You are on page 1of 8

International Communications in Heat and Mass Transfer 38 (2011) 1428–1435

Contents lists available at ScienceDirect

International Communications in Heat and Mass Transfer


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

Numerical study of mixed convection in an inclined two sided lid driven cavity filled
with nanofluid using two-phase mixture model☆
M. Alinia, D.D. Ganji ⁎, M. Gorji-Bandpy
Babol University of Technology, Department of Mechanical Engineering, Babol, P. O. Box 484, Iran

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

Available online 16 August 2011 Mixed convection of a nanofluid consisting of water and SiO2 in an inclined enclosure cavity has been studied
numerically. The left and right walls are maintained at different constant temperatures while upper and
Keywords: bottom insulated walls are moving lids. Two-phase mixture model has been used to investigate the thermal
Nanofluid behaviors of the nanofluid for various inclination angles of enclosure ranging from θ = − 60° to θ = 60°,
Mixed convection volume fraction from 0% to 8%, Richardson numbers varying from 0.01 to 100 and constant Grashof number
Inclined cavity
10 4. The governing equations are solved numerically using the finite-volume approach. Results are presented
Lid driven
Two-phase mixture model
in the form of streamlines, isotherms, distribution of nanoparticles and average Nusselt number. In addition,
effects of solid volume fraction of nanofluids on the hydrodynamic and thermal characteristics have been
investigated. The results reveal that addition of nanoparticles enhances heat transfer in the cavity remarkably
and causes significant changes in the flow pattern. Besides, effect of inclination angle is more pronounced at
higher Richardson numbers.
© 2011 Elsevier Ltd. All rights reserved.

1. Introduction fluid and the enhanced thermal conductivity of nanofluids. On the other
hand Kelbinski et al. [4] have studied four possible mechanisms that
Convective heat transfer is very important for many industrial contribute to the increase in nanofluid heat transfer: Brownian motion
heating or cooling equipments. The heat convection can passively be of the particles, molecular-level layering of the liquid/particle interface,
enhanced by changing flow geometry, boundary conditions or by heat transport in the nanoparticles and nanoparticles clustering. Wen
enhancing fluid thermophysical properties. One way is by adding small and Ding [5] focused on the entry region under laminar flow condition
solid particles in the fluid. The main idea backs Maxwell's study [1]. He using nanofluids containing γ-Al2O3 nanoparticles of various concen-
showed the possibility of increasing thermal conductivity of a fluid–solid trations. It is shown that the enhancement increases with the Reynolds
mixture by more volume fraction of solid particles. The particles with number as well as the volume concentration of nanoparticle. In a
micrometer or even millimeter dimensions were used. Those particles comparison between particle sizes it was observed that nanofluid with
caused several problems such as abrasion, clogging and pressure smaller particles shows higher heat transfer coefficient than that with
dropping. By improving the technology to make particles in nanometer larger particles [6]. Incorporating a dispersion model similar to that for
dimensions, a new generation of solid–liquid mixtures that is called the flow through porous media, Khanafer et al. [7] presented a two-
nanofluid, was appeared. The nanofluids are a new kind of heat transfer dimensional numerical simulation of natural convection of nanofluids in
fluid containing small quantity of nano-sized particles (usually less than a vertical rectangular enclosure. Ghasemi et al. [8] studied the natural
100 nm) that are uniformly and stably suspended in a liquid. The convection heat transfer in an inclined enclosure filled with a water-cuo
dispersion of a small amount of solid nanoparticles in conventional fluids nanofluid. They indicated that the inclination angle has a significant
changes their thermal conductivity remarkably. Choi [2] quantitatively impact on the flow and temperature fields and the heat transfer
analyzed some potential benefits of nanofluids for augmenting heat performance at high Rayleigh numbers. Ho et al. [9] studied natural
transfer and reducing size, weight and cost of thermal apparatuses, convection of nanofluid in a square enclosure numerically to identify the
while incurring little or no penalty in the pressure drop. Xuan and effects due to uncertainties in effective dynamic viscosity and thermal
Roetzel [3] have identified two causes of improved heat transfer by conductivity. Zhang et al. [10] studied effects of Brownian and
nanofluids: the increased thermal dispersion due to the chaotic thermophoretic diffusions of nanoparticles on nonequilibrium heat
movement of nanoparticles that accelerates energy exchanges in the conduction in a nanofluid layer with periodic heat flux. They showed
that the Brownian and thermophoretic diffusions only affect the
☆ Communicated by W.J. Minkowycz.
nanoparticle temperature, but their effect on the heat transfer
⁎ Corresponding author. enhancement is negligible. Talebi et al. [11] investigated the mixed
E-mail address: ddg_davood@yahoo.com (D.D. Ganji). convection flows in a square lid-driven cavity utilizing nanofluid which

