You are on page 1of 24

International Journal of Heat and Mass Transfer 116 (2018) 393–416

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


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

A numerical and experimental investigation of heat transfer and fluid


flow characteristics of an air-cooled oblique-finned heat sink
Bugra Kanargi, Poh Seng Lee ⇑, Christopher Yap
Department of Mechanical Engineering, National University of Singapore, 9 Engineering Drive 1, 117576, Singapore

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

Article history: The thermal and hydraulic performances of an air-cooled, planar, oblique-finned heat sink (OF HS) were
Received 26 January 2017 investigated for two oblique angles: 30° and 45°. The conjugate heat transfer between the heat sink and
Received in revised form 31 August 2017 the air flow were computed numerically in ANSYS Fluent for a range of air flow rates for the smallest peri-
Accepted 5 September 2017
odically repeating portion of the heat sink. The RNG k-e turbulence model with enhanced wall treatment
Available online 17 October 2017
was used to solve the fluid flow and heat transfer. Followed by the experimental validation, the numerical
results were scrutinized further to understand the effects of the flow field on the measured heat transfer
Keywords:
performances. Apart from boundary layer disruption, vortices generated within the secondary channels
Air cooling
Oblique fin
due to flow separation particularly improved the advection component of the convective heat transfer,
Advection resulting in a heat transfer enhancement, exceeding the pressure drop penalty. The strong flow mixing,
Secondary flow showing chaotic behavior, enabled a more uniform increase in the air temperature in the streamwise
Flow migration direction, utilizing the cooling potential of the air flow more effectively. 30° oblique-finned heat sink
Thermal-hydraulic performance factor was observed to induce higher rates of secondary flow rates and improve the heat transfer performance
more than its 45° counterpart. Due to the migration of the flow in the direction of the secondary channels,
the heat transfer enhancement was compromised at high Reynolds numbers, the pressure drop penalty
exceeded the heat transfer enhancement, causing a reduction in the thermal-hydraulic performance fac-
tor. The effects of flow migration on the flow and temperature fields were investigated with full domain
numerical simulations. Experimental investigations showed significant improvements in the junction
temperatures.
Ó 2017 Elsevier Ltd. All rights reserved.

1. Introduction faces that can lower the convective resistance. Inducing secondary
flows to reinitialize the boundary layers, or increasing turbulent
Air side heat transfer is the bottleneck of the heat transfer path intensity to enhance flow mixing via regular and/or chaotic advec-
in thermal management systems due to the inferior thermody- tion were commonly exploited flow phenomena. Achieving such
namic and transport properties of gases. Liquid cooling technolo- flow phenomena in a flow field requires non-conventional heat
gies are capable of managing higher cooling loads at smaller sink designs that introduce flow disturbances and inevitably oper-
footprints; however, they are expensive, more intricate in design, ate at higher pressure drops. A significant increase in the pressure
and require more comprehensive and frequent maintenance. drop may cause all heat transfer enhancement to be offset by an
Hence, ways are being sought to maintain the effectiveness of air increase in the pumping/fan power. However, a carefully tailored
cooling. To make better thermal management systems working flow field can improve the heat transfer at a relatively lower pres-
primarily with air; the entire heat transfer path, starting from sure drop penalty. Such attempts, reported in the literature, are
the source to the surrounding environment to which the heat is briefly summarized in the following paragraphs. Since these
sunk, should be optimized [1]. On the heat transfer path, the heat attempts have recently shifted from air cooling to single and two
sink is the most crucial component since the convective resistance phase liquid cooling, the literature review was not kept limited
makes the highest contribution to the total thermal resistance [2]. to air cooling studies.
Great effort has been made to design enhanced heat transfer sur- Developing thermal and hydraulic boundary layers yield higher
temperature and velocity gradients at the walls, providing more
rapid removal of thermal energy [3]. Zhou and Catton [4] intro-
⇑ Corresponding author.
duced pin fins of various shapes between plate fins to repeatedly
E-mail addresses: bugrakanargi@u.nus.edu (B. Kanargi), mpelps@nus.edu.sg
(P.S. Lee), mpecyap@nus.edu.sg (C. Yap).
disturb the boundary layers. They compared the overall perfor-

http://dx.doi.org/10.1016/j.ijheatmasstransfer.2017.09.013
0017-9310/Ó 2017 Elsevier Ltd. All rights reserved.
394 B. Kanargi et al. / International Journal of Heat and Mass Transfer 116 (2018) 393–416

Nomenclature

A area Q thermal power


cp specific heat Qair effective heating power of air flow
Dh hydraulic diameter Qloss the rate of the heat loss from the test section
DP pressure drop Qh heater power in numerical simulations
DTconv temperature difference that drives convective heat r radius
transfer Rcond thermal conduction resistance
deg degree Re Reynolds number
e uncertainty of a measured or a derived parameter Rec-o cross-over Reynolds number
ENu Nusselt number enhancement ratio: heat transfer SC HS straight channel heat sink
enhancement T temperature
Ef pressure drop penalty (friction factor penalty) u velocity
EWT enhanced wall treatment 
u average velocity
exp experimental V voltage
fF fanning friction factor W width, thickness
hlocal local convective heat transfer (convection) coefficient
hL average convection coefficient Greek letters
H height D difference
HS heat sink l dynamic viscosity
I current gth thermal-hydraulic performance factor
k thermal conductivity q density
k-e turbulence kinetic energy-rate of dissipation of h oblique angle
turbulence kinetic energy
L length Subscripts
LH hydrodynamic entrance length 0 reference heat sink design
LT thermal entry length a air
m_a air mass flow rate
avg average
m_ a;i air mass flow rate at the heat sink inlet ch channel
m_ a;sec secondary air mass flow rate cond conduction
num numerical
cr-sec cross section/cross-sectional
Nu0 average Nusselt number of the reference heat sink de- e enhanced heat sink design
sign hs heat sink
Nue average Nusselt number of the enhanced heat sink de- i inlet
sign
j junction
NuL average Nusselt number m mean
OF HS oblique-finned heat sink o outlet
P pressure sec secondary
Pr Prandtl number
w wall (wetted walls of a heat sink)
pfan fan power
pfin OF HS fin pitch
00
qconv convective heat flux

mances of various plate-pin fin heat sink combinations using the cooled, corrugated, straight microchannel heat sinks with triangu-
heat transfer effectiveness factor, proposed by Webb and Kim [5], lar and semi-circular cavities. Flow recirculation regions, induced
for equal pumping/fan power constraint. The square shaped pin within the cavities, were observed to reduce the thickness of the
fins showed the best heat transfer performance, however its pres- thermal boundary layer over the fin walls. The reentrant cavities
sure drop penalty offset the heat transfer enhancement compared increased the effective heat transfer area and the vortices provided
to the plate fins. The circular pin fins yielded a higher heat transfer better flow mixing via chaotic advection. A more uniform and
effectiveness, which was diminished at higher air flow rates due to lower substrate temperature was obtained at the same heat trans-
the vortices induced in the wake of the fins. The streamline-shaped fer rate. In contrast to the ribs that extend into the flow domain [7],
pin fins yielded the highest thermal effectiveness due to more the cavities [8,9] provided a heat transfer enhancement, exceeding
favorable pressure drop characteristics. The elliptic-shaped pin fins the pressure drop penalty. Compared to the conventional straight
and the NACA 0050 airfoil profile showed almost the same perfor- microchannel heat sink, the triangular reentrant cavities provided
mance, which was followed by the drop shaped pin fin. Similar a 40–60% increase in Nusselt number, while the friction factor pen-
findings on circular, elliptical and drop-shaped pin fins were also alty varied from 20% to 30% [8]. The corrugated microchannels
reported by Wang et al. [6]. with semi-circular reentrant cavities improved the fan power up
Chai et al. [7] introduced offset ribs of various shapes to the to 20% for constant thermal resistance [9].
sidewalls of a water cooled straight microchannel heat sink. The DeJong and Jacobi [10] performed an experimental study on
ribs of rectangular, backward triangular, isosceles triangular, for- louvered fin arrays that could enhance air side heat transfer by
ward triangular, and semicircular shapes were observed to repeat- boundary layer disruption and vortex shedding. An experimental
edly disturb the boundary layers at the fin walls and induce flow flow visualization was performed with dye injection method to
recirculation regions downstream. Heat transfer enhancement visualize the vortices induced between adjacent louvers. It was
was observed at higher pressure drop penalties. Xia et al. [8,9] observed that the vortices were encompassed within the louvers
investigated the thermal and hydraulic performances of water of the near wall region at low Reynolds numbers, whereas they
B. Kanargi et al. / International Journal of Heat and Mass Transfer 116 (2018) 393–416 395

started to be discharged to the main stream as Reynolds number The oblique-finned heat sink (OF HS) was created based on the
increased. In the spatially periodic region of the louvered fin arrays, conventional SC HS. The design parameters of the OF HS are pre-
the flow was mainly louver directed, except for very low Reynolds sented in Fig. 2. The oblique fin geometry is obtained by introduc-
numbers. Heat transfer enhancement, observed at high Reynolds ing secondary channels to the fins of the SC HS in a collinear
numbers, was due to boundary layer reinitialization in the periodic manner. The oblique angle h is the angle between the main and
section of the louvered fin array and unsteady flow field, generated the secondary channels. The fin pitch pfin is the distance between
by the vortex shedding in the near wall region. two consecutive fins in the streamwise direction. The width of
Lee et al. [11] studied the planar oblique-finned microchannel the oblique channels Woblique ch is the shortest distance between
heat sink. The secondary flows induced by the oblique channels the walls of an oblique channel. In this study, two OF HS configu-
reinitialized the thermal and hydraulic boundary layers at the rations were studied: h = 30° and h = 45°, where pfin was 4 mm for
leading edge of each fin section. This novel microchannel heat sink both cases. The rest of the dimensions of the OF HS are the same as
design allowed doubling the heat transfer performance, accompa- those of the SC HS. Fig. 3 shows the isometric view of the OF HS test
nied by a pressure drop penalty of approximately 50% relative to piece.
the straight microchannel heat sink of equivalent dimensions.
Fan et al. [12] studied the application of the oblique fins over a 3. Numerical analysis
cylindrical surface in minichannel size. A heat transfer enhance-
ment of 75% was obtained with respect to the cylindrical straight 3.1. Assumptions
fin minichannel heat sink at the maximum flow rate experimented
with an almost negligible pressure drop penalty. The adverse The numerical simulations of the OF HS were performed under
effects of the pressure drop observed in the planar oblique- the following assumptions:
finned microchannel [11] were overcome with the cylindrical obli-
que finned heat sink, since a complete flow recirculation zone was  Steady state fluid flow and heat transfer,
created around the cylindrical base of the heat sink.  Turbulent, incompressible air flow and heat transfer,
The advantages of the oblique-finned heat sinks were evident in  Temperature dependent air properties,
two-phase flow boiling, as well. Prajapati et al. [13] evaluated flow  Negligible viscous dissipation, radiation heat transfer, and body
boiling heat transfer characteristics of uniform cross section, forces.
diverging cross section, and segmented fin (oblique fin)
microchannels. For the entire range of operating conditions ana- 3.2. Governing equations
lyzed, segmented finned channels demonstrated the highest heat
transfer performance with negligible pressure drop penalty com- The OF HS simulations were initially attempted to be solved
pared to other configurations. The authors commented that the using the laminar flow and heat transfer equations for low Rey-
segmented fin microchannels totally eliminated the bubble clog- nolds numbers; however, a satisfactory convergence in the residu-
ging problem and no flow reversal was observed. Law et al. [14] als could not be obtained, unless turbulent flow models were
studied the effect of the oblique angle on two-phase flow and heat implemented. Shankar and Deshpande [16] summarized the fluid
transfer. A higher oblique angle promoted a higher heat transfer flow phenomena that can be observed in driven cavity flows such
performance since the number of oblique fin sections would be as complex three dimensional flow patterns, chaotic advection,
maximum, while the oblique angle did not have a profound effect flow instabilities, transition, turbulence etc. Therefore, in the sim-
on the pressure drop for two-phase flow. ulations of the OF HS; the vortices, induced within the secondary
In a recent review paper, summarizing the effects of channel channels due to flow separation, were suspected to have caused
design on the thermal and hydraulic performances of microchan- the inertial forces to exceed the viscous forces even at low Rey-
nel heat sinks, Ghani et al. [15] concluded that the oblique fins nolds numbers, resulting in an unsatisfactory convergence with
have achieved the best heat transfer performance among the the laminar flow and heat transfer equations. Among the variety
recently studied microchannel heat sink fin designs. So far, single of alternatives in ANSYS Fluent, the RNG k-e turbulence model
and two-phase liquid cooling performances of the oblique-finned [17], which is a refined version of the standard k-e model, was uti-
heat sinks were investigated for micro and minichannel sized, pla- lized. The RNG k-e turbulence model is capable of resolving rapidly
nar and cylindrical heat sinks; however, the air cooling perfor- strained flows. The RNG theory employs an analytical formulation
mance has never been tested. This study focuses particularly on for the calculation of the effective viscosity not only at high, but
the phenomenological understanding of air flow through planar, also at low Reynolds numbers. Hence, this turbulence model was
oblique-finned heat sink design for two oblique angles: 30° and appropriate by itself to resolve the entire range of air flow rates
45°. The thermal and hydraulic performance behaviors are investigated. The accuracy of the numerical results obtained with
explained by studying the flow and temperature fields from the the RNG k-e turbulence model for the solution of a similar flow
numerical results. The results and conclusions will be beneficial and heat transfer problem was validated experimentally by the
to researchers in understanding the usefulness of the flow phe- present authors in [18].
nomena for future optimization of heat sinks, and to the thermal The RNG k-e turbulence model is based on the Reynolds-
engineers in designing better performing thermal management averaged Navier-Stokes and energy equations [19]. The Reynolds-
systems for the electronics of the future. averaged continuity, momentum, and the fluid and solid domain
energy equations [17] can be respectively written as
@
2. Heat sink design ðqui Þ ¼ 0 ð1Þ
@xi
Fig. 1 shows the isometric view of the straight channel heat sink    
@ @p @ @u @u
(SC HS) test piece and its dimensions. The channel height Hch is 8 ðqui uj Þ ¼  þ l i þ j  qu0i u0j ð2Þ
mm, the channel width Wch is 4 mm, fin width Wfin is 1 mm, the @xj @xi @xj @xj @xi
total length of the heat sink LHS is 120 mm, the width of the heat sink  
@ @ l @T
WHS is 100 mm, and the length of the finned section Lch is 108 mm. ðqTuj Þ ¼  qT 0 u0j ð3Þ
Finally, the height of the base of the heat sink HHS is 17 mm. @xj @xj Pr @xj
396 B. Kanargi et al. / International Journal of Heat and Mass Transfer 116 (2018) 393–416

