Вы находитесь на странице: 1из 7

Applied Thermal Engineering 107 (2016) 10191025

Contents lists available at ScienceDirect

Applied Thermal Engineering


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

Research Paper

Investigation of thermal conductivity of multi walled carbon nanotube


dispersed in hydrogenated oil based drilling fluids
Yee Ho Chai a, Suzana Yusup a,, Vui Soon Chok b, Mohamad Taufiq Arpin b, Sonny Irawan c
a
Biomass Processing Lab, Center for Biofuel and Biochemical, Green Technology, Mission Oriented Research, Chemical Engineering Department, Universiti Teknologi PETRONAS,
Bandar Seri Iskandar, 31750 Tronoh, Perak, Malaysia
b
Scomi Platinum Sdn. Bhd., LOT15-19 & PT1409, Senawang Industrial Estate, Batu 4, Jalan Tampin, 70450 Seremban, Negeri Sembilan Darul Khusus, Malaysia
c
Petroleum Engineering Department, Universiti Teknologi PETRONAS, Bandar Seri Iskandar, 31750 Tronoh, Perak, Malaysia

h i g h l i g h t s
 Hydrodynamic and acoustic cavitation for nanoparticle dispersion is proposed.
 Dispersion of carbon nanotubes at very low particle loadings.
 Thermal conductivity of nanofluid is investigated.
 Thermal conductivity models underpredicted experimental data.

a r t i c l e

i n f o

Article history:
Received 20 April 2016
Revised 30 June 2016
Accepted 2 July 2016
Available online 4 July 2016
Keywords:
Carbon nanotube
Hydrogenated oil
Thermal conductivity
Hydrodynamic cavitation

a b s t r a c t
Thermal conductivity properties of non-polar hydrogenated based-oil dispersed with multi-walled
carbon nanotubes (MWCNT) are investigated. The dispersion of MWCNT are carried out via two-step
methods that utilizes the combination of hydrodynamic cavitation and ultrasonication dispersion at
weight concentration of 25 ppm, 50 ppm and 100 ppm respectively. MWCNT structure and
chemical compound was analyzed using TEM imaging and FTIR analysis. The thermal conductivity of
MWCNT-hydrogenated oil shows 9.8% enhancement at 100 ppm concentration as compared to 7.2%
and 4.5% enhancement at 50 ppm and 25 ppm respectively. Various conventional thermal conductivity
models were compared with experimental data obtained, of which Nan et al. model yields the closest
result approximations at 25 ppm while other models underpredicted the thermal conductivity of
MWCNT-hydrogenated oil nanofluid.
2016 Elsevier Ltd. All rights reserved.

1. Introduction
Thermal conductivity properties of drilling fluids plays an
important role in drilling activities as it helps to shift heat loads
generated between frictions of the walls of the reservoirs and the
drilling bit in an attempt to prevent the overheating of the bit.
Improvement of thermal conductivity properties will not only
reduce the damage of the equipment, but to decrease the downtime of the drilling operations.
Carbon nanotube (CNT) is comprised of single walled nanotube
(SWCNT) and multi walled nanotube (MWCNT), which comprises

Corresponding author.
E-mail addresses: yeeho.chai@gmail.com (Y.H. Chai), drsuzana_yusuf@petronas.
com.my (S. Yusup), calvinchok@platinumgroup.com.my (V.S. Chok), taufiq@
platinumgroup.com.my
(M.T.
Arpin),
drsonny_irawan@petronas.com.my
(S. Irawan).
http://dx.doi.org/10.1016/j.applthermaleng.2016.07.017
1359-4311/ 2016 Elsevier Ltd. All rights reserved.

of high L/D aspect ratio that gives high specific surface area
between the tubes for higher heat transfer rate. CNTs have
unusually high thermal conductivity on average of 3000 W/m K
[12,14] for MWCNT and 6000 W/m K [1] for SWCNT. However,
the dispersion and stability of MWCNT based nanofluids are
usually difficult as the tubes tend to entangle with one another
to form larger mass of aggregates and cause instabilities of the
nanofluid. This can be attributed to the high Van der Waals force
at a nano-scale that promotes particle-particle attractions.
Nanofluids have higher thermal conductivities as compared to
its base fluids. Liu et al. [3] obtained 12.4% thermal conductivity
enhancement when ethylene glycol is dispersed with 1 vol% CNT.
Ding et al. [1] obtained as much as 80% enhancement with only
1 wt% of CNT dispersed in water. Similarly, Jiang et al. [2] obtained
14% enhancement when deionized water was dispersed with 1 vol
% CNT. Farbod et al. [4] yielded approximately 15% enhancement
when 0.5 vol% CNT was dispersed in water. The wide variation in

