Sie sind auf Seite 1von 16

entropy

Article
Numerical Study of Double-Layered Microchannel
Heat Sinks with Different Cross-Sectional Shapes
Daxiang Deng 1,2, *, Guang Pi 1 , Weixun Zhang 1 , Peng Wang 3 and Ting Fu 4
1 Department of Mechanical & Electrical Engineering, Xiamen University, Xiamen 361005, China;
piguang95@163.com (G.P.); zhweixun@163.com (W.Z.)
2 Shenzhen Research Institute of Xiamen University, Shenzhen 518000, China
3 Tianjin Key Laboratory for Civil Aircraft Airworthiness and Maintenance Airworthiness and Maintenance,
Civil Aviation University of China, Tianjin 300300, China; xmdxhqsjy@163.com
4 Hubei Key Laboratory of Mechanical Transmission and Manufacturing Engineering in Wuhan University of
Science and Technology, Wuhan 430000, China; futing1234gh@163.com
* Correspondence: dxdeng@xmu.edu.cn; Tel.: +86-592-2186-383

Received: 23 October 2018; Accepted: 21 December 2018; Published: 25 December 2018 

Abstract: This work numerically studies the thermal and hydraulic performance of double-layered
microchannel heat sinks (DL-MCHS) for their application in the cooling of high heat flux
microelectronic devices. The superiority of double-layered microchannel heat sinks was assessed by a
comparison with a single-layered microchannel heat sink (SL-MCHS) with the same triangular
microchannels. Five DL-MCHSs with different cross-sectional shapes—triangular, rectangular,
trapezoidal, circular and reentrant Ω-shaped—were explored and compared. The results showed
that DL-MCHS decreased wall temperatures and thermal resistance considerably, induced much
more uniform wall temperature distribution, and reduced the pressure drop and pumping power in
comparison with SL-MCHS. The DL-MCHS with trapezoidal microchannels performed the worst with
regard to thermal resistance, pressure drop, and pumping power. The DL-MCHS with rectangular
microchannels produced the best overall thermal performance and seemed to be the optimum
when thermal performance was the prime concern. Nevertheless, the DL-MCHS with reentrant
Ω-shaped microchannels should be selected when pumping power consumption was the most
important consideration.

Keywords: double-layered microchannel heat sinks; microchannel shape; numerical simulation;


thermal resistance

1. Introduction
With the rapid development of microelectronic devices, the local heat flux inside has so far
increased to more than 300 W/cm2 [1], which is far beyond the heat dissipation limit of air cooling
schemes. Heat removal is thus vital for the safe and steady operation of microelectronic devices.
Microchannel heat sinks, which were proposed by Tuckerman and Pease [2] in 1981, have been
recognized to be an efficient means to dissipate high heat flux. Due to its high surface area to volume
ratio, large heat transfer coefficient, and small coolant inventory, the microchannel heat sink has been
used in recent years as a high-performance compact cooling method in thermal dissipation applications
of very-large-scale integrated (VLSI) circuits, microelectromechanical systems, and high power laser
diode arrays [3–5].
In microchannel heat sinks, coolant flows in parallel microchannels from inlet to outlet in a single
direction. Coolant temperature increases along the stream-wise direction and results in a poor heat
exchange process between the coolant and microchannel wall. As a result, the bottom wall temperatures

Entropy 2019, 21, 16; doi:10.3390/e21010016 www.mdpi.com/journal/entropy


Entropy 2019, 21, 16 2 of 16

of the heat sink increase along the flow length. Non-uniform wall temperature distribution induces
undesirable thermal stresses in microelectronic devices and hence reduces reliability and shortens the
lifetime of microelectronic devices. To address this issue, Vafai and Zhu [6] proposed the concept of a
double-layered microchannel heat sink (DL-MCHS) in which two single-layered microchannel heat
sinks (SL-MCHS) are stacked one on top of the other. It was found that the streamwise temperature
rise was substantially reduced for DL-MCHS and that the pressure drop was also smaller than that of
SL-MCHS. Motivated by these promising results, extensive studies have been subsequently conducted
to explore the thermal and fluid properties of DL-MCHS [7–21]. Early studies mainly focused on the
effectiveness verification of the DL-MCHS design. Chong et al. [7] numerically compared single-layered
and double-layered counter flow microchannel heat sinks with rectangular shapes by employing a
thermal resistance network and 3-D computational fluid dynamics (CFD) simulations. The superiority
of DL-MCHS was demonstrated again. Cheng [8] found in numerical simulations that the thermal
resistance of DL-MCHS can be reduced by 47.7% compared to that of a SL-MCHS.
Following this, research attention has been paid to the optimization design of geometric
parameters and the flow arrangement of DL-MCHS. Wei et al. [9] fabricated a stacked microchannel
heat sink using silicon micromachining techniques. The effects of flow direction and flow rate ratio
in each microchannel layer were explored both experimentally and numerically. Thermal resistance
as low as 0.09 ◦ C/(W·cm2 ) was obtained for a stacked microchannel heat sink with two layers of
microchannels. The counter-flow arrangement provided better wall temperature uniformity whereas
the parallel flow had the best performance in reducing the peak temperature. Hung et al. [10] identified
that several dominant parameters—substrate materials, coolants, and geometric parameters such
as channel number, channel width ratio, channel aspect ratio, substrate thickness, and pumping
power—play a notable role in the temperature distribution, pressure drop, and thermal resistance
of DL-MCHS with rectangular microchannels. They also provided an optimization procedure for
the geometric parameters of DL-MCHS via a simplified conjugate-gradient method and a three-
dimensional fluid flow and heat transfer model [11]. Xie et al. [12] explored the layout effect of
parallel-flow and counter-flow for inlet and outlet flow directions on the thermal performance of a
DL-MCHS with rectangular microchannels. The results showed that the parallel-flow layout presented
better heat dissipation when the flow rate was limited to a low value whereas the counter-flow layout
did better in high flow rate cases. The effect of the height of the upper-branch and lower-branch
channels on the thermal performance of DL-MCHS has also been assessed. Lin et al. [13] conducted
optimization studies on DL-MCHS with rectangular microchannels to search for a minimum of
the thermal resistance. Six design variables, including channel number, vertical rib width, bottom
channel height, thicknesses of two horizontal ribs, and coolant velocity in the bottom channel were
simultaneously optimized by a three-dimensional solid-fluid conjugated model coupled with a
simplified conjugate-gradient. Leng et al. [14] optimized channel number, channel width, bottom
channel height, and bottom coolant inlet velocity to improve bottom wall temperature uniformity and
to reduce the overall thermal resistance for DL-MCHS with rectangular microchannels at constant
pumping powers.
Recently, DL-MCHSs with advanced configurations or flow passage design have been developed
to enhance thermal and fluid flow characteristics. The research groups of Wang et al. have developed
improved designs of double-layered rectangular microchannel heat sinks with truncated top channels [17]
and porous fins [18] to promote the thermal and hydraulic performance of DL-MCHS. Such new
designs of DL-MCHS have been found to reduce thermal resistance and pumping power considerably.
Osanloo et al. [19] and Wong et al. [20] developed DL-MCHS with tapered microchannels or channel
contraction. Higher thermal performance was achieved for such improved DL-MCHS as compared
to that with the conventional design of straight channels, but pumping power was increased.
Zhai et al. [21] have developed double-layered microchannel heat sinks with cavities and ribs in the
rectangular microchannel flow passages. The comparisons among them and conventional DL-MCHSs
and single-layered micro heat sinks with simple structures revealed that DL-MCHS with cavities and
ribs showed better heat transfer characteristics under the same volumetric flow rate.
Entropy 2019, 21, 16 3 of 16

