You are on page 1of 10

international journal of refrigeration 70 (2016) 93102

Available online at www.sciencedirect.com

ScienceDirect
j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / i j r e f r i g

Investigation of thermal conductivity and


viscosity of Al2O3/PAG nanolubricant for
application in automotive air conditioning
system
M.Z. Sharif a, W.H. Azmi a,b,*, A.A.M. Redhwan a,c, R. Mamat a,b
a

Faculty of Mechanical Engineering, Universiti Malaysia Pahang, 26600 Pekan, Pahang, Malaysia
Automotive Engineering Centre, Universiti Malaysia Pahang, 26600 Pekan, Pahang, Malaysia
c
Faculty of Manufacturing Engineering Technology, TATI University College, 24000 Kemaman, Terengganu, Malaysia
b

A R T I C L E

I N F O

A B S T R A C T

Article history:

In this paper, thermal conductivity and viscosity of the Al2O3/polyalkylene glycol (PAG) 46

Received 25 March 2016

nanolubricants for 0.05 to 1.0% volume concentrations at temperatures of 303.15 to 353.15 K

Received in revised form 11 June

have been investigated. Al2O3 nanoparticles were dispersed in the PAG lubricant by a two

2016

step preparation. The measurement of thermal conductivity and viscosity was performed

Accepted 20 June 2016

using KD2 Pro Thermal Properties Analyzer and LVDV-III Rheometer, respectively. The results

Available online 21 June 2016

showed that the thermal conductivity of the nanolubricants increased by concentration,


but decreased by temperature. Besides, the viscosity of the nanolubricants sharply in-

Keywords:

creased for concentrations higher than 0.3%. However, this parameter diminished by

Nanolubricants

temperature. The highest thermal conductivity and viscosity ratio were observed to be 1.04

Thermal conductivity

and 7.58 times greater than the PAG lubricant for 1.0% and 0.4% concentrations, respec-

Viscosity

tively. As a conclusion, it was recommended to use the Al2O3/PAG nanolubricants with

Air conditioning system

concentration of less than 0.3% for application in automotive air conditioning system.
2016 Elsevier Ltd and IIR. All rights reserved.

tude de la conductivit thermique et de la viscosit de


nanolubrifiant Al2O3/PAG appliqu au systme de
conditionnement dair dautomobile
Mots cls : Nanolubrifiants ; Conductivit thermique ; Viscosit ; Systme de conditionnement dair

* Corresponding author. Faculty of Mechanical Engineering, Universiti Malaysia Pahang, 26600 Pekan, Pahang, Malaysia. Tel.: +6 09 4246338;
Fax: +6 09 4242202.
E-mail address: wanazmi2010@gmail.com (W.H. Azmi).
http://dx.doi.org/10.1016/j.ijrefrig.2016.06.025
0140-7007/ 2016 Elsevier Ltd and IIR. All rights reserved.

94

international journal of refrigeration 70 (2016) 93102

Nomenclature

English symbols
AD
average deviation
COP coefficient of performance
cSt
centistokes
FESEMfield emission scanning electron microscopy
h
nanolayer thickness
k
thermal conductivity [W(mK)1]
kr
thermal conductivity ratio [kNL/kL]
m
mass
n
number of layer
PAG polyalkylene glycol
POE polyolester
r
original radius of nanoparticle
SD
standard deviation
T
temperature [C]
TEM transmission electron microscopy
Greek symbols

ratio nanolayer thickness to the original


radius [h/r]

volume concentration [%]

volume concentration in fraction

dynamic viscosity [mPas]


viscosity ratio [NL/L]
r

density [kgm3]
Subscripts
bf
based fluid
eff
effective
eq
equation
exp
experiment
L
lubricant
NL
nanolubricant
p
nanoparticle
r
ratio

1.

Introduction

Masuda et al. (1993) measured the thermal conductivity of TiO2


water and Al2O3water nanofluids and discovered thermal
conductivity was increased by 11% and 32%, respectively. This
research served as a foundation for further thermal conductivity studies. Choi (1995) from the U.S. Argonne National
Laboratory introduced the term nanofluids (the mixture of solid
nanoparticles with a base fluid) as a promising heat transfer
fluid. This kind of heat transfer fluid has higher thermal conductivity in contrast to based conventional heat transfer fluids.
Nanofluids also possess superb heat transfer properties such
as high thermal conductivity, good stability, and homogeneity along with the minimal clogging in flow passages due to
very small size and tremendous specific surface area of the
nanoparticles (Chandrasekar et al., 2010).
Refrigerants are extensively used in commercial, industrial and automotive refrigeration and air conditioning systems.