1020

Y.H. Chai et al. / Applied Thermal Engineering 107 (2016) 10191025

thermal conductivity enhancement of CNT nanofluids can be


attributed to several factors such as Brownian motion, nanolayer
formation, nanoparticle clustering, size and volume fraction of
nanoparticles [33]. Furthermore, the influence of base fluids plays
an important role in thermal conductivity of nanofluids. Xie et al.
[5] compared three different base fluids, namely water and
ethylene glycol, dispersed with Al2O3 nanoparticles. In their
findings, ethylene glycol gave better performance by 10% thermal
conductivity enhancement when compared to water.
Most applications of nanofluids are pertained to heat transfer
applications. However, a different class of nanofluids is applicable
in oil and gas sector. Choi [6] predicted one of many applications of
nanofluids could be diversified to deep-hole drilling due to the
enhanced cooling and lubrication properties of nanofluids.
Al-Yasiri and Al-Sallami [7] stated that nanomaterial-enhanced
drilling fluid could deliver several prospective performance
including higher resistance to degradation at high temperature
high pressure (HTHP) conditions. Sedaghatzadeh et al. [8] delivered over 30% thermal conductivity enhancement in CNT-water
based drilling fluid.
Cavitation process can be classified into three stages, namely
generation, growth and collapse of bubbles as shown in Fig. 1.
The collapse of bubbles usually emits results in emission of high
densities from order 1 to 1018 kW/m3 [9] that can result in sudden
shear force energy around the bubbles capable of breaking bonds
of polymeric materials. Gogate et al. [9] compared different
cavitational equipment in terms of energy efficiency and concluded that hydrodynamic cavitation gives higher cavitation yield
and better energy efficiency for large nanofluid productions.
Baldyga et al. displayed deagglomeration process at extreme stresses that trigger the shattering mechanism [10].
Hydrodynamic cavitation is more preferred to ultrasonication
when carrying out in large scale operations [11] including breakdown of solids in wastewater treatment. Furthermore, a comparison study carried out by Sharma et al. [11] showed energy
efficiency of hydrodynamic cavitation to be second most efficient
when compared to ultrasonication processes while having the
highest cavitational yield.
To the authors knowledge, there has been no study on CNT
particles dispersion into hydrogenated oil-based fluid via
hydrodynamic cavitation and ultrasonication combination for the
production of nanofluids. Research work that relates similar to
the studys experimental method which utilizes high physical
energy for dispersion of nanoparticles was carried out by Kawano
and Kondo [30] for the production of hyperhydrophobic fullerene,
C60, nanofluid suspensions via aqueous counter collision. However,
high energy intensity ultrasonication is still considered the most
practiced method to produce nanofluids currently. Wang et al.
[31] sonicated multi-walled CNTs and graphene into ionic liquid
for 8 h with frequency of 40 kHz at 100 W power while Ruan and
Jacobi [32] dispersed CNT into ethylene glycol using tip ultrasonicator at 150 W and frequency of 20 kHz.

