Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

ICHMT-03422; No of Pages 9

International Communications in Heat and Mass Transfer xxx (2016) xxx–xxx

Contents lists available at ScienceDirect

International Communications in Heat and Mass Transfer

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

1 Heat transfer and friction factor of water and ethylene glycol mixture based TiO2 and
2 Al2O3 nanofluids under turbulent flow☆

F
3Q1 W.H. Azmi a,⁎, K. Abdul Hamid a, N.A. Usri a, Rizalman Mamat a, M.S. Mohamad b

O
4 a
Faculty of Mechanical Engineering, Universiti Malaysia Pahang, 26600 Pekan, Pahang, Malaysia
5 b
Faculty of Industrial Sciences & Technology, Universiti Malaysia Pahang, Lebuhraya Tun Razak, 26300 Gambang, Kuantan, Pahang, Malaysia
6

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

R
8
9 Available online xxxx It has been a great challenge in heat transfer to provide efficient thermal fluids for cooling purposes especially in 15
14
13
12
11
10 engineering practice. The concerns on various operating temperatures become the main concern in the present 16

P
30 Keywords: study to investigate the heat transfer and friction factor of titanium oxide (TiO2) and aluminium oxide (Al2O3) 17
31 Heat transfer under turbulent flow in a tube. The nanofluids were prepared using the two-step method and dilution process 18
32 Friction factor for volume concentrations of 0.5% to 1.0% in a mixture of water (W) and ethylene glycol (EG) at a volume ratio 19
33 Nanofluids
of 60:40 (W:EG). The convective heat transfer investigations were conducted at a constant heat flux boundary
D 20
34 Thermal conductivity
35 Viscosity
condition and operating temperatures of 30, 50 and 70 °C. The enhancement of thermal conductivity and viscos- 21
ity of Al2O3 was found to be influenced by the temperature while the enhancement of the TiO2 nanofluid prop- 22
erties was observed to be independent of temperature. Both Al2O3 and TiO2 nanofluids were observed to have 23
E
almost the same values of heat transfer coefficients for 1.0% concentration at 50 and 70 °C with an average en- 24
hancement of 24%. However, the heat transfer coefficients of Al2O3 nanofluids were found to be higher than 25
T

TiO2 nanofluids at the operating temperature of 30 °C. The heat transfer concentrations increased with volume 26
concentration and observed for both types of nanofluids at all operating temperatures. The friction factors for 27
28
C

both TiO2 and Al2O3 nanofluids slightly increased with volume concentration.
© 2016 Published by Elsevier Ltd. 29
39
37
36
E

38

40 1. Introduction The experimental determination of thermo-physical properties such 56


R

as viscosity and thermal conductivity of nanofluids has been explored 57


41 Forced convection heat transfer plays an important role in cooling by many researchers [8–16]. Sundar et al. [8] conducted a study on ther- 58
R

42 components and system especially in engineering practices. The process mal properties using Fe3O4 nanofluids in a mixture of EG and water and 59
43 involves heat movements such as heat added or removed from one found that the thermal conductivity of nanofluids increased as concen- 60
44 process to another. However, the poor thermal ability inherent by trations and temperatures of the nanofluids increased. Apart from these 61
O

45 conventional fluids puts a limitation on heat transfer to give the best two outstanding factors, the types of nanoparticles used also contribute 62
46 performance. Therefore, it is important to develop a new heat transfer to the thermal conductivity enhancement, as well as particle size and 63
C

47 fluid which can give high thermal performance compared to common the stability of the nanofluid [9,10]. Other factors that affect the thermal 64
48 fluids. In order to overcome the problem, the nanofluid was introduced conductivity are the particle shape and aggregation as found by Chen 65
49 by Masuda et al. [1]. Through Maxwell's [2] research on the possibilities et al. [11]. Thermal conductivity mostly increases with the addition of 66
N

50 of increasing thermal conductivity of a fluid–solid mixture, further re- nanoparticles into the base fluid. However, the enhancement may 67
51 search using particles with micrometre or even millimetre dimensions vary as proven by Turgut et al. [12] where the finding of their studies 68
U

52 was used [3]. With early research on using suspended nano-sized parti- showed that there was no temperature dependence related to the 69
53 cles in the base fluid [4–6], it is convinced that nanofluids have the enhancement of thermal conductivity. 70
54 potential to be the next-generation coolants and giving development For another important properties study, viscosity decreases with the 71
55 in achieving advance performance of the cooling system [7]. increase of temperature and increases with the increase of concentra- 72
tion [12]. The majority of previous studies conducted suggested that 73
the nanofluids have the best performance with the combination of 74
high thermal conductivity and low viscosity [13,14]. It also becomes a 75
☆ Communicated by Dr. W.J. Minkowycz.
major point in nanofluid studies because the nanofluid is expected to 76
⁎ Corresponding author.
show an increase in thermal conductivity without an increase in 77
E-mail addresses: wanazmi2010@gmail.com (W.H. Azmi), khamisah0301@gmail.com
(K. Abdul Hamid), nurashikinusri@gmail.com (N.A. Usri), rizalman@ump.edu.my pressure drop, which in turn is related to fluid viscosity [15]. Namburu 78
(R. Mamat), mohdsham@ump.edu.my (M.S. Mohamad). et al. [16] found a decrease in exponential pattern for viscosity 79

http://dx.doi.org/10.1016/j.icheatmasstransfer.2016.05.010
0735-1933/© 2016 Published by Elsevier Ltd.

Please cite this article as: W.H. Azmi, et al., Heat transfer and friction factor of water and ethylene glycol mixture based TiO2 and Al2O3 nanofluids
under turbulent flow, Int. Commun. Heat Mass Transf. (2016), http://dx.doi.org/10.1016/j.icheatmasstransfer.2016.05.010
2 W.H. Azmi et al. / International Communications in Heat and Mass Transfer xxx (2016) xxx–xxx

Rao [21] with nanoparticles of TiO2 proved that the use of the mixture 100
T1:1 Nomenclature
of water/EG can also provide heat transfer enhancement. 101
There are various types of nanoparticles that have been studied for 102
T1:2 A Area, m2
the investigation of nanofluid heat transfers up till today. However, 103
T1:3 Cp Specific heat, J/kg.K
the comparison of two different types of water/EG mixture based 104
T1:4 d Diameter, m
nanofluids for the evaluation on heat transfer performance under simi- 105
T1:5 EG Ethylene glycol
lar working temperature is still limited in literature. Hence, the present 106
T1:6 f Friction factor
work investigates the effect of TiO2 and Al2O3 nanoparticles with the 107
T1:7 FESEM Field emission scanning electron microscopy
average size of 13 and 50 nm, respectively dispersed in W:EG (60:40) 108
T1:8 h Heat transfer coefficient, W/m2.K
mixture on convective heat transfer coefficient at various working 109
T1:9 I Current, A
temperatures. 110
T1:10 k Thermal conductivity, W/m.K
T1:11 L Length, m
2. Methodology 111
T1:12 LPM Liter per minutes
T1:13 m _ Mass flow rate, kg/s