In previous reports of double-layered microchannel heat sinks, parallel microchannels with


conventional rectangular shapes have generally been employed, and the geometric parameters of
rectangular microchannels, such as channel number, channel width, channel height and channel
aspect ratio, have been focused on. Nevertheless, except for the studies of DL-MCHS with trapezoidal
and rectangular shapes in Sharma et al. [22] and for that including DL-MCHS with boot, diamond,
hexagonal, pentagonal, rectangular, rectangular wedge, and triangular shapes in Kulkarni et al. [23],
information about DL-MCHS with different cross-sectional shapes is still far from sufficient. To address
this, we in this study developed five double-layered microchannel heat sinks with different
microchannel cross-sectional shapes, i.e., rectangular, triangular, trapezoidal, circular, and reentrant
Ω-shaped. The thermal and hydraulic performance of these five DL-MCHSs were explored by
numerical simulations. Moreover, a comparison between DL-MCHS and SL-MCHS was also conducted
using the triangular microchannels. This study sheds some light on the design of double-layered
microchannel heat sinks and is believed to be of practical importance.

2. Model Description
The double-layered microchannel heat sink is illustrated in Figure 1a. Two layers of microchannels
with the same cross-sectional shape and geometric dimensions are stacked together. Due to the
symmetric and periodic arrangement of microchannels in double-layered microchannel heat sinks,
a unit cell containing a microchannel is chosen as the computational domain, as shown in Figure 1.
Five DL-MCHS with different microchannel cross-sectional shapes—triangular, rectangular, circular,
trapezoidal and reentrant Ω-shaped—were numerically studied to explore the effect of microchannel
cross-sectional shape on the performance of DL-MCHS. The microchannels were designed to have
nearly the same hydraulic diameter, as shown in Table 1. The geometric dimensions of each
microchannel are shown in Figure 2. The reentrant Ω-shaped microchannels, which feature large
circular cavities inside and an exit narrow slot at the top of the cross section, have been found to show
good heat transfer performance in convective flow in the single-layered microchannel heat sinks [24].
The thermal and fluid flow behaviors of these microchannels in double-layered microchannel heat
sinks are thus explored in this study. The counter flow arrangement of two layers of DL-MCHS was
adopted following the findings of many previous reports [9,12,25], i.e., the fluid flows through the
lower channel in the positive Z direction and through the upper channel in the negative Z direction.
For comparison, a single-layered microchannel heat sink with triangular shapes was also studied.
The unit cell of the SL-MCHS is of the same width of 2 mm and the same length of 45 mm but had
a height half of that of the DL-MCHS. All the geometric parameters of these samples are shown in
Figure 2 and Table 1. Uniform heat power q is supplied to the bottom of the heat sink to simulate a
heating element such as micro-processor chips. The microchannel heat sinks are made of copper and
the coolant used is water.

Table 1. Specification of geometric parameters of the computational domain of microchannel heat sinks.

Arrangement and Width of Height of Length of Thickness Cross-Sectional


Hydraulic
Cross-Sectional the Unit the Unit the Unit of Bottom Area of
Sample Diameter
Shape of Cell W Cell H Cell L Wall, tb Microchannel
Dh (mm)
Microchannels (mm) (mm) (mm) (mm) Ac (mm2 )
Double-layered
DL-TRI 2 4 45 0.79 0.81 0.849
triangular
Double-layered
DL-REC 2 4 45 0.8 0.8 0.72
rectangular
Double-layered
DL-CIR 2 4 45 1.35 0.79 0.664
circular
Double-layered
DL-TRA 2 4 45 1.33 0.8 0.77
trapezoidal
Double-layered
DL-REE 2 4 45 0.93 0.76 0.587
reentrant
Single-layered
SL-TRI 2 2 45 0.79 0.81 0.849
triangular
Entropy 2018,
Entropy 20, x20,
2018, FOR PEERPEER
x FOR REVIEW
REVIEW 4 of 17 4 of 17

Double-layered
Double-layered
DLDL
-REE
-REE 2 2 4 4 45 45 0.93 0.93 0.76 0.76 0.587 0.587
reentrant
reentrant
Entropy 2019, 21, Single-layered
16 Single-layered 4 of 16
SL SL
-TRI-TRI 2 2 2 2 45 45 0.79 0.79 0.81 0.81 0.849 0.849
triangular
triangular

(a) (b)
(a) (b)

(c) (d)

(c) (d)
Figure 1. Schematic of double-layered and single-layered micro-channel heat sinks: (a) a typical
Figure 1. Schematic
DL-MCHS; (b) a unitofcell
double-layered and single-layered
for the computational micro-channel
domain of DL-MCHS; heat sinks:
(c) a typical (a) a (d)
SL-MCHS; typical
a
Figure
DL-MCHS; 1 .
(b)the
unit cell for Schematic
a unit of double-layered
cell for thedomain
computational computationaland single-layered micro-channel heat sinks:
domain of DL-MCHS; (c) a typical SL-MCHS; (d)
of SL-MCHS. (a) a typical
a
DL-MCHS;
unit cell (b) a unit celldomain
for the computational for the computational
of SL-MCHS. domain of DL-MCHS; (c) a typical SL-MCHS; (d) a
unit cell for the computational domain of SL-MCHS.

(a) (b) (c) (d) (e)

Figure
Figure 2. Schematic
2. Schematic of DL-MCHS
of the the DL-MCHS with different
with different microchannel
microchannel cross-sectional
cross-sectional shapes: shapes: (a)
(a) triangular;
(a) (b) (c)
(b)triangular;
rectangular; rectangular; (b)
(c)
circular; (d) (c)(e)Ω-shaped.
circular; (d) trapezoidal;
trapezoidal; (e) reentrant (d)
reentrant Ω-shaped. (e)

Figure
3. Numerical 2. Schematic of the DL-MCHS with different microchannel cross-sectional shapes: (a)
Methods
triangular; (b) rectangular; (c) circular; (d) trapezoidal; (e) reentrant Ω-shaped.
3.1. Governing Equations and Models
A three-dimensional solid–fluid conjugate model is utilized to determine the thermal and
hydraulic performance of microchannel heat sinks. The following assumptions are taken into account
to simplify the analysis: (1) the flow is three-dimensional, incompressive, laminar, and in steady-state;
(2) the effect of gravity forces are considered; (3) the fluid thermophysical properties are temperature-
Entropy 2019, 21, 16 5 of 16

dependent; (4) radiation heat transfer is negligible; (5) all solid walls of the channel are no-slip and
impermeable; (6) viscous dissipation effects are negligible.
According to the above assumptions, the governing equations can be expressed as follows:
Continuity equation:

∇ · (ρ v ) = 0 (1)

Momentum equation:
→→ →
∇ · (ρ v v ) = −∇ P + ∇ · (µ∇ v ) (2)

Energy equation for the liquid:



∇ · (ρ v c p T ) = ∇ · (k f ∇ T ) (3)

Energy equation for the solid microchannel:

∇ · (k m ∇ T ) = 0 (4)

3.2. Boundary Conditions


The boundary conditions are provided as follows. In the solid region, a uniform heat flux
(12.67 W/cm2 ) was applied to the bottom surface of the unit cell. The top surface above the
microchannel fins and liquid is set to be adiabatic. A symmetry wall boundary condition is applied
on both outer lateral planes of the computational domain. In the fluid region, a fully developed and
uniform liquid velocity with two values (0.0625 m/s and 0.125 m/s) and a constant inlet temperature
(Tin = 33 ◦ C or 306 K) is applied to both inlets of the upper and lower layers of microchannels in
DL-MCHS. At the outlet of the microchannels, a pressure outlet boundary condition is specified as the
atmospheric pressure.