2. Experimental procedure
2.1. Materials
Multi-walled CNTs and hydrogenated oil-based drilling fluid
were provided by Platinum Green Chemicals Sdn. Bhd., Malaysia
and was used as received.
Hydrogenated oil contains a mixture of straight chain and
branched paraffins of carbon chain length from C15 to C18 that
constitute it as non-polar group intrinsically. The general thermophysical properties of hydrogenated oil as base fluid used in this
work are listed in Table 1.
The size and morphology of CNTs were taken through transmission electron microscopy (TEM, LIBRA Microscope 200). At 80,000
magnification, a bundle of CNT with an average outer diameter of
10 to12 nm is seen to entangle one another at Fig. 2(a). The entanglement could result in the aggregation of CNTs in nanofluid if not
dispersed homogenously. From Fig. 2(b), multi-walled CNT consists of inner diameter that are layered with several outer layers
of CNTs at an average thickness of 4.7 nm. The inner diameter of
the multi-walled CNTs ranges from 6 to 7 nm. The physical properties of MWCNT is summarised in Table 2.
Fourier Transform Infrared (FTIR) spectrometer (Perkin Elmer)
was used to detect the compositions present on the multi-walled
CNTs within the wavenumber range of 500 to 4000 cm1 as shown
in Fig. 3. Between 3300 cm1 and 3600 cm1, high absorbance of
OAH groups are seen attached to the surface of CNT. The peaks
between 2800 and 3500 cm1 are usually stretching vibrations of
OAH bonds [4]. At the range of 1500 and 1800 cm1, small peaks
are formed which confirms the presence of C@C bonds. There are
also sharp peaks ranging from 1600 to 1700 cm1 which denotes
the presence of C@O stretching for carboxylic groups (ACOOH).
The presence of hydroxyl and carboxylic groups confirms the
hydrophilicity behaviour of CNT [12]. Between 1000 and
1300 cm1, the CH2X group can be found on the surface of the
CNT. This could be attributed to the attachment of halogen atoms
to form alkyl halides during acid wash during the preparation of
CNT.
2.2. Nanofluid preparation
The dispersion of multi-walled CNTs into hydrogenated oilbased drilling fluid was carried out via two-step methods which
includes hydrodynamic cavitation dispersion and bath ultrasonication process.
A pilot scale hydrodynamic cavitation unit was set up as shown
in Fig. 4 for hydrodynamic cavitation dispersion for the production
of CNT-hydrogenated oil-based fluid. A variable frequency drive
(VFD) pump with a range of 050 Hz is used to control the inlet
flow pressure at the orifice. Table 3 below outlines the key parameters used for the HC dispersion process.
The minimum volume requirement for each batch is 1.5 L of
hydrogenated oil-based fluid to prevent the pump from
dry-running and was taken as the basis for each run. The weight
concentrations of MWCNTs selected in this study are 25 ppm,
50 ppm and 100 ppm respectively. Weighted MWCNTs were
dispersed into measured hydrogenated oil-based fluid for prestirring for 15 min at 800 RPM (VELP Scientifica Digital Overhead

Table 1
Thermophysical properties of hydrogenated oil.

Fig. 1. Schematic illustration of the stages of bubble implosion during hydrodynamic cavitation process.

Density
(kg/m3)

Viscosity (cP)

Flash
point (C)

Vapour
pressure (kPa)

780 (at 15 C)

1.52.0 (at 40 C)

90

<0.1 (at 40 C)

Y.H. Chai et al. / Applied Thermal Engineering 107 (2016) 10191025

1021

Fig. 2. TEM imaging of multi-wall CNT at 80,000 (a) and 630,000 (b) magnification.

Table 2
Physical properties of MWCNTs.
Parameters

Dimension
3

Density (g/cm )
Carbon content (%)
Oxygen (%)
XY dimensions (lm)
Z dimensions (lm)
Thermal conductance (W/m K)

2.12
>99.0
<1
0.060.1
110
1600

Fig. 3. Fourier Transform Infrared (FTIR) spectra of MWCNTs.

Stirrer DLS) before transferred into the HC unit as part of the


pre-homogenization process. Pre-homogenization process was
required as there were instances of CNTs settled down at the walls
of HC unit that rendered the dispersion process ineffective. The
transferred MWCNT-hydrogenated oil sample was circulated at
an average of 1.5 L min1 for 3 h continuously through a 1 mm
diameter orifice. A stirrer was installed at the feed tank and operated at 200 RPM to ensure the CNT-hydrogenated dispersed does
not settle down after each pass. Dispersed CNT-hydrogenated oil
was transferred to bath ultrasonicator (Bath Ultrasonic Branson
8510E-DTH) at 320 W power with 40 kHz frequency and sonicated
for 3 h to prevent multi-walled CNTs from entanglement which
may lead to aggregation.

