Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Applied Thermal Engineering 161 (2019) 114130

Contents lists available at ScienceDirect

Applied Thermal Engineering


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

Research Paper

Numerical study of TiO2-based nanofluids flow in microchannel heat sinks: T


Effect of the Reynolds number and the microchannel height
Víctor A. Martínez, Diego A. Vasco , Claudio M. García-Herrera, Roberto Ortega-Aguilera

Departamento de Ingeniería Mecánica, Universidad de Santiago de Chile, USACH, Av. Bernardo O’Higgins 3363, Santiago de Chile, Chile

HIGHLIGHTS

• Effect of temperature and TiO wt% on thermal conductivity and viscosity was tested.
• Both
2
properties variation with temperature and TiO wt% concentration was modeled.
• Laminar
2
flow of nanofluids in microchannels heat sink was computationally simulated.
• Nanofluids are more effective at low Reynolds numbers in microchannel heat sinks.
• Effect of the temperature-dependent properties is evident near to the heated wall.

ARTICLE INFO ABSTRACT

Keywords: In the present work, we have studied the laminar (200 Re 1200) three-dimensional flow of a water-based
Nanofluids nanofluid in rectangular microchannels of constant wide cross section (283 µ m) and three different heights
Microchannel heat sink (800 µ m, 600 µ m, and 400 µ m). Moreover, we analyse the concentration of dispersed TiO2 nanoparticles
Thermal conductivity (<6 nm) on the thermal and hydraulic performance of heat sinks with the microchannels. The numerical analysis
Viscosity
required the use of experimental models of the viscosity and the thermal conductivity with respect to tem-
perature TiO2 nanoparticles concentration (1 wt% and 3 wt%). Using a code based on the finite volume method
to solve the transport equations that govern the studied flows, the friction factor, the Nusselt number, the
convective heat transfer coefficient, and the mean temperature of the hot wall have been determined. The results
show that both the use of nanofluids and the reduction of the microchannel height favour heat transfer at low
Reynolds numbers, an improvement that decreases as this parameter increases. With a nanoparticles con-
centration of 3 wt% and a Reynolds of 200, we found a maximum 19.66% increase of the convective heat
transfer coefficient with respect to the pure base fluid. However, also at low Reynolds, the greatest increases of
the skin friction coefficient were registered, reaching a maximum increase of 137.68% with respect to the pure
base fluid in the case of the 1 wt% nanofluid.

1. Introduction traditional heat removal systems such as the method of cooling by air,
which in spite of their multiple advantages (high reliability, low initial
The development of high power and speed electronic devices re- maintenance and operating cost) became limited from the power dis-
quires the use of large-scale integrated circuits (VLSI), which implies sipation standpoint [3], with values between 10 W/cm2 and 50 W/cm2
the increase in highly focused thermal loads as a result of the minia- [4]. Based on the definition of the Nusselt number (Nu), two ap-
turization that they undergo. These charges must be dissipated effi- proaches appear that allow increasing the heat removal rate: (i) Opti-
ciently, since they are detrimental to their optimal performance [1]. mizing the relation between the heat transfer surface and the volume of
The above represents a challenge from the standpoint of heat man- the heat exchanger, decreasing its characteristic length, or (ii)
agement, particularly when it is foreseen that the generation of 14 nm mproving the thermal conductivity of the system’s cooling fluid [5].
chips would demand a heat removal of 100 W/cm2 [2], so compact and Tuckerman and Pease [1] were the first to study experimentally the
efficient cooling systems would be required in a way that can provide heat transfer from a heat sink made of silicone subjected to conditions
reliable functioning of those systems. Considering the above, and since of laminar flow, succeeding in dissipating a power density of


Corresponding author.
E-mail address: diego.vascoc@usach.cl (D.A. Vasco).

https://doi.org/10.1016/j.applthermaleng.2019.114130
Received 14 December 2018; Received in revised form 7 June 2019; Accepted 13 July 2019
Available online 15 July 2019
1359-4311/ © 2019 Elsevier Ltd. All rights reserved.
V.A. Martínez, et al. Applied Thermal Engineering 161 (2019) 114130

Nomenclature W microchannel width [m]

Cf skin friction factor Greek symbols


Cp specific heat [J/kg K]
Dh equivalent hydraulic diameter [m] mass concentration of dispersed nanoparticles
Err relative error [%] viscosity [Pa·s ]
H microchannel height [m] viscous dissipation function
h coefficient of heat transfer by convection [W/m2 K ] density [kg/m3]
k thermal conductivity [W/mK] shear stress [N/m2 ]
L length of microchannel [mm]
m mass [g] Subscripts
Nu Nusselt number
u velocity [m/s] bf base fluid
p pressure [Pa] c cold
q heat flux [W/m2 ] np nanoparticle
Re Reynolds number nf nanofluid
T temperature [K] w wall
x , y, z Cartesian coordinates