Fig. 1. Isometric view of straight channel heat sink test piece and its dimensions.

Main flow direction


Wch/2
pfin (fin pitch)
ϴ
Wfin
Woblique ch Secondary
flow direction
Wch/2

Fig. 2. Top view of oblique-finned heat sink and design parameters.

Fig. 3. Isometric view of oblique-finned heat sink test piece and its dimensions.
B. Kanargi et al. / International Journal of Heat and Mass Transfer 116 (2018) 393–416 397

 
@ @T The EWT option for the RNG k-e turbulence model necessitates
k ¼0 ð4Þ
@xj @xj that the first cell adjacent to a wall is placed at y+ < 1 so that the
viscous sublayer can be resolved accurately. In this study, the max-
where the turbulent stress and the turbulent heat flux are
imum y+ values of the cells adjacent to the walls were kept less
   
@ui @uj 2 @u than 0.5 to ensure that the first layer cells adjacent to the walls
qu0i u0j ¼ lt þ  qk þ lt k dij ð5Þ
@xj @xi 3 @xk are located well within the viscous sublayer.

 
lt @T 3.4. Solution domains and boundary conditions
qT 0 u0j ¼ ð6Þ
Pr t @xj
Fig. 4 shows the periodic solution domain of the OF HS. Only the
In order to calculate the turbulence kinetic energy k, and the rate of smallest periodically repeating section of the heat sink and the air
dissipation of turbulence kinetic energy e, transport equations are domain were modeled in order to reduce the computational cost.
solved. The transport equations of the RNG k-e model were modi- The inlet and outlet portions of the air flow domain were extended
fied from those of the standard k-e model and are given as by one third of the length of the heat sink to consider entrance
  effects at the heat sink air inlet and to overcome the convergence
@ @ @k
ðqkui Þ ¼ ak leff þ Gk  qe þ Sk ð7Þ issues stemming from the reversed flow at the heat sink air outlet.
@xi @xj @xj
The velocity at the air inlet was uniform and the air inlet temper-
  ature was 20 °C. Pressure outlet boundary condition was set for the
@ @ @e e e2
ðqeui Þ ¼ ae leff þ C 1e Gk  C 2e q ð8Þ air outlet with a gauge pressure of 0 Pa. All the side surfaces of the
@xi @xj @xj k k
computational domain were introduced to the solver as periodic
where Sk is a user defined source term for the turbulence kinetic boundaries. The wetted surfaces of the heat sink were coupled
energy and Gk is the turbulence kinetic energy generation due to with the corresponding surfaces of the air domain with no slip
the mean velocity gradients, which is given as boundary condition and conjugate heat transfer. A constant heat
  flux boundary condition of 1666.7 W/m2, corresponding to 20 W
@ui @uj @ui
Gk ¼ lt þ ð9Þ of heating power for the entire heat sink footprint, was set at the
@xj @xi @xj heater boundary. In the solid domain, the front and back walls of
The turbulent and effective viscosities, lt and leff, are defined as the base of the heat sink and the tips of the fins were attended
as adiabatic walls. The top wall of the air domain, and the top
lt ¼ qC l k2 =e ð10Þ and bottom walls of the extended air domains were also attended
as adiabatic walls with no slip boundary condition.
leff ¼ l þ lt ð11Þ The full domain computational model of the OF HS is presented
in Fig. 5. The inlet and outlet boundary conditions of the full
Finally C⁄2e, the modified version of the constant C2e, is domain computational model were the same as those of the peri-
C 2e ¼ C 2e þ C l g3 ð1  g=g0 Þ=ð1 þ bg3 Þ ð12Þ odic computational domain. The top, bottom, and side walls of
the air domain were introduced to the solver as adiabatic walls
where g is the ratio of the user defined turbulence kinetic energy to with no slip boundary condition. The wetted walls of the heat sink
the rate of dissipation of the turbulence kinetic energy, g = Sk/e. were coupled with the corresponding surfaces of the air domain
When the value of g is higher than g0, the flow is strongly with no slip boundary condition and conjugate heat transfer. The
strained and the value of C⁄2e is less than C2e, which would be side walls, the front and back walls, and the fin tips of the heat sink
predicted by the standard k-e turbulence model. A smaller value were attended as insulated walls. 20 W heat transfer rate was
of C⁄2e in (8) yields a lower destruction of e than normal. Since defined at the heat sink junction.
the destruction of e decreases, e actually increases, yielding a lower
k. An increase in e and a decrease in k results in a lower lt, which 3.5. Material properties
can be evaluated as (10). Finally, the analytically derived constants
of the RNG theory are Cl = 0.085, C1e = 1.42, C2e = 1.68, g0 = 4.38, The prototypes of the OF HS designs were fabricated out of alu-
b = 0.012, ak = ae = 1.393 [20]. minum alloy 6061-T6; density, thermal conductivity and specific
heat of which are q = 2700 kg/m3, k = 168 W/m °C and cp = 896 J/
3.3. Treatment of the wall kg °C [22,23], respectively. The air properties [22], incorporated
in 3rd degree polynomials, the coefficients of which are listed in
The enhanced wall treatment (EWT) was employed for the sim- Table 1, were introduced to the solver. The solver evaluated the
ulations of the turbulent flow and heat transfer. EWT is a near wall air properties for every cell at every iteration with an average per-
modeling that combines the two-layer model (the viscous sublayer cent deviation of 0.02%.
and the fully turbulent layer) with the enhanced wall functions. y+
is a significant near-wall modeling parameter [21], used in the con- 3.6. Solution grid and grid independence
text of wall treatment, which is a dimensionless wall coordinate
and defined as The computational grids of the periodic and the full domain OF
þ
pffiffiffiffiffiffiffiffiffiffiffi HS solution models, presented in Fig. 6, were created in ANSYS
y ¼ us y=m where us ¼ sw =q ð13Þ
Mesh. Structured quadrilateral cells were generated at the top wall
us is the friction velocity, y is the distance to the nearest wall, m is of the air domain with a face sizing of 0.1 mm. In the vicinity of the
the kinematic fluid viscosity, sw is the wall shear stress, and q is fin walls, where the no slip boundary condition is valid, a boundary
the fluid density. The viscous sublayer can be considered up to y+ layer mesh was generated with 14 slowly growing layers with a
= 11.225 with the least error, while the log law is applied for y+ > starting first layer thickness of 0.005 mm, which yielded a dimen-
30 [19]. The buffer zone, which lies in between these two regions, sionless wall distance y+ < 0.5. This is a requirement to resolve the
cannot be accurately resolved with neither of these models. EWT viscous boundary layer correctly with the selected turbulence
utilizes a hybrid wall function that is derived from these two mod- model and the enhanced wall treatment [17]. Next, the mesh pro-
els for the buffer layer [17]. file generated for the top surface of the air domain was swept
398 B. Kanargi et al. / International Journal of Heat and Mass Transfer 116 (2018) 393–416

Fig. 4. Periodic computational model of oblique-finned heat sink and boundaries.

heat sink
air inlet

air inlet

heat sink
air outlet

heater

air outlet

Fig. 5. Full domain computational model and boundaries.

Table 1
Thermal and transport properties of air as a function of temperature. than 0.22% for the average Nusselt number and 0.3% for the friction
U A1 A2 A3 factor. The mesh count was approximately 9 million for the peri-
odic OF HS solution domains.
q [kg/m3] 3.2894 0.010286 0.000010653
k [W/m K] 0.0007626595 0.0000965192 0.0000000371 Utilizing the same quality grid for the simulations of the full
cp [J/kg K] 1033.78 0.217226807 0.000421911 domain computational models as that of the periodic ones would
l [kg/m s] 0.0000012568043 0.00000006698901 0.0000000000313 have resulted in a mesh count of nearly 200 million, making it
U(T) = A1 + A2T + A3T2 where T is in Kelvin, and 273.15 K < T < 373.15 K impossible to handle the computations using our existing compu-
tational facilities. Therefore, a reverse grid independence analysis
was performed to coarsen the independent grid of the periodic
computational models. The quality of the mesh were reduced at
towards the bottom surface of the air domain with fine boundary locations, where strong gradients in temperature and velocity
layer cells in the vicinity of the top and bottom walls to satisfy fields did not exist. The number of sweep layers utilized for the
the y+ criteria. For the solid domain, a tetrahedral mesh was used channel height were reduced. The heat sink mesh sizing was set
to solve the conduction heat transfer within the heat sink. A non- to increase from 0.4 mm at the wetted surface to 3 mm at the hea-
conformal mesh was generated at the solid-fluid interface not to ter boundary. Nevertheless, the first layer thicknesses of the cells at
let the mesh count in the solid domain to increase abruptly due the walls of no slip boundary condition were kept the same since
to the fine boundary layer cells of the air domain. Instead, the aver- the y+ criteria had still to be satisfied. As a result, the mesh count
age mesh size in the solid domain was set to increase from 0.25 for a single channel was dropped from 9 million to approximately
mm at the wetted surfaces to 1 mm at the bottom of the heat sink. 2.1 million with a maximum deviation of 2.7% in Nusselt number
The aforementioned sizing parameters were determined after a and the friction factor compared to the periodic computational
comprehensive grid independence study. Compared to a higher model simulation results. The grid of the full domain solution
resolution grid, the final grid utilized did not yield a change more model comprised approximately 44 million elements.
B. Kanargi et al. / International Journal of Heat and Mass Transfer 116 (2018) 393–416 399

where n is the number of cells at the respective boundary. Finally,


top view-periodic model
the mass flow rate of air is calculated at the heat sink inlet as
X
n
_a¼
m ui :~
qi~ Ai ð16Þ
i¼1

The average air velocity through the finned section of the heat sink
is
¼m
u _ a =ðq@T m;a Aflow crsec Þ ð17Þ

where the mean air temperature and the flow cross sectional area
through the finned section of the heat sink are
T m;a ¼ ðT a;i þ T a;o Þ=2 ð18Þ
side view-
side view-
full
periodic Aflow cr-sec ¼ W ch  Hch ðperiodic numerical modelÞ ð19Þ
domain
model
model
Aflow cr-sec ¼ W ch  Hch
 #of channels ðfull domain & experimentalÞ ð20Þ

Thus, the average Reynolds number is given in (21), where the


hydraulic diameter is defined both for the SC HS and the OF HS as
in (22).
q@T m;a u Dh
Re ¼ ð21Þ
l@T m;a