2.3. Thermal conductivity analysis


Thermal conductivity analysis was carried out using KD2 Pro
Thermal Properties Analyzer equipped with sensor KS-1 type

(1.3 mm diameter  60 mm length). The KS-1 sensor are able to


detect thermal conductivity that ranges from 0.02 W m1 K1 to
2.00 W m1 K1 with a sensitivity of 0.001 C in temperature
change. KD2 Pro uses transient heat line source method that consists of the needle with a heater and a temperature sensor within.
To determine the effect of temperature on the effective thermal
conductivity of CNT-hydrogenated oil, KD2 Pro was calibrated with
glycerol at constant temperature beforehand. For a more accurate
and complete equilibrium of sample temperature and the external
surroundings, each samples were sonicated before each analysis
and were left to equilibrate in water bath (Memmert Water Bath)
at set temperature ranging from 30 C to 50 C for 20 min before
each analysis to prevent free convection within sample that might
result in error in thermal properties measurement. The upper limit
is set at 50 C as exceeding the upper limit will return fluctuating
results. This is attributed to the low viscosity of hydrogenated oil
as compared with other vegetable oils outlined in Table 4.
Sample temperature higher than the set temperature will be
subjected to free convection within the sample that can potentially
disrupt the accuracy of the measurement readings. At least four
measurements were taken for each CNT-hydrogenate oil concentration at each set temperature with 3% uncertainty of
measurements.
3. Results and discussions
3.1. Stability of nanofluids
MWCNT suspended in hydrogenated oil are stable for only four
days before fully settled down. MWCNT used in this study are
hydrophilic in nature while hydrogenated oil is a non-polar group
solvent. The dispersion of MWCNTs into hydrogenated oil yielded
poor performance in terms of stability due to the hydrophilicity of
nanoparticles. Fig. 5 shows the stability of MWCNT-hydrogenated
oil after dispersion.
However, the stability of MWCNT-hydrogenated oil could be
further improved by dispersing surfactants or using MWCNTs
which are hydrophobic in nature as a choice for better stability.
3.2. Thermal conductivity of nanofluids
Thermal conductivity enhancement of MWCNT-hydrogenated
oil with different concentrations and temperature was carried
out and investigated. The concentrations of MWCNT dispersed
are 25 ppm, 50 ppm and 100 ppm respectively. Fig. 6 shows the

1022

Y.H. Chai et al. / Applied Thermal Engineering 107 (2016) 10191025

Fig. 4. Schematic diagram of hydrodynamic cavitation unit (HDV: hydrodynamic vessel; MV: mixer vessel; PG: pressure gauge; RP: rotary pump; HDP: hydrodynamic pump).

Table 3
Key parameters of hydrodynamic cavitation unit.
Parameters

Dimensions

Orifice diameter
Orifice length
Pipe inlet diameter
Tank volume
Pump power

1 mm
30 mm
19 mm
10 L
0.4562 kW

Table 4
Viscosity comparison between hydrogenated oil and other vegetable oil.
Type

Viscosity (cP)

Reference

Hydrogenated oil
EFB-derived pyrolytic oil
Corn oil
Coconut oil

1.52.0 (at 40 C)
45.47 (at 40 C)
30.8 (at 37.8 C)
28.0 (at 37.8 C)

This work
[34]
[35]
[35]

Fig. 5. Dispersed CNT-hydrogenated oil at 25 ppm (left), 50 ppm (middle) and


100 ppm (right).

comparison of thermal conductivity of hydrogenated oil as base


fluid against thermal conductivity of CNT-hydrogenated oil at temperatures ranging from 30 C to 50 C with a step increase of 5 C.
From Fig. 6, thermal conductivity enhancement are not substantial at lower temperatures regardless of the concentrations of

Fig. 6. Comparison of thermal conductivity enhancement CNT-hydrogenated oil


samples against base fluid.