0735-1933/$ – see front matter © 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.icheatmasstransfer.2011.08.003
M. Alinia et al. / International Communications in Heat and Mass Transfer 38 (2011) 1428–1435 1429

Table 1
Nomenclature Thermophysical properties of water and nanoparticles [38].

Cp specific heat (J/kg K) Property Fluid phase (water) Solid phase (SiO2)

dp nanoparticle diameter (nm) cp(J/KgK) 4179 765


g gravitational acceleration (m s − 2) ρ(Kg/m3) 997.1 3970
gβf ΔTH 3 k(W/mK) 0.613 36
Gr Grashof number ( = v2f
) β × 10− 5(K− 1) 21 0.63
α × 10− 7(m2/s) 1.47 118.536
K thermal conductivity (W/m K) μ × 10− 4(Kg/ms) 8.9 –
Nu average Nusselt number dp(nm) – 100
P pressure (Pa)
μ
Pr Prandtl number( = kff )
Up H
Re Reynolds number( = v ) found that at the fixed Reynolds number, the solid concentration affects
Gr
Ri Richardson number( = Re 2) the flow pattern and thermal behavior particularly for a higher Rayleigh
T temperature (K) number. Shahi et al. [12] studied the mixed convective cooling in a
V velocity (m s − 1) square cavity ventilated and partially heated from the below utilizing
x, y Cartesian coordinates (m) nanofluid which results indicated that increase in solid concentration
leads to increase in the average Nusselt number at the heat source
Greek letters surface and decrease in the average bulk temperature. Oztop and
α thermal diffusivity Dagtekin [13] investigated the mixed convection in two-sided lid-driven
β volumetric expansion coefficient (K − 1) differentially heated square cavity which was performed for three
θ angular coordinate different cases characterized by the direction of movement of vertical
ϕ volume fraction walls. Tiwari and das [14] have performed same study as of [13]. Using
μ Dynamic viscosity (N s m 2) Cu-water nanofluid as working fluid, they studied the effect of various
ν kinematic viscosity ( = μρ(m 2s − 1)) solid volume fractions on the streamlines and isotherms. Their results
ρ density (kg m − 3) showed that nanoparticles are able to change the flow pattern of a fluid
from natural convection to forced convection regime. Aydin [15] studied
Subscripts numerically mechanisms of aiding and opposing forces in a shear and
b Bulk buoyancy-driven cavity. He carried out a parametric study and identified
dr Drift three kinds of heat transfer regime: forced convection dominated, mixed
eff Effective and buoyancy dominated regime. Raisi et al. [16] studied numerically
f base fluid forced convection of laminar nanofluid in a microchannel with both slip
k summation index and no-slip condition. They showed that the heat transfer rate is
m Mixture significantly affected by the solid volume fraction and slip velocity
p Particle coefficient at high Reynolds numbers. Convective heat transfer with
h Hot wall nanofluids can be modeled using the two-phase or single phase
c cold wall approach. The first provides the possibility of understanding the
functions of both the fluid phase and the solid particles in the heat
transfer process. The second assumes that the fluid phase and particles
are in thermal equilibrium and move with the same velocity. This
approach is simpler and requires less computational time. Thus it has
been used in several theoretical studies of convective heat transfer with
nanofluids [17–19]. However, due to the fact that the effective properties
of nanofluids are not known precisely, the numerical predictions of this
approach are, in general, not in good agreement with experimental
results. In addition, Ding and Wen [20] showed that the particle
concentration could only be assumed uniform if the corresponding
Peclet numbers are always less than 10. Therefore, the concerns in single
phase modeling consist of selecting the proper effective properties for
nanofluids and taking into account the chaotic movement of ultra fine
particles. To partially overcome this difficulty, some researchers have
used the dispersion model which takes into account the improvement of
heat transfer due to the random movement of particles in the main flow
[21–23]. Due to several factors such as gravity, friction between the fluid
and solid particles and Brownian forces, the phenomena of Brownian

Table 2
Grid independency test against the average Nusselt (Nu) number at ϕ = 8%, θ = 0∘ and
Ri = 0.01.

