Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

International Communications in Heat and Mass Transfer 58 (2014) 85–95

Contents lists available at ScienceDirect

International Communications in Heat and Mass Transfer


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

New thermophysical properties of water based TiO2 nanofluid—The


hysteresis phenomenon revisited☆
Z. Said a,b, R. Saidur a,c,⁎, A. Hepbasli d, N.A. Rahim c
a
Department of Mechanical Engineering, Faculty of Engineering, University of Malaya, 50603 Kuala Lumpur, Malaysia
b
Department of Mechanical and Materials Engineering (MME), Masdar Institute of Sicence and Technology, 54224 Abu Dhabi, United Arab Emirates
c
UM Power Energy Dedicated Advanced Centre (UMPEDAC), Level 4, Wisma R & D, University of Malaya, 50603 Kuala Lumpur, Malaysia
d
Department of Energy Systems Engineering, Faculty of Engineering, Yasar University, 35100 Bornova, Izmir, Turkey

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

Available online 3 September 2014 Homogeneous stable suspensions acquired by dispersing dry Al2O3 and TiO2 nanoparticles in controlled pH
solution and distilled water, respectively, were prepared and investigated in this study. First of all, the mean
Keywords: nanoparticle diameters were studied by dynamic light scattering (DLS) technique, and the nanofluid stability
Nanofluid was analyzed by zeta potential measurements. The nano-crystalline structures were characterized by scanning
Thermal conductivity electron microscope and transmission electron microscope. The rheological behavior was determined for both
Viscosity
nanofluids at nanoparticle volume concentrations up to 0.3%. The effect of temperature for the heating and
Hysteresis
Density
cooling phases was analyzed from 25 °C to 80 °C. Furthermore, the influence of temperature, pressure drop,
Pumping power pumping power, zeta potential, size and densities were analyzed for fresh prepared samples as well as for
samples used in a flat plate solar collector over a period of 30 days. The thermal conductivity enhancement of
the two nanofluids demonstrated a nonlinear relationship with respect to temperature and volume fraction,
with increases in the volume fraction and temperature. All resulted in an increase in the measured enhancement.
Existence of a critical temperature was observed beyond which the particle suspension properties altered dras-
tically, which in turn triggered a hysteresis phenomenon. The hysteresis phenomenon on viscosity measurement,
which is believed to be the first observed for Al2O3/water and TiO2/water-based nanofluids, has raised serious
concerns about the use of nanofluids for heat transfer enhancement. The pressure drop and pumping power of
the nanofluid flows are found to be very close to those of the base liquid for low volume concentration. It may
be concluded that nanofluids can be utilized as a working medium with a negligible effect of enhanced viscosity
and/or density. Our findings provide a view on the thermo physical properties of nanofluids that is compared
with that in the literature, and new findings (such as viscosity, hysteresis phenomenon and pumping power)
have been presented, which are not available in literature as yet.
© 2014 Elsevier Ltd. All rights reserved.

1. Introduction are immersed in base fluids. The main reason for this is to enhance
the heat transfer characteristics of conventional fluids by improving
Conventional fluids, such as water, engine oil, and ethylene their thermal conductivity. In the previous decade, nanofluids have
glycol, are usually used as heat transfer fluids. Their poor heat trans- achieved considerable devotion due to their enhanced thermal con-
fer rate is understood as an obstacle for enhancing efficiency of heat ductivities. In this regard, Eastman et al. [1] reported that the ther-
exchangers. A novel type of heat transfer fluids called “nanofluids” is mal conductivity of the conventional fluid upsurges by 40%, when
recognized for enhancement of heat exchanges for better perfor- 0.3% of copper nanoparticles were suspended in ethylene glycol.
mance. Nanofluids are two phase fluids where solid nanoparticles Pak and Cho [2] carried out an experimental work for the determina-
tion of forced convection heat transfer coefficients with 13 nm Al2O3
and 27 nm TiO2 sub micron particles dispersed in water. For a fixed
Abbreviations: FESEM, field emission scanning electron microscopy; SEM, scanning Reynolds number, the convective heat transfer coefficient was
electron microscopy; TEM, transmission electron microscopy. improved by 75% for an Al2O3 particle concentration of 2.78%. Heat
☆ Communicated by W.J. Minkowycz. transfer coefficients were also observed to increase with concentra-
⁎ Corresponding author at: Department of Mechanical Engineering, Faculty of
tion. Such outcomes have driven both the industrial and scientific
Engineering, University of Malaya, 50603 Kuala Lumpur, Malaysia.
E-mail addresses: saidur@um.edu.my, saidur912@yahoo.com, zaffar.ks@gmail.com community to examine the heat transfer and rheological properties
(R. Saidur). of nanofluids.

http://dx.doi.org/10.1016/j.icheatmasstransfer.2014.08.034
0735-1933/© 2014 Elsevier Ltd. All rights reserved.
86 Z. Said et al. / International Communications in Heat and Mass Transfer 58 (2014) 85–95

