You are on page 1of 25

International Journal of Heat and Mass Transfer 118 (2018) 129–153

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


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

Review

Optimization of thermal design of heat sinks: A review


Hamdi E. Ahmed a,⇑, B.H. Salman b,c, A.Sh. Kherbeet d, M.I. Ahmed e
a
Department of Mechanical Engineering, University of Anbar, Ramadi 31001, Iraq
b
DNV GL, Engineering Department, Las Vegas, NV 89146, USA
c
Department of Mechanical Engineering, University of Nevada, Las Vegas, NV 89154, USA
d
Department of Mechanical Engineering, KBU International College, 47800 Petaling Jaya, Selangor, Malaysia
e
Department of Mechanical Engineering, Faculty of Engineering, Islamic University in Medinah, Prince Naif Ibn Abdulaziz, Al Jamiah, Medinah 42351, Saudi Arabia

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

Article history: Heat sinks are a kind of heat exchangers used for cooling the electronic devices due to the simplicity of
Received 24 July 2017 fabrication, low cost, and reliability of heat dissipation. The extended surfaces from the heat sinks are
Received in revised form 20 October 2017 either flat-plate fins or pins fins shapes. In the last decades, intensive attentions were spent on miniatur-
Accepted 23 October 2017
izing the electronic devices because of the high sophisticated micro- and nano-technology development.
Available online 1 November 2017
But the heat dissipation is still the major problem of enhancing the thermal performance the heat sink. In
this article, a comprehensive review is carried out on the methods used for optimizing the hydrothermal
Keywords:
design of heat sinks. Therefore, available investigations regarding the passive and active techniques uti-
Optimization
Hydrothermal design
lized for enhancing the heat removal from heat sinks by modifying either the solid domain or fluid
Heat sink domain are covered. The purpose of this study is to summarize the investigational efforts spent for devel-
Active and passive technical enhancement oping the thermal performance of the heat sinks, limitations, and unsolved proposed solutions.
methods Ó 2017 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
2. Optimization of heat sinks in natural convection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
3. Optimization of heat sinks in unsteady flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
4. Optimization of heat sink shape . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
5. Location of inlet and outlet arrangement of heat sink. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
6. Rotating heat sinks. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
7. Optimization of the substrate material. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
8. Optimization of heat sinks with boiling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
9. Optimization of flat-plate fin heat sink (FPFHS). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
10. Optimization of pin-fin heat sink (PFHS) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
11. Optimization of heat sink by using porous media . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
12. Optimization of heat sinks by using turbulators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
13. Optimization of the shape of single-channel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
14. Optimization of heat sinks by working fluid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
15. Optimization of single- and double-layer heat sink. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
16. Optimization of heat sinks by varying the size . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
16.1. Mini-channel heat sink (MiCHS) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
16.2. Micro channel heat sink (MCHS) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
17. Temperature jump . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
18. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150

⇑ Corresponding author.
E-mail addresses: hamdi.ahmed@uoanbar.edu.iq (H.E. Ahmed), eng.bassam2007@yahoo.com (B.H. Salman).

https://doi.org/10.1016/j.ijheatmasstransfer.2017.10.099
0017-9310/Ó 2017 Elsevier Ltd. All rights reserved.
130 H.E. Ahmed et al. / International Journal of Heat and Mass Transfer 118 (2018) 129–153

Abbreviations

AR aspect ratio Nu Nusselt number


DLMCHS double-layer microchannel heat sink PF pin-fin
FPF flat-plate fin Pr Prandtl number
HS heat sink Ra Rayleigh number
MC micro-channel Re Reynolds number
MiCHS mini-channel heat sink Rth thermal resistance
MT micro-tube SLMCHS single-layer microchannel heat sink
NF Nanofluid

19. Recommendations for future works . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151


Conflict of interest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
Appendix A. Supplementary material . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151

1. Introduction the demand of high heat generated by electronic components [4].


Due to the development of the micro-technology, highly integrated
With the continuous development of electronic devices towards electronic circuits led to increased heat generation rates in elec-
high performance and miniaturization size, heat dissipation prob- tronic chips. In other words, the speed of chips is increasing, the heat
lem has become a major obstacle to their development. Besides, generated by the chip is going up, and the temperature of the chip is
the traditional air cooling method has been unable to meet the rising while the volume of the chip is miniaturized. The average heat
high-density heat dissipation requirement. Computer users prefer flux of the chip was about 50 W/cm2 in 2010 and rose to be around
computers that having high-speed processors. The thermal design 250 W/cm2 in 2012. Porous metal has been demonstrated that it can
optimization of the heat sinks leads to minimize the size and enhance the forced convection heat transfer strongly [5,6]. Heat sink
weight of the heat sink, and then improve the heat removal in con- with a porous medium can increase both the surface contact area
sequently increasing the speed of electronic devices. Electronic between the fins and coolant and the local mixing velocity of the
devices are increasingly miniaturized and the operating power of coolant, which ensures better heat removal. The thermal perfor-
CPU increases. Besides, a larger amount of data processed by the mance of a porous heat sink can be enhanced if porosity conditions
CPU at a time causes greater heat generation. This development and the geometric parameters of the channel are properly designed.
in the computer manufacturing makes the transfer of generated High-pressure drop is carrying out between the inlet and outlet of a
heat to the ambient becomes more difficult [1]. Generally, the heat porous microchannel heat sink (MCHS) having low porosity and per-
generated by the processors is typically transferred to a heat sink meability which needs more pumping power. Therefore, the high-
(HS) by heat conduction, and then to the ambient by natural, pressure drop associated with using porous medium plays a vital
mixed or forced convection. Low efficiency of heat removal of the role in the design of a porous MCHS [7,8].
heat sink possibly causing damage to the electronic component The advantages of micro-channel compact heat exchangers have
as the temperature rises [2]. This problem has motivated the com- gotten plenty of attention of investigators since the last two decades.
puter manufacturers to employ sophisticated technology to The ratio between the contact surface area with the refrigerant and
improve the speed of electronic elements with increasing the heat the heat exchanger volume increases with decreasing the channel
removal. In contrary, the smallest size of the computers increases hydraulic diameter. This characteristic permits minimizing the heat
the overall flow resistance for the system and eventually sup- exchanger size, low amount of material used in the heat exchanger
presses the fluid flow between fins of the heat sink. This signifi- manufacture, and reducing the refrigerant amount. These aspects
cantly influences the fan performance and affects its heat not only impact the fabrication cost but also environmental aspects.
removal capability. Therefore, the heat sink must be designed Flow boiling in mini- and microchannels can be used for cooling
properly to promote heat transfer and to avoid overheating of many high power density devices such as Micro Electro Mechanical
the electronic element. The effective thermal management of heat Systems (MEMS), microprocessors, laser diode arrays and Light
sinks is of priority concern of researchers. It is necessary to be Emitting Diodes (LEDs) [9]. Flow boiling in microchannels is very
mentioned that the common popular coolant of electronic systems effective in the thermal management of high-flux modern electron-
is air due to the ease of obtaining the coolant and the simplicity, ics. To avoid electric hazards of electronic equipment or integrated
high reliability and low cost of the required equipment [3]. circuit component, the dielectric fluids such as fluorocarbon fluids
With rapid developments in microelectronic techniques, the featuring excellent electrical and chemical properties, are the lead-
electronic devices are going to be more miniaturized having high ing candidates for such applications [10].
power, high performance, and higher temperature. The traditional The past two decades have witnessed intense interest among
heat transfer method of forced air convection is reached to its ther- researchers in the use of MCHS, which has been spurred by such
mal limit. Therefore, the challenge is how to develop effective meth- unique attributes as compactness, high power dissipation to vol-
ods for cooling high flux devices. Generally, the working limit ume ratio, and small coolant inventory. Due to their high area-
temperature of electronic devices is ranged from 85 to 100 °C. The to-volume ratios, the use of MCHS has been introduced, first by
literature has shown that the reliability of chip reduces 5% and the Tuckerman and Pease in 1981, as an alternative solution for remov-
life span significantly reduces when the temperature raises every ing high heat fluxes from small areas. The mode of flow in
1 °C above the limit temperature. Therefore, a huge threatens to chip microchannels mostly remains under laminar flow regime due to
reliability and service lifetime if the high heat generated by the elec- the small hydraulic diameter of the microchannel and eventually
tronic element cannot be removed in time. Thus, it is necessary to the pumping power at the micro-scale is still a limiting factor.
investigate and develop an effective cooling technology to meet One of the most common means for cooling electronic modules
H.E. Ahmed et al. / International Journal of Heat and Mass Transfer 118 (2018) 129–153 131

is a finned heat sink. There are a variety of heat sink types, depend-
ing on the fin geometry, configuration and orientation (flat-plate
fin, pin fin, interrupting fin, slotted fins, inline or staggered, same
or different height and width, etc.).
The lower thermal resistance, uniform temperature distribu-
tion, and lower maximum temperature on the base surface, lower
pumping power, higher compactness and lower fabrication cost
are still the essential requirements in MCHS. In recent years, the
single-layered microchannel heat sink (SLMCHS) has been widely
used in various electronic devices as a cooling system. With minia-
turizing the electronic chips, the optimization of the geometric
parameters of SLMCHS is still a hot and attractive research topic
to improve the overall performance [11].
The SLMCHS, firstly introduced by Tuckerman and Pease [1] in
1981. The relatively high and non-uniform temperature distribu-
tion along the channel is still the disadvantage of SLMCHS. This dis-
advantage produces thermal stresses in the chips and then reduces
the electrical performance and eventually their lifetime. The high Fig. 1. Schematic diagram of the orientation of the HS [18].
pressure drop between the inlet and outlet (due to the small
power inverters and electric vehicles reducing poluentes
hydraulic diameter of the channel), and the high temperature vari-
ation (due to the large amount of heat generated by chips cannot Jang et al. [19] investigated the effect of the orientation on the
be removed by the relatively small amount of coolant) restrict natural convection and radiation for a cylindrical HS used to cool
the SLMCHS to be applied in all engineering applications particu- an LED light bulb as shown in Fig. 2. As a result, the stagnation
larly with the development of microtechnology. However, increas- points and flow separation was appeared when the inclination
ing the coolant flow rate for enhancing the heat removal requires angle increased. Thus, the drag coefficient increased when the Nu
more pumping power which means higher noise associated number decreased. The drag coefficient increased steeply with
[12,13]. Therefore, in 1999, Vafai and Zhu [6] proposed a double- increasing the orientation angle by increasing either the number
layered microchannel heat sink (DLMCHS) design (top and bottom of fins or the fin length and eventually the orientation effect was
layers), as an alternative method for reducing the temperature intensified. Furthermore, there was no significant effect for the
variation along the heat sink. The proposed heat sink provides lar- fin height on the orientation effect. They derived a correlation to
ger hydraulic diameter can promise more uniform temperature predict the Nu number around an inclined cylindrical HS.
profile on the heated surface compared to the traditional one.
Therefore, further studies on the hydrothermal performance of
3. Optimization of heat sinks in unsteady flow
multi-layer MCHS design are necessary.
Xu et al. [20] carried out a CFD and experimental study on heat
2. Optimization of heat sinks in natural convection
transfer performance of a symmetrical fractal silicon microchannel
network under a pulsation flow. The results illustrated that the
Tari and Mehrtash [14,15] derived a set of correlations of Nu
heat transfer rate at pulsation frequency (2–10 and 30–40 Hz)
number (Nu) for both upward and downward natural convection
increased by 25–40% and it was higher than that for (10–20 Hz)
from PFHS by using large sets of experimental data from the liter-
when Re number was kept constant. By increasing Re number, a
ature. At small inclinations, it was noted that convection heat
decrease in the enhancement factors was recorded from 40% to
transfer rate remained almost the same. Their correlations covered
5% for the above frequency range, unlike the steady-state cases.
all possible inclination angles with accuracy less than 20%.
Increasing Re number caused a decrease in the wall temperature
Hassan [16] studied the natural convection heat transfer inside
and increase in the Nu number. There was no remarkable differ-
a horizontal and vertical enclosure HS having rectangular fins and
ence in the pressure drop between steady and pulsated flow cases.
filled with Cu–water nanofluid. He observed an increasing in the
Yu et al. [21] studied numerically the effect of agitator plates
Nu number with increasing Rayleigh number (Ra) and nanoparti-
inserted within the channels of a HS to enhance the heat transfer
cles concentration (maximum concentration was 10%) and with
by agitation. The periodic motion in a transverse direction gener-
decreasing diameter of nanoparticles (minimum nanoparticles
ated by the agitator plate enhanced the thermal performance.
diameter was 10 nm). At low Rayleigh number (Ra = 103), the Nu
Maximum heat transfer enhancement obtained was about 61%
number was increased with increasing number of fins and the
by agitation. They monitored a sharp increase in heat transfer on
length to the height ratio (L/H). At high Ra number (Ra = 106), there
the base surface as the tip gap size between the agitator plate
was no great effect for L/H on the Nu number, whereas the smallest
and the channel was decreased. In contrast, there was no effect
nanoparticle diameters and highest concentration covered showed
of tip gap size observed on heat transfer on the sidewalls. Higher
highest Nu number when the HS having one fin and L/H = 0.5 for
heat transfer rate from the channel wall was registered when the
both horizontal and vertical configuration.
amplitude or frequency of the agitator plate increased. It can be
Aminossadati and Ghasemi [17] recorded an increase in the
clearly seen that very limited investigations existed in this area
heat transfer rate of natural convection in a 2D square cavity filled
of research.
with water–CuO nanofluid by increasing Ra number and volume
fraction. Stronger circulating cells were associated with nanofluid
compared to pure water at high Ra numbers. Rectangular fin HS 4. Optimization of heat sink shape
is widely used in LED which encounters thermal challenges when
highly integrated into high power lighting system. Under natural Costa and Lopes [22] improved the thermal performance of the
convection mode, Shen et al. [18] found that intensive fin arrays HS for a light emitting diodes (LED) lamp operating under natural
were effectively affected by the orientation. Fig. 1 shows the convection conditions as shown in Fig. 3. The effect of the number,
degrees of orientation taken into the account of their study. thickness, length and height of fins was examined. Their results
132 H.E. Ahmed et al. / International Journal of Heat and Mass Transfer 118 (2018) 129–153

