Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Heat Transfer Engineering

ISSN: 0145-7632 (Print) 1521-0537 (Online) Journal homepage: https://www.tandfonline.com/loi/uhte20

Heat Transfer and Pressure Drop Characteristics


of Dilute Alumina–Water Nanofluids in a Pipe at
Different Power Inputs

Richa Saxena, Dasaroju Gangacharyulu & Vijaya Kumar Bulasara

To cite this article: Richa Saxena, Dasaroju Gangacharyulu & Vijaya Kumar Bulasara
(2016) Heat Transfer and Pressure Drop Characteristics of Dilute Alumina–Water Nanofluids
in a Pipe at Different Power Inputs, Heat Transfer Engineering, 37:18, 1554-1565, DOI:
10.1080/01457632.2016.1151298

To link to this article: https://doi.org/10.1080/01457632.2016.1151298

Accepted author version posted online: 18


Feb 2016.
Published online: 21 May 2016.

Submit your article to this journal

Article views: 5862

View related articles

View Crossmark data

Citing articles: 12 View citing articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=uhte20
Heat Transfer Engineering, 37(18):1554–1565, 2016
Copyright 
C Taylor and Francis Group, LLC
ISSN: 0145-7632 print / 1521-0537 online
DOI: 10.1080/01457632.2016.1151298

Heat Transfer and Pressure Drop


Characteristics of Dilute
Alumina–Water Nanofluids in a Pipe
at Different Power Inputs

RICHA SAXENA,1 DASAROJU GANGACHARYULU,2


and VIJAYA KUMAR BULASARA2
1
Department of Petroleum Engineering, DIT University, Dehradun, Uttarakhand, India
2
Department of Chemical Engineering, Thapar University, Patiala, Punjab, India

This work addresses the effect of temperature on the thermophysical properties (i.e., density, viscosity, thermal conductivity,
and specific heat capacity) of alumina–water nanofluid over a wide temperature range (25◦ C–75◦ C). Low concentrations
(0–0.5% v/v) of alumina nanoparticles (40 nm size) in distilled water were used in this study. The pressure drop and the
effective heat transfer coefficient of nanofluids were also estimated for different power inputs and at different flow rates
corresponding to Reynolds numbers in the range of 1500–6000. The trends in variation of thermophysical properties of
nanofluids with temperature were similar to that of water, owing to their low concentrations. However, the density, viscosity,
and thermal conductivity of nanofluids increased, while the specific heat capacity decreased with increasing the nanoparticle
concentration. The convective heat transfer coefficient of the nanofluid and the pressure drop along the test section increased
with increasing the particle concentration and flow rate of nanofluid. Results showed that the heat transfer coefficient
increases, while the pressure drop decreases slightly with increasing the power input. This is because of the fact that
increasing power input to heater increases the bulk mean temperature of nanofluids, resulting in a decreased viscosity. The
prepared nanofluids were found to be more effective under turbulent flow than in transition flow.

INTRODUCTION exchangers, microchannels, cooling systems, nuclear reactors,


concentrated solar power plants, fuel cells, automobiles, phar-
Nanofluids are engineered colloidal suspensions of maceutical processes, domestic refrigerators, chillers, grinding
nanometer-sized (1–100 nm) solid particles (called nanopar- processes, heating, ventilation, and air-conditioning systems,
ticles) in base fluids and have been proposed as a route for waste heat recovery devices for boilers and furnaces, and other
surpassing the performance of heat transfer fluids in thermal types of heat exchangers [4–6]. Therefore, knowledge of the
energy systems [1, 2]. Since solid materials have higher thermal physical, thermal, and flow properties of nanofluids is essential
conductivity than fluids, the suspended particles are supposed to in the selection of suitable nanofluids for specific applications.
increase the thermal conductivity and heat transfer performance The practical utility of a nanofluid as a heat transfer fluid is
of base fluids. The stability of suspension is also an impor- best determined by its convective heat transfer coefficient [7].
tant factor in the selection of nanoparticles [3]. Nanofluids have Nanofluids were found to exhibit enhanced thermal properties,
superior thermal conductivity and convective heat transfer coef- especially thermal conductivity, even at very low concentrations
ficient compared to the base fluid, which makes them potentially of suspended nanoparticles and are most suitable for rapid heat-
useful in many heat transfer applications such as compact heat ing and cooling applications [8]. Increase in effective thermal
conductivity and changes in density, specific heat, and viscos-
ity are important indications of improved heat transfer behav-
Address correspondence to Dr. Vijaya Kumar Bulasara, Department of
ior of nanofluids [9]; however, the net benefit of nanofluids
Chemical Engineering, Thapar University, Patiala 147004, Punjab, India. as heat transfer fluids is determined through the heat transfer
E-mail: vkbulasara@thapar.edu. coefficient and pressure drop. As a result of increased heat