3.3. Numerical Methods


The computations are carried out in a finite-volume based computational fluid dynamics (CFD)
software package, FLUENT. The simulation domains are generated in Gambit v2.3. A second-order
upwind scheme is used to discretize the convective term and a QUICK scheme is used for discretization
of the diffusion term. The coupling between pressure and velocity is implemented by a SIMPLE
algorithm. The solutions are thought to be converged when the normalized residuals are less than
10−3 for the flow equations and 10−7 for the energy equations. Unstructured grids with finer meshes
Entropy 2018, 20, x FOR PEER REVIEW 6 of 17
near the fluid-solid wall regions are utilized, as illustrated in Figure 3.

(a) (b) (c) (d) (e)


Figure 3.3.Illustration
Figure Illustrationofof grid
grid generations
generations for DL-TRI
for DL-TRI with with different
different microchannel
microchannel cross-sectional
cross-sectional shapes:
(a) triangular; (b) rectangular; (c) circular; (d) trapezoidal; (e) reentrant Ω-shaped.
shapes: (a) triangular; (b) rectangular; (c) circular; (d) trapezoidal; (e) reentrant Ω-shaped.

Grid sensitivity analysis was performed to check the independence of grids. The variations in
pressure drop from inlet to outlet of the heat sink were chosen for the evaluation of grid
dependence. Three non-uniform grid systems with denser grid clustering near the wall were
employed with 0.204, 0.412, and 0.628 million grids, respectively. The deviations in pressure drop
using 0.204 and 0.412 million grids as compared to that using 0.628 million grids were 5.6% and
1.1%, respectively. Excessive refinement of the grids seemed to provide no improvement to the
Entropy 2019, 21, 16 6 of 16

Grid sensitivity analysis was performed to check the independence of grids. The variations in
pressure drop from inlet to outlet of the heat sink were chosen for the evaluation of grid dependence.
Three non-uniform grid systems with denser grid clustering near the wall were employed with 0.204,
0.412, and 0.628 million grids, respectively. The deviations in pressure drop using 0.204 and 0.412
million grids as compared to that using 0.628 million grids were 5.6% and 1.1%, respectively. Excessive
refinement of the grids seemed to provide no improvement to the computational results. Therefore,
to save computing time and to maintain accuracy of the computed results, a grid number of 0.628
million was finally chosen for the simulations.

3.4. Data Reduction


When the temperature and pressure drop results are obtained by numerical simulation, they
could be processed to calculate the thermal and hydraulic characteristics.
The total thermal resistance of microchannel heat sinks is defined as [9,10]

Tmax − Tin
Rt = (5)
q

where Tmax is the maximum temperature of the microchannel base, Tin is the inlet coolant temperature,
and q is the heat flow carried away by the coolant.
Pumping power of the microchannel heat sink, P, is defined as

P = ∆pu · uu · Ac,u + ∆pd · ud · Ac,b (6)

where ∆p is the pressure drop of the coolant between the microchannel inlet and outlet of the upper
and lower layers,u is the inlet flow velocity, and Ac is the cross-sectional area of a microchannel.
The subscripts “u” and “d” represent the upper and down/lower channels of DL-MCHS, respectively.
The maximum temperature rise of ∆Tw,max is determined by the temperature difference between
the maximum and minimum wall temperatures in the microchannels. It is utilized to assess the wall
temperature uniformity.

3.5. Validation of Models and Methods


To validate the aforementioned numerical models and methods, a simulation of SL-MCHS with
triangular microchannels was first conducted. The obtained fluid temperature differences from
microchannel inlet to outlet with different inlet velocities were compared with the theoretical values
predicted by the energy balance equations using [25]
q
∆T = Tout − Tin = . (7)
ρc p V

It was found that the maximum relative deviation between the numerical results and theoretical
predictions in the simulated velocity range was less than 3.5%. This suggests that the present numerical
models and methods are reliable for the determination of the thermal and hydraulic performance of
microchannel heat sinks.

4. Results and Discussion

4.1. Heat Transfer Enhancement and Pressure Drop Reduction of DL-MCHS


To validate the superiority of DL-MCHS, a DL-MCHS sample with triangular shape (DL-TRI)
was compared with a single-layered microchannel heat sink with the same shape (SL-TRI). As the
flow area of double-layered microchannels was twice that of single-layered microchannels, the inlet
velocity of the DL-TRI was set to be half of that of the corresponding single-layered microchannels at
the same volumetric flow rate, i.e., 0.0625 m/s for DL-TRI, and 0.125 m/s for SL-TRI. Figure 4a shows
SL-TRI to 7 ℃ in DL-TRI. A much more uniform wall temperature distribution in the double layer
layout of microchannels helps to reduce thermal stresses caused by the temperature difference and
facilitates improvement of the reliability of microelectronic devices.
From Figure 4a, it can be noted that the wall temperature of SL-TRI increased monotonically
from the microchannel inlet to the outlet, which can also be seen in the temperature contours in
Entropy 2019, 21, 16 7 of 16
Figure 5a. This can be related to the deterioration of heat exchange between the wall and fluid with
increasing fluid temperature along the flow length. Conversely, for the DL-TRI, the wall
temperature
the first increased
wall temperature along
distributions thethe
along axial
axialflow
flowlength in both
length for the upstream region,
the upper layer andreached the
lower layer
maximum
of the DL-TRI wall temperature
and in the It
SL-TRI samples. middle
is cleartothat
downstream
the DL-MCHS region of the channels,
decreased and then tended
the wall temperature rise
to decrease in
significantly andthepresented
downstream muchregion. This canwall
more uniform be also seen in the
temperature temperature
distributions contours
along of DL-TRI
the stream-wise
in FigureThe
direction. 5b.maximum
The counter-flow cooling
wall temperature effects
of the of the coolant
microchannel from the
bottom surface inlet of
decreased the358
from upper
K in
microchannels
SL-TRI to 340–341 of Kthe DL-TRI In
in DL-TRI. contributed
addition, the to maximum
decreasingwallwall temperatures
temperature in thewas
difference downstream
decreased
region.
by 72% Thermal
for DL-TRI,resistance
that is,was also reduced
it decreased fromconsiderably
24 °C in SL-TRIfrom to0.41 × 10
7 °C K/(W/m2A
in−3DL-TRI. ) in SL-TRI
much more to
0.3 × 10 wall
uniform −3 K/(W/m 2 ) in DL-TRI,
temperature as shown
distribution in in
theFigure
double 4b.layer
The layout
double-layered microchannel
of microchannels helps heat sinks
to reduce
presented
thermal a 27%caused
stresses reduction in thermal
by the resistance,
temperature which
difference andisfacilitates
promisingimprovement
for the efficient cooling
of the of high
reliability of
heat flux microelectronic
microelectronic devices. devices.