Fig. 2. (a) Diagram of cylindrical heat sink for a LED light bulb and (b) computational domain [19].

standard HS, that the Rth of the optimized branched-fin HS


decreased by up to 30%. This reduction in the Rth was increased
with increasing the pumping power and decreasing the length of
the HS.
Türkakar and Özyurt [26] optimized the thermal design of sili-
con MCHS of Intel Core i7-900 Desktop Processor having a heat dis-
sipating of 130 W. They stated that when the hydraulic diameter
was reduced the pressure drop was increased across the channel
causing a decrease in the fluid velocity. They obtained as minimum
as of highest temperature 48.9 °C instead of 85 °C when the same
total heat dissipation equally distributed among four local heat
sources. They reported that the thermal performance of multi-
heat sources HS might be enhanced by intensification the channels
only around the hot spots while keeping coarser channels else-
where. They highlighted that their work could cover a wider range
of pumping powers and temperatures compared to original
designs of HS.
Hung and Yan [27] varied the channel height- or width-tapered
Fig. 3. The heat sink [22]. of MCHS numerically to improve the thermal performance. It was
found that the Rth decreased and then increased with the width-
tapered. A similar behavior was monitored with the height-
obtained by using the commercial code ANSYS-CFX illustrated the tapered. Briefly, the effect of height-tapered was negligible. A
remarkable importance of the geometric parameters on the heat lower and a relatively uniform temperature distribution were
removal for keeping the heat source maximum temperature main- screened for the width-tapered-channel design more than the par-
tained below the critical temperature. They reported that the allel or height-tapered channel design. Maximum thermal perfor-
objective core temperature of 65 °C was registered by increasing mance enhancement obtained from MCHS with width-tapered
the number of fins and their height and reducing the fin thickness. channel was around 16.7% over that of the parallel-channel for P.
These geometric parameters could cause an increase in the heat P > 0.4 W. While Hung et al. [28] examined the MCHS with
sink mass about 24% of the initial mass, but to a considerable tem- single-layered (SL), double-layered (DL) or tapered (T) channels
perature decrease. They recommended extending their proposed as shown in Fig. 5. They reported that the T-channel design has
methodology for exploring other types of heat sinks. the minimum temperature difference and the most uniform tem-
Kim [23] and Kim et al. [24] optimized the fin thickness of a ver- perature distribution (the best design compared to the others), fol-
tical FPFHS under natural convection when the varying was in the lowed by the DL-, and SL-channel configuration. The optimal Rth
direction normal to the fluid flow. The Rth was reduced by up to reduced with the pumping power for all cases, and the lowest Rth
10% and 15% when the fin thickness increased for the case of air was observed in the T-channel design.
and water coolant, respectively. However, the difference between Reyes et al. [29] optimized the hydrothermal performance of
the Rth of the HS with uniform thickness and the HS with variable MCHS using water by the implementation of tip clearance between
thickness decreased as the fin height and/or the heat flux the tip of fins and the upper cover of HS including different heights.
decreased. While he optimized in his research [25] the thermal The optimal tip clearance height showed a heat transfer enhance-
performance of a HS having fins branched in the direction normal ment about 83% at Re = 416 associated with a pumping power ris-
to water flow as shown in Fig. 4. It was illustrated, compared to a ing of 18%. At Re = 1300, the heat transfer behaved similarly while
H.E. Ahmed et al. / International Journal of Heat and Mass Transfer 118 (2018) 129–153 133

Fig. 4. Scheme of FPFHS, (a) Y-shape and (b) inverted Y-shape [25].

and the feasibility of enhancing 3D convection in the tip clearance


flow region experimentally.
Ramos et al. [30] stated that their novel channels arrangement
of MCHS exhibited suitable benefits for application in HSs. Some of
the proposed MCHS achieved excellent flow uniformity and tem-
perature distribution contours in the FFHS at remarkably low fric-
tional losses compared to the others as well as the conventional
configuration of HS.
Wang et al. [31] proposed a novel fin structure applicable to
compact air-cooled HS including fin having oblique dimples,
straight grooves, full oblique grooves, half oblique grooves with
dimple/cavity, and full oblique groove with dimple/cavity as well
as flat plate fins. The results indicated that the proposed full
grooved fin along with the dimple/cavity structure showed an
increase in the thermal performance of 25% accomplished with a
friction reduction of about 20%. A significant thermal enhancement
was registered for the proposed fin structure in the fully developed
flow region.

5. Location of inlet and outlet arrangement of heat sink


Fig. 5. A single channel of MCHS [28].
Hung et al. [8] examined numerically the effect of enlarging the
channel outlet (the channel outlet width and height) on the
the pumping power was 36% compared to the base case. They sug- hydrothermal performance of a porous MCHS. Their simulation
gested further works for future studies which the existence of an data indicated that the pressure drop across the MCHS was
optimum tip clearance height with stability issues within the fluid, reduced when the width or height enlargement ratio increased. It

Fig. 6. Temperature contour of the solid and liquid region of the MCHS, (a)–(d) new suggested design and (e) conventional design [32].
134 H.E. Ahmed et al. / International Journal of Heat and Mass Transfer 118 (2018) 129–153

Fig. 7. Schematic diagram of (A) inlet/outlet locations and (B) MCHS headers [35].

was observed that the thermal performance was enhanced and the by up to 26.7% at a dissipation power of 72 W compared with other
Rth was reduced with an enlarged channel outlet. Therefore, they reference inlet/outlet configurations. They concluded that the opti-
stated that the increase of the width or height enlargement ratio mum thermal performance of PFMCHS can be obtained by opti-
of the channel outlet played an important rule for improving the mization the inlet position and the geometrical parameters. Liu
hydrothermal performance of porous MCHS. et al. [34] showed that the inlet/outlet location played an impor-
Vinodhan and Rajan [32] investigated the effect of the location tant role in reducing the maximum temperature and increasing
of the inlet/outlet of the MCHS on the Rth and temperature distri- the heat transfer rate under uniform heat flux. They observed that
bution of the substrate. They obtained better Rth for their new the heater position at the inlet of MCHS showed higher maximum
designs compared to the conventional design. They motivated temperature than that being placed at the rear of the MCHS.
the rest of researchers saying the design may be further optimized Xia et al. [35] analyzed the effect of different inlet/outlet loca-
by varying channel dimensions. The maximum temperature of the tions, Fig. 7(A), and header shapes, Fig. 7(B) on the thermal resis-
substrate they obtained was 331, 325, 340, 333, and 392 K for the tance of MCHS. They highlighted that I-type provided better flow
configurations shown in Fig. 6, respectively. velocity uniformity and symmetrical flow distribution while Z-
Zhao et al. [33] explored the effect of inlet location, pin-fin type was the worst among them. It was screened that the rectan-
height, spacing distance and edge length on the thermal perfor- gular header shapes had better flow velocity uniformity than the
mance of pin-fin microchannel heat sink (PFMCHS) with an inlet trapezoidal followed by triangular headers.
near the midpoint of the diagonal and high AR micro pin-fin array Lelea [36] optimized the geometric parameters of the micro-
was designed. The maximum temperature of the chip was reduced tube heat sink (MTHS) on the thermal performance. The inlet

(a) (b)
Fig. 8. The single MT geometry, (a) side view and (b) front view [36].
H.E. Ahmed et al. / International Journal of Heat and Mass Transfer 118 (2018) 129–153 135

Fig. 9. Schematic diagram of the inlet and outlet positions of the HS [38].

cross-section has a rectangular shape and positioned tangentially Naphon and Klangchart [38] tested the effects of outlet
to the tube axis as shown in Fig. 8. The axial velocity was higher location on the jet impingement of fluid in mini-rectangular
near the tube wall and decreased in the center line of the tube. This fin HS as shown in Fig. 9. The monitored a significant effect
velocity behavior caused higher absorption of heat transfer and for the outlet port location on the fluid flow and temperature
eventually thermal enhancement. In addition, the lowest tempera- distribution.
ture was observed with a diameter of 225 lm and length of 1 mm. Chein and Chen [39] examined the MCHS with various inlet/
Also, Lelea [37] noted strong dependence of the position and num- outlet arrangements as illustrated in Fig. 10. It was monitored that
ber of the inlet jets on the hydrothermal performance of the MTHS. the highest temperature was taken place at the edge of the heat
For constant pumping power, the minimum temperature was sink due no heat dissipation by fluid convection. They observed
greater for multiple inlets compared with the single inlet jet mod- better uniformities for velocity and temperature in the HS when
ule. In addition, the multiple inlet jets designs had lower tempera- the fluid enters and exits vertically via inlet/outlet ports. It was
ture differences on the bottom surface between the inlet and outlet also found that V-type heat sink had better hydrothermal perfor-
of the MTHS. mance among these heat sinks types.

Fig. 10. Geometric configurations of MCHS [39].


136 H.E. Ahmed et al. / International Journal of Heat and Mass Transfer 118 (2018) 129–153

Fig. 11. Structure of MCHS; (a) PFHS, (b) Single-hole jet-cooling HS, and (c) DLMCHS [40].

Gong et al. [40] showed that the layout of MCHS was very 6. Rotating heat sinks
important to improve the thermal performance of HS by filling
metal foam in the inlet header to strengthen the flow distribu- Yang et al. [42] examined an air jet impingement rotating and
tion uniformity as illustrated in Fig. 11. The pin-fin array stationary HS under turbulent flow regime as illustrated in
arrangement enhanced the overall thermal performance of the Fig. 13. They found that Nu number increased with Re number
HS. They stated that although the jet-cooling HS possessed the for a stationary heat sink. The effect of Re number on the average
best cooling capacity, it was not proposed for the field of MCSH Nu number of a rotating HS with jet impingement was higher than
cooling because of a huge flowing resistance. The double-layer that for stationary HS for small Re, but it decreased with increasing
microchannel heat sink (DLMCHS) with the inlet at the bottom Re. They reported that by varying the fins geometry the hydrother-
layer was the second-best cooling capability with an acceptable mal performance of the HS can be optimized. From the open liter-
large pressure drop. Generally, there was substantial develop- ature, there is only one paper examining the thermal performance
ment on the thermal performance of the traditional MCHS with of rotating heat sink.
metal-foam inlet header and rectangle column fin MCHS under
the same pumping power. 7. Optimization of the substrate material
Chen et al. [41] explored different inlet/outlet arrangement of a
HS in order to obtain the lowest Rth and temperature distribution. Shkarah et al. [43] simulated a set of cases to show the superi-
Their proposed designs are illustrated in Fig. 12. ority of three types of the substrate; silicon, aluminum, and

Fig. 12. Proposed designs of HS with different inlet/outlet arrangement [41].


H.E. Ahmed et al. / International Journal of Heat and Mass Transfer 118 (2018) 129–153 137

having high thermal diffusivity with base fluids having a low


Prandtl number (Pr) showed better heat transfer enhancement.
Qu et al. [45] explored a hybrid HS with parallel plate fins sin-
tered onto the substrate top using three different metal types sat-
urated in the hollow basement of the substrate as shown in Fig. 14.
These three types are copper metal foam–paraffin composite, pure
paraffin, and solid copper basement. Their experiments showed
that the copper metal foam HS reduced the surface temperature,
and the time required to reach the melting point of the paraffin
was shorter. The surface temperature was reduced by either reduc-
ing foam porosity or foam pore density. They reported that during
the melting process, the temperature increased linearly for the
foam–PCM composite than the pure paraffin HS since the augmen-
tation in thermal conduction caused by the metal foam exceeded
the level of its suppression to natural convection of melted
paraffin.
Akhilesh et al. [46] studied the thermal design for proper sizing
of a composite heat sink (CHS), for maximizing the energy storage
and the time of operation until all of the latent heat storage is
exhausted. When the dimensions of the CHS were less than this
critical dimension, all of the PCM completely melted when the
CHS reached the set point temperature. They proposed a correla-
tion for chosen material properties within 10% average deviation.

Fig. 13. Rotating heat sink [42].


8. Optimization of heat sinks with boiling

Nascimento et al. [10] evaluated experimentally the flow boil-


ing of R134a inside a MCHS. It was noticed that by keeping the
graphene to provide the better thermal performance of MCHS. average vapor quality over the HS constant, the average heat trans-
They found that graphene effectively reduced the thermal resis- fer coefficient increased when the mass velocity increased. The HS
tance. Mohammed et al. [44] investigated the laminar flow and performance improved with decreasing the mass velocity and the
heat transfer characteristics of trapezoidal MCHS using various liquid sub-cooling at the microchannel inlet.
types of substrate materials; copper, aluminum, steel, and tita- McNeil et al. [9] studied the boiling of deionized water
nium. Their major findings confirmed that the substrate material and R113 in a pin-fins HS (in-line arrangement) at atmospheric

Fig. 14. Experimental setup of the three dissipation modes (a) metal foam–PCM composite; (b) pure PCM; (c) solid basement [45].

