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

c Pleiades Publishing, Ltd., 2017.

ISSN 0021-8944, Journal of Applied Mechanics and Technical Physics, 2017, Vol. 58, No. 1, pp. 103–115. 
c A. Ghafouri, M. Salari, A.F. Jozaei.
Original Russian Text 

EFFECT OF VARIABLE THERMAL CONDUCTIVITY MODELS


ON THE COMBINED CONVECTION HEAT TRANSFER
IN A SQUARE ENCLOSURE FILLED WITH
A WATER–ALUMINA NANOFLUID

A. Ghafouri, M. Salari, and A. F. Jozaei UDC 536.4

Abstract: In this numerical study, the effects of variable thermal conductivity models on the
combined convection heat transfer in a two-dimensional lid-driven square enclosure are investigated.
The fluid in the square enclosure is a water-based nanofluid containing alumina nanoparticles. The
top and bottom horizontal walls are insulated, while the vertical walls are kept at different constant
temperatures. Five different thermal conductivity models are used to evaluate the effects of various
parameters, such as the nanofluid bulk temperature, nanoparticle size, nanoparticle volume fraction,
Brownian motion, interfacial layer thickness, etc. The governing stream–vorticity equations are
solved by using a second-order central finite difference scheme coupled with the conservation of mass
and energy. It is found that higher heat transfer is predicted when the effects of the nanoparticle
size and bulk temperature of the nanofluid are taken into account.
Keywords: nanofluid, heat transfer enhancement, square enclosure, thermal conductivity model.
DOI: 10.1134/S0021894417010126

INTRODUCTION

Neoteric fondness in nanofluids emanates from the work of Choi [1] and Eastman et al. [2], who reported
large enhancements in the thermal conductivity of traditional heat transfer fluids when small amounts of metallic
and other nanoparticles were dispersed in these fluids. Nanofluids can be used in various engineering applications,
such as solar collectors, heat exchangers, materials processing, cooling of electronic devices, crystal growth, float
glass production, metal coating and casting, and nuclear system cooling [3–5]. Because of extremely complicated
behavior of nanofluids, several experimental correlations and theoretical models have been developed to estimate the
thermophysical properties of nanofluids. These models are based on the nanoparticle size, volume concentration of
nanoparticles in base fluids, nanofluid temperature, Brownian motion, nanoparticle shape, and interaction between
the pure fluid and nanoparticles [6–8].
There have been many investigations in the past decade on the natural and combined convective heat
transfer and fluid flow in a lid-driven square enclosure. Tiwari and Das [9] numerically investigated the heat
transfer augmentation in a lid-driven cavity filled with a nanofluid. They found that the presence of nanoparticles
in the base fluid is capable of increasing the heat transfer capacity of the base fluid. Oztop and Abu-Nada [10]
considered the buoyancy-driven heat transfer in a two-dimensional chamber with different aspect ratios and filled

Department of Mechanical Engineering, Ahvaz Branch, Islamic Azad University, Ahvaz, Iran;
a.ghafouri@iauahvaz.ac.ir; mehdi salari@iauahvaz.ac.ir; falavand@iauahvaz.ac.ir. Translated from Prikladnaya
Mekhanika i Tekhnicheskaya Fizika, Vol. 58, No. 1, pp. 117–131, January–February, 2017. Original article sub-
mitted May 5, 2014; revision submitted December 29, 2014.
0021-8944/17/5801-0103 
c 2017 by Pleiades Publishing, Ltd. 103
L
Um
äT =0
äy

Th Tc L
g

Um x
äT =0
äy

Fig. 1. Physical model and coordinate system.