790 W/cm2 with channels whose hydraulic diameter was 85.8 µ m and nanoparticles did not contribute to extra heat absorption at high fluid
using water as refrigerant. These authors concluded that the use of flow rates.
microchannels, ducts whose characteristic length is less than 1000 µ m To determine the effect of the height of the microchannels on the
[6], is appropriate as a dissipation system. heat transfer, Manay et al. [21] investigated experimentally the process
On the other hand, to improve heat transfer, it is currently possible of heat transfer by forced convection taking place in a microchannel of
to increase the thermal conductivity of the working fluid by adding low variable height subjected to a constant heat flow of 80 kW/m2 in the
concentrations (<1 vol%) of metallic and nonmetallic nanoparticles. bottom surface, through which a nanofluid composed of TiO2 nano-
This nanoparticle/base fluid composite system was named nanofluid by particles (0.25–2.0 vol%) and deionized water at concentrations
Choi et al. [7] and they have shown the ability of increasing the thermal flowed. From the tests, they determined that regardless of the con-
conductivity of the base fluid [8–10], presenting a strong relationship centration of dispersed nanoparticles, the use of nanofluids in general
between the magnitude of this improvement and the concentration of achieves higher heat transfer rates than pure water, but they estab-
dispersed nanoparticles [11–13], the size of the nanoparticles [14,15] lished that the heat transfer cannot be increased in microchannels with
and the temperature [16,17]. Therefore, to increase the heat removal water-based nanofluids whose concentration of dispersed nanoparticles
rate of a dissipation system, a reasonable strategy is to implement the was equal to or greater than 2.0 vol%, as the increase in the viscosity of
flow of nanofluids through microchannels. Sarafraz et al. [18] studying the nanofluid intensifies the viscous sublayer. On the other hand, they
experimentally the thermal performance of a heat sink through mi- determined that the height of the microchannel has a direct impact on
crochannels with rectangular cross-section (0.25 × 0.4 mm2) using na- the pressure drop, causing a 73% increase when the height was de-
nofluids synthesized with silver (Ag) nanoparticles dispersed in water, creased from 300 µ m to 200 µ m. However, the heat transfer coefficient
found that for a concentration of 0.1 wt% the system registers a 47% also increased 30–45% when the height of the microchannel was re-
improvement of the convective heat transfer coefficient when subjected duced from 500 µ m to 200 µ m.
to a constant heat flow at its base. However, due to the presence of Although the experimental results available in relation to the im-
nanoparticles in the fluid, there was also an increased pressure drop proved performance of heat sinks with microchannels due to the use of
with respect to the concentration of dispersed nanoparticles, although nanofluids are promising, this information is still limited and dispersed
for low Reynolds values (100 < Re < 350). The difference between because the experimental conditions differ, so to go deeper into un-
the pressure drop recorded with water compared to that obtained with derstanding the behavior of the flow as well as the heat transfer in the
the nanofluid at a 0.01 wt% concentration was less than 7%. microchannels it is basic to carry out numerical simulations of this
Peyghambarzadeh et al. [19] studied experimentally the heat issue. To determine the effect of the use of nanofluids on heat transfer
transfer by forced convection in a heat sink with 17 microchannels of as well as on characteristics of the flow in a rectangular microchannel of
rectangular cross section and dimensions of 400 × 560 µ m2, using two 430 × 280 µ m2, Mohammed et al. [22] studied numerically the perfor-
nanofluids as working fluids: Al2O3/water y CuO/water, forcing a la- mance of the dissipator with microchannels using nanofluids synthe-
minar flow (500 < Re < 2000) through the heat sink, which was sup- sized with alumina (Al2O3) nanoparticles suspended in water, at con-
plied with a constant heat flow of 19 W/cm2 . Under these conditions centrations of 1–5 vol% The numerical simulation was performed by
improvements of up to 49% were achieved in the convective heat discretizing the governing equations by means of the finite volume
transfer coefficient with respect to water, using a 1 vol% concentration method, with the hybrid discretization scheme, using the SIMPLE al-
of a Al2O3/water nanofluid, while using CuO/water (0.2 vol%) this gorithm to solve the velocity fields, showing that although the nano-
improvement amounted to 27%. An interesting result reported in this fluids are capable of improving the heat transfer coefficient of a dis-
paper is the relation between the Reynolds number and the heat sipator with microchannels, it depends on the volumetric fraction of
transfer performance, because increasing the Re values that perfor- nanoparticles suspended in the base fluid, registering a maximum at
mance decreases, and that is attributed by the authors to the accumu- 2.5 vol%, achieving it also with a slight increase of the friction factor.
lation and deposition of nanoparticles in the microchannels. Similar Similar results were reported by Abdollahi et al. [23] who im-
results were those reported by Chein et al. [20], who by studying ex- plementing various aqueous base nanofluids with different concentra-
perimentally the flow of nanofluids synthesized with copper nano- tions (0–2 vol%) as cooling fluid of a heat dissipator with micro-
particles dispersed in water (0.2–0.4 vol%) through microchannels of channels of rectangular cross section, found that, in general, an increase
trapezoidal cross section, determined that the addition of nanoparticles of the dispersed nanoparticles increases the Nusselt number, but to a
to water improves the heat absorption of the coolant, but the smaller extent, with the decrease of the average diameter of the

2
V.A. Martínez, et al. Applied Thermal Engineering 161 (2019) 114130

nanoparticles. It should be noted that in contrast with Mohammed et al. 2.3. Thermal conductivity measurement
[22], who only implemented static models to determine the thermal
conductivity of the studied nanofluid, Abdollahi et al. [23] determined The heat transfer process studied in the numerical simulation was
the thermal conductivity of the studied nanofluid by means of a model performed considering the nanofluids’ properties variable with tem-
that does consider the role of Brownian motion on this property [24], in perature, which requires determining the behavior of the thermal
addition to its static portion, adding both of them. Using this model, conductivity the nanofluid with respect to this variable, generating a
these authors considered in part the effect of the temperature on the model that describes such behavior. The nanofluids’ thermal con-
thermal conductivity, but not the one it has on the viscosity of the ductivity measurements have been developed using the Decagon
nanofluid, because the properties of the base fluid were assumed to be Devices thermal properties analyzer model KD2-Pro, with the KS-1 li-
constant. Shi et al. [25] simulated a laminar flow of water-based na- quid sample sensor, whose principle of operation corresponds to the hot
nofluids with nanoparticles of Al2O3 in a rectangular cross section mi- wire method, so it requires the measured sample to remain at constant
crochannel (215 × 341 µ m2) considering thetemperature-dependent temperature and devoid of natural convection. The measurements of
properties of the base fluid, and the Maxwell’s model to describe the the thermal conductivity are performed using the assembly illustrated
nanofluid’s thermal conductivity. Furthermore, this work proposes a in Fig. 2, for six nominal temperature levels, from 278 K to 303 K, with
simulation model that allows capturing the distribution of the nano- five temperature steps of 5 K between each level and making five
particle concentration in the flow field, which was found to be not measurements of the property per temperature level.
uniform, allowing the estimation with greater precision of the dis-
tribution of the nanofluid’s thermal conductivity in the studied domain. 2.4. Viscosity measurement
The objective of the present article is to study numerically the effect
of the height of a microchannel on the thermal performance of a heat To determine the behavior of viscosity with respect to the tem-
sink subjected to a constant heat flow of 50 W/cm2 through the bottom perature of the nanofluids, a Brookfield model DV2T-LV viscosimeter
surface. To that end, we considered the effect of the temperature on the was used. The 7 mL sample was placed in the SC4-13R small sample
thermal conductivity and the viscosity of nanofluids prepared with adapter, which is connected to the circulating bath to control the
nanoparticles of TiO2 (<6 nm) suspended in water at concentrations of sample’s temperature (Fig. 3). The SC14-18 spindle is used for mea-
1 wt% and 3 wt% using the two-step method. The mathematical models suring the deformation rate (1/s), sheer stress (N/mm2), and visc-
that describe these properties were incorporated in a validated code osity (Pa·s), evaluating them for five temperature levels from 283 K to
based on the finite volume method, with which a laminar three-di- 323 K, with four temperature steps of 10 K between each level, using
mensional flow study was made for Reynolds numbers between 200 and variable spindle rotational speeds from 70 RPM to 100 RPM. It should
1200. Three rectangular microchannels with 283 µ m wide cross-section be noted that for each the measurements were made in triplicate.
and heights of 800 µ m, 600 µ m, and 400 µ m were analysed.
3. Numerical analysis

