You are on page 1of 12

International Journal of Heat and Mass Transfer 81 (2015) 325–336

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


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

Fluid flow and heat transfer investigations on enhanced microchannel


heat sink using oblique fins with parametric study
Yong Jiun Lee, Pawan K. Singh ⇑, Poh Seng Lee
Department of Mechanical Engineering, Faculty of Engineering, National University of Singapore, 9 Engineering Drive 1, EA-07-08, Singapore 117575, Singapore

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

Article history: Enhanced microchannel heat sink with sectional oblique fin is used to modulate the flow in contrast to
Received 15 April 2014 continuous straight fin. The re-initialization of thermal boundary layer at the leading edge of each oblique
Received in revised form 28 August 2014 due to breakage of continuous fin into oblique sections and the secondary flow due to these oblique cuts
Accepted 10 October 2014
resulted in better heat transfer and a comparable pressure drop. Extensive experimental investigations
Available online 6 November 2014
are carried out with silicon test vehicle with hydraulic diameter of 100 lm and 200 lm and de-ionized
water as flowing fluid. A parametric study involving the oblique angle, fin pitch is also carried out. Appre-
Keywords:
ciable heat transfer augmentation is also achieved with maximum heat transfer performance enhance-
Enhanced microchannel
Oblique fins
ment at 47% when Re = 680. Comparable pressure drop to conventional microchannel is maintained up
Thermal management to Re = 500. Parametric study suggests that smaller oblique angle and smaller fin pitch are beneficial
Electronic cooling for heat transfer enhancement. The performance of the microchannel with 100 lm channel width and
Fin pitch 27° oblique angle is found to be optimum.
Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction 1250 W/cm2, still holds great promise of microchannels as a viable


cooling option for microelectronics cooling.
The advancement in packaging technology has led to smaller Extensive research on techniques to enhance the heat transfer
chip size associated with higher and more concentrated heat flux. performance of microchannel heat sinks has been conducted over
With smaller chip size and high heat flux comes the generation of the past decades. Based on the minimization of the flow resistance
high temperature with addition of hot spots which can accelerate between a volume (volumetric heater) and a point (cooling liquid
the meantime to failure (MTTF) and reduce the lifespan of elec- stream), a tree-shaped channel network was proposed by Bejan
tronic devices as described by Black0 s equation [1]. Therefore, cool- and Errera [5]. Chen and Cheng [6] proposed a right-angled bifur-
ing of electronic devices is a persistent challenge in package design. cation in a rectangular shaped heat sink while Pence [7] preferred a
Heat sinks with mini/microchannel are currently most widely used smaller bifurcation angle in a disk shaped heat sink. Both designs
for the cooling of small but highly heated electronic devices, due to deployed self-similar fractal-like branches in a heat sink at a fixed
their advantages such as compactness, light weight and higher ratio between the upstream and downstream channel width and
heat transfer surface area to fluid volume ratio compared with channel length, leading to identical bifurcating pattern at each
other macro-scale systems. Since the concept of microchannel heat level. On the other hand, the concept of deploying re-entrant space
sink was first proposed by Tuckerman and Pease [2] in 1981, in microchannel heat sink was explored experimentally and
numerous studies has been conducted to investigate its flow and numerically by Xu et al. [8–9]. Their studies indicated that the
heat transfer characteristic in microchannels. The study of Cope- thinning of boundary layers after every re-entrant space resulted
land et al. [3] revealed the application difficulties for the conven- local heat transfer enhancement and similar or reduced pressure
tional microchannel, which were associated with high pressure drop across the enhanced heat sink compared to that of the con-
drop and significant lateral temperature gradient. However, the ventional configuration. Lee et al. [10] proposed a hot spot mitiga-
report of Prasher and Chang [4] that single phase microchannel tion scheme by creating a recess at the cover lid just before hot
liquid cooling was capable of cooling heat flux as high as spot region.
As an effective method to promote better fluid mixing, secondary
flow has been listed as one of the efficient heat transfer augmenta-
tion techniques. Steinke and Kandlikar [11] suggested two potential
⇑ Corresponding author. Tel.: +65 6516 4657; fax: +65 6779 1459.
methods in generating secondary flow for microchannel application.
E-mail address: mpepks@nus.edu.sg (P.K. Singh).

http://dx.doi.org/10.1016/j.ijheatmasstransfer.2014.10.018
0017-9310/Ó 2014 Elsevier Ltd. All rights reserved.
326 Y.J. Lee et al. / International Journal of Heat and Mass Transfer 81 (2015) 325–336

Nomenclature

A convection heat transfer area, m2 q00 heat flux, W/m2


Ac fin area, m2 u velocity, m/s
Dh hydraulic diameter, m w width, lm
H channel height, lm
K loss coefficient Greek symbols
L heat sink length, mm a aspect ratio
L0 characteristic length h oblique angle, °
N number of fin D gradient
Nu Nusselt number l dynamic viscosity, Ns/m2
P fin perimeter, m q mass density, kg/m3
PP pump power, W
DP pressure drop, Pa Subscripts
Pr Prandtl number ave average
Q volumetric flow rate, m3/s b base (unfinned)
R thermal resistance, °C/W
c contraction
Rs summation of spreading resistance and 1D conduction ch channel
resistance, °C/W CM conventional microchannel
Re Reynolds number
conv convective
T temperature, °C EM enhanced microchannel
W heat sink width, mm e expansion
X, Y, Z global coordinate system f fluid
X0 dimensionless axial distance, X0 = X/L
fin fin
Y0 dimensionless channel height, Y0 = Y/H HS heat sink
Z0 dimensionless heat sink width, Z0 = Z/W in inlet
cp specific heat capacity, kJ/kgK max maximum
f friction factor
ob oblique channel
h heat transfer coefficient, W/m2K out outlet
k thermal conductivity, W/mK p plenum
l fin length, lm
si silicon
m _ mass flow rate, kg/s sp spreading
n number of fin row tot total
p fin pitch, lm w wall
q heat transfer rate, W
q⁄ heat transfer rate per unit heater, W