Fig. 15. (Left) inclined MCHS, and (right) symmetric/asymmetric elongated bubble and its size variation [47].
138 H.E. Ahmed et al. / International Journal of Heat and Mass Transfer 118 (2018) 129–153

pressure experimentally. The heat transfer coefficient of single- Continuously, the experiments of Li and Chao [3] indicated that
phase flow was associated with flow pattern and restricted, as higher Re number showed lower Rth. While this reduction in the Rth
the flow passage dimensions were much smaller than the capillary became smaller at the highest value of Re number. By keeping the
length for water where the slugs were observed in subcooled liquid fin width constant, the HS having a highest height of fins outper-
flows. They stated that the criterion of the slug flow onset was not formed the others in the thermal performance due to the largest
compatible with that occurred in macro-scale flows. The R113 dis- heat transfer area. In contrast, the optimal fin width (best thermal
played a slug and annular flow pattern and were not confined, with performance) increased with Re number by keeping the fin height
its capillary length similar to the channel flow dimensions. The unchanged. For wider fins, the flow passages between channels
refrigerant flow pattern transition criteria were comparable to became constricted. Conversely, for narrower fins, the heat transfer
macro-scale values. area of the HS was reduced. Briefly, both conditions reduced the
Wang et al. [47] explored the effect of the inclination angle of thermal performance of the HS. In general, the fin width should
MCHS on the convective boiling heat transfer of the dielectric fluid be optimized for every range of Re number.
HFE-7100 experimentally as illustrated in Fig. 15. They observed a The experiments of Li and Chen [49] showed that the Rth
similar heat transfer coefficient for the vertical upward and hori- decreased with increasing of Re number. The Rth was effectively
zontal, while for 45° upward, it considerably exceeded other con- reduced by adjusting the distance between the fins, and the opti-
figurations. Besides, the enhancement of heat transfer coefficient mal distance was increased with Re number. The influence of this
was reduced with increasing of the mass flux for inclined configu- distance on the Rth at high Re numbers was small because the Rth
rations. The downward arrangements always deteriorated the decreased with Re number. The Rth increased sharply when the
thermal performance, and more than 50% deterioration of heat fin width ratio W/L  0.135. Increasing the fin height caused an
transfer coefficient was encountered. The flow visualization increase the heat transfer area which eventually reduced the Rth.
emphasized that the thermal enhancement for upward inclination Generally, the influence of the fin height on the Rth was less obvi-
arrangement was due to the increase of slug velocity. ous than that of the fin width.
Sung and Mudawar [48] tested a single- (1P) and two-phase Lin et al. [1] tested oblique array of FPFHS to avoid the improve-
(2P) cooling performance of a microchannel having multi micro- ment of flow resistance caused by increasing the fin numbers in
jet impingement cooling scheme using the dielectric fluid HFE order to increase the heat transfer surface area as shown in
7100 as a coolant. As a result, the jet velocity has a significant effect Fig. 16. Their proposed design of HS showed better performance
on the 1P cooling performance. Higher jet fluid penetration in the than that with vertical plate-fins due to the larger surface area
axial microchannel flow was observed at high jet velocities, and a and accelerating flow between the fins. For high flow rates, the
strong impingement effect at the wall was recorded. At low jet extra cooling effect could yield in a reduction of 6 °C on the CPU
velocities, the effect of jets was highly compromised to the case temperature by introducing oblique fins. At low flow rates,
microchannel flow. During nucleate boiling, vapor layer develop- their new design yielded a better heat-dissipation capability than
ment along the microchannel with multi micro-jets was essentially the typical one.
different from that of the typical microchannels. Subcooled jet fluid
provided repeated regions of bubble growth followed by bubble
collapse unlike the continuous growth common in the typical ones. 10. Optimization of pin-fin heat sink (PFHS)
Their proposed design showed better wall temperature uniformity.
Huang et al. [50] optimized the thermal design of the rectangu-
lar PFHS with non-uniform fin widths. They reduced numerically
the Rth up to 12.98% compared to the original HS. Besides, Nu num-
9. Optimization of flat-plate fin heat sink (FPFHS) ber increased by 14.92% compared to the original one. Their exper-
imental results demonstrated that Rth decreased by 12.49% and Nu
Li et al. [2] tested experimentally the effects of the width, height increased by 14.21% compared to the original fin array. While
and number of fins on the thermal performance of vapor chamber Huang and Chen [51,52] optimized the HS design by varying the
of FPFHS. The results displayed that the maximum temperature fin widths and heights. The observed numerically and experimen-
was effectively reduced and heat transfer was more uniformly on tally a reduction in the Rth by 3.10%. In addition, Nu increased by
the base substrate by using the vapor chamber HS than that by alu- 3.20% compared to the original HS. Their designs of HS are shown
minum HS. The overall Rth of their new proposed HS reduced with in Fig. 17.
Re number, but this reduction became small as Re increased. At low Seyf and Feizbakhshi [53] examined the engineering application
Re number, the fin dimensions had a greater effect on the thermal of the using of nanofluids with different volume fractions,
performance, while the thermal performance increased with nanoparticle diameters and nanofluid types on the thermal resis-
increasing fin number at high Re number. tance and temperature distribution of the substrate of pin-fin

Fig. 16. (a) Vertical planar fins HS and (b) oblique fins HS [1].
H.E. Ahmed et al. / International Journal of Heat and Mass Transfer 118 (2018) 129–153 139

Fig. 17. HS with (a) fins with different widths, and (b) fins with different widths and heights [50–52].

microchannel heat sink. Hasan [54] examined the application of et al. [57] investigated the natural convection heat transfer of
nanofluids in PFMCHS having square, triangular or circular pin- square pins of PFHS. The results inferred that when the top of
fin shape. A higher heat transfer enhancement was noticed by the HS was downward facing orientation provided the lower heat
using nanofluids instead of water particularly by increasing the transfer rate. In addition, by increasing the fin number, the side-
nanoparticles concentrations associated with an increase in the ward arrangement had the same performance of downward
pressure drop. For all values of Re numbers, the circular pin-fins arrangement approximately. The optimal heat sink porosity was
outperformed the others shape of fins by removing more heat around 83% and 91% for the upward and sideward arrangement,
transfer, while a great pumping power was registered with the respectively. A higher number of fins played an important role in
square pin-fins shape. the downward arrangement, whereas a small effect was observed
Mohammadian and Zhang [55] tested numerically an alu- in the case of sideward arrangement. Maximum heat transfer per-
minum PFHS in which the heights of pin fins increase linearly with formance obtained was 3.2, 2.5 and 1.8 times for the downward,
the axial direction of the air flow for thermal management of upward and sideward arrangement HS, respectively, compared to
Lithium-ion battery pack. The results showed that their proposed the FPFHS.
design not only decreased the bulk temperature of the battery Chen et al. [58] carried out a set of simulations to study the
but also reduced the standard deviation of the temperature field reliability-based design optimization (RBDO) of inline and stag-
inside the battery as well compare to the HS without pin fins. gered PFHS for operating in an uncertain environment. They found
Higher inlet air temperature caused a decrease in the standard that the increase of the target reliability led to more symmetric and
deviation of the temperature field while an increase in the maxi- uniform temperature distribution around the core of the HS and
mum temperature of the battery was recorded. Besides, when then a greater ability to resist the environmental variations was
the inlet air velocity increased, the standard deviation of the tem- observed. In addition, the RBDO of HS outperformed the conven-
perature field first increased and persisted to the maximum point, tional HS in terms of more robust and reliable heat dissipation
and then decreased. Consequently, this increase in the velocity when faced with environmental uncertainties.
caused a decrease in the maximum temperature of the battery. Shafeie et al. [59] studied the laminar forced convection of
Liu et al. [56] carried out experiments to investigate the effect of PFMCHS cooled by water as shown in Fig. 18. They optimized the
an array of 625 staggered of square pin-fin on the performance of HS using entropy generation minimization. It was displayed that
PFMCHS considering deionized water as a coolant. It was also for the given pumping power, the heat transfer of finned PFMCHS
shown that the pressure drop and the average Nu number was slightly higher than PFMCHS. Short pins with oblique orienta-
increased with the fin Re number. The heat resistance decreased tion provided better heat removal under the corresponding pump-
with the pressure drop for a fixed surface temperature. Huang ing power. The height of the fins and channel played significant

Fig. 18. Schematics of different PFMCHSs, (a) oblique finned PFMCHS, (b) oblique PFMCHS, and (c) staggered PFMCHS [59].
140 H.E. Ahmed et al. / International Journal of Heat and Mass Transfer 118 (2018) 129–153

Fig. 19. Perforated pin-fins heat sink (in-line configuration) [61].

of strip fins are outperformed the in-line ones. The benefit of using
perforated fins was potential in the enhancing in heat transfer with
significant reduction in the pressure loss and heat sink mass. Fig. il-
lustrate one line of perforated fins in in-line arrangement. Al-
Damook et al. [62] investigated the thermal performance of pin-
fin heat sinks with a rectangular slotted or notched perforation
pin-fin heat sinks computationally, as illustrated in Fig. 20. They
monitored an increase in the heat transfer rate and a decrease in
the pressure drop when the size of the rectangular perforation
increased. The optimal heat transfer augmentation obtained was
10%, while the maximum reduction in the fan power consumption
registered was 30%.
Zhao et al. [63] could enhance numerically the cooling perfor-
mance of the micro square pin-fin heat sink by changing the pin-
fin porosity and pin-fin located angle as shown in Fig. 21. They
obtained an optimal porosity and located angle for thermal
performance.

11. Optimization of heat sink by using porous media

Hung et al. [8] concluded that for a certain width or height of


channel outlet enlargement ratios, the Rth of a porous MCHS is
Fig. 20. Slotted and notched perforation pin-fin heat sinks [62]. not necessarily to be less than that of plain MCHS if the pumping
power is not large enough to overcome the frictional loss.
Liu et al. [6] investigated a Lotus-type porous copper which is
role compared to the orientation of the pin fins. They reported that used for cooling of high-power electronic devices numerically
the heat removed by the finned PFMCHS was lesser than that for and experimentally. Their experimental results illustrated that this
the optimum simple MCHS at a corresponding pumping power. type of HS has an excellent heat transfer rate. The CFD data showed
Peles et al. [60] studied heat transfer and fluid flow over a bank that optimal porosity and pore diameter were existed for the HS to
of PFMCHS. They observed very low Rth in a pin fin heat sink com- conduct the maximal equivalent heat transfer coefficient. For the
pared to fin heat sink. A higher number of pin-fins was preferable same pressure drop conditions, the optimal pore diameter for the
to reduce the Rth at high Re number. PFHS provided more design water coolant was lesser than GaInSn coolant and the optimal
flexibility in terms of pin shapes and their spacing. porosity for water and GaInSn coolant was the same. Zhang et al.
Al-Sallami et al. [61] demonstrated that the perforated pin-fins [64] carried out experiments on a MCHS by using lotus-type por-
heat sink provided an effective heat transfer enhancement as ous (Gasar) metals with the long cylindrical pores formed during
shown in Fig. 19. They displayed that the staggered arrangements unidirectional solidification of metal–gas eutectic system. The

Fig. 21. Square pin-fin heat sink with (a) 0° rotated angle, and (b) 45° rotated angle [63].
H.E. Ahmed et al. / International Journal of Heat and Mass Transfer 118 (2018) 129–153 141

results showed that the lotus-type porous copper HS using water porous MCHS reduced significantly the wall superheat for the boil-
as a coolant has excellent heat transfer performance. They ing incipience, two to five folds augmentation in two-phase heat
obtained a larger heat transfer coefficient after simply cutting the transfer coefficient, and delay and mitigation of two-phase flow
porous copper along the vertical direction of pore axis into 4 or 8 instabilities at low to moderate heat fluxes. In addition, during
equal sections aligned in the pore axis direction. They attributed the flow boiling of the reentrant porous MCHS and at low heat
that to the increase in the penetrative porosity due to the reduction fluxes and vapor quality, nucleation boiling ruled the heat transfer
in the length of porous copper HS along the pore axis direction. mechanism. Besides and at moderate and high heat flux values,
Consequently, this yields to increase the flow rate and then forced convective boiling associated with thin film evaporation
enhance the heat transfer rate of the HS. They encouraged that was demonstrated. They concluded that the instabilities of the
some methods must be taken into account to improve the pore two-phase flow were accessed for the reentrant porous MCHS.
length and penetrative porosity during the fabricating of lotus- Wan et al. [4] investigated experimentally the thermal perfor-
type porous copper HS. mance of a porous MCHS particularly used for cooling electronic
Jeng and Tzeng [5] tested sintered copper beads smoothly with components having high heat fluxes. It was monitored that the
the radial plate fins of the copper HS by thin layers at high temper- heat flux of up to 140 W/cm2 was dissipated by the HS with the
ature to form a LED cooling device. They examined perforated and coolant pressure drop of about 34 kPa across the HS and the heater
not perforated HS as shown in Fig. 22 (only the perforated types). junction temperature of 62.9 °C at the coolant flow rate of 6.2 cm3/
Two modules were investigated; the plate-shape sintered-metal- s. It was also observed that Nu number was increased with increas-
bead (Model B) and strip-shape sintered-metal-HS (Model C) com- ing Re number. The increasing of coolant flow rate and heat fluxes
pared with the standard case of the pure copper finned HS (Model led to decrease the thermal resistance.
A). It was demonstrated that the thermal resistances of the Model Feng et al. [66] studied the thermal performance of finned metal
B and C were 29% and 16% higher than that of the Model A at cor- foam heat sink (FMFHS) and metal foam heat sink (MFHS) under
responding conditions without fan (pure natural convection heat impinging air jet cooling as shown in Fig. 24. Their experimental
transfer). It observed that the natural convection heat transfer results referred that the heat transfer of MFHS decreased monoto-
could not be strengthened by using the sintered-metal-bead layers. nously while that of FMFHS first increased and then slightly
The Rth was reduced by about 58%, 78% and 50% when a HS with decreased when the foam height increased at corresponding flow
the fan was used for the above three models, respectively. Under rates. When the pumping power was kept constant, the heat trans-
mixed convection heat transfer condition (HS with fan), the ther- fer of MFHS was disaffected to the foam height while it increased
mal resistance of Model B was 31% lower than that of Model A. as the foam height increased for FMFHS. The FMFHS removed a
Deng et al. [65] evaluate the two-phase boiling heat transfer heat about 1.5–2.8 times more than that for MFHS at the same
performance of a novel reentrant porous MCHS to show their fea- height, and same flow rate or pumping power.
sibility and applicability for cooling using deionized water as a Hashemi, et al. [7] examined numerically the effect of SiO2–wa-
coolant as illustrated in Fig. 23. They displayed that the reentrant ter nanofluid, channel AR and porosity of porous medium in a

Fig. 22. Plate shape heat sink (a) Model A; perforated HS without porous media, (b) Model B; perforated sintered-bronze-beads layers porous HS, and (c) Model C; perforated
sintered-bronze-beads layers porous HS [5].