2. Experimental procedure 3.1. Geometry, governing equations and boundary conditions

2.1. Materials A heat sink has been considered that has symmetric rectangular
microchannels, shown schematically in Fig. 4, where the selected
The studied nanofluids were synthesized through the dispersion of computational domain is bound to one of its own cavities (Fig. 5). The
titanium dioxide (TiO2) nanoparticles in the anatase crystal phase, with geometric dimensions of the microchannels are given in Table 1, where
99% purity, average diameter of 6 nm [26] and specific surface area of the width, the height, and the length of microchannel 1 (MCH-1) were
356 m2/g, which was supplied by Mknano (Toronto, Canada). The base determined based on the experimental work of Ho et al. [28], while for
fluid used was double distilled water. microchannels 2 (MCH-2) and 3 (MCH-3), although the width and
lengths are those used by these authors, the heights are 600 µ m and
2.2. Preparation of nanofluids 400 µ m, respectively. A constant heat flux is applied to the bottom wall
of the domain, while the top wall was kept at a constant temperature of
The preparation of the nanofluids used as working fluids in the heat 293.15 K. The remaining walls are considered to be perfectly insulated.
sink have been synthesized by the two-step ultrasonication method In the present study the conservation of mass, momentum, and
(Fig. 1). The volume of base fluid to be used has been defined (115 mL energy equations (considering the viscous dissipation term) in their
for measuring the thermal conductivity and 7 mL for measuring the
viscosity), to be then weighed on an analytical balance (Radwag, model
AS 82/220.R2), from which, after defining the concentration of interest
(1 wt% and 3 wt%), the mass of nanoparticles to be dispersed has been
determined through Eq. (1).
· mbf
mnp =
(1 ) (1)

Once the adequate mass of nanoparticles for the chosen con-


centration has been determined, they were dispersed in the base fluid
by means of a vertical axis ultrasonic probe (Dr. Hielscher GmbH,
model UP50H) whose working frequency is 30 kHz, exposing the
samples to a continuous ultrasonic irradiation process for 60 min, a
time that allows optimizing the conductivity of water due to the addi-
tion of TiO2 nanoparticles [27,26]. The sonication process was carried
out at constant temperature, connecting the container in which the
sample is deposited with a cooling circulation bath (JSR, model JSRC- Fig. 1. Experimental ultrasonication setting. (1) Cooling circulation bath. (2)
13C). Ultrasonic probe. (3) Nanofluid.

3
V.A. Martínez, et al. Applied Thermal Engineering 161 (2019) 114130

Fig. 5. Computational domain.

Fig. 2. Experimental setup for measuring the thermal conductivity. (1) Table 1
Circulating cooling bath. (2) Thermal properties analyzer. (3) KS1 sensor. Geometric dimensions of the studied microchannels.
Microchannel H µm W µm L mm

MCH-1 800 283 50


MCH-2 600 283 50
MCH-2 400 283 50

Table 2
Properties of titanium dioxide nanoparticles
[33].
Cp , J/kg K , kg/m3

710 3870

Table 3
Grid independence study.
MCH-1 MCH-2 MCH-3

Mesh Nu Error % Nu Error % Nu Error %

Fig. 3. Experimental setup for measuring the viscosity. (1) Circulating cooling 34 × 18 × 146 7.29522 – 6.79037 – 5.90315 –
bath. (2) Viscosimeter. (3) Nanofluid. (4) Spindle. (5) Cooling fluid. 50 × 26 × 218 7.69968 5.54419 7.19901 6.01799 6.43436 8.99863
58 × 30 × 254 7.87761 2.31092 7.39418 2.71108 6.73785 4.71679

( uk ui ) p ui
= +
xk xi xk xk (3)
Conservation of energy:

( ui T ) k T
= +
xi x i cp x i cp (4)
where is the viscous dissipation function and is defined by Eq. (5).

ui uj ui
= +
xj xi xj (5)
The associated boundary conditions are:
Fig. 4. Schematic of a heat sink with microchannels. Re·
u= at z=0
Dh (6)
traditional form (Eqs. (2)–(4)) comprise the mathematical model used
ui = 0 at x = 0, x = H, y = 0, y=W (7)
for predicting the behavior of the flow of the nanofluid, which is con-
sidered incompressible, Newtonian, and single phase. ui = 0 at x = 0, x = H, y = 0, y=W (8)
Conservation of mass:
T = Tc at z = 0, y=H (9)
( ui )
=0 T
xi (2) q = knf at y=0
y (10)
Conservation of momentum: q = 0 at x = 0, x = H, y = 0, y=W (11)

4
V.A. Martínez, et al. Applied Thermal Engineering 161 (2019) 114130

Fig. 6. Validation of the FVM code according to (a) the Nusselt number and (b) the mean friction factor obtained by Ho et al. [28].

Fig. 7. (a) Thermal conductivity, (b) relative thermal conductivity, (c) viscosity and (d) relative viscosity with respect to temperature of pure base fluid [37] and
nanofluids with 1 wt% and 3 wt%.

For the relative thermal conductivity (Fig. 7b), it was fitted to Eq. (13),
3.2. Thermophysical properties of nanofluids which is a linear model proposed by Palm et al. [30]. Both models have
been selected considering that they have been used to describe the
The thermophysical properties of the base fluid (double-distilled viscosity and thermal conductivity of nanofluids prepared with oxide-
water) are modified by the addition of TiO2 nanoparticles in mass based nanoparticles.
concentrations of 1 wt% and 3 wt%. According to this, the behavior of
the relative viscosity and relative thermal conductivity with respect to
nf
= a 0 + a1 T + a2 T 2
temperature of the prepared nanofluids has been determined. The data bf (12)
of relative viscosity has been fitted to Eq. (12), which is presented in
Fig. 7d and corresponds to the model proposed by Nguyen et al. [29].