4W ch Hch
top view-full domain model Dh ¼ ð22Þ
2ðW ch þ Hch Þ
Fig. 6. Top and side views of computational grids of periodic and full domain The average convection coefficient is
computational models of oblique-finned heat sink.
Qh
hL ¼ ð23Þ
Aw DT conv
where the temperature difference that drives the convective heat
3.7. Solver settings
transfer is calculated as
After importing the grid into ANSYS Fluent, the steady state sol- DT conv ¼ T w  T m;a ð24Þ
ver with double precision was selected. The boundary conditions
The fan power is given in (25) where the pressure drop through the
were introduced. The SIMPLEC algorithm was used for pressure-
heat sink is calculated by taking the difference between the static
velocity coupling with second order upwind discretization scheme
pressures at heat sink air inlet and outlet.
for momentum, turbulent kinetic energy, and turbulent dissipation
rate. The absolute convergence criteria was 1010 for energy and _ a DP
m
106 for the remaining parameters. Besides the residuals, the
pfan ¼ ð25Þ
q@T m;a
mass-weighted average values of the air inlet pressure and air out-
let temperature were ensured to reach a steady state value before DP ¼ Pa;i  Pa;o ð26Þ
the calculations were terminated. Running on Intel Xeon X5690
processor at 3.47 GHz with 8 cores, the simulations of the periodic Finally, the performance parameters, the average Nusselt number
OF HS models were completed within 24–72 h for each case. The and the Fanning friction factor, are calculated as
simulations converged relatively faster at higher air flow rates.
hL Dh
Each full domain solution took up 72 h to fully converge. NuL ¼ ð27Þ
k@T m;a
3.8. Numerical data reduction
DPDh
fF ¼ ð28Þ
 2 Lhs
2q@T m;a u
The air inlet and outlet temperatures and pressures (Ta,i, Ta,o,
Pa,i, and Pa,o) were calculated as mass-weighted, while the average
temperatures of the heat sink wetted walls Tw and junction Tj were
4. Experimental analysis
calculated as area-weighted averages of all the respective values of
the cells at the respective boundaries. The mass- and area-
The thermal and hydraulic performances of the SC HS and the
weighted average of a parameter, /, is calculated in ANSYS Fluent
OF HS were measured using the compact wind tunnel experimen-
respectively as
tal setup.
Pn
ui :~
/i qi j~ Ai j
/mass-weighted ¼ Pi¼1 ð14Þ
n
~ ~ 4.1. Experimental setup
i¼1 i i i j
q j u : A

Fig. 7a shows the schematic view of the compact wind tunnel


1X n
/area-weighted ¼ / jAi j ð15Þ experimental setup. EBM-Papst rg148 centrifugal fan was installed
A i¼1 i
at the inlet of the wind tunnel. A variac transformer served the
400 B. Kanargi et al. / International Journal of Heat and Mass Transfer 116 (2018) 393–416

(a) Reducer & expander


Long circular duct Test section
L = 1 m, Dh = 44 mm

Air
Flow
discharge Air outlet Ta,o Air inlet Ta,i Air blower
direction

(b) low high (c)


ΔP Flow cross sectional area at heat sink outlet

Test section Whs = 100 mm


Hch
8 mm
Heat sink Tair outlet

(d)

Air discharge
Ths 8 7 6 5 4 3 2 1

Velocity probe
Heater
Fig. 7. Schematic view of desktop wind tunnel experimental setup.

function of powering the fan and adjusting its rotational speed. An 4.2. Data acquisition and reduction
expander section connected the fan outlet to the wind tunnel inlet.
The inlet section of the wind tunnel was extended to ensure flow Data was acquired using a National Instrument cDAQ 9178 USB
uniformity at the heat sink inlet. The air inlet temperature was chassis, equipped with appropriate thermocouple, voltage, and
measured at the beginning of the extended section in order not current reading modules. A LabView program was created to auto-
to disturb the air flow just before the heat sink inlet. The test sec- matically acquire, pre-process, and record the data. The tempera-
tion was made out of Teflon, which is a polymer that can withstand ture measurements were monitored real time while a steady
high temperatures of the heater. It held the heater and the test state was sought, in which the temperature variations were within
piece firmly in place and provided thermal insulation against the ±0.1 °C over five-minute periods. All the readings of a given vari-
extensive heat losses from the test section. A thin layer of high able, recorded for five minutes at every second, were time-
thermal conductivity thermal interface material, OMEGATHERM averaged to obtain a single data.
201, was applied on the entire surface of the heater before bringing Although, the experimental data presented in this study were
it in contact with the heat sink junction. A bottom cover made out merely of steady state, the transient heat transfer characteristics
of Teflon was screwed to the heat sink housing to maintain tight were observed while reaching a steady state. The observations
contact between the plate heater and the heat sink. The plate hea- aligned well with those of Prajapati et al. [24]. The air flow rate sig-
ter was powered by the Sorensen XTR-150–5.6 power supply, nificantly affected the time required to reach a steady state. As the
which allowed fine adjustments of the supply voltage and current. air flow rate was increased, the temperatures reached steady state
The temperature of the test piece was measured using eight T-type faster. For the lowest air flow rate studied, it took longer than an
thermocouples, placed into the holes within the test pieces, opened hour for the temperature deviations to drop below ±0.1 °C over
in the streamwise direction at mid-height of the heat sink base five-minute periods, while it took less than 10 min for the highest
(Fig. 7b). The pressure drop of the air flow through the heat sink air flow rate. No observable change in the steady state times
was measured using differential pressure transducers. Considering occurred with a change in the heating power. The OF HS experi-
the possibility of having a non-uniform air outlet temperature, ments reached steady state operating conditions faster than the
nine T-type thermocouples, tips of which had been spot welded, SC HS. Finally, it was observed that cooling the heat sink helped
were placed at the air outlet (Fig. 7a and c). The last section of to reach steady state faster than heating the heat sink. Therefore,
the compact wind tunnel was a long circular duct, which was con- the experimental sets were initialized with the lowest Reynolds
nected to the test section outlet with a reducer. This long circular numbers.
tube helped to obtain a fully developed velocity profile at the air The average heat sink temperature Ths, the heat sink air inlet tem-
discharge so that the central velocity could be accurately measured perature Ta,i, and the heat sink air outlet temperature Ta,o were cal-
using a probe type velocimeter (Fig. 7d). culated by averaging the time-averaged readings of the respective
B. Kanargi et al. / International Journal of Heat and Mass Transfer 116 (2018) 393–416 401

thermocouples as in (29), where n represents the number of thermo- Since the heat sink temperature Ths was measured at the mid-
couples utilized for each measurement. Pressure drop and air outlet height of the base of the heat sink, the average wetted wall tem-
velocity measurements were performed using a single sensor at a perature Tw vary from Ths as
time, and the voltage and current supplied to the heater plate were
T w ¼ T hs  Rcond;hs Q air =2 ð36Þ
recorded according to the digital output of the power supply.
! where the 1D thermal conduction resistance of the heat sink base is
X
n
Ti ¼ T iðjÞ =n ð29Þ calculated as
j¼1
Hhs 0:017 m
Rcond;hs ¼ ¼
To be able to impose the thermal boundary condition as accurately khs  Ahs;base 168W=m K ð0:1  0:2Þm2
as possible, the heat loss from the test section to the surroundings K
had to be immediately calculated and compensated during the ¼ 8:4325  103 ð37Þ
W
experiments. Therefore, the exact amount of heat loss from the test
During the experiments, the air inlet temperature varied by up to
section at various testing conditions had to be known. Since the
±1.5 °C depending on the environment temperature and the fan
temperature of the surroundings remained constant, the heat loss
speed. As the fan speed increased, the air inlet temperature
was a function of the temperature of the test section, therefore
increased since the heat dissipated by the fan motor increased.
the average heat sink temperature. For that reason, the heat loss
For the numerical and experimental junction temperatures to be
of the test section was characterized by measuring the heat sink
comparable, the average heat sink temperature Ths was corrected
temperature at various heating powers with the absence of air flow.
as Ths,corrected by offsetting Ths according to the deviation of the air
A similar heat loss characterization method was utilized by Jeng
inlet temperature Ta,i from 20 °C. Ths,corrected was evaluated accord-
et al. [25]. The relation between the heat loss and the average heat
ing to Eq. (38) regardless of whether Ta,i was smaller or higher than
sink temperature was formulated as a linear equation as
20 °C.
Q loss ¼ 0:3461T HS av g  6:7736 ð30Þ
T hs;corrected ¼ T hs  ðT a;i  20 CÞ ð38Þ
Once the above correlation is obtained, the effective heating power
Next, the junction temperature was evaluated from Ths,corrected as
of the air flow is calculated as
T j ¼ T hs;corrected þ Rcond;hs Q air =2 ð39Þ
Q air ¼ V heater Iheater  ð0:3461T HS av g  6:7736Þ ð31Þ
Finally; fan power, Nusselt number, and the Fanning friction factor
where V and I are the voltage and the current supplied to the heater.
were evaluated as in (25), (27), and (28), respectively.
The expression in (31) was calculated by the LabView software in
real time during the experiments and its value was set to
4.3. Error analysis
20 W ± 0.1 W or 60 W ± 0.5 W during the steady state.
The mass flow rate of air through the wind tunnel is calculated
In this study, a standard error analysis [27] was utilized to
by using the air velocity measurement at the discharge of the cir-
quantify the uncertainties of the derived parameters. The uncer-
cular duct. The calculation of the Reynolds number of flow through
tainties of the sensors and the power supply, documented by the
the circular duct yielded that the flow through the duct was in the
manufacturers, are listed in Table 2.
transition region for the low air flow rates and in the turbulent
Assuming that a, b, c . . . would be the sensor measurements or
region for medium to high air flow rates investigated. Therefore,
some other derived parameter with the given or calculated uncer-
the velocity profile at the outlet of the circular duct was predicted
tainties of ea, eb, ec, . . ., the uncertainty of a derived parameter, f,
with the power law velocity profile for turbulent flow [26] given as
given in the functional form of f = f(a, b, c, . . .), would be
uðrÞ ¼ umax ð1  r=RÞ1=n ð32Þ sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 2  2  2
@f @f @f
where umax is the maximum velocity at the center of the outlet of df ¼ ea þ eb þ ec þ . . . ð40Þ
@a @b @c
the duct, measured with the velocity probe, and u(r) is the velocity
profile with respect to the radius. After integrating (32) over area, The uncertainties of the reported Nusselt numbers vary from 2 to
the relation between the average and maximum velocities at the 9%. The largest contributors to the uncertainties of the Nusselt num-
circular duct outlet, uduct and umax, becomes: bers are the uncertainties of the air inlet temperature (±0.3 °C), the
average heat sink temperature (±0.18 °C) and the air outlet temper-
 duct
u 2n2
¼ ð33Þ ature (±0.17 °C) measurements. The uncertainties in the Nusselt
umax ðn þ 1Þð2n þ 1Þ numbers increase with increasing air flow rate because the convec-
The value of n is taken as 7, which satisfies a wide range of flow con- tive heat transfer coefficient, which is an independent parameter of
ditions. The air mass flow rate through the wind tunnel is then cal- the Nusselt number, is a function of DTconv, defined as in (24). As the
culated as air flow rate is increased, DTconv decreases, while its uncertainty,
relative to its absolute value, increases.
_ a ¼ q@T a;o u
m  duct Aduct ð34Þ The most significant source of uncertainty for the reported fric-
tion coefficients are the uncertainties of the pressure transducers,
where Aduct = p D2/4, and D is 44 mm.
The average air velocity through the finned section of the heat
sink is evaluated as in (17). The definition of the experimental Table 2
average Reynolds number is the same as (21). Sensor uncertainties.
The average convection coefficient is Sensor Uncertainty