1554
R. SAXENA ET AL. 1555

transfer coefficient, the nanofluids facilitate reduction in size convective cooling performance of a nanofluid in a tube at a
of thermal energy system (heat transfer equipment) and lead to given heat flux depends strongly on the flow rate and particle
increased energy and fuel efficiency, reduced emissions, and im- fraction, and the gains in forced convection performance can
proved reliability. On the other hand, an increase in the density be significantly outweighed by the pressure drop penalty re-
and viscosity of a basefluid with the addition of nanoparticles sulted from their greatly enhanced viscosity [30]. The convec-
would contribute to pressure drop, thereby necessitating a high tive movement of nanoparticles in the turbulent core region and
pumping power [10–12]. Therefore, pressure drop during the the diffusion mechanism such as thermophoresis in the laminar
flow of nanofluids is also an important consideration in the eco- sublayer are believed to be the reasons for enhanced heat trans-
nomic evaluation and final application of a nanofluid. fer in turbulent flow conditions [31]. Ho and Lin [32] demon-
Extensive research has been carried out on water-based strated that the forced convection heat transfer performance of
nanofluids containing alumina (Al2 O3 ), TiO2 , and CuO nanopar- alumina–water nanofluid in turbulent flow can be greatly en-
ticles, besides a few studies on Cu, SiO2 , SiC, ZnO nanopar- hanced by elevating its inlet temperature. Although nanofluids
ticles, and carbon nanotubes [1–5, 9, 13–16]. It has been well enhance the heat transfer in forced convection, the Al2 O3 –water
established that the thermal conductivity of nanofluids increases nanofluids have shown poor heat transfer performance in natu-
with increasing volume fraction of nanoparticles, and their per- ral convection and nucleate boiling as compared to pure water
formance is also affected by particle size, geometry and mi- [33, 34].
crostructure, type of solid material and base fluid, and solu- Since all previous studies on convective heat transfer using
tion pH [17]. One of the key studies on thermal conductivity nanofluids have been conducted at a particular power input, the
of nanofluids is an international nanofluid benchmark exercise present study assumes significance in that the experiments are
[18], in which the thermal conductivities of various nanoflu- performed at different power inputs in order to study the effect
ids were measured and compared by 34 different organizations of nanofluid temperature on the heat transfer and pressure drop
worldwide using a variety of experimental approaches. This characteristics. Alumina nanoparticles were chosen in this work
study revealed that the Maxwell’s effective medium theory for as they are less expensive and easily dispersible in water without
dispersed particles could predict well the experimental data. needing additives such as surfactants and possess slightly higher
In another study by Keblinski et al. [10], almost all the pre- specific heat capacity and thermal conductivity than CuO, titania
viously published data on thermal conductivity of nanofluids and zirconia.
have been found to lie between the well-known Hashin and
Shtrikman effective medium bounds, suggesting that the ther-
mal conductance behavior of nanofluids is similar to that of
binary solid composites or liquid mixtures. In addition to ther- METHODOLOGY
mal conductivity, the thermal dispersion and Brownian motion
of particles also play an important role in heat transfer enhance- Since high volume concentrations of nanoparticles result in
ment [19–21]. Further, particle migration is also a reason for low specific heat capacity and increased pressure drop [35], di-
the enhancement of thermal conductivity and lowering of the lute solutions of Al2 O3 –water nanofluids were considered in
thermal boundary-layer thickness [22]. this work. The uncertainty in the measurement of temperature
Very few articles described the effect of temperature on the was ±0.1◦ C and pressure was ±0.2 Pa. The accuracy of the
thermal conductivity of nanofluids, that too at high particle flowmeter and other units was ±1%. Each experiment on esti-
concentrations (1–5% v/v) [23]. While the study of effective mation of thermophysical properties of nanofluids was repeated
thermal conductivity of nanofluids has received much atten- at least four times, while the forced convection experiments
tion in the last decade [24, 25], recent research focuses on ef- were repeated at least three times, and average values have been
fective viscosity, which influences the flow and heat transfer reported with a standard deviation of ±5% (formulas for cal-
characteristics of nanofluids [26]. Besides thermal conductiv- culating the standard deviation were adopted from Buongiorno
ity and viscosity, other properties such as specific heat capacity et al. [18]).
should also be considered to examine their influence on flow
and heat transfer properties of nanofluids because the increase
of thermal conductivity might be offset by a large increase Preparation of Nanofluids
in viscosity and decreasing of effective specific heat capacity
[27, 28]. Unlike several studies that have focused on thermal Nanofluids are generally prepared by one of two methods,
conductivity, the convective heat transfer coefficient and pres- namely, single-step or two-step [36]. The single-step method
sure drop of nanofluids still need to be explored in more detail. involves simultaneous synthesis and dispersion of nanoparti-
Williams et al. [12] showed that the Dittus–Boelter correlation cles in the base liquid to produce stable nanofluids, while the
can be used to predict the Nusselt number for alumina and two-step method involves dispersion of nanoparticles (obtained
zirconia nanofluids. The enhancement of convective heat trans- by different methods) into the base liquid. In the present work,
fer in nanofluids at high Reynolds numbers is attributed to a water-based alumina nanofluid (0.1–0.5% v/v) was prepared
decrease in thermal boundary-layer thickness [29]. The forced by a two-step method by adding a known amount of Al2 O3
heat transfer engineering vol. 37 no. 18 2016
1556 R. SAXENA ET AL.

nanoparticles (of 40 nm nominal size) into distilled water (with- viscosity (μ) can be calculated as
out using a surfactant) and keeping the solution in ultrasonic bath tnf ρnf
for 2–3 h for proper mixing and stable dispersion. The prepared μnf = μw (2)
nanofluids remained stable for 21 days, during which no signif- tw ρw
icant changes in absorbance (determined by ultraviolet–visible where the subscript w represents distilled water.
[UV-Vis] spectroscopy) and turbidity were noticed. A similar An attempt has been made to fit a power-law model to the
observation has been reported in the literature [31, 37], wherein experimental data based on Eq. (3) using a nonlinear regression
the authors dispersed alumina nanoparticles (∼40 nm) in water method to evaluate the exponents “a” and “b”:
in an ultrasonic bath without using additives and observed that
dilute (0.3–0.5% v/v) nanofluids are stable for several weeks. μnf = μw (1 + φ)a (Tnf )b (3)
However, in the present work, freshly prepared nanofluids were
Here, the temperature of nanofluid (T nf ) is in ◦ C and Eq. (3)
used in the measurement of thermophysical properties and in
is valid only for 20◦ C ≤ T nf ≤ 80◦ C.
forced convection experiments.

Specific Heat Capacity


Measurement of Nanofluid Properties
The specific heat capacity of nanofluids was determined by
differential scanning calorimetry (DSC 821e, Mettler Toledo)
The reliability and accuracy of the heat transfer coefficient
using sapphire as the standard reference for calibration (mea-
depend greatly on the experimental system calibration rather
surement error ±0.3%). The three-step DSC procedure was
than on the thermophysical property models of nanofluids [38].
followed to measure specific heat capacity. In the first step,
Therefore, all the properties of nanofluids were determined ex-
a measurement was taken with two empty sample pans made of
perimentally. The accuracy associated with the measurements
aluminum. In the second step, measurement was taken with the
of density, viscosity, specific heat capacity, and thermal conduc-
reference sample in one pan, keeping the other pan empty. And
tivity are 99.8%, 99.5%, 99.7%, and 99.0%, respectively.
in the final step, the measurement was taken for the nanofluid
sample in one pan, keeping the other pan empty. A detailed
procedure for DSC analysis is available elsewhere [39, 40].
Density
The experimental heat capacity values were also compared with
The density of nanofluids at different temperatures is mea- those calculated from energy balance around the heat exchanger
sured by using a pycnometer, which is also known as specific (Eq. (4)), assuming perfect heat transfer between the two fluids
gravity bottle or relative density bottle. It is a glass flask (5 ml without accounting for heat losses [41]. The average energy bal-
working volume) with a close-fitting ground-glass stopper hav- ance error was 8.2%, while the minimum and maximum errors
ing a capillary hole through it. The precision of the pycnometer were 4.3% and 12.8%, respectively:
used in this work is ±0.2% (±2 kg/m3). Distilled water was
ṁ w C P,w Tw
used as the working liquid since the temperature-dependent val- C P,nf = (4)
ues of density are known. The measured values of density (ρ) of ṁ nf Tnf
nanofluid were also compared with those calculated using the Most of the previous studies used the ideal gas mixture cor-
following expression: relation (Eq. (5)) for estimating the specific heat capacity of
nanofluids from the individual heat capacities [42]:
ρnf = (1 − φ)ρbf + φρnp (1)
C P,nf = (1 − φ)C P,bf + φC P,np (5)
Here, φ is the volume fraction of nanoparticles (0.1 vol.%
means φ = 0.001) and the subscripts nf , bf, and np represent the However, the relation described by Eq. (5) deviates consider-
nanofluid, base fluid, and solid nanoparticles respectively. ably from experimental observations [43]; therefore, it was not
used in the present work. The specific heat capacity of nanoflu-
ids can be closely estimated using the volume fraction mixture
Viscosity rule [44, 45] given by Eq. (6):