(a) (b)

Figure 4. Comparison of wall temperature distribution and thermal resistance between SL-TRI and
DL-TRI with the same triangular shape: (a) wall temperature distribution of the microchannel bottom
surface along the axial flow direction; (b) thermal resistance.

From Figure 4a, it can be noted that the wall temperature of SL-TRI increased monotonically from
the microchannel inlet to the outlet, which can also be seen in the temperature contours in Figure 5a.
This can be related to the deterioration of heat exchange between the wall and fluid with increasing
fluid temperature along the flow length. Conversely, for the DL-TRI, the wall temperature first
increased along the axial flow length in the upstream region, reached the maximum wall temperature
in the middle to downstream region of the channels, and then tended to decrease in the downstream
region. This can be also seen in the temperature contours of DL-TRI in Figure 5b. The counter-flow
cooling effects of the coolant from the inlet of the upper microchannels of the DL-TRI contributed
to decreasing wall temperatures in the downstream region. Thermal resistance was also reduced
considerably from 0.41 × 10−3 K/(W/m2 ) in SL-TRI to 0.3 × 10−3 K/(W/m2 ) in DL-TRI, as shown
in Figure 4b. The double-layered microchannel heat sinks presented a 27% reduction in thermal
resistance, which is promising for the efficient cooling of high heat flux microelectronic devices.
Figure 6 shows the comparison of pressure drop characteristics of SL-TRI and both the upper and
lower layers of DL-TRI. Despite the same volumetric flow rate being applied for both microchannel
heat sinks, and the inlet flow velocity of SL-TRI being equal to the sum of both the upper and lower
layers of DL-TRI, the SL-TRI still presented a larger ∆P than the sum of both the upper and lower layers
of DL-TRI (122.2 Pa). The above results are consistent with those of previous reports [6,12], suggesting
that the DL-TRI far outperformed its single-layered counterpart in the reduction of both pressure
drop and pumping power. In this regard, the double-layered microchannel heat sink showed its
merits in both heat transfer enhancement and pressure drop reduction compared to the single-layered
microchannel heat sink.
Entropy 2018, 20, x FOR PEER REVIEW 8 of 17

Figure 4. Comparison of wall temperature distribution and thermal resistance between SL-TRI and
DL-TRI with the same triangular shape: (a) wall temperature distribution of the microchannel
bottom
Entropy 2019, surface along the axial flow direction; (b) thermal resistance.
21, 16 8 of 16

(a)

(b)
Entropy Figure
2018,
Figure 5. x5.
20, Temperature
FOR contours
PEER REVIEW
Temperature contours of of
thethe middle
middle plane
plane along
along thethe axial
axial flow
flow direction
direction of SL-TRI
of SL-TRI andand
9 of 17
DL-TRI
DL-TRI with
with thethe same
same triangular
triangular shape:
shape: (a) (a) SL-TRI;
SL-TRI; (b)(b) DL-TRI.
DL-TRI.

Figure 6 shows the comparison of pressure drop characteristics of SL-TRI and both the upper
and lower layers of DL-TRI. Despite the same volumetric flow rate being applied for both
microchannel heat sinks, and the inlet flow velocity of SL-TRI being equal to the sum of both the
upper and lower layers of DL-TRI, the SL-TRI still presented a larger △P than the sum of both the
upper and lower layers of DL-TRI (122.2 Pa). The above results are consistent with those of
previous reports [6,12], suggesting that the DL-TRI far outperformed its single-layered counterpart
in the reduction of both pressure drop and pumping power. In this regard, the double-layered
microchannel heat sink showed its merits in both heat transfer enhancement and pressure drop
reduction compared to the single-layered microchannel heat sink.

Figure 6. Comparison of pressure


pressure drop
drop between
between SL-TRI
SL-TRI and
and DL-TRI
DL-TRI with
with the
the same
same triangular
triangularshape.
shape.

4.2. Effects of Microchannel Shape on the Performance of DL-MCHS

4.2.1. Thermal Characteristics


The wall temperature distributions along the axial flow length for all five DL-MCHS with
different cross-sectional shapes are shown in Figures 7 and 8 at flow velocities of 0.0625 m/s and
0.125 m/s, respectively. All DL-MCHS samples first presented an increase in the Tw, then reached
the plateau in the middle to downstream region of the channels, and then tended to decrease in the
Figure 6. Comparison of pressure drop between SL-TRI and DL-TRI with the same triangular shape.
Entropy 2019, 21, 16 9 of 16

4.2. Effects of Microchannel Shape on the Performance of DL-MCHS


4.2. Effects of Microchannel Shape on the Performance of DL-MCHS
4.2.1. Thermal Characteristics
4.2.1. Thermal Characteristics
The wall temperature distributions along the axial flow length for all five DL-MCHS with
The wall
different temperatureshapes
cross-sectional distributions
are shown along in
theFigures
axial flow length
7 and 8 atforflow
all five DL-MCHS
velocities with different
of 0.0625 m/s and
cross-sectional shapes areAll
0.125 m/s, respectively. shown in Figures
DL-MCHS 7 andfirst
samples 8 atpresented
flow velocities of 0.0625
an increase m/sTand
in the 0.125
w, then m/s,
reached
respectively. All DL-MCHS samples first presented
the plateau in the middle to downstream region of the channels, and w an increase in the T , then reached the plateau
then tended to decrease in the in
the middle to downstream
downstream region. For both regioninlet
of the channels,
velocity anditthen
cases, tended
is clear to decrease
that the DL-TRI in the downstream
with triangular
region.
microchannels and the DL-REC with rectangular microchannels showed notably lower and
For both inlet velocity cases, it is clear that the DL-TRI with triangular microchannels wall
the DL-REC with
temperatures rectangular
in both the upper microchannels
and lower layers showed notably lowerbases
of microchannel wallin temperatures
comparison in to both the
the other
upper
three and lower These
samples. layers two
of microchannel
samples were bases
of in comparison
large to the other
microchannel heightthree samples.
(about 1.2 mm),Thesewhich
two
samples
inducedwere of large
the small microchannel
thickness of the height
bottom(about
wall, as 1.2shown
mm), which
in Tableinduced
1. As the small heat
uniform thickness
fluxesofwere
the
bottom
imposed wall,
on as theshown
bottom in Table
surface1. As uniform
of the heat fluxes
samples, heat was weretransported
imposed onvertically
the bottom in surface
the upperof
the samples,
direction byheatheatwas transported
conduction. The vertically
DL-TRI inand theDL-REC
upper direction
with smallby heat conduction.
thicknesses The DL-TRI
of bottom walls
and DL-REC
facilitated thewith small thicknesses
reduction of the lengthofofbottom walls facilitated
heat conduction, the reduction
and thus reduced the of conducted
the length of heat
thermal
conduction,
resistance in and thus
the reduced
solid basesthe [9].conducted
The heatthermal resistance much
was dissipated in the solid
quickerbasesby [9].
theThe heatinwas
fluid the
dissipated much quicker by the fluid in the microchannels, and
microchannels, and smaller wall temperatures can be noted. This can be also seen in the smaller wall temperatures can be
noted. This can
temperature be also seen
contours of theinDL-MCHS
the temperature samplescontours of the
in Figures DL-MCHS
9 and 10. samples in Figures 9 and 10.