F
2.1. Sample preparation 112
T1:14 Nu Nusselt number
T1:15 Pr Prandtl number

O
Two types of oxide materials used in the present study are titanium 113
T1:16 Q Rate of heat transfer
oxide (TiO2) and aluminium oxide (Al2O3) procured from US Research 114
T1:17 Re Reynolds number
Nanomaterials, Inc. (USA) and Sigma Aldrich (USA), respectively. The 115
T1:18

O
T Temperature, °C
TiO2 nanoparticles were suspended in water with 40 wt.% concentration 116
T1:19 TEM Transmission electron microscopy
(13.6% volume concentration) and 50 nm in size. On the other hand, 117
T1:20 v Velocity, m/s2
Al2O3 was in powder form with 13 nm in size. Eq. (1) is used to convert 118

R
T1:21 V Volume, m3
weight concentration of TiO2 nanofluids to volume concentration. 119
T1:22
T1:23 Vt Voltage, V
Meanwhile, Eq. (2) is used to find the mass of Al2O3 nanoparticles to 120

P
the desired volume concentration by the two step preparation. Both 121
T1:24 Greek symbols
TiO2 and Al2O3 nanofluids were diluted to low concentrations by adding 122
T1:25 ΔP Pressure drop, Pa
the base fluid into the solution using Eq. (3). The samples were observed 123
T1:26 ϕ Volume concentration, % D
to be stable for over two months for both Al2O3 and TiO2 nanofluids. The 124
T1:27 ϕ Volume fraction
summary of Al2O3 and TiO2 nanoparticle properties [14], the range of pH 125
T1:28 ω Weight concentration, %
and nanofluid concentrations are shown in Table 1. 126
T1:29 ρ Density, kg/m3
E

T1:30
T1:31 μ Dynamic viscosity, kg/m.s
ωρbf
ϕ¼ ω  ð1Þ
T

ω
T1:32 Subscripts 1− ρ þ ρ
100 p 100 bf
T1:33 bf Base fluid 128
C

T1:34 DB Dittus–Boelter 129


T1:35 exp Experimental mp =ρp
ϕ¼  100 ð2Þ
T1:36 mp =ρp þ mbf =ρbf
E

nf Nanofluid
131
T1:37 p Particle
132
T1:38 s Surface  
ϕ
R

T1:39 B Bulk ΔV ¼ ðV 2 −V 1 Þ ¼ V 1 1 −1 ð3Þ


ϕ2
T1:40 1 Initial
R

T1:41
T1:42 2 Final 134
A mixture of water and ethylene glycol (EG) at a volume ratio of
60:40 was used as the base fluid in the present study. A volume of 135
O

22 L of nanofluids is required to conduct the experiment. The nanofluids 136


prepared were subjected to a mixing process using mechanical stirrer 137
80 properties using CuO nanoparticles dispersed in 60:40 (EG:W) when for 30 min and underwent sonication process using Fisherbrand ultra- 138
C

81 the temperature elevated from 10 to 50 °C. While for base fluids, the vis- sonic bath for 2 h for each concentration prepared to enhance the 139
82 cosity also decreases by about 73.21% while the temperature increases dispersion and stability of nanofluids [22]. The thermal conductivity 140
from −35 to 50 °C.
N

83 and dynamic viscosity of nanofluids were measured with KD2 Pro 141
84 The experimental studies on nanofluid heat transfer were started by Thermal Analyser and Brookfield LVDV-III Rheometre, respectively at 142
85 Pak and Cho [17] under turbulent flow using Al2O3–water and TiO2– temperatures from 30 to 70 °C. The mixture relations for density and 143
U

86 water nanofluids in a circular pipe and continued by many researchers specific heat of nanofluids are given as Eqs. (4) and (5), respectively. 144
87 after that. Suresh et al. [18] examined the convective heat transfer be-
88 haviour of Al2O3 nanoparticles dispersed in water based in a circular
ρnf ¼ φρp þ ð1−φÞρbf ð4Þ
89 tube under a laminar region. The convective heat transfer coefficient 146
90 can be predicted by means of the traditional correlations and observed
91 for positive heat transfer enhancements compared to the base fluid.
92 Such studies by Duangthongsuk and Wongwises [19] used a horizontal Table 1 t1:1
Properties of TiO2 and Al2O3 nanomaterials. t1:2
93 double-tube counter flow heat exchanger with TiO2 nanofluid of 21 nm
94 and a concentration of 0.2%. The base fluid used was water and tested at Nanoparticles TiO2 Al2O3 t1:3
95 turbulent region where the Reynolds number ranges from 4000 to Size, dp (nm) 50 13 t1:4
96 18,000. The heat transfer coefficient of nanofluids in the study where Density, ρp(kg/m3) [14] 4230 4000 t1:5
97 nanofluid was present was about 6%–11% higher than the base liquid. Thermal conductivity, kp (W/m.K) [17] 8.4 36 t1:6
98 The heat transfer coefficient also increased with an increase in the Volume concentration, ϕ (%) 0.5, 0.7 and 1.0 0.6, 0.8 and 1.0 t1:7
pH 6.70–6.93 5.55–5.85 t1:8
99 mass flow rate. Both researches by Bhanvase et al. [20] and Reddy and

Please cite this article as: W.H. Azmi, et al., Heat transfer and friction factor of water and ethylene glycol mixture based TiO2 and Al2O3 nanofluids
under turbulent flow, Int. Commun. Heat Mass Transf. (2016), http://dx.doi.org/10.1016/j.icheatmasstransfer.2016.05.010
W.H. Azmi et al. / International Communications in Heat and Mass Transfer xxx (2016) xxx–xxx 3

147 ð1−φÞðρCÞbf þ φðρCÞp insulated with ceramic fibre insulators. A thermocouple is placed at 154
C nf ¼ ð5Þ each inlet and outlet of the test section. Five thermocouples were 155
ð1−φÞρbf þ φρp
welded to the body of the tube wall at 0.25, 0.5, 0.75, 1.0, and 1.25 m 156
149 from the inlet. A constant heat flux of 7955 W/m2 was supplied by 157
voltage regulators at 600 W input power. 158
2.2. Forced convection heat transfer apparatus The setup included a chiller to cool the working fluid after being 159
heated. The thermocouples and pressure transducers were connected 160
150 The forced convection heat transfer experiment was conducted to the data acquisition system. The experimental setup and its schemat- 161
151 using the setup that consists of a main test section with an inner ic diagram are shown in Fig. 1. The nanofluid was placed in the 162
152 diameter of 16 mm and outer diameter of 19 mm. The test section collecting tank and circulated by the 0.5 hp pump to the entire test rig 163
153 was installed with two nichrome heaters each of 1500 W rating and through the piping system. The nanofluid then entered the tube test 164

F
O
O
R
P
D
E
T
C
E
R
R
O
C
N
U

Fig. 1. Forced convection heat transfer test rig.