5
V.A. Martínez, et al. Applied Thermal Engineering 161 (2019) 114130

Table 4 The properties of the base fluid have been modeled with respect to
Results of the fittings parameters of the viscosity and thermal conductivity temperature according to Eqs. (16)–(19) proposed by Zografos et al.
according to Eqs. (12) y (13). Standard error (SE) and determination coefficient [32] at atmospheric pressure and a temperature range of 273.2 K and
(R2). 600 K, while those of the nanoparticles are considered constant
Property Parameter S.E. R2 (Table 2) [33]. The thermophysical properties of the nanofluids have
been determined by applying a correction to the thermal conductivity
1%wt nf / bf a0 -35.12478 0.00913 0.99420 and the viscosity of the base fluid, which depend on temperature.
a1 0.23559
a2 −3.83676 × 10 4
bf = 3.8208 × 10 2 (T 252.33) 1
(16)
knf /kbf b0 0.07071 0.00897 0.93874
b1 3.17364 × 10 3
kbf = 4.2365 × 10 9T 3 1.1440 × 10 5T 2 + 7.1959 × 10 3T 0.63262

multirow590pt3%wt a0 −9.95202 0.00336 0.99961 (17)


nf / bf
a1 0.07658 6T 3 4T 2
bf = 3.0115 × 10 + 9.6272 × 10 0.11052T + 1022.4 (18)
a2 −1.32743 × 10 4

knf /kbf b0 −1.09226 0.01366 0.93817


b1 7.22651 × 10 3 Cp, bf = 1.7850 × 10 7T 3 1.9149 × 10 4T 2 + 6.7953 × 10 2T 3.7559
(19)

knf
= b0 + b1· T 3.3. Data processing
kbf (13)
where a0 , a1, a2 , b0 , and b1, are fitting constants that have been de- The computational domain has been subjected to a constant heat
termined for each group of measurements, which are only valid for the flux in the lower surface so the Nusselt number can be determined by
corresponding concentration level. Both the density (14) and the spe- the Eq. (20).
cific heat (15) of the nanofluid have been calculated considering the q Dh
conventional solid/liquid mixtures theory [31]. Nu =
knf (Tw Too) (20)
nf = np + bf (1 ) (14)
The coefficient of heat transfer by convection is given by (21).
np Cp, np + bf Cp, bf (1 ) Nu·knf
Cp, nf = h=
np + bf (1 ) (15) Dh (21)

Fig. 8. Variation of the Nusselt number for (a) MCH-


1, (b) MCH-2, (c) MCH-3, and (d) effect of the mi-
crochannel’s height for = 1%wt.

6
V.A. Martínez, et al. Applied Thermal Engineering 161 (2019) 114130

Fig. 9. Variation of the coefficient of heat transfer by convection for (a) 1 wt% and (b) 3 wt%.

Fig. 10. Variation of the relative Nusselt number for


(a) MCH-1, (b) MCH-2, (c) MCH-3, and (d) effect of
the microchannel’s height for = 1 wt%.

The friction factor is calculated by the Eq. (22). volume method (FVM) has been used for the resolution of the equations
w
of continuity, momentum, and energy [34]. To solve the pressure-speed
Cf = coupled system, the SIMPLE coupling algorithm has been used, where
1/2 nf U2 (22)
the applied sub-relaxation parameters are 0.5 for the speed, 0.3 for the
where pressure, and 0.1 for the temperature. On the other hand, the power-
law scheme is applied to discretize the diffusive and convective terms of
u
w = ·
(23) the transport equations involved. In the present study, it has been
x
considered that a solution converges when the maximum residual mass
is less than 10−6 for the speed in the x and y direction and less than
3.4. Numerical method 10−12 for the speed in the z-direction, while the criterion of con-
vergence used for the energy equation has been 10−6.
A home-made Open-MP parallelized code based on the finite

7
V.A. Martínez, et al. Applied Thermal Engineering 161 (2019) 114130

Fig. 11. Variation of the mean temperature of the


bottom wall for (a) MCH-1, (b) MCH-2, (c) MCH-3
and (d) effect of microchannel height for = 1 wt%.

3.5. Grid independence and code validation 4. Results and discussion

A structured grid with rectangular hexahedral elements was used to 4.1. Experimental results
discretize the domain. A grid independence study was carried out based
on the average Nusselt number, to determine the number of elements From the measurements of the thermal conductivity as well as of the
that produce precise results with the lowest requirement of computa- viscosity of nanofluids prepared at a mass concentration of 1 wt% and
tional resources. This test has been carried out with four grids im- 3 wt%, the curves are obtained with respect to the temperature of those
plemented in the MCH-1 case, with Re = 500 and q = 100 W/cm2. relative properties (Fig. 7b and d). These results show that compared to
Based on the results shown in Table 3, and since the difference between the base fluid, an average increase of 1.62% in thermal conductivity is
the Nusselt number achieved with the third and the second grids is registered for the nanofluid prepared with a concentration of 1 wt%, an
lower than 5%, we decided to use a grid with 50 nodes in the direction improvement that rises to 5.65% for that synthesized at a concentration
x, 26 in the y, direction and 218 in the z direction. of 3 wt% of TiO2 nanoparticles. Regarding the effect of temperature on
To establish the validity of the results obtained with the developed this property, it is found that although between 275 K and 290 K a
numeric code, a comparison has been made with the experimental work combined effect is seen, that is, an increase and decrease of thermal
of Ho et al. [28], based on the Nusselt number and the Fanning friction conductivity with respect to temperature, above 290 K the behavior
factor obtained under the same flow conditions with a constant heat becomes increasing until 305 K, temperature at which the last mea-
flow of 100 W/cm2 on the bottom of the microchannel, using water as surement was carried out. Regarding the measurements made of the
the cooling fluid. Based on the results shown in Fig. 6, it can be es- viscosity of the studied nanofluids, it is again found that there is a
tablished that the numerical results of the Nusselt number present a combined effect with respect to temperature (increase and decrease),
lower mean relative error of 3.78%, when the Reynolds number is with very insignificant increases in the magnitude of this property with
lower than 900. While at higher Reynolds numbers (Re > 900), the respect to the base fluid, recording a maximum increase of 2.86% at
mean relative error increases to 11%. Regarding the fanning friction 313.6 K for the nanofluid synthesized at a concentration of 1 wt%,
factor a similar tendency was observed, the mean relative error is while for that with a concentration of 3 wt% the maximum increase
9.79% for Re < 900, and 16.37% for Re > 900. The found difference of registered was 9.45% at a temperature of 293.4 K. It should be noted
the mean relative errors may be explained by the acquired relevance of that the behavior exhibited by the effective viscosity of the studied
the turbulent effects at higher Reynolds numbers [35]. nanofluids does not imply that the temperature does not have an effect,
but simply that it affects the viscosity of both water and the nanofluid
[36]. Therefore, an increase in the effective viscosity with respect to the
temperature is explained by a predominance of the effect of this

8
V.A. Martínez, et al. Applied Thermal Engineering 161 (2019) 114130

Fig. 12. Variation of the Fanning friction coefficient for (a) MCH-1, (b) MCH-2, (c) MCH-3 and (d) effect of the height of the microchannel for 1 wt%.