The idea of nanorefrigerants has been proposed based on the


idea of nanofluids, which were prepared by mixing the
nanoparticles into the conventional refrigerant. So far, at least
three main advantages of using nanoparticles in refrigerant
are obtained (Bi et al., 2011): (1) Nanoparticle as additives can
increase the solubility between the refrigerant and the lubricant; (2) Thermal conductivity and heat transfer characteristics
of the refrigerant can be increased; and (3) Nanoparticle dispersion into lubricant might reduce the friction coefficient and
wear rate. Nanorefrigerants have the prospective to boost heat
transfer rate thus more compact of heat exchanger in air conditioning and refrigeration systems are achievable. Studies on
nanorefrigerants (Mahbubul et al., 2013a, 2013b; Park and Jung,
2007; Peng et al., 2009; Sun and Yang, 2013) have shown that
adding nanoparticles to refrigerants can improve the heat transfer of the base refrigerant.
The effect of carbon nanotubes (CNT)/R134a on nucleate
boiling heat transfer has been studied by Park and Jung (2007).
They found that large enhancement up to 36.6% was observed. Peng et al. (2009) explored the heat transfer features
of refrigerant-based nanofluid flow boiling inside a horizontal smooth tube and found the maximum heat transfer
coefficient increased to 29.7%. Then, they projected a heat transfer correlation and found that the difference between the
predicted and experimental data is 20%. Sun and Yang (2013)
did the research on nanorefrigerant material type and vapour
quality in horizontal pipe for pure copper (Cu), copper oxide
(CuO), pure aluminium (Al) and aluminium oxide (Al2O3). According to their research, the heat transfer coefficient of CuR141b was about 1.26 times higher than pure R141b and the
highest among nanoparticles studied. Mahbubul et al. (2013a,
2013b) investigated heat transfer of Al2O3-R141b and Al2O3R134a nanorefrigerant in horizontal smooth circular tube. They
concluded that with the increment of nanorefrigerant volume
concentration, the heat transfer characteristics increased
significantly.
Apart from nanorefrigerants, the studies on mixture of
nanolubricants and nanorefrigerants (Bartelt et al., 2008; Bobbo
et al., 2010; Henderson et al., 2010; Kedzierski, 2009, 2012;
Kedzierski and Gong, 2009; Peng et al., 2010a, 2010b; Wang et al.,
2003) have also shown better performance compared to their
based fluid. Wang et al. (2003) studied the refrigeration system
using HFC134a and mineral lubricant appended with TiO2
nanoparticle as working fluids. They found that it performs
better by returning more lubricant back to the compressor compared by using HFC134a and POE oil. Bartelt et al. (2008)
examined the heat transfer effect on horizontal flow boiling
condition of R-134a/POE/CuO. They concluded that a maximum
enhancement of 101% was obtained for 2% mass fraction.
Kedzierski and Gong (2009) studied the effect of CuO-R134a pool
boiling heat transfer and found that the nanoparticles improved the heat transfer by 50 to 275% compared with the heat
transfer of pure R134a/polyolester (99.5/0.5). The effect of CuO
nanoparticles dispersed in R134a-lubricant pool boiling heat
transfer has been investigated by Kedzierski (2009). He found
that 2.0% volume concentration of CuO nanoparticle dispersed in R134a/nanolubricant mixtures had smaller boiling
heat flux than the mixtures with 1% volume concentration.
Henderson et al. (2010) studied the effect of nanoparticles
on the heat transfer of R-134a and R-134a/POE mixtures. They

international journal of refrigeration 70 (2016) 93102

concluded that the average heat transfer enhancement was


up to 76%. The influence of nanoparticles dispersion in POE
oils on lubricity and R134a solubility has been studied by Bobbo
et al. (2010). Solubility and tribology of TiO2/SW32 oil mixture
showed the best performance among mixtures that have been
studied. Peng et al. (2010b) studied nucleate pool boiling heat
transfer characteristics of refrigerant/oil dispersed with diamond
nanoparticles. They found that the nucleate pool boiling heat
transfer coefficient increased up to 63.4%. The effect of carbon
nanotubes (CNTs) on nucleate pool boiling heat transfer features of refrigerant oil mixture have also been studied by Peng
et al. (2010a). They found that nucleate pool boiling heat transfer coefficient increased up to 61%. Kedzierski (2012) done the
research on the influence of Al2O3 nanoparticles and R134a/
polyolester mixtures on the pool boiling performance on a
rectangular finned surface. He found that the boiling performance enhanced up to 113% on a rectangular finned surface.
Considering the potential of nanolubricants for improving
the efficiency of air conditioning and refrigeration systems,
thermal conductivity and viscosity measurements of potential nanolubricants gain advantages not only on fundamental
research, but also on the design consideration. But to the best
of the authors knowledge, there are only few literatures (Jiang
et al., 2009a, 2009b; Kedzierski, 2013; Mahbubul et al., 2013c)
that are available for the experimental works on thermal conductivity of nanorefrigerant or nanolubricant. Jiang et al. (2009a)
studied thermal conductivity of CNT/R113 and found that the
thermal conductivity was increased between 50 and 104%. Jiang
et al. (2009b) also investigated the thermal conductivity of R113
with Cu, Al, Ni, CuO and Al2O3 with controlled volume concentrations of 0.1 to 1.2%. They found that the thermal
conductivity of nanorefrigerant increased tremendously with
the increment of volume fraction. They also concluded that
thermal conductivities of nanorefrigerants with various types
of nanoparticles are quite similar to one another if the
nanoparticle volume fractions are similar. In heat transfer applications, apart from thermal conductivity, viscosity also plays
a vital parameter. Pumping power and pressure drop are directly associated to viscosity of any fluid, especially in laminar
flow (Mahbubul et al., 2012).
Thermal conductivity and viscosity of the Al2O3/R141b
nanorefrigerants for 0.5 to 2% volume concentrations at temperatures of 278.15 to 293.15 K were studied by Mahbubul et al.
(2013c). They found that the thermal conductivity was up to
1.63 times greater than the base R141b. Meanwhile, Kedzierski
(2013) investigated the viscosity and density of Al 2 O 3
nanolubricants. He came up with the viscosity model, which
was able to predict the kinematic viscosity of the nanolubricant
within 15% of the measurement was developed. Most of the
researchers are using classical models for thermal conductivity estimation such as Hamilton and Crosser (1962), Maxwell
(1904), Timofeeva et al. (2007) and Yu and Choi (2003). Meanwhile, the model for viscosity are based on the models of
Brinkman (1952), Pak and Cho (1998) and Wang et al. (1999).
Both the thermal conductivity and viscosity models are widely
used for nanofluids properties estimated at low volume concentration. As different base fluids have different thermophysical properties, the model implemented may not suit for
lubricants purposes. Therefore if experimental data of the
thermo-physical properties of the nanolubricants are obtained,