Please cite this article as: W.H. Azmi, et al., Heat transfer and friction factor of water and ethylene glycol mixture based TiO2 and Al2O3 nanofluids
under turbulent flow, Int. Commun. Heat Mass Transf. (2016), http://dx.doi.org/10.1016/j.icheatmasstransfer.2016.05.010
4 W.H. Azmi et al. / International Communications in Heat and Mass Transfer xxx (2016) xxx–xxx

165 section. The outlet of the test section was connected to a chiller to cool Chon et al. [32] where the smaller particle size affects the thermal 228
166 down the nanofluid at a desired bulk temperature before it entered conductivity of the nanofluids as a result of the increasing Brownian 229
167 the inlet. The flow rate was controlled using a bypass regulator, placed motion. The Al2O3 nanoparticles being the smaller sized compared to 230
168 at various locations. TiO2 resulted in higher surface area to volume ratio hence, the increase 231
169 All thermocouples recorded the data of inlet, outlet and five surface in thermal conductivity [33]. In addition, the thermal conductivity of 232
170 temperatures. The differential pressure transducer of 0.5 psi recorded Al2O3 nanoparticles is also naturally higher than TiO2 nanoparticles 233
171 the pressure drop. All pressure and temperature data were recorded according to Pak and Cho [17] as presented in Table 1. Therefore, the 234
172 by the ADAMView Advantech Data Acquisition and preceded with anal- nanofluid solution with high thermal conductivity of nanoparticles 235
173 ysis. The setup was used in previous experimental evaluations [14,23, exhibits higher effective thermal conductivity. 236
174 24] and modified accordingly. The total length of the fluid flow in the Fig. 4 shows the variation of dynamic viscosity with temperature for 237
175 tube was 4 m, which ensures a fully developed turbulent flow at the TiO2 and Al2O3 nanofluids. Al2O3 nanofluids exhibit higher relative vis- 238
176 entry of the test section. The requirements to ensure that the flow of cosity as compared to TiO2 for the range of concentrations studied. 239
177 working fluid is turbulent are based on the equations of Lh ≈ 10D [25] The average enhancement in viscosity for TiO2 nanofluids was found 240
178 and L/D ≥ 60 [26]. The range of uncertainties in the measuring to be 12.6% which is 3.6% smaller than Al2O3 nanofluids. The viscosity 241

F
179 instrumentations was from 0.01% to 0.73% as shown in Appendix 1. ratio of TiO2 nanofluids was almost independent of the temperature. 242
180 The forced convection experiments were undertaken with base fluid However for Al2O3, there was a fluctuation along the temperature 243

O
181 operating at 30, 50 and 70 °C to determine the heat transfer coefficient range with higher enhancements recorded at low temperatures of 244
182 and pressure drop for flow rate from 2 to 20 LPM. The temperature and 30 °C. The viscosity enhancements of the two nanofluids increased 245
183 pressure drop were recorded under steady state condition. The determi- with concentration. The fluctuation of the relative viscosity in the 246

O
184 nation of heat transfer coefficients and friction factors was according to range of the temperature studied is possibly related to the difference 247
185 the Newton's law of cooling and Darcy equation, respectively. The ex- in the structure and thickness of the diffused fluid layers around the 248
186 perimental data for water/EG (60:40) mixture was compared with nanoparticles in the base fluids, which affects the effective volume con- 249

R
187 equations in literature to ensure the reliability of the test rig and the ex- centration, and ultimately the viscosity of the suspension [34]. Interest- 250
188 perimental results. The experiments using TiO2 and Al2O3 nanofluids ingly, Al2O3 nanofluids exhibit high effective thermal conductivity but 251

P
189 were conducted for volume concentrations of less than 1.0% at a wide on the other hand, penalized with high relative viscosity. Contradictory 252
190 range of flow rates and operating temperatures of 30, 50 and 70 °C for to TiO2 nanofluids, which is low in effective thermal conductivity but 253
191 the evaluation of heat transfer coefficients and friction factors. The max- also low in relative viscosity. This behaviour will influence the overall
D 254
192 imum and minimum error in the experimental data is presented in heat transfer performance of TiO2 and Al2O3 nanofluids. Consequently, 255
193 Appendix 2. The stability and dispersion of nanofluids were ensured further investigations on the forced convection heat transfer were 256
194 through pH measurements and FESEM or TEM imaging techniques, conducted to compare the heat transfer and friction factors between 257
E
195 respectively. TiO2 and Al2O3 nanofluids at various working temperatures. 258
T

196 3. Results and discussion 3.2. Forced convection heat transfer 259
C

197 3.1. Dispersion and properties Eqs. (6) to (12) are considered in the present heat transfer analysis. 260
The rate of heat transfer, Q to fluid flowing in a tube is expressed as 261
198 The stability of TiO2 and Al2O3 nanofluids was observed through Eq. (6) based on Newton's law of cooling. The input power, Q from the 262
E

199 pH and micrograph evaluation by FESEM and TEM. The pH of TiO2 heater is supplied from the electrical power, expressed as Eq. (7). For 263
200 nanofluids for concentrations of less than 1.0% is found to be 6.70 to energy balance, the heat transfer from the heater is equal to the heat 264
R

201 6.93 which is close to neutral. The results were in good agreement transfer into the fluid flow, with the assumption of no heat loss, 265
202 with literatures such as Murshed et al. [27], Duangthongsuk and expressed by Eq. (8). Hence, the heat transfer coefficient, h is derived 266
203 Wongwises [28], Chakraborty et al. [29] and Azmi et al. [14] for the from Eq. (9). 267
R