temperature detecting hysteresis phenomenon [5,24–26]. The viscosity


Nomenclature of the nanofluid is among one of the physical properties, which is ex-
tremely related to temperature. During the heating phase outside the
d Diameter (m) critical temperature, nanofluids might become more viscous, which
DH hydraulic diameter (m) designates a rather radical change of the nanofluids' rheological proper-
f friction factor ties resulting in hysteresis. Nguyen et al. [25] experimentally examined
K loss coefficient temperature and concentration effects of aqueous Al2O3 nanofluids.
k thermal conductivity, W/mK Hysteresis conduct of dynamic viscosity due to temperature effect was
Re Reynolds number detected, which was with higher volume concentrations. They disclosed
T temperature, °C that at any operating temperature, the viscosity values were greater for
Δp pressure drop, Pa the cooling phase compared to the value of the heating phase.
Recently, a review on the application of nanofluids in numerous
kinds of heat exchangers has been reported by Huminic and Huminic
Greek symbols [27]. They concluded that both the thermo physical properties and
ρ density, kgm−3 form of flow within the heat exchanger played an imperative role in
ϕ volumetric fraction the efficiency of the nanofluid as a coolant. Furthermore, the heat
μ volumetric fraction transfer fluid was not immobile in most of the practical applications
η Intrinsic viscosity [27]. Thus, the rheological properties and their study were also cru-
δ thickness of nano-layer cial to obtain the enhancements on the average heat transfer coeffi-
cients of the flowing system suitably, which usually increased with
the volume fraction of the nanoparticles as well as with the Reynolds
Subscripts number [28].
bf base fluid Viscosity is directly related to pumping power and pressure
nf nanofluid drop for a solar collector. Numerical results indicate that high con-
p Particle centration of nanoparticles (Al2O3 or TiO2) in water displays a higher
heat transfer improvement and also higher pressure drops [29].
Fotukian et al. [30] experimentally examined the influence of dilute
Al2O3/water nanofluid on the heat transfer, and pressure drop inside
In recent years, a lot of work has been done on solar energy. Fossil a circular tube under turbulent flow condition. Their work showed
fuels and their utilization will be coming limited due to their lack in that with the addition of a little volume of nanoparticles to the
reserve as well as for the environmental deliberations. Due to these base fluid improved the heat transfer coefficient extraordinarily.
reasons, researchers are inspired to look for other potential sources While testing the viscosity of TiO2/water nanofluids, Newtonian be-
of energy. One of the foremost concerns of energy saving and havior for all the concentrations was observed by Bobbo et al. [14]
compact design in solar thermal collectors is the heat transfer im- and Penkavova et al. [31]. However, He et al. [32] stated that aqueous
provement. Solar thermal energy has been extensively utilized for TiO 2 nanofluids, with anatase phase and a little amount of rutile
numerous applications, such as electricity generation, chemical pro- phase, resulted in a shear thinning behavior where the shear viscos-
cessing, and thermal heating, due to its renewable and nonpolluting ity inclined to be continuous at shear rates above 100/s and also the
nature [3]. The literature available on using nanofluids as heat trans- pressure drop of these nanofluids was very close to that of the base
fer fluids in solar collectors are very few in numbers up to date liquid. However, Tseng and Lin [33] reported a pseudo plastic flow
[4–12]. for most of the shear rates inspected, from 10 to 1000/s for 0.0 to
Al2O3 and TiO2 based nanofluids are essential because of their avail- 0.12 vol.% of anatase TiO2 nanoparticles.
ability and unique thermal properties. Al2O3 and TiO2 nanofluids, which As pointed out above, few studies were focused on the thermo phys-
are prepared by a two-step method using an ultrasonic vibrator, result ical behavior of Al2O3/water and TiO2/water nanofluids. The objective of
unstable. Different methods have been chosen by a number of investiga- this present investigative effort is to examine the effect of density, ther-
tors for making stable nanofluids using numerous surfactants, improving mal conductivity and viscosity of water based alumina and titanium
the pH, temperature for numerous nanofluids, and by surface modifica- nanofluids (0.05 % b φ b 0.3%) on pressure drop and pumping power
tion of the particles. However, reports on Al2O3 and TiO2 based nanofluids for a flat plate solar collector. The properties are measured for the
prepared by high pressure homogenizer are very few [13,14]. solutions prepared before using them in the system and after using
Various researchers [15–18] have studied for thermal conductivity them in the system for 30 days. No such work has been carried out till
improvements. An enhancement was noticed in the thermal conductiv- date to the best of the authors' knowledge.
ity with adding of nanoparticles in the fluid from all the investigation
data [19]. An improvement of 60% in the thermal conductivity asso- 2. Materials and methodology
ciated with the consistent base fluids for 5%v/v of nanoparticles
concentration was obtained by Eastman et al. [20]. The effective In this section, details of the experimental procedure have been stat-
thermal conductivity of Al2O3 and CuO nanoparticles suspended in ed while the physical properties of Al2O3 and TiO2 nanoparticles, and
base fluids (water, ethylene glycol, vacuum pump oil and engine oil) the base fluid have been indicated. Materials, data collection, nanoparti-
was investigated by Wang et al. [21]. An improvement of 12% was cle characterization, preparation of nanofluids, experimental data
achieved in the effective thermal conductivity by them with 3%v/v con- acquisition and analytical study of effect of nanoparticles on flat plate
centration of nanoparticles; however, 20% improvement was reported solar collectors are discussed in the subsections below.
by Masuda et al. [22], and 8% was reported by Lee et al. [23] for the
identical volume concentration of particles. Hence, dissimilarities in 2.1. Materials
the experimental reports regarding the thermal conductivity are still
present in great extent, which make room to investigate properly Al2O3 (99.8% trace metal's basis) and TiO2 (99.5% trace metal's basis)
about the property. were purchased from Sigma-Aldrich. Distilled water was used as base
Very few studies described the influence of temperature on fluids. Hydrochloric acid (HCl-37%) was also purchased from Sigma-
viscosity for both at higher and lower volume fraction and at a higher Aldrich as well.
Z. Said et al. / International Communications in Heat and Mass Transfer 58 (2014) 85–95 87

Table 1
Analytical models on the thermal conductivity of nanofluids [23,26–30].