MWCNT dispersed. Weight concentration of MWCNT in this study


are much lesser as compared to other researchers. Ijam et al. [13]
proved that not more than 1% enhancement is seen at 0.01 wt%
graphene oxide nanoparticle in water. Liu et al. [3] showed only
1.6% enhancement ratio with 0.2 vol% CNT when dispersed in ethylene glycol but 12.4% increment ratio can be seen with 1.0 vol%
CNT. Various researchers performed studies of low volume
concentration of CNT in various base fluids with negligible thermal
conductivity enhancement [8]. At low nanoparticle concentrations,
thermal conductivity enhancement are improved by a small
margin at lower temperatures as compared to higher temperature.
This observation confirmed that the volume fraction of nanoparticles play a crucial role in thermal enhancement even at low
temperatures.
MWCNT-hydrogenated oil showed the highest enhancement of
9.8% for loading of 100 ppm concentration at 50 C. There are slight
inconsistencies such as the sudden performance drop at 50 ppm at
40 C and sudden jumps in thermal conductivity values from
35 C to 40 C at 25 ppm and 50 ppm concentration. These fluctuations are affected by the difference in Brownian motion intensity
caused by micro-convection within the fluid. This prompted CNTs
with longer length to move slower compared to CNTs with shorter
length that can result in entanglement of CNTs [22]. This will affect
the thermal conductivity of nanofluids as particle size plays a crucial role in thermal conductance as well.
Nanoparticle aggregation can be represented as a local
percolation structure which improves the effective thermal conductivity of nanofluids but in return destabilizing the stability

1023

Y.H. Chai et al. / Applied Thermal Engineering 107 (2016) 10191025

and homogeneity of nanofluids [33]. Aggregations decreases the


specific surface area and induces weak thermal transfer between
CNTs [2] which will subsequently settle down due to the increase
in density of aggregates. Similar observations by Nasiri et al. [12]
showed declining trend of thermal conductivity of nanofluids due
to settlement of aggregated clusters which reduce the overall
nanofluid thermal conductivity.
3.3. Comparisons of experimental results and model predictions

bf

bf

HamiltonCrosser (HC) model is used to predict the thermal


conductivity of solid-liquid mixture. The model is given by



knf
kp n  1kbf n  1kp  kbf u
1
kbf
kp n  1kbf kp  kbf u





kp  keff
kbf  keff
1  u
0
kp keff
kbf 2keff

Classical thermal conductivity models are used to predict the


thermal conductivity based on nanoparticle volume fraction,
thermal conductivity of nanoparticles and thermal conductivity
of base fluid. However, results from these models are generally
underdetermined as compared to experimental data [2729]. Several conventional models based on solid-liquid mixture were used
to predict the thermal conductivity enhancement of nanofluids.
Conventional thermal conductivity models usually predict based
on several main key parameters such as volume fraction (u),
thermal conductivity of nanoparticle (kp), thermal conductivity of
base fluid (kbf), and shape factor (n) for nanoparticle types.
Experimental data results are compared with theoretical predictions by Maxwell model [23], Hamilton and Crosser model
[24], Bruggeman model [25], and Nan et al. model [26] as shown
in Fig. 7.
Maxwell model predicts well with dilute solid-liquid mixture
with particle at micro size and with spherical solid particles [15].
Maxwell model is given by



k
3 kp  1 u
knf
bf 


kp
k
kbf

2
 kp  1 u
k

taken for spherical particles while n = 6 is considered for cylindrical


shape particles.
Bruggeman model predicts binary mixtures which was developed based on symmetrical effective medium theory (EMT). The
model does not have any concentration limitations for spherical
particles. However, at low particle concentration, Bruggeman
model predict similarly to the Maxwell model. The model is given
by

HC model takes account of the shape factor, n = 3/w, of the


nanoparticles where w is the sphericity factor. Generally, n = 3 is

Nan et al. model predicts the effective thermal conductivity of


solid-liquid composites by taking account of interfacial thermal
resistance. The model is given by

knf 3 b11 b33 u

kbf
3  b11 u

where b11 and b33 are given by


c

b11

2k11  kbf
;
c
k11 kbf

b33

k33
1
kbf

The parameters k11 and k33 are transverse and longitudinal


equivalent thermal conductivities of the nanotube that takes
account of the length, l, and diameter, d, of the nanotube.
c

k11

kp
1 2dak

kp
kbf

k33

kp
1 2dak

kp
kbf

where ak is the Kapitza radius and is expressed by ak = kbfRk.