204 same type of particle and range of concentration. The Al2O3 nanofluid
205 however is more acidic with pH of 5.55 to 5.83 and found agreeable Q ¼ hAðT s −T b Þ ð6Þ 269
O

206 with Chandrasekar et al. [30]. FESEM images of TiO2 and Al2O3 showed 270
207 that the nanopowder was in clusters under atmospheric conditions. The _ p ΔT
Q ¼ V t I ¼ mC ð7Þ
208 TEM image showed that the particles dispersed into the liquid medium 272
C

209 and ensured a good nanofluid suspension. Both FESEM and TEM images Heat from tube = Heat in fluid flow
210 are shown in Fig. 2.
_ p ΔT
Q ¼ hAðT s −T b Þ ¼ mC ð8Þ 274
N

211 The variation of thermal conductivity with temperature for TiO2 and
212 Al2O3 nanofluids in water/EG mixture is shown in Fig. 3. Based on the 275
213 figure, there is a considerable increase in the thermal conductivity en- Q
h exp ¼ ð9Þ
U

214 hancement from 30 to 70 °C for Al2O3 nanofluids. The Al2O3 nanofluids AΔT
277
215 for all volume concentrations provided enhancements of more than 10%
The dimensionless parameters, Reynolds number, Nusselt num-
216 for temperatures of higher than 40 °C. The thermal conductivity en-
ber, and Dittus–Boelter [26] for Nusselt number, are presented by 278
217 hancement for Al2O3 nanofluids was also discovered by Das et al. [31]
Eqs. (10) to (12) [25]. 279
218 with concentrations of 1.0% and temperature ranges of 21 to 51 °C.
Reynolds number, Re is given by Eq. (10). 280
219 However, the TiO2 nanofluids only recorded enhancements of less
220 than 5% and the ratios that are nearly independent to the temperature.
ρvd
221 The enhancement variations of the two types of nanofluids became Re ¼ ð10Þ
μ
222 more noticeable when the temperature exceeds 40 °C. However, the
282
223 thermal conductivity enhancements show increment with volume con-
Nusselt number, Nu is given by Eq. (11).
224 centrations for both types of nanofluids. The difference in enhancement
225 values between the two sets of data is possibly due to the particle size
226 difference as given in Table 1. The Al2O3 nanoparticle size is much small- hd
Nu exp ¼ ð11Þ
227 er than TiO2 nanoparticles. This is in agreement with the findings by k 284

Please cite this article as: W.H. Azmi, et al., Heat transfer and friction factor of water and ethylene glycol mixture based TiO2 and Al2O3 nanofluids
under turbulent flow, Int. Commun. Heat Mass Transf. (2016), http://dx.doi.org/10.1016/j.icheatmasstransfer.2016.05.010
W.H. Azmi et al. / International Communications in Heat and Mass Transfer xxx (2016) xxx–xxx 5

F
TiO2 at 200,000 magnifications Al2O3 at 250,000 magnifications
(a) FESEM images of TiO2 and Al2O3

O
O
R
P
D
E

TiO2 at 140,000 magnifications and 1.0% volume concentration


T

(b) TEM images of TiO2


C

Fig. 2. Micrograph images of nanoparticles (TiO2 and Al2O3) and nanofluid (TiO2).
E

Dittus–Boelter [26] equation which is applicable for Re N 104 and shown in Fig. 5. The reliability of the experimental setup was confirmed 289
R

285 0.6 b Pr b 200 is presented as Eq. (12). with the estimated values from the Dittus–Boelter [26] equation as in 290
Eq. (12) and experimental data of water/EG. The experimental values 291
NuDB ¼ 0:023Re0:8 Pr0:4 ð12Þ of water/EG at three bulk temperatures showed good agreement with 292
R

287 the Dittus–Boelter relation. The maximum deviation between the 293
Validation of the experimental Nusselt number data was conducted experimental values and Dittus–Boelter were 5.8%, 3.8% and 2.9%, 294
O

288 for base fluid (W/EG) at three bulk temperatures of 30, 50 and 70 °C and respectively for 30, 50 and 70 °C. The validation test for Nusselt number 295
C

1.25
TiO2 Al 2 O3 2.4
φ [%] φ [%] TiO2 Al2O3
φ [%] φ [%]
N

0.5 0.6
1.20 2.2
0.7 0.8 0.5 0.6
Thermal conductivity ratio, kr

1.0 1.0 0.7 0.8


Dynamic viscosity ratio, μ r

2.0
U

1.0 1.0
1.15
1.8

1.10
1.6

1.05 1.4

1.2
1.00
20 30 40 50 60 70 80 1.0
o 20 30 40 50 60 70 80
Temperature, T [ C]
Temperature, T [oC]
Fig. 3. Variation of thermal conductivity ratio of TiO2 and Al2O3 nanofluids with
temperature. Fig. 4. Variation of viscosity ratio of TiO2 and Al2O3 nanofluids with temperature.

Please cite this article as: W.H. Azmi, et al., Heat transfer and friction factor of water and ethylene glycol mixture based TiO2 and Al2O3 nanofluids
under turbulent flow, Int. Commun. Heat Mass Transf. (2016), http://dx.doi.org/10.1016/j.icheatmasstransfer.2016.05.010
6 W.H. Azmi et al. / International Communications in Heat and Mass Transfer xxx (2016) xxx–xxx

240 3500
W/EG at 30oC 1
W/EG at 50oC 3000
200 2
W/EG at 70oC

Heat transfer coefficient, h [W/m .K]


2
3
2500
160
Nusselt number, Nu

2000
120

1500 TiO 2 Al 2 O 3
80 φ [%] φ [%]
Dittus-Boelter [26] 0.5 0.6
1000 0.7 0.8
1 30oC
40 1.0 1.0
2 50oC
3 70oC 500
water/EG
0

F
0 4000 8000 12000 16000 20000 24000 28000 0 2000 4000 6000 8000 10000 12000
Reynolds number, Re Reynolds number, Re

O
(a) temperature of 30 oC
Fig. 5. Comparison of experimental Nusselt number W/EG with Dittus–Boelter relation.
5000

O
296 using the Dittus–Boelter relation was also consistent with other 4000

Heat transfer coefficient, h [W/m .K]


297 researchers [14,35,36].