(a) (b)
Figure
Figure
Entropy 7.7. xComparison
2018, 20, Comparison of wall
of wall temperature
FOR PEER REVIEW temperature distributions
distributions along
along the
the axial
axial flow
flow length
length for
for all
all five
five
10 of 17
DL-MCHSswith
DL-MCHSs withflow
flowvelocities
velocitiesofof0.0625m/s:
0.0625m/s: (a)
(a) upper
upper layer;
layer; (b)
(b) lower
lower layer.
layer.

(a) (b)
Figure 8.8. Comparison of wall temperature distributions along the axial flow
Figure flow length
length of
of all
all five
five
DL-MCHSswith
DL-MCHSs withflow
flowvelocities
velocitiesof
of0.125m/s:
0.125m/s: (a) upper layer; (b) lower layer.
layer.
(a) (b)
Figure 8. Comparison of wall temperature distributions along the axial flow length of all five
Entropy 2019, 21, 16 10 of 16
DL-MCHSs with flow velocities of 0.125m/s: (a) upper layer; (b) lower layer.

Entropy 2018, 20, x FOR PEER REVIEW 11 of 17

Table 2 lists the maximum wall temperature differences along the flow length for DL-REE.
It was found that DL-CIR and DL-TRA presented the worst wall temperature uniformity, followed
by DL-TRI and DL-REE. The DL-REC with rectangular microchannels presented the most uniform
temperature distributions, indicating that it is more suitable for reducing thermal stresses in the
cooling of microelectronic devices. The total thermal resistance of the five DL-MCHS samples is
shown in Figure 12. For both flow velocities cases, the DL-TRA with trapezoidal microchannels
showed the largest thermal resistance, followed by DL-CIR and DL-REE. This can be related to the
fact that these three samples presented large wall temperatures as discussed above. The DL-REC
with rectangular microchannels showed the smallest thermal resistance. The rectangular
microchannels with flat bottom walls and large channel height seemed to help in the heat transfer
process of coolants and presented the best overall thermal performance.
Figure
Figure 9. Temperaturecontours
9. Temperature contours in the
in the middle
middle planeplane
along along theflow
the axial axial flow direction
direction of all five
of all five DL-MCHSs.
DL-MCHSs.

For the other three DL-CIR, DL-TRA, and DL-REE samples, the Tw of DL-CIR was close to
those of the DL-TRA in general. The DL-REE with reentrant Ω-shaped microchannels showed the
highest wall temperatures in the upstream to middle regions. Nevertheless, the DL-REE presented a
more rapid decrease in the Tw in the downstream region, which induced lower wall temperatures
than for DL-CIR and DL-TRA in this region. For the DL-REE with reentrant Ω-shaped
microchannels, as the main flow was located inside the large circular portion of the reentrant
cavities, the heat exchange between the hot wall and fluid mainly occurred in this circular cavity
area. Since the vertical distances from the microchannel bottom surface to the circular cavity of
DL-REE were smaller than those of DL-CIR and DL-TRA, the conducted thermal resistance was
reduced for DL-REE. Therefore, more heat was dissipated from the microchannel solid wall to the
fluid in the reentrant Ω-shaped microchannels, and smaller wall temperatures were induced in the
downstream regions. Such behaviors can also be seen in the temperature contours in Figures 9–11.
The DL-CIR and DL-TRA showed more portions of high wall temperature regions in the middle to
downstream
Figure10.
areas, which iscontours
10.Temperature
especially
Temperaturecontours
notable inheight
the microchannel base of the lower layer. For the
Figure ininthe
the middle
middle height plane
plane of microchannels
of microchannels in the
in the lower
lower layerlayer of
of the
DL-REE, however,
the five significantly
DL-MCHSs. fewer portions of high wall temperature regions can be seen only in
five DL-MCHSs.
the middle region.
For the other three DL-CIR, DL-TRA, and DL-REE samples, the Tw of DL-CIR was close to those
of the DL-TRA in general. The DL-REE with reentrant Ω-shaped microchannels showed the highest
(a)
wall temperatures in the upstream to middle regions. Nevertheless, the DL-REE presented a more
rapid decrease in the Tw in the downstream region, which induced lower wall temperatures than for
DL-CIR and DL-TRA in this region. For the DL-REE with reentrant Ω-shaped microchannels, as the
main flow was located inside the large circular portion of the reentrant cavities, the heat exchange
between the hot wall and fluid mainly occurred in this circular cavity area. Since the vertical distances
from the microchannel bottom surface to the circular cavity of DL-REE were smaller than those of
DL-CIR and DL-TRA, the conducted thermal resistance was reduced for DL-REE. Therefore, more heat
was dissipated from the microchannel solid wall to the fluid in the reentrant Ω-shaped microchannels,
and smaller wall temperatures were induced in the downstream regions. Such behaviors can also be
seen in the temperature contours in Figures 9–11. The DL-CIR and DL-TRA showed more portions of
high wall temperature regions in the middle to downstream areas, which is especially notable in the
microchannel base of the lower layer. For the DL-REE, however, significantly fewer portions of high
wall temperature regions can be seen only in the middle region.
Entropy 2019, 21, 16 11 of 16
Entropy 2018, 20, x FOR PEER REVIEW 12 of 17

Entropy 2018, 20, x FOR PEER REVIEW (b) (c) 12 of 17


(a)

(d) (e)
(c)
(b) (z = 20, 22.5, and 25 mm) in the middle
Figure 11. Temperature (a)
contours in three cross-sectional planes
regions
Figure of
11.five DL-MCHSscontours
Temperature with flowin velocities of 0.125m/s:
three cross-sectional (a) DL-TRI;
planes (z = 20,(b) DL-REC;
22.5, and 25(c)
mm)DL-CIR;
in the
(d) DL-TRA; and (e) DL-REE.
middle regions of five DL-MCHSs with flow velocities of 0.125m/s: (a) DL-TRI; (b) DL-REC; (c)
DL-CIR; (d) DL-TRA; and (e) DL-REE.
Table 2 lists the maximum wall temperature differences along the flow length for DL-REE.
It was found that DL-CIR and DL-TRA presented the worst wall temperature uniformity, followed
by DL-TRI and DL-REE. The DL-REC with rectangular microchannels presented the most uniform
temperature distributions, indicating that it is more suitable for reducing thermal stresses in the cooling
of microelectronic devices. The total thermal resistance of the five DL-MCHS samples is shown in
(d)
Figure 12. For both flow velocities cases, the DL-TRA with trapezoidal(e) microchannels showed the
largest thermal resistance, followed by DL-CIR and DL-REE. This can be related to the fact that these
three samples presented large wall temperatures as discussed above. The DL-REC with rectangular
Figure 11. Temperature contours in three cross-sectional planes (z = 20, 22.5, and 25 mm) in the
microchannels showed the smallest thermal resistance. The rectangular microchannels with flat bottom
middle regions of five DL-MCHSs with flow velocities of 0.125m/s: (a) DL-TRI; (b) DL-REC; (c)
walls DL-CIR;
and large channel height seemed to help in the heat transfer process of coolants and presented
(d) DL-TRA; and (e) DL-REE.
the best overall thermal performance.

(a) (b)
Figure 12. Thermal resistance of five DL-MCHSs for test cases of: (a) 0.0625 m/s; (b) 0.125 m/s.

Table 2 Temperature results for the five DL-MCHSs.