with different types of nanoparticles for the Rayleigh (Ra) numbers 103 to 5 · 105 using the finite volume method.
They concluded that the existence of nanoparticles resulted in an increase in the rate of heat transfer and average
Nusselt number for the whole range of Rayleigh numbers. Chamkha and Abu-Nada [11] conducted a study of a
steady laminar mixed convection flow in single- and double-lid square cavities filled with a water–alumina nanofluid
using two viscosity models. Their results showed that significant heat transfer enhancement could be obtained
by increasing the nanoparticle volume fraction at moderate and large Richardson numbers using both nanofluid
models. Sheikhzadeh et al. [12] performed a numerical study of the heat transfer performance of nanofluids inside
two-dimensional rectangular enclosures. Their results indicated that increasing the volume fraction of nanoparticles
produced significant enhancement of the average rate of heat transfer. Their results also showed that the variation
of the average Nusselt number with increasing volume fraction of nanoparticles is linear in two studied cases.
Oztop et al. [13] investigated a laminar natural convection flow through a square inclined enclosure filled with a
CuO nanofluid. They found that the rate of heat transfer in the cavity increases due to addition of nanoparticles.
Moreover, the rate of this increase is greater for enclosures with low Rayleigh numbers. Ghafouri and Salari [14]
numerically studied the laminar mixed convection in a lid-driven square enclosure filled with CuO nanoparticles
using eight different nanofluid viscosity models. They found that the rate of heat transfer is accentuated moderately
by decreasing the Richardson number and rising the solid volume fraction.
Something that is common in most of the cited numerical studies on nanofluid combined convection heat
transfer problems is the use of the Maxwell-Garnett thermal conductivity model [15] for the nanofluid, which does
not consider the main mechanisms for heat transfer in nanofluids, such as the Brownian motion, and does not
consider the nanoparticle size or the temperature dependence. A number of researchers performed comparative
studies of variable thermal conductivity models in various configurations. Abu-Nada [16] investigated the effects of
the thermal conductivity of the Al2 O3 –water nanofluid on heat transfer enhancement in natural convection. Two
different thermal conductivity models, namely, those of Chon et al. [17] and Maxwell-Garnett [15], were evaluated
by comparing their predicted results for the Nusselt number. At Ra  104 , the difference in the Nusselt numbers
predicted by both models was fairly small. However, there was a deviation in the predictions at Ra = 103 , and this
deviation became more significant at high volume fractions of nanoparticles. These two models were also evaluated
by Sebdani et al. [18] for the mixed convection heat transfer in a cavity. The vertical cold side walls were moving
downward and the heat source was located on the bottom wall. They concluded that significant differences existed
between the calculated overall heat transfer calculated by two different formulas. Moreover, the difference increased
with an increase in the nanoparticle volume fraction. Similar comparative investigations were reported by Abu-Nada
et al. [19], Terekhov et al. [20], Sheikhzadeh et al. [21], Parvin et al. [22], and Pourmahmoud et al. [23, 24].
The objective of this study is to compare numerically the thermal conductivity assessor models of the water–
alumina nanofluid and their effects on natural, combined, and forced convection heat transfer in a square enclosure.
The experimental data reported by Das et al. [25] are used to evaluate the results obtained.
104
Table 1. Physical properties of the base fluid and nanoparticles [16]

Nanofluid component cp , J/(kg · K) ρ, kg/m3 k, W/(m · K) β · 10−5 , K−1


Water 4179 997.1 0.613 21.00
Al2 O3 nanoparticles 765 3970.0 25.000 0.85

Here cp is the heat capacitance at constant pressure, ρ is the density, k is the thermal conductivity, and β is
the thermal expansion coefficient.

1. MODEL DESCRIPTION AND GOVERNING EQUATION

Figure 1 shows a schematic view of a two-dimensional square enclosure with a height and width of L
considered in this study. The square enclosure is filled with a suspension of alumina nanoparticles in water. The
top wall is moving rightwards with a uniform velocity Um , while the bottom wall is moving leftwards with the
same velocity. The left boundary is heated and maintained at a constant temperature Th higher than the right cold
boundary temperature Tc , whereas the horizontal walls are adiabatic. The nanofluid in the enclosure is considered
to be an incompressible Newtonian fluid, and the flow is laminar. The physical properties of pure water and
nanoparticles at a temperature of 25◦ C are assumed to be constant (Table 1) [16], whereas the density variation in
the buoyancy force term is handled by the Boussinesq approximation.
The steady-state equations that govern the conservation of mass, momentum, and energy can be written in
the dimensionless form as
∂U ∂V
+ = 0,
∂X ∂Y

∂U ∂U ∂P 1 μnf  ∂ 2 U ∂2U 
U +V =− + + ,
∂X ∂Y ∂X Re ρnf νf ∂X 2 ∂Y 2

∂V ∂V ∂P μnf 1  ∂ 2 V ∂ 2 V  (ρβ)nf
U +V =− + + + Ri θ,
∂X ∂Y ∂Y ρnf νf Re ∂X 2 ∂Y 2 ρnf βf

∂θ ∂θ αnf 1  ∂2θ ∂2θ 


U +V = + ,
∂X ∂Y αf Re Pr ∂X 2 ∂Y 2
where μ is the dynamic viscosity, ν is the kinematic viscosity, α is the thermal diffusivity of the fluid, and
x y u v p T − Tc
X= , Y = , U= , V = , P = 2
, θ= ;
L L Um Um ρnf Um Th − Tc
the subscripts nf and f refer to the nanofluid and fluid, respectively. The Reynolds number (Re), Prandtl num-
ber (Pr ), Richardson Number (Ri), and Rayleigh number (Ra) are defined as
Um L νf Ra gβf (Th − Tc )L3
Re = , Pr = , Ri = , Ra = .
νf αf Pr Re2 αf νf
In terms of the stream function–vorticity (Ψ–Ω) formulation, the governing equations are written as
∂2Ψ ∂2Ψ
+ = −Ω; (1)
∂X 2 ∂Y 2