Viscosity of nanofluids is less investigated than thermal con- (ρC P )nf = (1 − φ)(ρC P )bf + φ(ρC P )np (6)
ductivity. A Ubbelohde viscometer was used for measuring the
viscosity of nanofluids at different temperatures (measurement
error ±0.5%). This apparatus is cheap and easily available. A
Thermal Conductivity
Ubbelohde-type viscometer or suspended-level viscometer is
a measuring instrument that uses a capillary-based method of Thermal conductivity is an important parameter in enhancing
measuring viscosity. By noting the time (t) the fluid takes to pass the heat transfer performance of a base fluid with a suspension
through a known length (volume) of the capillary, the dynamic of nanoparticles. Among the different techniques available for
heat transfer engineering vol. 37 no. 18 2016
R. SAXENA ET AL. 1557

tures T1 to T12 , wherein T1 and T8 represent the temperatures of


nanofluid at the inlet (T1 ≈ 35◦ C) and outlet, respectively, and
T2 through T7 represent the outside temperatures of the cop-
per pipe (test section) at different locations separated by 0.2 m
to each other. Here, T1 and T8 were measured by temperature
sensors immersed in the mixing chambers provided at the two
ends of the test section. A rotameter (with 0.02 L/min accu-
racy) and an inclined tube manometer were used to measure
the volumetric flow rate and pressure difference across the test
section, respectively. A stainless-steel heat exchanger (pipe-in-
pipe type) with countercurrent flow pattern was used to cool
the nanofluid (from T9 to T10 ) at outlet using cooling water
(T11 ≈ 25◦ C). Each experiment was conducted for at least
2 hours until the attainment of steady state with a minimum
Figure 1 Schematic diagram of experimental setup (not to scale).
of three repetitions, and the data were acquired manually. After
attaining steady state, the temperature measurement at the outlet
measurement of thermal conductivity of nanofluids, the tran- (T8 ) remained stable with good accuracy and precision (<1%
sient hot-wire method has been extensively used in literature. error). The flow rate of the nanofluid was adjusted using a globe
Recently, the KS-1 probe of the KD2 Pro thermal property an- valve and the experiments were repeated for different power in-
alyzer, which is also based on the transient hot-wire method, puts. The uncertainty analysis was carried out using the method
has emerged as the simplest and most convenient way of de- described by Cieśliński and Kaczmarczyk [15], considering the
termining the thermal conductivity of low-viscosity fluids such measurement errors. Calculations indicated that the uncertain-
as water and other aqueous solutions [18, 46, 47]. Therefore, ties involved in the estimation of friction factor, heat transfer
the thermal conductivity was measured by using the KD2 Pro coefficient, Reynolds number, and Nusselt number are ±3.2%,
thermal properties analyzer supplied by Decagon Devices Inc., ±2.7%, ±1.6% and ±2.4%, respectively. The constant heat flux
USA (measurement error ±1%). The KD2 Pro is a battery- boundary condition was applied in the present work.
operated device consisting of a hand-held controller and sensor
that can be inserted into the medium for measuring the thermal
properties. Heat Transfer Coefficient
The rate of heat transfer (Q) to the nanofluid can be calculated
Forced Convection Experiments as
Experimental Setup Q = ṁ nf C P,nf (T8 − T1 ) (7)
The experimental setup used to perform convective heat The heat transfer rate in the cylindrical test section is given
transfer experiments is shown in Figure 1. It consisted of a by
1.4-m-long copper test section (12.7 mm outer diameter and
(Ts,o − Ts,i )
0.5 mm thickness) with heating facility. A transition length Q = 2πk L (8)
(calm length) of 1.6 m was used (prior to the test section) to ln(do /di )
eliminate the entrance effect and ensure fully developed flow in Here, k is the thermal conductivity of copper, L is the length
the test section. The storage tank had a capacity of 3 L and of the test section, T s,o is the outside surface temperature of the
was fitted with a proportional-integral-derivative (PID) con- copper pipe, T s,i is the inside surface temperature of the copper
trolled heater. A magnetic drive pump was used to circulate pipe, and do and di are the outer and inner diameters of the pipe.
the nanofluid to the test section. The test section (copper pipe) The outside surface temperature of the copper pipe, T s,o is
was first wound with sheet mica followed by Nichrome heating evaluated as follows:
wire, giving a maximum power of 500 W. It was then insulated
T2 + T3 + T4 + T5 + T6 + T7
by layers of ceramic fiber, glass wool, and asbestos rope to pre- Ts,o = (9)
vent radial heat loss. The temperature difference between the 6
insulated surfaces of the experimental facility and the ambient The value of T s,i can be obtained by equating Eqs. (7) and
was found to be less than 10◦ C. The terminals of the Nichrome (8).
wire were attached to a tapped autotransformer, by which the Now the heat transfer coefficient (hi ) of nanofluid can be
heat flux can be varied by varying the voltage. The power input found using the following expression:
to heater is measured by a wattmeter with ±5 W accuracy. In
Q
total, 12 calibrated resistance temperature detector (RTD) Pt100 hi = (10)
sensors with 0.1◦ C accuracy were used to measure the tempera- πdi L(Ts,i − Tb )

heat transfer engineering vol. 37 no. 18 2016


1558 R. SAXENA ET AL.

Here, T b is the bulk mean temperature of nanofluid, given by 1040


0 vol%, experiment
T1 + T8 1030
Tb = (11) 0.1 vol%, experiment
2 1020
0.5 vol%, experiment
Equation (1)

Density (kg/m3)
The Reynolds and Nusselt numbers are evaluated at the bulk
1010
mean temperature of nanofluid as
1000
di u nf ρnf
Re = (12)
μnf 990

980
h i di
Nu = (13) 970
knf
960
Here, unf is the velocity of nanofluid inside the test section. 20 30 40 50 60 70 80
In the present work, Q was calculated by using Eq. (7), and was
Temperature (°C)
substituted in Eqs. (8) and (10) to obtain T s,i and hi , respectively.
Figure 2 Variation in density of Al2 O3 –water nanofluid with temperature
The heat transfer coefficient (hi ) and the Nusselt number (Nu)
(0 vol% refers to water).
presented here are the surface-averaged values and not the local
values.
Eq. (1) can be applied for estimating the density of dilute alu-
mina nanofluids with good accuracy. However, the magnitude
Pressure Drop and Friction Factor of deviation gradually increases with increase in temperature
(>60◦ C) as well as particle concentration. A similar obser-
The pressure drop across the test section (p) was obtained
vation was made by Teng and Hung [40], who studied the
from manometer readings and the friction factor (f ) was evalu-
density of nanofluids at low temperatures (10–40◦ C). The au-
ated using the Darcy–Weisbach equation as follows:
thors used a value of 3880 kg/m3 for the density of nanopar-
 
2di p ticles (ρnp ) in Eq. (1), versus 3970 kg/m3 in the present study.
f = (14) This could be the reason for the observation of large deviation
ρnf u 2nf L
(>1%) of calculated values from the experimental values in their
Hwang et al. [48] showed that the friction factor correlation study.
for the single-phase flow can be extended to water-based Al2 O3 Figure 3 shows the variation of viscosity of alumina–water
nanofluids. Under fully developed laminar conditions, the Darcy nanofluid with temperature. It was observed that the dynamic
friction factor of dilute Al2 O3 –water nanofluid follows Eq. (15): viscosity of nanofluid decreases with increasing the temper-
64 ature. As the temperature increases, the intermolecular forces
f = (15) decrease because of the increased velocities of individual
Re
molecules. This could be the reason for lowering of nanofluid
Under turbulent conditions, the friction factor of dilute viscosity with increase in temperature. On the other hand, the
Al2 O3 –water nanofluid can be closely related to Colebrook dynamic viscosity increased with an increase in the particle
equation [49], given here:
  
1 ε di 2.51 0.0014
√ = −2 log10 + √ (16) 0 vol%, experiment
f 3.7 Re f 0.1 vol%, experiment
0.0012
 0.1 vol%, literature [27]
Here, ε di is the relative roughness of the pipe. 0.5 vol%, experiment
Viscosity (Pa·s)