Fig. 23. SEM photographs of reentrant porous MCHS, (a) cross section, and (b) top view of the microchannel [65].
142 H.E. Ahmed et al. / International Journal of Heat and Mass Transfer 118 (2018) 129–153

Fig. 24. Test samples, (a) FMFHS, and (b) MFHS, at height of 30 mm [66].

miniature FPFHS. They pointed out that the nanofluid-cooled HS increase in the pressure drop between the inlet and outlet of the
outperformed considerably the water-cooled one. It was demon- HS. For the smallest fin height, the Rth increased with increasing
strated that an increase in the AR or the porosity of the miniature inclination angle. The best hydrothermal performance obtained
FPFHS enhanced the heat transfer coefficient. was at an inclination angle of 90° for small fin height, whereas lar-
ger height, the benefit of choosing of the inclined angle reduced.
12. Optimization of heat sinks by using turbulators They concluded that for getting lowest Rth and lower increase in
the pressure drop, the inclination angle must be chosen carefully.
Ahmed et al. [67] optimized the thermal design of FPFHS by As a continuation work, Li and Chiang [69] focused on vapor
inserting ribs in between channels in different sizes, locations, chamber of FPFHS with shield experimentally. They pointed out
numbers and orientations as shown in Fig. 25. They examined that the maximum surface temperature of the HS was effectively
the effect of inserting the ribs in the channels by keeping the num- reduced by adding the shield, which pushed more cold air into
ber of fins constant and reducing them. They reduced the amount the channels and thus reduce the thermal resistance. However,
of substrate material by reducing the number of fins and adding using the shield caused a higher pressure drop across the HS. It
tiny ribs simultaneously and reduced the pumping power by keep- was also shown that the thermal enhancement increased with Re
ing the number of fins and inserting ribs with reducing the flow number, but this improvement declined as Re increased. The pro-
rate to get the same thermal performance of the original smooth posed ribbed heat sink with thinner, higher or more fins exhibited
HS. Ribbed flat-plate fin heat sink (RFPFHS) provided the thermal better performance in terms of both heat transfer and pressure
performance of 1.55 times greater than FPFHS under correspond- drop compared to typical ones. Li et al. [70] experiments and sim-
ing conditions. While this enhancement reduced with increasing ulations investigate a FPFHS with a pair of vortex generators (VG)
the number of ribs. For the same thermal performance, the pump- installed in a cross flow channel as shown in Fig. 27. It was
ing power of RFPFHS was reduced up to 69.65% compared to FPFHS observed that when the distance between the trailing edges of
case. In addition, RFPFHS having five channels with 15 ribs showed the VG was negligible, the thermal performance was worse due
the hydrothermal performance of 1.37 times better than FPFHS to the difficulty of air flowing into the HS. The optimal thermal per-
having nine channels, with a reduction in the substrate material formance was observed when the distance between the VG trailing
of 27.24%. edges equals the length of the HS and the distance between the
Tsai et al. [68] investigated the effect of an inclination angle of a trailing edge of each VG and the front end of the HS was zero. Fur-
plate heat shield on the hydrothermal performance of a rectangu- thermore, the lowest Rth and low increase in the pressure drop was
lar heat sink as shown in Fig. 26. The flow field was effectively registered at attack angle of 30°. Both the heat transfer and pres-
affected by the variation of the angle in both upstream and down- sure loss was observed increased with increasing the height of
stream of the HS because of the blockage effect causing an obvious the VG. A common-flow-up configuration of VG outperformed a

Fig. 25. Schematic diagram of (a) geometric parameters of FPFHS, and (b) computational domain [67].
H.E. Ahmed et al. / International Journal of Heat and Mass Transfer 118 (2018) 129–153 143

Fig. 26. An inclined plate shield to a FPFHS [68].

Fig. 27. (a) Heat sink with VG, (b) VG in common-flow-up configuration, and (c) VG in common-flow-down configuration [70].

common- flow-down configuration in term of thermal perfor-


mance. A great heat transfer enhancement and low-pressure drop
increase by using VG at low values of Re number compared to that
without VG.
As a continuation work, they examined the performance of a
FPFHS with a shield in a cross flow [71]. An obvious thermal
enhancement was obtained by using the shield due to more cool-
ant enter the channels of the HS. Low Rth was recorded by increas-
ing the Re number. The highest height of the fins showed the best
thermal performance for a given width. The optimal fin width (best
thermal performance) increased with increasing Re number. The
wider fins constructed the entrance of coolant air whereas the nar-
rower fins reduced the heat transfer surface area of the HS and con-
sequently both of them increased the Rth of the HS.
In the same research area, Shaalan et al. [72] studied the ther-
mos fluid performance of a FPFHS with a shield. It was found that
the shield enhanced the thermal performance of the HS particu-
larly at low Re number. In addition, the inclination angle of the
shield played an important role in enhancing the thermal perfor-
mance. The best inclination angle that provided minimum Rth var-
ied with the HS height. They derived correlations to estimate the Fig. 28. Front view of single microchannel with Y-shaped bifurcation [73].
mean Nu number in non-dimensional form for both HS with and
without the shield.
Li et al. [73] proposed a new design for water-cooled MCHS
with vertical Y-shaped bifurcation plates (with angles ranged from
60° to 180°) in order to enlarge the area of heat removal surface as
illustrated in Fig. 28. The results showed better thermal perfor-
mance for MCHS with Y-shaped bifurcation compared to the corre-
sponding straight MCHS. The optimal thermal performance
equipped by the Y-shaped bifurcation at the angle between the Fig. 29. Top view of a single microchannel with multistage bifurcations plates
[74,75].
two arms of the Y-shaped plates was kept at 90°. The Rth decreased
with the increase of Re number and it was smaller compared to the Singh and Patil [76] modelled an embossed heat sink having
smooth MCHS. Besides, a great increase in the pressure drop was repeated impressions on the fin surface under natural convection
recorded with the increase of Re number. condition as shown in Fig. 30. A maximum enhancement in Nu
Xie et al. [74,75] modelled a MCHS having multistage bifurca- number obtained was 2.86 at the impression angle of 45° and
tions plates under laminar flow regimes as shown in Fig. 29. They the impression pitch of 12 mm compared to smooth ones. Finally,
monitored better thermal performance for the MCHS with multi- the embossed fin effectiveness declined with the increase in the
stage bifurcation, especially the heat sink with longer length bifur- heat input.
cation compared to the corresponding traditional ones. The
multistage bifurcated reduced the overall Rth and provided more
temperature uniformity of the heated surface. An obvious 13. Optimization of the shape of single-channel
enhancement of the thermal performance was observed when
the multistage bifurcations were used in the MCHS and the maxi- Wang et al. [77] studied an inverse geometric optimization for
mum number of stages of bifurcations was recommended to be 2. Al2O3-water nanofluid-cooled MCHS. They tested the parameters
144 H.E. Ahmed et al. / International Journal of Heat and Mass Transfer 118 (2018) 129–153

and staggered arrangement) conditions given the simultaneous


consideration of both heat transfer and viscous dissipation. The
effect of diameter, density, height and thermal conductivity of pins
and approach velocity of fluid were tested. It was demonstrated
that the optimum pin density was existed for each material. For
low thermal conductivity material, no optimum could be found
in the staggered arrangement for a given number of pins. The heat
sinks having high thermal conductivity and high pin density out-
performed other types.
Kuppusamy et al. [83] investigated the effect of the secondary
flows in MCHS by introducing slanted passage in rectangular chan-
nels walls in alternating orientation as shown in Fig. 31. These pas-
sages might cause disruption in the hydrodynamic boundary layer
and redevelopment at the leading edge of the following wall. Thus,
the thermal performance enhanced with minor pressure drop.
They recorded an increase in the overall performance of rectangu-
lar MCHS with alternating slanted passage by up to 146% and a
reduction in the Rth up to 76.8% compared to the simple MCHS.
The thermal enhancement was accomplished with a reduction in
the pressure drop up to 6% compared to the simple one.
Kim and Mudawar [84] developed an analytical methodology
for heat diffusion effects on rectangular, inverse trapezoidal, trian-
gular, trapezoidal, and diamond-shaped MCHS. The rectangular
microchannel with a high AR and small fin spacing outperformed
the others in the overall thermal performance under corresponding
conditions. They proved that the analytical methodology was very
effective tool for the design and Rth prediction of MCHS.
Fig. 30. Embossed fin geometry [76]. Mohammed et al. [85] found that the thermal performance of the
wavy rectangular MCHS was much better than the straight
microchannels under same cross-section condition and using
of the number, aspect ratio (AR), and the width ratio of the channel. water as a coolant. The heat transfer enhancement obtained from
They concluded that the cooling performance was enhanced by their new design of MCHS was higher than the increase in the pres-
increasing the pumping power; however, this enhancement was sure drop penalty. A proportional increase in both friction factor
reduced when the pumping power increased highly. and wall shear stress was monitored with increasing the wavy
Kim et al. [78] derived correlations for optimal fin thickness and amplitude. Chen et al. [86] found that Nu number had much higher
optimal channel width which reduce the Rth for given width, values at the inlet region, and then quickly reduced and
height and length of vertical flat-plate fin heat sink (FPFHS) under approached the constant fully developed value. The temperature
natural convection in a fully-developed-flow regime. From their in both substrate and fluid increased along the axial flow direction.
correlations, the optimal fin thickness depends on the height, the As a conclusion, the triangular microchannel showed the best ther-
solid and fluid conductivity only and is independent of Ra number, mal efficiency. Kroeker et al. [87] concluded that the rectangular
fluid viscosity, and the length of the HS. Biswal et al. [79] optimized MCHS outperformed the circular one in terms of Rth, while circular
analytically various design parameters such as eccentricity and sinks dissipated more heat per unit pumping power. Mohammed
footprint of heat source, the thickness of the MCHS base, channel et al. [88] carried out a numerical study to evaluate the effect of
aspect ratio, the number of fins, coolant flow rate, and thermal con- channels shape on the performance of MCHS by keeping the
ductivity of HS material on HS thermal resistances and pressure cross-sectional area constant. Channel shapes such as zigzag,
drop. They have got good agreement with experimental data. Their curvy, step, straight and wavy microchannels were covered. The
predicted analytical method was helpful for reducing the design temperature and the heat transfer coefficient of the zigzag MCHS
cycle time and time-to- market.
Before more than twenty yours, Knight et al. [80] determined
the HS dimensions that display the lowest thermal resistance.
Their parameters examined were; fin thickness/channel width
ratio, and the AR of the fluid channel for laminar and turbulent
flow regimes. Their results indicated that for the low pressure drop
across the inlet/outlet portion, laminar flow yielded lower Rth than
turbulent flow. Conversely, the optimal Rth was observed in the
turbulent region for high-pressure drop. Their configurations and
dimensions provided considerable improvement in the Rth from
(10–35%) over the previous studies. Before fifteen yours, Copeland
[81] estimated the optimum dimensions of fin thickness and pitch
of parallel plate heat sink. They stated that under specific operating
conditions (pressure drop or pumping power), the optimum outer
dimensions of the heat sink can be calculated, while not necessary
to estimate the fin thickness or pitch exactly to achieve high per-
formance. Khan et al. [82] carried out a study on parametric opti-
mization including geometric parameters, material properties
and flow conditions for determining optimum PFHS (both in-line Fig. 31. Schematic diagram of MCHS [83].
H.E. Ahmed et al. / International Journal of Heat and Mass Transfer 118 (2018) 129–153 145

Fig. 32. Geometrical parameters of MCHS varied in the study [89].

Fig. 33. (a) Structure of MCHS and (b) MCHS with offset fan-shaped reentrant cavities [90].

was the least and greatest, respectively, compared to the others


whereas the pressure drop penalty for all channel shapes was
higher than the conventional straight microchannel. Besides, the
zigzag shape showed the highest pressure losses and friction factor
followed by the curvy and step MCHS, respectively. Kuppusamy
et al. [89] optimized the thermal performance of trapezoidal
grooved MCHS using nanofluids. They claimed that maximum
value of (a) with a minimum value of (b), as shown in Fig. 32, of
the trapezoidal grooves showed the maximum thermal perfor-
mance. In contrast, they stated that the triangular groove would
be favorable compared to rectangular shape MCHS. From their Fig. 34. Structure of single microchannel [92].
results and conclusion, it is obviously understood that the pre-
ferred triangular groove cannot be obtained unless (a) becomes
as small as possible and (b) as much as possible. Alfaryjat et al. [93] found that the smallest hydraulic diameter
Chai et al. [90] explored experimentally and numerically the of the hexagonal microchannels showed the highest pressure
effect of periodic expansion–constriction cross-sections on the drop and heat transfer coefficient among circular and rhombus
heat transfer of MCHS as shown in Fig. 33(a). They remarked an MCHS. In contrast, the rhombus shape had the lowest friction fac-
increase in Nu number up to 1.8 for their proposed design of HS. tor and Rth followed by the hexagonal and circular cross-sections.
The Nu number was observed to decrease in the expansion region Generally, hexagonal microchannel was the best channel shape
and increase in the constriction region. While Chai et al. [91] ana- for the heat transfer and pressure drop, while rhombus
lyzed the behavior of MCHS with offset fan-shaped reentrant cav- microchannel was the best channel in terms of wall temperature,
ities hydrothermally as shown in Fig. 33(b). This configuration had friction factor, and Rth.
improved heat transfer performance with moderate pressure drop. Chai et al. [94] examined numerically rectangular ribs inserted
They attributed the enhancement in the heat transfer to the peri- in the transverse (interrupted) microchannels as shown in Fig. 35.
odic thermal developing flow and increase in the heat transfer sur- The variable parameters were the rib length and width with vary-
face area. Whereas the throttling and jet enhanced the heat ing the arrangement the rib position. They concluded that the
transfer and increased the pressure drop simultaneously. Besides, interrupted microchannel with ribs was a good passive method
a pressure drop reduction and a dramatic decrease of heat transfer for heat transfer enhancement for Re < 600 only. They reported
attributed to the slipping over the reentrant cavities. that according to the International Technology Roadmap for Semi-
While Xia et al. [92] focused their investigation on the MCHS conductors (ITRS), the peak power consumption of high perfor-
configuration shown in Fig. 34. They obtained an enhancement in mance desktops will rise by 96% (147–288 W) in 2016, and by
Nu number by 1.3–3 times with 6.5 times increase in the friction 95% (91–158 W) in lower-end desktops in 2016. They motivated
factor compared to rectangular channels. They stated that the the researchers to apply this technique in other engineering appli-
effect of rib height was stronger than the individual effect of the cations, such as pharmaceutical applications, refrigeration, and
size or arrangement of the reentrant cavity for Re > 300. chemical engineering industry.
146 H.E. Ahmed et al. / International Journal of Heat and Mass Transfer 118 (2018) 129–153

Fig. 35. The new interrupted microchannel with rectangular ribs [94].