The value of Rk depicts the thermal resistance which is
8.3  108 K m2 W1 [16]. For this study, the length-to-diameter
(L/D) ratio for multi-walled CNT is taken to be 2500 [17], assuming the multi-walled CNT suspended possessed extensive length.
The thermal conductivity values of hydrogenated oil, kbf, and
multi-walled CNT, kp, is taken as 0.1825 W m1 K1 and
2000 W m1 K1 respectively.
The calculated values predicted by Maxwell model, HC model
and Bruggeman model underpredicted the experimental data
obtained. This is in agreement with other findings that proved that

Fig. 7. Comparison of classical models against experimental data at 25 ppm (top left), 50 ppm (top right) and 100 ppm (bottom).

1024

Y.H. Chai et al. / Applied Thermal Engineering 107 (2016) 10191025

classical models predicted lower values compared to experimental


data [18,20,22]. Interestingly, the predicted values approached the
experimental data when the particle concentration increases.
However, Xue claimed that Nan et al. model overestimated the
thermal conductivity of the experimental data due to the inconsideration of the aggregation effects [16]. Agglomerated particles tend
to have lower heat transfer between particles as specific surface
area of the agglomerates reduces significantly. It is interesting to
note that at very low particle loadings, Nan model is able to predict
closely to the experimental data at higher temperature as compared to other classical models but deviates further when particle
loading increases.
Classical models such as HC model and Bruggeman model are
extensions of Maxwells model. There has been several attempts by
other researchers to fine-tune the Maxwell model to take in the
effect of particle size in a homogenous system [21,22]. Most classical models does not take account of particle aggregations, resulting
in underprediction of thermal conductivity data.
3.4. Effective thermal conductivity of nanoparticle
Kole and Dey [27] explained that a benchmark study on thermal
conductivity findings by other researchers was generalized by Nan
et al. Therefore, Nan et al. model is used to predict the effective
thermal conductivity of composites for completely misoriented
ellipsoidal particles [16], which is given by

knf kbf

3 u2b11 1  L11 b33 1  L33 


3  u2b11 L11 b33 L33

where u and Lii are volume fraction and geometrical factor respectively. bii is defined as

bii

kp  kbf
kbf Lii kp  kbf

where kp and kbf are thermal conductivity of nanoparticles and base


fluids respectively. Since multi-walled CNTs possessed high aspect
ratio, it is assumed L11 = 0 and L33 = 1. A least square fitting of the
experiment data was carried out based on Eq. (7) to obtain the inplane thermal conductivity of multi-walled CNT at 50 C to be at
3.09 0.9 W m1 K1. As observed from Fig. 8, the effective thermal
conductivity of CNT has been drastically reduced as compared to
intrinsic thermal conductivity of suspended CNT [12,14]. Nan
et al. model takes account of matrix-additive interface contact resistance into consideration [19] which yielded lower effective thermal
conductivity of CNT when dispersed in hydrogenated oil. Apart from
that, low effective thermal conductivity could be affected by the
possibility of comprising various structural defects on CNT. Hydrodynamic cavitation dispersion requires high pressure and power to

Fig. 8. Estimation of effective thermal conductivity of multi-wall CNT in the


prepared nanofluids by Nan et al. model.