95

it would be used for better understanding on the enhancement of heat transfer, coefficient of performance (COP), energy
saving, lubricity and others.
The application of nanolubricants in the refrigeration and
air condition systems done by Lee et al. (2009), Sabareesh et al.
(2012) and Xing et al. (2014) are very much relevant to this study
and should be used as references accordingly. Sabareesh et al.
(2012) used R12 refrigerant as working fluid in experimenting
the effect of dispersing low concentration of TiO2 nanoparticles
in the mineral oil based lubricant, on its viscosity and lubrication characteristics, as well as on the overall performance
of a vapour compression refrigeration system. Lee et al. (2009)
mixed fullerene nanoparticles of 0.1% volume concentration
in mineral oil and evaluate the friction coefficient by a diskon-disk tribotester. They found that, the friction coefficient of
the nano-oil decreased by 90% in comparison with raw oil. While
Xing et al. (2014) found that the friction coefficients of the
Fullerene C60 nano-oil significantly decreased with increasing the concentration of nanoparticles in the mineral oil.
Hence, the objective of the present work is to study the
thermal conductivity and viscosity of Al2O3 nanoparticles suspended in polyalkylene glycol (PAG) 46 synthetic oil for 0.05
to 1.0% volume concentrations and working temperature of
303.15 to 353.15 K. Further, the optimum volume concentration of nanolubricants needs to be identified thoroughly by
considering the thermal conductivity and viscosity of
nanolubricants for application in automotive air conditioning system. Simultaneously, the regression equation for each
properties were developed using the measured data.

2.

Methodology

The experimental procedures are thoroughly discussed in this


section. This segment is divided into the characterization of
the Al2O3 nanoparticle and the PAG lubricant. Subsequently,
steps for preparation and stability observation of nanolubricants.
Finally, steps to measure the thermal conductivity and viscosity of the nanolubricants are further discussed in detail.

2.1.

Materials and preparation of the nanolubricants

Al2O3 nanoparticles are used and procured with 99.8% purity


and 13 nm in size. The characterization of Al2O3 nanoparticle
is obtained using field emission scanning electron microscopy (FESEM) and transmission electron microscopy (TEM)
imaging technique. Fig. 1(a) shows the image of FESEM in
300,000 magnification. The Al2O3 nanoparticles shape is observed spherically. From the FESEM image, it has been observed
that the nanoparticles are spherical in shape and the sizes
are approximately 13 nm. TEM analysis was further undertaken to determine the condition of nanolubricant
agglomeration, dispersion and also to confirm the particle size
under suspended form. Fig. 1(b) shows the TEM image of Al2O3
nanoparticle suspended in PAG lubricant in 88,000 magnification. From the TEM images, the nanoparticle is dispersed
well in the lubricant. However, small agglomeration and
minimum clustering of nanoparticle is observed in the solution as displayed in Fig. 1(b). The number-based size of the

96

international journal of refrigeration 70 (2016) 93102

Fig. 1 FESEM and TEM images of Al2O3 nanoparticle.

nanoparticles from TEM image is used to determine the average


size of the nanoparticle. As a result, it is confirmed that Al2O3
nanoparticle size is approximately 13 nm and nearly spherical in shape. The properties of Al2O3 are shown in Table 1
(Aldrich, 2013; Zakaria et al., 2015).
The present nanolubricant was intended to be tested in automotive air conditioning system that uses compressor and
being lubricated by polyalkylene gycol (PAG). PAG lubricants have
better tribology performance over mineral oils when used together with HFCs according to Matlock et al. (1999). At high
pressures and temperatures, PAG has low solubility in the
gaseous refrigerant and provide excellent lubricity. PAG have
been used mainly in automotive air-conditioning systems due
to the compatibility characteristic with most of elastomers

Table 1 Properties of nanoparticles used in this


experiment (Aldrich, 2013; Zakaria et al., 2015).
Property
Molecular mass, gmol1
Average particle diameter, nm
Density, kgm3
Thermal conductivity, W(mK)1
Specific heat, J(kgK)1

Al2O3
101.96
13
4000
36
773

(Matlock et al., 1999). Table 2 shows the properties of the PAG


46 lubricant at atmospheric pressure (Brown, 1993; Dow, 2013).
Two step method suggested by Yu and Xie (2012) is used
in preparation of nanolubricants. Eastman et al. (1997), Lee et al.
(1999) and Wang et al. (1999) used the same method in their
preparation of Al2O3 nanofluids. Eq. (1) used to calculate the
volume concentration of the nanolubricants.