MiCHS had lower Rth with nanofluid compared to water. The


experimental study of Sohel et al. [98] on the thermal perfor-
mances of a copper MiCHS showed that the thermal performance
was highly improved when alumina oxide-water nanofluid (0.1–
0.25 vol%) used rather than pure distilled water. Greater enhance-
ment in heat transfer coefficient, lowest HS base temperature, and
smallest Rth were found about 18%, 2.7 °C, and 15.72%, respectively
at 0.25 vol%. Moraveji et al. [99] examined the effect of TiO2- and
SiC–water nanofluids (0.8–4 vol%) on the thermal performance of
MiCHS. They showed that the heat transfer rate was higher using
nanofluids particularly with increasing nanoparticle concentration
and Re number. They correlated two correlations; one for Nu num-
ber and another for friction factor with an error about 5% for both
equations. They included the nanoparticles concentrations into the
account in their correlations.
Fig. 36. Offset strip-fin microchannels [95].
Moraveji and Ardehali [100] examined the precision of four CFD
modeling (single phase, VOF, mixture, Eulerian) of two-phase lam-
inar forced convection of Al2O3 nanofluid (0.5–6 vol%, and 32 nm
Hong and Cheng [95] optimized numerically the thermal per- particle diameter) flows in MiCHS. They stated that the two-
formance of offset strip-fin MCHS using water as a coolant as phase models were more precise in both hydrodynamic and ther-
shown in Fig. 36. They stated that the convection heat transfer mal performance than single phase model compared to experi-
was enhanced by mixing the cold and hot fluid due to the period- mental reference data. They reported that the mixture model
ical change of the flow direction. The periodical strip-fins was a was most precise and requires less CPU usage and run time com-
factor for breaking the thermal boundary layer. They observed a pared to the others. Ho et al. [101] an experimental study is carried
reduction in the required mass flow rate to keep the maximum out to examine the cooling performance of a copper MiCHS using
wall temperature when the ratio of (K = fin interval/fin length) Al2O3–water nanofluid (0–10 vol%) and/or microencapsulated
was kept constant and the fin numbers (M) was increased. There phase change material (MEPCM) particles. They registered better
is an optimal K and M to minimize the pumping power, and this heat transfer enhancement when Al2O3–water nanofluid was used
optimal point depends on the input heat flux and maximum wall rather than the hybrid water-based suspensions of Al2O3 nanopar-
temperature. ticles and MEPCM particles at higher flow rates.
Fani et al. [102] found, by using Eulerian–Eulerian two-phase
14. Optimization of heat sinks by working fluid model, that copper oxide nanoparticles (1–4 vol%) caused a
decrease in the heat transfer and increase in the pressure drop
The experiments of Ho et al. [96] on using nanofluids in copper across a trapezoidal MCHS due to the large nanoparticle diameter
MCHS proved the remarkable increase in the dynamic viscosity (i.e., higher than 150 nm). The heat transfer results obtained by
due to dispersing the alumina nanoparticles in water base fluid two-phase approach were higher than that of the single phase
(0–2 vol%), while a slight increase in the friction factor was (homogeneous) model. As a consequence, the base fluid has more
recorded. Besides, a significant heat transfer enhancement was effect on thermal performance than nanoparticles.
registered and thereby lower Rth and wall temperature was The set of experiments carried out by Rimbault et al. [103]
observed. Continuously, the experiments of Ho and Chen [97] were showed that CuO–water (0.24–4.5 vol%, and 29 nm) nanofluid
focused on the effect of alumina oxide-water nanofluid (0–10 vol%, flowing in a rectangular MCHS, under both laminar and turbulent
and 85–100 nm particle diameter) on thermal performance of a conditions, caused an increase in the friction factor as high as
copper mini-channel heat sink (MiCHS). It was noticed that the 70% for 4.5 vol% with respect to the base fluid. They claimed that
H.E. Ahmed et al. / International Journal of Heat and Mass Transfer 118 (2018) 129–153 147

a slight heat transfer enhancement observed with nanofluids at was observed with Al2O3 nanofluid. Highest pressure drop and wall
low concentrations, while a clear decrease of heat transfer was reg- shear stress was recorded with SiO2 nanofluid while the opposite
istered at above volume fraction. In terms of Performance Ener- was monitored with Ag nanofluid. They concluded that diamond
getic Coefficient (PEC: is a ratio of heat transferred/pumping and Ag nanofluids are recommended to get higher thermal
power) was lower than that of water for a given Re number. In enhancement and low-pressure losses, respectively, compared
addition, it decreased with increasing the volume fraction and Re with pure water.
number. Finally, they stated that best results were obtained by Lelea and Laza [111] analyzed the thermal behavior of MTHS
water compared to all the nanofluids tested, and they encouraged with five multiple tangential inlet jets using Al2O3–water nanofluid
authors to focus of PEC of nanofluids flow inside a MCHS. (0–5 vol%, and 13–47 nm in diameter) as a working fluid. They sta-
Hung et al. [104] reported that by using base fluids (having a ted that when Re is constant in the analysis, the maximum surface
lower dynamic viscosity such as water) and substrate materials temperature was better with nanofluid rather than water. For
(having high thermal conductivity) enhanced the thermal perfor- higher viscosity, the fluid velocity was greater in order to keep
mance of the MCHS. This performance was further enhanced and Re number fixed. Therefore, the velocity must be determined if
then reduced when the volume fraction increased. Smallest parti- the improvement is based on higher thermal conductivity or
cle diameter with adjusting their concentration and the pumping increasing fluid velocity. When the mass flow rate or pumping
power could provide lowest thermal resistance. Anyway, nanoflu- power was fixed in the analysis, the maximum temperature of
ids (1–5 vol%, and 5–38 nm particle diameter) enhanced the heat the substrate was very close to each other. Nanofluids did not have
transfer of MCHS better than pure water. Mohammed et al. [105] an effective effect on the maximum temperature difference when
revealed that when the volume fraction increased, both the heat the imposed constraint was the surface temperature difference.
transfer coefficient and wall shear stress were increased. However, They observed an effect for the nanofluid particle diameter which
Al2O3–water nanofluid (0–5 vol%) could not enhance the heat minimized the maximum temperature at constant pumping power
transfer, and the performing almost was the same of pure water. for variable nanofluid viscosity with particle diameter and fixed
They registered only a slight increase in the pressure drop across volume fraction.
the MCHS compared to the base fluid. Tokit et al. [112] examined the effect of nanofluids flow through
Yang et al. [106] studied the incompressible steady laminar and an interrupted rectangular MCHS. It was pointed out that Nu num-
turbulent nanofluid flow of a trapezoidal MCHS numerically. The ber was higher than the conventional MCHS with a slight increase
CuO/water nanofluid (0–0.4 vol%) was simulated considering in the pressure drop. It was screened that highest thermal
single-phase and two-phase (mixture) models. They monitored enhancement was registered with Al2O3, followed by CuO, and
that the two-phase model was more precise than the single- finally for SiO2. They emphasized that Nu number behaves propor-
phase model comparing to the experimental results. The Rth of tionally with nanoparticle concentration (1–4 vol%) and reversely
nanofluid was smaller than that of water, which decreased as the with the nanoparticle diameter (30–60 nm in diameter).
particle concentration and Re number increased, while the pres- Seyf et al. [113] investigated the hydrothermal performance of a
sure drop was increased slightly with the concentration. MTHS with tangential impingement with nano-encapsulated
Halelfadl et al. [107] optimized a rectangular MCHS using phase change materials (NEPCM) slurry as coolant with the base
nanofluids (aqueous carbon nanotubes 1 vol%, and 9–10 nm in fluid of Octadecane for NEPCM and polyalphaolefin (PAO). Their
diameter) as a coolant. They investigated the effects of the temper- results revealed that adding NEPCM to the base fluid provided
ature, the channel AR, the channel wall ratio and nanofluid on the greater heat transfer rate. However, the pressure drop increased
Rth and the pumping power of MCHS. They observed that nanoflu- with mass concentration and Re number when NEPCM was used
ids reduced the total Rth and enhanced significantly the thermal while better temperature uniformity and lower Rth was obtained.
performances of HS at high temperatures. A remarkable effect of different parameters on slurry entropy
Peyghambarzadeh et al. [108] found experimentally that alu- was demonstrated. A decrease in generated total entropy was reg-
mina oxide (0.5–1 vol%) and copper oxide (0.1–0.2 vol%, and 20 istered with increasing mass concentration and Re number. Seyf
nm in diameter) nanofluids showed greater thermal performance and Feizbakhshi [53] obtained a significant enhancement of heat
and higher pressure drop compared to pure water applied in rect- transfer in the in PFMCHS due to replace water by a nanofluid. This
angular MCHS. They recorded a higher heat transfer enhancement enhancement increased with increasing the nanoparticles concen-
by using CuO-water nanofluid more than another one. The higher tration and Re number. However, higher concentration caused a
enhancement was obtained with increasing the nanoparticle con- higher pressure drop. The Nu number was increased with decreas-
centration but they claimed and stated that it cannot be said that ing the particle diameter of Al2O3 nano-powder while the trend
the enhancement increased with Re number. The lower Nu num- was opposite for CuO particles.
bers and higher heat transfer coefficients would be obtained in Naphon and Nakharintr [114] studied the turbulent heat trans-
the MCHS in comparison with the conventional diameter heat fer and nanofluids (0.4 vol% and 21 nm in diameter) flow charac-
exchangers. It showed the greater importance of both conductive teristics in a copper MCHS considering single-phase, two-phase
and convective heat transfer mechanisms in HSs. (mixture and VOF) models in their set of simulations. They
Chen and Ding [109] modelled the MCHS as a nanofluid (1–2 vol obtained a reasonable agreement between their predicted data
%)-saturated porous medium. They found that the temperature dis- and the measured results. Two-phase models were observed more
tribution of the channel wall was practically not sensitive to the appropriate than the homogeneous (single phase) model. They
inertial effect. In addition, the nanofluid temperature distribution have found better heat transfer enhancement by using nanofluids
and the total Rth changed remarkably, attributed to the inertial rather than base fluid.
force effect. Moreover, the effect of fluid inertia caused a reduction
in the total Rth and the temperature difference between the chan-
nel wall and nanofluid phase. 15. Optimization of single- and double-layer heat sink
Mohammed et al. [110] investigated the effect of Al2O3, Ag, CuO,
diamond, SiO2, and TiO2 nanoparticles (2 vol%) suspended in water Hung et al. [13] explored numerically the heat transfer charac-
on the thermal performance of aluminum triangular MCHS. They teristics of a DLMCHS as shown in Fig. 37. Their predictions dis-
implied that lowest temperature and highest heat transfer played that the substrate materials having a higher thermal
enhancement was obtained by diamond nanofluid, the opposite conductivity ratio provided the higher thermal performance of
148 H.E. Ahmed et al. / International Journal of Heat and Mass Transfer 118 (2018) 129–153

Fig. 37. Schematic of DLMCHS [13].

the HS. Better heat transfer was observed with a coolant having increased with increasing the channel number, channel width
high thermal conductivity and low dynamic viscosity. They ratio, or channel aspect ratio. The thermal performance of the
recorded a decrease in the pressure loss when the channel AR channel was augmented by decreasing the channel number or
and channel width ratio increased. They stated that the Rth of the channel aspect ratio, or increasing the channel width ratio. They
MCHS can be reduced by optimizing the geometrical parameters. highlighted that the overall Rth was reduced with increasing the
They proved that the thermal performance of the DLMCHS was pumping power. The main statement in their research is that the
better than that of the single-layered one for the same geometric thermal resistance of MCHS with nanofluids was not necessarily
dimensions, by about 6.3%. lower than that of water under all pumping power ranges, but it
Hung and his group [115] continued in their research to opti- can be reduced by adjusting the geometric parameters of the HS.
mize the geometric parameters of the DLMCHS involving the fol- Wu et al. [11] investigated numerically the hydrothermal perfor-
lowing geometric parameters. The minimum Rth of MCHS they mance of DLMCHS under different geometrical parameters and
obtained was about 0.12 °C/Wm2 for a bottom area of 100 mm2 flow conditions. They stated that the optimal width ratio of
and heat flux of 100 W/cm2. By optimizing the geometric parame- DLMCHS should be increased when the microchannel AR was
ters; the number of channels, channel width ratio, lower channel increased. They also proved that the overall thermal performance
aspect ratio, and upper channel aspect ratio, they recorded maxi- of MCHS depends on the pumping power. DLMCHS with higher
mum reduction in the overall Rth by 52.8% compared to their orig- AR and smaller width ratio is suited to the situation when higher
inal case. The reduction in the Rth increased with the pumping pumping power is provided. They observed that when the inlet
power until to reach to a constant value, which means there is velocity of upper channels was smaller than that of bottom chan-
no power saving by increasing the pumping power. nels enhanced the overall performance of DLMCHS for a given
Ahmed et al. [116] tested a new innovative design of aluminum velocities, particularly for low pumping power values.
rectangular (RDLMCHS) and triangular (TDLMCHS) using different Wong and Muezzin [117] conducted a numerical study to pre-
types of nanofluids. They varied the parameters of channel dimen- dict the thermal performance of a parallel flow DLMCHS compared
sions, nanoparticles concentrations, and types and pumping pow- to the case of counter flow. Their concrete findings revealed that
ers throughout the experiments. They firstly observed an the parallel flow configuration provided better heat transfer for
excellent hydrothermal performance for DLMCHS over traditional low Re numbers and low channel aspect ratios. The Rth can be
SLMCHS. In addition, the sequential TDLMCHS provided a 27.4% reduced when the smaller thickness of middle rib is used in a par-
reduction in the wall temperature comparing to the RDLMCHS allel flow DLMCHS. Shao et al. [118] optimized the size of DLMCHS
and had better temperature uniformity across the HS length with numerically. The optimal Rth they obtained was about 0.4025 °C/W
less than 2 °C. Sequential TDLMCHS provided 16.6% total Rth lesser corresponding to their conditions. They have got a good agreement
than the RDLMCHS at low pumping power. No significant differ- between their numerical data and the analysis result, which
ences monitored in the pressure drop between the two designs. emphasized the reasonability of using the theory analysis of Rth
A larger number of channels and a smaller fin thickness inferred network model.
less Rth rather than only increasing the pumping power. Higher
particle volume fraction provided better thermal stability for both 16. Optimization of heat sinks by varying the size
nanofluids than pure water with the best thermal performance of
Al2O3–H2O nanofluid (0.9 vol%) was registered with a temperature 16.1. Mini-channel heat sink (MiCHS)
difference of 1.6 °C.
Hung and Yan [12] investigated the effect of nanofluids and Fan et al. [119] examined a novel cylindrical oblique fin MiCHS
geometric parameters on the thermal performance of a DLMCHS. in order to optimize the thermal performance of the HS for laminar
They pointed out that Al2O3–water nanofluid (1 vol%) showed the flow regime as shown in Fig. 38. The oblique angle was varied from
better heat transfer enhancement about 26% compared to water. 20° to 45° with Re number range from 200 to 900. They found that
They reported that the Rth behavior was decreased and then Nu number depended on the geometric parameters of the oblique
H.E. Ahmed et al. / International Journal of Heat and Mass Transfer 118 (2018) 129–153 149

