Вы находитесь на странице: 1из 15

International Journal of Heat and Mass Transfer 100 (2016) 646–660

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


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

Effects of varying oblique angles on flow boiling heat transfer


and pressure characteristics in oblique-finned microchannels
Matthew Law, Omer Bugra Kanargi, Poh-Seng Lee ⇑
Department of Mechanical Engineering, National University of Singapore, 9 Engineering Drive 1, Singapore 117575, Singapore

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

Article history: Flow boiling experiments are conducted in oblique-finned microchannels with different oblique angles to
Received 26 August 2015 investigate its effects on two-phase heat transfer, pressure drop and instabilities. These studies are car-
Received in revised form 21 April 2016 ried out with FC-72 dielectric fluid at three different oblique angles of 10°, 30° and 50°. Three mass fluxes
Accepted 21 April 2016
ranging from 194 kg/m2 s to 386 kg/m2 s, and effective heat fluxes from 14.9 W/cm2 to 70.0 W/cm2 are
Available online 14 May 2016
tested with high-speed flow visualisations to explore the physical boiling phenomenon in the microchan-
nels. Results show that heat transfer performance increases with oblique angles from 10° to 50° due to
Keywords:
the increase in secondary channels which promote flow mixing. This causes an increase in pressure drop
Oblique fins
Oblique angles
from 10° to 30° oblique angles due to the increase in flow separation and flow diversion. Pressure drop,
Flow boiling however, is almost negligible when the oblique angles are increased from 30° to 50°. As a result, the 50°
Pressure drop oblique-finned microchannels presents an attractive option for further parametric studies.
FC-72 Ó 2016 Elsevier Ltd. All rights reserved.

1. Introduction parameters as the porous structure promotes bubble nucleation.


The heat transfer performance is found to be dependent on heat
Endless development in the semiconductor and electronics sec- and mass fluxes and is linked to the transition of flow boiling
tors has triggered an upsurge in the miniaturisation of electronic mechanisms. Pressure drop and flow instabilities are also found
components. As a result, heat flux dissipation of these components to be significantly dependent on inlet subcooling and heat flux.
increased exponentially, and the conventional heat management Sarwar et al. [7] studied the effects of micro/nano porous sur-
techniques are no longer effective. This is where two-phase face coated vertical tubes on CHF enhancement in subcooled flow
microchannels cooling is steadily gaining widespread attention boiling using filtered degassed water. The authors reported that
due to its ability in high heat flux removal over a relatively com- the microporous surface coated vertical tubes showed greater
pact area. In contrast with its macrochannels counterpart, CHF enhancement as compared to the case of the nanoporous sur-
microchannels present large heat transfer areas per unit volume. face coated vertical tubes, which is found to have insignificant
Due to the vast potential of this technology, various research enhancement. One of the factors which contribute to this finding
efforts have been undertaken to better understand the flow boiling is that the larger particle sized coating produced more nucleation
heat transfer phenomena in microchannels [1–5]. Besides that, sites, and as a result, create pores with better water entrapment
numerous heat transfer enhancement techniques, which mainly capabilities.
look into stabilising the flow boiling process and reducing pressure Prajapati et al. [8] experimentally investigated and compared
drop penalty, can be found in the literature. the flow boiling performance of three different microchannels con-
Deng et al. [6] tested a novel type of reentrant porous figurations using deionised water. The microchannels configura-
microchannels with X-shaped configurations using ethanol, which tions tested are uniform cross-section, diverging cross-section
looked into the effects of operating parameters on flow boiling heat and segmented (oblique) finned microchannels. It was found that
transfer, pressure drop and two-phase flow instabilities. The the segmented finned microchannels demonstrates the highest
authors found that the onset of nucleate boiling of ethanol is initi- heat transfer coefficient with negligible higher pressure drop com-
ated at relatively low wall temperature superheats at all operating pared to other two configurations of channels for the entire range
of operating conditions. At higher heat flux, the authors reported
bubble clogging and flow reversal in uniform and diverging
⇑ Corresponding author. Tel.: +65 6516 4187; fax: +65 6779 1459.
cross-section microchannels, whereas the segmented finned
E-mail addresses: matthew.law@u.nus.edu (M. Law), bugrakanargi@u.nus.edu
(O.B. Kanargi), mpelps@nus.edu.sg (P.-S. Lee).

http://dx.doi.org/10.1016/j.ijheatmasstransfer.2016.04.077
0017-9310/Ó 2016 Elsevier Ltd. All rights reserved.
M. Law et al. / International Journal of Heat and Mass Transfer 100 (2016) 646–660 647

Nomenclature

A area Greek symbols


cp specific heat capacity g fin efficiency
d distance from temperature probe to channel surface h oblique angle
Dh hydraulic diameter q density
f friction factor r standard deviation
FOM figure of merit
G mass flux Subscripts
h heat transfer coefficient amb ambient
H height av g average
I current ch channel
k thermal conductivity cs cross section
l length Cu copper
L length of heat sink eff effective
m fin parameter f fluid
m _ mass flow rate fin fin
N number of in inlet
p perimeter l liquid
P pressure loc3 location of the third (most downstream) temperature
DP pressure drop sensor
q heat transfer rate loss loss
q00 heat flux max maximum
Rth thermal resistance ob oblique cut
t time of flow visualisation frame out outlet
t fin fin thickness sat saturated
T temperature sub subcooled
V voltage supplied supplied
w width total total
W width of heat sink unfin unfinned
z distance from microchannels inlet to the location of the wall wall
most downstream temperature sensor wall3 third (most downstream) wall
x¼0 location of zero thermodynamic equilibrium quality