0.0010 Equation (3)

RESULTS AND DISCUSSION 0.0008

Density and Viscosity 0.0006

Figure 2 shows the variation in density of nanofluid with 0.0004


temperature from 25◦ C to 75◦ C. It can be observed from this
20 30 40 50 60 70 80
figure that the density of nanofluids decreases with increase
in temperature, while it increases with volume fraction of Temperature (°C)
nanoparticles. The difference between the experimental and Figure 3 Variation in viscosity of Al2 O3 –water nanofluid with temperature
calculated density values was less than 0.5%, suggesting that (0 vol% refers to water).

heat transfer engineering vol. 37 no. 18 2016


R. SAXENA ET AL. 1559

concentration. Similar trends in the variation of viscosity with 0.80

Thermal conductivity (W/m·°C)


temperature and concentration have been reported by Adio et al. 0.78 0 vol%, KD2 Pro
[26] for alumina–glycerol nanofluid. The viscosity data re- 0.1 vol%, KD2 Pro
0.76 0.5 vol%, KD2 Pro
ported by Sekhar and Sharma [27] for 0.1 vol.% alumina–water Model 6 (Table 1)
0.74
nanofluid match closely with the measured values of viscosity
and also with Eq. (3) proposed in the present work (as shown in 0.72

Figure 3). However, the variation in viscosity was predominant 0.70


at low temperatures and the deviation in viscosity of nanofluid 0.68
from the base fluid decreased gradually as the temperature 0.66
was raised. It is also evident that Eq. (3) fits well with the 0.64
experimental data with less than 4% error. The values of “a”
0.62
and “b” obtained by nonlinear regression analysis were 59.556
and 0.00023, respectively (R2 = 0.9992). A large value of “a” 0.60
20 30 40 50 60 70 80
indicates that the viscosity is very much sensitive to particle
Temperature (°C)
concentration and a small value of “b” implies that the tempera-
ture dependence of viscosity of nanofluids is the same as that of Figure 5 Effect of temperature on thermal conductivity of Al2 O3 –water
nanofluids (0 vol% = water).
the base fluid. It should be noted that the use of Eq. (3) is limited
to nanofluid temperatures (T nf ) between 20◦ C and 80◦ C only.
deviation (up to 2%) as the entropy change and heat losses were
ignored.
Specific Heat Capacity

The effect of temperature on the specific heat capacity of Thermal Conductivity


water based alumina nanofluid is presented in Figure 4. From
this figure, it can be noted that the heat capacity of nanofluids It has been well established that the thermal conductivity of
is more susceptible to particle concentration than to tempera- nanofluids is more temperature sensitive than that of the base
ture in the temperature range 25–75◦ C. The specific heat of the fluid [13]. Figure 5 shows the effect of temperature on the ther-
base fluid (water) was almost constant; however, a slight (0.6%) mal conductivity of Al2 O3 –water nanofluids. The values of ther-
increase in the values of CP (from DSC) with temperature was mal conductivity increased with increasing the temperature as
noticed for the nanofluids. The heat capacity of nanofluids de- well as particle concentration. It can be observed that the effec-
creased with increasing the particle concentration, and the CP tive thermal conductivity varies nonlinearly with temperature,
values of 0.1 vol.% nanofluid were 0.5–1.0% lower than for the and it is more sensitive to temperature at high temperatures. At
base fluid, while, the CP values of 0.5 vol.% nanofluid were low temperatures (below 40◦ C), a small increase (<5%) in the
2.1–2.7% lower than those of the base fluid. The heat capac- values of thermal conductivity was noticed for nanofluids, and a
ity values calculated using Eq. (6) are very close to those ob- considerable increase (>10%) was noticed at high temperatures
tained from DSC with a negligible deviation (<0.5%). On the (above 40◦ C) in comparison with the base fluid. The decrease
other hand, Eq. (4) yielded very irregular results with a large in the difference of thermal conductivities approaching 20◦ C
indicates that the thermal conductivity is less sensitive to tem-
4.40
perature and particle concentration at low temperatures. This
0 vol%, DSC observation is in accordance with the findings of Das et al. [23],
4.35 A: 0 vol%
0.1 vol%, DSC who noticed increased thermal conductivity difference with in-
Heat capacity (kJ/kg·°C)

B: 0.1 vol%
C: 0.5 vol% 0.5 vol%, DSC creased temperature for Al2 O3 –water nanofluids (1 and 4 vol.%)
4.30
Equation (6)
Equation (4)
over the temperature range of 21–51◦ C. They suggested that the
4.25 strong temperature dependence of nanofluid thermal conduc-
4.20 tivity at high temperatures is due to the stochastic motion of
(A)
(B) nanoparticles. Chandrasekar and Suresh [19] explored that ad-
4.15 (A)
(B)
ditional energy transport in nanofluids can arise from the Brown-
4.10 (C) ian motion of particles induced by the effect of stochastic forces.
(C) Jang and Choi [20] also explained that the Brownian motion of
4.05
nanoparticles in the base fluid dominates at high temperatures
4.00 resulting in enhanced thermal conductivity. Therefore, it is ob-
20 30 40 50 60 70 80
vious to observe an increase in the difference of thermal con-
Temperature (°C) ductivity of a nanofluid from its base fluid at high temperatures.
Figure 4 Specific heat capacity of Al2 O3 –water nanofluids (0 vol% refers to Although several studies have been reported on the thermal
water). conductivity of nanofluids [13, 16, 19, 23], none of them has
heat transfer engineering vol. 37 no. 18 2016
1560 R. SAXENA ET AL.

Table 1 Model parameters for thermal conductivity (with T nf in ◦ C)

Model parameters
Model Expression k0 k1 k2 m n R2 value

Model 1 k = k0 (1 + φ)(Tnf − 20) 0.018 — — — — 0.5701


Model 2 k = k0 + k1 (1 + φ)(Tnf − 20) 0.601 2.31 × 10—3 — — — 0.7689
Model 3 k = k0 + k1 φ + k2 (Tnf − 20) 0.585 8.045 2.35 × 10—3 — — 0.9824
Model 4 k = k0 (1 + φ)m (Tnf − 20)n 0.521 — — 12.09 0.071 0.9704
Model 5 k = k0 (1 + φ)m + k1 (Tnf − 20)n 0.614 2.14 × 10—5 — 12.63 2.147 0.9896
Model 6 k = k0 + k1 (1 + φ)m + k2 (Tnf − 20)n 0.088 0.521 7.35 × 10−5 14.63 1.841 0.9960