Grid dimension (X by Y) Nu

21 × 21 27.932
41 × 41 28.020
61 × 61 28.005
81 × 81 28.048
101 × 101 28.099
Fig. 1. Schematic for physical model.
1430 M. Alinia et al. / International Communications in Heat and Mass Transfer 38 (2011) 1428–1435

Table 3
Comparison of the present code with natural convection [7,31–33] and mixed convection [34–37].

Natural convection

Ra Present Khanafer et al.[7] De Vahl Davis [31] Fusegi et al. [32] Barakos and Mitsoulis [33]
3
10
Nu 1.118 1.118 1.118 1.105 1.114
Umax at (Y/H) 0.138 (0.808) 0.137 (0.812) 0.136 (0.813) 0.132 (0.833) 0.153 (0.806)
Vmax at (X/L) 0.140 (0.181) 0.139 (0.173) 0.138 (0.178) 0.131 (0.200) 0.155 (0.181)
104
Nu 2.244 2.245 2.243 2.302 2.245
Umax at (Y/H) 0.193 (0.819) 0.192 (0.827) 0.192 (0.823) 0.201 (0.817) 0.193 (0.818)
Vmax at (X/L) 0.234 (0.121) 0.233 (0.123) 0.234 (0.119) 0.225 (0.117) 0.234 (0.119)
105
Nu 4.519 4.522 4.519 4.646 4.510
Umax at (Y/H) 0.132 (0.850) 0.131 (0.854) 0.153 (0.855) 0.147 (0.855) 0.132 (0.859)
Vmax at (X/L) 0.259 (0.063) 0.258 (0.065) 0.261 (0.066) 0.247 (0.065) 0.258 (0.066)
106
Nu 8.817 8.826 8.799 9.012 8.806
Umax at (Y/H) 0.077 (0.850) 0.077 (0.854) 0.079 (0.850) 0.084 (0.856) 0.077 (0.859)
Vmax at (X/L) 0.263 (0.035) 0.262 (0.039) 0.262 (0.038) 0.259 (0.033) 0.262 (0.039)

Mixed convection

Parameter Present Nu Waheed [34] Khanafer et al. [35] Sharif [36] Iwatsu et al. [37]

Re = 100 2.02 2.03 2.02 – 1.94


Re = 400 4.03 4.02 4.01 4.05 3.84
Re = 1000 6.48 6.55 6.42 6.55 6.33

diffusion, sedimentation, and dispersion may coexist in the main flow of Two sidewalls are heated and cooled at constant temperatures Th
a nanofluid. This means that the slip velocity between the fluid and and Tc, respectively and the other walls are thermally insulated which
particles may not be zero [23]; therefore it seems that the two-phase are moving lids. The computational results were obtained for inclination
approach is better model to apply the nanofluid [24,25]. In this work, angles θ = −60°, −30°, 0°, 30°, 60° and Richardson numbers varying
mixed convection flow of SiO2–water nanofluid in an inclined enclosed from 0.01 to 100. The Prandtl number of the base fluid (water) is 6.2, and
square cavity is analyzed numerically using two phase mixture model to the nanoparticle volume fraction ϕ was, variably, 0%,2%, 4%, 6%and 8%.
evaluate the influence of control parameters on the heat transfer The corresponding thermophysical properties of the fluid and solid
characteristics of nanofluid. The effective thermal conductivity of phases (nano-particles) are shown in Table 1.
nanofluid has been calculated using the model by Maxwell–Garnetts
[26] and to determine the viscosity of nanofluid, the model proposed by 3. Mathematical formulation
Brinkman [27] has been employed.
The mixture model, based on a single fluid two-phase approach, is
2. Problem definition and mathematical model employed in the simulation by assuming that the coupling between
phases is strong, and particles closely follow the flow. The two-phases
Mixed convection of a nanofluid consisting of water and SiO2 in a are assumed to be interpenetrating, meaning that each phase has its
square inclined cavity considered in this study is shown in Fig. 1. own velocity vector field, and within any control volume there is a
volume fraction of primary phase and also a volume fraction of the
secondary phase. Instead of utilizing the governing equations of each
phase separately, the continuity, momentum and energy equations
for the mixture are employed. A nanofluid consisting of water and
SiO2 nanoparticles in a square inclined cavity is considered. The two
sidewalls are in different constant temperatures and the others are
moving lids which are thermally insulated.
The physical properties of the fluid are assumed constant except
for the density in the body force, which varies linearly with the
temperature (Boussinesq's hypothesis). Dissipation and pressure
work are neglected. Therefore, the dimensional equations for steady
state mean conditions are:
Continuity equation:
 