Fig. 38. Full domain configuration for cylindrical oblique fin MiCHS [119].

fins, Re number and Pr number. The secondary flows were


strengthened when the length of the secondary channels
increased. They monitored a re-developing boundary layer at each
leading edge of the oblique fin and consequently enhanced the
heat transfer and intensified the recirculation zones, but suffered
a high-pressure drop penalty at higher Re numbers. The enhance-
ment of heat transfer reduced at higher values of Re numbers
and larger values of oblique angles due to the reduction in heat
transfer surface area causing additional pressure losses.
Ho et al. [120] explored the heat transfer characteristics of
water-based suspensions of phase-change nano-capsules in a nat-
ural circulation loop with a rectangular copper MiCHS and heat
source. The total HS length was 50 mm and a hydraulic diameter
was 0.96 mm. They revealed that water-based suspensions of
phase-change nano-capsules enhanced remarkably the heat trans-
fer rate of the natural circulation loop. Higher phase-change mate-
rial concentration was required to enhance heat transfer
Fig. 39. The microchannel with trapezoidal cavities [123].
adequately in increased wattages. Smaller Rth was observed when
modified Ra numbers increased led to improving Nu number at
both MiCHS and heat source. While they found that Al2O3–water
that the trapezoidal groove with groove tip length ratio of 0.5,
nanofluid enhanced the thermal performance of the natural circu-
groove depth ratio 0.4, groove pitch ratio of 3.334, grooves orienta-
lation loop at the heat source and heat sink about 3.5–22% and 9.5–
tion ratio of 0.00 and Re = 100 was the optimum thermal design for
62%, respectively [121].
GMCHS with Nu number augmentation of 51.59% and friction fac-
Jajja et al. [122] studied the effect of fin spacing on the thermal
tor increasing of 2.35%.
design of flat plate copper MiCHS for high heat generating micro-
Xie et al. [74,75] modelled a MCHS having multistage bifurca-
processors using water as a coolant. Five different fin spacing were
tions plates under laminar flow regimes. They monitored better
considered ranged from 0.2 mm to 1.5 mm. They obtained a lowest
thermal performance, lower overall Rth and more temperature uni-
heat sink base temperature when fin spacing of 0.2 mm was con-
formity of the MCHS with multistage bifurcation, especially the
sidered in the experiments. The thermal resistance and HS base
heat sink with longer length bifurcation compared to the conven-
temperature reduced by decreasing the fin space and increasing
tional ones. Li et al. [73] proposed a new design for water-cooled
Re number.
MCHS with vertical Y-shaped bifurcation plates (with angles ran-
ged from 60° to 180°) in order to enlarge the area of heat removal
16.2. Micro channel heat sink (MCHS) surface. They observed better thermal performance for MCHS with
the bifurcation compared to the corresponding straight ones. The
Ahmed and Ahmed [123] optimized the geometric parameters optimal thermal performance obtained when the angle between
of the depth, tip length, pitch and orientation of the cavities of the two arms of the Y-shaped plates was kept at 90°. Lower Rth
grooved microchannel heat sink (GMCHS) under laminar forced and higher pressure drop were monitored with increasing Re num-
convection heat transfer as shown in Fig. 39. The variation in the ber and they were smaller compared to the smooth MCHS.
tip length of the groove could change the cavity shape from trian- The experiments of Ho et al. [96] registered a significant
gular to trapezoidal and then to a rectangular shape. They revealed enhancement in heat transfer rate and on in copper MCHS by using
150 H.E. Ahmed et al. / International Journal of Heat and Mass Transfer 118 (2018) 129–153

nanofluids thereby lower Rth and wall temperature was observed. phase change materials (NEPCM) slurry as coolant with the base
Fani et al. [102] examined the copper oxide nanoparticles in trape- fluid of Octadecane for NEPCM and polyalphaolefin (PAO). Their
zoidal MCHS and they claimed that the heat transfer was decreased results revealed that adding NEPCM to the base fluid provided
with an increase in the pressure drop across the microchannel. The greater heat transfer rate. However, the pressure drop increased
same claim was pronounced by Rimbault et al. [103], when they with mass concentration and Re number when NEPCM was used
examined CuO–water nanofluid in a rectangular MCHS under lam- while better temperature uniformity and lower Rth was obtained.
inar and turbulent conditions in terms of Performance Energetic Seyf and Feizbakhshi [53] obtained a significant enhancement of
Coefficient, it was lower than that of water for a given Re number. heat transfer in the in PFMCHS due to replace water by a nanofluid.
Hung et al. [104] reported that the base fluid having a lower This enhancement and the pressure drop increased with increasing
dynamic viscosity such as water and substrate materials having the nanoparticles concentration and Re number. Naphon and
high thermal conductivity could enhance the thermal performance Nakharintr [114] obtained a reasonable agreement between their
of the MCHS. They stated that nanofluids enhanced the heat trans- predicted data and the measured results of turbulent nanofluids
fer of MCHS better than pure water. Mohammed et al. [105] con- flow in a copper MCHS. They have found better heat transfer
firmed the increase in the thermal and hydraulic performance of enhancement by using nanofluids rather than base fluid.
MCHS by using nanofluid while Al2O3–H2O nanofluid (5 vol%) It can be concluded in this section that most investigations
showed the same performance of pure water. existed in the open literature examine the effect of nanofluid flow-
For laminar and turbulent flow regimes of nanofluids in trape- ing in microchannel heat sink. No paper has been published testing
zoidal MCHS, Yang et al. [106] observed lower Rth than using water, the effect of the micro-size on the hydrothermal performance of
which decreased as the particle concentration and Re number the heat sink.
increased. A slight increase in the pressure drop was monitored In 2017, Sidik et al. [124] reviewed more than ninety papers
with the concentration. that investigate the passive cooling techniques only, such as chan-
The optimization study of nanofluids in a rectangular MCHS nel types, surface roughness, fluid additives, and Reynolds number,
done by Halelfadl et al. [107] by examining the effects of the tem- which were used for removing heat from heat sinks in micro-sizes
perature, the channel AR, the channel wall ratio and nanofluid on only (MCHS). They concluded that many challenges most be settled
the Rth and the pumping power of MCHS. They stated that the opti- for different applications. The stability and synthesis cost are the
mized AR (i.e., narrower channel and higher exchange surface) was main obstacle for using the nanofluids in MCHS.
lower for nanofluids compared to water. Peyghambarzadeh et al.
[108] found experimentally that alumina oxide and copper oxide
17. Temperature jump
nanofluids showed greater thermal performance and higher pres-
sure drop compared to pure water applied in rectangular MCHS.
Bushehri et al. [125] explored the effect of temperature jump
The higher enhancement was obtained with increasing the
boundary condition (TJBC) on heat transfer and temperature distri-
nanoparticle concentration but they claimed and stated that it can-
bution in FFHS. They proposed a new method for coupling equa-
not be said that the enhancement increased with Re number. They
tions between solid and fluid domains with TJBC in the open-
claimed that lower Nu numbers and higher heat transfer coeffi-
source CFD package, Open FOAM. They considered boundary con-
cients would be obtained in the MCHS compared to the conven-
ditions, constant wall temperature and constant heat flux. They
tional diameter heat exchangers.
observed a significant effect for the temperature jump on temper-
Chen and Ding [109] modelled the MCHS as a nanofluid-
ature field and heat transfer in HS. For uniform heat flux boundary
saturated porous medium. They found that the temperature distri-
condition, higher thermal conductivity ratios reduced the mean
bution of the channel wall was practically not sensitive to the iner-
temperature in the solid and fluid domain, while an adverse trend
tial effect. In addition, the nanofluid temperature distribution and
was monitored in another boundary condition.
the total Rth changed remarkably, attributed to the inertial force
Pang et al. [126] evaluated the thermal behaviors of a hybrid
effect. Moreover, the effect of fluid inertia caused a reduction in
thermoelectric/photovoltaic (PV/TE) system. They noticed more
the total Rth and the temperature difference between the channel
effective structure through combining a HS with a TE module prof-
wall and nanofluid phase. Mohammed et al. [110] investigated
its heat dissipation by cooling down the whole cell by 8 °C,
the superiority of different types of nanofluids which provide
wherein the TE module itself demonstrated the cooling perfor-
higher thermal performance of aluminum triangular MCHS. Dia-
mance by 27% enhancement in addition to its traditional one for
mond nanofluid was the first while Al2O3 nanofluid was the worst.
electricity generation. They stated that the PV/TE with a proper
Highest pressure drop was recorded with SiO2 nanofluid while the
design can be used as a passive method for improving the cell effi-
opposite was monitored with Ag nanofluid. They concluded that
ciency and alleviating hot spot. They encouraged the researchers to
diamond and Ag nanofluids are recommended to get higher ther-
do further investigations to develop the stability of power genera-
mal enhancement and low-pressure losses, respectively, compared
tion of a hybrid PV/TE system and high-powered light emit diode.
with pure water. Lelea and Laza [111] analyzed the thermal behav-
It is worth to be mentioned that only two papers have been pub-
ior of MTHS with five multiple tangential inlet jets using Al2O3–
lished in this field of research.
water nanofluid as a working fluid. They stated that when Re
was constant, the maximum surface temperature was better with
nanofluid rather than water. For higher viscosity, the fluid velocity 18. Conclusion
was greater in order to keep Re number fixed. The surface temper-
ature difference is not very much improved by using the nanofluid. In the present paper, the techniques which are used for improv-
The smallest nanoparticle diameter reduced the maximum tem- ing the hydrothermal performance of heat sinks are reviewed. The
perature at constant pumping power and fixed volume fraction. review includes the optimization of the thermal design of the heat
Tokit et al. [112] examined the effect of nanofluids flow through sinks by examining the pulsating flow and agitation, shapes and
an interrupted rectangular MCHS. They have found that the Nu orientations of fins, the inlet/outlet of the heat sinks, static and
number for interrupted MCHS was higher than that of straight rotating heat sinks, substrate material of the heat sinks, channel
MCHS with a slight increase in the frictional losses. configurations, adding of foams and porous media, using of turbu-
Seyf et al. [113] investigated the hydrothermal performance of a lators, additives to the conventional working fluids, single- and
MTHS with tangential impingement with nano-encapsulated double-layer heat sinks, and sizing of the heat sinks.
H.E. Ahmed et al. / International Journal of Heat and Mass Transfer 118 (2018) 129–153 151