Researcher Formula (keff/kb) Relation no. Remarks

Maxwell [26]   Relates to the thermal conductivity of spherical particles,


kp þ 2kb þ 2 kp −kb ϕ
  ð1Þ solid volume fraction and base fluid
kp þ 2kb − kp −kb ϕ

 
3
Yu and Choi [27,28] kpe þ 2kb þ 2 kpe −kb ð1−β Þ ϕ Modified Maxwell model, modified Hamilton–Crosser model
  ;
kpe þ 2kb − kpe −kb ð1 þ β Þ3 ϕ

nϕeff A
1þ ð2Þ
1−ϕeff A

2
Xie et al. [29] 2Θ2 ϕT Effect of nanolayer is included
1 þ 3ΘϕT þ ð3Þ
1−ΘϕT

knf
Effective medium theory [23,30] ¼ 1 þ 3φ ð4Þ
k0

2.2. Nanofluid preparation and characterization 2.3. Thermal conductivity calculation and measurement

The nanofluids were prepared following a two-step method. Ultra- Presently, there is no reliable theory to predict the anomalous ther-
sonicator and high pressure homogenizer (capacity of up to 2000 bar) mal conductivity of nanofluids. From the experimental results of many
were used to dissolve the nanoparticles (0.1% and 0.3%v/v) into distilled researchers, it is recognized that the thermal conductivity of nanofluids
water. depends on parameters, including the thermal conductivities of base
Field emission scanning electron microscopy (FESEM) from the fluid and nanoparticles, volume fraction, surface area, shape of nanopar-
SIGMA Zeiss instrument (Carl Zeiss SMT Ltd., UK) and transmission ticles, and temperature. To predict the effective thermal conductivity of
electron microscope (TEM) were used to achieve the morphological nanofluids, a number of theoretical and empirical models have been
characterization of the nanoparticles. A zeta-seizer Nano ZS (Malvern) proposed [12,22–25]. Table 1 contains some popular models, which
was used to analyze the average size of the nanoparticles in base will be used to compare our experimental results later. KD2 Pro thermal
mediums. It gave the hydrodynamic radius of the particles dispersed property analyzer (Decagon, USA) was used to measure thermal con-
in a medium using DLS approach. ductivity of the nanofluids.

Table 2
Most cited correlations of nanofluid viscosity [25,34–39].

Model Correlation Correlation no. Remarks

Einstein [34] μ nf ¼ μ bf :ð1 þ 2:5ϕÞ ð5Þ Valid for very low volume concentrations
(ϕ ≤0.02) and spherical particles
 
Brinkman [35] 1 Formulated by two corrections of Einstein's model
μ nf ¼ μ bf : ð6Þ
ð1−ϕÞ2:5

 
Batchelor [36] 2 Considered the effect of Brownian motion.
μ nf ¼ μ bf : 1 þ 2:5ϕ þ 6:5ϕ ð7Þ
Extension of the Einstein model.

Abu-Nada et al. [37] μ Al2 O3 ¼ expð3:003−0:04203T Includes temperature T and volume fraction ϕ.
2 Modified model of Nguyen.
−0:5445ϕ þ 0:0002553T
2 −1
−0:0534ϕ −1:622ϕ Þ ð8aÞ

μ H2 O ¼ −81:1 þ 98:75 ln ðT Þ
2 3
−45:23 ln ðT Þ þ 9:71 ln ðT Þ
4 5
−0:946 ln ðT Þ þ 0:03 ln ðT Þ ð8bÞ

  
Masoud Hossein et al. [38] T Based on Nguyen et al. [25] experimental data.
μ nf ¼ μ bf : exp m þ α þ β ð ϕÞ
T0
 
dp
þγ ð9aÞ
1þR

α ¼ −0:485; β ¼ 14:94; γ ¼ 0:0105;



m ¼ 0:72; T 0 ¼ 20 C; ð9bÞ

R ¼ 1nm ð9cÞ

 
2
Maiga model [39] μ nf ¼ μ bf : 123ϕ þ 7:3ϕ þ 1 ð10Þ Derived from Al2O3/water nanofluids.
88 Z. Said et al. / International Communications in Heat and Mass Transfer 58 (2014) 85–95

2.5. Pumping power calculation

The flow of nanofluid in the system was considered as forced flow. A


pump is required to circulate nanofluids all the way through the system.
The pump would need electrical energy. It is critical to understand the
entire energy required by the pump to sustain a constant flow across
the collector. The pumping power and pressure drop were calculated
using the following procedure [40,41]:

ρV 2 Δl ρV 2
Δp ¼ f þK ð11Þ
2 d 2

The loss coefficient K is often taken from table's results from tests or
calculated using formulas that consist of the density and kinematic
Fig. 1. Particle diameter distribution, according to the intensity, for water–Al2O3 0.1% viscosity of the heat transfer fluid. V is the mean flow velocity of
nanofluids with pH 9, just after preparation, after 7 days and after 30 days. nanofluids in the system, and is given by


V¼ ð12Þ
ρnf πD2H =4

In current investigational consideration, DH = pipe diameter (d). ρnf


was calculated from

ρnf ¼ φρnp þ ð1−φÞρbf ð13Þ

The frictional factor, f, for laminar and turbulent flow, correspond-


ingly, is as follows [41]

64
f ¼ for laminar flow
Re

0:079
Fig. 2. Particle diameter distribution, according to the intensity, for water–TiO2 0.1% f ¼ for turbulent flow:
nanofluids, just after preparation, after 7 days and after 30 days. ðReÞ1=4

The Reynolds number is


2.4. Viscosity calculation and measurement
ρVDH
Re ¼ ð14Þ
μ
The viscosities of nanofluids were calculated using the most cited for-
mulas of viscosities to compare the experimental results, as shown in
Now, the pumping power can be calculated using the following
Table 2. The viscosity of nanofluid was measured using a Brookfield Vis-
equation:
cometer (DV-II + Pro Programmable Viscometer), which was connected
to a temperature-controlled bath. For experimental uncertainties in vis- !

cosities the measured data was checked against the data from ASHARE Pumping power ¼  Δp
ρnf
Handbook [31].