Maximum Temperature of Maximum Wall


Inlet Velocity
Sample Layer Microchannel Base Temperature Difference
(m/s)
Tmax (K) ΔTw,max (K)
Upper 7.2
0.0625 344
Lower 6.8
DL-TRI
Upper (a) (b) 5.4
0.125 332
Lower 5.2
Figure12.
Figure 12.Thermal
Thermalresistance
resistanceofoffive
fiveDL-MCHSs
DL-MCHSsfor
fortest
testcases
casesof:of:(a)
(a)0.0625
0.0625m/s;
m/s;(b)
(b)0.125
0.125m/s.
m/s.
Upper 5.6
0.0625 342
Lower Table 2 Temperature results for the five DL-MCHSs. 5.3
DL-REC
Upper 3.9
0.125 328
Maximum Temperature of Maximum
Lower Inlet Velocity 3.6 Wall
Sample Layer Microchannel Base Temperature Difference
(m/s)
Tmax (K) ΔTw,max (K)
Upper 7.2
0.0625 344
Lower 6.8
Entropy 2019, 21, 16 12 of 16

Table 2. Temperature results for the five DL-MCHSs.

Maximum Temperature Maximum Wall


Inlet Velocity
Sample Layer of Microchannel Base Temperature Difference
(m/s)
T max (K) ∆T w,max (K)
Upper 7.2
0.0625 344
Lower 6.8
DL-TRI
Upper 5.4
0.125 332
Lower 5.2
Entropy 2018, 20, x FOR PEER REVIEW
Upper 5.6 13 of 17
0.0625 342
Lower 5.3
DL-REC
Upper Upper 3.9 8.2
0.0625 0.125 348 328
Lower 3.6
Lower 8.1
DL-CIR Upper 8.2 5.9
Upper 0.0625
0.125
Lower 333 348 8.1
Lower
DL-CIR 5.8
Upper 5.9
Upper 0.125 333
Lower
0.0625 356 5.8 8.0
Lower 8.4
DL-TRA Upper 8.0
Upper 0.0625 356
Lower
0.125 339 8.4 5.6
DL-TRA
Lower Upper 5.6 5.8
0.125 339
Upper Lower 5.8 5.9
0.0625 350
Lower Upper 5.9 5.6
DL-REE Lower
0.0625 350
5.6 4.2
Upper
DL-REE 0.125
Upper 333 4.2
Lower 0.125 333 4
Lower 4
SL-TRI - 0.125 358 24.3
SL-TRI - 0.125 358 24.3

4.2.2. Pressure Drop and Pumping Power


4.2.2. Pressure Drop and Pumping Power
The pressure drop characteristics of both layers in the five double-layered microchannel heat
sinksThearepressure
shown in drop characteristics
Figure 13a,b. Theofupper
both layers
layer in the five double-layered
presented larger pressuremicrochannel
drops than theheatlower
sinks
are shown in Figure 13a,b. The upper layer presented larger pressure drops
layer for all five DL-MCHSs. This was especially notable in the DL-TRA with trapezoidal than the lower layer for all
five DL-MCHSs. This was especially notable in the DL-TRA with trapezoidal microchannels.
microchannels. As the microchannel wall and fluid temperatures in the bottom base were larger As the
microchannel
than wall and
in the upper fluid
layer, thetemperatures in the
fluid viscosity in bottom
the lowerbase weredecreased
layer larger thanwith
in the upper layer,
elevation the
of fluid
fluid viscosity in the lower layer decreased with elevation of fluid temperatures. This
temperatures. This resulted in the reduction of viscous forces and thus induced a smaller pressure resulted in the
reduction
drop in theoflower
viscous forces and thus induced a smaller pressure drop in the lower layer.
layer.

(a) (b)
Figure13.
Figure 13. Pressure
Pressuredrop
dropof
offive
fiveDL-MCHSs
DL-MCHSsin
inthe
thetest
testcases
casesof:
of:(a)
(a)0.0625
0.0625m/s;
m/s; (b) 0.125m/s.
0.125m/s.

considering the sum of total pressure drops in both layers of the double-layered
When considering
microchannel heat sinks, it can be found that the DL-TRA
DL-TRA with
with trapezoidal
trapezoidal microchannels induced
the largest pressure drop, followed by DL-REE, DL-CIR, and DL-REC. The DL-TRI with triangular
microchannels definitively presented the smallest pressure drop. The above trend can be related to
the flow morphologies inside the microchannels with different shapes. Figure 14 illustrates the
velocity streamlines in the cross-sectional plane of the outlet of the five DL-MCHSs, and Figure 15
shows the velocity streamlines along the middle x plane in the streamwise directions in the
Entropy 2019, 21, 16 13 of 16

the largest pressure drop, followed by DL-REE, DL-CIR, and DL-REC. The DL-TRI with triangular
microchannels definitively presented the smallest pressure drop. The above trend can be related
to the flow morphologies inside the microchannels with different shapes. Figure 14 illustrates the
velocity streamlines in the cross-sectional plane of the outlet of the five DL-MCHSs, and Figure 15
shows the velocity streamlines along the middle x plane in the streamwise directions in the entrance
regions (z = 0–5 mm). Since the DL-TRA and DL-CIR featured large width and small height in the
flow passages, they played a confinement role in the development of flow streamlines in the vertical
direction. The velocity magnitude quickly decreased in the vertical direction from the core region
to the boundary
Entropy layer
2018, 20, x FOR PEERdue to the action of the molecular viscous force drag [26]. Therefore, due
REVIEW 14 ofto
17
the confinement of the trapezoidal and circular configurations, a large pressure drop can be obtained
from the DL-TRA
microchannels, and DL-CIR.
given For the DL-REE
that the unique reentrantwith reentrant-shaped
Ω-shaped microchannelsmicrochannels, given that
featured narrow slotstheat
unique and largeΩ-shaped
the topreentrant circular microchannels
cavities below,featured
the abruptnarrow slots at the
protrusion top and
between thelarge circular
circular cavities
cavity and
below,
verticalthe abrupt
slot protrusion
interrupted between the
continuous flowcircular cavity and
streamlines in the vertical
crossslot interrupted
section continuous
and exerted flow
throttling
streamlines
effects on the in fluid
the cross
flow.section and exerted
This accelerated thethrottling
main flow effects
in the oncircular
the fluid flow. This
cavities [22] accelerated
but slowed thethe
main flow in the circular cavities [22] but slowed the flow in the narrow slot,
flow in the narrow slot, as shown in Figure 13e. Large friction flow resistance was thus induced. as shown in Figure 13e.
Large friction
Therefore, theflow resistance
DL-REE with was thus induced.
reentrant Therefore,
microchannels alsothe DL-REE awith
presented largereentrant
pressure microchannels
drop. When
also presented
comparing thea large
DL-TRIpressure
withdrop.
DL-REC,Whenthecomparing
DL-REC thefeatured
DL-TRI with more DL-REC,
cornersthethan DL-REC
the featured
DL-TRI.
more corners than the DL-TRI. Moreover, the spanwise extension
Moreover, the spanwise extension of the flow streamlines in the rectangular microchannels of the flow streamlines in was
the
rectangular
restricted bymicrochannels
its two vertical was restricted
walls, by itsin
as shown two vertical
Figure 13b.walls,
The as shown
flow in Figure
friction 13b. was
resistance The flow
thus
friction
enhanced, resistance
and a was thuspressure
larger enhanced, and can
drop a larger pressure
be noted fordrop can be The
DL-REC. notedDL-TRI
for DL-REC.
with The DL-TRI
rectangular
with rectangularmay
microchannels microchannels
reach a good may reach awithout
balance good balance
notable without
channelnotable channel in
confinement confinement in both
both vertical and
vertical
spanwise and spanwisewhich
directions, directions,
wouldwhich would for
be helpful be helpful
the fluid forflow
the with
fluid aflow
smallwith a smalldrop.
pressure pressure drop.