∂Ω ∂Ω μnf /μf 1  ∂2Ω ∂2Ω   βs  Ra ∂θ


U +V = 2
+ 2
+ 1 − ϕ + ϕ 2 ; (2)
∂X ∂Y 1 − ϕ + ϕρs /ρf Re ∂X ∂Y βf Pr Re ∂X

∂θ ∂θ knf /kf 1  ∂2θ ∂2θ 


U +V = + , (3)
∂X ∂Y 1 − ϕ + ϕ(ρcp )s /(ρcp )f Re Pr ∂X 2 ∂Y 2
where ϕ is the nanoparticle volume fraction, and
ψ ωL ∂Ψ ∂Ψ
Ψ= , Ω= , U= , V =− .
Um L Um ∂Y ∂X
105
The appropriate dimensionless boundary conditions can be written as
∂2Ψ
U = V = Ψ = 0, θ = 1, Ω=− (4)
∂X 2
on the left wall,
∂2Ψ
U = V = Ψ = 0, θ = 0, Ω=− (5)
∂X 2
on the right wall,
∂θ ∂2Ψ
V = Ψ = 0, U = 1, = 0, Ω=− (6)
∂Y ∂Y 2
on the top wall, and
∂θ ∂2Ψ
U = 0, V = Ψ = 0, = 0, Ω=− (7)
∂Y ∂Y 2
on the bottom wall.

2. NANOFLUID MODELING

There are several theoretical formulas and experimental correlations that can be used to estimate the ther-
mophysical properties of nanofluids. In Eqs. (2) and (3), the effective density, heat capacitance, thermal diffusivity,
and thermal expansion coefficient of the nanofluid are borrowed from [26]:
ρnf = (1 − ϕ)ρf + ϕρs , (ρcp )nf = (1 − ϕ)(ρcp )f + ϕ(ρcp )s ,

αnf = knf /(ρcp )nf , (ρβ)nf = (1 − ϕ)(ρβ)f + ϕ(ρβ)s .


The effective dynamic viscosity of the Al2 O3 –water nanofluid is estimated by the correlation proposed by
Abu-Nada et al. [19] based on the detailed experimental results reported by Nguyen et al. [27]:
μAl2 O3 = exp (3.003 − 0.042 03 T − 0.5445ϕ + 0.000 255 3 T 2 + 0.0524ϕ2 − 1.622ϕ−1).
The viscosity in this equation is expressed in centipoise and the temperature in centigrade degrees. The viscosity
of the base fluid (water) is assumed to vary with temperature as
μf = 2.414 · 10−5 · 10247.8/(T −140) .
The effective thermal conductivity knf is determined by the Maxwell-Garnett model as
knf ks + 2kf − 2ϕ(kf − ks )
= , (8)
kf ks + 2kf + ϕ(kf − ks )
which implies that the effective thermal conductivity of nanofluids relies on the thermal conductivity of spherical
particles, the base fluid, and the volume fraction of solid particles.
Koo and Kleinstreuer [28] in their theoretical model considered the effects of the particle size, particle
volume fraction, and temperature dependence, as well as the properties of the base fluid and the Brownian motion
of particles. The corresponding expression for the effective thermal conductivity is

knf = kst + 5 · 104 βϕρs cp kB T /(ρs ds ) f (T, ϕ),
where β = 0.0017(100ϕ)−0.0841, f (T, ϕ) = (−6.04ϕ + 0.4705)T + 1722.3ϕ − 134.63, ds is the solid particle diameter,
kB is the Boltzmann constant, and kst is calculated by Eq. (8). This equation is valid for temperatures in the range
300 K < T < 325 K and solid volume fractions in the range ϕ = 0.01–0.04.
Xie et al. [29] considered an interfacial nanolayer with a linear thermal conductivity distribution and proposed
an effective thermal conductivity model to account for the effects of the nanolayer thickness, nanoparticle size,
nanoparticle volume fraction, and thermal conductivities of the fluid, nanoparticles, and nanolayer. The nanolayer
thickness is considered as 1 nm for the purpose of calculation in this paper. Thus, the formula for the effective
thermal conductivity is
knf 3Θ2 ϕ2T
= 1 + 3ΘϕT + ,
kf 1 − ΘϕT
106
where
Blf [(1 + γ)3 − Bsl /Bf l ] ks − kl kf − kl kl − kf
Θ= , Bsl = , Bf l = , Blf = ,
(1 + γ)3 + 2Bsl Blf ks + 2kl kf + 2kl kl + 2kf