mp p
100
m p p + mL L

(1)

where, is the volume concentration in percent, mp and mL


are the masses of the nanoparticle and lubricant, respectively; and p and L are the density of the nanoparticle and
density of the lubricant, respectively. The initial mixing process

Table 2 Properties of PAG 46 lubricant (Brown, 1993;


Dow, 2013).
Property

PAG 46
3

Density, gcm @ 293.15 K


Flash point, K
Kinematic viscosity, cSt @ 313.15 K
Pour point, K

0.9954
447.15
41.450.6
222.15

international journal of refrigeration 70 (2016) 93102

97

Fig. 2 Nanolubricant samples after a month of preparation.

is done by using a magnetic stirrer. The required mass of the


Al2O3 nanoparticles to be dispersed in lubricant was precisely measured utilizing a high accuracy electronic balance.
The mixture then is subjected to ultrasonic homogenization
for an hour to ensure good dispersions of nanolubricants. Dispersion stability is observed visually after a month of
preparation and found that no sedimentations were occurred in the samples as shown in Fig. 2. It should be noted
that no surfactant has been used in this preparation.

Table 3 Thermal conductivity model for nanofluids.


Model

Thermal conductivity

Maxwell (1904)

Hamilton and
Crosser (1962)
Timofeeva et al. (2007)

2.2.

Thermal conductivity measurements

Thermal conductivity is measured using KD2 Pro thermal property analyzer as shown in Fig. 3. Azmi et al. (2016a), Lee et al.
(2011), Mahbubul et al. (2013c) and Zakaria et al. (2015) are some
of the researchers who used KD2 Pro in their thermal conductivity measurement. This apparatus uses the transient line heat
source to determine the thermal properties of liquids and solids.
The apparatus meets the standards of both ASTM D5334 and
IEEE 442 1981. A single needle sensor (KS-1) in the range of
0.002 to 2.00 W(mK)1 is used. A water bath of WNB7L1 model
is utilised to maintain a constant temperature of the sample
with accuracy of 0.1 K (Zakaria et al., 2015). The thermal conductivity of 0.05 to 1.0% volume concentrations of Al2O3/PAG
nanolubricants were measured for temperature range of 303.15
to 353.15 K. The sensor was validated by measuring the thermal
conductivity of the verification liquid (glycerin) given by the
supplier. The measured value of glycerin at 298.15 K is
0.286 W(mK)1, which is in agreement with the calibrated data

Yu and Choi (2003)

kr =

keff kp + 2kbf + 2 (kp kbf )


=

kbf kp + 2kbf (kp kbf )

kr =

keff kp + (n 1) kbf (n 1) (kbf kp )


=

kbf
kp + (n 1) kbf + (kbf kp )

kr =

keff
= (1 + 3 )
kbf

kr =

keff kp + 2kbf + 2 (kp kbf ) (1 + )3


=

kbf kp + 2kbf (kp kbf ) (1 + )3

of 0.285 W(mK)1 and within 0.35% accuracy. The validation


process was done each time before the thermal conductivity
measurement was taken. In order to ensure the consistency
of data measurement, minimum five data were taken for every
concentration at a specific temperature. The thermal conductivity models (Hamilton and Crosser, 1962; Maxwell, 1904;
Timofeeva et al., 2007; Yu and Choi, 2003) are shown in Table 3
and used to verify the results of nanolubricants thermal
conductivity.

2.3.

Viscosity measurements

Fig. 4 shows LVDV-III (low viscosity digital viscometer) Ultra


Programmable Rheometer. The viscometer is able to measure
the viscosity of liquid sample within the range of 1.0 to

Fig. 3 KD2 Pro thermal properties analyzer.

Fig. 4 LVDV III Ultra Programmable Rheometer.

98

international journal of refrigeration 70 (2016) 93102

Table 4 Viscosity model for nanofluids.


Model

Viscosity

Brinkman (1952)

Wang et al. (1999)

Pak and Cho (1998)

r =

eff
1
=
bf (1 )2.5

r =

eff
= 123 2 + 7.3 + 1
bf

r =

eff
= 1 + 39.11 + 533.9 2
bf

6,000,000 mPas with accuracy of 1.0% and temperature accuracy is 0.1 K ranging from 253.15 to 373.15 K by utilizing the
UL Adapter. The viscosity of 0.05 to 0.40% volume concentrations of Al2O3/PAG were measured for a temperature range of
303.15 to 353.15 K. A spindle was used to measure the viscosity of suspensions. The viscometer drives a spindle immersed
in nanolubricants. By means of rotation of the spindle, a viscous
drag of the fluid in opposition to the spindle is created, which
is measured by the deflection of the calibrated spring. Each measurement was conducted three times to get more reliable data.
The mean value of the three data was considered for analysis. The dynamic viscosity models (Brinkman, 1952; Pak and
Cho, 1998; Wang et al., 1999) are listed in Table 4 and are used
to compare the measured values of viscosity for the different
volume concentrations. For both thermo-physical measurements, the confidence level is 99.9% with 0.1% uncertainties.

3.

Results and discussion

3.1.