Figure
Figure 14. Velocity streamlines in the cross-sectional planes of outlets in the lower layer of five
14. Velocity five
DL-MCHSs
DL-MCHSs with aa flow flowvelocity
velocityofof0.125
0.125m/s:
m/s:(a) (a) DL-TRI;
DL-TRI; (b) (b) DL-REC;
DL-REC; (c) DL-CIR;
(c) DL-CIR; (d) DL-TRA;
(d) DL-TRA; and
and (e) DL-REE.
(e) DL-REE.

When pumping power P, that is, the integration of pressure drop with cross-sectional area,
is considered for the five DL-MCHSs with the same inlet velocity, it can be noted from Figure 16
that the DL-TRA with trapezoidal microchannels induced the largest pumping power consumption.
For the other DL-MCHSs, the orders of P were just the opposite to those of the pressure drop, i.e.,
the smallest pumping power needed was for the DL-REE with reentrant microchannels, followed
by DL-CIR, DL-REC, and DL-TRI. This trend, however, is consistent with the cross-sectional areas,
as shown in Table 1. This indicates that cross-sectional area may play a more significant role in the
total pumping power able to drive the fluid flow in double-layered microchannel heat sinks.

Figure 15. Velocity vectors in the middle x plane of five DL-MCHSs in the entrance regions
Figure 14. Velocity streamlines in the cross-sectional planes of outlets in the lower layer of five
DL-MCHSs with a flow velocity of 0.125 m/s: (a) DL-TRI; (b) DL-REC; (c) DL-CIR; (d) DL-TRA; and
Entropy(e)
2019, 21, 16
DL-REE. 14 of 16

Entropy 2018, 20, x FOR PEER REVIEW 15 of 17

pressure drop, i.e., the smallest pumping power needed was for the DL-REE with reentrant
microchannels, followed by DL-CIR, DL-REC, and DL-TRI. This trend, however, is consistent with
the cross-sectional areas, as shown in Table 1. This indicates that cross-sectional area may play a
moreFigure
significant
Figure 15.15. role vectors
in vectors
Velocity
Velocity theintotal pumping
theinmiddle
the middle power
x plane fiveable
xofplane of to
fivedrive
DL-MCHSs theentrance
DL-MCHSs
in the fluidin flow
the in (zdouble-layered
entrance
regions regions
= 0–5 mm)
microchannel
(z = 0–5 heat
mm) sinks.
with a flow velocity of 0.125 m/s: (a) DL-TRI; (b) DL-REC; (c) DL-CIR;
with a flow velocity of 0.125 m/s: (a) DL-TRI; (b) DL-REC; (c) DL-CIR; (d) DL-TRA; and (e) DL-REE.(d) DL-TRA;
and (e) DL-REE.

When pumping power P, that is, the integration of pressure drop with cross-sectional area, is
considered for the five DL-MCHSs with the same inlet velocity, it can be noted from
Figure 16 that the DL-TRA with trapezoidal microchannels induced the largest pumping power
consumption. For the other DL-MCHSs, the orders of P were just the opposite to those of the

(a) (b)
Figure16.
Figure 16.Pumping
Pumpingpower
powerof
ofthe
thefive
fiveDL-MCHSs
DL-MCHSsininthe
thetest
testcases
casesof:
of:(a)
(a)0.0625
0.0625m/s;
m/s; (b) 0.125 m/s.
m/s.

4.2.3.
4.2.3. Overall
Overall Performance
performance Evaluation
evaluation
When
When considering
considering both
both thermal
thermal and
and hydraulic
hydraulic characteristics
characteristics of
of the
the DL-MCHS,
DL-MCHS, it it can
can bebe noted
noted
that the DL-TRA with trapezoidal microchannels performed the worst in
that the DL-TRA with trapezoidal microchannels performed the worst in terms of thermalterms of thermal resistance,
pressure
resistance,drop, and pumping
pressure drop, and power.
pumpingIt should
power.notItbe selected
should notfor
bedouble-layered microchannel
selected for double-layered
heat sinks. On the other hand, the DL-REC with rectangular microchannels
microchannel heat sinks. On the other hand, the DL-REC with rectangular microchannels presented the smallest
thermal resistance and the best wall temperature uniformity. It also presented
presented the smallest thermal resistance and the best wall temperature uniformity. It alsosmall wall temperatures
along the flow
presented smalldirection. As such, italong
wall temperatures showedthe the
flowbest overall thermal
direction. As such,performance
it showed the outbest
of the five
overall
DL-MCHSs. The DL-TRI with triangular microchannels also showed fairly good
thermal performance out of the five DL-MCHSs. The DL-TRI with triangular microchannels also thermal performance
with
showedlowfairly
wall temperatures
good thermaland small thermal
performance withresistance. In addition, these
low wall temperatures two DL-MCHSs
and small showed
thermal resistance.
small pressure drops. Nevertheless, when pumping power is considered,
In addition, these two DL-MCHSs showed small pressure drops. Nevertheless, when pumping these two DL-MCHSs
are no longer
power superior.these
is considered, The DL-REE
two DL-MCHSs are noΩ-shaped
with reentrant microchannels
longer superior. exhibited
The DL-REE withthereentrant
smallest
pumping
Ω-shapedpower. Therefore, exhibited
microchannels the optimum theDL-MCHS
smallest is dependent
pumping on which
power. issue is considered
Therefore, the optimum the
most important. When thermal performance is the prime concern,
DL-MCHS is dependent on which issue is considered the most important. When thermalthe DL-REC with rectangular
microchannels
performance isseems to beconcern,
the prime the optimum choice. with
the DL-REC Whenrectangular
pumping power consumption
microchannels seemsis tothebemost
the
important
optimum consideration,
choice. When the DL-REEpower
pumping with reentrant Ω-shaped
consumption is themicrochannels
most important should be selected as
consideration, a
the
double-layered microchannel heat sink.
DL-REE with reentrant Ω-shaped microchannels should be selected as a double-layered
microchannel heat sink.

5. Conclusions
This work has presented a computational study of the thermal and hydraulic performance of
double-layered microchannel heat sinks with different cross-sectional shapes. The superiority of
Entropy 2019, 21, 16 15 of 16

5. Conclusions
This work has presented a computational study of the thermal and hydraulic performance of
double-layered microchannel heat sinks with different cross-sectional shapes. The superiority of
double-layered microchannel heat sinks has also been assessed via a comparison of a single-layered
and double-layered microchannel heat sink with triangular microchannels. Compared to the SL-TRI,
the DL-TRI decreased wall temperatures considerably and presented a 27% reduction in thermal
resistance. It also induced much more uniform wall temperature distribution, and significantly reduced
the pressure drop and pumping power. For the five DL-MCHSs with different cross-sectional shapes,
the DL-TRA with trapezoidal microchannels performed the worst in terms of thermal resistance,
pressure drop, and pumping power. The DL-TRI with triangular microchannels and the DL-REC
with rectangular microchannels presented small wall temperatures and thermal resistance, which
are promising for maintaining high cooling efficiency. The DL-REC showed the most uniform
temperature distribution and presented the best overall thermal performance. The DL-TRI and
DL-REC also outperformed others in the reduction of pressure drop. Nevertheless, when pumping
power is considered, the DL-REE with reentrant Ω-shaped microchannels performed the best.
As such, the DL-REC with rectangular microchannels seems to be the optimum choice when thermal
performance is the prime concern, whereas the DL-REE with reentrant Ω-shaped microchannels should
be selected when pumping power consumption is considered to be the most important issue.