Q air T-type thermocouples ±0.5 °C


hL ¼ ð35Þ
Aw DT conv Differential pressure transducer (Omega PX653-0.25D5V) ±0.15 Pa
Differential pressure transducer (Unik 5000) ±2.8 Pa
where the temperature difference that drives the convective heat Velocity probe (Velocicalc 9565) ±1.5%
Power supply voltage and current (Sorensen XTR 150-5.6) ±0.5%
transfer is given in (24).
402 B. Kanargi et al. / International Journal of Heat and Mass Transfer 116 (2018) 393–416

which are constant values for each transducer. As the air flow rate for the hydrodynamic entrance region x+ = Lch/Dh.Re, by Shah and
is increased, the relative values of the uncertainties decrease com- London [3]. The friction factors measured with the present exper-
pared to the absolute values of the measurements. In order to keep imental facility agreed with those predicted by Curr et al. [28] very
the uncertainty of the friction coefficients within reasonable limits, well for the entire range of air flow rates.
two pressure transducers were utilized: OMEGA PX653 for the low The thermal entrance length LT for laminar flow through ducts
air flow rates and UNIK 5000 for the high air flow rates (Table 2). is approximated as LT = Re.Dh.Pr/20. Since the Prandtl number for
Thus, the uncertainties in the experimental friction factors could air flow is approximately 0.7, the thermal entrance length is
be maintained within 1.5–12% for the measurements with the shorter than the hydrodynamic entrance length. For the combined
OMEGA PX653, and 2.5–20% for the measurements with the UNIK entrance length, NuL was given by Sieder and Tate [29] as
5000 pressure transducer.  0:33  0:14
RePrD lf
NuL ¼ 1:86 ð41Þ
Lch lw
5. Results and discussion
Eq. (41) is valid for simultaneously developing flows when
Numerical simulations of the OF HS designs were performed 0.60  Pr  5 and 0.0044  lf/lw  9.75. For Reynolds numbers
using ANSYS Fluent 16.2. The heater power was taken as 20 W above 2300, Dittus-Boelter equation can be used to determine the
for the entire heat sink footprint. A wide range of air flow rates NuL [30].
were simulated with the periodic solution domains, while only
three air flow rates were simulated with the full domain computa- NuL ¼ 0:023Re0:8 Pr0:4 ð42Þ
tional models due to extended run time. The numerical simulations Finally, to ensure that the data obtained with the compact wind
were experimentally validated on the desktop wind tunnel exper- tunnel experimental setup were free from random errors, the entire
imental setup. 20 W and 60 W heater powers were experimented. sets of experiments were repeated. The performance parameters
The heat loss from the test section was calculated and immediately obtained from both pairs of data sets are presented in Fig. 9. It
compensated during the experiments to match the experimental was observed that the deviations between each pair of data sets
and the numerical boundary conditions as good as possible. For were within acceptable limits, within the experimental
performance benchmarking purposes, the conventional SC HS uncertainties.
design was also tested for both heater powers for the entire range
of air flow rates.
5.2. Validation of the numerical results

5.1. Validation of the experimental setup and experimental Fig. 10 shows that the numerical junction temperatures of the
repeatability periodic computational domains were observed to match the
experimental measurements with reasonable accuracy. However,
The thermal and hydraulic performances of the SC HS were val- NuL and fF were overpredicted by the periodic computational
idated against the well known correlations in the literature so as to domains as can be seen in Fig. 11.
ensure the reliability of the experimental setup. The comparison of The numerical methodology utilized for the simulations of the
the experimental SC HS data and correlations is presented in Fig. 8. OF HS was already validated experimentally by the present authors
The theoretical hydrodynamic entrance length approximation for the cross-connected alternating converging-diverging channel
for laminar flow, LH = Re.Dh/20, given by Kays et al. [21] estimates heat sink with the use of the same experimental facility [18], in
that the air flow through the SC HS becomes hydrodynamically which case no systematic deviation between the numerical and
fully developed after half of the length of the heat sink for the experimental performance parameters was observed. Consid-
Re = 200. For Re = 400 onwards, the flow through the entire length ering the fact that the experimental facility is able to produce reli-
of the SC HS remains in the hydrodynamically developing regime. able data (Figs. 8 and 9), the suspicion was given to the validity of
The graphical data, provided by Curr et al. [28] for developing flow the fully periodic flow assumption rather than to the accuracy of
in square ducts, was converted to f.Re multiplications as a function the numerical and the experimental data produced.
of x+, the dimensionless axial distance in the direction of the flow Mou et al. [31] and Fan et al. [32] performed numerical investi-
gations on liquid-cooled OF HS in micro and minichannel size
within wall bounded flow domains. The authors reported that
30 1
the periodicity of the secondary flows was not maintained in the
Nu_L (exp) current study
Nu_L Sieder & Tate (41) vicinity of the edges due to the migration of the flow in the direc-
25 Nu_L Dittus & Boelter (42) tion of the oblique channels. In contrast, the secondary channels in
f_F Curr et al. [28] the middle region were reported to have relatively uniform distri-
20 f_F (exp) current study 0.1 bution of the secondary flow rates, preserving the periodic flow
conditions. Therefore, in the current investigation, a deviation
was already anticipated between the experimental and numerical
NuL

fF

15
data obtained from the single channel, periodic computational
domains of the OF HS; however, the extent of the migration of
10 0.01 the air flow and its adverse effects on the overall heat transfer
and the pressure drop performances were not yet clear.
5 Fig. 11 shows that the differences between the experimental
and numerical NuL and fF are more pronounced for the 30° OF
HS, compared to the 45° OF HS. It can be commented that a lower
0 0.001
oblique angle, which makes a secondary channel more inline with
0 1000 2000 3000 4000 5000 6000 7000
the main channel, is capable of inducing higher rates of secondary
Re
flows, hence escalating the flow migration in the direction of the
Fig. 8. Comparison of experimental Nusselt numbers and friction factors of straight oblique channels. A stronger flow migration from one side of the
channel heat sink with well-known correlations. heat sink (draining edge) to the other side (filling edge) will disturb
B. Kanargi et al. / International Journal of Heat and Mass Transfer 116 (2018) 393–416 403

35 0.7
30deg OF HS Set 1 (a) (b) 30deg OF HS Set 1
30deg OF HS Set 2 30deg OF HS Set 2
30 0.6
45deg OF HS Set 1 45deg OF HS Set 1
45deg OF HS Set 2
45deg OF HS Set 2
25 0.5 SC HS Set 1
SC HS Set 1 SC HS Set 2
SC HS Set 2
20 0.4
NuL

fF
15 0.3

10 0.2

5 0.1

0 0.0
0 1000 2000 3000 4000 0 1000 2000 3000 4000
Re Re

Fig. 9. Experimental repeatability of straight channel, 30° and 45° oblique-finned heat sink data.

70 To support these claims on the flow migration and the period-


45deg OF HS (num, periodic) icity of the secondary flows through the OF HS, the need for a full
65
45deg OF HS (exp) domain solution arouse. Due to the high computational load and
60 30deg OF HS (num, periodic) long run time, each OF HS configuration was simulated in full
30deg OF HS (exp) domain at three Reynolds numbers. The performance parameters
55 obtained from the full domain numerical simulations are pre-
50 sented in Fig. 11 along with those obtained from the periodic com-
Tj [°C]

putational domains and the experiments. It was observed that the


45 full domain simulations, especially of the 30° OF HS, could success-
40 fully capture the experimental Nusselt numbers and friction fac-
tors with reasonable accuracy both in trend and magnitude.
35
30
5.3. Distribution of air flow rate through the main and secondary
25 channels
20
0 1000 2000 3000 4000 The full domain simulation results were investigated to quanti-
Re tatively identify the severity of the flow migration on the unifor-
mity of the air flow rates through the main channels and the
Fig. 10. Numerical & experimental junction temperatures of 30° and 45° oblique- secondary channels in the spanwise direction of the OF HS.
finned heat sink. The air mass flow rates at the heat sink inlet and heat sink out-
let of each of the main channels were calculated as in (16) at the
respective boundaries of the full domain OF HS simulation model,
the uniformity of the air flow rate distribution in the lateral direc- shown in Fig. 5.
tion and reduce the number of channels that follow the periodic Fig. 12 shows the distribution of the air flow rates at the heat
flow conditions. sink inlet and outlet at low, moderate and high Reynolds numbers.

50 0.18
30deg OF HS (num, periodic) 30deg OF HS (num, periodic)
45 30deg OF HS (num, full domain) (a) 0.16 (b) 30deg OF HS (num, full domain)
30deg OF HS (exp) 30deg OF HS (exp)
40 45deg OF HS (num, periodic)
0.14 45deg OF HS (num, periodic)
45deg OF HS (num, full domain) 45deg OF HS (num, full domain)
35 45deg OF HS (exp) 45deg OF HS (exp)
0.12
30
0.1
Nu L

fF

25
0.08
20
0.06
15
10 0.04

5 0.02

0 0
0 1000 2000 3000 4000 0 1000 2000 3000 4000
Re Re

Fig. 11. Numerical and experimental (a) Nusselt numbers and (b) Fanning friction factors of 30° and 45° OF HS.
404 B. Kanargi et al. / International Journal of Heat and Mass Transfer 116 (2018) 393–416

While the 45° OF HS did not show a significantly non-uniform inlet mid-channel height at identical air inlet velocities, corresponding
and outlet mass flow rate distributions in the spanwise direction, to Re = 989.
the 30° OF HS experienced approximately 14% decrease and 40% Fig. 15 shows that the 45° OF HS has a higher flow velocity in
increase in the air flow rates at the heat sink inlet and outlet, the main channel compared to the 30° OF HS, since 30° oblique
respectively, for Re = 3117. channels carry higher rates of secondary flows. As can be observed
m_ a;sec is the cumulative secondary mass flow rate of air, migrat- from the zoom-in version of Fig. 15a, a large portion of the sec-
ing from one main channel to the next through the oblique chan- ondary channels of the 45° OF HS is occupied by a flow field at a
nels. m _ a;sec for any fin row was calculated through the plane, lower velocity. These lower velocity regions are the vortices,
passing through the secondary channels of that particular fin induced by flow separation, the details of which will be studied
row, extending from the heat sink inlet to the heat sink outlet, as in the next section. The vortices within the secondary channels
shown in Fig. 13. Fig. 14 shows the distribution of m _ a;sec in the of the 30° OF HS are smaller in size and stretched, therefore they
spanwise direction of the (a) 45° and (b) 30° OF HS at various Rey- do not block the secondary channels and allow higher rates of sec-
nolds numbers. It was observed that certain portion of the chan- ondary flows to be induced.
nels in the middle section of the OF HS sustained almost Fig. 16 shows that the 30° OF HS is capable of maintaining lower
identical cumulative secondary flow rates, unaware of the exis- fin temperatures compared to the 45° OF HS at equal air flow rates
tence of the edges of the heat sink. The uniformity of the secondary and heater powers. Based on the observations of the velocity and
flow rates was much worse near the edges, similar to the findings temperature fields, it can be commented that the heat transfer
of DeJong and Jacobi [10] on the wall bounded flows through lou- enhancement depends strongly on the secondary flow rates. The
vered fin arrays. The 45° OF HS was always capable of maintaining air temperature within the secondary channels of the 45° OF HS
a more uniform secondary flow rate distribution compared to the (Fig. 16a) is higher than that of the 30° OF HS (Fig. 16b) due to
30° OF HS in the spanwise direction since the cumulative sec- the fact that the advection performance of the former is lower than
ondary flow rates of the former were much lower than those of the latter due to insufficient amount of secondary flow rates. A
the latter. As the Reynolds number increased, the severity of the higher air flow temperature requires a higher fin temperature for
flow migration increased significantly for the 30° OF HS; therefore, the same heat transfer rate to be sustained. Moreover, the 30° OF
the number of channels, through which the uniformity of the sec- HS enables a more uniform air temperature build up in the stream-
ondary flow rates was maintained, decreased. wise direction compared to the 45° OF HS by inducing stronger
The results obtained from the periodic computational models flow mixing, which is again a result of the higher rates of secondary
are free from the edge effects, which reduce the cumulative sec- flows.
ondary flow rates in the vicinity of the edges. Therefore, the peri- Finally, Fig. 17 shows that similar to the heat transfer enhance-
odic computational models cannot predict the adverse effects of ment, the pressure drop penalty is also proportional to the sec-
flow migration on Nusselt numbers, and yield higher Nusselt num- ondary flow rates. For identical outlet pressures, the inlet
bers compared to the experimental data and the full domain pressure of the 30° OF HS (Fig. 17b) is 26.3 Pa, whereas it is
numerical simulation results. 13.3 Pa for the 45° OF HS (Fig. 17a). Sustaining flow momentum
in a direction other than the natural direction of the flow increases
5.4. Pressure, velocity and temperature fields in the periodic region of the pressure drop penalty, along with the heat transfer enhance-
the oblique-finned heat sink ment. As could be observed from the zoom-in images in Fig. 17,
there exist adverse pressure gradients around the edges of the obli-
To understand the thermal and hydraulic performance charac- que fins, where the boundary layer disruption and flow separation
teristics of air flow through the OF HS, the flow and temperature occurs.
fields were studied in detail. Due to the superior grid quality, the This section reported merely the preliminary observations of
single channel, periodic computational models were studied to the flow and temperature fields of the planar, air-cooled OF HS in
obtain better resolution. The velocity, temperature, and pressure the periodic region. More detailed explanations will be given in
contours of the 45° and 30° OF HS designs, are presented in the following section to fully understand the coupled characteris-
Figs. 15–17, respectively. The contour plots were taken at the tics of the air flow and heat transfer.