Table 5 4.2. Numerical results


Results of the fittings parameters of relative Nusselt number according to Eq.
(24). Standard error (SE) and determination coefficient (R2). To study the heat transfer process that occurred in a heat sink with
Microchannel A B C S.E. microchannels and the combined effect of the height of the micro-
R2
channel and that of the concentration of nanoparticles (TiO2) dispersed
MCH-1 1%wt 0.29020 311.73249 0.98221 0.00353 0.99781 in water-based nanofluids, we have carried out simulations of three
3%wt 0.20119 753.05881 0.98295 0.00557 0.99126 microchannels (MCH-1, MCH-2 and MCH-3) subjected to a constant
MCH-2 1%wt 0.18615 309.94451 0.99803 0.00189 0.99847 heating rate of 50 W/cm2, using 1 wt% and 3 wt% mass concentrations
3%wt 0.15856 425.65229 1.03999 0.00090 0.99963 of nanofluids as cooling fluids with variable properties with respect to
MCH-3 1%wt 0.16503 398.80714 0.99265 0.00254 0.99717 temperature. The nanofluid enters the microchannels at a temperature
3%wt 0.18280 599.43224 1.01311 0.00332 0.99643 of 293.15 K, with a constant Reynolds number (100 Re 1200),
which was studied for each microchannel and nanoparticle concentra-
tion level, obtaining for each case the corresponding velocity and
variable on the viscosity of water in comparison with the effect on the temperature field, from which the Nusselt number, the convective heat
viscosity of the nanofluid, even though individually the viscosity of transfer coefficient, the Fanning friction factor, and the average surface
both the water and the studied nanofluids decreases strictly with the temperature have been determined (Figs. 8–12). To appreciate the ef-
temperature increase (Fig. 7c). Similarly, the above reasoning may fect of the microchannel’s height, the simulation performed with the
explain the behavior registered by the effective thermal conductivity 1 wt% nanofluid concentration has been selected arbitrarily. From the
where, despite registering a decrease with respect to temperature be- results shown in Fig. 8a–c, it is corroborated that for the three micro-
tween 285 K and 290 K, individually the thermal conductivity of both channel heights studied, the Nusselt number presents a strictly in-
the base fluid as well as of the nanofluids showed a strictly increasing creasing behavior with respect to the Reynolds number for all the na-
behavior with respect to temperature (Fig. 7a). noparticle concentrations, in addition to the base fluid, registering in
The experimental results were fitted to Eqs. (12) and (13) as ap- the latter case a magnitude of 5.15 for Re = 200, and 10.58 for
propriate, in the temperature interval between 290 K and 305 K for Re = 1200, the above in the highest microchannel (MCH-1). In general,
relative thermal conductivity and 290 K and 325 K for relative viscosity, it is seen that the use of nanofluids as cooling fluids promotes the heat
which correspond to the operating temperatures of the studied micro- transfer in microchannels, regardless of their height. However, this
channels. The curve fitting of the experimental data has provided the improvement is not sustained with respect to the Reynolds number; in
constants of each correlation, whose values are given in Table 4. fact, as shown in Fig. 8a, for microchannel MCH-1 the addition of na-
noparticles at 1 wt% concentration increases the Nusselt number by