Fig. 3. Particle diameter distribution (Z-Average size: 196.4 nm), according to the intensity, for water–Al2O3 0.1% nanofluids with pH 9, after running for 30 days in a system.
Z. Said et al. / International Communications in Heat and Mass Transfer 58 (2014) 85–95 89

Fig. 4. Particle diameter distribution (Z-Average size: 225.9 nm) according to the intensity, for water–TiO2 0.1%v/v nanofluids, after running for 30 days in a system.

3. Results and discussions for water–Al2O3 was (Z-Average size: 109.4 nm) with an initial zeta
potential value of 58.4 (mV), whereas the particle diameter increased
Fig. 1 shows the particle size distribution according to the intensity to (Z-Average size: 196.4 nm), with an increment of 87 nm and zeta
obtained from the zeta-seizer at a different interval of days while TEM potential value reduced to 40.2 (mV) after the solution was used in
and SEM images are indicated in Figs. 2 and 3, respectively. a flat plate solar collector for 30 days. In the case of water–TiO2 0.1%
Fig. 1 shows a size distribution graph of the aggregates in the nanofluids, the initial particle diameter was (Z-Average size: 126.9 nm)
nanofluid at different times after preparation. The peak of the distri- with an initial zeta potential value of 48.6 (mV), whereas the particle
butions is at almost identical horizontal position. Their intensity, diameter increased to (Z-Average size: 225.9 nm), with an increment of
however, increases vertically witnessing an increase in population 99 nm and zeta potential value reduced to 36.7 (mV) after the solution
of aggregates. was used in a flat plate solar collector for 30 days.
Similar pattern as in Fig. 1 is observed here in Fig. 2. The elevation of
the distributions is at almost identical horizontal position; however, 3.1. Thermal conductivity
their intensity rises vertically witnessing a growth in population of
aggregates, which is larger compared to Al2O3–water. Therefore, it can Experimentally calculated thermal conductivity of the pure base
be said from the particle diameter graph that Al2O3–water nanofluid is fluids with our device presented a good agreement with the available
more stable compared to TiO2–water. literature data. From Figs. 7 and 8, it is witnessed that the thermal con-
Figs. 3 and 4 represent the particle diameter distribution of water– ductivity grows with the increment in volume fraction, which is almost
Al2O3 0.1% nanofluids with pH 9 and water–TiO2 0.1% nanofluids, after linear. A small dissimilarity is detected in measured thermal conductiv-
running for 30 days in a system, respectively. Initial particle diameter ity associated with the projected values obtained from the effective

Fig. 5. (a) SEM image of Al2O3 nanoparticles, (b) SEM image of TiO2 nanoparticle, (c) TEM image of Al2O3 nanoparticles with controlled pH and (d) TEM image of TiO2 nanoparticles.
90 Z. Said et al. / International Communications in Heat and Mass Transfer 58 (2014) 85–95

Fig. 6. Visual images of Al2O3 & TiO2 right after preparation and after being used in a flat plate solar collector for 30 days with 0.1% and 0.3% volume fraction.

medium theory [19,42]. As perceived from the TEM images in Fig. 5, that material properties of both particle and carrier fluid was attributed to
the particles are not wholly spherical, thus, resulting in a lower thermal the long impact range of the antiparticle potential, which influenced the
conductivity. For spherical nanoparticles and a lower volume fraction, particle motion. Visual images of Al2O3 & TiO2 right after preparation
thermal conductivity was projected to be greater compare to other and after being used in a flat plate solar collector for 30 days with 0.1%
shaped nanoparticles. and 0.3% volume fraction are presented in Fig. 6. No sign of sedimentation
It is perceived from these models that the projected values were not was noticed for this period of study, through visual means as well.
nearly comparable to the experimental values. The thermal behavior of Fig. 7 shows the thermal conductivity enhancement in water with
the nanofluids is depended on the Brownian motions at nano-scale and Al2O3 nanoparticles suspended in it with respect to different volume
molecular level, which was reported by Jang et al. [43]. concentration, which is similar with the results published in the
In summary, it is problematic to classify a recognized theory to predict literature [23,46,47]. The deviation of the experimental values of the
precisely heat transfer characteristics of nanofluids. Numerous investiga- estimated values of thermal conductivity of Al2O3/water nanofluids is
tors deal with the nanofluids as single-phase fluid rather than a two phase considerably high. Effective medium theory estimates a high value. Al-
mixture. The particle–liquid interaction and the movement between the though Timofeeva et al. [51] concluded in their work that the effective
particle and liquids, however, play significant parts in affecting the con- medium theory is adequate for measuring the thermal conductivity of
vective heat transfer performance of nanofluids [44]. The strong depen- nanofluids, which was found to be applicable for Al2O3/EG nanofluids
dence of the effective thermal conductivity on the temperature and [5] rather than Al2O3/water nanofluids.
Fig. 8 shows the effects of temperature and different volume concen-
trations. It is observed from the figure that the thermal conductivity

Fig. 7. Thermal conductivity of Al2O3/water nanofluids at different volume fractions and at


25 °C. Fig. 8. Effect of temperature on thermal conductivity of base fluids [45].
Z. Said et al. / International Communications in Heat and Mass Transfer 58 (2014) 85–95 91

Fig. 9. Uncertainty in experimental data of viscosity of Al2O3–water nanofluid with 13 nm, 0.1% volume concentration and at 35 °C.