It can be concluded that the inclination angle, orientation, and [5] T.-M. Jeng, S.-C. Tzeng, Heat transfer measurement of the cylindrical heat sink
with sintered-metal-bead-layers fins and a built-in motor fan, Int. Commun.
Ra number play an important role in increasing the heat removal
Heat Mass Transf. 59 (2014) 136–142.
from the heat sink. The natural heat transfer convection still [6] Y. Liu, H.F. Chen, H.W. Zhang, Y.X. Li, Heat transfer performance of lotus-type
requires more attentions from the researchers to increase the heat porous copper heat sink with liquid GaInSn coolant, Int. J. Heat Mass Transf.
transfer rate of the heat sinks. Applying an external force such as 80 (2015) 605–613.
[7] S.M.H. Hashemi, S.A. Fazeli, H. Zirakzadeh, M. Ashjaee, Study of heat transfer
agitation or pulsation flow could increase the heat removed from enhancement in a nanofluid-cooled miniature heat sink, Int. Commun. Heat
the heat sink. The varying of the fin number, fin geometry, channel Mass Transf. 39 (6) (2012) 877–884.
shape, channel aspect ratio, grooved channel, inlet/outlet location, [8] T.C. Hung, Y.X. Huang, W.M. Yan, Thermal performance of porous
microchannel heat sink: Effects of enlarging channel outlet, Int. Commun.
and ribs and turbulators in between channels have an attractive Heat Mass Transf. 48 (2013) 86–92.
effect for getting the optimal thermal design of the heat sinks. It [9] D.A. Mcneil, A.H. Raeisi, P.A. Kew, R.S. Hamed, International Journal of
can be said that there is a very lack in the available data for Multiphase Flow An investigation into flow boiling heat transfer and
pressure drop in a pin – finned heat sink, Int. J. Multiph. Flow 67 (2014)
enhancing the thermal performance of rotating heat sinks. Filling 65–84.
the base of the substrate of the heat sink modified the thermal [10] F.J. do Nascimento, H.L.S.L. Leão, G. Ribatski, An experimental study on flow
design of the heat sink. Heat sink with porous media and/or perfo- boiling heat transfer of R134a in a microchannel-based heat sink, Exp. Therm.
Fluid Sci. 45 (2013) 117–127.
rated fins showed better heat transfer but the increase in the pres- [11] J.M. Wu, J.Y. Zhao, K.J. Tseng, Parametric study on the performance of double-
sure drop must be taken into the account particularly with dense layered microchannels heat sink, Energy Convers. Manage. 80 (2014) 550–
fluids such as nanofluids having high concentrations. Baffles, vor- 560.
[12] T.C. Hung, W.M. Yan, Enhancement of thermal performance in double-
tex generators, and ribs showed better hydrothermal performance
layered microchannel heat sink with nanofluids, Int. J. Heat Mass Transf. 55
more than smooth heat sinks. Double-layer heat sink outper- (11–12) (2012) 3225–3238.
formed the typical ones (single-layer) but cannot be applied for [13] T.-C. Hung, W.-M. Yan, W.-P. Li, Analysis of heat transfer characteristics of
high-pressure drops. The miniaturization of the heat sinks showed double-layered microchannel heat sink, Int. J. Heat Mass Transf. 55 (11–12)
(2012) 3090–3099.
excellent results in both hydraulic and thermal performance. A [14] I. Tari, M. Mehrtash, Natural Convection heat transfer from horizontal and
large numbers of papers published in mini- and micro-channel slightly inclined plate-fin heat sinks, Appl. Therm. Eng. 61 (2) (2013) 728–
considering traditional fluids and nanofluids. 736.
[15] M.M. Ilker Tari, Natural convection heat transfer from inclined plate-fin heat
sinks, Int. J. Heat Mass Transf. 56 (2013) 574–593.
19. Recommendations for future works [16] H. Hassan, Heat transfer of Cu–water nanofluid in an enclosure with a heat
sink and discrete heat source, Eur. J. Mech. – B/Fluids 45 (2014) 72–83.
[17] S.M. Aminossadati, B. Ghasemi, Natural convection of water–CuO nano fluid
A few numbers of investigations published in using active in a cavity with two pairs of heat source – sink, Int. Commun. Heat Mass
methods for thermal performance augmenting of the heat sinks. Transf. 38 (5) (2011) 672–678.
[18] Q. Shen, D. Sun, Y. Xu, T. Jin, X. Zhao, Orientation effects on natural convection
Zhang et al. [64] suggested taking the pore length and penetrative heat dissipation of rectangular fin heat sinks mounted on LEDs, Int. J. Heat
porosity into the account during fabricating the lotus-type porous Mass Transf. 75 (2014) 462–469.
copper HS. Further investigations are needed for enhancing the [19] D. Jang, S.J. Park, S.J. Yook, K.S. Lee, The orientation effect for cylindrical heat
sinks with application to LED light bulbs, Int. J. Heat Mass Transf. 71 (2014)
thermal design of MCHS by using more than two layers. Making
496–502.
grooves in the flat-plate fins or using slotted fins in order to [20] S. Xu, W. Wang, K. Fang, C.-N. Wong, Heat transfer performance of a fractal
improve the heat transfer, reducing the frictional losses and using silicon microchannel heat sink subjected to pulsation flow, Int. J. Heat Mass
Transf. 81 (2015) 33–40.
a smaller amount of substrate material lacks to extreme
[21] Y. Yu, T. Simon, T. Cui, A parametric study of heat transfer in an air-cooled
researches. Different foams and fillets are proposed to be filled at heat sink enhanced by actuated plates, Int. J. Heat Mass Transf. 64 (2013)
the base of the substrate of the heat sink for raising the heat 792–801.
removal. The major problem in the using of nanofluids in such tiny [22] V.A.F. Costa, A.M.G. Lopes, Improved radial heat sink for led lamp cooling,
Appl. Therm. Eng. 70 (1) (2014) 131–138.
channels as microchannels is the sedimentation which causes an [23] D. Kim, Thermal optimization of plate-fin heat sinks with fins of variable
agglomeration of the nanopowder and then clogging the passages thickness under natural convection, Int. J. Heat Mass Transf. 55 (4) (2012)
of the MCHS. This problem becomes more complex when the por- 752–761.
[24] D.K. Kim, J. Jung, S.J. Kim, Thermal optimization of plate-fin heat sinks with
ous media is used due to the porosity of the foam. variable fin thickness, Int. J. Heat Mass Transf. 53 (25–26) (2010) 5988–5995.
[25] D.-K. Kim, Thermal optimization of branched-fin heat sinks subject to a
Conflict of interest parallel flow, Int. J. Heat Mass Transf. 77 (2014) 278–287.
[26] G. Türkakar, T. Okutucu-Özyurt, Dimensional optimization of microchannel
heat sinks with multiple heat sources, Int. J. Therm. Sci. 62 (2012) 85–92.
The authors declared that there is no conflict of interest. [27] T.-C. Hung, W.-M. Yan, Effects of tapered-channel design on thermal
performance of microchannel heat sink, Int. Commun. Heat Mass Transf. 39
(9) (2012) 1342–1347.
Appendix A. Supplementary material [28] T.-C. Hung, T.-S. Sheu, W.-M. Yan, Optimal thermal design of microchannel
heat sinks with different geometric configurations, Int. Commun. Heat Mass
Supplementary data associated with this article can be found, in Transf. 39 (10) (2012) 1572–1577.
[29] M. Reyes, J.R. Arias, A. Velazquez, J.M. Vega, Experimental study of heat
the online version, at https://doi.org/10.1016/j.ijheatmasstransfer. transfer and pressure drop in micro-channel based heat sinks with tip
2017.10.099. clearance, Appl. Therm. Eng. 31 (5) (2011) 887–893.
[30] B. Ramos-Alvarado, P. Li, H. Liu, A. Hernandez-Guerrero, CFD study of liquid-
cooled heat sinks with microchannel flow field configurations for electronics,
References fuel cells, and concentrated solar cells, Appl. Therm. Eng. 31 (14–15) (2011)
2494–2507.
[1] S.-C. Lin, F.-S. Chuang, C.-A. Chou, Experimental study of the heat sink [31] C.-C. Wang, K.-S. Yang, Y.-P. Liu, I.Y. Chen, Effect of cannelure fin configuration
assembly with oblique straight fins, Exp. Therm. Fluid Sci. 29 (5) (2005) 591– on compact aircooling heat sink, Appl. Therm. Eng. 31 (10) (2011) 1640–
600. 1647.
[2] H.Y. Li, M.H. Chiang, C.I. Lee, W.J. Yang, Thermal performance of plate-fin [32] V. Leela Vinodhan, K.S. Rajan, Computational analysis of new microchannel
vapor chamber heat sinks, Int. Commun. Heat Mass Transf. 37 (7) (2010) 731– heat sink configurations, Energy Convers. Manage. 86 (2014) 595–604.
738. [33] J. Zhao, Y. Wang, G. Ding, Y. Sun, G. Wang, Design, fabrication and
[3] H.-Y.Y. Li, S.-M.M. Chao, Measurement of performance of plate-fin heat sinks measurement of a microchannel heat sink with a pin-fin array and optimal
with cross flow cooling, Int. J. Heat Mass Transf. 52 (13–14) (2009) 2949– inlet position for alleviating the hot spot effect, J. Micromech. Microeng. 24
2955. (11) (2014) 115013.
[4] Z.M. Wan, G.Q. Guo, K.L. Su, Z.K. Tu, W. Liu, Experimental analysis of flow and [34] C.-K. Liu, S.-J. Yang, Y.-L. Chao, K.Y. Liou, C.-C. Wang, Effect of non-uniform
heat transfer in a miniature porous heat sink for high heat flux application, heating on the performance of the microchannel heat sinks, Int. Commun.
Int. J. Heat Mass Transf. 55 (15–16) (2012) 4437–4441. Heat Mass Transf. 43 (2013) 57–62.
152 H.E. Ahmed et al. / International Journal of Heat and Mass Transfer 118 (2018) 129–153

[35] G.D. Xia, J. Jiang, J. Wang, Y.L. Zhai, D.D. Ma, Effects of different geometric [64] H. Zhang, L. Chen, Y. Liu, Y. Li, Experimental study on heat transfer
structures on fluid flow and heat transfer performance in microchannel heat performance of lotus-type porous copper heat sink, Int. J. Heat Mass Transf.
sinks, Int. J. Heat Mass Transf. 80 (2015) 439–447. 56 (1–2) (2013) 172–180.
[36] D. Lelea, Effects of inlet geometry on heat transfer and fluid flow of tangential [65] D. Deng, Y. Tang, D. Liang, H. He, S. Yang, Flow boiling characteristics in
micro-heat sink, Int. J. Heat Mass Transf. 53 (17–18) (2010) 3562–3569. porous heat sink with reentrant microchannels, Int. J. Heat Mass Transf. 70
[37] D. Lelea, The tangential micro-heat sink with multiple fluid inlets, Int. (2014) 463–477.
Commun. Heat Mass Transf. 39 (2) (2012) 190–195. [66] S.S. Feng, J.J. Kuang, T. Wen, T.J. Lu, K. Ichimiya, An experimental and
[38] P. Naphon, S. Klangchart, Effects of outlet port positions on the jet numerical study of finned metal foam heat sinks under impinging air jet
impingement heat transfer characteristics in the mini-fin heat sink, Int. cooling, Int. J. Heat Mass Transf. 77 (2014) 1063–1074.
Commun. Heat Mass Transf. 38 (10) (2011) 1400–1405. [67] H.E. Ahmed, Optimization of thermal design of ribbed flat-plate fin heat sink,
[39] R. Chein, J. Chen, Numerical study of the inlet/outlet arrangement effect on Appl. Therm. Eng. 102 (2016) 1422–1432.
microchannel heat sink performance, Int. J. Therm. Sci. 48 (8) (2009) 1627– [68] G.-L. Tsai, H.-Y. Li, C.-C. Lin, Effect of the angle of inclination of a plate shield
1638. on the thermal and hydraulic performance of a plate-fin heat sink, Int.
[40] L. Gong, J. Zhao, S. Huang, Numerical study on layout of micro-channel heat Commun. Heat Mass Transf. 37 (4) (2010) 364–371.
sink for thermal management of electronic devices, Appl. Therm. Eng. (2014). [69] H.-Y. Li, M.-H. Chiang, Effects of shield on thermal-fluid performance of vapor
[41] Y. Chen, B. Peng, X. Hao, G. Xie, Fast approach of Pareto-optimal solution chamber heat sink, Int. J. Heat Mass Transf. 54 (7–8) (2011) 1410–1419.
recommendation to multi-objective optimal design of serpentine-channel [70] H.Y. Li, C.L. Chen, S.M. Chao, G.F. Liang, Enhancing heat transfer in a plate-fin
heat sink, Appl. Therm. Eng. 70 (1) (2014) 263–273. heat sink using delta winglet vortex generators, Int. J. Heat Mass Transf. 67
[42] Y.-T. Yang, S.-C. Lin, Y.-H. Wang, J.-C. Hsu, Numerical simulation and (2013) 666–677.
optimization of impingement cooling for rotating and stationary pin-fin [71] H.-Y. Li, G.-L. Tsai, S.-M. Chao, Y.-F. Yen, Measurement of thermal and
heat sinks, Int. J. Heat Fluid Flow 44 (2013) 383–393. hydraulic performance of a plate-fin heat sink with a shield, Exp. Therm. Fluid
[43] A. Jassim, M. Yusoff, B. Sulaiman, R. Bin, H. Ayob, H. Togun, A 3D numerical Sci. 42 (2012) 71–78.
study of heat transfer in a single-phase micro-channel heat sink using [72] M.R. Shaalan, M.A. Saleh, O. Mesalhy, M.L. Elsayed, Thermo/fluid performance
graphene, aluminum and silicon as substrates, Int. Commun. Heat Mass of a shielded heat sink, Int. J. Therm. Sci. 60 (2012) 171–181.
Transf. 48 (2013) 108–115. [73] Y. Li, F. Zhang, B. Sunden, G. Xie, Laminar thermal performance of
[44] H.A. Mohammed, P. Gunnasegaran, N.H. Shuaib, In fl uence of various base microchannel heat sinks with constructal vertical Y-Shaped bifurcation
nano fluids and substrate materials on heat transfer in trapezoidal plates, Appl. Therm. Eng. 73 (1) (2014) 185–195.
microchannel heat sinks, Int. Commun. Heat Mass Transf. 38 (2) (2011) [74] G. Xie, F. Zhang, B. Sundén, W. Zhang, Constructal design and thermal analysis
194–201. of microchannel heat sinks with multistage bifurcations in single-phase
[45] Z.G. Qu, W.Q. Li, J.L. Wang, W.Q. Tao, Passive thermal management using liquid flow, Appl. Therm. Eng. 62 (2) (2014) 791–802.
metal foam saturated with phase change material in a heat sink, Int Commun [75] G. Xie, S. Li, B. Sunden, W. Zhang, H. Li, A numerical study of the thermal
Heat Mass Transf 39 (2012) 1546–1549. performance of microchannel heat sinks with multiple length bifurcation in
[46] R. Akhilesh, A. Narasimhan, C. Balaji, Method to improve geometry for heat laminar liquid flow, Numer. Heat Transf. Part A Appl. 65 (2) (2014) 107–126.
transfer enhancement in PCM composite heat sinks, Int. J. Heat Mass Transf. [76] P. Singh, A.K. Patil, Experimental investigation of heat transfer enhancement
48 (13) (2005) 2759–2770. through embossed fin heat sink under natural convection, Exp. Therm. Fluid
[47] C.C. Wang, W.J. Chang, C.H. Dai, Y.T. Lin, K.S. Yang, Effect of inclination on the Sci. 61 (2015) 24–33.
convective boiling performance of a microchannel heat sink using HFE-7100, [77] X. Wang, B. An, L. Lin, D. Lee, Inverse geometric optimization for geometry of
Exp. Therm. Fluid Sci. 36 (2012) 143–148. nano fluid-cooled microchannel heat sink, Appl. Therm. Eng. 55 (1–2) (2013)
[48] M.K. Sung, I. Mudawar, Single-phase and two-phase heat transfer 87–94.
characteristics of low temperature hybrid micro-channel/micro-jet [78] T.H. Kim, K.H. Do, D.-K. Kim, Closed form correlations for thermal
impingement cooling module, Int. J. Heat Mass Transf. 51 (15–16) (2008) optimization of plate-fin heat sinks under natural convection, Int. J. Heat
3882–3895. Mass Transf. 54 (5–6) (2011) 1210–1216.
[49] H.-Y. Li, K.-Y. Chen, Thermal performance of plate-fin heat sinks under [79] L. Biswal, S. Chakraborty, S.K. Som, Design and optimization of single-phase
confined impinging jet conditions, Int. J. Heat Mass Transf. 50 (9–10) (2007) liquid cooled microchannel heat sink, IEEE Trans. Components Packag.
1963–1970. Technol. 32 (4) (2009) 876–886.
[50] C.H. Huang, Y.H. Chen, H.Y. Li, An impingement heat sink module design [80] R.W. Knight, D.J. Hall, J.S. Goodling, R.C. Jaeger, Heat sink optimization with
problem in determining optimal non-uniform fin widths, Int. J. Heat Mass application to microchannels, IEEE Trans. Compon. Hybrids Manuf. Technol.
Transf. 67 (2013) 992–1006. 15 (5) (1992) 832–842.
[51] Cheng-Hung Huang, Yu-Hsiang Chen, An impingement heat sink module [81] D. Copeland, Optimization of parallel plate heat sinks for forced convection,
design problem in determining simultaneously the optimal non-uniform fin in: Proc. 16th IEEE Semit. Symp., 2000, pp. 266–272.
widths and heights, Int. J. Heat Mass Transf. 73 (2014) 627–633. [82] J.R. Culham, Y.S. Muzychka, Optimization of plate fin heat sinks using entropy
[52] Cheng-Hung Huang, Yu-Hsiang Chen, An optimal design problem in generation minimization, IEEE Trans. Compon. Packag. Technol. 24 (2) (2001)
determining non-uniform fin heights and widths for an impingement heat 159–165.
sink module, Appl. Therm. Eng. 63 (2014) 481–494. [83] N. Raja, R. Saidur, N.N.N. Ghazali, H.A. Mohammed, International Journal of
[53] H.R. Seyf, M. Feizbakhshi, Computational analysis of nanofluid effects on Heat and Mass Transfer Numerical study of thermal enhancement in micro
convective heat transfer enhancement of micro-pin-fin heat sinks, Int. J. channel heat sink with secondary flow, Int. J. Heat Mass Transf. 78 (2014)
Therm. Sci. 58 (2012) 168–179. 216–223.
[54] M.I. Hasan, Investigation of flow and heat transfer characteristics in micro pin [84] S.M. Kim, I. Mudawar, Analytical heat diffusion models for heat sinks with
fin heat sink with nanofluid, Appl. Therm. Eng. 63 (2) (2014) 598–607. circular micro-channels, Int. J. Heat Mass Transf. 53 (21–22) (2010) 4552–
[55] S.K. Mohammadian, Y. Zhang, Thermal management optimization of an air- 4566.
cooled Li-ion battery module using pin-fin heat sinks for hybrid electric [85] H.A. Mohammed, P. Gunnasegaran, N.H. Shuaib, Numerical simulation of heat
vehicles, J. Power Sources 273 (2015) 431–439. transfer enhancement in wavy microchannel heat sink, Int. Commun. Heat
[56] M. Liu, D. Liu, S. Xu, Y. Chen, Experimental study on liquid flow and heat Mass Transf. 38 (1) (2011) 63–68.
transfer in micro square pin fin heat sink, Int. J. Heat Mass Transf. 54 (25–26) [86] Y. Chen, C. Zhang, M. Shi, J. Wu, Three-dimensional numerical simulation of
(2011) 5602–5611. heat and fluid flow in noncircular microchannel heat sinks, Int. Commun.
[57] R.T. Huang, W.J. Sheu, C.C. Wang, Orientation effect on natural convective Heat Mass Transf. 36 (9) (2009) 917–920.
performance of square pin fin heat sinks, Int. J. Heat Mass Transf. 51 (9–10) [87] C.J. Kroeker, H.M. Soliman, S.J. Ormiston, Three-dimensional thermal analysis
(2008) 2368–2376. of heat sinks with circular cooling micro-channels, Int. J. Heat Mass Transf. 47
[58] C.-T. Chen, M.-H. Chen, W.-T. Horng, Reliability-based design optimization of (22) (2004) 4733–4744.
pin-fin heat sinks using a cell evolution method, Int. J. Heat Mass Transf. 79 [88] H.A. Mohammed, P. Gunnasegaran, N.H. Shuaib, Influence of channel shape
(2014) 450–467. on the thermal and hydraulic performance of microchannel heat sink, Int.
[59] H. Shafeie, O. Abouali, K. Jafarpur, G. Ahmadi, Numerical study of heat transfer Commun. Heat Mass Transf. 38 (4) (2011) 474–480.
performance of single-phase heat sinks with micro pin-fin structures, Appl. [89] R. Kuppusamy, H.A. Mohammed, C.W. Lim, Numerical investigation of
Therm. Eng. 58 (1–2) (2013) 68–76. trapezoidal grooved microchannel heat sink using nanofluids, Thermochim.
[60] Y. Peles, A. Kosßar, C. Mishra, C.J. Kuo, B. Schneider, Forced convective heat Acta 573 (2013) 39–56.
transfer across a pin fin micro heat sink, Int. J. Heat Mass Transf. 48 (17) [90] L. Chai, G. Xia, L. Wang, M. Zhou, Z. Cui, Heat transfer enhancement in
(2005) 3615–3627. microchannel heat sinks with periodic expansion-constriction cross-sections,
[61] W. Al-Sallami, A. Al-Damook, H.M. Thompson, A numerical investigation of Int. J. Heat Mass Transf. 62 (2013) 741–751.
thermal airflows over strip fin heat sinks, Int. Commun. Heat Mass Transf. 75 [91] L. Chai, G. Xia, M. Zhou, J. Li, Numerical simulation of fluid flow and heat
(2016) 183–191. transfer in a microchannel heat sink with offset fan-shaped reentrant cavities
[62] A. Al-Damook, N. Kapur, J.L. Summers, H.M. Thompson, Computational design in sidewall, Int. Commun. Heat Mass Transf. 38 (5) (2011) 577–584.
and optimisation of pin fin heat sinks with rectangular perforations, Appl. [92] G. Xia, Y. Zhai, Z. Cui, Numerical investigation of thermal enhancement in a
Therm. Eng. 105 (2016) 691–703. micro heat sink with fan-shaped reentrant cavities and internal ribs, ATE 58
[63] J. Zhao, S. Huang, L. Gong, Z. Huang, Numerical study and optimizing on micro (1–2) (2013) 52–60.
square pin-fin heat sink for electronic cooling, Appl. Therm. Eng. 93 (2015) [93] A.A. Alfaryjat, H.A. Mohammed, N.M. Adam, M.K.A. Ariffin, M.I. Najafabadi,
1347–1359. Influence of geometrical parameters of hexagonal, circular, and rhombus
H.E. Ahmed et al. / International Journal of Heat and Mass Transfer 118 (2018) 129–153 153