geometry completely eliminated the problem of bubble clogging drop. Boiling curves, two-phase heat transfer coefficients, pressure
allowing smooth and easy passage of growing bubbles. drops and flow boiling pressure instabilities are presented for each
Wang et al. [9] examined the effect of inclination on the case of oblique angle and compared for a range of mass fluxes. The
convective boiling performance of a microchannel heat sink using figure of merit to characterise the overall performance of each obli-
the dielectric fluid HFE-7100. The inclinations span from +90° (ver- que angle is also discussed.
tical upward) to 90° (vertical downward), with a 45° angle incre-
ment. Experimental investigations found that heat transfer 2. Experimental setup
performance for the upward and horizontal configurations is
comparable for low mass flux and that for the 45° upward config- 2.1. Flow loop
uration is significantly better than the other inclinations. Further-
more, flow visualisations suggested that heat transfer A schematic diagram of the flow loop is given in Fig. 1. The
enhancement for the upward inclination is mainly due to the descriptions are similar to Law et al. [12] and Law and Lee [10],
increase in vapour slug velocity. however, there are a few changes made in the temperature and
Law and Lee [10] compared the oblique-finned microchannels pressure sensors, as listed in Table 1. The accuracy of the various
with the straight-finned microchannels using FC-72 in terms of sensors and equipment used in this study is given in Table 3 under
their two-phase heat transfer performance, critical heat flux, tem- Section 3.3.
perature uniformity, pressure drop and flow boiling instability
characteristics. The authors reported that the oblique fins not only 2.2. Microchannels test section
offer improved heat transfer performance, they also highly sta-
bilise the flow boiling process by allowing parallel microchannels The test section consists of three main components, as labelled
to ‘‘communicate” with their neighbours. However, these various in Fig. 2.
enhancements come with quite a substantial two-phase pressure
drop penalty, which can be controlled by careful modifications in 1. Top cover: Made of transparent polycarbonate to enable unob-
the oblique fin structure. Lee et al. [11] found that the oblique structed high-speed visualisation of the microchannels.
angle is one of the parameters which greatly influenced heat trans- 2. Housings: Consists of a top housing and bottom housing which
fer and pressure drop performances during single-phase flow in hold the microchannel heat sink, and a base, which holds the
the oblique-finned microchannels. Therefore, this forms the moti- entire assembly in place. The housings are made of Teflon to
vation behind the present study, where the oblique angles are var- reduce heat loss.
ied to investigate its effect on two-phase heat transfer and pressure 3. Microchannel test piece: Made of copper.
648 M. Law et al. / International Journal of Heat and Mass Transfer 100 (2016) 646–660

Fig. 1. Schematic diagram of the experimental flow loop.

In the test section, degassed FC-72 (Tsat = 56 °C at 1 atm) flows are tapped from a shunt resistor connected to the heaters for a
into the microchannels via the inlet port to the inlet plenum and more accurate measurement of the heat input.
exits the microchannels via the outlet plenum and outlet port. A
pair of inlet and outlet manifolds with depth, length and width 2.3. Microchannels geometry
of 1.2 mm, 23 mm and 32 mm, respectively, is fabricated next to
the inlet and outlet of the microchannels to ensure that the flow The oblique angle of the oblique-finned microchannels in the
is evenly distributed across the microchannel arrays. Two temper- present study are varied at 10°, 30° and 50° to examine its effect
ature measurements are taken at the inlet and outlet plenums as on heat transfer and pressure drop. This variation of oblique angles
the inlet and outlet fluid temperatures, respectively. Three holes resulted in 585, 897 and 1014 oblique fins, respectively. The fins
are drilled until the centre of the test piece at 8.8 mm below the are formed on the surface of a copper block using wire-cut electri-
surface of the microchannels to measure the streamwise wall tem- cal discharge machining (EDM) process. The surface roughness of
perature distribution. Two small holes of 1/8 inch (approx. 3.2 mm) the microchannels is not measured, however, it is assumed to be
are drilled on the other side of the inlet and outlet plenums for the similar for all three microchannels as they are fabricated on the
differential pressure ports. The location of the various ports is same wire-cut machine using the same spool of metal wire. Typical
shown in Fig. 3. surface roughness for wire-cut EDM is around 2.0 lm [13].
Heating is simulated in the test section by four cartridge heaters Schematic diagrams in Fig. 4 show the differences in the oblique
inserted into the four holes at the bottom of the test piece, as angles of the three different oblique-finned geometries. Apart from
shown in the schematic in Fig. 3. Voltage and current readings the oblique angles and the number of fins, the other dimensions of
M. Law et al. / International Journal of Heat and Mass Transfer 100 (2016) 646–660 649

Table 1
List of equipment and sensors.

Device Brand Model Specification


Fluid reservoir McMaster-Carr Portable wide mouth ASME 01 EA 1 gallon (0.0038 m3)
pressure tank
Filter Swagelok SS-4FW-15 15 microns
Gear pump Micropump GA-T23.PFS.A –
Pump drive Cole Parmer 75211-15 –
Flow metre McMillan 104 Flo-Sen type 5HY Flow: 50–500 ml/min
(0.00005–0.0005 m3/min)
Max temp: 85 °C
RTD Omega P-M-1/3-1/8-3-0-TS-3 50 °C–250 °C
Inlet and outlet pressure transmitters Omega PX409-015G5V 0–100 kPa
Diff pressure transducer Omega PX409-2.5DWU5V 0–17 kPa
Power supply unit Ametek Sorensen XG 300-5.6 Voltage: 0–300 V
Current: 0–5.6 A
Condenser Thermatron 735SPC2A01 –
Degassing unit Gastorr PG-32S –
Data acquisition system National Instruments Pressure and flow data: cDAQ-9178 –
Temperature data: PXIe-1073 –
Agilent Voltage and current: 34970A –
High-speed camera Photron FASTCAM SA5 1000K-M3 7000 fps
1024  1024 pixels

A total of three flow rates are tested at 100 ml/min (0.0001 m3/
min), 150 ml/min (0.00015 m3/min) and 200 ml/min (0.0002 m3/
min), which corresponds to an average inlet mass flux of 194 kg/
m2 s, 290 kg/m2 s and 386 kg/m2 s, respectively. The inlet temper-
ature of FC-72 is maintained at a subcooled condition at
29.5 °C ± 0.5 °C, for heat fluxes of 14 W/cm2 up to 70 W/cm2.