increases with the increasing volume concentration as well as the rising Uncertainty in viscosity measurement of water based alumina
temperature. These results followed a similar pattern as that of Fedele nanofluid is presented in Fig. 9 and uncertainty in viscosity measure-
et al. [45]. The experimental results demonstrated a peak in the en- ment of water based titanium nanofluid is presented in Fig. 10. The ef-
hancement factor in this range of volume fractions in the temperature fect of temperature and volume fraction on viscosities of Al2O3–water
range evaluated, which implies that an optimal size exists for different nanofluid is presented in Fig. 11. For the nanofluid samples after the
nanoparticle and base fluid combinations. This phenomenon can be use in a system over a period of 30 days, the water-based alumina
neither predicted nor explained using the theoretical models currently nanofluid at 0.3%v/v showed Newtonian behavior below temperature
available within the literature. 30 °C, whereas above this temperature, non-Newtonian behavior was
observed, as shown in Fig. 11.
3.2. Rheological behavior of nanofluids For the nanofluid sample with 0.1%v/v, the water-based alumina
nanofluid behaved as a Newtonian fluid above 40 °C, whereas below
As discussed earlier, only a few numbers of studies have been carried this temperature, it behaved as a non-Newtonian fluid, presented in
out about the rheological behavior of nanofluids in the past decade, and Fig. 12. As it is observed from the figure, with continuous running of
there are contradictions such as Newtonian and non-Newtonian behav- the solution in the system, the viscosity of the solution became almost
iors stated for the similar nanofluid as well as inconsistencies in the effects a constant value with few changes compared with the viscosity of the
of temperature, particle size, shape and great shear viscosity values solution, which was taken right after preparation with higher changes
[48–51]. In this context, an important matter is to achieve nanofluid phys- in values with the same temperature.
ical information, and one of the possible techniques is through a compre- From 55 °C, the TiO2 nanofluid with 0.1% volume fractions was
hensive rheological investigation [52]. In this experiment, three kinds of Newtonian, whereas below this temperature, TiO2 nanofluid with 0.1%
studies, namely viscosity as a function of shear rate, temperature as well was non-Newtonian, presented in Fig. 14. For 0.3%v/v of TiO2, a
as at different mass concentrations, have been conducted. Newtonian behavior was observed in the entire temperature range in
It is worth noting that the aforementioned equations were Fig. 13.
established to relative viscosity as a function of particle volume fraction
only. There is no equation to narrate temperature. Furthermore, these 3.3. Hysteresis loop
equations are for homogenous fluid and do not take into account parti-
cle agglomeration effect. Nanofluids showed lower viscosities when the We have noted that the viscosity of nanofluid reduces with an in-
nanoparticles were dispersed in base fluids. crease in the temperature without coming back through the same
In a Newtonian fluid, the relation between the shear stress and the path but was deviated upward when it was allowed to cool down. The
shear rate is linear, passing through the origin; the constant of propor- phenomena are presented in Figs. 15 and 16.
tionality is called the coefficient of viscosity. In a non-Newtonian fluid, For water based alumina nanofluids with 0.1%v/v concentrations, the
the line never passes through the origin. intersection point occurs at 80 °C, whereas in case of water based

Fig. 10. Uncertainty in experimental data of viscosity of TiO2–water nanofluid with 13 nm, 0.1% volume concentration and at 35 °C.
92 Z. Said et al. / International Communications in Heat and Mass Transfer 58 (2014) 85–95

the impact of hysteresis was very low, compared to higher volume con-
centrations, which was investigated by Nguyen et al. [25,26]. The temper-
ature difference from the point of interaction was noticed for the same
volume concentration before and after using the solutions in the system.
The reason behind this shift could be a possible result from the aggrega-
tion of the nanoparticles, which have shown increment in their particle
size diameter over a period during a month (as indicated in Figs. 1 to 4
as well as the effect of rusting on the nanoparticles, which affected the
pH of the solution and hence resulted in particle aggregation and
affecting the stability of the solution for the long-term in a system.
The experimental results reveal that the path for 0.1% intersects at a
temperature of 65.2 °C for the TiO2/water nanofluid and for 0.3%, the
path intersects at 68 °C. Such a phenomenon still remains not very
well understood. It was noted from the measured results that the
extraordinary improvement of nanofluid viscosity happened through
Fig. 11. Effect of temperature and volume fraction on viscosities of Al2O3–water
nanofluids. the cooling phase of the suspensions. No work is accessible to clarify
this behavior of nanofluids.
alumina nanofluids with 0.1% & 0.3%v/v concentrations, the intersection
point occurs at 67 °C and 82 °C, respectively. Hence, one can say that the 3.4. Pumping power and pressure drop
temperature level, at which the suspensions were heated, was very sig-
nificant for the possible impact of the viscous behavior of nanofluids. It is vital to study the flow resistance of nanofluids in order to im-
Since very low volume concentrations of nanoparticles was considered, prove their heat-transfer characteristics, so that the nanofluid can be

Fig. 12. Viscosity and shear stress versus shear strain rate for 0.05%v/v & 0.1%v/v Al2O3/H2O nanofluids at 35 °C.