The first suggestion was to add smaller channels at a certain angle ior. Dimension is another important aspect of the oblique fin chan-
between two main liquid channels. Alternatively, secondary flow nel which can affect the flow and heat transfer. In a recent review
can also be generated by a venturi effect. Both methods can lead to article, it has been observed by Kandlikar et al. [16] that the heat
heat transfer enhancement phenomenon without or with slight transfer coefficients of these oblique finned copper microchannels
additional pumping power consumption. This technique can be were relatively low due to large hydraulic diameters and suggested
applied by various kinds of fins and plates arrangement. The study it will be interesting to see performance at smaller level. In addi-
of Tatsumi et al. [12] on parallel plate fin arrays with oblique notches tion, there are a few key design parameters that greatly influence
suggested that the presence of oblique notches resulted in heat the heat transfer and pressure drop performance in oblique finned
transfer enhancement through interrupting thermal boundary lay- microchannels. Therefore, a parametric study is essential to
ers and promoting the generation of spanwise flow (secondary explore the outcome of varying these parameters.
flow). The present work tries to address these challenges by experi-
The concept of oblique fins in microchannel heat sinks was pro- mentally investigating the oblique fin microchannels of 100 micron
posed by Lee et al. [13] By breaking the continuous fins into obli- and 200 micron nominal width. The local flow and heat transfer
que sections, significant local and global heat transfer enhanced behavior is observed and analyzed. A parametric study looking into
was achievable with little or negligible pressure drop penalty. the effect of oblique angles and fin pitch is carried out. At last, an
Fan et al. [14] introduced the concept of oblique fins into the optimum geometry is suggested based on the results.
design of cylindrical heat sinks and obtained a much larger average
Nusselt number compared with that of conventional straight fin
heat sink. Detailed numerical and experimental studies were sub- 2. Experimental setup
sequently performed by Lee et al. [15] to further look into the
hydrodynamics and thermal development along the oblique finned A schematic of the experimental flow loop is shown in Fig. 1.
microchannel heat sink, providing some insights to the fundamen- Deionized water from a reservoir tank is driven through the flow
tal mechanism of this technology. These all works focus on overall loop using a micro-annular gear pump. This pump forces the cool-
flow and heat transfer in oblique fin channel. Since the change in ant through a 15 lm filter and a flow meter before entering the
flow and heat transfer behavior in oblique fin can be attributed microchannel test section. A heat exchanger (connected to water
to local geometry change such as oblique fin and oblique channel, bath) is used to regulate the water temperature before coolant
it will be interesting to see the local flow and heat transfer behav- enters the test section. Pressure transmitters are attached to the
Y.J. Lee et al. / International Journal of Heat and Mass Transfer 81 (2015) 325–336 327

Fig. 1. Schematic for experimental flow loop.

manifolds immediately before and after the test piece to measure ness for the current test vehicles. Each thermal test die has a heater
the pressure drop across the test section. The water temperatures (doped silicon well) at the bottom wall of the chip at 2 mm  2 mm
immediately before, Tf,in and after, Tf,out the microchannel heat sink and a series of thermal sensors (5 diodes connected in series), which
in the manifolds are also measured with two T-type thermocouples enable independent heater control and local temperature measure-
(copper-constantan). Heated water that exits the test section is ment, providing greater insight into local heat transfer behavior and
cooled by passing it through a liquid-to-air heat exchanger before the overall temperature mapping. The resistive heating of the sili-
it returns to the reservoir tank. con thermal test dies is accomplished by driving the current
After the test section is assembled and properly sealed, the gear through the doped silicon well, where the input power is controlled
pump is switched on and the desired flow rate through the flow by a DC power supply unit. Fig. 2(b), on the other hand, shows the
loop is controlled with the digital control on the gear pump. As test piece and the enlarged view of the 5  5 arrays of thermal test
the flow rate is stabilized, the power supply to the heaters is dies. A coordinate numbering scheme (X, Z) for thermal test dies
switched on, and steady state is usually reached in 30–45 min. identification is also indicated in the figure.
The power input to the heaters in the test piece is controlled by Temperatures on the thermal test chip are indicated as voltage
a DC power supply unit. The voltage across the heaters is measured drop across the thermal diode sensors. Prior to the experiment, the
directly while the high current through the heaters is calculated voltage-temperature response of these thermal diode sensors are
from Ohm0 s law based on the voltage measured across a shunt established through calibration. The calibration is performed in a
resistor that is connected in series with the heater. Input power convection oven from 30 °C to 90 °C, in steps of 10 °C. Temperatures
to the heaters is calculated as the product of measured voltage and voltage drop are then recorded when both temperatures of oven
and calculated current. Steady-state readings from the thermocou- and thermal sensors reach steady state, typically in an hour time.
ples, differential pressure transmitter and others are recorded by
data acquisition system and stored in a computer throughout the 2.2. Data reduction
duration of the experiment. Each steady state value was calculated
based on the average of 100 readings sampled at 0.1 Hz. Fig. 3 shows the representative image of oblique fin channel. In
the schematic, l is fin length, p is fin pitch, h is oblique angle, wch is
2.1. Test sections channel width, ww is fin width, wob is oblique channel width.
The sensible heat gained by the coolant is determined from an
Silicon-based microchannel heat sinks are created on flip chip energy balance:
packages for experimental investigation. Channels are cut on the q ¼ qcp Q ðT f ;out  T f ;in Þ ð1Þ
backside of the silicon thermal test chips using wafer dicing saw
blades of different width. Measurement is performed with a 3-axis It is found that more than 95% of the heat input is transferred to the
measurement microscope (Mitutoyo MF-B1010C) with 200 mag- coolant across all the experimental runs, where the heat loss is
nification at 9 different points (3  3 grid) on the microchannel deemed minimum or negligible.
heat sink. The average surface roughness (Ra) measured at the bot- As for material properties, silicon heat sink are assumed to have
tom wall of the channel is 178 nm. constant thermal conductivity, ksi = 148 W/mK. The density, spe-
Fig. 2(a) displays the test section of microchannel heat sinks in cific heat capacity, thermal conductivity and dynamic viscosity of
this experiment. It consists of a polycarbonate manifold that is water are evaluated at the mean fluid temperature (average of
mounted onto a flip chip package, where the microchannels are the fluid inlet and outlet temperatures) based on the formulas sta-
laid. Each thermal test die is 0.100  0.100 (2.54 mm  2.54 mm) in ted in [17].
size and when diced in 5  5 grid array, this results in an overall As the pressure transmitters are located at the manifolds, the
footprint of 0.500  0.500 (12.7 mm  12.7 mm) with 0.65 mm thick- pressure drop measurement represents the combined losses due
328 Y.J. Lee et al. / International Journal of Heat and Mass Transfer 81 (2015) 325–336