presented any correlation for dilute nanofluids over a wide range input to heater resulted in significant increase in the bulk mean
of temperatures. Therefore, some simple model equations (pre- temperature of nanofluids. It can also be observed from this
sented in Table 1) were considered to correlate the experimental table that the higher the power input, the lower is the pressure
values of thermal conductivity with temperature (◦ C). In Ta- drop. This is probably because the increase of power input leads
ble 1, k0 , k1 , and k2 are the coefficients and “m” and “n” are the to increasing of fluid temperature, which, in turn, reduces the
exponents. Among the six different model equations, the one viscosity of nanofluid (as in Figure 3) and results in the reduction
described by Model 6 fits best (error < 2%) to the experimental of pressure drop. In comparison with the base fluid, the pressure
data with an R2 value of 0.9960. The values of the exponents drop values (p < 12%) observed in the present study (for
and coefficients obtained for Model 6 suggest that the thermal Re ≈ 6000 and φ = 0.1%) are well below that reported (30%)
conductivity of dilute alumina–water nanofluids is affected pri- in the literature [49] for 0.135 vol.% Al2 O3 –water nanofluid at
marily by the particle concentration rather than the temperature a Reynolds number of 20,000.
(since k1 > k2 and m > n) in the studied range of temperatures. A plot of friction factor versus Reynolds number is shown in
Figure 6. A comparison of the experimental data with Eqs. (15)
and (16) is also made in the figure. The relative roughness of
pipe (ε/di ) was found to be 0.0012 (using water at Re > 10,000)
Pressure Drop and Friction Factor following the procedure outlined by Fotukian and Esfahany [49].
The results for the friction factor are in accordance with those
The values of the pressure drop of nanofluids during the reported by Sharma et al. [43] for 0.1 vol.% and Suresh et al. [37]
convective heat transfer experiments are shown in Table 2. From for 0.5 vol.%. Increase in nanoparticle concentration resulted in
this table, it is clear that the pressure drop increases nonlinearly increase in the friction factor values due to increased pressure
as the flow rate increases. The pressure drop of nanofluids was drop. On the other hand, increase in the power input to the
slightly higher than the base fluid and increased with increasing heater (or increase in nanofluid temperature) caused a slight
the concentration of nanoparticles due to increased viscosity. decline in the friction factor values. At low Reynolds number
This observation is in good agreement with that reported by (Re ≈ 1500), the experimental values matched closely with Eq.
Duangthongsuk and Wongwises [50] for TiO2 –water nanofluids. (15), and at moderate Reynolds numbers (3000 and 4500) they
However, for any particular flow rate, an increase in the power deviated significantly from Eq. (15). However, from moderate

Table 2 Experimental data of pressure drop and heat transfer coefficient for alumina–water nanofluid

Heat transfer coefficient,


Bulk mean temperature, T b (◦ C) Pressure drop, p (Pa) hi (W/m2-◦ C)
Power Flow rate Reynolds
input (W) (L/min) number, Re 0% 0.1% 0.5% 0% 0.1% 0.5% 0% 0.1% 0.5%

218.4 0.5 1512 ± 17 41.3 42.7 43.6 15.7 19.1 19.7 1586 1739 1868
218.4 1.0 3025 ± 34 38.2 39.6 40.4 70.6 82.3 88.5 1667 1835 1981
218.4 1.5 4537 ± 50 37.1 38.5 39.3 143.5 169.2 182.6 2026 2416 2683
218.4 2.0 6049 ± 67 36.6 38.0 38.8 225.4 263.9 287.2 2793 3436 4195
312.9 0.5 1538 ± 23 44.1 45.5 46.9 15.7 18.2 19.2 1680 1823 1946
312.9 1.0 3064 ± 34 39.6 40.9 42.3 68.7 78.1 81.1 1824 2006 2220
312.9 1.5 4582 ± 37 38.0 39.4 40.8 138.2 163.4 174.7 2199 2455 2933
312.9 2.0 6110 ± 54 37.3 38.7 40.1 210.6 248.2 265.8 2917 3788 4618
475.7 0.5 1557 ± 28 48.8 50.6 52.3 14.7 17.2 17.3 1847 1995 2126
475.7 1.0 3094 ± 40 41.9 43.7 45.3 66.3 69.2 73.2 2012 2215 2514
475.7 1.5 4617 ± 38 39.6 41.4 43.0 127.4 142.5 152.1 2354 2601 3295
475.7 2.0 6155 ± 53 38.4 40.3 41.9 195.8 227.8 236.3 3196 4050 5172

heat transfer engineering vol. 37 no. 18 2016


R. SAXENA ET AL. 1561

0.10 Rescaled and error bars added


0.065
Vol%
0.09
0% 0.1% 0.5%
0.060

Power
0.08 218 W

input
313 W

Friction factor, f
0.07 476 W 0.055

0.06 0.050
0.05
0.045
0.04
0.040
0.03
Equation (15)
0.02 0.035
Equation (16)
0.01
0.030
0 1000 2000 3000 4000 5000 6000 7000
1000 2000 3000 4000 5000 6000
Reynolds number, Re Axis titles and legend are the same as those on the left

Figure 6 Variation in friction factor with Reynolds number (0 vol% refers to water).

to high Reynolds numbers (3000 to 6000), the experimental effect of power input to the heater. However, the present study
values approached Eq. (16). Therefore, it can be observed that makes an attempt to quantify the effect of power input on the
Eq. (15) is only valid for laminar flow conditions, while the heat transfer performance of water-based alumina nanofluids.
Colebrook relation described by Eq. (16) is equally valid for It was observed that the convective heat transfer using dilute
transition as well as turbulent flows. The lowest values of friction alumina–water nanofluids is more effective at high power inputs
factor were observed at the maximum flow rate corresponding to (or high fluid temperatures). An increase of power input from
Re ≈ 6000. 218 W to 313 W caused an improvement up to 12%, and further
increase in power input to 476 W resulted in an additional 15%
increment, leading to an overall enhancement of heat transfer
coefficient up to 27% (evaluated as hen = hnf /hbf – 1, based on
Convective Heat Transfer Coefficient
fixed volumetric flow rate and particle concentration). The ob-
served increase in heat transfer coefficient with increasing the
The values of the convective heat transfer coefficient of power input is probably due to the beneficial effects of enhance-
nanofluids at different flow rates and volume concentrations ment of thermal conductivity of the nanofluid and correspond-
for three different power inputs are presented in Table 2. The ing decrease of viscosity at elevated temperatures (Figures 3
heat transfer coefficient was observed to increase with increas- and 5).
ing the nanoparticle concentration as well as flow rate. This A plot showing the variation of Nusselt number with increas-
observation is in accordance with the findings of Ho and Lin ing Reynolds number is presented in Figure 7. The values of the
[32], who reported an increase in the values of convective heat Nusselt number obtained in this work are comparable with those
transfer coefficient with increasing the fluid flow rate as well as reported by Sharma et al. [43]. From this figure it is clear that the
inlet temperature. Under laminar flow condition (Re ≈ 1500), Nusselt number of dilute Al2 O3 –water nanofluids is positively
the heat transfer coefficient values increased by 8–10% and correlated to all the three parameters, namely, Reynolds num-
15–18%, respectively, for 0.1 vol.% and 0.5 vol.% nanofluids. ber, nanoparticle concentration, and power input. The values of
This is in good agreement with the reported increase of 8% for Nu increased nonlinearly with the increase in Re, φ, and power
0.3 vol.% Al2 O3 –water nanofluid under Re < 600 [48]. Under input (W in ). A correlation described by Eq. (17) was fitted to
turbulent flow condition (Re ≈ 6000), the heat transfer coef- the experimental data in order to quantify the dependence of
ficient values increased by 23–30% and 50–62%, respectively, Nusselt number on these parameters.
for 0.1 vol.% and 0.5 vol.% nanofluids. This observation is
also in accordance with the 48% increase reported by Fotukian
and Esfahany [49] for 0.05 vol.% Al2 O3 –water nanofluid under N u = a(Re)b (1 + 100φ)c (Win )d (17)
Re > 10,000. A similar observation has been recently reported
by Zhang et al. [51]: that the enhancement in heat transfer co- The values obtained for exponents a, b, c, and d by a nonlinear
efficient (hi ) for SiO2 –water nanofluid increases with Reynolds regression analysis were 0.309, 0.451, 0.455, and 0.203, respec-
number because of Brownian motion of nanoparticles, ther- tively, with R2 = 0.9319. The values of the exponents reveal
mophoresis, and thermal diffusion. The authors observed that that the Nusselt number is governed by Reynolds number and
only 2% enhancement in hi is possible in laminar flow (Re ≈ volume fraction of nanoparticles rather than the power input. A
1500), while nearly 60% enhancement can be achieved in tur- parity chart comparing the values of Nu evaluated experimen-
bulent flow (Re ≈ 9000) for 2% v/v nanofluid [51]. A detailed tally (Nuexp ) and Nu obtained from the preceding correlation
discussion of the mechanisms of heat transport in nanofluids (Nucal ) is shown in Figure 8. From this figure, it can be observed
is available elsewhere [16, 19, 24, 31]. Existing literature re- that the preceding relation holds well with acceptable devia-
ports on the heat transfer in nanofluids do not describe the tion (all data points are within ±10% region). Therefore, the