0.0006 0.0006
Re=3117 inlet Re=3117 outlet Re=3117 inlet Re=3117 outlet
(a) Re=1865 inlet Re=1865 outlet (b) Re=1865 inlet Re=1865 outlet
0.0005 0.0005 Re=1239 inlet Re=1239 outlet
Re=1239 inlet Re=1239 outlet
air mass flow rate [kg/s]
air mass flow rate [kg/s]

0.0004 θ=45° 0.0004 θ=30°

0.0003 0.0003

0.0002 0.0002

0.0001 0.0001

0 0
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20
Channel number Channel number

Fig. 12. Air flow rate distributions at heat sink inlet and outlet of (a) 30° and (b) 45° OF HS.
B. Kanargi et al. / International Journal of Heat and Mass Transfer 116 (2018) 393–416 405


#4
#5
#6
#7
#8

Fig. 13. Secondary flow planes.

0.00022 0.001
Re=3117
θ=45º (a) Re=3117
θ=30º (b)
cumulative secondary flow rate [kg/s]

cumulative secondary flow rate [kg/s]


0.0009
Re=1865 Re=1865
0.00018
Re=1239 0.0008 Re=1239
0.0007
0.00014
0.0006
0.0001 0.0005
0.0004
0.00006
0.0003
0.0002
0.00002
0.0001
-0.00002 0
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20
Fin rows Fin rows

Fig. 14. Cumulative secondary flow rates w.r.t. the fin rows.

(a)
inlet

(b)

(a)

(b)

Fig. 15. Velocity contours of flow through (a) 45° and (b) 30° oblique-finned heat sink.

(a)
inlet

(b)

(a)

(b)

Fig. 16. Temperature contours of flow and fins through (a) 45° and (b) 30° oblique-finned heat sink.
406 B. Kanargi et al. / International Journal of Heat and Mass Transfer 116 (2018) 393–416

(a)
inlet
(b)

(a)

(b)

Fig. 17. Pressure contours of flow through (a) 45° and (b) 30° oblique-finned heat sink.

5.5. Thermal and hydraulic performances number regime. Beyond that point, the pressure drop penalty Ef
continued to increase consistently, maintaining its superior values
In order to evaluate the heat transfer enhancement and the over ENu until the highest air flow rate investigated. Meanwhile,
pressure drop penalty of the 30° and 45° OF HS designs relative beyond the second ENu-Ef cross-over, a decrease in the ENu was
to the SC HS, the Nusselt number enhancement ratio ENu and the observed for the 30° and 45° OF HS designs.
pressure drop penalty Ef were calculated as in (43) and (44), From Fig. 18, the Rec-o were observed to lie between 239 and
respectively. The single channel, periodic computational domain 489 for the 45° OF HS, and 489 and 739 for the 30° OF HS. In the
simulation results of the OF HS and the experimental SC HS data previous section, the secondary flow rates were suspected to be a
were used to evaluate ENu and Ef. The results are presented in crucial parameter, which control the heat transfer and the pressure
Fig. 18. It was observed for both OF HS designs that the pressure drop performances. Therefore, with an intention to validate the
drop penalty Ef exceeded the heat transfer enhancement ENu at observed Rec-o intervals quantitatively, a secondary flow ratio
very low Reynolds numbers; however, as the air flow rate was was defined as the ratio of the cumulative secondary flow rate
increased, ENu was observed to exceed Ef. Although the exact val- m _ a;i as
_ a;sec to the inlet air flow rate m
ues of the Reynolds numbers at which ENu-Ef cross-over occurred
were not known for neither of the OF HS designs; the intervals in
_ a;sec
m
secondary flow ratio ¼ ð45Þ
which those particular Reynolds numbers lied were known. This m_ a;i
particular Reynolds numbers, which caused the ENu-Ef cross-over,
The single channel, periodic computational domain simulation
has been named as the cross-over Reynolds number Rec-o.
 results were used for the calculation of (45). The secondary flow
NuLOF HS  ratios of the 30° and 45° OF HS designs were presented in the right
ENu ¼ ð43Þ
NuLSC HS Re side vertical axis of Fig. 18 as percentage, along with ENu and Ef
curves, for ease of observation. The secondary flow ratios showed

f F OF HS  an increasing-decreasing-increasing trend for both 30° and 45° OF
Ef ¼  ð44Þ HS designs. This trend is much clearer for the 30° OF HS since it is
f F SC HS 
Re capable of inducing higher rates of secondary flows, for which the
secondary flow ratio is above 100% for the entire range of inlet air
For a wide range of air flow rates, ENu maintained its superior posi-
flow rates, while the 45° OF HS is not capable of inducing cumula-
tion over Ef until another cross-over occurred at the high Reynolds
tive secondary flow rates higher than the inlet air flow rates. It was
observed that the ENu-Ef cross-over occurred within the Reynolds
5.0 450 number intervals in which the secondary flow ratios started show-
30deg OF HS E_Nu ing a decreasing trend. The decrease in the percent secondary flow
30deg OF HS E_f
4.5 45deg OF HS E_Nu 400 ratios at the aforementioned Reynolds numbers can also be read
45deg OF HS E_f
30deg OF HS sec flow ratio from Table 3.
4.0 350
45deg OF HS sec flow ratio Some flow phenomena, triggered at Rec-o, cause a dramatic
increase in the heat transfer performance, while a decrease in the
secondary flow ratio (%)

3.5
300 secondary flow rates is observed, accompanied by a negligible
3.0 increase in the pressure drop penalty. The combined effect of these
250
Ef, ENu

phenomena emerges as a heat transfer enhancement, exceeding


2.5
the pressure drop penalty of the OF HS relative to the SC HS. To
200
2.0
150
1.5
Table 3
1.0 100
Secondary flow ratios of 45° and 30° OF HS w.r.t. Reynolds number.

0.5 50 Secondary flow ratio (%)


Re 45° OF HS 30° OF HS
0.0 0
177 70 213
114
177
239
489
739
989
1239
1865
2490
3116
3741

239 71 262
489 55 280
Re 739 38 245
989 31 233
Fig. 18. Heat transfer enhancement and pressure drop penalty of 30° and 45° 1239 27 242
oblique-finned heat sink w.r.t. Reynolds number.
B. Kanargi et al. / International Journal of Heat and Mass Transfer 116 (2018) 393–416 407

understand the reason for such thermal and hydraulic performance the vortex within the secondary channel becomes larger and stron-
characteristics, the flow field around the 30° and 45° OF HS designs ger, it reduces the effective oblique channel width, through which
were studied in detail at Reynolds numbers just below and above the secondary flow can pass. As a result, the quantity of the portion
the respective Rec-o of each design. of the flow, diverted from its natural flowing direction, is reduced.
Thus, the pressure drop penalty becomes relatively lower than the
heat transfer enhancement, which is a result of improved advec-
5.6. Analysis of the secondary flow field
tion due to the induced vortex.
Similar secondary flow behavior was observed for the 30° OF HS
The secondary flow field analysis was performed around the
with slight differences. As can be observed from Fig. 20a, a lower
oblique fins #13, located at the halfway in the streamwise direc-
oblique angle of 30° already causes a flow separation and a very
tion of each OF HS configuration. Therefore, the flow fields, pre-
weak vortex below Rec-o, similar to its 45° counterpart. This vortex
sented in Figs. 19 and 20, are free from entrance effects.
comprises air streams from the downstream of the fin even below
Fig. 19a shows the velocity vectors around a 45° oblique fin at
Rec-o; however, it is not strong and large enough to reduce the sec-
the mid-channel height at Re = 239 and Re = 489, just below and
ondary flow ratio. Therefore, Rec-o for the 30° OF HS cannot be
above Rec-o. At Re = 239, the velocity field shows a very weak vor-
lower than Re = 489. It can be observed that the vortex becomes
tex generated within the secondary channel due to the separation
very strong above Re = 489 and below Re = 739, where it becomes
of the flow over the separation edge. At Re = 489, however, the vor-
large enough to reduce the secondary flow ratio and improve the
tex within the secondary channel becomes much larger and stron-
advection heat transfer.
ger. The effect of this vortex on the flow field is apparent from the
streamlines (Fig. 19b), starting from the indicated surface, adjacent
to the trailing edge. When the Reynolds number is below Rec-o, the 5.7. Local heat transfer performance
streamlines starting from the downstream of the 45° oblique fin
mix with the main stream without any disturbance. However, The contribution of the vortices to convective heat transfer can
when Rec-o is exceeded, the vortex within the secondary channel be quantitatively studied by comparing the distribution of local
becomes powerful enough to create an adverse pressure gradient convection coefficient hlocal over the oblique fin walls for Reynolds
and draw the air flow from the downstream of the fin back into numbers below and above Rec-o. The local convection coefficient is
the oblique channel. The portion of the air flow, drawn back into defined as
the secondary channel, follows a circular path around the vortex q00local
and then joins back to the main stream, removing the high temper- hlocal ¼ ð46Þ
T w;local  T ref
ature air away from the fin wall.
The explanation for the decreasing secondary flow rates with The local wall heat flux and the local wall temperature can only be
increasing inlet air flow rate is trivial: once Rec-o is exceeded and downloaded from ANSYS Fluent simulation results as long as the

θ=45°, Re=239<Re c-o θ=45°, Re=489>Re c-o


Top view

(a)

Flow direction Flow direction

trailing
Top view

edge
(b)
separation leading
Streamline starting surface edge edge
Side view

(c)

Fig. 19. Flow field around 45° oblique fin below and above cross-over Reynolds number.
408 B. Kanargi et al. / International Journal of Heat and Mass Transfer 116 (2018) 393–416

θ=30°, Re=489<Re c-o θ=30°, Re=739>Re c-o


Top view

(a)

Flow direction Flow direction


trailing
edge
Top view

(b)
separation leading
Streamline starting surface
edge edge
Side view

(c)

Fig. 20. Flow field around 30° oblique fin below and above cross-over Reynolds number.

mesh between the solid and fluid domains are conformal, in are reduced, and hlocal increases regularly on Surface #1 and #2 for
another word: matching, so that the variables of Eq. (46) can be both OF HS designs.
accurately calculated across the solid-fluid interface. In this regard, The vortices induced within the secondary channels after
the numerical simulation results of the periodic, single channel exceeding Rec-o, act as obstacles against the secondary flows,
computational domains of the 30° and 45° OF HS were not useful resulting in a decrease in the secondary flow ratios. However, the
due to the non-conformal mesh interface between the heat sink vortices induced within the secondary channels decrease the effec-
and the air domain. To overcome this problem, merely the air flow tive channel width, especially at the inlet of the secondary chan-
domains were remodeled without the heat sink, the fin walls were nels, resulting in a local increase in the flow velocity. Thus, hlocal
attended with constant wall temperature of 60 °C, and the simula- on Surface #3 shows a slightly better improvement at the entry
tions were repeated for the 30° and 45° OF HS designs at Reynolds of the channel due to increased flow velocity and reduced bound-
numbers below and above Rec-o. Hence, the thermal boundary con- ary layer thicknesses.
dition was directly introduced to the fluid domain mesh elements Finally, the effect of the vortices on convective heat transfer can
for an accurate post-processing of the data. be observed from hlocal on Surface #4. Upon exceeding Rec-o, hlocal
The reference temperature selection is important in local con- on Surface #4 increases not only in magnitude, but also in trend,
vection coefficient calculations [33]. If a very fine boundary layer the increase making a peak around the Corners: A & B. Once the
mesh is used at the wall, the wall adjacent fluid temperature can- downstream flow is drawn back into the secondary channel, hlocal
not be used as Tref. Since the first layer thickness of the mesh is makes the first peak at Corner-B. The increase in hlocal at Corner-
very small, the wall adjacent temperature would be very close to B is very much clear especially for the 45° OF HS, while a trendwise
the wall temperature, resulting in an unrealistically high hlocal increase in hlocal can be observed for the 30° OF HS. hlocal also
value. Therefore, in this study, the average bulk temperature of shows a slight increase at Corner-A on Surface #4, where the vor-
the air domain was taken as Tref. As a final note, the entire param- tex within the secondary channel separates from Surface#4 and
eters of Eq. (46) were calculated at the mid-channel height, and removes the heat away from the surface perpendicularly.
hlocal data was calculated for the fins #13, which are located at
the halfway in the streamwise direction. Therefore, the flow fields
were free from the entrance effects. 5.8. Advection characteristics
Fig. 21 shows hlocal over the four walls of the 30° and 45° obli-
que fins. The arrows shown around the oblique fins in the legend A better flow mixing improves the advection characteristics of a
represent both the flow direction and the distance on the fin walls, heat sink. It is possible to study the flow mixing characteristics of a
located on the abscissa of each subplot. heat sink with qualitative methods. The trajectories of the massless
When the Reynolds number exceeds Rec-o, the velocity of the air particles, released at the inlet of the flow domain, can be tracked at
flow in the main channel increases, the boundary layer thicknesses different locations in the streamwise direction. These trajectories
B. Kanargi et al. / International Journal of Heat and Mass Transfer 116 (2018) 393–416 409

