study on the effect of
perforations shapes on vertical heated fins performance under forced
of Korea (NRF), and visiting professorship provided by KMUTT. convection heat transfer, Int. J. Heat Mass Transf. 118 (2018) 832–846.
[30] H.K. Versteeg, W. Malalasekera, Computational fluid dynamics, Longman
References Group, 1995.
[31] P.H. Oosthuizen, D. Nayler, An introduction to convective heat transfer
analysis, Mc-Graw-Hill, 1999.
[1] S.J. Kim, D.-K. Kim, H.H. Oh, Comparison of fluid flow and thermal
[32] F.R. Menter, Two-equation eddy-viscosity turbulence models for engineering
characteristics of plate-fin and pin-fin heat sinks subject to a parallel flow,
applications, AIAA J. 32 (8) (1994) 1298.
Heat Transf. Eng. 29 (2) (2008) 169–177.
[33] F. Zhou, N. Hansen, I. Catton, Numerical predictions of thermal and hydraulic
[2] H.Y. Li, S.M. Chao, Measurement of performance of plate-fin heat sinks with
performances of heat sinks with enhanced heat transfer capability,
cross flow cooling, Int. J. Heat Mass Transf. 52 (2009) 2949–2955.
Proceedings of the ASME/JSME 2011 8th Thermal Engineering Joint
[3] H.H. Wu, Y.Y. Hsiao, H.S. Huang, P.H. Tang, S.L. Chen, A practical plate-fin heat
Conference AJTEC2011, March 13–17, 2011, Honolulu, Hawaii, USA, 2011.
sink model, Appl. Therm. Eng. 31 (2011) 984–992.
[34] O.B. Kanargi, P.S. Lee, C. Yap, A numerical and experimental investigation of
[4] C.T. Chen, H.I. Chen, Multi-objective optimization design of plate-fin heat sink
heat transfer and fluid flow characteristics of a cross-connected alternating
using a direction-based genetic algorithm, J. Taiwan Inst. Chem. Eng. 44 (2013)
converging–diverging channel heat sink, Int. J. Heat Mass Transf. 106 (2017)
257–265.
449–464.
[5] R.V. Rao, G.G. Waghmare, Multi-objective design optimization of a plate-fin
[35] J. van der Kindere, B. Ganapathisubramani, Effect of length of two-dimensional
heat sink using a teaching-learning-based optimization algorithm, Appl.
obstacles on characteristics of separation and reattachment, J. Wind Eng. Ind.
Therm. Eng. 76 (2015) 521–529.
Aerod. 178 (2018) 38–48.
[6] J.R. Culham, Y.S. Muzychka, Optimization of plate-fin heat sink using entropy
[36] T.B. Gatski, C.E. Grosch, Embedded cavity drag in steady and unsteady flows,
generation minimization, IEEE Trans. Compon. Packag. Manuf. Technol. 24 (2)
NASA Contractor Report, 1983.
(2001) 159–165.

You might also like