kf M 2 2δ ks
kl = , ϕT = ϕ(1 + γ)3 , γ= , M= (1 + γ) − 1.
(M − γ) ln (1 + M ) + γM ds kf
Chon el al. [17] used the Buckingham π-theorem and the linear regression analysis to develop an empirical
model with R2 = 95% based on experimental results. The formula for the effective thermal conductivity is
knf  d 0.3690  k 0.7476
f s
= 1 + 64.7ϕ0.7640 Pr 0.9955
T Re1.2321
T ,
kf ds kf
where
μf ρf kB T 1
Pr T = , ReT = , lf = √ ,
ρf αf 3πμ2f lf 2m πd2f
kB = 1.3807 · 10−23 J/K is the Boltzmann constant, df is the fluid molecule size (df = 2 Å for water), and
lf = 0.17 nm is the mean path of base fluid particles [17]. This model considers the effect of the nanoparticle
size and temperature on the nanofluid thermal conductivity with a wide range of temperature Ts = 21–70◦ C. The
accuracy of this model was confirmed by the experiments of Angue Minsta et al. [30].
Patel et al. [31] proposed the expression for the effective thermal conductivity of the nanofluid in the form
knf ks As As
=1+ + cks Pe ,
kf kf Af kf Af
where Peclet (Pe) number, the velocity of the Brownian motion us , and the ratio As /Af are defined as
us ds 2kB T As df ϕ
Pe = , us = , = ,
αf πμf d2s Af ds 1 − ϕ
and c is an experimental constant, which is considered as 25 000 for the purpose of calculation [31]. In this model,
the increase in the specific surface area as well as the Brownian motion are assumed to be the most significant
reasons for the anomalous enhancement in the thermal conductivity of nanofluids.
Das et al. [25] examined the effects of the temperature and nanoparticle diameter on thermal conductivity
enhancement for nanofluids containing Al2 O3 particles (38.4 nm) in the temperature range T = 21–51◦ C through
an experimental investigation by using the temperature oscillation method. They observed that an increase in the
thermal conductivity ratio knf /kf equal to 11–24% can take place if the nanoparticle volume fraction increases
from 1% to 4% [24] at T = 51◦ C. In this paper, five thermal conductivity models proposed in [15, 17, 28, 29, 31]
are evaluated by comparing their predicted results against the experimental data of Das et al. [25] on the combined
convection heat transfer in a square enclosure.

3. NUMERICAL IMPLEMENTATION

The classical theory of single-phase fluids can be applied if the physical properties of the nanofluid are taken as
functions of the properties of both constituents and their concentrations. In order to numerically solve the governing
equations for the two-dimensional velocity and temperature fields, Eqs. (1)–(3) with the corresponding boundary
conditions (4)–(7) are approximated by a second-order central difference scheme. The convergence criterion is
defined by the expression
j=M
 i=N
  j=M
 i=N

ε= |ζ n+1 − ζ n | |ζ n+1 |  10−6 ,
j=1 i=1 j=1 i=1

where M and N are the numbers of grid points in the X and Y directions, respectively. The symbol ζ denotes any
scalar transport quantity: Ψ, Ω, or θ. The local and average heat transfer rates of the enclosure can be presented by
means of the local and average Nusselt numbers. The local Nusselt number Nu is calculated along the left heated
107
Nuavg o
9 1.0
1
8 2
0.8 3
1
7
2
0.6
6
3 0.4
5

0.2
4

3
21 41 61 81 101 121 N 0 0.2 0.4 0.6 0.8 1.0 X

Fig. 2. Fig. 3.

Fig. 2. Average Nusselt number versus the number of grid nodes for different values of the Richard-
son number: Ri = 0.1 (1), 1 (2), and 10 (3).
Fig. 3. Temperature distribution on the left wall (Ra = 105 and Pr = 0.7): the solid curve shows
the data of the present work; the other curves represent the data of [10] (1), [26] (2), and [33] (3).

wall, and the average Nusselt number Nuavg is determined by integrating the local Nusselt number along the heated
wall:
1
knf ∂θ 
Nu(X) =  , Nuavg = Nu(X) dY.
kf ∂X X=0
0

The normalized average Nusselt number is defined as the ratio of the Nusselt number at any volume fraction of
nanoparticles to that of pure water [32]:
Nuavg (ϕ)
Nu∗avg (ϕ) = .
Nuavg |ϕ=0

4. GRID INDEPENDENCE AND VALIDATION

The grid independence test was performed using six distinct uniform grids, namely, 21 × 21, 41 × 41, 61 × 61,
81 × 81, 101 × 101, and 121 × 121. As it can be observed from Fig. 2, the 81 × 81 uniform grid is sufficiently fine
to ensure a grid-independent solution. Hence, this grid was used to perform all the subsequent calculations.
The present numerical solution was further validated by comparing the present code results for the tem-
perature distribution at Ra = 105 and Pr = 0.7 against the experiment of Krane and Jessee [33] and numerical
simulations of Khanafer et al. [26] and also Oztop and Abu-Nada [10]. It is clear that the present code is in good
agreement with other works reported in the literature (Fig. 3). In addition, the governing equations were also
solved for the natural convection flow in an enclosed chamber filled by a pure fluid in order to compare the results
with those of [21, 26, 34–36]. This comparison revealed good agreement between the results (Table 2).

5. RESULTS AND DISCUSSION

The two-dimensional combined convection is studied numerically for the Al2 O3 –water nanofluid in a square
enclosure. The calculations are accomplished for Ri = 0.1, 1.0, and 10.0, Re = 50, 100, and 150, and ϕ = 0.01,
0.02, 0.03, and 0.04.
108
Y (a) Y (b)
1.0 1.0

0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0 0.2 0.4 0.6 0.8 1.0 X 0 0.2 0.4 0.6 0.8 1.0 X

Y (c) Y (d)
1.0 1.0

0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0 0.2 0.4 0.6 0.8 1.0 X 0 0.2 0.4 0.6 0.8 1.0 X

Y (e) Y (f)
1.0 1.0

0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0 0.2 0.4 0.6 0.8 1.0 X 0 0.2 0.4 0.6 0.8 1.0 X

Fig. 4. Isotherms predicted by the Maxwell-Garnett model [15] (a, c, and e) and by the model of
Chon et al. [17] (b, d, and f) for Ri = 0.1 (a and b), 1 (c and d), and 10 (e and f); the solid and
dashed curves show the data for ϕ = 0 and 4%. respectively.