heat transfer engineering vol. 37 no. 18 2016


1562 R. SAXENA ET AL.

100 temperature, while it increases rapidly with particle concentra-


Vol% tion. Model 6 presented in Table 1 can be used for approximate
90 0% 0.1% 0.5% interpretation of the thermal conductivity values of water-based
Power

476 W
Nusselt number, Nu

alumina nanofluids at low concentrations. The specific heat ca-


input

80 313 W

70
218 W pacity (CP ) values of dilute alumina–water nanofluids can be
estimated using Eq. (6) with high accuracy (error < 0.5%). The
0.1%, literature [43]
60 thermal conductivity of nanofluid increased while the specific
50 heat capacity decreased with increasing the particle concen-
tration, because of higher thermal conductivity and lower heat
40 capacity of nanoparticles compared to the base fluid (water).
30 The thermal conductivity of nanofluids increases with tempera-
ture, and temperature has little effect on the CP of nanofluids in
20
0 1000 2000 3000 4000 5000 6000 7000
the range investigated (25◦ C–75◦ C). The experimental values of
friction factor for Reynolds numbers 3000 and 4500 were very
Reynolds number, Re
close to those interpreted by the Colebrook equation. Therefore,
Figure 7 A plot of Nusselt number versus Reynolds number (0 vol% refers to the Colebrook equation deduced for turbulent flow of single flu-
water).
ids can be extended to estimate the pressure drop characteristics
of dilute alumina nanofluids in transition flow. Pressure drop
tends to decrease with increasing the power input (or heat flux)
because increase in nanofluid temperature decreases its viscos-
ity. At the maximum Reynolds number (Re ≈ 6000), the friction
factor was found to be minimum and the Nusselt number was
maximum. Therefore, the prepared nanofluids are more effec-
tive under turbulent flow conditions than under transition flow.
It was also observed that the convective heat transfer coefficient
increases (and pressure drop decreases slightly) with increasing
power input.

NOMENCLATURE

CP specific heat capacity (J/kg-◦ C)


d pipe diameter (m)
Figure 8 Comparison of calculated Nusselt number with experimentally DSC differential scanning calorimetry
observed Nusselt number. f friction factor
h heat transfer coefficient (W/m2-◦ C)
correlation given by Eq. (17) is supposed to predict the depen-
k thermal conductivity (W/m-◦ C)
dence of Nusselt number on various parameters (Re, φ, and
L length of test section (m)
W in ) for different nanofluids, provided that the values of the co-
ṁ mass flow rate (kg/s)
efficient and exponents (a, b, c, and d) are correctly evaluated.
Nu Nusselt number
PID proportional-integral-derivative
Q heat transfer rate (J/s)
Re Reynolds number
CONCLUSIONS T temperature (◦ C)
t time (s)
This work uniquely reports the effect of power input on the u fluid velocity (m/s)
heat transfer and pressure drop characteristics of nanofluids. The W power input to heater (W)
experimental results indicated that the addition of nanoparticles
does not generate anomalous behavior. It was observed that the
thermal conductivity and viscosity of nanofluids are highly sen- Greek Symbols
sitive to temperature, while the density and heat capacity are
little affected by the fluid temperature. The density of nanoflu- p pressure drop (Pa)
ids increases linearly with the volume fraction of nanoparti- T temperature difference (◦ C)
cles. However, it decreases nonlinearly with increase in the ε pipe roughness (m)
nanofluid temperature. The viscosity decreases with increase in μ fluid dynamic viscosity (Pa-s)
heat transfer engineering vol. 37 no. 18 2016
R. SAXENA ET AL. 1563