Author Contributions: D.D. finished the ideation and manuscript. G.P. and W.Z. conducted the numerical
simulations. G.P. accomplished all the Figures. P.W. and T.F. are responsible for checking and revising.
Funding: This research was funded by the Science and Technology Planning Project of Guangdong Province, China
(No. 2017A010104002), the open fund of Tianjin Key Laboratory for Civil Aircraft Airworthiness and Maintenance
Airworthiness and Maintenance, and the open fund of Hubei Key Laboratory Mechanical Transmission and
Manufacturing Engineering in the Wuhan University of Science and Technology (No. 2017A01).
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Mahajan, R.; Chia-pin, C.; Chrysler, G. Cooling a microprocessor chip. Proc. IEEE 2006, 94, 1476–1486.
[CrossRef]
2. Tuckerman, D.B.; Pease, R.F.W. High-performance heat sinking for VlSI. IEEE Electron. Device Lett. 1981, 2,
126–129. [CrossRef]
3. Deng, D.; Wan, W.; Tang, Y.; Wan, Z.; Liang, D. Experimental investigations on flow boiling performance of
reentrant and rectangular microchannels—A comparative study. Int. J. Heat Mass Transf. 2015, 82, 435–446.
[CrossRef]
4. Li, P.; Xie, Y.; Zhang, D.; Xie, G. Heat Transfer Enhancement and Entropy Generation of Nanofluids Laminar
Convection in Microchannels with Flow Control Devices. Entropy 2016, 18, 134. [CrossRef]
5. Deng, D.; Wan, W.; Qin, Y.; Zhang, J.; Chu, X. Flow boiling enhancement of structured microchannels with
micro pin fins. Int. J. Heat Mass Transf. 2017, 105, 338–349. [CrossRef]
6. Vafai, K.; Zhu, L. Analysis of two-layered micro-channel heat sink concept in electronic cooling. Int. J. Heat
Mass Transf. 1999, 31, 1176–1186. [CrossRef]
7. Chong, S.H.; Ooi, K.T.; Wong, T.N. Optimization of single and double layer counter flow microchannel heat
sinks. Appl. Therm. Eng. 2002, 22, 1569–1585. [CrossRef]
8. Cheng, Y.J. Numerical simulation of stacked microchannel heat sink with mixing-enhanced passive structure.
Int. Commun. Heat Mass Transf. 2007, 34, 295–303. [CrossRef]
9. Wei, X.J.; Joshi, Y.; Patterson, M.K. Experimental and numerical study of a stacked microchannel heat sink
for liquid cooling of microelectronic devices. ASME J. Heat Transf. 2007, 129, 1432–1444. [CrossRef]
10. Hung, T.C.; Yan, W.M.; Li, W.P. Analysis of heat transfer characteristics of doublelayered layered
microchannel heat sink. Int. J. Heat Mass Transf. 2012, 55, 3090–3099. [CrossRef]
11. Hung, T.C.; Yan, W.M.; Wang, X.D.; Huang, Y.X. Optimal design of geometric parameters of double-layered
microchannel heat sinks. Int. J. Heat Mass Transf. 2012, 55, 3262–3272. [CrossRef]
Entropy 2019, 21, 16 16 of 16

12. Xie, G.N.; Liu, Y.Q.; Sunden, B.; Zhang, W.H. Computational study and optimization of laminar heat transfer
and pressure loss of double-layer microchannels for chip liquid cooling. ASME J. Therm. Sci. Eng. Appl. 2013,
5, 011004. [CrossRef]
13. Lin, L.; Chen, Y.Y.; Zhang, X.X.; Wang, X.D. Optimization of geometry and flow rate distribution for
double-layer microchannel heat sink. Int. J. Therm. Sci. 2014, 78, 158–168. [CrossRef]
14. Leng, C.; Wang, X.D.; Wang, T.H.; Yan, W.M. Multi-parameter optimization of flow and heat transfer for a
novel double-layered microchannel heat sink. Int. J. Heat Mass Transf. 2015, 84, 359–369. [CrossRef]
15. Jing, D.; He, L. Thermal Characteristics of Staggered Double-Layer Microchannel Heat Sink. Entropy 2018,
20, 537. [CrossRef]
16. Kulkarni, K.; Afzal, A.; Kim, K. Multi-objective optimization of a double-layered microchannel heat sink
with temperature-dependent fluid properties. Appl. Therm. Eng. 2016, 99, 262–272. [CrossRef]
17. Leng, C.; Wang, X.D.; Wang, T.H. An improved design of double-layered microchannel heat sink with
truncated top channels. Appl. Therm. Eng. 2015, 79, 54–62. [CrossRef]
18. Wang, S.; Li, X.; Wang, X.; Lu, G. Flow and heat transfer characteristics in double-layered microchannel heat
sinks with porous fins. Int. J. Heat Mass Transf. 2018, 93, 41–47. [CrossRef]
19. Osanloo, B.; Mohammadi-Ahmar, A.; Solati, A.; Baghani, M. Performance enhancement of the double-layered
micro-channel heat sink by use of tapered channels. Appl. Therm. Eng. 2016, 102, 1345–1354. [CrossRef]
20. Wong, K.C.; Ang, M.L. Thermal hydraulic performance of a double-layer microchannel heat sink with
channel contraction. Int. Commun. Heat Mass Transf. 2017, 81, 269–275. [CrossRef]
21. Zhai, Y.L.; Xia, G.D.; Liu, X.F.; Wang, J. Characteristics of entropy generation and heattransfer in
double-layered micro heat sinks with complex structure. Energy Convers. Manag. 2015, 103, 477–486.
[CrossRef]
22. Sharma, D.; Singh, P.P.; Garg, H. Numerical analysis of trapezoidal shape double layer microchannel Heat
Sink. Int. J. Mech. Ind. Eng. 2013, 3, 10–15.
23. Kulkarni, K.; Khan, A.A.; Kim, K. Performance analysis of double-layer microchannel heat sink with various
microchannel shapes. Heat Transf. Res. 2018, 49, 349–368. [CrossRef]
24. Deng, D.; Wan, W.; Tang, Y.; Shao, H.; Huang, Y. Experimental and numerical study of thermal enhancement
in reentrant copper microchannels. Int. J. Heat Mass Transf. 2015, 91, 656–670. [CrossRef]
25. Wu, J.M.; Zhao, J.Y.; Tseng, K.J. Parametric study on the performance of double-layered microchannels heat
sink. Energy Convers. Manag. 2014, 80, 550–560. [CrossRef]
26. Zhang, S.; Cui, N.; Xiong, Y.; Feng, Y.; Qin, J.; Bao, W. Effect of channel aspect ratio on chemical recuperation
process in advanced aeroengines. Energy 2017, 123, 9–19. [CrossRef]

© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

Das könnte Ihnen auch gefallen