3.1. Heat loss characterisation

Detailed heat loss characterisation procedures, as described in


Law et al. [12], are employed in the present study. As the base of
the microchannels and the inlet and outlet manifolds are inline,
the contact between the test section material and the fluid in the
manifolds are very small. Thus, the heat loss due to the contact
between these two surfaces is negligible. Moreover, the test sec-
tion is properly insulated with Teflon housings, which further min-
imises heat loss through the housings. Therefore, the heat loss
considered in the present study is assumed to be a function of
the average wall temperatures of the microchannels and the ambi-
ent temperature.
Fig. 2. Test section. The heat loss for the three microchannels geometries is found to
have a strong linear relationship with the difference between the
average wall and ambient temperatures, as shown in Fig. 5. The
the oblique-finned microchannels are kept constant, as given in corresponding heat loss equations are given in Eqs. (1)–(3).
Table 2. The accuracy of the measurements is given in Table 3
10 : qloss ¼ 0:1785ðT wall;av g  T amb;av g Þ½W ð1Þ
under Section 3.3.

30 : qloss ¼ 0:1745ðT wall;av g  T amb;av g Þ½W ð2Þ


3. Experimental procedures
50 : qloss ¼ 0:1754ðT wall;av g  T amb;av g Þ½W ð3Þ
Detailed experimental procedures are listed below:

1. Valves on the main flow loop are opened and those on the 3.2. Data reduction
degassing loop closed before activating the pump.
2. The required flow rate is set by varying the speed of the pump 3.2.1. Pressure drop
and the power supplied to the cartridge heaters is set to the As the differential pressure ports are located upstream and
desired value after the flow rate and the inlet fluid temperature downstream of the microchannels in the inlet and outlet plenum,
are stabilised. the pressure drop measurement represents the combined losses
3. A 15-min waiting time is observed in each test run, when all due to the frictional loss in the microchannels and minor losses
temperature readings became stable. Temperature and pressure across the bends from the inlet plenum to the inlet manifold and
data are logged at a frequency of 25 Hz and averaged over a 60-s outlet manifold to the outlet plenum, as well as abrupt contraction
period. and expansion from the inlet and manifolds to the microchannels.
4. The heat flux is then increased for the next test condition, and Calculations were made based on the methods described in [14]
procedure 3 is repeated. and it was identified that the other pressure losses apart from the
650 M. Law et al. / International Journal of Heat and Mass Transfer 100 (2016) 646–660

Fig. 3. Location of inlet, outlet, temperature and pressure ports (dimensions in mm).

Table 2
Actual dimensions of microchannels.

Parameter Unit Nominal Oblique angle, h


dimension
10° 30° 50°
Material – – Copper
Footprint area, A mm2 – 25  25
Heat transfer area, Atotal cm2 – 35.5 31.8 30.0
Primary channels
Number of channels, – – 40
Nch
Channel width, wch mm 0.30 0.32 0.30 0.30
Channel depth, Hch mm 1.20 1.18 1.19 1.19
Secondary channels
Oblique-cut width, wob mm 0.15
Fins
Number of fins, Nfin – – 585 897 1014
Fin thickness, tfin mm 0.30 0.29 0.30 0.30
Fin length, Lfin mm 0.70 0.65 0.68 0.68

Table 3
List of uncertainties.

Variable Uncertainty
Channel and fin dimensions ±0.001 cm
Flow rate 1.0% full scale
Temperature measurements ±0.14 °C at 25 °C to ± 0.44 °C at 200 °C
Voltage measurement 0.1% rated
Current measurement 0.2% rated
Inlet pressure transmitter 0.08% full scale
Differential pressure transducer 0.08% full scale

frictional drop were less than 5% of the total drop and hence were
neglected. Therefore, the pressure drop data reported here are as
directly obtained from the differential pressure transducer.

3.2.2. Heat transfer


The effective heat absorbed by the fluid, after taking into
account heat loss, as described in Section 3.1, is given by
qeff ¼ qsupplied  qloss ð4Þ
Fig. 4. Schematic diagrams of different oblique angles.
M. Law et al. / International Journal of Heat and Mass Transfer 100 (2016) 646–660 651

between the inlet and outlet pressures, which can be justified


based on very low measured pressure drop, as in Eq. (11).
T f ¼ T sat in saturated region; where T sat ¼ f ðPout Þ ð11Þ
The thermal properties of FC-72 are shown in Eqs. (12) and (13),
whereas the fluid density is given in Eq. (14), obtained from Celata
et al. [15] and Hua et.al [16]. Units of P is bar and T is °C.

T sat ¼ 51:6304 þ 24:098 ln P þ 0:5967=P  4:2899P þ 0:2159P2


ð12Þ

cp ¼ 1000ð1:0096 þ 1:55  103 TÞ ð13Þ

q ¼ 1:618  1011 T 6  1:269  109 T 5 þ 1:027  106 T 4


 4:239  104 T 3 þ 4:973  102 T 2  4:014T þ 1:755  103
Fig. 5. Heat loss characterisation curves. ð14Þ
The local heat transfer coefficient can then be calculated using
where Newton’s law of cooling in Eq. (15).
qeff
qsupplied ¼ VI ð5Þ hloc3 ¼ ð15Þ
Atotal ðT wall3  T f Þ
V and I are obtained by measuring the voltage drop and current
where Atotal is the total convective heat transfer area of the heat
across the shunt resistor.
sink, given by
The associated heat flux in Eq. (6) is calculated based on the
base area of the microchannel heat sink, which is equal to the foot- Atotal ¼ Aunfin þ gAfin ð16Þ
print area of the test section. This is also the reported heat flux that
Aunfin is the unfinned surface area at the bottom of the channels
the heat sink can dissipate.
and Afin is fin area, as in Eqs. (17) and (18).
qeff
q00eff ¼ ð6Þ Aunfin ¼ Base area of heat sink  Base area of oblique fins
A
¼ WL  Nfin wfin lfin ð17Þ
where