ρ density (kg/m3) [9] Celata, G. P., D’Annibale, F., Mariani, A., Saraceno, L.,
φ particle volume fraction D’Amato, R., and Bubbico, R., Heat Transfer in Water-
Based SiC and TiO2 Nanofluids, Heat Transfer Engineer-
ing, vol. 34, no. 13, pp. 1060–1072, 2013.
Subscripts [10] Keblinski, P., Prasher, R., and Eapen, J., Thermal Con-
ductance of Nanofluids: Is the Controversy Over?, Journal
b bulk mean value of Nanoparticle Research, vol. 10, no. 7, pp 1089–1097,
bf base fluid (water) 2008.
cal calculated [11] Moghari, R. M., Talebi, F., Rafee, R., and Shariat, M., Nu-
en enhancement merical Study of Pressure Drop and Thermal Characteris-
exp experimental tics of Al2 O3 –Water Nanofluid Flow in Horizontal Annuli,
i inside Heat Transfer Engineering, vol. 36, no. 2, pp. 166–177,
in input 2015.
nf nanofluid [12] Williams, W., Buongiorno, J., and Hu, L.-W., Experimen-
np nanoparticles tal Investigation of Turbulent Convective Heat Transfer
o outside and Pressure Loss of Alumina/Water and Zirconia/Water
s surface Nanoparticle Colloids (Nanofluids) in Horizontal Tubes,
w water Journal of Heat Transfer, vol. 130, no. 4, pp. 042412–7,
2008.
[13] Yu, W. France, D. M., Routbort, J. L., and Choi, S. U.
REFERENCES S., Review and Comparison of Nanofluid Thermal Con-
ductivity and Heat Transfer Enhancements, Heat Transfer
[1] Das, S. K., Choi, S. U. S., and Patel, H. E., Heat Transfer in Engineering, vol. 29, no. 5, pp. 432–460, 2008.
Nanofluids—A Review, Heat Transfer Engineering, vol. [14] Sarkar, J., Ghosh, P., and Adil, A., A Review on Hybrid
27, no. 10, pp. 3–19, 2006. Nanofluids: Recent Research, Development and Applica-
[2] Choi, S. U. S., Nanofluids: A New Field of Scientific Re- tions, Renewable and Sustainable Energy Reviews, vol. 43,
search and Innovative Applications, Heat Transfer Engi- pp. 164–177, 2015.
neering, vol. 29, no. 5, pp. 429–431, 2008. [15] Cieśliński, J. T., and Kaczmarczyk, T. Z., Pool Boiling of
[3] Trisaksri, V., and Wongwises, S., Critical Review of Heat Water–Al2 O3 and Water–Cu Nanofluids Outside Porous
Transfer Characteristics of Nanofluids, Renewable and Coated Tubes, Heat Transfer Engineering, vol. 36, no. 6,
Sustainable Energy Reviews, vol. 11, no. 3, pp. 512–523, pp. 553–563, 2015.
2007. [16] Minea, A. A., Numerical Simulation of Nanoparticles Con-
[4] Chandrasekar, M., Suresh, S., and Bose, A. C., Experi- centration Effect on Forced Convection in a Tube With
mental Studies on Heat Transfer and Friction Factor Char- Nanofluids, Heat Transfer Engineering, vol. 36, no. 13,
acteristics of Al2 O3 /Water Nanofluid in a Circular Pipe pp. 1144–1153, 2015.
Under Laminar Flow With Wire Coil Inserts, Experimen- [17] Timofeeva, E. V., Routbort, J. L., and Singh, D., Particle
tal Thermal and Fluid Science, vol. 34, no. 2, pp. 122–130, Shape Effects on Thermophysical Properties of Alumina
2010. Nanofluids, Journal of Applied Physics, vol. 106, no. 1,
[5] Mohammed, H. A., Bhaskaran, G., Shuaib, N. H., and pp. 014304–10, 2009.
Saidur, R., Heat Transfer and Fluid Flow Characteristics [18] Buongiorno, J., Venerus, D. C., Prabhat, N., McKrell, T.,
in Microchannels Heat Exchanger Using Nanofluids: A Townsend, J., Christianson, R., Tolmachev, Y. V., Keblin-
Review, Renewable and Sustainable Energy Reviews, vol. ski, P., Hu, L., Alvarado, J. L., Bang, I. C., Bishnoi, S.
15, no. 3, pp. 1502–1512, 2011. W., Bonetti, M., Botz, F., Cecere, A., Chang, Y., Chen, G.,
[6] Mohammed, H. A., Gunnasegaran, P., and Shuaib, N. H., Chen, H., Chung, S. J., Chyu, M. K., Das, S. K., Paola, R.
Heat Transfer in Rectangular Microchannels Heat Sink D., Ding, Y., Dubois, F., Dzido, G., Eapen, J., Escher, W.,
Using Nanofluids, International Communications in Heat Funfschilling, D., Galand, Q., Gao, J., Gharagozloo, P. E.,
and Mass Transfer, vol. 37, no. 10, pp. 1496–1503, 2010. Goodson, K. E., Gutierrez, J. G., Hong, H., Horton, M.,
[7] Liu, L., Kim, E. S., Park, Y., -G., and Jacobi, A. M., The Hwang, K. S., Iorio, C. S., Jang, S. P., Jarzebski, A. B.,
Potential Impact of Nanofluid Enhancements on the Per- Jiang, Y., Jin, L., Kabelac, S., Kamath, A., Kedzierski, M.
formance of Heat Exchangers, Heat Transfer Engineering, A., Kieng, L. G., Kim, C., Kim, J.-H., Kim, S., Lee, S. H.,
vol. 33, no. 1, pp. 31–41, 2012. Leong, K. C., Manna, I., Michel, B., Ni, R., Patel, H. E.,
[8] Nguyen, C. T., Roy, G., Gauthier, C., and Galanis, N., Heat Philip, J., Poulikakos, D., Reynaud, C., Savino, R., Singh,
Transfer Enhancement Using Al2 O3 –Water Nanofluid for P. K., Song, P., Sundararajan, T., Timofeeva, E., Tritcak,
an Electronic Liquid Cooling System, Applied Thermal T., Turanov, A. N., Vaerenbergh, S. V., Wen, D., With-
Engineering, vol. 27, no. 8–9, pp. 1501–1506, 2007. arana, S., Yang, C., Yeh, W.-H., Zhao, X.-Z. and Zhou,
heat transfer engineering vol. 37 no. 18 2016
1564 R. SAXENA ET AL.

S.-Q., A Benchmark Study on the Thermal Conductivity mina/Water Nanofluid Under Laminar and Turbulent Flow
of Nanofluids, Journal of Applied Physics, vol. 106, no. 9, Conditions, Experimental Heat Transfer, vol. 24, no. 3, pp.
pp. 094312–14, 2009. 234–256, 2011.
[19] Chandrasekar, M., and Suresh, S., A Review on the Mech- [32] Ho, C. J., and Lin, Y.J., Turbulent Forced Convection Effec-
anisms of Heat Transport in Nanofluids, Heat Transfer tiveness of Alumina–Water Nanofluid in a Circular Tube
Engineering, vol. 30, no. 14, pp. 1136–1150, 2009. With Elevated Inlet Fluid Temperatures: An Experimental
[20] Jang, S. P., and Choi, S. U. S., Role of Brownian mo- Study, International Communications in Heat and Mass
tion in the Enhanced Thermal Conductivity of Nanofluids, Transfer, vol. 57, pp. 247–253, 2014.
Applied Physics Letters, vol. 84, no. 21, pp. 4316–4318, [33] Ciloglu, D., and Bolukbasi, A., A Comprehensive Review
2004. on Pool Boiling of Nanofluids, Applied Thermal Engineer-
[21] Heris, S. Z., Esfahany, M. N., and Etemad, S. G., Ex- ing, vol. 84, pp. 45–63, 2015.
perimental Investigation of Convective Heat Transfer of [34] Wen, D., and Ding, Y., Experimental Investigation Into the
Al2 O3 /Water Nanofluid in Circular Tube, International Pool Boiling Heat Transfer of Aqueous Based γ-Alumina
Journal of Heat and Fluid Flow, vol. 28, no. 2, pp. Nanofluids, Journal of Nanoparticle Research, vol. 7, no.
203–210, 2007. 2–3, pp. 265–274, 2005.
[22] Wen, D., and Ding, Y., Experimental Investigation Into [35] Sundar, L. S., and Sharma, K. V., Turbulent Heat Transfer
Convective Heat Transfer of Nanofluids at the Entrance and Friction Factor of Al2 O3 Nanofluid in Circular Tube
Region Under Laminar Flow Conditions, International With Twisted Tape Inserts, International Journal of Heat
Journal of Heat and Mass Transfer, vol. 47, no. 24, pp. and Mass Transfer, vol. 53, no. 7–8, pp. 1409–1416, 2010.
5181–5188, 2004. [36] Haddad, Z., Abid, C., Oztop, H. F., and Mataoui, A., A
[23] Das, S. K., Putra, N., Thiesen, P., and Roetzel, W., Temper- Review on How the Researchers Prepare Their Nanofluids,
ature Dependence of Thermal Conductivity Enhancement International Journal of Thermal Sciences, vol. 76, pp.
For Nanofluids, Journal of Heat Transfer, vol. 125, no. 4, 168–189, 2014.
pp. 567–574, 2003. [37] Suresh, S., Selvakumar, P., Chandrasekar, M., and Raman,
[24] Wang, X.-Q., and Mujumdar, A. S., Heat Transfer Char- V. S., Experimental Studies on Heat Transfer and Friction
acteristics of Nanofluids: A Review, International Journal Factor Characteristics of Al2 O3 /Water Nanofluid Under
of Thermal Sciences, vol. 46, no. 1, pp. 1–19, 2007. Turbulent Flow With Spiraled Rod Inserts, Chemical En-
[25] Wen, D., Lin, G., Vafaei, S., and Zhang, K., Review of gineering and Processing: Process Intensification, vol. 53,
Nanofluids for Heat Transfer Applications, Particuology, pp. 24–30, 2012.
vol. 7, no. 2, pp. 141–150, 2009. [38] Duangthongsuk, W., and Wongwises, S., Effect of Ther-
[26] Adio, S. A., Sharifpur, M., and Meyer, J. P., Investigation mophysical Properties Models on the Predicting of the
Into Effective Viscosity, Electrical Conductivity, and pH Convective Heat Transfer Coefficient for Low Concentra-
of γ-Al2 O3 -Glycerol Nanofluids in Einstein Concentration tion Nanofluid, International Communications in Heat and
Regime, Heat Transfer Engineering, vol. 36, no. 14-15, pp. Mass Transfer, vol. 35, no. 10, pp. 1320–1326, 2008.
1241–1251, 2015. [39] Chieruzzi, M., Cerritelli, G. F., Miliozzi, A., and Kenny, J.
[27] Sekhar, Y. R., and Sharma, K. V., Study of Viscosity M., Effect of Nanoparticles on Heat Capacity of Nanoflu-
and Specific Heat Capacity Characteristics of Water-Based ids Based on Molten Salts as PCM For Thermal Energy
Al2 O3 Nanofluids at Low Particle Concentrations, Journal Storage, Nanoscale Research Letters, vol. 8, no. 1, pp.
of Experimental Nanoscience, vol. 10, no. 2, pp. 86–102, 448–449, 2013.
2015. [40] Teng, T.-P., and Hung, Y.-H., Estimation and Experimen-
[28] Pak, B., and Cho, Y. I., Hydrodynamic and Heat Transfer tal Study of the Density and Specific Heat for Alumina
Study of Dispersed Fluids With Submicron Metallic Oxide Nanofluid, Journal of Experimental Nanoscience, vol. 9,
Particles, Experimental Heat Transfer, vol. 11, no. 2, pp. no. 7, pp. 707–718, 2014.
151–170, 1998. [41] Sommers, A. D., and Yerkes, K. L., Experimental Inves-
[29] Torii, S., and Yang, W.-J., Heat Transfer Augmentation tigation Into the Convective Heat Transfer and System-
of Aqueous Suspensions of Nanodiamonds in Turbulent Level Effects of Al2 O3 –Propanol Nanofluid, Journal of
Pipe Flow, Journal of Heat Transfer, vol. 131, no. 4, pp. Nanoparticle Research, vol. 12, no. 3, pp. 1003–1014,
043203–5, 2009. 2010.
[30] Ho, C. J., Huang, J. B., Tsai, P. S., Yang, Y. M., Water- [42] Sarkar, J., A Critical Review on Convective Heat Transfer
Based Suspensions of Al2 O3 Nanoparticles and MEPCM Correlations of Nanofluids, Renewable and Sustainable
Particles on Convection Effectiveness in a Circular Tube, Energy Reviews, vol. 15, no. 6, pp. 3271–3277, 2011.
International Journal of Thermal Sciences, vol. 50, no. 5, [43] Sharma, K. V., Sundar, L. S., and Sarma, P. K., Estimation
pp. 736–748, 2011. of Heat Transfer Coefficient and Friction Factor in the
[31] Chandrasekar, M., and Suresh, S., Experiments to Explore Transition Flow With Low Volume Concentration of Al2 O3
the Mechanisms of Heat Transfer in Nanocrystalline Alu- Nanofluid Flowing in a Circular Tube and With Twisted
heat transfer engineering vol. 37 no. 18 2016
R. SAXENA ET AL. 1565