Fig. 13. Viscosity of TiO2–water nanofluid with different volume concentrations and different temperatures.
Z. Said et al. / International Communications in Heat and Mass Transfer 58 (2014) 85–95 93

Fig. 14. Viscosity and shear stress versus shear strain rate of 0.05% to 0.5%v/v TiO2/H2O nanofluids at 40 °C.

appropriately used in the solar collector. The pressure drops of water at temperatures ranging from 25 °C to 80 °C. We have also listed some
and EG/water mixture based alumina nanofluids in a flat plate solar col- key concluding remarks as follows:
lector were theoretically examined considering the laminar flow.
Nanofluids with Al2O3 & TiO2 nanoparticle volume concentrations a. The effect of pH on the stability of the Al2O3–H2O suspension was
are engaged in the pressure drop calculation. Fig. 17 presents a variation critical. At pH 9.0, a good dispersion of alumina particles was obtain-
of the pumping power and the pressure drop with volume concentra- ed, which was attributed to the charge build up on the surface of
tion, respectively, while the effect of the volume flow rate on the alumina particles due to the controlled pH solution. The highly
pumping power, and the pressure drop is shown in Fig. 18 for the lam- charged formation around alumina particles and a valid dispersion
inar flow. These two parameters were calculated using Eqs. (11)–(14) of suspension was verified at pH 9.0.
and (Tables 3 and 4). b. Nanofluids containing small quantities of nanoparticles (below
Apparently, the friction factors of the nanofluids are nearly the same 0.3%) have substantially higher thermal conductivity than those of
to those of water under the comparable nanoparticle volume fraction. base fluids. The use of Al2O3 and TiO2 nanoparticles can significantly
Therefore, the effect of pumping power using nanofluids with a low improve the thermal conductivity of the solution, and the improve-
volume fraction is comparable to that of water [53]. No considerable ment rises with the growing particle concentration in this subject
accumulation to the pressure drop for the nanofluids is found on all area.
sequences of the investigation, which discloses that nanofluids will c. It has been found that the nanofluid viscosity strongly depends upon
not require an added disadvantage over the pumping power. It is the volume concentration and temperature as well as the base-fluids
witnessed that the friction factor relationship of the single-phase flow used. The hysteresis behavior was also observed for all the nano-
can be used for nanofluids. fluids.
d. Almost negligible effect in the pumping power and pressure drop is
4. Conclusions noticed for low concentration nanofluids. The zeta potential is an im-
portant basis for selecting the conditions of dispersing particles.
In this paper, we have experimentally investigated the preparation, There is a good correlation between the stability and the zeta poten-
thermal conductivity, viscosity and hysteresis phenomenon of water tial. The higher the value of the zeta potential is, the greater the
based Al2O3 & TiO2 nanofluids of low concentrations (0.05 to 0.3%v/v) stability of the solution is.

Fig. 15. Hysteresis observed for Al2O3/water for 13 nm, particle volume fractions of 0.1% and 0.3%, for fresh samples and samples after being used for a month in a flat plate solar collector.
94 Z. Said et al. / International Communications in Heat and Mass Transfer 58 (2014) 85–95

Table 3
Physical characteristic of metal oxides nanofluids and Water [9,42,43,54].

Particle & Average Actual Cp (J/kg K) K (W/m K) Viscosity


base fluid particle density (m Pa s)
size (nm) (kg/m3)

Al2O3 (gamma) 13 3960 773 40


TiO2 21 4230 692 8.4
Water 997.1 4179 0.605 0.89

Table 4
Environmental and analytical conditions for the flat plate solar collector.

Fig. 16. Hysteresis observed for TiO2/water for ~21 nm, particle volume fractions of 0.1% Parameters of collector Value
and 0.3%.
Type Black paint flat plate
Glazing Single glass
Agent fluids Water, Al2O3 and TiO2 nanofluids
Absorption area, Ap 1.51 m2
Wind speed 2 m/s
Collector tilt, βo 20°
Apparent sun temperature, Ts 4350 K
Optical efficiency, o [44] 0.84
Glass thickness, t 4 mm
Insulation thermal conductivity, ki 0.06 W/m K
Inner diameter of pipes, d 0.01 m

Acknowledgement

This research is supported by UM High Impact Research Grant UM-


MOHE UM.C/HIR/MOHE/ENG/40 from the Ministry of Higher Education
Fig. 17. Effect of the volume fractions on the pressure drop and pumping power for Al2O3–
Malaysia.
H2O & TiO2–H2O.