R
2
298 The experimental heat transfer coefficients of water/EG, TiO2 and
299 Al2O3 nanofluids are shown plotted in Fig. 6(a) to (c). The operating 3000

P
300 temperatures for the experiments were 30, 50 and 70 °C; the TiO2 and
301 Al2O3 particle sizes were 50 nm and 13 nm respectively. The heat trans-
302 fer coefficient variation with Reynolds number at the temperature of 2000 TiO 2 Al 2 O 3
φ [%] φ [%]
303 30 °C is shown in Fig. 6(a). The heat transfer coefficient of TiO2
D 0.5 0.6
304 nanofluids was lower than W/EG for volume concentrations of 0.5% to 0.7 0.8
1000
305 1.0%. The cause for this behaviour is complex involving the combination 1.0 1.0
E
306 of enhancements in thermal conductivity and viscosity of the nanofluids
307 [24]. The decrease in heat transfer coefficient is due to the particle mi- water/EG
0
T

308 gration mechanism where the particles tend to concentrate in the 0 4000 8000 12000 16000 20000
309 pipe centre, which resulted in the decrease of the boundary layer thick- Reynolds number, Re
C

310 ness [37]. However, Al2O3 nanofluids show favourable advantages of (b) temperature of 50 oC
311 being higher than W/EG for all concentrations. This is due to the higher
6000
312 enhancement in thermal conductivity and the effect of smaller particle
E

313 size as compared to TiO2 nanofluids. The numerical study by Saha


314 et al. [38] described the reason for such augmentation between TiO2 5000
Heat transfer coefficient, h [W/m .K]
R

315 and Al2O3 water based nanofluids where the aspects such as enhance-
2

316 ment of thermal conductivity, nanoparticle size and shape, Brownian 4000
317 motion of particles, decrease in boundary layer thickness and delay in
R

318 boundary layer growth may contribute to the behaviour. This is a fit 3000
319 with the present study since Saha et al. [38] conducted the study for
TiO 2 Al 2 O 3
O

320 inlet temperature of 20 °C, and particle size of 10 nm was the best choice
2000 φ [%] φ [%]
321 in order to achieve a higher heat transfer rate. 0.5 0.6
322 On the other hand, it is interesting to see that the heat transfer coef- 0.7 0.8
C

323 ficients for both nanofluids were almost similar for operating tempera- 1000 1.0 1.0
324 tures of 50 and 70 °C as shown in Fig. 6(b) and (c). Despite the fact that
water/EG
N

325 Al2O3 nanofluids possessed higher enhancement in thermal conductiv- 0


326 ity than TiO2, this effect is reduced as the heat transfer coefficients of 0 4000 8000 12000 16000 20000 24000
327 both TiO2 and Al2O3 nanofluids at 50 and 70 °C for a concentration of Reynolds number, Re
U

328 1.0% were almost the same values. The possible reason for this behav- (c) temperature of 70 oC
329 iour is due to the fact that as the temperature increased, the nanofluid
330 viscosity decreased and directly affects the fluid internal shear stress Fig. 6. Variation of heat transfer coefficient with Reynolds number of TiO2 and Al2O3
nanofluids at various temperatures.
331 as well as weakened the inter-particle and inter-molecular adhesion
332 forces [39]. With the increase in thermal conductivity and decrease in
333 viscosity, the effect to the heat transfer is almost equal even though
334 the particle size differs. In addition, with higher temperature and higher 3.3. Friction factor 342
335 Reynolds number range for 50 and 70 °C, the intensity of the flow
336 turbulence is higher due to the increase in the collision among particle The Darcy friction factor of nanofluid was evaluated according to 343
337 loading [40] and resulted in the enhancement of heat transfer rates. Eq. (13) with the values recorded by the pressure transducer. 344
338 The similarity for these three data plots is that the heat transfer
ΔP exp
339 coefficient increased with the increase in concentration and Reynolds f exp ¼   2  ð13Þ
340 number for both nanofluids and in agreement with other researchers L ρv
341 [23,35,41]. d 2 346

Please cite this article as: W.H. Azmi, et al., Heat transfer and friction factor of water and ethylene glycol mixture based TiO2 and Al2O3 nanofluids
under turbulent flow, Int. Commun. Heat Mass Transf. (2016), http://dx.doi.org/10.1016/j.icheatmasstransfer.2016.05.010
W.H. Azmi et al. / International Communications in Heat and Mass Transfer xxx (2016) xxx–xxx 7

0.06 increasing temperature. Therefore, the internal friction forces between 359
TiO 2 Al 2 O 3
molecules decreased. Consequently, the flow resistance decreased and 360
φ [%] φ [%]
0.5 0.6 the distribution of nanofluids at all concentrations and temperatures 361
0.05 0.7 0.8 was within reach of each other [43]. Combination of the effects of 362
1.0 1.0 temperature and flow rate (increased Reynolds number) would lead 363
to decrement in viscosity due to the weakening of the link of nanopar- 364
Friction factor, f

ticles and shrinkage of the nanoparticle cluster sizes [44]. These results 365
0.04
were consistent with those of some other researchers [45,46]. 366

4. Conclusions 367
0.03

The thermal conductivity enhancement of Al2O3 nanofluids is 368


Blasius [42]
water/EG higher than TiO2 nanofluids. The ratio of thermal conductivity for TiO2 369
0.02 nanofluids was almost constant with the temperature; however, it in- 370

F
0 4000 8000 12000 16000 20000 24000
creased with the concentration for both nanofluids. The dynamic viscos- 371
Reynolds number, Re ity ratio of Al2O3 nanofluids fluctuated with temperature variation with 372

O
the maximum value at 30 °C. Meanwhile, the TiO2 nanofluids were in- 373
Fig. 7. Friction factor of base fluid, TiO2 and Al2O3 nanofluids at bulk temperature 30, 50
dependent of temperature. The heat transfer coefficients of Al2O3 374
and 70 °C.
375

O
nanofluids were higher than TiO2 nanofluids at operating temperature
of 30 °C. In addition, the heat transfer coefficients of TiO2 nanofluids 376
The experimental values of friction factor for water/EG estimated were lower than the based mixture of water/EG for concentrations up 377

R
347 with Eq. (13) were compared with the values calculated using Blasius to 1.0% at the same working temperature. However, the heat transfer 378
348 [42] in Eq. (14), which is valid for the range of ReN4000. coefficients were observed to be nearly equal at temperatures of 50 379
and 70 °C. The heat transfer coefficient increased with concentration 380