9
V.A. Martínez, et al. Applied Thermal Engineering 161 (2019) 114130

13.87% compared to the base fluid for Re = 200, an improvement that cooling fluid does not produce a significant increase in the Nusselt
decreases to less than 5% when it is considered that the Nusselt number number, since only a 36.79% of the maximum Nusselt number im-
is independent of the nanoparticle concentration, which in this case provement (A) has been reached. Therefore, parameter B is a mea-
occurs starting from Re = 500, while for the case of the nanofluid with surement of the effectiveness of using nanofluids based on TiO2 and
a 3 wt% of dispersed nanoparticles this phenomenon is recorded from water in microchannel heat sinks. Despite, the height of the micro-
Re = 1000, in spite of getting a 14.24% increase at Re = 200 with re- channel affects the value of parameter B, it has not been possible to
spect to the base fluid. The above suggests that the use of nanofluids as verify a clear relationship between them.
cooling fluids is desirable from the view point of heat transfer when the Another variable to consider when evaluating the performance of
microchannel operates at low Reynolds numbers, that is, the addition of nanofluids as cooling agents in microchannels is the average tempera-
TiO2 nanoparticles increases heat transfer significantly when the pre- ture of the bottom wall, a variable presented in Fig. 11a–c for each
dominant heat transfer mechanism is conduction, an improvement that nanoparticle concentration with respect to the Reynolds number,
is not relevant when convection predominates in the heat transfer. A showing that for increasing this parameter, the average temperature of
similar behavior is seen in Fig. 8b and c, where the Nusselt number for the bottom of the microchannels decreases strictly, decreasing with the
the nanofluids with 1 wt% and 3 wt% becomes independent of the addition of nanoparticles to the base fluid, which is most significant at
concentration of dispersed nanoparticles for Reynolds 400 and 1200, low Reynolds numbers for all the microchannels studied. That is how,
respectively, in the case of microchannel MCH-2, and for Re = 500 and for example, for the microchannel with the greatest height (MCH-1),
Re = 1000, respectively, in the case of microchannel MCH-3. the addition of nanoparticles at a 1 wt% concentration induces a de-
Regarding the influence of the microchannel’s height on the Nusselt crease of the average temperature of the base of the microchannel of
number, Fig. 8d shows that a decrease of the height negatively affects 8.08 K at Re = 200 with respect to the base fluid, and a decrease of only
the Nusselt number, a difference that becomes greater as the Reynolds 0.33 K by increasing the Reynolds to 1200. A similar behavior is re-
number increases. That is why for Re = 200, decreasing the height of corded for the synthesized nanofluid with a 3 wt% concentration
the microchannel by 200 µ m reduces the Nusselt number by 5.12% a where, referred to the base fluid, the nanofluid produces a decrease of
difference that increases until it reaches a maximum of 9.97% for the average temperature of the bottom wall of the microchannel of
Re = 1000. Similarly, for the same nanofluid, decreasing the height of 10.78 K for Re = 200, a decrease that is only 1.72 K when Re = 1200.
the microchannel by 400 µ m reduces the Nusselt number by 8.33% for The behavior described above is also seen for microchannels MCH-2
Re = 200, a difference that increases until it reaches a maximum of and MCH-3, from which it can be established that although the addition
18.52% for Re = 1000. The above indicates that decreasing the height of nanoparticles to water decreases the temperature of the bottom wall
of a microchannel and therefore Dh inecessarily implies reducing the of the at equal Reynolds numbers, this improvement is more significant
convective influence of the studied heat transfer process, giving greater for low Reynolds numbers, that is, when conduction is prevalent in the
importance to conduction. It should be noted that the above does not heat transfer. However, further to the above, it is seen that the lower
entail a decrease of the transferred heat, on the contrary, as shown in temperatures of the bottom wall of the microchannels were recorded, in
Fig. 9a and b, the convective heat transfer coefficient in general in- all cases, for Re = 1200, that is, when neither the addition of nano-
creases with the reduction of the equivalent hydraulic diameter, particles nor the height of the microchannel are important, as shown in
reaching 2.91% in the case of the nanofluid with a 1 wt% concentration Fig. 11d, where for Re= 1200 and referred to MCH-1, decreasing by
when the height of the microchannel is decreased by 200 µ m, an im- 200 µ m and 400 µ m the height of the microchannel only produces a
provement that reaches 9.43% when the height is reduced to 400 µ m. temperature decrease of 0.40 K and 0.85 K, respectively.
The Nusselt number is proportional to Dh , which is reduced successively Fig. 12a–c show the behavior of the Fanning friction coefficient with
by 8.01% and 13.81% for the microchannel 2 and 3, respectively. In respect to the Reynolds number for the three studied microchannels,
this way, since the increase of the convective heat transfer coefficient is presenting in each of them the three fluids used for cooling, to verify
less than the decrease of Dh , the Nusselt number increases with the the influence of the concentration of nanoparticles on that variable.
height of the microchannel, a conclusion that fully fits that found in the These figures show a decreasing behavior of the coefficient of the
available literature [38,39,21]. Fanning friction coefficient with respect to the Reynolds number, with
Considering the behavior of the Nusselt number with respect to the an asymptotic tendency in all the modeled microchannels and con-
Reynolds number and the effect of the concentration of nanoparticles, centration levels of nanoparticles, suggesting that this variable is in-
in Fig. 10, the relative Nusselt number is exposed, that is, the ratio dependent of the Reynolds number when it is sufficiently high. As re-
between the Nusselt numbers obtained with the nanofluid and the base gards the addition of nanoparticles to the base fluid, it is seen that
fluid at the corresponding Reynolds number. This Nusselt number ratio regardless of the height of the microchannel, this addition induces an
strictly decreases as the Reynolds number is increased, exhibiting an increase of the friction coefficient, which decreases with a growing
asymptotic behavior. In all the studied cases, the nanofluids with a TiO2 Reynolds number. This is how, for example, for the MCH-1 micro-
concentration of 3%wt presents higher improvements. According to channel (Fig. 12a), the addition of nanoparticles at a 1 wt% con-
such a tendency, the relative Nusselt number as a function of the centration increases the coefficient of friction by 137.68% at Re = 200,
Reynolds number has been modeled through Eq. (24) with adjustable a difference that decreases continuously to become less than 5%, where
parameters A, B and C (Table 5). it is considered that this variable is independent of the concentration of
nanoparticles, a phenomenon that happened at Re = 600 for this case,
Nunf Re / B while for the nanofluid with a concentration of 3 wt%, this behavior
= A·e +C
Nubf (24) was recorded at Re = 400. It should be noted that although with the
microchannels MCH-2 and MCH-3 no independence of the coefficient of
Based on the obtained adjustment results, it has been found that friction under the criterion stated above was recorded, the results show
there is a Reynolds number after which the addition of TiO2 nano- a trend that approaches significantly the one mentioned above. In re-
particles does not improve the Nusselt number. The asymptotic beha- lation to the increase of the coefficient of friction, it is found that the
vior of the proposed model corresponds this observation, i.e., the maximum, for all the modelled heights, was recorded at a nanoparticle
parameter C is close to unity in all the modeled cases. The parameter A concentration of 1 wt% for the entire range of Reynolds numbers stu-
corresponds to the maximum improvement of the Nusselt number due died. This is explained in part because, referred to the fluid base, the
to the addition of TiO2 nanoparticles, in the range of Reynolds number addition of nanoparticles at a concentration of 1 wt%, induces a de-
of 200–1200. Finally, the parameter B represents a critical value of the crease in the average flow speed, because it depends directly on the
Reynolds number, from which the implementation of nanofluids as a viscosity of the fluid (6), which for microchannel MCH-1 is on the