Thermal conductivity of the nanolubricants

Fig. 5 shows the thermal conductivity of the Al 2 O 3 /PAG


nanolubricants at 303.15 K for 0.05 to 1.0% volume concentrations. The experimental results of the present study were

Fig. 5 Variation of thermal conductivity ratio as the


function of volume concentration at 303.15 K.

compared with the estimated values obtained from previously published models in literature. The figure shows that the
thermal conductivity of the Al 2 O 3 /PAG nanolubricant increases with volume concentration. The experimental values
for this study were found to be slightly higher than the three
models of Hamilton and Crosser (1962), Maxwell (1904), and
Timofeeva et al. (2007). However, the model by Yu and Choi
(2003) seem to agree with the experimental value to some
extent. The mean and maximum deviation of the experimental values compared to Yu and Choi (2003) is 0.05% and 0.12%,
respectively. Mahbubul et al. (2013c) have compared the experiment value of thermal conductivity of Al 2 O 3 -R141b
nanorefrigerant with Maxwell (1904). The results showed that
their experimental value also is much higher compared to
Maxwell (1904) by 34% deviation. On the other hand, Peng et al.
(2009) used the Hamilton and Crosser model (1962) to calculate the thermal conductivity of CuO/R113 nanorefrigerants.
By comparing results with others researchers, the thermal conductivity result in this study is found to be in a good agreement
with most models from literature.
Fig. 6 shows the thermal conductivity of the Al2O3/PAG
nanolubricant as a function of temperature. It clearly shows
that thermal conductivity increases as volume concentrations increase. The highest thermal conductivity achievable is
0.153 W(mK)1 at 1.0% volume concentration and temperature of 303.15 K. In addition, the enhancement ratio of
nanolubricant is 1.04 when compared to pure PAG under the
same volume concentration and temperature. The measured
thermal conductivity for all volume concentrations decreased
with the increasing in temperature. The pattern is agreed well
with the pure PAG behaviour as plotted using data from Booser
(1994), which is presented by the solid straight line in Fig. 6.
This behaviour can be explained when liquid is heated, the molecules of the liquid move apart, hence increasing the mean
path. Consequently the probability of collision of molecules will
be reduced. As a result, thermal conductivity decreases as temperature increases. It can be concluded that the thermal
conductivity of the nanolubricant increases by volume concentration, but in contrast, decreases by temperature. In
addition, the thermal conductivity enhancement of Al2O3/PAG

Fig. 6 Thermal conductivity as a function of temperature


for different volume concentrations.

international journal of refrigeration 70 (2016) 93102

Fig. 7 Comparison of nanolubricant thermal conductivity


value between present data and proposed Eq. (2).

nanolubricant is observed to be insignificant with less than 5%


for the volume concentration up to 1.0%.
Consequently, Eq. (2) is developed to estimate the thermal
conductivity of nanolubricants for different volume concentrations and a wide range of temperature. The correlation has
an average deviation of 0.34% and standard deviation of 0.26%.
The equation is in good agreement within 1.5% deviation compared to the experimental data as shown in Fig. 7 and applicable
for 0 1.0% and 303.15 T 353.15 K .

kr =

3.2.

4
kNL
T 273.15 0.05
= 0.15 1 +
1+

kL
100
80

(2)

Viscosity of the nanolubricants

Viscosity of the Al2O3/PAG nanolubricants for 0.05 to 0.4%


volume concentrations and temperature of 313.15 K have been
plotted in Fig. 8. The figure shows that the viscosity of the

Fig. 8 Variation of dynamic viscosity ratio as the function


of volume concentration at 313.15 K.

99

Fig. 9 Viscosity against shear strain rate for volume


concentration at 303.15 K.

nanolubricant increases exponentially with the increase of


volume concentration. The model of Brinkman (1952), Pak and
Cho (1998) and Wang et al. (1999) were utilized to compare with
the present values for different volume concentrations and temperature of 313.15 K. It can be seen that all models were largely
under prediction of the viscosity of nanolubricants. The classical models show that the effective viscosity depends on the
volume concentration and viscosity of the base fluid. While the
experimental results shows that the temperature, the particle diameter and the type of material also contributes to the
enhancement of the effective viscosity of a nanofluids (Azmi
et al., 2016b). Kole and Dey (2010) whom studied viscosity of
Al2O3 in car engine coolant compared their viscosity result with
Batchelor (1977), Brinkman (1952), Chen et al. (2007), Einstein
(1956) and Krieger and Dougherty (1959). They found that all
of the models failed to predict their measured viscosity of the
nanofluids. Similar trend results also had been found by other
researches (Mahbubul et al., 2013c; Namburu et al., 2007; Peng
et al., 2009).
For further investigation towards the behaviour of the
nanolubricant, the viscosity of the pure PAG lubricant and
nanolubricant were been measured as a function of shear strain
rate for 303.15 K as depicted in Fig. 9. It clearly shows that, the
viscosity of the pure PAG lubricant is independent of the shear
strain rate, indicative of Newtonian behaviour. Adding the Al2O3
nanoparticles up to certain extend did not affect the behaviour
of the lubricant as shown by 0.1 and 0.2% volume concentration. However, the nanolubricant tends to show non-Newtonian
behaviour at high volume concentrations. At 0.3% volume concentration, the graph shows downward trend reflecting the
shear thinning behaviour. Therefore, the nanolubricant behaves
non-newtonian fluids for volume concentration above 0.3%. The
same phenomenon on nanolubricant behaviour was also been
observed by Kole and Dey (2011).
Fig. 10 shows the temperature dependence of Al 2 O 3
nanolubricants viscosity for different volume concentrations. It clearly showed that the viscosity of nanolubricants
decrease exponentially at elevated temperature. The highest