A ¼ WL ð7Þ Afin ¼ Nfin Hfin pfin ð18Þ

Subcooled FC-72 (Tf,in < Tsat) is pumped into the test section for Local fin efficiency is used to account for the drop in tempera-
all test conditions. The microchannels can therefore be divided into ture along the fin. An adiabatic fin tip condition is assumed due
two regions: an upstream subcooled region and a downstream sat- to the low thermal conductivity of the Polycarbonate top cover
urated region. The location of zero thermodynamic equilibrium and the corresponding fin efficiency is given as
quality (x = 0) serves as a dividing point between the two regions.
tanhðmHfin Þ
The length of the two regions can be evaluated based on energy g¼ ð19Þ
mHfin
balance as

_ p ðT sat;x¼0  T f ;in Þ where m is the fin parameter, given by


mc
Lsub ¼ ð8Þ  1
q00eff W hloc3 pfin 2
m¼ ð20Þ
kCu Acs
and
Eqs. (19) and (20) are iteratively solved to obtain hloc3. This
Lsat ¼ L  Lsub ð9Þ
method is widely used in the literature by Qu and Siu-Ho [17], Har-
where Tsat,x=0 is the saturation temperature at the location of zero irchian and Garimella [18], and Balasubramanian et al. [19].
thermodynamic equilibrium quality. In the present study, Tsat,x=0 As direct wall temperature measurements at the bottom of the
is evaluated as a function of the measured inlet pressure, Pin, as channel are not available, the one-dimensional heat conduction
pressure drop across the subcooled region is small. assumption is used to extrapolate the temperature readings of
Local heat transfer coefficient is computed only at the location the temperature sensor. Thus, the local wall temperature at the
of the most downstream temperature sensor, which relates to most downstream location is given by Eq. (21). This assumption
the greatest amount of saturated boiling. For uniform heat flux is available in the literature and is used by Prajapati et al. [8], Wang
conditions, the bulk fluid temperature in the single-phase region et al. [9], Lee et al. [20], Qu and Mudawar [21].
vary linearly with the energy balance.
q00eff d
q00eff Wz T wall3 ¼ T loc3  ð21Þ
T f ¼ T f ;in þ ð10Þ kCu
mc _ p

where z is the distance from the inlet of the microchannels to the 3.3. Experimental uncertainties
location of the most downstream temperature sensor.
Within the saturated region, the bulk fluid temperature is The uncertainties of the various sensors used in this study are
equals to the local saturation temperature which is taken corre- listed in Table 3. A standard error analysis adopted from Taylor
sponding to the local pressure obtained as a linear interpolation [22] which used the Kline–McClintock method [23] revealed the
652 M. Law et al. / International Journal of Heat and Mass Transfer 100 (2016) 646–660

 
Table 4 @ @ @e e e2
Range of uncertainties in heat transfer coefficient and pressure measurements under ðqeui Þ ¼ ae leff þ C 1e Gk  C 2e q ð26Þ
@xi @xj @xj k k
various conditions.

G ðkg=m2 sÞ Range of uncertainty where Sk is a user defined source term for the turbulence kinetic
q00eff (%) hloc3 (%) P in (%) DP (%) energy and Gk is the turbulence kinetic energy generation due to
the mean velocity gradients, which is given as
194 0.23 5.2–14.5 0.50–1.2 0.57–1.3
 
290 0.23 4.6–9.0 0.29–1.0 0.39–0.89 @ui @uj @ui
386 0.23 3.8–6.6 0.19–0.78 0.27–0.64 Gk ¼ lt þ ð27Þ
@xj @xi @xj
The turbulent and effective viscosities, lt and leff are defined as
2
k
lt ¼ qC l ð28Þ
range of experimental uncertainty in the heat transfer coefficient e
calculations and pressure data as given in Table 4 for various mass
fluxes. The maximum uncertainty in heat transfer coefficient, inlet leff ¼ l þ lt ð29Þ
pressure and pressure drop measurements are 14.5%, 1.2% and
1.3%, respectively, all occurring at the smallest mass flux condition, Finally, C 2e , the modified version of the constant C2e is written
G = 194 kg/m2 s. The uncertainty in the heat transfer coefficients as
and pressure drop data generally decrease with an increase in mass
C l g3 ð1  gÞ=g0
flux and heat flux. C 2e ¼ C 2e þ ð30Þ
1 þ bg3
4. Numerical analysis where g is the ratio of the user defined turbulence kinetic energy to
the rate of the dissipation of the turbulence kinetic energy, g ¼ Sk =e.
As the modelling of two-phase flow is not straightforward, it is When the value of g is not very high; in other words, when the
assumed that the flow in the two-phase region follows that in the flow is weakly or moderately strained, the values of C 2e and C2e are
single-phase region. Furthermore, two-phase flow modelling with
close to each other. Therefore, the RNG model yields results very
phase change is presently still not readily available, thus the
close to the standard k–e turbulence model. When the value of g
authors are willing to infer that the flow field in single-phase might
is higher than g0, on the other hand, the flow is strongly strained
be valid for two-phase condition as well. Apart from that, as the
and the value of C 2e is less than C2e that would be predicted by
fluid enters the microchannels in single-phase, the authors believe
the standard k–e turbulence model. A smaller value of C 2e in Eq.
that it is reasonable for the adoption of single-phase modelling to
(26) yields a lower destruction of e than normal. Since the destruc-
represent the two-phase flow phenomena in the microchannels.
tion of e decreases, e actually increases, yielding a lower k. An
Thus, single-phase numerical simulation for flow in a channel-fin
pair is employed in the present study to investigate the flow veloc-
ity and amount of secondary flow in the oblique cuts so as to better
explain the heat transfer performance of the different oblique
angles.