Fig. 2. (a) Microchannel test section, (b) test piece with 5  5 array of thermal test dies.

DPch ¼ DP  ðDP c1 þ DP c2 þ DP e2 þ DP e1 Þ
where DPc1 and DPc2 are the contraction pressure losses from the
deep plenum to the shallow plenum, and from the shallow plenum
to the microchannel. These minor losses can be expressed as [19]:
1   K
c1
DPc1 ¼ qf u2p2;in  u2p1;in þ q u2 ð3Þ
2 2 f p2;in
1   K c2
DPc2 ¼ qf u2in  u2p2;in þ q u2 ð4Þ
2 2 f in
where p1 and p2 denote the deep plenum and shallow plenum,
respectively, and Kc1 and Kc2 are the loss coefficients for the abrupt
contractions. On the other hand, DPe2 and DPe1 express the pressure
losses from the microchannel to the shallow plenum and from the
shallow plenum to the deep plenum, which can be written as
follow:
1   K
e2
DPe2 ¼ qf u2p2;out  u2out þ q u2 ð5Þ
2 2 f out
Fig. 3. Schematic of the oblique fin and its nomenclature. 1   K
e1
DPe1 ¼ qf u2p1;out  u2p2;out þ q u2 ð6Þ
to the frictional loss in microchannels and minor losses due to 2 2 f p2;out
abrupt contraction and expansion at the inlet and outlet [18], where Ke1 and Ke2 represent the losses coefficients due to the abrupt
which can be written as follow: expansion. For the present heat sink test section geometry, the
value of Kc1, Kc2, Ke1 and Ke2 are close to unity [20].
DP ¼ DPc1 þ DPc2 þ DPch þ DPe2 þ DPe1 ð2Þ
In addition to the above data reduction, thermal spreading resis-
The pressure drop across microchannel can then be calculated as tance is calculated to better assess the channel wall temperature of
Y.J. Lee et al. / International Journal of Heat and Mass Transfer 81 (2015) 325–336 329

silicon-based microchannel heat sinks. Thermal spreading resistance measured across microchannel heat sinks are at 1.8–12.6%. The
occurs whenever heat leaves a heat source of finite dimensions and nominal uncertainty in each parameter is shown in Table 1.
enters into a larger region. The thermal test dies used in the
experiment has 64% (80% length  80% width) of its bottom 3. Results and discussion
surface area covered by heater. Thus, heat spreading takes place
when heat is transported from the bottom of the dies to the base of 3.1. Silicon-based microchannel heat sink with 100 lm nominal
the microchannel. A 3D thermal spreading resistance model proposed channel width
by Yovanovich et al. [21] is adopted to account for the effect of
thermal spreading. This model assumes a planar rectangular heat The experiments are conducted over the flow rates ranging
source situated on one end of a flux channel, where it is cooled on from 100 mL/min to 500 mL/min, which corresponds to Reynolds
the other end through an average heat transfer coefficient, h while numbers of 180–680. By employing silicon thermal test chip in
the flux channel is laterally adiabatic. these experiments, the local behavior (local temperature and local
The summation of the thermal spreading and 1D conduction heat transfer coefficient profiles) of the enhanced heat sink can be
resistances across the thermal test chip is then found as Rs,(x,z) investigated. In this case, the performance of the enhanced micro-
while the temperature at the base of microchannel can be deter- channel heat sink with 100 lm nominal width under the experi-
mined as the heat supplied by each individual heater, q⁄ and tem- mental conditions of 160 mL/min (Re = 252) total coolant flow
perature at the bottom surface of thermal test dies, Th are known. rate and total 273 W heater power of will be presented in detail.
Total resistance = Thermal spreading resistance + 1D conduc- The measured temperatures of the thermal test chips are plot-
tion resistance ted in Fig. 4 based on the spanwise and streamwise position on
the heat sink. The figure is plotted for different Z where Z is Dimen-
Rs;ðx;zÞ ¼ Rsp;ðx;zÞ þ R1D ð7Þ
sionless heat sink width. The wall temperature is lowest at Z0 = 0.5