100

international journal of refrigeration 70 (2016) 93102

Fig. 10 Comparison of experimental value of viscosity at


various temperatures.

viscosity ratio in this study were found to be 7.58 for


nanolubricants with 0.4% volume concentration compared to
the pure PAG at 313.15 K. The differences became closer align
with the increment of temperature. The pattern again is agreed
well when compared to pure PAG behaviour as plotted in Fig. 10
using the data from Booser (1994). High nanolubricants temperature is believed to intensify the Brownian motion as
suggested by Namburu et al. (2007). This is the reason why the
viscosity of nanolubricants is reduced as the temperature increases. Based on the result in Fig. 10, it is recommended to
use the Al2O3/PAG nanolubricant with volume concentration
of less than 0.3% for application in automotive air conditioning system. The effective viscosity of nanolubricants is
exponentially increased for volume concentration higher
than 0.3%. This can be due to the significant agglomeration
that occurs in nanolubricants at high volume concentrations. The physical agglomeration of the suspended Al2O3
nanoparticle in lubricants was confirmed by TEM image in
Fig. 1(b). Furthermore, the Al2O3/PAG nanolubricants exhibit
non-Newtonion behaviour for volume concentration above
0.3% and confirmed by the result in Fig. 9. Higher viscosity of
nanolubricants with more than 100 mPa.s will penalty with
extra load, work under performance and lesser life cycle to the
compressor. Simultaneously, the pumping power and pressure drop of the overall system are directly associated to the
viscosity of nanolubricants (Mahbubul et al., 2012).
There is no theoretical correlation available in literature for
Al2O3 nanoparticle dispersed in PAG lubricant for designated
volume concentration and temperature. However, there is
similar work done by Kedzierski (2013) to investigate the viscosity of Al2O3 nanoparticle dispersed in POE lubricant for high
concentration. Kedzierski (2013) has generalized the model for
viscosity estimation of various types of lubricant dispersed by
Al2O3 nanoparticle. Fig. 11 depicted the parity graph of Kedzierski
model against the present viscosity data. The model is able to
predict the viscosity of present measurements with the deviation between the experimental value and the model within
15% as suggested by the author. Even though the equation
could predict the viscosity of the Al2O3/PAG nanolubricant, but

Fig. 11 Comparison of nanolubricant viscosity value


between present data and Kedzierski (2013) model.

Fig. 12 Comparison of nanolubricant viscosity value


between present data and proposed Eq. (3).

the deviation is consider high. Hence, Eq. (3) has been proposed to estimate the viscosity of nanolubricants for different
volume concentrations and a wide range of temperature. The
correlation has an average deviation of 3.88% and standard deviation of 3.33%. The equation is in good agreement within 10%
compared to the present data as shown in Fig. 12 and it is applicable for 0 0.3% and 303.15 T 353.15 K .

r =

4.

T 273.15 0.3
NL
368
= 1 +

0 .1 +

L
100
80

(3)

Conclusions

Thermal conductivity and viscosity of Al2O3/PAG nanolubricants


have been studied. Like other nanofluids, the experimental investigation found that the thermal conductivity of the
nanolubricants increases with volume concentration. But contrary to other nanofluids, the thermal conductivity of

international journal of refrigeration 70 (2016) 93102

nanolubricants decreases as temperature increases. Similar to


thermal conductivity, volume concentration and temperature have significant effects over the viscosity of nanolubricants.
The results showed that by increasing the volume concentration, the viscosity of nanolubricants increases. However, it will
decrease by increment of temperature.
From the observation, for the same volume concentration, the viscosity increment rate was found to be higher
compared to the increment rate of thermal conductivity. The
highest thermal conductivity ratio was found to be 1.04. On
the other hand, the highest viscosity was observed to be 7.58
compared to the based lubricant. Hence it is important to find
the ideal volume concentration of Al2O3/PAG nanolubricants
in order to get optimum performance of the automotive vapour
compression system. Therefore, it is recommended to use the
Al2O3/PAG nanolubricants with volume concentration of less
than 0.3% for application in automotive air conditioning
systems. Further investigations on the performance of automotive air conditioning system using Al2O3/PAG nanolubricants
are required to extend the present work.

Acknowledgements
The authors are grateful to the Universiti Malaysia Pahang (UMP)
and Automotive Engineering Centre (AEC) for financial support
given under RDU151411 (RAGS/1/2015/TK0/UMP/03/2).

REFERENCES

Aldrich, S., 2013. Safety data sheet. Aluminium oxide.