break down agglomerates into smaller sizes. The shear pressure in


breaking down the structure of CNT at micro level could possibly
cause severe defects on the surface of CNT that strained the structure itself into bends and kinks.
4. Conclusions
This study investigated the effects of multi-walled CNT
nanoparticle addition into hydrogenated oil towards the thermal
conductivity of the nanofluid. Multi-walled CNTs were dispersed
into hydrogenated oil-based fluid via two-step methods combining
hydrodynamic cavitation and ultrasonication processes. It is
observed that thermal conductivity enhancement by nearly 10%
is obtainable at a very low nanoparticle loadings but the effective
thermal conductivity of multi-walled CNTs are drastically reduced
at 3.09 0.9 W m1 K1 as compared to their intrinsic values. This
can be accounted to the structural defects that were formed during
cavitation dispersion process. While other thermal conductivity
models underpredicted the experimental values, Nan et al. model
gave a more accurate prediction at lower particle loadings at
25 ppm but deviates greatly at 50 ppm and 100 ppm.
Acknowledgements
The authors would like to thank Platinum Green Chemicals Sdn.
Bhd. for supporting the required materials and equipment to carry
out this investigation. This research is in collaboration with Platinum Green Chemicals Sdn. Bhd., Malaysia. The authors would also
like to thank Centralized Analytical Laboratory (CAL), Universiti
Teknologi PETRONAS for their kind assistance in carrying out FTIR
analysis for this study.
References
[1] Y. Ding, H. Alias, D. Wen, R.A. Williams, Heat transfer of aqueous suspensions
of carbon nanotubes (CNT nanofluids), Int. J. Heat Mass Transfer 49 (2006)
240250.
[2] H. Jiang, Q. Zhang, L. Shi, Effective thermal conductivity of carbon nanotubebased nanofluid, J. Taiwan Inst. Chem. Eng. 55 (2015) 7681.
[3] M. Liu, M.C. Lin, I. Huang, C. Wang, Enhancement of thermal conductivity with
carbon nanotube for nanofluids, Int. Commun. Heat Mass Transfer 32 (2005)
12021210.
[4] M. Farbod, A. Ahangarpour, S.G. Etemad, Stability and thermal conductivity of
water-based carbon nanotube nanofluids, Particuology 22 (2015) 5965.
[5] H. Xie, J. Wang, T. Xi, Y. Li, F. Ai, Dependence of the thermal conductivity of
nanoparticle-fluid mixture on the base fluid, J. Mater. Sci. Lett. 21 (2002) 1469
1471.
[6] S.U.S. Choi, Nanofluids: from vision to reality through research, J. Heat Transfer
131 (2009).
[7] M.S. Al-Yasiri, W.T. Al-Sallami, How the drilling fluids can be made more
efficient by using nanomaterials, Am. J. Nano Res. Appl. 3 (3) (2015) 4145.
[8] M. Sedaghatzadeh, A.A. Khodadadi, M.R.T. Birgani, An improvement in thermal
and rheological properties of water-based drilling fluids using multiwall
carbon nanotube (MWCNT), Iran. J. Oil Gas Sci. Technol. 1 (2012) 5565.
[9] P.R. Gogate, R.K. Tayal, A.B. Pandir, Cavitation: a technology on the horizon,
Curr. Sci. 91 (1) (2006).
[10] J. Baldyga, L. Makowski, W. Orciuch, C. Sauter, H.P. Schuchmann, Agglomerate
dispersion in cavitating flows, Chem. Eng. Res. Des. 97 (2009) 474484.
[11] A. Sharma, P.R. Gogate, A. Mahulkar, A.B. Pandit, Modelling of hydrodynamic
cavitation reactors based on orifice plates considering hydrodynamic and
chemical reactions in a bubble, Chem. Eng. J. 143 (2008) 201209.
[12] A. Nasiri, M.S. Niasar, A. Rashidi, A. Amrollahi, R. Khodafarin, Effect of
dispersion method on thermal conductivity and stability of nanofluid, Exp.
Thermal Fluid Sci. 35 (2011) 717723.
[13] A. Ijam, R. Saidur, P. Ganesan, A.M. Golsheikh, Stability, thermo-physical
properties, and electrical conductivity of graphene oxide-deionized water/
ethylene glycol based nanofluid, Int. J. Heat Transfer 87 (2015) 92103.
[14] N.M. Phan, H.T. Bui, M.H. Nguyen, H.K. Phan, Carbon-nanotube-based liquids:
a new class of nanomaterials and their applications, Adv. Nat. Sci.: Nanosci.
Nanotechnol. 5 (1) (2014).
[15] S.M.S. Murshed, K.C. Leong, C. Yang, Enhanced thermal conductivity of TiO2water based nanofluids, Int. J. Therm. Sci. 44 (2005) 367373.
[16] Q.Z. Xue, Model for the effective thermal conductivity of carbon nanotube
composites, Nanotechnology 17 (2006) 16551660.