P
0:3164
fB ¼ ð14Þ and was higher than the based mixture. The maximum enhancements 381
Re0:25
of 24.2% and 23.8% were observed for TiO2 and Al2O3 nanofluids, consec- 382
350
utively at a 1.0% concentration and a temperature of 70 °C. The friction 383
Fig. 7 presents the friction factor of base fluid, TiO2 and Al2O3
351 nanofluids. The friction factor decreased with the increase of Reynolds
D
factors for both TiO2 and Al2O3 nanofluids slightly increased with 384
volume concentrations. 385
352 number. It can be seen that nanofluids that were less and equal to
E
353 1.0% in volume concentration in the present study did not influence
354 the friction factor substantially. The friction factor distribution of TiO2 Acknowledgements 386
T

355 and Al2O3 nanofluids was close to the Blasius [42] line for concentra-
356 tions of 0.5% to 1.0%, hence the friction factor insignificantly increases The authors are grateful to the Universiti Malaysia Pahang (UMP) 387
357 with the concentration. The free volume in the nanofluids internal and Automotive Engineering Centre (AEC) for financial supports given 388
C

358 structure increased due to the decreasing viscosity caused by the under RDU1403153 and RDU151411 (RAGS/1/2015/TK0/UMP/03/2). 389
E

390 Appendix 1. Uncertainty of instruments


R

t2:1
R

t2:2 No. Name of instrument Range of instrument Variables measured Least division in measuring instrument Values measured in % uncertainty
experiment

t2:3 Min Max Max Min


O


t2:4 1 Thermocouple 0–300 °C Bulk temperature, Tb UT = 0.1 C 28.55 71.35 0.49535 0.19821
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
UTb ¼
t2:5 0:12 þ 0:12 =0.14142
UT = 0.1∘ C
C

t2:6 2 Thermocouple 0–300 °C Average surface temperature, Tw 30.73 84.63 0.72765 0.26422
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
U T w ¼ 5  0:12 =0.22361
t2:7
t2:8 3 Flow metre 2–30 LPM Volume flow rate, V_ 0.01 1.87 21.15 0.53476 0.04728
N

t2:9 4 Voltage 0–240 V Voltage, Vt 0.01 110.1 110.1 0.00908 0.00908


t2:10 5 Current 0–15 A Current, I 0.01 5.45 5.45 0.18349 0.18349
t2:11 6 Pressure transducer 0–60 mV Pressure drop, ΔP 0.01 0.32 42.26 3.125 0.02366
U

t2:12 0–6894.8 Pa – 54. 44 3272 3.125 0.02366


t2:13 0–1.0 psi – 0.00790 0.47456 3.125 0.02366

391

392 Appendix 2. Uncertainty of physical quantities


393
t3:1

t3:2 No. Heat transfer and friction Maximum uncertainty (%) Minimum uncertainty (%)
parameter
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
t3:3 1 Reynolds number, Re U 2 U 2 U 2 U 2 U 2 U 2
U Re U Re
¼ ð ρρ Þ þ ð vv Þ þ ð μμ Þ ¼ ð ρρ Þ þ ð vv Þ þ ð μμ Þ
Re ¼ ρvd
μ
Re
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Re
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
¼ ð0:1Þ2 þ ð0:53476Þ2 þ ð0:1Þ2 ¼ ð0:1Þ2 þ ð0:04728Þ2 þ ð0:1Þ2
¼ 0:55314% ¼ 0:14912%
(continued on next page)

Please cite this article as: W.H. Azmi, et al., Heat transfer and friction factor of water and ethylene glycol mixture based TiO2 and Al2O3 nanofluids
under turbulent flow, Int. Commun. Heat Mass Transf. (2016), http://dx.doi.org/10.1016/j.icheatmasstransfer.2016.05.010
8 W.H. Azmi et al. / International Communications in Heat and Mass Transfer xxx (2016) xxx–xxx

Appendix 2 (continued)
(continued)

No. Heat transfer and friction Maximum uncertainty (%) Minimum uncertainty (%)
parameter
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
t3:4 2 Heat flux, q Uq U 2 2 Uq U 2 2
Q Vt I ¼ ð VVtt Þ þ ðUI I Þ ¼ ð VVtt Þ þ ðUI I Þ
q¼ ¼ q q
A πdL qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
¼ ð0:00908Þ2 þ ð0:18349Þ2 ¼ ð0:00908Þ2 þ ð0:18349Þ2
¼ 0:18349 ¼ 0:18371%
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
t3:5 3 Heat transfer coefficient, h U 2 U ðT −T Þ 2 U 2 U ðT 2
Uh Uh −T Þ
q
h ¼ ðT w −T h
¼ ð qq Þ
þ ððT ww−T bb ÞÞ h
¼ ð qq Þ þ ððT ww−T bb ÞÞ
bÞ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
U ðT w −T b Þ U ðT w −T b Þ
ðT w −T b Þ ¼ ð0:72765Þ2 þ ð0:49535Þ2 ðT w −T b Þ ¼ ð0:26422Þ2 þ ð0:19821Þ2
¼ 0:88025% ¼ 0:33030%
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Uh
h
¼ ð0:18371Þ2 þ ð0:88025Þ2 Uh
h
¼ ð0:18371Þ2 þ ð0:35269Þ2
¼ 0:89922%
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ 0:397795%
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

F
t3:6 4 Nusselt number, Nu 2 2 2 2
U Nu
Nu ¼ ðUhh Þ þ ðUkk Þ U Nu
Nu ¼ ðUhh Þ þ ðUkk Þ
Nu ¼ hd
k
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

¼ ð0:89922Þ2 þ ð0:1Þ2 ¼ ð0:37795Þ2 þ ð0:1Þ2

O
¼r
0:90476%
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼r
0:39096%
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
t3:7 5 Friction factor, f Uf 2 U 2 U 2 Uf 2 U 2 U 2
¼ ðUPΔP Þ þ ð ρρ Þ þ ð vv Þ ¼ ðUPΔP Þ þ ð ρρ Þ þ ð vv Þ
f ¼ L ΔPρvd f f
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

O
ðdÞð 2
Þ
¼ ð3:125Þ2 þ ð0:1Þ2 þ 2  ð0:53476Þ2 ¼ ð0:02366Þ2 þ ð0:1Þ2 þ 2  ð0:04728Þ2
¼ 3:21676% ¼ 0:12260%
t3:8 6 Thermo-physical properties 0.1 0.1