Corner-A θ=45° Corner-A θ=30°


1 1
4 4 4
3 4 3

2 2
Corner-B Corner-B

A
A B B

Fig. 21. Distributions of local convection coefficient on 30° and 45° oblique fin walls below and above cross-over Reynolds numbers.

are then plotted at cross sections of interest in the downstream, of the air flow domains through the 30° and 45° OF HS designs. The
preferably at the heat sink outlet, for a qualitative understanding two groups of particles did not mix with each other as if there
of the system [34,35]. These maps are known as the Poincaré sec- existed a membrane at the mid-channel height. Although, Fig. 22
tions and they provide qualitative information whether the flow gives the impression that there was not effective flow mixing
regime is laminar, transitional or chaotic [36]. For steady state through the channels, flow mixing actually occurred within the
flows, the pathlines of the particle trajectories coincide with the upper and lower halves of the channels. When the streamlines
streamlines [26], therefore tracking the coordinates of the stream- were initialized merely from the upper and lower one eighth por-
lines at different locations in the downstream would yield the tions of the channels at the heat sink inlet, the Poincaré sections in
same results as tracing the trajectories of massless particles. Fig. 23b–e at the heat sink outlets showed that streamlines actually
The streamline trajectories of the 30° and 45° OF HS designs at dispersed within their respective halves of the channel. The flow
Reynolds numbers below and above Rec-o, were plotted at heat sink mixing was observed to be more pronounced for the 30° OF HS
inlet and outlet as Poincaré sections in Figs. 22 and 23. compared to the 45° OF HS, which can be explained by the higher
The streamlines released at the heat sink inlet in Fig. 22a were secondary flow rates and stronger vortices of the former than those
distinguished with blue and red colors for the upper and lower of the latter. On the other hand, exceeding Rec-o improved the flow
halves of the channels, respectively. Fig. 22b–e present the loca- mixing characteristics of the 45° OF HS more significantly. Because,
tions of the streamlines at the heat sink outlet. It was observed that below Rec-o, the 45° OF HS could not induce a vortex strong enough
there occurred no flow mixing between the upper and lower halves to draw streams of air from the downstream of the fin back into the
410 B. Kanargi et al. / International Journal of Heat and Mass Transfer 116 (2018) 393–416

Fig. 22. Poincaré maps at the inlet and outlets of 30° and 45° oblique-finned heat sink below and above cross-over Reynolds numbers.

Fig. 23. Poincaré maps at the inlet and outlets of 30° and 45° oblique-finned heat sink below and above cross-over Reynolds numbers (streamlines were initialized from the
top and bottom 1/8th portion of heat sink inlet).

secondary channel (Fig. 19c), while the 30° OF HS could do so random distribution was observed at the Poincaré sections at the
(Fig. 20c). heat sink outlet in Fig. 23d-e. On the other hand, the Poincaré sec-
The Poincaré sections presented in Figs. 22 and 23 also provide tions at the heat sink outlets of the 45° OF HS in Fig. 23b-c did not
information on the flow regimes through the OF HS. To distinguish show a streamline distribution as random as that of the 30° OF HS.
between the regular or chaotic advection regimes, the trajectories The relative positions of the streamlines at the heat sink inlet were
of the adjacent particles/streamlines should be tracked. If the still observable at the heat sink outlet of the 45° OF HS. It can be
neighboring particles/streamlines maintain their relative positions said that the heat transfer enhancement of the 30° OF HS is based
with respect to each other as they advance in the streamwise not only on the repetitive thermal boundary layer disruption over
direction, the advection regime is said to be regular. If the parti- the oblique fin sections, but also on chaotic advection. The chaotic
cles/streamlines disperse to the flow cross sectional area randomly, advection regime improves flow mixing significantly and enables a
then the advection regime is termed as chaotic [36]. The streamli- uniform temperature build up (Fig. 16b).
nes, released around the mid-channel height of the heat sink inlet
in Fig. 22a, were observed in the Poincaré sections at the OF HS 5.9. Secondary flow field
outlets to have stayed relatively close to each other, which sug-
gests that the advection regime was closer to being regular. As Fig. 24 shows the velocity and temperature fields of the 30° and
for the streamlines starting from the top and bottom one eighth 45° OF HS designs, captured from the full domain simulation
portions of the heat sink inlet of the 30° OF HS as in Fig. 23a, a results in the high Reynolds number regime. At Re = 3117, the
B. Kanargi et al. / International Journal of Heat and Mass Transfer 116 (2018) 393–416 411

θ=30°, Re=3117 θ=45°, Re=3117

Filling Counter
rotating
edge vortices

Middle
channel
(periodic
behavior)

Draining Counter
rotating
edge vortices

Flow direction

Fig. 24. Flow and temperature fields at filling edge, draining edge, and central portions of 30° and 45° oblique-finned heat sink at high Reynolds numbers.

pressure drop penalties of both of the OF HS designs exceed their constraint other than the identical mass flow rate should be
heat transfer enhancement relative to the SC HS (Fig. 18). applied, such as identical pressure drop or identical fan/pumping
It was observed that a pair of counter rotating vortices were power, as suggested by Webb and Kim [5]. In addition, the perfor-
induced in the secondary channels of the 45° OF HS in the vicinity mance evaluation criterion has to be selected, as well, since the
of the draining and filling edges. Since sufficient rates of secondary effectiveness of an enhanced heat transfer surface could be charac-
flow rates could not be induced, the vortex induced right after the terized by a variety of definitions [37] such as:
separation edge of the 45° OF HS had to be counter-balanced by a
smaller vortex within the secondary channel. The counter rotating  reduced surface area (for a given heat duty and fan/pumping
pair of vortices created recirculation regions, trapping higher tem- power)
perature air within the secondary channels, compromising the  increased heat duty (for identical surface area and fan/pumping
advection performance, as can be observed from the temperature power)
contour. As for the 30° OF HS, a counter rotating vortex was not  reduced fan/pumping power (for a given heat duty and surface
visible since the 30° oblique channels were more inline with the area)
main channels and could still sustain small amounts of secondary
The most commonly used overall performance parameter in the
flow rates at the draining and filling edges (Fig. 14b). At Re = 3117,
literature [4,38–45], given in (47), calculates the ratio of the Nus-
a significant amount of flow migrates from the draining edge to the
selt numbers of an enhanced ‘‘e” heat transfer surface to a refer-
filling edge, causing an almost entirely secondary channel oriented
ence ‘‘0” one at identical fan powers. In other words, it calculates
flow in the periodic region of the heat sink and improving the heat
how good the heat transfer performance of an enhanced heat
transfer performance significantly. However, lower secondary air
sink/exchanger is relative to a reference one, once the former is
flow rates in the near edge region compromises the heat transfer
operated at identical fan/pumping power as the latter. The overall
enhancement by compromising the advection performance.
performance parameter will hereafter be referred to as the
thermal-hydraulic performance factor gth for conciseness.
5.10. Thermal-hydraulic performance factor 
 Nue 
Nue  Nu0 
Evaluating the heat transfer enhancement ratio ENu of the OF HS gth ¼ ¼   Re1=3 ð47Þ
Nu0 pfan fe
f 0
relative to the SC HS helped to discover the cross-over Reynolds
Re
number, beyond which the heat transfer performance of the OF
HS exceeds its pressure drop penalty. However, the heat transfer To calculate the thermal-hydraulic performance factor, the average
enhancement ratio that is calculated at identical mass flow rates Nusselt numbers and friction factors of the reference surface should
does not represent the real heat transfer enhancement of enhanced be available in the form of tabulated data or, more conveniently, in
heat transfer surfaces, since they generally induce higher pressure the form of correlations; while the performance data for the
drop penalties and operate at higher fan/pumping powers. The ref- enhanced surface can be obtained either via numerical simulations
erence heat transfer surface would have performed better, if it or via experiments [46].
were allowed to operate at identical fan/pumping power as the In this study, the thermal-hydraulic performance factors
enhanced heat transfer surface. Therefore, for a fair evaluation of calculated by (47) were intended to be validated. For this purpose,
the overall performance of an enhanced heat transfer surface, a performance factors were directly evaluated as the ratios of
412 B. Kanargi et al. / International Journal of Heat and Mass Transfer 116 (2018) 393–416

experimental Nusselt numbers of the OF HS and SC HS at identical


fan powers. First, the experimental NuL data of the SC HS were cor-
related against the fan power as

Nu0 ðpfan Þ ¼ 13:69p0:3941


fan þ 2:79 ð48Þ

Next, the performance factors were calculated using the expression,


given in (49). Nue values and the fan powers of the OF HS were
taken pointwise from the experimental data, while the Nu0 values
that corresponded to the fan powers of the OF HS were evaluated
in the denominator.

Nue  Nue ðpfan Þ
gth;v alidation ¼ ¼ ð49Þ
Nu0 pfan 13:69p0:3941
fan þ 2:79

The two sets of thermal-hydraulic performance data were plotted in


Fig. 25. It was observed that the conventional expression (47)
approximated the validated performance factors (49) more accu-
rately for moderate to high Reynolds numbers; however, it under-
predicted for low Reynolds numbers. Eq. (47) is widely used in
Fig. 26. Derivation parameters of thermal-hydraulic performance factor.
the literature without an emphasis on the Reynolds number range.
During the derivation of the overall performance factor for the
roughened tubes relative to the smooth tubes, Webb [37] utilized
the friction factor and Stanton number correlations for turbulent The Reynolds numbers Ree and Re0, indicated in Fig. 26, are selected
flow through roughened tubes, reported in their previous papers in a way that the fan power of the enhanced heat sink at Ree is equal
[47,48] for Reynolds numbers well beyond 10,000. Therefore, an to the fan power of the reference heat sink at Re0. Rewriting Eq. (28)
overall performance parameter in the form of Eq. (47) may not for DP and substituting it into (25), fan power can be written as
always hold true for the laminar regime since the correlations for  3 fAflow cr-sec Lhs
2qu
the reference and the enhanced heat transfer surfaces may differ. pfan ¼ ð52Þ
Dh
Therefore, in this section, a low Reynolds number correction to
Eq. (47) was made. A general form of the thermal-hydraulic perfor- In a case where the enhanced and the reference heat sink geome-
mance factor expression, previously derived by Fan et al. [46] for tries are geometrically similar (Aflow cr-sec, Lhs, Dh) and the fluid
various constraints, was rederived for the case at hand for identical properties (q, l) are comparable, identical fan power condition
fan power constraint using a simpler, graphical procedure, which for the enhanced and the reference heat sink designs can be written
was explained below. as
Fig. 26 shows the hypothetical trends of the friction factors and
f e ðRee Þ:Re3e ¼ f 0 ðRe0 Þ:Re30 ð53Þ
the Nusselt numbers for any reference and enhanced heat sink
design, represented with the subscripts ‘‘0” and ‘‘e”, respectively; The thermal-hydraulic performance factor, defined as the ratio of
for two consecutive Reynolds numbers: Ree and Re0. The friction the Nusselt numbers of the enhanced heat sink design to those of
factors and the Nusselt numbers of the reference geometry are the reference one at identical fan powers [5] can be expressed
given by the following correlations: according to Fig. 26 as:
f 0 ¼ c1 Rem1 ð50Þ Nue ðRee Þ
gth ¼ ð54Þ
Nu0 ðRe0 Þ
m2
Nu0 ¼ c2 Re ð51Þ
Using the chain rule, Eq. (54) can be expanded as
Nue ðRee Þ Nue ðRee Þ Nu0 ðRee Þ
gth ¼ ¼ ð55Þ
Nu0 ðRe0 Þ Nu0 ðRee Þ Nu0 ðRe0 Þ
Substituting Nu0 terms in the second fraction using Eq. (51),
Nue ðRee Þ c2 Rem 2
Nue ðRee Þ=Nu0 ðRee Þ
gth ¼ m2 ¼
e
ð56Þ
Nu0 ðRee Þ c2 Re0
m
c2 Re0 2
m
c2 Ree 2