109
Table 2. Average Nusselt number on the hot wall

Nuavg
Reference
Ra = 103 Ra = 104 Ra = 105 Ra = 106
Present work 1.123 2.246 4.521 8.984
[21] 1.120 2.242 4.514 8.790
[26] 1.118 2.245 4.522 8.826
[34] 1.052 2.302 4.646 9.012
[35] 1.108 2.201 4.430 8.754
[36] 1.118 2.243 4.519 8.799

Figure 4 shows the representative contour maps for isotherms inside a double lid-driven square enclosure
filled with pure water (ϕ = 0) and with the Al2 O3 –water nanofluid with the particle volume fraction ϕ = 4% for
several values of the Richardson number by using the Maxwell-Garnett model [15] and the model of Chon et al.
[17]. For small Richardson numbers (Ri = 0.1, purely forced convection regime), the buoyancy effect is overwhelmed
by the mechanical or shear effect due to the movement of the top and bottom lids (see Figs. 4a and 4b). The
isotherms are clustered close to the vertical walls of the enclosure, i.e., the hot fluid moves horizontally toward the
right vertical wall. In this case, oval-shaped isotherm lines were observed in the central region of the enclosure.
For moderate values of the Richardson number (Ri = 1, combined convection dominated regime), the buoyancy
effect is relatively comparable in magnitude to the shear effect due to the sliding walls (see Figs. 4c and 4d);
however, there is no variation in the isotherm patterns. At high Richardson numbers (Ri = 10, natural convection
dominated regime), the solid particle concentration is more effective to increase the heat penetration. Consequently,
the isotherms become more uniformly distributed over the entire enclosure (see Figs. 4e and 4f).
It is interesting to note that the isotherms for Ri = 0.1 are slightly affected by addition of 4% of alumina
nanoparticles in the Maxwell-Garnett model [15], while they are significantly affected in the model of Chon et al.
[17], where the thermal gradient at the right vertical wall increases. A similar feature is observed for Ri = 1.
At higher Richardson numbers (Ri = 10), obviously, the solid concentration has a more pronounced effect on the
isotherms and leads to an increase in the thermal gradient at the right vertical wall. Moreover, this increase is more
significant if the model of Chon et al. [17] is used than the corresponding value predicted by the Maxwell-Garnett
model [15].
Figure 5 illustrates the variation of the average Nusselt number and normalized Nusselt number as functions
of the nanoparticle volume fraction at Re = 100. As is seen from the figure, the average Nusselt number increases
as the nanoparticle volume fraction increases from ϕ = 0 to ϕ = 4%. At ϕ = 2%, the result predicted by the model
of Chon et al. [17] ensures the best agreement with the experimental data of Das et al. [25]. At ϕ = 3%, the
Nusselt numbers predicted by the model of Chon et al. [17] are higher than the results of the experimental data of
Das et al. [25]. At ϕ = 4% and Ri = 0.1, the highest and lowest experimental values of the Nusselt number differ
from the corresponding values predicted by the model of Chon et al. [17] and by the Maxwell-Garnett model [15]
by 13% and 5%, respectively (see Figs. 5a and 5b). In addition, the highest average Nusselt number (Nuavg = 8.1)
is predicted at forced convection heat transfer by the model of Chon et al. [17], which takes into account the role of
the Brownian motion, temperature, and nanoparticles size. Furthermore, Figs. 5c and 5d indicate that the highest
and lowest experimental values of the Nusselt numbers at Ri = 1 differ from the corresponding values predicted by
the model of Chon et al. [17] and by the Maxwell-Garnett model [15] by 13% and 5%, respectively. At Ri = 10,
these percentages reach the maximum value. Figures 5e and 5f indicate that the highest and lowest experimental
values differ from those predicted by the model of Chon et al. [17] and by the Maxwell-Garnett model [15] by 26%
and 18%, respectively. In this case, the experimental data of Das et al. [25] and the calculations by the model of
Koo and Kleinstreuer [28] are the same as those predicted by the model of Chon et al. [17]. The predictions of the
model of Patel et al. [31] are higher by 23%.
For further validation, the average Nusselt numbers on the left wall of the square enclosure predicted by
five studied thermal conductivity models have been compared with the experimental results of Ho et al. [37] in the
natural convection regime with the Rayleigh number varied in the interval Ra = 7.55 · 105 –9.70 · 108 (Table 3). The
results of this comparison show that the results predicted by the models of Koo and Kleinstreuer [28] and Chon et
al. [17] are more consistent with the experimental result of Ho et al. [37] because these models involve the effects
of the nanoparticle size and bulk temperature of the nanofluid.
110
Nuavg (a) Nu*avg (b)
8.4 1 1.17
2
8.2 3 1.14
4
8.0 5
6 1.11
7.8
1.08
7.6