q Rs;ðx;zÞ ¼ T h;ðx;zÞ  T w ; ðx; zÞ ð8Þ followed Z0 = 0.3 and 0.7. The heat transfer is almost same in this
region and the plots overlap each other. Higher temperatures are
where Thhðx;zÞ is the measured temperature at bottom of thermal
observed beyond Z0 = 0.3 and 0.7. Wall temperature of heaters at
test dies and Thhðx;zÞ is calculated channel wall temperature.
Z0 = 0.9 is the highest followed by heaters at Z0 = 0.1. The difference
As the local temperatures of the silicon-based microchannel
is clearly due to the lateral heat conduction from discrete heat
heat sink can be measured by individual thermal diode sensor in
sources towards the peripheral of thermal test chips.
a 5  5 grid array across the heat sink, local heat transfer coeffi-
The enhanced microchannel shows significant reduction in total
cient of the heat sink section on top of each thermal test chip is
thermal resistance compared to the conventional microchannel.
then determined using the correlation:
The total thermal resistance comprises of conductive, spreading,
q convective and caloric thermal resistances. Generally, the conduc-
hðx;zÞ ¼ ð9Þ
Atot ðT wðx;zÞ  T f ;x Þ tive thermal resistance remains constant while spreading, convec-
tive and caloric thermal resistances reduce with the increasing
where q⁄ is the local heat transfer rate evaluated as the average sen-
Reynolds number, resulting in lower total thermal resistance. The
sible heat gain by coolant per unit thermal test dies. A⁄tot on the
improved heat transfer performance of the enhanced microchannel
other hand, is the average convective heat transfer area per unit
is demonstrated in Fig. 5, where the total thermal resistance of the
thermal test dies.
enhanced microchannel heat sink is consistently lower than the
Total thermal resistance of the heat sink is defined as
conventional configuration in the range of Reynolds numbers stud-
T max  T f ;in ied. At low Reynolds number 180, the difference in thermal resis-
Rtot ¼ ð10Þ tances of both the channels cannot be claimed. As Reynolds
q
number rises, the effectiveness of enhanced microchannel
where Tmax is the maximum measured temperature of the heat sink, increases and the percentage of reduction in total thermal resis-
Tf is the inlet coolant temperature and q is the heat supplied into the tance quickly increase. The maximum total thermal resistance
heat sink. reduction achieved at Reynolds number 690 is as much as 25%
The streamwise local heat transfer coefficient, hx is defined as (Rtot,EM = 0.089 °C/W versus Rtot,CM = 0.119 °C/W). In the context of
the average heat transfer coefficient of thermal test chips in the current experiment, the highest percentage of total thermal
spanwise direction at the identical streamwise distance from the
inlet.
Table 1
hðx;1Þ þ hðx;2Þ þ hðx;3Þ þ hðx;4Þ þ hðx;5Þ Uncertainties for various critical parameters for silicon-based microchannel heat
hx ¼ ð11Þ sinks.
5
Subsequently, the average heat transfer coefficient of the heat sink Variable Absolute uncertainties/fractional
uncertainties
is calculated as the average value from all the twenty-five thermal
test dies. Channel width, wc, fin width, ww and 2 lm
fin length, l
The corresponding Nusselt number is then calculated as
Channel height, H 5 lm
Nu ¼ hDkf
h
Oblique angle, h 0.15°
Standard error analysis adopting the principles by Taylor [22] is Volumetric flow rate, Q 1.9–9.8%
performed with the uncertainties in the reported Nusselt numbers Temperature, Tf, Tcu 0.3 °C
falling between 6.6% and 15.6%. The highest percentage of error Reynolds number, Re 3.2–10.3%
Wall temperature, Tw 0.9–1.2%
occurs at the maximum flow rate, mainly due to the lowest sensi-
Hydraulic diameter, Dh 1.4%
ble water temperature rise from the inlet to the outlet of micro- Fin efficiency, g 3.7–15.8%
channel. The main contributor of uncertainties in this experiment Total heat transfer area, Atot 3.4–13.7%
is identified as the measurement of water temperature (±0.3 °C). Heat transfer coefficient, h 6.4–15.6%
Nusselt number, Nu 6.6–15.6%
Sequential perturbation method [23] is used to address the uncer-
Pressure, DP 1.8–12.6%
tainty of spreading resistance, due to the complexity of the Friction factor, f 2.7–13.6%
equation. On the other hand, uncertainties for pressure drop
330 Y.J. Lee et al. / International Journal of Heat and Mass Transfer 81 (2015) 325–336