∇⋅ ρeff Vm = 0 ð1Þ

Momentum equation:
   n

∇⋅ ρeff Vm Vm = −∇P + ∇:½τ −ρeff βeff ðT−T0 Þg + ∇⋅ ∑ ϕk ρk Vdr;k Vdr;k
k=1

Energy equation: ð2Þ

Fig. 2. Comparison of the present results (solid lines) with that of [38]. (▲) Represents  n   
the data obtained based on the experimental thermal conductivity and (■) shows the ∇⋅ ∑ ðρk ck Þϕk Vk T = ∇⋅ keff ∇T ð3Þ
results calculated by Hamilton and Crosser model [39]. k=1
M. Alinia et al. / International Communications in Heat and Mass Transfer 38 (2011) 1428–1435 1431

θ = -60
θ = -30 ψ max nf = 0.14412, ψ max bf = 0.14447 ψ max nf = 0.12697, ψ max bf = 0.12682 ψ max nf = 0.00432, ψ max bf = 0.04974

ψ max nf = 0.14403,ψ max bf = 0.14436 ψ max nf = 0.12566, ψ max bf = 0.12439 ψ max nf = 0.05596, ψ max bf = 0.05310
θ=0

ψ max nf = 0.14412,ψ max bf = 0.14442 ψ max nf = 0.12667, ψ max bf = 0.12488 ψ max nf = 0.06318,ψ max bf = 0.06246
θ = 30

ψ max nf = 0.14435,ψ max bf = 0.14466 ψ max nf = 0.12986,ψ max bf = 0.12854 ψ max nf = 0.07063,ψ max bf = 0.07027
θ = 60

ψ max nf = 0.14466,ψ max bf = 0.14500 ψ max nf = 0.13355,ψ max bf = 0.13291 ψ max nf = 0.14065,ψ max bf = 0.07827

Ri = 0.1 Ri = 1 Ri = 10

Fig. 3. Stream lines for various inclination angles and Richardson numbers for pure fluid (Dashed lines - - - -) and nanofluid with ϕ = 8% (solid lines _______).

Volume fraction: The slip velocity (relative velocity) is defined as the velocity of a
    secondary phase (p) relative to the velocity of the primary phase (f):
∇⋅ ϕp ρp Vm = −∇⋅ ϕp ρp Vdr;p ð4Þ
Vpf = Vp −Vf ð8Þ
where
The drift velocity is related to the relative velocity
n
∑ ϕ k ρk V k
k=1
Vm = ð5Þ n ϕk ρk
ρeff Vdr;p = Vpf − ∑ V ð9Þ
k=1 ρeff fk

τ = μeff ∇Vm ð6Þ


The relative velocity is determined from Eq. (10) proposed by
are the mean velocity and shear stress respectively and ϕ is the Manninen et al. [28] while Eq. (11) by Schiller and Naumann [29] is
volume fraction of phase k. used to calculate the drag function fdrag as follows
In Eq. (2), Vdr. k is the drift velocity for the secondary phase k, i.e.
 
the nanoparticles in the present study. 2
ρp dp ρp −ρeff
Vpf = a ð10Þ
Vdr:k = Vk −Vm ð7Þ 18μf fdrag ρp
1432 M. Alinia et al. / International Communications in Heat and Mass Transfer 38 (2011) 1428–1435

θ = -60
θ = -30
θ=0
θ = 30
θ = 60

Ri = 0.1 Ri = 1 Ri = 10

Fig. 4. Isotherms for various inclination angles and Richardson numbers for pure fluid (Dashed lines - - - -) and nanofluid with ϕ = 8%(solid lines _______).