e. To improve the thermal efficiency of a variety of systems, the en- References


hancement of the thermal conductivity of the base fluid will be a [1] J. Eastman, S. Choi, S. Li, W. Yu, L. Thompson, Anomalously increased effective
substantial requirement. Experiments dealing with the measure- thermal conductivities of ethylene glycol-based nanofluids containing copper nano-
ment of thermal conductivity at low range temperatures have not particles, Appl. Phys. Lett. 78 (6) (2001) 718–720.
[2] B.C. Pak, Y.I. Cho, Hydrodynamic and heat transfer study of dispersed fluids with
been stated yet. The behavior of thermal conductivity of nanofluids submicron metallic oxide particles, Exp. Heat Transf. Int. J. 11 (2) (1998) 151–170.
at lower temperatures can direct to a new course in the field of [3] Z. Said, M. Sajid, R. Saidur, M. Kamalisarvestani, N. Rahim, Radiative properties of
research. nanofluids, Int. Commun. Heat Mass Transf. 46 (2013) 74–84.
[4] E. Natarajan, R. Sathish, Role of nanofluids in solar water heater, Int. J. Adv. Manuf.
f. Hysteresis phenomena in viscosity of nanofluids have elevated seri- Technol. (2009) 1–5.
ous alarms regarding their use for enhancing the heat-transfer agil- [5] Z. Said, M.H. Sajid, M.A. Alim, R. Saidur, N.A. Rahim, Experimental investigation of
ity. Researchers can emphasize on the influence of temperature and the thermophysical properties of Al2O3–nanofluid and its effect on a flat plate
solar collector, Int. Commun. Heat Mass Transf. 48 (2013) 99–107.
the hysteresis behavior for various nanofluids and can try to upsurge [6] T.P. Otanicar, J.S. Golden, Comparative environmental and economic analysis of con-
the temperature tolerating capacity of various nanofluids, which are ventional and nanofluid solar hot water technologies, Environ. Sci. Technol. 43 (15)
yet to be proven. From the attained results, clearly nanofluids can (2009) 6082–6087.
[7] T.P. Otanicar, P.E. Phelan, R.S. Prasher, G. Rosengarten, R.A. Taylor, Nanofluid-based
play an active role in enhancing the heat transfer and are suitable
direct absorption solar collector, J. Renew. Sust. Energ. 2 (2010) 033102.
for the applications in applied heat transfer expansions. [8] T. Yousefi, E. Shojaeizadeh, F. Veysi, S. Zinadini, An experimental investigation on
the effect of pH variation of MWCNT–H2O nanofluid on the efficiency of a flat-
plate solar collector, Sol. Energy 86 (2) (2011) 771–779.
[9] T. Yousefi, F. Veysi, E. Shojaeizadeh, S. Zinadini, An experimental investigation on
the effect of Al2O3–H2O nanofluid on the efficiency of flat-plate solar collectors,
Renew. Energy 39 (1) (2012) 293–298.
[10] R.A. Taylor, P.E. Phelan, T.P. Otanicar, R. Adrian, R. Prasher, Nanofluid optical proper-
ty characterization: towards efficient direct absorption solar collectors, Nanoscale
Res. Lett. 6 (1) (2011) 1–11.
[11] A.K. Tiwari, P. Ghosh, J. Sarkar, Solar water heating using nanofluids—a comprehen-
sive overview and environmental impact analysis, Int. J. Emerg. Technol. Adv. Eng. 3
(3) (2013) 221–224.
[12] R. Saidur, T. Meng, Z. Said, M. Hasanuzzaman, A. Kamyar, Evaluation of the effect of
nanofluid-based absorbers on direct solar collector, Int. J. Heat Mass Transf. 55
(21–22) (2012) 5899–5907.
[13] V. Sridhara, L.N. Satapathy, Al2O3-based nanofluids: a review, Nanoscale Res. Lett. 6
(1) (2011) 1–16.
[14] S. Bobbo, L. Fedele, A. Benetti, L. Colla, M. Fabrizio, C. Pagura, S. Barison, Viscosity of
water based SWCNH and TiO2 nanofluids, Exp. Thermal Fluid Sci. 36 (2011) 65–71.
[15] M. Reddy, V.V. Rao, B. Reddy, S.N. Sarada, L. Ramesh, Thermal conductivity measure-
ments of ethylene glycol water based TiO2 nanofluids, Nanosci. Nanotechnol. Lett. 4
(1) (2012) 105–109.
[16] C.T. Wamkam, M.K. Opoku, H. Hong, P. Smith, Effects of pH on heat transfer
Fig. 18. Effect of volume flow rate on the pumping power and pressure drop for Al2O3–H2O nanofluids containing ZrO2 and TiO2 nanoparticles, J. Appl. Phys. 109 (2011)
& TiO2–H2O. 024305.
Z. Said et al. / International Communications in Heat and Mass Transfer 58 (2014) 85–95 95