highly non-uniform heat transfer performance within the heat


sink. Instead of declining, the enhanced microchannel keeps the
Nusselt number at a much elevated level and consistent across
the enhanced microchannel heat sink. The profile of the stream-
wise local Nusselt number is very uniform, confined to a narrow
range of 7.0–8.2 and 12.4–14.4 for both Reynolds number cases
respectively from the inlet of the microchannel to its outlet. This
phenomenon is highly due to the frequent re-initialization of ther-
mal boundary layers and generation of secondary flows. This com-
bination ensures that the flow is in a constant state of development
thus having a sustainable performance close to that of the flow
upstream.
The average heat transfer performance of the microchannel
heat sinks is plotted in Fig. 7. Generally, the average Nusselt num-
ber, Nuave increases with Reynolds number as the thermal bound-
ary layer thickness decreases with the increased fluid velocity.
However, the heat transfer for the enhanced microchannel with
oblique fins is highly augmented in comparison with conventional
microchannel heat sink. At Reynolds number 180, the average
Nusselt number of the enhanced microchannel is 15% higher than
Fig. 4. Measured heat sink wall temperature of the enhanced microchannel heat that of conventional microchannel. As Reynolds number rises, the
sink.
Nusselt number increases by almost 47%, from 9.0 to 13.2. The
appreciable enhancement in heat transfer is due to the combined
effects of thermal boundary layer re-development at the leading
edge of each oblique fin and the secondary flows generated by flow
diversion through the oblique channels. As the flow rate induced
through the heat sink is increased, more fluid is diverted from
the main channel into the oblique channel. This secondary flow
thus carries higher momentum and further disrupts the boundary
layers development and augments the heat transfer.
Nevertheless, the most interesting feature of the enhanced
microchannel heat sink is that the significant heat transfer aug-
mentation is achieved with small pressure drop penalty. The
enhanced microchannel heat sink, which employs secondary flow
to enhance its heat transfer, still managed a comparable pressure
drop with the conventional microchannel heat sink for Reynolds
number lower than 500, as displayed in Fig. 8. This distinguishes
the proposed scheme from the conventional heat transfer enhance-
ment scheme, where trade-off in terms of pressure drop penalty is
inevitable. As the Reynolds number increases, a higher percentage
of coolant will be diverted into oblique channels. This creates a sec-
ondary flow with stronger momentum, which further augments
Fig. 5. Comparison of total thermal resistance between the microchannel heat the heat transfer but incurs additional pressure drop penalty.
sinks. Therefore, the pressure drop for enhanced microchannel starts to
deviate and increases more than the conventional configuration.
However, the magnitude of increment of pressure drop is consid-
resistance reduction translates to an 8.1 °C lower in maximum ered manageable for the same micropump.
chip temperature for the enhanced microchannel heat sink
(Tmax,EM = 50.5 °C versus Tmax,CM = 58.6 °C). 3.2. Parametric study
As discussed, the introduction of oblique fins and channels
leads to a much uniform heat removal capability across the micro- As discussed in introduction, there are few key design parame-
channel heat sink. Fig. 6(a) and (b) illustrate the comparison of ters. As these parameters vary, the overall fin layout will change.
local heat transfer performance of microchannel heat sinks for This change has an enormous effect on the flow field and conse-
two of the experimental runs at Re  250 and Re  680. The mark- quently it affects the performance of the heat sink. Therefore, a
ers denote the location of thermal diode sensors on the silicon test parametric study is essential to explore the outcome of varying
chip in the streamwise direction, where the temperature measure- these parameters. The two critical design parameters identified
ment is made. It is noticed that both conventional and enhanced for the parametric study are; oblique angle and oblique fin pitch.
microchannel heat sinks display a relatively close Nusselt number A total of seven configurations with three oblique angles (15°,
at the upstream of the heat sink, X0 = 0.1 when the boundary layers 27° and 45°) coupling with three fin pitches (400 lm, 800 lm
are thin. The distinguishing point for these two heat sinks is the and 1500 lm) are fabricated on silicon-based microchannel heat
development of the flow regime and boundary layers as coolant sink for the performance evaluation. The benchmarking is based
travels downstream. As boundary layers continue to thicken with on two conventional microchannel configurations, which have
the flow distance, heat transfer performance of the conventional 100 lm and 200 lm nominal channel width respectively. Detailed
microchannel deteriorates rapidly, as showed in the figures for dimensions of various microchannel heat sinks are tabulated in
both Reynolds number cases. A maximum 48% drop in Nusselt Tables 2 and 3. Microchannel heat sinks with the similar channel
number is observed between X0 = 0.1 and X0 = 0.9, resulting in a width, fin pitch but different oblique angles are grouped together
Y.J. Lee et al. / International Journal of Heat and Mass Transfer 81 (2015) 325–336 331

Fig. 6. Comparison of local heat transfer performance of microchannel heat sinks for (a) Re  250 (b) Re  680.

While the combination of a different oblique angle and pitch,


can result in a wholly new enhanced microchannel heat sink lay-
out, it also re-distributes the fin and un-finned surface areas avail-
able for convective heat transfer. Table 4 compares the fin, un-
finned and total surface areas available for convective heat transfer
of different microchannel configurations. From Table 4, it is clear
that the introduction of oblique fins and channels in microchannel
heat sink increases the total heat transfer area by 5–30% with
smaller fin pitch and oblique angle leading to larger surface areas.

3.2.1. The effect of oblique angle variation


The effect of a change in oblique angle to heat transfer, and
pressure drop performance, is diagnosed with three groups of
microchannel heat sinks. The first group are three microchannel
heat sinks with 100 lm nominal channel width, which consists
of a conventional straight channel heat sink (Con-100) as bench-
mark and two enhanced heat sink (400 lm nominal fin pitch)
Fig. 7. Comparison of heat transfer performance for microchannel heat sinks.
with 27° and 45° oblique angle respectively (27° – NP0 and 45° –
NP0 ). The experimental results of both heat transfer and pressure
drop are presented in Fig. 9. Fig. 9(a) compares the average Nusselt
number of three microchannel heat sinks. It is observed that the
average Nusselt numbers of Con-100 and 45° – NP0 almost overlap
each other, while that of 27° – NP0 is significantly higher. This sug-
gests that 45° – NP0 is not as effective as 27° – NP0 in augmenting
heat transfer performance. In contrast, the heat transfer augmenta-
tion for 27° – NP0 rises with Reynolds number. As a result, this con-
figuration records the lowest total thermal resistance among the
threes, as shown in On the other hand, the pressure drop across
Con-100 and 45° – NP0 overlap each other as they did in the Nusselt
number plot, as shown in Fig. 9(b). The 27° – NP0 also displays a
pressure drop that is comparable to the other two for Reynolds
numbers lower than 500, signaling that the heat transfer enhance-
ment can be achieved with negligible pressure drop penalty.
The second group of microchannel heat sinks for comparison
has the same features as the first group, with exception that they
are 200 lm in nominal channel width. In this group, conventional
microchannel (Con-200) is adopted as the baseline for comparison
with two enhanced microchannel heat sinks (800 lm nominal fin
Fig. 8. Comparison of pressure drop for microchannel heat sinks. pitch) with 27° and 45° oblique angle respectively (27° – NP and
45° – NP). A similar trend is observed from the heat transfer and
to compare the effect of oblique angle variation and it is vice versa pressure drop performance of this group, as displayed in Fig. 10,
when evaluating the effect of fin pitch variation. in comparison to the first group, which has 100 lm nominal
332 Y.J. Lee et al. / International Journal of Heat and Mass Transfer 81 (2015) 325–336

Table 2
Dimensional details of microchannel heat sinks with 100 lm channel width.