(
1 + 0:15Re0:687 Rep ≤1000 The effective viscosity of a fluid containing a dilute suspension of
p
fdrag = ð11Þ small rigid spherical particles is given by Brinkman [27] as
0:15Rep Rep > 1000

μf
μeff = ð15Þ
dp ρp jVpf j ð1−ϕÞ0:25
where Rep = μm and the acceleration a in Eq. (10) is:

The effective thermal conductivity of the nanofluid restricted to


a = g−ðVm ⋅∇ ÞVm : ð12Þ
spherical nanoparticles is approximated by the Maxwell–Garnetts
model [26]:
The physical properties in the equations above are effective
density:  
keff kp + 2kf −2ϕ kf −kp
=  : ð16Þ
kf kp + 2kf + ϕ kf −kp
ρeff = ð1−ϕÞρf + ϕρp ð13Þ

and thermal expansion coefficient [7]: The local Nusselt number of the nanofluid along the hot wall can
be expressed as

2 3
!
1 βp 1 keff ∂θ
βeff = 4 + 5β
ϕ ρp ⋅ f
ð14Þ Nuloc =
Q
=− ð17Þ
1 + ϕρ f βf
ð1−ϕÞρ
1 + ð1−ϕÞ ρf Qcond;f luid kf ∂X
p
M. Alinia et al. / International Communications in Heat and Mass Transfer 38 (2011) 1428–1435 1433

θ = -60
θ = -30
θ=0
θ = 30
θ = 60

Ri = 0.1 Ri = 1 Ri = 10

Fig. 5. Contours of Nanoparticles distribution for various inclination angles and Richardson numbers at ϕ = 8 %.

The average Nusselt number throughout the cavity, are evaluated is adequate to describe the flow and heat and mass transfer processes
as correctly. The absolute convergence criterion of (|Ωn − Ωn − 1| b 10− 8) is
used for the termination of all computations, where Ω(u, v, T, ϕ) is the
H
dependant variable in the partial differential equations and n is the
Nu = ∫ Nuloc ðyÞdy: ð18Þ iteration number. An under-relaxation parameter of 0.4 is used in order
0
to obtain a stable convergence for the solution of momentum and
energy equations. The validation of present computer programming
Grashof, Prandtl and Reynolds numbers for the based fluid are
code has been verified with the natural convection in a square cavity by
given by:
[7,31–33] besides the results obtained by the present code have been
compared with that of [34–37].
μf gβf ΔTH 3 Up H As seen in Table 3 comparison of the present results for average
Pr = ; Gr = ; Re = : ð19Þ
kf v2f v Nusselt number showed excellent agreement with previous works.
Moreover, in order to demonstrate the precision of the model, results
4. Numerical procedure are compared with those obtained by Jahanshahi et al. [38]. They have
performed a study on natural convection of a water-SiO2 nanofluid
System of equations above is numerically solved by finite-volume using two different models. In the first model they have employed a set
approach using SIMPLE algorithm [30]; central differencing is adopted of experimental data for thermal conductivity of the nanofluid and in
for the diffusion terms and upwind scheme is utilized for the nonlinear the second method they have calculated the thermal conductivity from
convective terms. The algebraic system resulting from numerical the equation proposed by Hamilton and Crosser [39]. As seen in Fig. 2,
discretization is solved by TDMA (Tridiagonal Matrix Algorithm) there is a good agreement between the results. It should be noted that
applied in a line going through all volumes in the computational present (mixture model) results are in much better agreement with the
domain. Non-uniform grid in x and y directions is used for all solution based on experimental thermal conductivity; it is an interesting
computations. To test and assess grid independence of the solution feature of mixture model especially for high Rayleigh numbers since
scheme, numerical experiments are performed as shown in Table 2. higher Rayleigh numbers could result in non-uniform distribution and
These experiments show that an unequally spaced grid mesh of 81 × 81 for which the single phase approach no longer would be precise.
1434 M. Alinia et al. / International Communications in Heat and Mass Transfer 38 (2011) 1428–1435

effect. As seen, a singular cell is formed at low Richardson (Ri = 0.1, 1)