R
P
394 References containing TiO2 nanoparticles, Chem. Eng. Process. Process Intensif. 82 (2014) 448
123–131. 449
395 [1] H. Masuda, A. Ebata, K. Teramae, N. Hishinuma, Alteration of thermal conductivity [21] M. Chandra Sekhara Reddy, V. Vasudeva Rao, Experimental investigation of heat 450
396 and viscosity of liquid by dispersing ultra fine particles, Netsu Bussei 4 (4) (1993) transfer coefficient and friction factor of ethylene glycol water based TiO2 nanofluid 451
452
397
398
227–233.
[2] J.C. Maxwell, A Treatise on Electricity and Magnetism, Clarendon press, 1881.
Din double pipe heat exchanger with and without helical coil inserts, Int. Commun.
Heat Mass Transfer 50 (2014) 68–76. 453
399 [3] M. Izadi, A. Behzadmehr, D. Jalali-Vahida, Numerical study of developing laminar [22] W. Yu, H. Xie, A review on nanofluids: preparation, stability mechanisms, and 454
400 forced convection of a nanofluid in an annulus, Int. J. Therm. Sci. 48 (11) (2009) applications, J. Nanomater. 2012 (2012) 1. 455
E
401 2119–2129. [23] W.H. Azmi, K.V. Sharma, P.K. Sarma, R. Mamat, S. Anuar, Comparison of convective 456
402 [4] U.S. Choi, Enhancing Thermal Conductivity of Fluids with Nanoparticles, American heat transfer coefficient and friction factor of TiO2 nanofluid flow in a tube with 457
403 Society of Mechanical Engineers (ASME), New York, 1995. twisted tape inserts, Int. J. Therm. Sci. 81 (2014) 84–93. 458
T

404 [5] J.A. Eastman, S.U.S. Choi, S. Li, L.J. Thompson, S. Lee, Enhanced thermal conductivity [24] W.H. Azmi, K.V. Sharma, P.K. Sarma, R. Mamat, S. Anuar, V. Dharma Rao, Experimen- 459
405 through the development of nanofluids, Proc. Symposium Nanophase and Nano- tal determination of turbulent forced convection heat transfer and friction factor 460
406 with SiO2 nanofluid, Exp. Thermal Fluid Sci. 51 (2013) 103–111. 461
C

composite Materials II, Materials Research Society, Boston, MA 1997, pp. 3–11.
407 [6] K.V. Wong, O. De Leon, Applications of nanofluids: current and future, Adv. Mech. [25] Y.A. Cengel, Ghajar, J. Afshin, Heat and Mass Transfer-Fundamental and Applica- 462
408 Eng. 2 (2010) 519659. tions, 4th (SI Units) ed. McGraw Hill, New York, 2011. 463
409 [7] S.K. Das, S.U.S. Choi, W. Yu, T. Pradeep, Nanofluids: Science and Technology, John [26] F.W. Dittus, L.M.K. Boelter, Heat transfer in automobile radiators of the tubular type, 464
E

410 Wiley & Sons, Inc., Hoboken, New Jersey, 2008. University of California Publications on Engineering, 2 1930, pp. 443–461. 465
411 [8] L.S. Sundar, M.K. Singh, A.C.M. Sousa, Thermal conductivity of ethylene glycol and [27] S.M.S. Murshed, K.C. Leong, C. Yang, Enhanced thermal conductivity of TiO2–water 466
412 water mixture based Fe3O4 nanofluid, Int. Commun. Heat Mass Transfer 49 (2013) based nanofluids, Int. J. Therm. Sci. 44 (4) (2005) 367–373. 467
R

413 17–24. [28] W. Duangthongsuk, S. Wongwises, An experimental study on the heat transfer per- 468
414 [9] G. Paul, J. Philip, B. Raj, P.K. Das, I. Manna, Synthesis, characterization, and thermal formance and pressure drop of TiO2–water nanofluids flowing under a turbulent 469
415 property measurement of nano-Al95Zn05 dispersed nanofluid prepared by a two- flow regime, Int. J. Heat Mass Transf. 53 (1–3) (2010) 334–344. 470
R

416 step process, Int. J. Heat Mass Transf. 54 (15–16) (2011) 3783–3788. [29] S. Chakraborty, J. Mukherjee, M. Manna, P. Ghosh, S. Das, M.B. Denys, Effect of Ag 471
417 [10] F.S. Javadi, S. Sadeghipour, R. Saidur, G. BoroumandJazi, B. Rahmati, M.M. Elias, M.R. nanoparticle addition and ultrasonic treatment on a stable TiO2 nanofluid, Ultrason. 472
418 Sohel, The effects of nanofluid on thermophysical properties and heat transfer char- Sonochem. 19 (5) (2012) 1044–1050. 473
419 [30] M. Chandrasekar, S. Suresh, A. Chandra Bose, Experimental studies on heat 474
O

acteristics of a plate heat exchanger, Int. Commun. Heat Mass Transfer 44 (2013)
420 58–63. transfer and friction factor characteristics of Al2O3/water nanofluid in a circular 475
421 [11] H. Chen, Y. Ding, A. Lapkin, Rheological behaviour of nanofluids containing tube/ pipe under laminar flow with wire coil inserts, Exp. Thermal Fluid Sci. 34 (2) 476
422 rod-like nanoparticles, Powder Technol. 194 (1–2) (2009) 132–141. (2010) 122–130. 477
C

423 [12] A. Turgut, I. Tavman, M. Chirtoc, H. Schuchmann, C. Sauter, S. Tavman, Thermal [31] S.K. Das, N. Putra, P. Thiesen, W. Roetzel, Temperature dependence of thermal con- 478
424 conductivity and viscosity measurements of water-based TiO2 nanofluids, Int. J. ductivity enhancement for nanofluids, J. Heat Transf. 125 (4) (2003) 567–574. 479
425 Thermophys. 30 (4) (2009) 1213–1226. [32] C.H. Chon, K.D. Kihm, S.P. Lee, S.U.S. Choi, Empirical correlation finding the role of 480
481
N

426 [13] J. Garg, B. Poudel, M. Chiesa, J.B. Gordon, J.J. Ma, J.B. Wang, Z.F. Ren, Y.T. Kang, H. temperature and particle size for nanofluid (Al2O3) thermal conductivity enhance-
427 Ohtani, J. Nanda, G.H. McKinley, G. Chen, Enhanced thermal conductivity and viscos- ment, Appl. Phys. Lett. 87 (15) (2005) 1531071–1531073. 482
428 ity of copper nanoparticles in ethylene glycol nanofluid, J. Appl. Phys. 103 (7) [33] U.S. Choi, Enhancing thermal conductivity of fluids with nanoparticles, in: D.A. 483
429 (2008) 074301–074306. Siginer, H.P. Wang (Eds.), Developments and Applications of Non-Newtonian 484
U