Characteristic Conventional microchannel #1 (Con-100) Enhanced microchannel #1 (27° – NP0 ) Enhanced microchannel #2 (45° – NP0 )
Material Silicon
Footprint, width  length (mm2) 12.7  12.7
Number of fin row, n 61
Main channel width, wc (lm) 115 113 110
Fin width, ww (lm) 85 87 91
Channel depth, H (lm) 387 379 351
Aspect ratio, a 3.37 3.35 3.19
Oblique channel width, wob (lm) – 49 48
Fin pitch, p (lm) – 405 400
Fin length, l (lm) – 292 332
Oblique angel, h (°) – 26.3 45.0

– NP0 denotes nominal pitch (400 lm) for enhanced microchannel heat sink with 100 lm channel width.

Table 3
Dimensional details of microchannel heat sinks with 200 lm channel width.

Characteristic Conventional Enhanced Enhanced Enhanced Enhanced Enhanced


microchannel #2 (Con- microchannel #3 (27° – microchannel #4 (45° – microchannel #5 (15° microchannel #6 (27° microchannel #7 (45°
200) NP) NP) – LP) - LP) - SP)
Material Silicon
Footprint, 12.7  12.7
width  length
(mm2)
Number of fin row, 30
n
Main channel 205 205 203 205 206 204
width, wc (lm)
Fin width, ww (lm) 195 195 197 194 194 196
Channel depth, H 450 418 445 455 405 449
(lm)
Aspect ratio, a 2.20 2.04 2.19 2.17 1.97 2.20
Oblique channel – 104 103 101 103 103
width, wob (lm)
Fin pitch, p (lm) – 783 800 1502 1536 400
Fin length, l (lm) – 550 653 1104 1301 253
Oblique angel, h (°) – 27.1 45.0 14.9 26.3 45.0

– NP denotes nominal pitch (800 lm) for enhanced microchannel heat sink with 200 lm channel width.
– LP denotes large pitch (1500 lm) for enhanced microchannel heat sink with 200 lm channel width.
– SP denotes small pitch (400 lm) for enhanced microchannel heat sink with 200 lm channel width.

Table 4
Comparison of convective heat transfer areas for microchannel heat sinks.

Configuration Oblique angle (°) Fin pitch (lm) Oblique fin perimeter (lm) Fin area (mm2) Un-finned area (mm2) Total heat transfer area (mm2)
Con-100 – – – 609.5 90.6 700.1
27° – NP0 26.3 405 976.6 718.7 107.5 826.2
45° – NP0 45.0 400 919.8 635.6 98.1 733.7
Con-200 – – – 354.0 80.8 434.8
27° – NP 27.1 783 1956.9 409.1 102.8 511.9
45° – NP 45.0 800 1864.8 406.4 93.6 500.0
15° – LP 14.9 1502 3718.2 441.6 100.5 542.1
27° – LP 26.3 1536 3476.4 359.1 92.4 451.5
45° – SP 45.0 400 1059.3 464.5 107.8 572.3

channel width. Both Con-200 and 45° – NP result in comparable aver- angle can lead to better heat transfer performance with higher aver-
age Nusselt numbers and pressure drop for Reynolds number up to age Nusselt number. 27° – LP also stands out from the comparison
1000, while that of 27° – NP is consistently higher. This again with its comparable pressure drop with Con-200, within the range
indicates that smaller oblique angle is better for heat transfer of Reynolds number examined, even though the heat transfer
enhancement. The gradient for 45° – NP in the average Nusselt num- enhancement is not as high as 15° – LP.
ber increases beyond Re 1000, suggesting that this configuration can From the assessment employing multiple microchannel groups,
be more effective in heat transfer for higher Reynolds number. it is consistently demonstrated that a smaller oblique angle con-
The third microchannel group is three microchannel heat sinks tributes to higher heat transfer performance, with a generally
with 200 lm nominal channel width, with the conventional micro- higher pressure drop. Even though the smaller oblique angle
channel heat sink (Con-200) up against two enhanced microchannel results in lesser flow resistance in secondary oblique channel than
(1500 lm large fin pitch) with 15° and 27° oblique angle (15° – LP the larger oblique angle, the longer oblique length contribute to
and 27° – LP). Fig. 11(a) again demonstrates that smaller oblique higher pressure drop for smaller oblique angle fin. Similar to liquid
Y.J. Lee et al. / International Journal of Heat and Mass Transfer 81 (2015) 325–336 333

Fig. 9. Comparison of (a) average Nusselt number, (b) pressure drop for microchannel heat sinks with different oblique angles (100 lm nominal channel width and 400 lm
nominal fin pitch).

Fig. 10. Comparison of (a) average Nusselt number, (b) pressure drop for microchannel heat sinks with different oblique angles (200 lm nominal channel width and 800 lm
nominal fin pitch).