numbers and the shape of the main cell is circular as a dominant
characteristic of the fluid flow; the recirculation is anticlockwise and
some perturbations are seen in streamlines in the lower left and upper
right corners which are characteristics of a lid-driven cavity flow
problem. The shape of the cell becomes ellipsoidal with increase of
Richardson number. For Richardson number Ri = 1 the central vortex
tends to become elliptic and eventually breaks up into three and four
vortices for Richardson number of Ri= 10. As inclination angle enhances
the size of two extra cell increase and finally for θ N 0 they join together.
Here the important phenomenon happening is that with the increased
concentration inside the fluid, nanoparticles help in minimizing the
natural convection effect and precipitate in merging all cells into one for
Ri = 10 and θ = 60.
Fig. 4 illustrates the isotherms at various inclination angles and
Richardson numbers for both the pure fluid and nanofluid with
ϕ = 8%; as shown, at low Richardson numbers effect of inclination
angle is negligible due to the domination of forced convection. As
Richardson number increases, natural convection effect is comparable
with the forced convection effect. As a result, the deflection of
isotherms decreases and isotherms become more stratified. As stated
above the main feature of the isotherms is seen for Ri = 10 and θ = 60
in which with increased concentration of particles inside the fluid,
nanofluid helps in minimizing the natural convection effect.
Using two-phase approach, gives significant information on the
nanoparticle distributions. Fig. 5 depicts the nanoparticle distribution
based on (ϕp −ϕp0)×103 for various inclination angles and Richardson
numbers. The figure demonstrates that nanoparticle distribution becomes
more severe for higher Richardson numbers. This increase in variation of
nanoparticle concentration shows none uniform properties of nanofluid
in the cavity at higher Richardson numbers. At lower Richardson numbers
the particle distribution is fairly uniform and single phase approach could
be well adopted. While increasing the Richardson numbers non-
uniformity on the particle distribution becomes more important and
single phase approach could not well predict the flow parameters.
Fig. 6 (a, b) comprehensively presents the effects of volume
fraction, Richardson number and inclination angles on heat transfer.
Fig. 6 (a) shows the effects of volume fraction on the average Nusselt
number for various Richardson numbers at θ = 0. The figure shows
Fig. 6. Average Nusselt number (a) for various volume fractions and Richardson numbers that Nusselt number increases with growth of volume fraction at any
at θ = 0 and (b) for various inclination angles and Richardson numbers at ϕ = 8%. Richardson. However, the increment is more considerable in low
Richardson numbers. Fig. 6 (b) illustrates the effects of inclination
angles on the average Nusselt number for various Richardson
numbers at ϕ = 8%. The figure demonstrates that the inclination
5. Results and discussion angle has no remarkable effects on heat transfer except forRi = 10,
100 in which natural convection mechanism is dominated; the
A numerical analysis using two-phase mixture model has been Nusselt number reaches its maximum value atθ = 30°.
carried out to investigate the effects of Richardson number, particle
concentration and inclination angle on SiO2–water nanofluid filling a 6. Conclusions
two dimensional inclined square enclosure. The enclosure is differen-
tially heated and the other two opposite moving walls are adiabatic. This study investigates mixed convection of a nanofluid consisting of
Investigations through the cavity are made for various values of volume water and SiO2 in an inclined enclosure using two-phase mixture
fraction of nanoparticle (0≤ ϕ ≤ 8%), inclination angles (− 60° ≤ θ ≤ 60°) model. The numerical results are obtained for different inclination
and Richardson numbers (0.01 ≤ Ri ≤ 100). Richardson number which is angles and various Richardson number at constant Grashof number 104.
defined as Ri = Gr/Re 2, characterizes the relative importance of Results have clearly indicated that the addition of SiO2 nanoparticles has
buoyancy to forced convection. To vary Richardson number, Grashof produced a remarkable enhancement on heat transfer with respect to
number is fixed at Gr = 104 while changing Reynolds number through that of the pure fluid. In addition, nanoparticles are able to change the
the plate velocity Up. The calculations are done with Reynolds number flow pattern of a fluid from natural convection to forced convection
identical at both sides of the cavity. regime. Inclination angle is more pronounced at high Richardson
Fig. 3 illustrates streamlines at various Richardson number and numbers due to domination of natural convection.
inclination angles for both the pure fluid and nanofluid with ϕ = 8%. It is
noted that forces due to moving lids and buoyancy act in opposite Acknowledgments
directions. As seen for low Richardson numbers, the forced convection
becomes dominant and a strong circulation is observed in the cavity. We would like to acknowledge Dr. Seyed Farid Hosseinizadeh for
With increase in Richardson number from Ri = 0.01 to Ri = 100, the the many valuable discussions that helped us understand our research
natural convection effect will be comparable with the forced convection area better.
M. Alinia et al. / International Communications in Heat and Mass Transfer 38 (2011) 1428–1435 1435