10
V.A. Martínez, et al. Applied Thermal Engineering 161 (2019) 114130

average equal to 0.000872 Pa·s for a concentration of 1 wt% and Appendix A. Supplementary material
Re = 200, while for water the viscosity is 0.000902 Pa·s, which implies
that the nanofluid has on the average a 3.25% lower viscosity, since due Supplementary data associated with this article can be found, in the
to the increased thermal conductivity of the nanofluid, its average online version, at https://doi.org/10.1016/j.applthermaleng.2019.
temperature was 295.14 K, while that of the base fluid was 294.15 K for 114130.
the same Reynolds number. Similarly, the decrease in the friction
coefficient due to the increased concentration of nanoparticles at 3 wt% References
with respect to the synthesized nanofluid at a concentration of 1 wt%,
for the same case explained above (MCH-1), at a Reynolds number of [1] D.B. Tuckerman, R. Pease, High-performance heat sinking for vlsi, IEEE Electron.
200, the nanofluid with a concentration of 3 wt% has an average Dev. Lett. 2 (5) (1981) 126–129.
[2] H. Ghaedamini, P. Lee, C. Teo, Developing forced convection in con-
viscosity of 0.000978 Pa·s, which represents a 12.11% increase with verging–diverging microchannels, Int. J. Heat Mass Transfer 65 (Supplement C)
respect to the nanofluid with a concentration of 1 wt%, which is be- (2013) 491–499, https://doi.org/10.1016/j.ijheatmasstransfer.2013.06.036.
cause the nanofluid with higher concentration of nanoparticles has an [3] S.G. Kandlikar, W.J. Grande, Evaluation of single phase flow in microchannels for
high heat flux chip cooling—thermohydraulic performance enhancement and fab-
average temperature in the microchannel of 294.62 K. rication technology, Heat Transfer Eng. 25 (8) (2004) 5–16, https://doi.org/10.
Regarding the influence of the height of the microchannel on the 1080/01457630490519772.
coefficient of friction (Fig. 12d), it is seen that for Reynolds lower than [4] G. Takács, G. Bognár, E. Bándy, G. Rózsás, P.G. Szabó, Fabrication and character-
ization of microscale heat sinks, Microelectron. Reliab. 79 (2017) 480–487.
400, the decrease of the height of the microchannel induces a decrease [5] H. Mohammed, G. Bhaskaran, N. Shuaib, R. Saidur, Heat transfer and fluid flow
of the coefficient of friction, a reduction which rises to 19.27% when characteristics in microchannels heat exchanger using nanofluids: a review, Renew.
the height is decreased by 200 µ m and to 33.58% when it is reduced by Sust. Energy Rev. 15 (3) (2011) 1502–1512.
[6] A.J. Shkarah, M.Y.B. Sulaiman, M.R.B.H. Ayob, H. Togun, A 3d numerical study of
400 µ m, in the case of MCH-1. However, for Reynolds numbers that are
heat transfer in a single-phase micro-channel heat sink using graphene, aluminum
the same as or greater than 400, an increase in the coefficient of friction and silicon as substrates, Int. Commun. Heat Mass Transfer 48 (2013) 108–115.
is recorded with the reduction of the height of the microchannel, [7] S.U. Choi, J.A. Eastman, Enhancing thermal conductivity of fluids with nano-
reaching for the case mentioned above, a maximum increase of 28.76% particles, Tech. rep., Argonne National Lab., IL (United States), 1995.
[8] M. Sahooli, S. Sabbaghi, Investigation of thermal properties of cuo nanoparticles on
when the height was reduced by 200 µ m and 37.29% with a reduction the ethylene glycol–water mixture, Mater. Lett. 93 (Supplement C) (2013) 254–257,
of 400 µ m, both registered for a Reynolds of 1000. https://doi.org/10.1016/j.matlet.2012.11.097.
[9] R. Agarwal, K. Verma, N.K. Agrawal, R.K. Duchaniya, R. Singh, Synthesis, char-
acterization, thermal conductivity and sensitivity of cuo nanofluids, Appl. Therm.
Eng. 102 (Supplement C) (2016) 1024–1036, https://doi.org/10.1016/j.
5. Conclusions applthermaleng.2016.04.051.
[10] M. Abdolbaqi, N.A.C. Sidik, A. Aziz, R. Mamat, W. Azmi, M.N.A.W.M. Yazid,
G. Najafi, An experimental determination of thermal conductivity and viscosity of
In the present study we have investigated the effect of the con- bioglycol/water based tio2 nanofluids, Int. Commun. Heat Mass Transfer 77
centration of nanoparticles of TiO2 in water based nanofluids and the (Supplement C) (2016) 22–32, https://doi.org/10.1016/j.icheatmasstransfer.2016.
height of the microchannel (800 µ m, 600 µ m and 400 µ m) on the 07.007.
[11] S.S. Harandi, A. Karimipour, M. Afrand, M. Akbari, A. D’Orazio, An experimental
thermal performance of a heat sink, by making a numerical study of a study on thermal conductivity of f-mwcnts–fe3o4/eg hybrid nanofluid: effects of
three-dimensional laminar flow (200 Re 1200) at concentrations of temperature and concentration, Int. Commun. Heat Mass Transfer 76 (2016)
1 wt% and 3 wt%. The simulations were developed considering tem- 171–177.
[12] H. Jiang, H. Li, C. Zan, F. Wang, Q. Yang, L. Shi, Temperature dependence of the
perature dependent thermophysical properties of the base fluid as well
stability and thermal conductivity of an oil-based nanofluid, Thermochim. Acta 579
as the nanofluids. Specifically, we studied experimentally the behavior (Supplement C) (2014) 27–30, https://doi.org/10.1016/j.tca.2014.01.012.
of the viscosity and the thermal conductivity with respect to the tem- [13] M. Ghanbarpour, E.B. Haghigi, R. Khodabandeh, Thermal properties and rheolo-
perature to determine the models that describes them. From the results, gical behavior of water based al2o3 nanofluid as a heat transfer fluid, Exp. Therm.
Fluid Sci. 53 (Supplement C) (2014) 227–235, https://doi.org/10.1016/j.
it can be established that using nanofluids as coolants in heat sinks with expthermflusci.2013.12.013.
microchannels in general favors heat transfer significantly when the [14] M.H.K. Darvanjooghi, M.N. Esfahany, Experimental investigation of the effect of
predominant transfer mechanism is conduction, which happens at low nanoparticle size on thermal conductivity of in-situ prepared silica–ethanol nano-
fluid, Int. Commun. Heat Mass Transfer 77 (Supplement C) (2016) 148–154,
Reynolds numbers, recording a maximum increase of the convective https://doi.org/10.1016/j.icheatmasstransfer.2016.08.001.
heat transfer coefficient of 19.66% compared to the base fluid for mi- [15] P. Maheshwary, C. Handa, K. Nemade, A comprehensive study of effect of con-
crochannel MCH-3, with a concentration of 3 wt% and Re = 200. With centration, particle size and particle shape on thermal conductivity of titania/water
based nanofluid, Appl. Therm. Eng. 119 (Supplement C) (2017) 79–88, https://doi.
respect to the effect of the height of the microchannel, it is seen that org/10.1016/j.applthermaleng.2017.03.054.
when it is decreased and thereby Dh is reduced, the Nusselt number [16] S.S. Harandi, A. Karimipour, M. Afrand, M. Akbari, A. D’Orazio, An experimental
undergoes a decline in its magnitude, which indicates that this favors study on thermal conductivity of f-mwcnts–fe3o4/eg hybrid nanofluid: effects of
temperature and concentration, Int. Commun. Heat Mass Transfer 76 (Supplement
heat transfer by conduction, reducing the importance of convection,
C) (2016) 171–177, https://doi.org/10.1016/j.icheatmasstransfer.2016.05.029.
which is not detrimental to heat transfer, since the convective heat [17] K.A. Hamid, W. Azmi, M. Nabil, R. Mamat, K. Sharma, Experimental investigation
transfer coefficient generally increases with decreasing the hydraulic of thermal conductivity and dynamic viscosity on nanoparticle mixture ratios of
tio2-sio2 nanofluids, Int. J. Heat Mass Transfer 116 (Supplement C) (2018)
diameter. In spite of the improvements achieved in heat transfer with
1143–1152, https://doi.org/10.1016/j.ijheatmasstransfer.2017.09.087.
the addition of nanoparticles, as well as with the decrease of the height [18] M. Sarafraz, V. Nikkhah, M. Nakhjavani, A. Arya, Thermal performance of a heat
of the microchannel, which is favored at low Reynolds numbers in these sink microchannel working with biologically produced silver-water nanofluid: ex-
cases, also the largest increases in the coefficient of friction were ob- perimental assessment, Exp. Therm. Fluid Sci. 91 (Supplement C) (2018) 509–519,
https://doi.org/10.1016/j.expthermflusci.2017.11.007.
tained due to the addition of nanoparticles, for the three microchannel [19] S. Peyghambarzadeh, S. Hashemabadi, A. Chabi, M. Salimi, Performance of water
heights studied, recording a maximum increase with respect to the base based cuo and al2o3 nanofluids in a cu–be alloy heat sink with rectangular mi-
fluid of 137.68% for the case of the nanofluid with 1 wt% in micro- crochannels, Energy Convers. Manage. 86 (Supplement C) (2014) 28–38, https://
doi.org/10.1016/j.enconman.2014.05.013.
channel MCH-1 and Re = 200. [20] R. Chein, J. Chuang, Experimental microchannel heat sink performance studies
using nanofluids, Int. J. Therm. Sci. 46 (1) (2007) 57–66, https://doi.org/10.1016/
j.ijthermalsci.2006.03.009.
[21] E. Manay, B. Sahin, The effect of microchannel height on performance of nano-
Acknowledgements fluids, Int. J. Heat Mass Transfer 95 (Supplement C) (2016) 307–320, https://doi.
org/10.1016/j.ijheatmasstransfer.2015.12.015.
The authors acknowledge with thanks the support through Proyecto [22] H. Mohammed, P. Gunnasegaran, N. Shuaib, Heat transfer in rectangular micro-
channels heat sink using nanofluids, Int. Commun. Heat Mass Transfer 37 (10)
Código 051916VC_DAS, Dirección de Investigación, Científica y
(2010) 1496–1503, https://doi.org/10.1016/j.icheatmasstransfer.2010.08.020.
Tecnológica, Dicyt, Universidad de Santiago de Chile.