To be able to express the thermal-hydraulic performance factor as


in (56), the ðRe0 =Ree Þm2 expression should be substituted with
ηth

ηth, validation (Eqn. 49)


appropriate terms. Those terms will be obtained from the equal
ηth, conventional (Eqn. 47) fan power condition stated in (53). Applying the chain rule to the
friction factors:
f e ðRee Þ f e ðRee Þ f 0 ðRee Þ
¼ ð57Þ
f 0 ðRe0 Þ f 0 ðRee Þ f 0 ðRe0 Þ

If both sides are multiplied by Re3e =Re30 , the equality becomes equal
to unity due to the equal fan power condition. Recollecting,
Ree
f e ðRee Þ f 0 ðRee Þ Re3e
Fig. 25. Comparison of thermal-hydraulic performance factors calculated using ¼1 ð58Þ
conventional expression (47) and validation expression (49).
f 0 ðRee Þ f 0 ðRe0 Þ Re30
B. Kanargi et al. / International Journal of Heat and Mass Transfer 116 (2018) 393–416 413

Substituting f0 terms of the second fraction with Eq. (50), value as 0.161 for low Reynolds numbers, and 0.335 for the high
Reynolds numbers, according to the NuL and fF correlations of the
f e ðRee Þ c1 Rem 1
Re3e
e
m1 ¼1 ð59Þ SC HS experimental data given (62)–(64). The derived expression
f 0 ðRee Þ c1 Re0 Re30 (61) is almost identical to the conventional expression (47) in the
turbulent flow regime; however, the varying heat transfer behavior
From here, Re0 =Ree can be written as
has to be taken into consideration not to misevaluate the thermal-
  1
Re0 f e ðRee Þ 3þm1 hydraulic performance factor at low Reynolds numbers.
¼ ð60Þ Fig. 28 shows the experimental gth (61) of the 30° and 45° OF
Ree f 0 ðRee Þ
HS designs, in comparison with the Nusselt number enhancement
Substituting Eq. (60) into (56), the thermal-hydraulic performance ratios ENu, evaluated at identical air flow rates (43). gth of the 30°
factor becomes OF HS deviate from ENu more significantly compared to the 45° OF

Nue 
HS, due to the fact that the 30° OF HS incurs a higher pressure drop
Nu0  penalty. gth of the 30° and 45° OF HS designs were observed to
Ree
gth;deriv ed ¼  3þm ð61Þ
 m2
experience local maximums within the range of air flow rates
fe 1

f 0 investigated. The maximum gth were observed at moderate Rey-


Ree
nolds numbers, around Re = 1500 for the 30° OF HS, and around
The thermal-hydraulic performance factor evaluates the enhance- Re = 2500 for the 45° OF HS. For higher Reynolds numbers, the
ment of the convection coefficient of the OF HS relative to the SC pressure drop penalty increases more significantly than the heat
HS at identical fan powers since the hydraulic diameters were taken transfer enhancement, resulting in a decreasing trend in gth.
to be equal and the changes in the thermal and transport properties
of the coolant are negligible. The coefficients m1 and m2, should be
5.11. Heat sink junction temperature improvement
substituted from the friction factor and the Nusselt number correla-
tions of the SC HS, which are to be written in the form of (50) and
Although the Nusselt number and the friction factor are the
(51). The experimental data for the SC HS, presented in Fig. 8, were
most commonly used dimensionless performance metrics for the
correlated as
evaluation and the benchmarking of the thermal and hydraulic
f F 0 ¼ 3:543Re0:6917 ð62Þ performances of heat sinks; from a commercial product design
point of view, the most important performance metrics for the
NuL0 ¼ 0:4876Re0:3717 for Re < 2000 ð63Þ thermal management systems are the device junction tempera-
tures and solder temperatures [49]. So as to ensure the commercial
success of a product, these two temperature limits must be safely
NuL0 ¼ 0:023Re0:774 for Re > 2000 ð64Þ
and reliably satisfied, regardless of the coolant type selected and
Fig. 27 presents the thermal-hydraulic performance factors calcu- the extent of the heat dissipation. Therefore, it is vital to know
lated using the conventional expression (47), the derived expres- the average junction temperature Tj improvement achievable by
sion (61) along with the validation expression (49). The horizontal the OF HS design with respect to the conventional SC HS. To make
axis was aptly chosen as Ree since the Nusselt numbers and friction a perfectly fair comparison of the junction temperatures of
factors of the SC HS were evaluated at Ree in (62)–(64) for the eval-
uation of the Nusselt number enhancement and pressure drop pen-
alty ratios in (61). ENu 30º OF HS (43)
It was observed that the derived expression (61) could predict
the thermal-hydraulic performance factors, evaluated by (49),
ηth, derived 30º OF HS (61)
accurately for the entire range of Reynolds numbers. The conven-
tional expression (47) takes the value of 0.333 as the power of
ENu 45º OF HS (43)
the ðf e =f 0 ÞjRee term, while the derived expression calculates this

ηth, derived 45º OF HS (61)

ηth, validation (Eqn. 49)


ηth

ηth

ηth, derived (Eqn. 61)

ηth, conventional (Eqn. 47)

Ree
Ree
Fig. 27. Comparison of thermal-hydraulic performance factors, evaluated the using
conventional expression (47), validation expression (49), and derived expression Fig. 28. Comparison of Nusselt number enhancement ratios ENu (43) and thermal-
(61). hydraulic performance factors gth (61) of the 30° and 45° OF HS.
414 B. Kanargi et al. / International Journal of Heat and Mass Transfer 116 (2018) 393–416

enhanced and reference heat sink designs, the heat sink foot prints Table 4
and the heater powers should be identical; therefore, the results of Operational parameters of SC HS & 30° OF HS at identical fan powers.

the current study provides a fair comparison. SC HS 30° OF HS


Fig. 29 shows the junction temperatures obtained with the 30° Fan power [W] 0.219 ± 0.0035 0.221 ± 0.0034
and 45° OF HS designs and the conventional SC HS at 20 W and Qheater [W] 60 ± 0.48 60 ± 0.46
60 W heating powers, with respect to (a) Reynolds number and Tjunction avg [°C] 51.74 ± 0.34 41.01 ± 0.34
(b) fan power. The error bars were not shown for clarity. Each junc- Tair out [°C] 32.32 ± 0.17 34.29 ± 0.17
m_ air [kg/s] 0.00554178 ± 0.000083 0.00411426 ± 0.000062
tion temperature measurement has an uncertainty of ±0.34 °C.
Re 2490 ± 38 1848.7 ± 28
Compared to the SC HS, at identical air flow rates, the 45° and
30° OF HS designs can improve the junction temperature by up
to 4.5 °C and 9 °C, respectively, at 20 W; and by up to 9 °C and
20 °C, respectively, at 60 W effective heating powers. The improve- 0.219 W and 0.054 W, respectively. Therefore, the 30° OF HS pro-
ment in junction temperature is more significant at moderate Rey- vided a fan power improvement of 75% relative to the SC HS.
nolds numbers since the benefits of the boundary layer disruption The detailed investigations of the thermal and hydraulic perfor-
and flow mixing phenomena are more pronounced at these air flow mances of the OF HS designs showed that the use of the oblique
rates. As the air flow rate is increased, the improvement in junction fins is an effective way of improving the air cooling performance.
temperature decreases; because at higher air flow rates, the SC HS The oblique fin design can effectively reduce the boundary layer
can already sustain thinner boundary layers and turbulent flow thicknesses over the walls of an oblique fin, while the vortex
regime improves the flow mixing and the advection properties. induced within the secondary channels improves the advection
In addition, as the air flow rate is increased, heat sink temperatures characteristics, yielding a heat transfer enhancement exceeding
inevitably approach to the air inlet temperature. the pressure drop penalty. This study showed that the effective-
Fig. 29b shows the improvements in junction temperature ness of air cooling can still be maintained by a carefully tailored
obtained with the OF HS designs relative to the SC HS at identical air flow, and air cooling can be utilized for the thermal manage-
fan powers. Compared to the SC HS, the 45° and 30° OF HS could ment systems of the future.
improve the junction temperature by up to 8 °C and 16 °C for
60 W effective heating power at identical fan powers. 6. Conclusions
Table 4 summarizes the operational parameters of the SC HS
and the 30° OF HS for 60 W effective heater powers and at equal The heat transfer and pressure drop performances of the 30°
fan powers of 0.2 W. Due to the increase in the pressure drop and 45° OF HS designs were evaluated and compared with those
penalty of the 30° OF HS, there was a 25% drop in the air flow rate of the conventional SC HS of equivalent dimensions. The SC HS
compared to the SC HS. A reduction in the air flow rate caused an and OF HS prototypes were tested on the compact wind tunnel
increase in the air outlet temperature by approximately 2 °C; while experimental setup for a wide range of Reynolds numbers. Upon
at the same time, lowering junction temperature by more than successful validation of the numerical results, detailed flow and
10 °C. temperature field analysis were performed. The key findings are
The improved performance of the enhanced heat transfer sur- listed as follows:
faces can also be quantified from the improvement in the fan
power. A lower fan power to maintain the same junction temper-  The single channel, periodic computational model, comprising
ature can improve the battery life of mobile equipment at a single the smallest periodically repeating portion of the OF HS, overes-
charge due to reduced energy consumption, or allows the installa- timated the experimental Nusselt numbers and friction factors
tion of smaller fans, which can help reducing the outer dimensions of the 30° and 45° OF HS designs due to the fully periodic flow
of the devices, improving the material and manufacturing costs, assumption throughout the heat sink in the spanwise direction.
weight and mobility. For example, to maintain a junction temper-  The 3D full domain numerical simulations captured the flow
ature of 52 °C, the approximate fan powers required by the SC HS migration predicting the experimental Nusselt numbers and
and the 30° OF HS for effective heating power of 60 W were friction factors of the OF HS more accurately.

90 90
SC HS 60W SC HS 60W

80 (a) 45deg OF HS 60W


30deg OF HS 60W 80 (b)
45deg OF HS 60W
30deg OF HS 60W
SC HS 20W SC HS 20W
45deg OF HS 20W 45deg OF HS 20W
70 30deg OF HS 20W
70 30deg OF HS 20W

60 60
Tj [°C]
Tj [°C]

50 50

40 40

30 30

20 20
0 1000 2000 3000 4000 0.0001 0.001 0.01 0.1 1 10
Re Fan power [W]

Fig. 29. Junction temperature improvements obtained with 30° and 45° OF HS at 20 W and 60 W heating powers as a function of (a) Reynolds number and (b) fan power.
B. Kanargi et al. / International Journal of Heat and Mass Transfer 116 (2018) 393–416 415