4.1. Assumptions and models

The numerical analyses of the single phase flow of FC-72 dielec-


tric liquid through oblique-finned microchannels are performed at
steady state. Reynolds averaged continuity and Navier–Stokes
equations are solved using the RNG k–e turbulence model [24],
which is capable of resolving rapidly strained flows more accu-
rately than the standard k–e turbulence model.
Depending on the stated assumptions and the turbulent model
selected, the simplified, Reynolds averaged continuity and Navier–
Stokes equations can be written in tensor notation as
@
ðqui Þ ¼ 0 ð22Þ
@xi
   
@ @p @ @u @u
ðqui uj Þ ¼  þ l i þ j  qu0i u0j ð23Þ Fig. 6. Velocity contours of flow inside the oblique-finned microchannels.
@xj @xi @xj @xj @xi
where the turbulent stress term is expanded as Table 5
 
@ui @uj Velocity distribution at different locations of an oblique fin.
q u0i u0j ¼ lt þ ð24Þ
@xj @xi
In order to obtain the turbulence kinetic energy, k and the rate
of dissipation of the turbulence kinetic energy, e, transport equa-
tions are solved. The transport equations of the RNG k–e model
Average velocity [m/s]
are modified from those of the standard k–e model and are given as
  10° 30° 50°
@ @ @k
ðqkui Þ ¼ ak leff þ Gk  qe þ Sk ð25Þ 0.137 0.158 0.172
@xi @xj @xj
M. Law et al. / International Journal of Heat and Mass Transfer 100 (2016) 646–660 653

Fig. 7. Velocity vectors of flow through a secondary channel.

Fig. 9. Pressure drop data repeatability for G = 194 kg/m2 s.


Fig. 8. Temperature data repeatability for G = 194 kg/m2 s.

increase in e and a decrease in k results in a lower lt, as can be Instead of using the RNG k–e turbulence model with the stan-
evaluated with Eq. (29). dard wall functions, enhanced wall treatment (EWT) option, pro-
The analytically derived constants of the RNG theory are vided by ANSYS Fluent, is selected. EWT is a near wall modelling,
Cl = 0.085, C1e = 1.42, C2e = 1.68, g0 = 4.38, b = 0.012, ak = ae = 1.393 which is a combination of the two-layer model (the viscous sub-
[25]. layer and the fully turbulent layer) and the enhanced wall func-
654 M. Law et al. / International Journal of Heat and Mass Transfer 100 (2016) 646–660

Table 6
Velocity and mass flow through secondary channels.

Parameter 10° 30° 50°


Average velocity [m/s] 0.090 0.055 0.032
Mass flow rate [kg/s] 0.277 0.255 0.097

Fig. 11. Effect of different oblique angles on heat transfer coefficient with effective
heat flux (a) low, (b) medium and (c) high mass fluxes.

addition to that, utilising a very thin mesh in the vicinity of the


wall, makes it possible to accurately resolve the viscous sublayer.

4.2. Boundary conditions

To simplify the problem and reduce computational efforts, a


Fig. 10. Boiling curves as a function of oblique angles for the oblique-finned single channel with a single row of oblique fins are modelled on
microchannels at (a) low, (b) medium and (c) high mass fluxes. 2-D domains and periodic boundary condition is employed for all
three cases considered. The no-slip boundary condition is applied
on the fin walls and pressure outlet boundary condition is assigned
tions. The viscous sublayer can be considered up to y+ = 11.225 to the outlets of the microchannel. The inlet and outlet sections are
with the least error, while the log law is applied for y+ > 30. The extended in order to simulate the experimental conditions as accu-
buffer zone, which lies in between these two regions, cannot be rate as possible for the calculation of the pressure drop. An average
resolved accurately with any of these models. EWT, on the other inlet velocity of 0.089 m/s is set at the entrance of the microchan-
hand, utilises a hybrid wall function for the buffer layer [24]. In nel, which corresponds to a flow rate of 3.75 ml/min per
M. Law et al. / International Journal of Heat and Mass Transfer 100 (2016) 646–660 655

Fig. 12. High-speed visualisations of nucleate boiling in the oblique-finned microchannels for different oblique angles at G = 194 kg/m2 s.

Fig. 13. High-speed visualisations of the transition of nucleate boiling to convective boiling at G = 194 kg/m2 s and q00eff = 30.64 W/cm2.

of the flow domain. A comparison of the mass flow rates through


the main channels showed that almost 60% of the channels of an
oblique-finned minichannel heat sink were unaffected by the flow
maldistribution and the use of the periodic boundary condition for
the modelling of a single channel in the numerical analysis yielded
very close results to the experiments. Therefore, in this study,
merely a single channel is modelled with periodic boundary condi-
tions to model the flow through 2-D oblique-finned microchannel
heat sink.

4.3. Computational grid

The grids for the computational domains are generated in


ANSYS Mesh 15. Unstructured quadrilateral cells are generated
Fig. 14. Number of oblique fins and heat transfer area at different oblique angles. with an average mesh sizing of 7.5 lm. In the vicinity of the fin
walls, a very fine boundary layer mesh with 13 layers with a first
layer thickness of 1.0 lm is created. The thickness of the first layer
microchannel and 150 ml/min for the entire heat sink for 1.2 mm of the boundary layer mesh is specifically determined so that the
channel height. value of y+ is ensured to be less than 0.5, which is a strict require-
In planar oblique-finned heat sinks, regardless of mini or ment for the EWT option. This requirement is vital for the EWT to
microchannel size, the negative effects of flow migration are resolve the viscous boundary layer accurately [24]. The grid inde-
observed due to the strong secondary velocity components. This pendence of the numerical results is also validated by refining
phenomenon, entitled as the edge effect, is studied on cylindrical the utilised mesh a few times. The final grid did not yield a change
oblique-finned minichannel heat sink by Fan et al. [26]. It is shown in DP for more than 0.12% compared to a lower resolution grid,
that the negative effects of flow migration had more adverse which results in a final mesh count for the models of around
effects on the rows close to the edges than the rows in the middle 750,000 for each case.
656 M. Law et al. / International Journal of Heat and Mass Transfer 100 (2016) 646–660

Importing the grid data into Fluent; 2D, double precision, steady rate in the secondary channels then reduces with oblique angle
state solver is selected. The pressure–velocity coupling is per- increment, as depicted in Fig. 7(b) and (c). Average flow velocity
formed by the SIMPLEC algorithm and the second order upwind and mass flow in the secondary channels is also the highest in
scheme is used to discretise the domain. The absolute convergence the 10° angle case, followed by 30° and 50°, as listed in Table 6.
criterion for continuity, x-velocity and y-velocity are 106. Apart Although the 10° oblique angle provides the most amount of sec-
from the convergence criterion for termination, the mass- ondary flow, it does not translate to better heat transfer perfor-
weighted values of the inlet pressure are also monitored at every mance, as shown in the boiling curves in Fig. 10. Therefore, it can
iteration to make sure that the steady state conditions are reached. be inferred that the amount of secondary flow needs to be regu-
lated in order to obtain maximum heat transfer enhancement from
the oblique fins. Apart from that, lesser secondary flow in the 50°
5. Results and discussion oblique fins could bring about higher heat flux in the fin walls.