11
V.A. Martínez, et al. Applied Thermal Engineering 161 (2019) 114130

[23] A. Abdollahi, H. Mohammed, S. Vanaki, A. Osia, M.G. Haghighi, Fluid flow and heat Appl. Therm. Eng. 26 (17) (2006) 2209–2218, https://doi.org/10.1016/j.
transfer of nanofluids in microchannel heat sink with v-type inlet/outlet arrange- applthermaleng.2006.03.014.
ment, Alex. Eng. J. 56 (1) (2017) 161–170, https://doi.org/10.1016/j.aej.2016.09. [31] A. Sakanova, C.C. Keian, J. Zhao, Performance improvements of microchannel heat
019. sink using wavy channel and nanofluids, Int. J. Heat Mass Transfer 89 (2015)
[24] L. Godson, B. Raja, D.M. Lal, S. Wongwises, Enhancement of heat transfer using 59–74, https://doi.org/10.1016/j.ijheatmasstransfer.2015.05.033.
nanofluids—an overview, Renew. Sust. Energy Rev. 14 (2) (2010) 629–641. [32] A.I. Zografos, W.A. Martin, J. Sunderland, Equations of properties as a function of
[25] X. Shi, S. Li, Y. Wei, J. Gao, Numerical investigation of laminar convective heat temperature for seven fluids, Comput. Meth. Appl. Mech. Eng. 61 (2) (1987)
transfer and pressure drop of water-based al2o3 nanofluids in a microchannel, Int. 177–187, https://doi.org/10.1016/0045-7825(87)90003-X.
Commun. Heat Mass Transfer 90 (2018) 111–120, https://doi.org/10.1016/j. [33] G. Wypych, Handbook of Fillers, Elsevier, 2016.
icheatmasstransfer.2017.11.007. [34] D.A. Vasco, N.O. Moraga, G. Haase, Parallel finite volume method simulation of
[26] V.A. Martínez, D.A. Vasco, C.M. García-Herrera, Transient measurement of the three-dimensional fluid flow and convective heat transfer for viscoplastic non-
thermal conductivity as a tool for the evaluation of the stability of nanofluids newtonian fluids, Numer. Heat Transfer, Part A: Appl. 66 (9) (2014) 990–1019.
subjected to a pressure treatment, Int. Commun. Heat Mass Transfer 91 (2018) [35] B. Rimbault, C.T. Nguyen, N. Galanis, Experimental investigation of cuo–water
234–238, https://doi.org/10.1016/j.icheatmasstransfer.2017.12.016. nanofluid flow and heat transfer inside a microchannel heat sink, Int. J. Therm. Sci.
[27] S.S. Sonawane, R.S. Khedkar, K.L. Wasewar, Effect of sonication time on en- 84 (2014) 275–292, https://doi.org/10.1016/j.ijthermalsci.2014.05.025.
hancement of effective thermal conductivity of nano tio2–water, ethylene glycol, [36] S.S. Murshed, P. Estellé, A state of the art review on viscosity of nanofluids, Renew.
and paraffin oil nanofluids and models comparisons, J. Exp. Nanosci. 10 (4) (2015) Sust. Energy Rev. 76 (2017) 1134–1152.
310–322, https://doi.org/10.1080/17458080.2013.832421 Available from: [37] W. Duangthongsuk, S. Wongwises, Measurement of temperature-dependent thermal
arXiv:https://doi.org/10.1080/17458080.2013.832421. conductivity and viscosity of tio2-water nanofluids, Exp. Therm. Fluid Sci. 33 (4)
[28] C.-J. Ho, L. Wei, Z. Li, An experimental investigation of forced convective cooling (2009) 706–714.
performance of a microchannel heat sink with al2o3/water nanofluid, Appl. Therm. [38] Y. Islamoglu, C. Parmaksizoglu, The effect of channel height on the enhanced heat
Eng. 30 (2–3) (2010) 96–103. transfer characteristics in a corrugated heat exchanger channel, Appl. Therm. Eng.
[29] C. Nguyen, F. Desgranges, N. Galanis, G. Roy, T. Maré, S. Boucher, H.A. Mintsa, 23 (8) (2003) 979–987, https://doi.org/10.1016/S1359-4311(03)00029-2.
Viscosity data for al2o3–water nanofluid—hysteresis: is heat transfer enhancement [39] P.-S. Lee, S.V. Garimella, Thermally developing flow and heat transfer in rectan-
using nanofluids reliable? Int. J. Therm. Sci. 47 (2) (2008) 103–111. gular microchannels of different aspect ratios, Int. J. Heat Mass Transfer 49 (17)
[30] S.J. Palm, G. Roy, C.T. Nguyen, Heat transfer enhancement with the use of nano- (2006) 3060–3067, https://doi.org/10.1016/j.ijheatmasstransfer.2006.02.011.
fluids in radial flow cooling systems considering temperature-dependent properties,

12

You might also like