Tape Insert, International Communications in Heat and [51] Zhang, L., Lv, J., Bai, M., and Guo, D., Effect of Vibra-
Mass Transfer, vol. 36, no. 5, pp. 503–507, 2009. tion on Forced Convection Heat Transfer for SiO2 –Water
[44] Heris, S. Z., Nassan, T. H., Noie, S. H., Sardarabadi, H., Nanofluids, Heat Transfer Engineering, vol. 36, no. 5, pp.
and Sardarabadi, M., Laminar Convective Heat Transfer of 452–461, 2015.
Al2 O3 /Water Nanofluid Through Square Cross-Sectional
Duct, International Journal of Heat and Fluid Flow, vol. Richa Saxena is an assistant professor in the De-
partment of Petroleum Engineering, DIT Univer-
44, pp. 375–382, 2013.
sity, Dehradun, Uttarakhand, India. She received her
[45] Murshed, S. M. S., Determination of Effective Specific M.Tech. degree from Thapar University, Patiala, In-
Heat of Nanofluids, Journal of Experimental Nanoscience, dia, in 2013 and B.Tech. degree from Moradabad
vol. 6, no. 5, pp. 539–546, 2011. Institute of Technology, Moradabad, India, in 2010.
[46] Rashmi, W., Ismail, A. F., Sopyan, I., Jameel, A. T., Yusof, She worked on heat transfer and pressure drop char-
F., Khalid, M., and Mubarak, N. M., Stability and Thermal acteristics of nanofluids during her master’s degree
as her thesis work.
Conductivity Enhancement of Carbon Nanotube Nanofluid
Using Gum Arabic, Journal of Experimental Nanoscience,
Dasaroju Gangacharyulu is a professor in the De-
vol. 6, no. 6, pp. 567–579, 2011. partment of Chemical Engineering, Thapar Univer-
[47] Amrollahi, A., Rashidi, A. M., Meibodi, M. E., and sity, Patiala, India. He is also the Controller of Ex-
Kashefi, K., Conduction Heat Transfer Characteristics and aminations for the university. He has more than
Dispersion Behaviour of Carbon Nanofluids as a Func- 20 years of experience in teaching, research and in-
dustry. He has specialized in heat transfer and fluid
tion of Different Parameters, Journal of Experimental
flow, nanofluids, heat pipes, hydrogen energy, en-
Nanoscience, vol. 4, no. 4, pp. 347–363, 2009. ergy management, energy storage, design of heat ex-
[48] Hwang, K. S., Jang, S. P., and Choi, S. U. S., Flow changers, thermal engineering, process design, pro-
and Convective Heat Transfer Characteristics of Water- cess modeling, and simulation. His research interests
Based Al2 O3 Nanofluids in Fully Developed Laminar Flow include heat transfer enhancement using nanofluids and energy conservation and
integration.
Regime, International Journal of Heat and Mass Transfer,
vol. 52, no. 1–2, pp. 193–199, 2009.
[49] Fotukian, S. M., and Esfahany, M. N., Experimental Inves- Vijaya Kumar Bulasara is an assistant professor
in the Department of Chemical Engineering, Thapar
tigation of Turbulent Convective Heat Transfer of Dilute University, Patiala, India. He received his M.Tech.
γ-Al2 -O3 /Water Nanofluid Inside a Circular Tube, Inter- and Ph.D. degrees in chemical engineering from In-
national Journal of Heat and Fluid Flow, vol. 31, no. 4, dian Institute of Technology, Guwahati, India, in the
pp. 606–612, 2010. years 2008 and 2011, respectively. He has special-
ized in heat transfer in nanofluids, design of heat
[50] Duangthongsuk, W., and Wongwises, S., An Experi-
exchanger networks, electroless plating, reaction en-
mental Study on the Heat Transfer Performance and gineering and catalysis, adsorption, and membrane
Pressure Drop of TiO2 –Water Nanofluids Flowing Un- separation. His research interests include heat trans-
der a Turbulent Flow Regime, International Journal of fer enhancement using nanofluids and wastewater treatment. He is an editorial
Heat and Mass Transfer, vol. 53, no. 1–3, pp. 334–344, board member for the International Journal of Chemical Research, Journal of
Catalyst & Catalysis, and Trends in Chemical Engineering.
2010.

heat transfer engineering vol. 37 no. 18 2016

You might also like