5.1. Numerical results 5.2. Repeatability test

Fig. 6 shows the velocity contours of the oblique-finned Temperature and pressure drop data repeatability are shown in
microchannels for 10°, 30° and 50° angles. The average velocity Figs. 8 and 9, respectively, for the two extreme cases of 10° and 50°.
of the flow in the main channel increases with oblique angle, as All four sets of data are found to be repeatable at maximum devi-
seen in Table 5. Due to this higher velocity, it is deduced that more ations of 3.7 °C for the most downstream temperature reading and
heat can be transferred from the fins to the fluid as oblique angle 350.2 Pa for the differential pressure reading. The trends for the
increases, which contributes to the heat transfer enhancements data are also found to be repeatable.
for larger oblique angles. The higher flow velocity will bring about
increased rate of heat transfer for a fixed mass flux and heat flux, 5.3. Heat transfer performance
which in turn, will reduce the wall temperature of the
microchannels. 5.3.1. Boiling curves
Fig. 7 shows the velocity vectors of flow through one secondary Fig. 10 shows the comparison of the boiling curves at
channel for 10°, 30° and 50° angles. It can be observed that the 10° G = 194 kg/m2 s, G = 290 kg/m2 s and G = 386 kg/m2 s for the
configuration corresponds to the highest amount of secondary oblique-finned microchannels at 10°, 30° and 50° angles. The boil-
flow, which presents significant contribution to the flow field. Flow ing curves are plotted between wall temperature at the most

Fig. 15. Effect of oblique angles on the liquid-film thickness surrounding the oblique fins at G = 194 kg/m2 s.
M. Law et al. / International Journal of Heat and Mass Transfer 100 (2016) 646–660 657

downstream location (Twall3), instead of wall superheat (DTsat), and boiling dominated mechanism. The high-speed visualisation
applied effective heat flux for direct and easier comparison. images in Fig. 12, taken at the location of the most downstream tem-
The boiling curves tend to merge into a single plot during perature sensor where the heat transfer coefficient is calculated,
single-phase flow and diverge after the onset of nucleate boiling confirm this observation. At higher effective heat fluxes, nucleation
(ONB). It can be seen that the 10° oblique-finned microchannels begins to be suppressed as bubbles elongate and form vapour slugs.
provides the least effective heat transfer compared to the other This is reported by Law et al. [12] as the ‘‘tail-end of nucleate boil-
two cases at all mass fluxes tested. Heat transfer is enhanced as ing” which is a combination of nucleate boiling and convective boil-
the oblique angle is increased from 10° to 50° for all mass fluxes, ing via thin-film evaporation, as seen in the high-speed visualisation
however, this enhancement becomes less apparent as the oblique images in Fig. 13 for the case of 10° oblique angle.
angle is increased from 30° to 50° at low mass flux. The high-speed images in Fig. 12 suggest that nucleation activity
is the most significant in the 10° oblique angle configuration, char-
5.3.2. Heat transfer coefficient acterised by the large number of bubbles in the microchannels. This
Fig. 11 shows the variation of heat transfer coefficient for differ- might be caused by the additional fin area offered by the 10°
ent oblique angles at low, medium and high mass fluxes. The heat oblique-finned configuration (Fig. 14). As a result, the larger
transfer coefficients show a strong relationship with heat flux at nucleation activity may be the contributing factor behind the much
low effective heat fluxes, which is an indication of a nucleate

Fig. 16. Effect of different oblique angles on pressure drop with effective heat flux Fig. 17. Inlet pressure fluctuations for different oblique angles at (a) low, (b)
at (a) low, (b) medium and (c) high mass fluxes. medium and (c) high mass fluxes.
658 M. Law et al. / International Journal of Heat and Mass Transfer 100 (2016) 646–660

smoother decline in the heat transfer coefficient trends for the 10° with the straight fins. The amplitude of fluctuation of the inlet
case in the low effective heat flux region, where nucleate boiling pressure is measured using the standard deviation equation, and
dominates. their values are included in the plots.
It can be seen in Fig. 11 that oblique angles have quite a strong It can be seen that the instabilities in the oblique-finned
influence on heat transfer coefficient. Heat transfer coefficient microchannels at all oblique angles are much lower than their
increases with oblique angle from 10° to 50°. This heat transfer straight-finned counterpart. At low, medium and high mass fluxes,
coefficient enhancement at larger oblique angles may be due to the inlet pressure instabilities is reduced from 1.6 to 2.0 times, 1.7
the higher number of oblique fins, as shown in Fig. 14. This is likely to 2.3 times and 2.2 to 3.1 times, respectively. This indicates that
to produce a greater degree of disruptions and re-initialisations of the oblique-finned microchannels configuration is able to offer
the thin liquid-film, which is revealed to be beneficial for two- more stable flow boiling, regardless of oblique angles. It is also
phase convective boiling heat transfer [10]. As shown in Fig. 15, noted from Fig. 17 that the 10° oblique angle geometry has the
the liquid-film thickness, characterised by the dark regions
adjoined to the fin walls, diminishes at larger oblique angles at
the same operating conditions, indicative of a more effective con-
vective boiling mechanism. Apart from the number of fins, the
more slender profile of the 10° oblique fins causes a significant por-
tion of heat to be transferred from the base of the microchannels to
the flowing fluid, instead of from the fins. This can be attributed to
the lowest fin efficiency for the smallest oblique angle relative to
its counterparts, which bring about less effective heat transfer
[27]. Thus, this can also be considered as one of the reasons behind
the much lower heat transfer performance of this particular
oblique-finned configuration compared to its counterparts.