flow through a bend, loss coefficient increases for sharper bend as and 1500 lm fin pitch respectively (27° – NP and 27° – LP). The
area of flow separation becomes extensive [24]. Thus, a smaller second group of microchannel heat sinks for comparison comprises
oblique angle helps to translate the diffusive oblique channel into of conventional microchannel (Con-200) and two enhanced micro-
a smoother and smaller flow area expansion from the main chan- channel heat sink (both have 45° oblique angle), with 400 lm and
nel. This can help to reduce the boundary layer separation that 800 lm fin pitch (45° – SP and 45° – NP). The experimental heat
occurs in the oblique channel and increase the secondary flow gen- transfer and pressure drop findings of the first group is presented
eration. However, one should be aware that for a smaller oblique in Fig. 12. The average Nusselt number is the highest for 27° –
angle with similar fin pitch, oblique fins will become thinner and NP followed by 27° – LP while that of Con-200 is the lowest. It also
this might compromise the structural integrity. In fact, the author suggests that the maximum temperature is lower than conven-
has struggled to fabricate the enhanced microchannel heat sinks tional microchannel. On the other hand, the pressure drop for
with 15° oblique angle through mechanical wafer cutting. 27° – NP is the highest while both 27° – LP and Con-200 display
comparable pressure drop. This indicates that shorter fin pitch is
3.2.2. The effect of oblique fin pitch variation beneficial for the heat transfer but incurs some pressure drop
The impact of oblique fin pitch variation to the enhanced micro- penalty.
channel performance is evaluated in two groups of microchannel On the other hand, Fig. 13 displays the comparison of heat trans-
heat sinks, with 200 lm nominal channel width, by varying the fer and pressure drop between Con-200, 45° – SP and 45° – NP
oblique fin pitch while other parameters including oblique angle microchannel heat sinks. It is found that 45° – SP prevails with the
are fixed. The first group of microchannel heat sinks includes a highest heat transfer performance followed by 45° – NP and Con-
conventional microchannel (Con-200) and two enhanced micro- 200. Both 45° – SP and 45° – NP heat sinks demonstrate an incre-
channel heat sinks (both have 27° oblique angle) with 800 lm ment in the heat transfer enhancement as Reynolds number rises
334 Y.J. Lee et al. / International Journal of Heat and Mass Transfer 81 (2015) 325–336

Fig. 11. Comparison of (a) average Nusselt number, (b) pressure drop for microchannel heat sinks with different oblique angles (200 lm nominal channel width and 1500 lm
nominal fin pitch).

Fig. 12. Comparison of (a) average Nusselt number, (b) pressure drop for microchannel heat sinks (200 lm nominal channel width and 27° oblique angle) with different fin
pitches.

above 1000. In addition, the pressure drop for these two heat sinks is heat sinks under the constraint of pressure drop. Among all the
higher than Con-200. Based on the evaluation with two different heat sinks tested, the enhanced microchannel with 100 lm chan-
microchannel groups, it can be concluded that smaller fin pitch is nel width and 27° oblique channel (27° – NP0 ), achieves the lowest
beneficial for higher heat transfer performance with generally total thermal resistance at the lowest pressure drop. This observa-
higher pressure drop. The introduction of oblique fins and channels tion demonstrates the effectiveness of the enhanced microchannel
aims to promote thermal boundary layer re-development and sec- through the integration of smaller channel size (hydraulic diame-
ondary flow generation, which are the critical elements in augment- ter) that has large surface area-to-volume ratio with short sec-
ing heat transfer performance. With the smaller fin pitch, more tional oblique fins. For instance, the pressure drop that is
oblique fins and channels can be inserted in a heat sink and this required to achieve 0.119 °C/W employing conventional micro-
enables the higher occurrence of boundary layer re-development channel (Con-100) can be reduced by 55% (from 45 kPa to
and secondary flow generation. 25 kPa) by adopting the enhanced microchannel (27° – NP0 ).
A similar trend is observed for microchannel heat sinks with
3.2.3. Overall performance comparison 200 lm nominal channel width. 45° – SP and 15° – LP are the
Besides evaluating the heat transfer of microchannel heat sinks two microchannel configurations that record the lowest total ther-
to the Reynolds number, the total thermal resistance of the indi- mal resistance, in comparison to other microchannel heat sinks
vidual heat sink is always compared between each other under with 200 lm nominal channel width. This finding also points out
the constant pressure drop, and, also, pump power constraint, in the effectiveness of a smaller fin pitch and a smaller oblique angle
the selection process to determine the best performing heat sink in enhancing heat transfer performance. More importantly, these
that meets the operating constraints. Fig. 14 demonstrates the configurations perform better than conventional microchannel
comparison of total thermal resistances of various microchannel heat sink with 100 lm nominal channel width at the similar
Y.J. Lee et al. / International Journal of Heat and Mass Transfer 81 (2015) 325–336 335

Fig. 13. Comparison of (a) average Nusselt number, (b) pressure drop for microchannel heat sinks (200 lm nominal channel width and 45° oblique angle) with different
oblique fin pitches.

1. Appreciable heat transfer augmentation is achieved with sili-


con-based enhanced microchannel with maximum heat trans-
fer performance enhancement at 47% when Re = 680.
Comparable pressure drop to conventional microchannel is
maintained up to Re = 500.
2. The parametric study shows that a smaller oblique angle is ben-
eficial for heat transfer enhancement. The smaller oblique angle
results in smaller flow resistance that could generate larger sec-
ondary flow rate through it.
3. Smaller fin pitch enables higher occurrence of thermal bound-
ary layer re-development and secondary flow generation, thus
further improve the heat transfer performance.
4. Among all the configurations tested, configuration 27° – NP0 is
the most effective microchannel heat sink through the combi-
nation of small hydraulic diameter (104 lm), small oblique
angle (27°) and small oblique fin pitch (400 lm). On the other
hand, enhanced microchannel heat sinks with 200 lm nominal
channel width demonstrate the capability to either equal or
enhance the thermal resistance of conventional microchannel
Fig. 14. Comparison of total thermal resistance to the pressure drop across
with 100 lm nominal channel width. This provides an opportu-
microchannel heat sinks.
nity for channel width to be relaxed in the micro-fabrication
without compromising the heat transfer performance.

pressure drop. By incorporating sectional oblique fins, both chan- Conflicts of interest
nel width and fin width of a microchannel heat sink can be relaxed,
allowing a much simpler, and economical, fabrication process. None declared.
In summary, the heat transfer of microchannel heat sinks can be
ranked as follow under constant pressure drop or pump power
constraint: Acknowledgement
27° – NP0 > 45° – NP0 > Con-100, 45° – SP and 15° – LP > 27° –
NP > 27° – LP > 45° – NP > Con-200 The authors gratefully acknowledge the MOE Academic
Research Fund (AcRF) Tier 2 research project Singapore for their
financial support. The Project title is Investigations on Micro scale
Transport Phenomena in Novel Oblique Structures with Project No.
3.3. Conclusion MOE2011-T2-2-126 (WBS No: R265-000-423-112).