 The secondary flows, induced through the oblique channels of [7] L. Chai, G.D. Xia, H.S. Wang, Numerical study of laminar flow and heat transfer
in microchannel heat sink with offset ribs on sidewalls, Appl. Therm. Eng. 92
the OF HS, effectively disturbed and reinitialized the boundary
(2016) 32–41, http://dx.doi.org/10.1016/j.applthermaleng.2015.09.071.
layers on three walls of the oblique fins, reducing the thermal [8] G. Xia, L. Chai, H. Wang, M. Zhou, Z. Cui, Optimum thermal design of
convective resistance for the entire range of air flow rates. microchannel heat sink with triangular reentrant cavities, Appl. Therm. Eng. 31
 At very low Reynolds numbers, the heat transfer enhancement (2011) 1208–1219, http://dx.doi.org/10.1016/j.applthermaleng.2010.12.022.
[9] G. Xia, D. Ma, Y. Zhai, Y. Li, R. Liu, M. Du, Experimental and numerical study of
ENu of the OF HS, was less than its pressure drop penalty Ef. fluid flow and heat transfer characteristics in microchannel heat sink with
Upon exceeding the cross-over Reynolds number Rec-o, ENu complex structure, Energy Convers. Manage. 105 (2015) 848–857, http://dx.
was observed to exceed Ef for both OF HS designs. The flow field doi.org/10.1016/j.enconman.2015.08.042.
[10] N.C. DeJong, A.M. Jacobi, Flow, heat transfer, and pressure drop in the near-
investigations showed that after exceeding Rec-o, secondary wall region of louvered-fin arrays, Exp. Therm. Fluid Sci. 27 (2003) 237–250,
flows entering the secondary channels separated and induced http://dx.doi.org/10.1016/S0894-1777(02)00224-8.
vortices that contributed to advection heat transfer. [11] Y.J. Lee, P.S. Lee, S.K. Chou, Enhanced thermal transport in microchannel using
oblique fins, J. Heat Transfer 134 (2012) 101901, http://dx.doi.org/10.1115/
 The secondary flows enabled a proper mixing of air streams 1.4006843.
flowing through the core of the main channels and in the vicin- [12] Y. Fan, P.S. Lee, L.-W. Jin, B.W. Chua, A simulation and experimental study of
ity of the fin walls, enabling a more uniform increase in the air fluid flow and heat transfer on cylindrical oblique-finned heat sink, Int. J. Heat
Mass Transf. 61 (2013) 62–72, http://dx.doi.org/10.1016/j.
temperature in the streamwise direction. A lower air tempera- ijheatmasstransfer.2013.01.075.
ture in the vicinity of the fin walls helped to maintain lower [13] Y.K. Prajapati, M. Pathak, M. Kaleem Khan, A comparative study of flow boiling
wall temperatures. heat transfer in three different configurations of microchannels, Int. J. Heat
Mass Transf. 85 (2015) 711–722, http://dx.doi.org/10.1016/j.
 At very high Reynolds numbers, the cumulative secondary air
ijheatmasstransfer.2015.02.016.
flow rates through the secondary channels in the vicinity of [14] M. Law, O.B. Kanargi, P.S. Lee, Effects of varying oblique angles on flow boiling
the edges of the heat sink were diminished significantly due heat transfer and pressure characteristics in oblique-finned microchannels,
to the draining and accumulation of flow through the main Int. J. Heat Mass Transf. 100 (2016) 646–660, http://dx.doi.org/10.1016/j.
ijheatmasstransfer.2016.04.077.
channels. Therefore, the heat transfer enhancements of both [15] I.A. Ghani, N.A.C. Sidik, N. Kamaruzaman, Hydrothermal performance of
OF HS designs were compromised at high Reynolds numbers. microchannel heat sink: the effect of channel design, Int. J. Heat Mass Transf.
 The advection heat transfer was determined to be a combina- 107 (2017) 21–44, http://dx.doi.org/10.1016/j.ijheatmasstransfer.2016.11.031.
[16] P.N.N. Shankar, M.D.D. Deshpande, Fluid mechanics in the driven cavity, Annu.
tion of regular and chaotic regimes for the 30° OF HS by the Rev. Fluid Mech. 32 (2003) 93–136, http://dx.doi.org/10.1146/annurev.
use of Poincaré maps, while the 45° OF HS showed rather regu- fluid.32.1.93.
lar advection characteristics. [17] ANSYS Fluent 14.0 Theory Guide, ANSYS, Inc., 2011.
[18] O.B. Kanargi, P.S. Lee, C. Yap, A numerical and experimental investigation of
 The thermal-hydraulic performance factor of the 30° OF HS heat transfer and fluid flow characteristics of a cross-connected alternating
exceeded that of the 45° OF HS. The thermal-hydraulic perfor- converging-diverging channel heat sink, Int. J. Heat Mass Transf. (2016) 1–24,
mance factors of both OF HS designs peaked at moderate Rey- http://dx.doi.org/10.1016/j.ijheatmasstransfer.2016.08.057.
[19] ANSYS Fluent 14.0 User’s Guide, ANSYS, Inc., 2011.
nolds numbers. [20] V. Yakhot, L.M. Smith, The renormalization group, the e-expansion and
 Significant improvement in junction temperature was observed. derivation of turbulence models, J. Sci. Comput. 7 (1992) 35–61, http://dx.
The 30° OF HS, which outperformed the 45° OF HS with stronger doi.org/10.1007/BF01060210.
[21] W.M. Kays, M.E. Crawford, B. Weigand, Convective Heat and Mass Transfer, 4th
secondary air flow rates, provided up to 16 °C reduction in the
ed., McGraw-Hill, 2005.
junction temperature relative to the SC HS at identical fan pow- [22] VDI-Gesellschaft Verfahrenstechnik und Chemieingenieurwesen, ed., VDI Heat
ers for 60 W effective heating power. The OF HS also showed a Atlas, 2nd ed., Springer Berlin Heidelberg, 2010.
significant improvement of up to 75% in the fan power require- [23] J.R. Davis (Ed.), Aluminum and Aluminum Alloys, ASM International, 1993.
[24] Y.K. Prajapati, M. Pathak, M.K. Khan, Transient heat transfer characteristics of
ment to maintain a constant junction temperature. segmented finned microchannels, Exp. Therm. Fluid Sci. 79 (2016) 134–142,
http://dx.doi.org/10.1016/j.expthermflusci.2016.07.004.
[25] T.M. Jeng, S.C. Tzeng, Q.Y. Huang, Heat transfer performance of the pin-fin heat
Acknowledgements sink filled with packed brass beads under a vertical oncoming flow, Int. J. Heat
Mass Transf. 86 (2015) 531–541, http://dx.doi.org/10.1016/j.
ijheatmasstransfer.2015.03.049.
The authors gratefully acknowledge the financial support given [26] Y.A. Çengel, J.M. Cimbala, Fluid Mechanics: Fundamentals and Applications,
by the Mechanical Engineering Department of the National Univer- 3rd ed., McGraw-Hill Education, 2010.
sity of Singapore (WBS No: R265-000-800-733). [27] J. Taylor, Introduction to Error Analysis: The Study of Uncertainties in Physical
Measurements, 2nd ed., University Science Books, 1997.
[28] R.M. Curr, D. Sharma, D.G. Tatchell, Numerical predictions of some three-
Conflicts of interest dimensional boundary layers in ducts, Comput. Methods Appl. Mech. Eng. 1
(1972) 143–158, http://dx.doi.org/10.1016/0045-7825(72)90001-1.
[29] F.P. Incropera, D.P. DeWitt, T.L. Bergman, A.S. Lavine, Fundamentals of Heat
The authors of this submission have no conflicts of interest to and Mass Transfer, 6th ed., John Wiley & Sons, 2002.
declare. [30] Y.A. Çengel, Heat Transfer: A Practical Approach, McGraw-Hill Education,
2003.
[31] N. Mou, Y. Jiun Lee, P. Seng Lee, P.K. Singh, S.A. Khan, Investigations on the
References influence of flow migration on flow and heat transfer in oblique fin
microchannel array, J. Heat Transfer 138 (2016) 102403, http://dx.doi.org/
[1] P. Rodgers, V. Eveloy, M.G. Pecht, Limits of air-cooling: status and challenges, 10.1115/1.4033540.
in: Semicond. Therm. Meas. Manag. Symp. 2005 IEEE, IEEE, 2005, pp. 116–124. [32] Y. Fan, P.S. Lee, B.W. Chua, Investigation on the influence of edge effect on flow
doi: 10.1109/STHERM.2005.1412167. and temperature uniformities in cylindrical oblique-finned minichannel array,
[2] N. Goldberg, Narrow channel forced air heat sink, in: IEEE Trans. Components, Int. J. Heat Mass Transf. 70 (2014) 651–663, http://dx.doi.org/10.1016/j.
Hybrids, Manuf. Technol., 1984, pp. 154–159. doi: 10.1109/ ijheatmasstransfer.2013.11.072.
TCHMT.1984.1136326. [33] R.J. Moffat, What’s new in convective heat transfer?, Int J. Heat Fluid Flow 19
[3] R. Shah, A. London, Laminar Flow Forced Convection In Ducts: A Source Book (1998) 90–101, http://dx.doi.org/10.1016/S0142-727X(97)10014-5.
for Compact Heat Exchanger Analytical Data, Academic Press, 1978. [34] H. Ghaedamini, P.S. Lee, C.J. Teo, Developing forced convection in converging-
[4] F. Zhou, I. Catton, Numerical evaluation of flow and heat transfer in plate-pin diverging microchannels, Int. J. Heat Mass Transf. 65 (2013) 491–499, http://
fin heat sinks with various pin cross-sections, Numer. Heat Transf. Part A Appl. dx.doi.org/10.1016/j.ijheatmasstransfer.2013.06.036.
60 (2011) 107–128, http://dx.doi.org/10.1080/10407782.2011.588574. [35] Y. Sui, P.S. Lee, C.J. Teo, An experimental study of flow friction and heat transfer
[5] R.L. Webb, N.H. Kim, Principles of Enhanced Heat Transfer, 2nd ed., Taylor & in wavy microchannels with rectangular cross section, Int. J. Therm. Sci. 50
Francis, 2005. (2011) 2473–2482, http://dx.doi.org/10.1016/j.ijthermalsci.2011.06.017.
[6] F. Wang, J. Zhang, S. Wang, Investigation on flow and heat transfer [36] C.H. Amon, A.M. Guzmán, B. Morel, A.M. Guzmán, B. Morel, Lagrangian chaos,
characteristics in rectangular channel with drop-shaped pin fins, Propuls. Eulerian chaos, and mixing enhancement in converging–diverging channel
Power Res. 1 (2012) 64–70, http://dx.doi.org/10.1016/j.jppr.2012.10.003. flows, Phys. Fluids 8 (1996) 1192–1206, http://dx.doi.org/10.1063/1.868910.
416 B. Kanargi et al. / International Journal of Heat and Mass Transfer 116 (2018) 393–416

[37] R.L. Webb, E.R.G. Eckert, Application of rough surfaces to heat exchanger [43] L. Gong, K. Kota, W. Tao, Y. Joshi, Parametric numerical study of flow and heat
design, Int. J. Heat Mass Transf. 15 (1972) 1647–1658, http://dx.doi.org/ transfer in microchannels with wavy walls, J. Heat Transfer 133 (2011) 51702,
10.1016/0017-9310(72)90095-6. http://dx.doi.org/10.1115/1.4003284.
[38] A. Kumar, S. Chamoli, M. Kumar, Experimental investigation on thermal [44] Z.G. Mills, A. Warey, A. Alexeev, Heat transfer enhancement and thermal-
performance and fluid flow characteristics in heat exchanger tube with solid hydraulic performance in laminar flows through asymmetric wavy walled
hollow circular disk inserts, Appl. Therm. Eng. 100 (2016) 227–236, http://dx. channels, Int. J. Heat Mass Transf. 97 (2016) 450–460, http://dx.doi.org/
doi.org/10.1016/j.applthermaleng.2016.01.081. 10.1016/j.ijheatmasstransfer.2016.02.013.
[39] M.K. Abdolbaqi, W.H. Azmi, R. Mamat, N.M.Z.N. Mohamed, G. Najafi, [45] T. Yeom, T. Simon, T. Zhang, M. Zhang, M. North, T. Cui, Enhanced heat transfer
Experimental investigation of turbulent heat transfer by counter and co- of heat sink channels with micro pin fin roughened walls, Int. J. Heat Mass
swirling flow in a flat tube fitted with twin twisted tapes, Int. Commun. Heat Transf. 92 (2016) 617–627, http://dx.doi.org/10.1016/j.
Mass Transf. 75 (2016) 295–302, http://dx.doi.org/10.1016/j. ijheatmasstransfer.2015.09.014.
icheatmasstransfer.2016.04.021. [46] J.F. Fan, W.K. Ding, J.F. Zhang, Y.L. He, W.Q. Tao, A performance evaluation plot
[40] I.A. Ghani, N. Kamaruzaman, N.A.C. Sidik, Heat transfer augmentation in a of enhanced heat transfer techniques oriented for energy-saving, Int. J. Heat
microchannel heat sink with sinusoidal cavities and rectangular ribs, Int. J. Mass Transf. 52 (2009) 33–44, http://dx.doi.org/10.1016/j.
Heat Mass Transf. 108 (2017) 1969–1981, http://dx.doi.org/10.1016/j. ijheatmasstransfer.2008.07.006.
ijheatmasstransfer.2017.01.046. [47] R.L. Webb, E.R.G. Eckert, R.J. Goldstein, Generalized heat transfer and friction
[41] L. Chai, G.D. Xia, H.S. Wang, Laminar flow and heat transfer characteristics of correlations for tubes with repeated-rib roughness, Int. J. Heat Mass Transf. 15
interrupted microchannel heat sink with ribs in the transverse (1972) 180–184, http://dx.doi.org/10.1016/0017-9310(72)90179-2.
microchambers, Int. J. Therm. Sci. 110 (2016) 1–11, http://dx.doi.org/ [48] R.L. Webb, E.R.G. Eckert, R.J. Goldstein, Heat transfer and friction in tubes with
10.1016/j.ijthermalsci.2016.06.029. repeated-rib roughness, Int. J. Heat Mass Transf. 14 (1971) 601–617, http://dx.
[42] A.F. Al-Neama, N. Kapur, J. Summers, H.M. Thompson, An experimental and doi.org/10.1016/0017-9310(71)90009-3.
numerical investigation of the use of liquid flow in serpentine microchannels [49] K. Azar, Cooling technology options (accessed on March 22nd, 2016) Electron.
for microelectronics cooling, Appl. Therm. Eng. 116 (2017) 709–723, http://dx. Cool. Mag. (2003). http://www.electronics-cooling.com/2003/08/cooling-
doi.org/10.1016/j.applthermaleng.2017.02.001. technology-options/.

You might also like