5.4. Pressure characteristics

5.4.1. Two-phase pressure drop


The effect of oblique angle on two-phase pressure drop is
shown in Fig. 16 for different operating conditions. Generally, it
can be seen that the pressure drop trend is linearly increasing with
effective heat flux for all three oblique angles, which is the result of
the vapour acceleration component and the two-phase frictional
component. In addition, the slope of the pressure drop plots for
the 10° oblique-finned microchannels is not as steep as that of
its 30° and 50° counterparts.
The increment of oblique angle from 10° to 30° causes a rather
substantial increase in the pressure drop at all mass fluxes, which
is more pronounced as effective heat flux increases. Interestingly
however, as the oblique angle is further varied from 30° to 50°,
the pressure drop is seen to be almost negligible, especially for
the case of medium and high mass fluxes. There are three explana-
tions for this observation. One explanation is that on the increase
in the occurrence of boundary layer interruptions brought about
by the increase in oblique fins, although this might be beneficial
for heat transfer. Another explanation is that of the larger increase
in flow velocity when the oblique angle is varied from 10° to 30°,
compared to 30° to 50°. Another possible explanation might be
the case of minimum cross sectional area of the flow. Based on
Fig. 15, the minimum cross sectional area corresponds to the area
of the main channels (0.30 mm) for oblique fins with angles of 30°
and 50°. On the other hand, for the 10° oblique fins, the minimum
cross sectional area is found to be 1.5 times larger (0.45 mm) than
its 30° and 50° counterparts. Thus, this can be an additional reason
for the sudden increase in pressure drop.
The effort taken by the two-phase liquid and vapour compo-
nents to flow into the secondary channel also contributes to the
measured pressure drop. For the case of the 10° oblique angles,
the fairly ‘‘streamlined” fins makes it easier to generate secondary
flow and thus, more secondary flow can be obtained. Whereas for
the cases of the 30° and 50° oblique angles, fluid flowing into the
secondary channels has to undergo a sharp turn into the oblique
cuts, which require more effort and thus increasing pressure drop.

5.4.2. Inlet pressure instabilities


The inlet pressure instabilities of the oblique-finned microchan- Fig. 18. FOM parameters for different oblique angles at (a) low, (b) medium and (c)
nels at different oblique angles are given in Fig. 17 as a comparison high mass fluxes.
M. Law et al. / International Journal of Heat and Mass Transfer 100 (2016) 646–660 659