7.4 1.05

7.2 1.02
7.0
0 0.01 0.02 0.03 0.04 f 0 0.01 0.02 0.03 0.04 f
Nuavg (c) Nu*avg (d)

7.0
1.16

6.8
1.12
6.6
1.08
6.4

1.04
6.2

6.0 1.00
0 0.01 0.02 0.03 0.04 f 0 0.01 0.02 0.03 0.04 f
Nuavg (e) Nu*avg (f)

6.75 1.30
6.50
1.25
6.25
1.20
6.00
1.15
5.75
1.10
5.50

5.25 1.05

5.00 1.00
0 0.01 0.02 0.03 0.04 f 0 0.01 0.02 0.03 0.04 f

Fig. 5. Average (a, c, and e) and normalized (b, d, and f) Nusselt numbers versus the nanoparticle
volume fraction as calculated by the Maxwell-Garnett model [15] (1), calculated by the model [28]
(2), measured in [25] (3), calculated by the model [31] (4), calculated by the model [17] (5), and
calculated by the model [29] (6).

111
Nuavg (a) Nu*avg (b)
5.7 1
2 1.30
5.4 3
4
5 1.25
5.1 6
7 1.20
4.8
4.5 1.15

4.2 1.10
3.9 1.05
3.6
1.00
0.1 1.0 10.0 Ri 0.1 1.0 10.0 Ri

Nuavg (c) Nu*avg (d)


8.4
1.30
8.0
7.6 1.25
7.2 1.20
6.8
1.15
6.4
6.0 1.10
5.6 1.05
5.2
1.00
0.1 1.0 10.0 Ri 0.1 1.0 10.0 Ri

Nuavg (e) Nu*avg (f)

10.5
1.30
10.0
9.5 1.25
9.0 1.20
8.5
8.0 1.15
7.5 1.10
7.0
1.05
6.5
6.0 1.00
0.1 1.0 10.0 Ri 0.1 1.0 10.0 Ri

Fig. 6. Average (a, c, and e) and normalized (b, d, and f) Nusselt numbers versus the Richardson
number for Re = 100 as calculated in the present work for the pure fluid (1), calculated by the
model [15] (2), calculated by the model [28] (3), measured in [25] (4), calculated by the model [31]
(5), calculated by the model [17] (6), and calculated by the model [29] (7).

112
Table 3. Average Nusselt numbers predicted by different models and in experiments [37] for ϕ = 0.1 %

Experiment Nuavg
and numerical model Ra = 7.55 · 105 Ra = 1.05 · 107 Ra = 9.7 · 108 δavg , %
Experiment [37] 7.6587 14.8723 29.9072 —
Model [15] 7.3296 14.2410 28.1299 5.22
Model [28] 7.4534 14.7344 29.7715 0.91
Model [29] 7.3739 14.3631 28.4961 4.20
Model [17] 7.6803 15.0611 29.9447 0.47
Model [31] 7.3739 14.3631 28.4961 1.07
Here δavg is the average error.

Nu*avg
1.30
1
2
1.25 3
4
5
6
1.20
7
8
9
1.15

1.10

1.05

0.1 1.0 10.0 Ri


Fig. 7. Normalized Nusselt number versus the Richardson number for different values of the Reynolds
number: the curves show the results for Re = 50 (solid curves), Re = 100 (dot-and-dashed curves), and
Re = 150 (dashed curves); the points are the data calculated by the model [15] (points 1, 4, and 7),
experimental data [25] (2, 5, and 8), and the data predicted by the model [31] (3, 6, and 9).

Figure 6 demonstrates the effect of various nanofluid thermal conductivity models on the average Nusselt
number (Fig. 6a) and normalized Nusselt number (Fig. 6b) as functions of the Richardson number for Re = 50.
It can be seen that the average Nusselt number in this flow regime decreases with increasing Ri. Nonetheless,
the normalized Nusselt number Nu∗avg decreases at the early stage, and then it increases and reaches its peak
value. At Ri = 0.1, the maximum and minimum experimental values of the average Nusselt number differ from the
corresponding predictions of the model of Chon et al. [17] by 5.6% and 5.2%, respectively. Moreover, at Re = 50,
the highest enhancement (26%) of the average Nusselt number is observed in the natural convection heat transfer
regime as predicted by the model of Chon et al. [17].
Figures 6c–6f depict the effect of nanofluid thermal conductivity models on the average Nusselt number and
normalized Nusselt number as functions of the Richardson number for Re = 100 and 150. As can be observed in
these figures, the average Nusselt number decreases with increasing Ri, whereas the normalized Nusselt number
increases. This results again shows that the values predicted by the model of Chon et al. [17] are significantly
higher than the average and normalized Nusselt numbers obtained by using other models and in experiments.
It is observed that the variation of the Reynolds numbers leads to a disparate behavior in different convection
heat transfer regimes (Fig. 7). An interesting point to note is that the normalized Nusselt numbers for Ri = 10
are only slightly affected by the variation of the Reynolds number, whereas they are significantly affected by the
Reynolds number variation at Ri = 0.1.
113
CONCLUSIONS

Laminar combined convection in a two-dimensional square enclosure filled with the Al2 O3 –water nanofluid
was numerically studied. The main findings are listed below.
The Nusselt number and, thus, heat transfer from the left wall of the enclosure is enhanced with increasing
volume fraction of nanoparticles.
The heat transfer rate in the square enclosure is enhanced by decreasing the Richardson number, while
keeping the other parameters constant.
It is diagnosed that the normalized Nusselt number is more sensitive to the Reynolds number and the thermal
conductivity models at low Richardson numbers.
The lowest assess of heat transfer enhancement is provided by the Maxwell-Garnett model [15]. This model
is not suitable to use in future investigations according to the result of the present study.
It is found that the results of using the models [17, 28] ensure better agreement with the experimental data
[25, 37] than the results predicted by other models.