Y.H. Chai et al. / Applied Thermal Engineering 107 (2016) 10191025


[17] R.K.A. Al-Rub, A.I. Ashour, B.M. Tyson, On the aspect ratio effect of multiwalled carbon nanotube reinforcements on the mechanical properties of
cementitious nanocomposites, Constr. Build. Mater. 35 (2012) 647655.
[18] W. Yu, S.U.S. Choi, The role of interfacial layers in the enhanced thermal
conductivity of nanofluids: a renovated Maxwell model, J. Nanopart. Res. 5
(2003) 167171.
[19] W. Yu, H. Xie, X. Wang, X. Wang, Significant thermal conductivity
enhancement for nanofluids containing graphene nanosheets, Phys. Lett. A
375 (2011) 13231328.
[20] P. Estell, S. Halelfadl, T. Mar, Thermal conductivity of CNT water based
nanofluids: experimental trends and models overview, J. Therm. Eng. 1 (2)
(2015) 381390.
[21] M. Shaker, E. Birgersson, A.S. Mujumdar, Extended Maxwell model for the
thermal conductivity of nanofluids that accounts for nonlocal heat transfer,
Int. J. Therm. Sci. 84 (2014) 260266.
[22] F.C. Li, J.C. Yang, W.W. Zhou, Y.R. He, Y.M. Huang, B.C. Jiang, Experimental
study on the characteristics of thermal conductivity and shear viscosity of
viscoelastic-fluid-based nanofluids containing multiwalled carbon nanotubes,
Thermochim. Acta 556 (2013) 4753.
[23] J.C. Maxwell, A Treatise on Electricity and Magnetism, 1st ed., vol. 1, Clarendon
Press, Oxford, UK, 1873.
[24] R.L. Hamilton, O.K. Crosser, Thermal conductivity of heterogeneous twocomponent systems, Ind. Eng. Chem. Fundam. 1 (1962) 187191.
[25] D.A.G. Bruggeman, Berechnung verschiedener physikalischer Konstanten von
heterogenen Substanzen. I. Dielektrizittskonstanten und Leitfhigkeiten der
Mischkrper aus isotropen Substanzen, Annalen der Physik. Leipzig, vol. 24,
1935, pp. 636679.

1025

[26] C.W. Nan, R. Birringer, D.R. Clarke, H. Gleiter, Effective thermal conductivity of
particulate composites with interfacial thermal resistance, J. Appl. Phys. 81
(10) (1997) 66926699.
[27] M. Kole, T.K. Dey, Investigation of thermal conductivity, viscosity, and
electrical conductivity of graphene based nanofluids, J. Appl. Phys. 113 (2013).
[28] J.H. Lee, S.H. Lee, C.J. Choi, S.P. Jang, S.U.S. Choi, A review of thermal
conductivity data, mechanisms and models for nanofluids, Int. J. Micro-Nano
Scale Transport 1 (4) (2010).
[29] S.M. Hosseini, A. Moghadassi, D. Henneke, Modelling of the effective thermal
conductivity of carbon nanotube nanofluids based on dimensionless groups,
Can. J. Chem. Eng. 89 (2011) 183186.
[30] Y. Kawano, T. Kondo, Preparation of aqueous carbon material suspensions by
aqueous counter collision, Chem. Lett. 43 (2014) 483485.
[31] F. Wang, L. Han, Z. Zhang, X. Fang, J. Shi, W. Ma, Surfactant-free ionic liquidbased nanofluids with remarkable thermal conductivity enhancement at very
low loading of graphene, Nanoscale Res. Lett. 7 (2012).
[32] B. Ruan, A.M. Jacobi, Ultrasonication effects on thermal and rheological
properties of carbon nanotube suspensions, Nanoscale Res. Lett. 7 (2012).
[33] H.S
. Aybar, M. Sharifpur, M.R. Azizian, M. Mehrabi, J.P. Meyer, A review of
thermal conductivity models for nanofluids, Heat Transfer Eng. (2015).
[34] Y.H. Chan, K.V. Dang, S. Yusup, M.T. Lim, A.M. Zain, Y. Uemura, Studies on
catalytic pyrolysis of empty fruit bunch (EFB) using Taguchis L9 Orthogonal
Array, J. Energy Inst. 87 (2014) 227234.
[35] H. Noureddini, B.C. Teoh, L.D. Clements, Viscosities of vegetable oils and fatty
acids, Papers in Biomaterials (1992).

Вам также может понравиться