References [19] A. Akbarinia, A. Behzadmehr, Numerical study of laminar mixed convection of a


nanofluid in a horizontal curved tube, Applied Thermal Engineering 27 (2007)
[1] J.C. Maxwell, Electricity and Magnetism, Clarendon Press, Oxford, UK, 1873. 1327–1337.
[2] S.U.S. Choi, Enhancing thermal conductivity of fluid with nanoparticles, De- [20] Y. Ding, D. Wen, Particle migration in a flow of nanoparticle suspensions, Powder
velopments and Applications of Non-Newtonian Flow, 66, ASME, 1995, Technology 149 (2005) 84–92.
pp. 99–105, FED 231/MD. [21] S.Z. Heris, M.N. Esfahany, G. Etemad, Numerical investigation of nanofluid laminar
[3] Y.M. Xuan, W. Roetzel, Conceptions for heat transfer correlation of nanofluids, convective heat transfer through a circular tube, Numerical Heat Transfer; Part A:
International Journal of Heat and Mass Transfer 43 (2000) 3701–3707. Applications 52 (2007) 1043–1058.
[4] P. Keblinski, S.R. Phillpot, S.U.S. Choi, J.A. Eastman, Mechanisms of heat flow in [22] Y.M. Xuan, Q. Li, Heat transfer enhancement of nanofluids, International Journal of
suspensions of nano-sized particles (nanofluid), International Journal of Heat and Heat and Fluid Flow 21 (2000) 58–64.
Mass Transfer 45 (2002) 855–863. [23] Y.M. Xuan, Q. Li, Investigation on convective heat transfer and flow features of
[5] D. Wen, Y. Ding, Experimental investigation into convective heat transfer of nanofluids, ASME Journal of Heat Transfer 125 (2003) 151–155.
nanofluids at the entrance region under laminar flow conditions, International [24] A. Behzadmehr, M. Saffar-Avval, N. Galanis, Prediction of turbulent forced
Journal of Heat and Mass Transfer 47 (2004) 5181–5188. convection of a nanofluid in a tube with uniform heat flux using a two-phase
[6] K.B. Anoop, T. Sundararajan, S.K. Das, Effect of particle size on the convective heat approach, International Journal of Heat and Fluid Flow 28 (2007) 211–219.
transfer in nanofluid in the developing region, International Journal of Heat and [25] R. Lotfi, Y. Saboohi, A.M. Rashidi, Numerical study of forced convective heat
Mass Transfer 52 (2009) 735–743. transfer of nanofluids: comparison of different approaches, International
[7] K. Khanafer, K. Vafai, M. Lightstone, Buoyancy-driven heat transfer enhancement Communications in Heat and Mass Transfer 37 (2010) 74–78.
in a two dimensional enclosure utilizing nanofluids, International Journal of Heat [26] J. Maxwell, A Treatise on Electricity and Magnetism, second ed, Oxford University
and Mass Transfer 46 (2003) 3639–3653. Press, Cambridge, UK, 1904, pp. 435–441.
[8] B. Ghasemi, S.M. Aminossadati, Natural convection heat transfer in an inclined [27] H.C. Brinkman, The viscosity of concentrated suspensions and solutions, Journal of
enclosure filled with a water-cuo nanofluid, Numerical Heat Transfer; Part A: Chemical Physics 20 (1952) 571–581.
Applications 55 (2009) 807–823. [28] M. Manninen, V. Taivassalo, S. Kallio, On the mixture model for multiphase flow,
[9] C.J. Ho, M.W. Chen, Z.W. Li, Numerical simulation of natural convection of Technical Research Center of Finland, 288, VTT Publications, 1996, pp. 9–18.
nanofluid in a square enclosure: effects due to uncertainties of viscosity and [29] L. Schiller, A. Naumann, A drag coefficient correlation, Zeitschrift Des Vereines
thermal conductivity, International Journal of Heat and Mass Transfer 51 (2008) Deutscher Ingenieure 77 (1935) 318–320.
4506–4516. [30] S.V. Patnkar, Numerical Heat Transfer and Fluid Flow, Hemisphere Pub., New York,
[10] Y. Zhang, L. Li, H.B. Ma, M. Yang, Effects of Brownian and thermophoretic 1980.