highest inlet pressure fluctuation among the three configurations, 3. The increment of oblique angle from 10° to 30° causes a sub-
which might be due to greater amount of secondary flow that stantial increase in the pressure drop at all mass fluxes, which
causes disturbance to the flow in the main channels. This higher is more pronounced as effective heat flux increases. However,
instability translates to a lower heat transfer performance for the as the oblique angle is further varied from 30° to 50°, the pres-
10° oblique-finned microchannels. sure drop is almost negligible.
4. The inlet pressure instabilities for the oblique fins are reduced
5.5. Figure of merit compared to the straight fins, regardless of oblique angles.
5. The 10° oblique-finned microchannels has the highest figure of
The figure of merit (FOM) for the three oblique angle configura- merit than its 30° and 50° counterparts. Interestingly, however,
tions with fixed mass flux, as depicted in Fig. 18, is defined as the 30° and 50° oblique angles have almost comparable figures
follows: of merit.
1
FOM ¼ ð31Þ
Rth  f
Acknowledgements
where
T max  T in The authors would like to thank the Ministry of Education Sin-
Rth ¼ ð32Þ gapore for the financial support granted through the Academic
qeff
Research Fund (AcRF) Tier 2 (WBS No: R-265-000-423-112).
DPDh ql
f ¼ ð33Þ References
2G2 L
According to Ghaedamini et al. [28], the FOM parameter pro- [1] I. Pranoto, K.C. Leong, An experimental study of flow boiling heat transfer from
porous foam structures in a channel, Appl. Therm. Eng. 70 (2014) 100–114.
vides an indication of whether or not the heat transfer enhance- [2] C. Choi, J.S. Shin, D.I. Yu, M.H. Kim, Flow boiling behaviors in hydrophilic and
ment offered by a particular oblique angle is worth the incurred hydrophobic microchannels, Exp. Thermal Fluid Sci. 35 (2011) 816–824.
pressure drop. As the range of effective heat flux is constant for a [3] C. Huh, J. Kim, M.H. Kim, Flow pattern transition instability during flow boiling
in a single microchannel, Int. J. Heat Mass Transfer 50 (2007) 1049–1060.
given mass flux, the FOM parameter in Fig. 18 is therefore plotted
[4] S.G. Kandlikar, Heat transfer mechanisms during flow boiling in
against effective heat flux as a tool for comparison for the three microchannels, J. Heat Transfer 126 (2004) 8–16.
oblique angle configurations. [5] R. Zhuan, W. Wang, Boiling heat transfer characteristics in a microchannel
From the FOM plots, it is clearly seen that at the given range of array heat sink with low mass flow rate, Appl. Therm. Eng. 21 (2013) 65–74.
[6] D. Deng, W. Wan, H. Shao, Y. Tang, J. Feng, J. Zeng, Effects of operation
effective heat flux, the 10° oblique-finned microchannels has the parameters on flow boiling characteristics of heat sink cooling systems with
highest FOM. Although its heat transfer performance is inferior reentrant porous microchannels, Energy Convers. Manage. 96 (2015) 340–351.
compared to the other two oblique angles, this is compensated [7] M.S. Sarwar, Y.H. Jeong, S.H. Chang, Subcooled flow boiling CHF enhancement
with porous surface coatings, Int. J. Heat Mass Transfer 50 (2007) 3649–3657.
by its low operating requirements, i.e. low pumping power to deli- [8] Y.K. Prajapati, M. Pathak, M.K. Khan, A comparative study of flow boiling heat
ver the necessary cooling performance. transfer in three different configurations of microchannels, Int. J. Heat Mass
An interesting observation in Fig. 18 is that the 30° and 50° obli- Transfer 85 (2015) 711–722.
[9] C.-C. Wang, W.-J. Chang, C.-H. Dai, Y.-T. Lin, K.-S. Yang, Effect of inclination on
que angles have almost comparable FOM, particularly at medium the convective boiling performance of a microchannel heat sink using HFE-
and high mass fluxes. This is largely due to the almost negligible 7100, Exp. Thermal Fluid Sci. 36 (2012) 143–148.
pressure drop of the 50° oblique fins compared to its 30° counter- [10] M. Law, P.-S. Lee, A comparative study of experimental flow boiling heat
transfer and pressure characteristics in straight- and oblique-finned
part (Fig. 16), but with considerable improvement in heat transfer microchannels, Int. J. Heat Mass Transfer 85 (2015) 797–810.
(Fig. 11). As a result, the 50° oblique-finned geometry presents an [11] Y.J. Lee, P.K. Singh, P.S. Lee, Fluid flow and heat transfer investigations on
attractive option for further parametric studies. enhanced microchannel heat sink using oblique fins with parametric study,
Int. J. Heat Mass Transfer 81 (2015) 325–336.
[12] M. Law, P.-S. Lee, K. Balasubramanian, Experimental investigation of flow
6. Conclusions boiling heat transfer in novel oblique-finned microchannels, Int. J. Heat Mass
Transfer 76 (2014) 419–431.
[13] K. Balasubramanian, P.S. Lee, L.W. Jin, S.K. Chou, C.J. Teo, S. Gao, Experimental
Flow boiling experiments on oblique-finned microchannels are investigations of flow boiling heat transfer and pressure drop in straight and
conducted using FC-72 to investigate the effects of varying oblique expanding microchannels – a comparative study, Int. J. Therm. Sci. 50 (2011)
angles on heat transfer, pressure drop and instability characteris- 2413–2421.
[14] R.D. Blevins, Applied Fluid Dynamics Handbook, Krieger Publishing Company,
tics. Three different oblique angles are considered in the present New York, 2003.
work: 10°, 30° and 50° and their flow boiling performances are [15] G.P. Celata, S.K. Saha, G. Zummo, D. Dossevi, Heat transfer characteristics of
compared with each other. High-speed visualisations are per- flow boiling in a single horizontal microchannel, Int. J. Therm. Sci. 49 (2010)
1086–1094.
formed to better characterise the boiling phenomenon in the
[16] X. Hua, G. Lin, Y. Cai, D. Wen, Experimental study of flow boiling of FC-72 in
microchannels. The following conclusions summarise the key find- parallel minichannels under sub-atmospheric pressure, Appl. Therm. Eng. 31
ings of the present study: (2011) 3839–3853.
[17] W. Qu, A. Siu-Ho, Experimental study of saturated flow boiling heat transfer in
an array of staggered micro-pin-fins, Int. J. Heat Mass Transfer 52 (2009)
1. The 10° oblique-finned microchannels provides the least effec- 1853–1863.
tive heat transfer, followed by 30° and 50°. This is likely due to [18] T. Harirchian, S.V. Garimella, Microchannel size effects on local flow boiling
the increasing number of oblique fins, as well as increasing flow heat transfer to a dielectric fluid, Int. J. Heat Mass Transfer 51 (2008) 3724–
3735.
velocity in the main channels for larger oblique angles. [19] K. Balasubramanian, P.S. Lee, C.J. Teo, S.K. Chou, Flow boiling heat transfer and
2. Oblique angles have quite a strong influence on heat transfer pressure drop in stepped fin microchannels, Int. J. Heat Mass Transfer 67
coefficient. Heat transfer coefficient increases with oblique (2013) 234–252.
[20] Y.J. Lee, P.S. Lee, S.K. Chou, Enhanced thermal transport in microchannel using
angle from 10° to 50°. This heat transfer coefficient enhance- oblique fins, J. Heat Transfer 134 (2012) 101901-1–101901-10.
ment at larger oblique angles may be due to larger number of [21] W. Qu, I. Mudawar, Flow boiling heat transfer in two-phase micro-channel
oblique fins, which is likely to produce a greater degree of heat sinks – I. Experimental investigation and assessment of correlation
methods, Int. J. Heat Mass Transfer 46 (2003) 2755–2771.
boundary layer disruptions and re-initialisations of the thin [22] J.R. Taylor, An Introduction to Error Analysis, Second ed., University Science
liquid-film. Books, California, 1997.
660 M. Law et al. / International Journal of Heat and Mass Transfer 100 (2016) 646–660

[23] S.J. Kline, F.A. McClintock, Describing uncertainties in single-sample [27] S. Krishnamurthy, Y. Peles, Flow boiling heat transfer on micro pin fins
experiments, Mech. Eng. 75 (1953) 3–8. entrenched in a microchannel, J. Heat Transfer 132 (2010) 041007-1–041007-
[24] ANSYS, ANSYS Fluent Theory Guide, Canonsburg, Pennsylvania, ANSYS Inc., 10.
2013. [28] H. Ghaedamini, P.S. Lee, C.J. Teo, Forced pulsatile flow to provoke chaotic
[25] V. Yakhot, L.M. Smith, The renormalization group, the e-expansion and advection in wavy walled microchannel heat sinks, in: IEEE Intersociety
derivation of turbulence models, J. Sci. Comput. 7 (1992) 35–61. Conference on Thermal and Thermomechanical Phenomena in Electronic
[26] Y. Fan, P.S. Lee, B.W. Chua, Investigation on the influence of edge effect on flow Systems (ITherm), Orlando, Florida, 2014, pp. 680–687.
and temperature uniformities in cylindrical oblique-finned minichannel array,
Int. J. Heat Mass Transfer 70 (2014) 651–663.

Вам также может понравиться