Azmi, W.H., Abdul Hamid, K., Usri, N.A., Mamat, R., Mohamad,
M.S., 2016a. Heat transfer and friction factor of water and
ethylene glycol mixture based TiO2 and Al2O3 nanofluids
under turbulent flow. Int. Commun. Heat Mass Transf. 76, 24
32.
Azmi, W.H., Sharma, K.V., Mamat, R., Najafi, G., Mohamad, M.S.,
2016b. The enhancement of effective thermal conductivity
and effective dynamic viscosity of nanofluids a review.
Renew. Sustain. Energy Rev. 53, 10461058.
Bartelt, K., Park, Y., Liu, L., Jacobi, A., 2008. Flow-boiling of R-134a/
POE/CuO nanofluids in a horizontal tube. In: International
Refrigeration and Air Conditioning Conference. Purdue
University, p. 928.
Batchelor, G.K., 1977. The effect of Brownian motion on the bulk
stress in a suspension of spherical particles. J. Fluid Mech. 83,
97117.
Bi, S., Guo, K., Liu, Z., Wu, J., 2011. Performance of a domestic
refrigerator using TiO2-R600a nano-refrigerant as working
fluid. Energy Convers. Manag. 52, 733737.
Bobbo, S., Fedele, L., Fabrizio, M., Barison, S., Battiston, S., Pagura,
C., 2010. Influence of nanoparticles dispersion in POE oils on
lubricity and R134a solubility. Int. J. Refrigeration 33, 1180
1186.
Booser, E.R., 1994. CRC Handbook of Lubrication and Tribology,
Volume III: Monitoring, Materials, Synthetic Lubricants and
Applications. CRC Press LLC, USA.
Brinkman, H.C., 1952. The viscosity of concentrated suspensions
and solutions. J. Chem. Phys. 20, 571.
Brown, W.L., 1993. Polyalkylene glycols. In: CRC Handbook of
Lubrication and Tribology, vol. 3. pp. 253267.

101

Chandrasekar, M., Suresh, S., Bose, A.C., 2010. Experimental


investigations and theoretical determination of thermal
conductivity and viscosity of Al2O3/water nanofluid. Exp.
Therm. Fluid Sci. 34, 210216.
Chen, H., Ding, Y., He, Y., Tan, C., 2007. Rheological behaviour of
ethylene glycol based titania nanofluids. Chem. Phys. Lett.
444, 333337.
Choi, U.S., 1995. Enhancing thermal conductivity of fluids with
nanoparticles. In: Siginer, D.A., Wang, H.P. (Eds.),
Developments and Applications of Non-Newtonian Flows.
American Society of Mechanical Engineers (ASME), New York,
pp. 99105.
Dow, 2013. Material Safety Data Sheet. Ucon Refrigerant
Lubricant 213.
Eastman, J.A., Choi, U.S., Li, S., Thompson, L.J., Lee, S., 1997.
Enhanced thermal conductivity through the development of
nanofluids. In: Proceedings of Symposium Nanophase and
Nanocomposite Materials II. Materials Research Society,
Boston, MA, pp. 311.
Einstein, A., 1956. Investigations on the Theory of the Brownian
Movement. Courier Corporation.
Hamilton, R.L., Crosser, O.K., 1962. Thermal conductivity of
heterogeneous two-component systems. Ind. Eng. Chem.
Fund. 1, 187191.
Henderson, K., Park, Y.G., Liu, L., Jacobi, A.M., 2010. Flow-boiling
heat transfer of R-134a-based nanofluids in a horizontal tube.
Int. J. Heat Mass Transf. 53, 944951.
Jiang, W., Ding, G., Peng, H., 2009a. Measurement and model on
thermal conductivities of carbon nanotube nanorefrigerants.
Int. J. Therm. Sci. 48, 11081115.
Jiang, W., Ding, G., Peng, H., Gao, Y., Wang, K., 2009b. Experimental
and model research on nanorefrigerant thermal conductivity.
HVAC&R Res 15, 651669.
Kedzierski, M.A., 2009. Effect of CuO nanoparticle concentration
on R134a/lubricant pool-boiling heat transfer. J. Heat Transfer
131, 043205.
Kedzierski, M.A., 2012. R134a/Al2O3 nanolubricant mixture pool
boiling on a rectangular finned surface. J. Heat Transfer 134,
121501.
Kedzierski, M.A., 2013. Viscosity and density of aluminum oxide
nanolubricant. Int. J. Refrigeration 36, 13331340.
Kedzierski, M.A., Gong, M., 2009. Effect of CuO nanolubricant on
R134a pool boiling heat transfer. Int. J. Refrigeration 32, 791799.
Kole, M., Dey, T.K., 2010. Viscosity of alumina nanoparticles
dispersed in car engine coolant. Exp. Therm. Fluid Sci. 34, 677
683.
Kole, M., Dey, T.K., 2011. Effect of aggregation on the viscosity of
copper oxidegear oil nanofluids. Int. J. Therm. Sci. 50, 1741
1747.
Krieger, I.M., Dougherty, T.J., 1959. A mechanism for nonNewtonian flow in suspensions of rigid spheres. Trans. Soc.
Rheol 3, 137152.
Lee, K., Hwang, Y., Cheong, S., Kwon, L., Kim, S., Lee, J., 2009.
Performance evaluation of nano-lubricants of fullerene
nanoparticles in refrigeration mineral oil. Curr. Appl. Phys. 9,
e128e131.
Lee, S., Choi, U.S., Li, S., Eastman, J.A., 1999. Measuring thermal
conductivity of fluids containing oxide nanoparticles. J. Heat
Transfer 121, 280289.
Lee, S.W., Park, S.D., Kang, S., Bang, I.C., Kim, J.H., 2011.
Investigation of viscosity and thermal conductivity of SiC
nanofluids for heat transfer applications. Int. J. Heat Mass
Transf. 54, 433438.
Mahbubul, I.M., Saidur, R., Amalina, M.A., 2012. Latest
developments on the viscosity of nanofluids. Int. J. Heat Mass
Transf. 55, 874885.
Mahbubul, I.M., Fadhilah, S.A., Saidur, R., Leong, K.Y., Amalina,
M.A., 2013a. Thermophysical properties and heat transfer