[17] H. Xie, W. Yu, W. Chen, MgO nanofluids: higher thermal conductivity and lower [35] H. Brinkman, The viscosity of concentrated suspensions and solutions, J. Chem. Phys.
viscosity among ethylene glycol-based nanofluids containing oxide nanoparticles, 5 (20) (1952) 71.
J. Exp. Nanosci. 5 (5) (2010) 463–472. [36] G. Batchelor, The effect of Brownian motion on the bulk stress in a suspension of
[18] A. Turgut, I. Tavman, M. Chirtoc, H. Schuchmann, C. Sauter, S. Tavman, Thermal spherical particles, J. Fluid Mech. 83 (01) (1977) 97–117.
conductivity and viscosity measurements of water-based TiO2 nanofluids, Int. J. [37] E. Abu-Nada, Z. Masoud, H.F. Oztop, A. Campo, Effect of nanofluid variable properties
Thermophys. 30 (4) (2009) 1213–1226. on natural convection in enclosures, Int. J. Therm. Sci. 49 (3) (2010) 479–491.
[19] X.Q. Wang, A.S. Mujumdar, Heat transfer characteristics of nanofluids: a review, Int. [38] S. Masoud Hosseini, A. Moghadassi, D.E. Henneke, A new dimensionless group
J. Therm. Sci. 46 (1) (2007) 1–19. model for determining the viscosity of nanofluids, J. Therm. Anal. Calorim. 100 (3)
[20] J. Eastman, U. Choi, S. Li, L. Thompson, S. Lee, Enhanced thermal conductivity (2010) 873–877.
through the development of nanofluids, Materials Research Society Symposium [39] S.E.B. Maiga, C.T. Nguyen, N. Galanis, G. Roy, Heat transfer behaviours of nanofluids
Proceedings, Cambridge Univ Press, 1997. in a uniformly heated tube, Superlattice. Microst. 35 (3) (2004) 543–557.
[21] X. Wang, X.S. Xu, S.U. Choi, Thermal conductivity of nanoparticle–fluid mixture, [40] H. Garg, R. Agarwal, Some aspects of a PV/T collector/forced circulation flat plate
J. Thermophys. Heat Transf. 13 (4) (1999) 474–480. solar water heater with solar cells, Energy Convers. Manag. 36 (2) (1995) 87–99.
[22] H. Masuda, A. Ebata, K. Teramae, N. Hishinuma, Alteration of thermal conductivity [41] F.M. White, Fluid Mechanics. 5th, McGraw-Hill Book Company, Boston, 2003.
and viscosity of liquid by dispersing ultra-fine particles, Netsu Bussei 7 (2) (1993) [42] S.K. Das, S.U. Choi, H.E. Patel, Heat transfer in nanofluids—a review, Heat Transf. Eng.
227–233. 27 (10) (2006) 3–19.
[23] S. Lee, S.U. Choi, S. Li, J. Eastman, Measuring thermal conductivity of fluids contain- [43] S.P. Jang, S.U. Choi, Role of Brownian motion in the enhanced thermal conductivity
ing oxide nanoparticles, J. Heat Transf. 121 (2) (1999). of nanofluids, Appl. Phys. Lett. 84 (21) (2004) 4316–4318.
[24] B. Aladag, S. Halelfadl, N. Doner, T. Maré, S. Duret, P. Estellé, Experimental investiga- [44] X.Q. Wang, A.S. Mujumdar, A review on nanofluids-part I: theoretical and numerical
tions of the viscosity of nanofluids at low temperatures, Appl. Energy 97 (2012) investigations, Braz. J. Chem. Eng. 25 (4) (2008) 613–630.
876–880. [45] L. Fedele, L. Colla, S. Bobbo, Viscosity and thermal conductivity measurements of
[25] C. Nguyen, F. Desgranges, G. Roy, N. Galanis, T. Mare, S. Boucher, H. Angue Mintsa, water-based nanofluids containing titanium oxide nanoparticles, Int. J. Refrig. 35
Temperature and particle-size dependent viscosity data for water-based nanofluids— (5) (2012) 1359–1366.
hysteresis phenomenon, Int. J. Heat Fluid Flow 28 (6) (2007) 1492–1506. [46] H. Xie, J. Wang, T. Xi, Y. Liu, F. Ai, Dependence of the thermal conductivity of
[26] C.T. Nguyen, N. Galanis, T. Maré, E. Eveillard, New viscosity data for CuO–water nanoparticle–fluid mixture on the base fluid, J. Mater. Sci. Lett. 21 (19) (2002)
nanofluid—the hysteresis phenomenon revisited, Adv. Sci. Technol. 81 (2013) 1469–1471.
101–106. [47] S. Kabelac, J. Kuhnke, Heat transfer mechanisms in nanofluids-Experiments and
[27] G. Huminic, A. Huminic, Application of nanofluids in heat exchangers: a review, theory, Annals of the assembly for international heat transfer conference, 2006.
Renew. Sust. Energ. Rev. 16 (8) (2012) 5625–5638. [48] Y. Ding, H. Alias, D. Wen, R.A. Williams, Heat transfer of aqueous suspensions of
[28] X. Fan, H. Chen, Y. Ding, P.K. Plucinski, A.A. Lapkin, Potential of ‘nanofluids’ to further carbon nanotubes (CNT nanofluids), Int. J. Heat Mass Transf. 49 (1) (2006) 240–250.
intensify microreactors, Green Chem. 10 (6) (2008) 670–677. [49] K. Kwak, C. Kim, Viscosity and thermal conductivity of copper oxide nanofluid
[29] H. Demir, A. Dalkilic, N. Kürekci, W. Duangthongsuk, S. Wongwises, Numerical in- dispersed in ethylene glycol, Korea Aust. Rheol. J. 17 (2) (2005) 35–40.
vestigation on the single phase forced convection heat transfer characteristics of [50] R. Prasher, D. Song, J. Wang, P. Phelan, Measurements of nanofluid viscosity and its
TiO2 nanofluids in a double-tube counter flow heat exchanger, Int. Commun. Heat implications for thermal applications, Appl. Phys. Lett. 89 (13) (2006) 133108.
Mass Transf. 38 (2) (2011) 218–228. [51] H. Chen, Y. Ding, C. Tan, Rheological behaviour of nanofluids, New J. Phys. 9 (10)
[30] S. Fotukian, M. Nasr Esfahany, Experimental investigation of turbulent convective (2007) 367.
heat transfer of dilute γ-Al2O3/water nanofluid inside a circular tube, Int. J. Heat [52] H. Chen, Y. Ding, Heat transfer and rheological behaviour of nanofluids—a review,
Fluid Flow 31 (4) (2010) 606–612. Advances in Transport Phenomena, Springer, 2009, pp. 135–177.
[31] V. Penkavova, J. Tihon, O. Wein, Stability and rheology of dilute TiO2–water [53] I. Gherasim, G. Roy, C.T. Nguyen, D. Vo-Ngoc, Heat transfer enhancement and
nanofluids, Nanoscale Res. Lett. 6 (1) (2011) 1–7. pumping power in confined radial flows using nanoparticle suspensions
[32] Y. He, Y. Jin, H. Chen, Y. Ding, D. Cang, H. Lu, Heat transfer and flow behaviour of (nanofluids), Int. J. Therm. Sci. 50 (3) (2011) 369–377.
aqueous suspensions of TiO2 nanoparticles (nanofluids) flowing upward through [54] Z. Said, A. Kamyar, R. Saidur, Experimental investigation on the stability and density of
a vertical pipe, Int. J. Heat Mass Transf. 50 (11) (2007) 2272–2281. TiO2, Al2O3, SiO2 and TiSiO4, IOP Conference Series: Earth and Environmental Science,
[33] W.J. Tseng, K.C. Lin, Rheology and colloidal structure of aqueous TiO2 nanoparticle Vol. 16. No. 1, IOP Publishing, 2013.
suspensions, Mater. Sci. Eng. A 355 (1) (2003) 186–192.
[34] A. Einstein, Calculation of the viscosity-coefficient of a liquid in which a large num-
ber of small spheres are suspended in irregular distribution, Ann. Phys. Leipzig 19
(1906) 286–306.

You might also like