REFERENCES

1. S. U. S. Choi, “Enhancing Thermal Conductivity of Fluids with Nanoparticles,” in Developments and Applied
of Non-Newtonian Flows (ASME, New York, 1995), Vol. 231, pp. 99–105.
2. J. A. Eastman, S. U. S. Choi, S Li, et al., “Anomalously Increased Effective Thermal Conductivities of Ethylene
Glycol-Based Nanofluids Containing Copper Nanoparticles,” Appl. Phys. Lett. 78 (6), 718–720 (2001).
3. S. Kakac and A. Pramuanjaroenkij, “Review of Convective Heat Transfer Enhancement with Nanofluids,” Int.
J. Heat Mass Transfer 52, 3187–3196 (2009).
4. R. Saidur, K. Y. Leong, and H. A. Mohammad, “A Review on Applications and Challenges of Nanofluids,”
Renewable Sustainable Energy Rev. 15, 1646–1668 (2011).
5. K. Khanafer and K. Vafai, “A Critical Synthesis of Thermophysical Characteristics of Nanofluids,” Int. J. Heat
Technol. 54, 4410–4428 (2011).
6. V. I. Terekhov, S. V. Kalinina, and V. V. Lemanov, “The Mechanism of Heat Transfer in Nanofluids: State of
the Art (Review). 1. Synthesis and Properties of Nanofluids,” Thermophys. Aeromech. 17 (1), 1–14 (2010).
7. I. M. Mahbubul, R. Saidur, and M. A. Amalina, “Latest Developments on the Viscosity of Nanofluids,” Int.
J. Heat Mass Transfer 55, 874–885 (2012).
8. P. C. Mukesh Kumar, J. Kumar, and S. Suresh, “Review on Nanofluid Theoretical Viscosity Models,” Int.
J. Eng Innovat. Res. 1 (2), 128–134 (2012).
9. R. K. Tiwari and M. K. Das, “Heat Transfer Augmentation in a Two-Sided Lid-Driven Differentially Heated
Square Cavity Utilizing Nanofluids,” Int. J. Heat Mass Transfer 50, 2002–2018 (2007).
10. H. F. Oztop and E. Abu-Nada, “Numerical Study of Natural Convection in Partially Heated Rectangular
Enclosures Filled with Nanofluids,” Int. J. Heat Fluid Flow 29, 1326–1336 (2008).
11. A. J. Chamkha and E. Abu-Nada, “Mixed Convection Flow in Single and Double-Lid Driven Square Cavities
Filled with Water–Al2 O3 Nanofluid: Effect of Viscosity Models,” Eur. J. Mech., B: Fluids 36, 82–96 (2012).
12. G. A. Sheikhzadeh, N. Hajialigol, M. Ebrahim Qomi, and A. Fattahi, “Laminar Mixed Convection of Cu–Water
Nano-Fluid in Twosided Lid-Driven Enclosures,” J. Nanostructure 1, 44–53 (2012).
13. H. F. Oztop, M. Mobedi, E. Abu-Nada, and I. Pop, “A Heatline Analysis of Natural Convection in a Square
Inclined Enclosure Filled with a CuO Nanofluid under Non-Uniform Wall Heating Condition,” Int. J. Heat
Mass Transfer 55, 5076–5086 (2012).
14. A. Ghafouri and M. Salari, “Numerical Investigation of the Heat Transfer Enhancement using Various Viscosity
Models in Chamber Filled with Water–CuO Nanofluid,” J. Brazilian Soc. Mech. Sci. Eng. 36 (4), 825–836 (2014);
DOI: 10.1007/s40430-013-0091-1.
15. J. C. Maxwell-Garnett, “Colours in Metal Glasses and in Metallic Films,” Philos. Trans. Roy. Soc., London,
Ser. A 203, 385–420 (1904).
16. E. Abu-Nada, “Effects of Variable Viscosity and Thermal Conductivity of Al2 O3 –Water Nanofluid on Heat
Transfer Enhancement in Natural Convection,” Int. J. Heat Fluid Flow 30, 679–690 (2009).
114
17. C. H. Chon, K. D. Kihm, S. P. Lee, and S. U. S. Choi, “Empirical Correlation Finding the Role of Temperature
and Particle Size for Nanofluid (Al2 O3 ) Thermal Conductivity Enhancement,” Appl. Phys. Lett. 87 (15), 153107
(2005).
18. S. M. Sebdani, M. Mahmoodi, and S. M. Hashemi, “Effect of Nanofluid Variable Properties on Mixed Convection
in a Square Cavity,” Int. J. Thermal Sci. 52, 112–126 (2012).
19. E. Abu-Nada, Z. Masoud, H. F. Oztop, and A. Campo, “Effect of Nanofluid Variable Properties on Natural