430 [14] W.H. Azmi, K.V. Sharma, P.K. Sarma, R. Mamat, G. Najafi, Heat transfer and friction Flows, American Society of Mechanical Engineers (ASME), New York 1995, 485
431 factor of water based TiO2 and SiO2 nanofluids under turbulent flow in a tube, Int. pp. 99–105. 486
432 Commun. Heat Mass Transfer 59 (2014) 30–38. [34] E.V. Timofeeva, W. Yu, D.M. France, D. Singh, J.L. Routbort, Base fluid and tempera- 487
433 [15] Y. Li, J.e. Zhou, S. Tung, E. Schneider, S. Xi, A review on development of nanofluid ture effects on the heat transfer characteristics of SiC in ethylene glycol/H2O and 488
434 preparation and characterization, Powder Technol. 196 (2) (2009) 89–101. H2O nanofluids, J. Appl. Phys. 109 (1) (2011) (014914–014914-014915). 489
Q2
435 [16] P.K. Namburu, D.P. Kulkarni, D. Misra, D.K. Das, Viscosity of copper oxide nanoparti- [35] D.K. Agarwal, A. Vaidyanathan, S. Sunil Kumar, Investigation on convective heat 490
436 cles dispersed in ethylene glycol and water mixture, Exp. Thermal Fluid Sci. 32 (2) transfer behaviour of kerosene–Al2O3 nanofluid, Appl. Therm. Eng. 84 (2015) 491
437 (2007) 397–402. 64–73. 492
438 [17] B.C. Pak, Y.I. Cho, Hydrodynamic and heat transfer study of dispersed fluids with [36] S. Mossaz, J.-A. Gruss, S. Ferrouillat, J. Skrzypski, D. Getto, O. Poncelet, P. Berne, Ex- 493
439 submicron metallic oxide particles, Exp. Heat Transfer 11 (2) (1998) 151–170. perimental study on the influence of nanoparticle PCM slurry for high temperature 494
440 [18] S. Suresh, M. Chandrasekar, P. Selvakumar, T. Page, Experimental studies on heat on convective heat transfer and energetic performance in a circular tube under im- 495
441 transfer and friction factor characteristics of Al2O3/water nanofluid under laminar posed heat flux, Appl. Therm. Eng. 81 (2015) 388–398. 496
442 flow with spiralled rod inserts, Int. J. Nanoparticles 5 (1) (2012) 37–55. [37] H. Chen, W. Yang, Y. He, Y. Ding, L. Zhang, C. Tan, A.A. Lapkin, D.V. Bavykin, Heat 497
443 [19] W. Duangthongsuk, S. Wongwises, Heat transfer enhancement and pressure drop transfer and flow behaviour of aqueous suspensions of titanate nanotubes 498
444 characteristics of TiO2–water nanofluid in a double-tube counter flow heat exchang- (nanofluids), Powder Technol. 183 (1) (2008) 63–72. 499
445 er, Int. J. Heat Mass Transf. 52 (7–8) (2009) 2059–2067. [38] G. Saha, M.C. Paul, Numerical analysis of the heat transfer behaviour of water based 500
446 [20] B.A. Bhanvase, M.R. Sarode, L.A. Putterwar, A. K.A, M.P. Deosarkar, S.H. Sonawane, In- Al2O3 and TiO2 nanofluids in a circular pipe under the turbulent flow condition, Int. 501
447 tensification of convective heat transfer in water/ethylene glycol based nanofluids Commun. Heat Mass Transfer 56 (2014) 96–108. 502

Please cite this article as: W.H. Azmi, et al., Heat transfer and friction factor of water and ethylene glycol mixture based TiO2 and Al2O3 nanofluids
under turbulent flow, Int. Commun. Heat Mass Transf. (2016), http://dx.doi.org/10.1016/j.icheatmasstransfer.2016.05.010
W.H. Azmi et al. / International Communications in Heat and Mass Transfer xxx (2016) xxx–xxx 9

503 [39] C.T. Nguyen, F. Desgranges, G. Roy, N. Galanis, T. Maré, S. Boucher, H. Angue Mintsa, [43] T.G. Mezger, The Rheology Handbook, third ed. Vincentz Network, Hanover, 515
504 Temperature and particle-size dependent viscosity data for water-based nanofluids— Germany, 2011. 516
505 hysteresis phenomenon, Int. J. Heat Fluid Flow 28 (6) (2007) 1492–1506. [44] P. Samira, Z. Saeed, S. Motahare, K. Mostafa, Pressure drop and thermal performance 517
506 [40] M. Hemmat Esfe, S. Saedodin, O. Mahian, S. Wongwises, Thermophysical properties, of CuO/ethylene glycol (60%)–water (40%) nanofluid in car radiator, Korean J. Chem. 518
507 heat transfer and pressure drop of COOH-functionalized multi walled carbon nano- Eng. 32 (4) (2015) 609–616. 519
508 tubes/water nanofluids, Int. Commun. Heat Mass Transfer 58 (2014) 176–183. [45] J. Lee, I. Mudawar, Assessment of the effectiveness of nanofluids for single-phase 520
509 [41] R.S. Vajjha, D.K. Das, D.P. Kulkarni, Development of new correlations for convective and two-phase heat transfer in micro-channels, Int. J. Heat Mass Transf. 50 (3–4) 521
510 heat transfer and friction factor in turbulent regime for nanofluids, Int. J. Heat Mass (2007) 452–463. 522
511 Transf. 53 (21–22) (2010) 4607–4618. [46] K.V. Sharma, L.S. Sundar, P.K. Sarma, Estimation of heat transfer coefficient and fric- 523
512 [42] H. Blasius, Das Aehnlichkeitsgesetz bei Reibungsvorgängen in Flüssigkeiten, tion factor in the transition flow with low volume concentration of Al2O3 nanofluid 524
513 Mitteilungen über Forschungsarbeiten auf dem Gebiete des Ingenieurwesens, 131, flowing in a circular tube and with twisted tape insert, Int. Commun. Heat Mass 525
514 1913. Transfer 36 (5) (2009) 503–507. 526

F
O
O
R
P
D
E
T
C
E
R
R
O
C
N
U

Please cite this article as: W.H. Azmi, et al., Heat transfer and friction factor of water and ethylene glycol mixture based TiO2 and Al2O3 nanofluids
under turbulent flow, Int. Commun. Heat Mass Transf. (2016), http://dx.doi.org/10.1016/j.icheatmasstransfer.2016.05.010

You might also like