102

international journal of refrigeration 70 (2016) 93102

performance of Al2O3/R-134a nanorefrigerants. Int. J. Heat


Mass Transf. 57, 100108.
Mahbubul, I.M., Saidur, R., Amalina, M.A., 2013b. Heat transfer
and pressure drop characteristics of Al2O3-R141b
nanorefrigerant in horizontal smooth circular tube. Procedia
Eng. 56, 323329.
Mahbubul, I.M., Saidur, R., Amalina, M.A., 2013c. Influence of
particle concentration and temperature on thermal
conductivity and viscosity of Al2O3/R141b nanorefrigerant. Int.
Commun. Heat Mass Transf. 43, 100104.
Masuda, H., Ebata, A., Teramae, K., Hishinuma, N., 1993.
Alteration of thermal conductivity and viscosity of liquid by
dispersing ultra fine particles. Netsu Bussei 4, 227233.
Matlock, P.L., Brown, W.L., Clinton, N.A., 1999. Polyalkylene
Glycols. Chemical Industries, New York, Marcel Dekker, pp.
159194.
Maxwell, J.C., 1904. A Treatise on Electricity and Magnetism, 2nd
ed. Oxford University Press, Cambridge, UK.
Namburu, P.K., Kulkarni, D.P., Misra, D., Das, D.K., 2007. Viscosity
of copper oxide nanoparticles dispersed in ethylene
glycol and water mixture. Exp. Therm. Fluid Sci. 32, 397
402.
Pak, B.C., Cho, Y.I., 1998. Hydrodynamic and heat transfer study
of dispersed fluids with submicron metallic oxide particles.
Exp. Heat Transfer 11, 151170.
Park, K.J., Jung, D., 2007. Boiling heat transfer enhancement with
carbon nanotubes for refrigerants used in building airconditioning. Energy Build. 39, 10611064.
Peng, H., Ding, G., Jiang, W., Hu, H., Gao, Y., 2009. Heat transfer
characteristics of refrigerant-based nanofluid flow boiling
inside a horizontal smooth tube. Int. J. Refrigeration 32, 1259
1270.
Peng, H., Ding, G., Hu, H., Jiang, W., 2010a. Influence of carbon
nanotubes on nucleate pool boiling heat transfer
characteristics of refrigerantoil mixture. Int. J. Therm. Sci. 49,
24282438.

Peng, H., Ding, G., Hu, H., Jiang, W., Zhuang, D., Wang, K., 2010b.
Nucleate pool boiling heat transfer characteristics of
refrigerant/oil mixture with diamond nanoparticles. Int. J.
Refrigeration 33, 347358.
Sabareesh, R.K., Gobinath, N., Sajith, V., Das, S., Sobhan, C., 2012.
Application of TiO2 nanoparticles as a lubricant-additive for
vapor compression refrigeration systemsAn experimental
investigation. Int. J. Refrigeration 35, 19891996.
Sun, B., Yang, D., 2013. Experimental study on the heat transfer
characteristics of nanorefrigerants in an internal thread
copper tube. Int. J. Heat Mass Transf. 64, 559566.
Timofeeva, E.V., Gavrilov, A.N., McCloskey, J.M., Tolmachev, Y.V.,
Sprunt, S., Lopatina, L.M., et al., 2007. Thermal conductivity
and particle agglomeration in alumina nanofluids:
experiment and theory. Phys. Rev. E 76, 061203.
Wang, B.X., Zhou, L.P., Peng, X.F., 2003. A fractal model for
predicting the effective thermal conductivity of liquid with
suspension of nanoparticles. Int. J. Heat Mass Transf. 46, 2665
2672.
Wang, X., Xu, X., Choi, U.S., 1999. Thermal conductivity of
nanoparticle-fluid mixture. J. Thermophys. Heat Transfer 13,
474480.
Xing, M., Wang, R., Yu, J., 2014. Application of fullerene C60 nanooil for performance enhancement of domestic refrigerator
compressors. Int. J. Refrigeration 40, 398403.
Yu, W., Choi, U.S., 2003. The role of interfacial layers in the
enhanced thermal conductivity of nanofluids: a renovated
Maxwell model. J. Nanopart. Res. 5, 167171.
Yu, W., Xie, H., 2012. A review on nanofluids: preparation,
stability mechanisms, and applications. J. Nanomater. 2012,
435873.
Zakaria, I., Azmi, W.H., Mohamed, W.A.N.W., Mamat, R., Najafi, G.,
2015. Experimental investigation of thermal conductivity and
electrical conductivity of Al2O3 nanofluid in water-ethylene
glycol mixture for proton exchange membrane fuel cell
application. Int. Commun. Heat Mass Transf. 61, 6168.

You might also like