Convection in Enclosures,” Int. J. Thermal Sci. 49, 479–491 (2010).
20. V. I. Terekhov, S. V. Kalinina, and V. V. Lemanov, “The Mechanism of Heat Transfer in Nanofluids: State of
the Art (Review). 2. Convective Heat Transfer,” Thermophys. Aeromech. 17 (2), 157–171 (2010).
21. G. A. Sheikhzadeh, M. Ebrahim Qomi, N. Hajialigol, and A. Fattahi, “Numerical Study of Mixed Convection
Flows in a Lid-Driven Enclosure Filled with Nanofluid using Variable Properties,” Results Phys. 2, 5–13 (2012).
22. S. Parvin, R. Nasrin, M. A. Alim, et al., “Thermal Conductivity Variation on Natural Convection Flow of
Water–Alumina Nanofluid in an Annulus,” Int. J. Heat Mass Transfer 55, 5268–5274 (2012).
23. N. Pourmahmoud, A. Ghafouri, and I. Mirzaee, “Numerical Study of Mixed Convection Heat Transfer in Lid-
Driven Cavity Using Nanofluid; Effect of Type and Model of Nanofluid,” J. Thermal Sci. 19 (5), 1575–1590
(2015).
24. N. Pourmahmoud, A. Ghafouri, and I. Mirzaee, “Numerical Comparison of Viscosity Models on Mixed Con-
vection in Double Lid-Driven Cavity Utilized CuO–Water Nanofluid,” J. Thermal Sci. 20 (1), 347–358 (2016).
25. S. K. Das, N. Putra, P. Thiesen, and W. Roetzel, “Temperature Dependence of Thermal Conductivity En-
hancement for Nanofluids,” J. Heat Transfer 125, 567–574 (2003).
26. K. Khanafer, K. Vafai, and M. Lightstone, “Buoyancy-Driven Heat Transfer Enhancement in a Two-Dimensional
Enclosure Utilizing Nanofluids,” Int. J. Heat Mass Transfer 46, 3639–3653 (2003).
27. C. T.Nguyen, F. Desgranges, G. Roy, et al., “Temperature and Particle-Size Dependent Viscosity Data for
Water Based Nanofluids-Hysteresis Phenomenon,” Int. J. Heat Fluid Flow 28, 1492–1506 (2007).
28. J. Koo and C. Kleinstreuer, “A New Thermal Conductivity Model for Nanofluids,” J. Nanoparticle Res. 6,
577–588 (2004).
29. H. Xie, M. Fujii, and X. Zhang, “Effect of Interfacial Nanolayer on the Effective Thermal Conductivity of
Nanoparticle-Fluid Mixture,” Int. J. Heat Mass Transfer 48, 2926–2932 (2005).
30. H. Angue Minsta, G. Roy, C. T. Nguyen, and D. Doucet, “New Temperature and Conductivity Data for
Water-Based Nanofluids,” Int. J. Thermal Sci. 48 (2), 363–373 (2008).
31. H. E. Patel, T. Pradeep, T. Sundararajan, et al., “A Micro-Convection Model for Thermal Conductivity of
Nanofluid,” Pramana J. Phys. 65, 863–869 (2005).
32. E. Abu-Nada, “Effects of Variable Viscosity and Thermal Conductivity of CuO–Water Nanofluid on Heat
Transfer Enhancement in Natural Convection: Mathematical Model and Simulation,” J. Heat Transfer 132,
052401 (2010).
33. R. J. Krane and J. Jessee, “Some Detailed Field Measurements for a Natural Convection Flow in a Vertical
Square Enclosure,” in Proc. of the 1st ASME–JSME Thermal Engineering Joint Conf. (Honolulu, 1983), Vol. 1,
pp. 323–329.
34. T. Fusegi, J. M. Hyun, K. Kuwahara, and B. Farouk, “A Numerical Study of Three Dimensional Natural
Convection in a Differentially Heated Cubical Enclosure,” Int. J. Heat Mass Transfer 34, 1543–1557 (1991).
35. N. C. Markatos and K. A. Pericleous, “Laminar and Turbulent Natural Convection in an Enclosed Cavity,” Int.
J. Heat Mass Transfer 27, 772–775 (1984).
36. G. De Vahl Davis, “Natural Convection of Air in a Square Cavity, a Benchmark Numerical Solution,” Int. J.
Numer. Methods Fluids 3, 249–264 (1983).
37. C. J. Ho, W. K. Liu, Y. S. Chang, and C. C. Lin, “Natural Convection Heat Transfer of Alumina–Water
Nanofluid in Vertical Square Enclosures: An Experimental Study,” Int. J. Thermal Sci. 49, 1345–1353 (2010).

115

You might also like