Experimental validation is performed on the proposed heat


References
transfer augmentation concept, using silicon-based test vehicles.
Both local and overall characteristics in heat transfer are examined, [1] J.R. Black, Electromigration – a brief survey and some recent results, IEEE
as well as pressure drop across the heat sinks. A parametric study Trans. Electron Devices 16 (4) (1969) 338–347.
is then conducted to study the effect of variation in two critical [2] D.B. Tuckerman, R.F.W. Pease, High-performance heat sinking for VLSI, IEEE
Electron Device Lett. 2 (5) (1981) 126–129.
design parameters; oblique angle and oblique fin pitch. The follow- [3] D. Copeland, H. Takahira, W. Nakayama, Manifold Microchannel Heat Sinks
ing key conclusions are drawn from this study: Theory and Experiment, ASME – PUBLICATIONS – EEP, 1995.
336 Y.J. Lee et al. / International Journal of Heat and Mass Transfer 81 (2015) 325–336

[4] R. Prasher, J. Chang, Cooling of electronic chips using microchannel and micro- [13] Y.J. Lee, P.S. Lee, S.K. Chou, Enhanced microchannel heat sinks using oblique
pin fin heat exchangers, in: Proceedings of the 6th International Conference on fins, ASME (2009).
Nanochannels, Microchannels, and Minichannels, ICNMM, 2008, pp. 1881– [14] Y. Fan, P.S. Lee, L. Jin, B.W. Chua, A simulation and experimental study of fluid
1887. flow and heat transfer on cylindrical oblique-finned heat sink, Int. J. Heat Mass
[5] A. Bejan, M.R. Errera, Convective trees of fluid channels for volumetric cooling, Transfer 61 (2013) 62–72.
Int. J. Heat Mass Transfer 43 (17) (2000) 3105–3118. [15] Y.J. Lee, P.S. Lee, S.K. Chou, Enhanced thermal transport in microchannel using
[6] Y. Chen, P. Cheng, Heat transfer and pressure drop in fractal tree-like oblique fins, J. Heat Transfer 134 (10) (2012) 101901.
microchannel nets, Int. J. Heat Mass Transfer 45 (13) (2002) 2643–2648. [16] S.G. Kandlikar, S. Colin, Y. Peles, S. Garimella, R.F. Pease, J.J. Brandner, D.B.
[7] D. Pence, Reduced pumping power and wall temperature in microchannel heat Tuckerman, Heat transfer in microchannels – 2012 status and research needs,
sinks with fractal-like branching channel networks, Microscale Therm. Eng. 6 J. Heat Transfer 135 (9) (2013) 091001.
(4) (2003) 319–330. [17] Y.J. Lee, P.S. Lee, S.K. Chou, Numerical study of fluid flow and heat transfer in
[8] J.L. Xu, Y.H. Gan, D.C. Zhang, X.H. Li, Microscale heat transfer enhancement the enhanced microchannel with oblique fins, J. Heat Transfer 135 (4) (2013)
using thermal boundary layer redeveloping concept, Int. J. Heat Mass Transfer 41901.
48 (9) (2005) 1662–1674. [18] W. Qu, Transport phenomena in single-phase and two-phase micro-channel
[9] J. Xu, Y. Song, W. Zhang, H. Zhang, Y. Gan, Numerical simulations of interrupted heat sinks, J. Electron. Packag. 126 (2004) 213–226.
and conventional microchannel heat sinks, Int. J. Heat Mass Transfer 51 (25– [19] N.E. Todreas, M.S. Kazimi, Nuclear Systems I – Thermal Hydraulic
26) (2008) 5906–5917. Fundamentals, Hemisphere Publishing Corporation, USA, 1989.
[10] Y.J. Lee, P.S. Lee, S.K. Chou, Hotspot mitigating with obliquely finned [20] W.M. Rohsenow, J.P. Hartnett, E.N. Ganić, Handbook of Heat Transfer
microchannel heat sink – an experimental study, IEEE Trans. Compon. Fundamentals, second ed., McGraw-Hill publication, 1985.
Packag. Manuf. Technol. 3 (8) (2013) 1332–1341. [21] M.M. Yovanovich, Y.S. Muzychka, J.R. Culham, Spreading resistance of isoflux
[11] M.E. Steinke, S.G. Kandlikar, Single-phase heat transfer enhancement rectangles and strips on compound flux channels, J. Thermophys. Heat
techniques in microchannel and minichannel flows, in: Second International Transfer 13 (4) (1999) 495–500.
Conference on Microchannels and Minichannels, Rochester, NY, June, 2004, pp. [22] J.R. Taylor, An Introduction to Error Analysis: The Study of Uncertainties in
17–19. Physical Measurements, University Science Books, 1997.
[12] K.Tatsumi, M.Yamaguchi, Y. Nishio, K. Nakabe, Heat transfer and pressure loss [23] R.J. Moffat, Describing the uncertainties in experimental results, Exp. Therm
characteristics of obliquely-arranged cut-fins, in: Proceedings of the 13th Fluid Sci. 1 (1) (1988) 3–17.
International Heat Transfer Conference 2006, CD ROM (No. THE-15), http:// [24] J.F. Douglas, J.M. Gasiorek, J.A. Swaffield, L.B. Jack, Fluid Mechanics, fifth ed.,
dx.doi.org/10.1615/IHTC13.p17.150. Pearson, 2005. p. 389.

You might also like