diffusions of nanoparticles on nonequilibrium heat conduction in a nanofluid [31] G. De Vahl Davis, Natural convection of air in a square cavity, a benchmark numerical
layer with periodic heat flux, Numerical Heat Transfer; Part A: Applications 56 solution, International Journal of Numerical Methods Fluids 3 (1962) 249–264.
(2009) 325–341. [32] T. Fusegi, J.M. Hyun, K. Kuwahara, B. Farouk, A numerical study of three-
[11] F. Talebi, A.H. Mahmoudi, M. Shahi, Numerical study of mixed convection flows in dimensional natural convection in a differentially heated cubical enclosure,
a square lid-driven cavity utilizing nanofluid, International Communications in International Journal of Heat and Mass Transfer 34 (1991) 1543–1557.
Heat and Mass Transfer 37 (2010) 79–90. [33] N.C. Markatos, K.A. Pericleous, Laminar and turbulent natural convection in an
[12] M. Shahi, A.H. Mahmoudi, F. Talebi, Numerical study of mixed convective cooling enclosed cavity, International Journal of Heat and Mass Transfer 27 (1984)
in a square cavity ventilated and partially heated from the below utilizing 772–775.
nanofluid, International Communications in Heat and Mass Transfer (2009), doi: [34] M.A. Waheed, Mixed convective heat transfer in rectangular enclosures driven by
10.1016/j.icheatmasstransfer.2009.10.002. a continuously moving horizontal plate, International Journal of Heat and Mass
[13] H.F. Oztop, I. Dagtekin, Mixed convection in two-sided lid-driven differentially Transfer 52 (2009) 5055–5063.
heated square cavity, International Journal of Heat and Mass Transfer 47 (2004) [35] K.M. Khanafer, A.M. Al-Amiri, I. Pop, Numerical simulation of unsteady mixed
1761–1769. convection in a driven cavity, using an externally excited sliding lid, European
[14] Raj Kamal Tiwari, Manab Kumar Das, heat transfer augmentation in a two-sided Journal of Mechanics B Fluids 26 (2007) 669–687.
lid-driven differentially heated square cavity utilizing nanofluids, International [36] M.A.R. Sharif, Laminar mixed convection in shallow inclined driven cavities with
Journal of Heat and Mass Transfer 50 (2007) 2002–2018. hot moving lid on top and cooled from bottom, Applied Thermal Engineering 27
[15] O. Aydin, Aiding and opposing mechanisms of mixed convection in a shear-and (2007) 1036–1042.
buoyancy-driven cavity, International Communications in Heat and Mass Transfer [37] R. Iwatsu, J. Hyun, K. Kuwahara, Mixed convection in a driven cavity with a stable
26 (1999) 1019–1028. vertical temperature gradient, International Journal of Heat and Mass Transfer 36
[16] J. Koo, C. Kleinstreuer, Laminar nanofluid flow in microheat-sinks, International (1993) 1601–1608.
Journal of Heat and Mass Transfer 48 (2005) 2652–2661. [38] M. Jahanshahi, S.F. Hosseinizadeh, M. Alipanah, A. Dehghani, G.R. Vakilinejad,
[17] A. Raisi, B. Ghasemi, S.M. Aminossadati, A numerical forced convection of laminar Numerical simulation of free convection based on experimental measured
nanofluid in a microchannel with both slip and no-slip condition, Numerical Heat conductivity in a square cavity using Water/SiO2 nanofluid, International
Transfer; Part A: Applications 59 (2011) 114–129. Communications In Heat and Mass Transfer 37 (2010) 687–694.
[18] M. Akbari, A. Behzadmehr, F. Shahraki, Fully developed mixed convection in [39] R.L. Hamilton, O.K. Crosser, Thermal conductivity of heterogeneous two
horizontal and inclined tubes with uniform heat flux using nanofluid, Interna- component systems, I&EC Fundamentals 1 (3) (1962) 187–191.
tional Journal of Heat and Fluid Flow 29 (2008) 545–556.

You might also like