microchannel heat sinks on the thermohydraulic characteristics, Int. [110] H.A. Mohammed, P. Gunnasegaran, N.H. Shuaib, The impact of various nano
Commun. Heat Mass Transf. 52 (2014) 121–131. fluid types on triangular microchannels heat sink cooling performance, Int.
[94] L. Chai, G. Xia, M. Zhou, J. Li, J. Qi, Optimum thermal design of interrupted Commun. Heat Mass Transf. 38 (6) (2011) 767–773.
microchannel heat sink with rectangular ribs in the transverse [111] D. Lelea, I. Laza, The water based Al2O3 nanofluid flow and heat transfer in
microchambers, Appl. Therm. Eng. 51 (1–2) (2013) 880–889. tangential microtube heat sink with multiple inlets, Int. J. Heat Mass Transf.
[95] F. Hong, P. Cheng, Three dimensional numerical analyses and optimization of 69 (2014) 264–275.
offset strip-fin microchannel heat sinks, Int. Commun. Heat Mass Transf. 36 [112] E.M. Tokit, H.A. Mohammed, M.Z. Yusoff, Thermal performance of optimized
(7) (2009) 651–656. interrupted microchannel heat sink (IMCHS) using nanofluids, Int. Commun.
[96] C.J. Ho, L.C. Wei, Z.W. Li, An experimental investigation of forced convective Heat Mass Transf. 39 (10) (2012) 1595–1604.
cooling performance of a microchannel heat sink with Al2O3/water nanofluid, [113] H.R. Seyf, Z. Zhou, H.B. Ma, Y. Zhang, Three dimensional numerical study of
Appl. Therm. Eng. 30 (2–3) (2010) 96–103. heat-transfer enhancement by nano-encapsulated phase change material
[97] C.J. Ho, W.C. Chen, An experimental study on thermal performance of Al2O3/ slurry in microtube heat sinks with tangential impingement, Int. J. Heat Mass
water nano fluid in a minichannel heat sink, Appl. Therm. Eng. 50 (2013) Transf. 56 (1–2) (2013) 561–573.
516–522. [114] P. Naphon, L. Nakharintr, Turbulent two phase approach model for the
[98] M.R. Sohel, S.S. Khaleduzzaman, R. Saidur, A. Hepbasli, M.F.M. Sabri, I.M. nanofluids heat transfer analysis flowing through the minichannel heat sinks,
Mahbubul, An experimental investigation of heat transfer enhancement of a Int. J. Heat Mass Transf. 82 (2015) 388–395.
minichannel heat sink using Al2O3–H2O nanofluid, Int. J. Heat Mass Transf. 74 [115] T.-C. Hung, W.-M. Yan, X.-D. Wang, Y.-X. Huang, Optimal design of geometric
(2014) 164–172. parameters of double-layered microchannel heat sinks, Int. J. Heat Mass
[99] M. Keshavarz, R. Mohammadi, A. Ijam, CFD investigation of nano fluid effects Transf. 55 (11–12) (2012) 3262–3272.
(cooling performance and pressure drop) in mini-channel heat sink, Int. [116] B.H.S. Hamdi, E. Ahmed, M.I. Ahmed, Islam M.F. Seder, Experimental
Commun. Heat Mass Transf. 40 (2013) 58–66. investigation for sequential triangular double-layered microchannel heat
[100] M.K. Moraveji, R.M. Ardehali, CFD modeling (comparing single and two- sink with nanofluids, Int. Commun. Heat Mass Transf. 77 (2016) 104–115.
phase approaches) on thermal performance of Al2O3/water nano fluid in [117] K.-C. Wong, F.N.A. Muezzin, Heat transfer of a parallel flow two-layered
mini-channel heat sink, Int. Commun. Heat Mass Transf. 44 (2013) 157–164. microchannel heat sink, Int. Commun. Heat Mass Transf. 49 (2013) (2013)
[101] C.J. Ho, W. Chen, W. Yan, Correlations of heat transfer effectiveness in a 136–140.
minichannel heat sink with water-based suspensions of Al2O3 nanoparticles [118] B. Shao, L. Wang, H. Cheng, J. Li, Optimization and numerical simulation of
and/or MEPCM particles, Int. J. Heat Mass Transf. 69 (2014) 293–299. multi-layer microchannel heat sink, Procedia Eng. 31 (2012) 928–933.
[102] B. Fani, A. Abbassi, M. Kalteh, Effect of nanoparticles size on thermal [119] Y. Fan, P. Seng, L. Jin, W. Chua, D. Zhang, International Journal of Heat and
performance of nanofluid in a trapezoidal microchannel-heat-sink, Int. Mass Transfer A parametric investigation of heat transfer and friction
Commun. Heat Mass Transf. 45 (2013) 155–161. characteristics in cylindrical oblique fin minichannel heat sink, Int. J. Heat
[103] B. Rimbault, C.T. Nguyen, N. Galanis, Experimental investigation of CuO- Mass Transf. 68 (2014) 567–584.
water nanofluid flow and heat transfer inside a microchannel heat sink, Int. J. [120] C.J. Ho, Y.Z. Chen, F. Tu, C. Lai, Thermal performance of water-based
Therm. Sci. 84 (2014) 275–292. suspensions of phase change nanocapsules in a natural circulation loop
[104] T.-C. Hung, W.-M. Yan, X.-D. Wang, C.-Y. Chang, Heat transfer enhancement with a mini-channel heat sink and heat source, Appl. Therm. Eng. 64 (2014)
in microchannel heat sinks using nanofluids, Int. J. Heat Mass Transf. 55 (9– 376–384.
10) (2012) 2559–2570. [121] C.J. Ho, Y.N. Chung, C. Lai, Thermal performance of Al2O3/water nanofluid in
[105] H.A. Mohammed, P. Gunnasegaran, N.H. Shuaib, Heat transfer in rectangular a natural circulation loop with a mini-channel heat sink and heat source,
microchannels heat sink using nanofluids, Int. Commun. Heat Mass Transf. 37 Energy Convers. Manage. 87 (2014) 848–858.
(10) (2010) 1496–1503. [122] S. Ayub, W. Ali, A. Maryam, Water cooled minichannel heat sinks for
[106] Y. Yang, K. Tsai, Y. Wang, S. Lin, Numerical study of microchannel heat sink microprocessor cooling: effect of fin spacing, Appl. Therm. Eng. 64 (2014) 76–
performance using nano fluids, Int. Commun. Heat Mass Transf. 57 (2014) 82.
27–35. [123] H.E. Ahmed, M.I. Ahmed, Optimum thermal design of triangular, trapezoidal
[107] S. Halelfadl, A. Mohammed, N. Mohd-ghazali, T. Maré, P. Estellé, R. Ahmad, and rectangular grooved microchannel heat sinks, Int. Commun. Heat Mass
Optimization of thermal performances and pressure drop of rectangular Transf. 66 (2015) 47–57.
microchannel heat sink using aqueous carbon nanotubes based nano fluid, [124] N.A.C. Sidik, M.N.A.W. Muhamad, W.M.A.A. Japar, Z.A. Rasid, An overview of
Appl. Therm. Eng. 62 (2) (2014) 492–499. passive techniques for heat transfer augmentation in microchannel heat sink,
[108] S.M. Peyghambarzadeh, S.H. Hashemabadi, A.R. Chabi, M. Salimi, Int. Commun. Heat Mass Transf. 88 (2017) 74–83.
Performance of water based CuO and Al2O3 nanofluids in a Cu–Be alloy [125] M. Reza, S. Bushehri, H. Ramin, M. Reza, International Journal of Thermal
heat sink with rectangular microchannels, Energy Convers. Manage. 86 Sciences A new coupling method for slip- fl ow and conjugate heat transfer in
(2014) 28–38. a parallel plate micro heat sink, Int. J. Therm. Sci. 89 (2015) 174–184.
[109] C. Chen, C. Ding, Study on the thermal behavior and cooling performance of a [126] W. Pang, Y. Liu, S. Shao, X. Gao, Empirical study on thermal performance
nano fluid-cooled microchannel heat sink, Int. J. Therm. Sci. 50 (3) (2011) through separating impacts from a hybrid PV/TE system design integrating
378–384. heat sink, Int. Commun. Heat Mass Transf. 60 (2015) 9–12.

You might also like