You are on page 1of 11

MIXED CONVECTION FLOWS IN A SQUARE LID-DRIVEN

CAVITY WITH HEAT SOURCE AT THE BOTTOM UTILISING


NANOFLUID
M. A. Mansour1 and Sameh E. Ahmed2 *
1. Department of Mathematics, Faculty of Sciences, Assuit University, Assuit, Egypt
2. Department of Mathematics, Faculty of Sciences, South Valley University, Qena, Egypt

This paper presents a numerical investigation of laminar mixed convection cooling of heat source embedded on the bottom wall of an enclosure
filled with nanofluids. The transport equations for a Newtonian fluid are solved numerically with a finite volume approach using the SIMPLE
algorithm. The influences of governing parameters, namely, Rayleigh number location and geometry of the heat source, the type of nanofluid and
solid volume fraction of nanoparticles on the cooling performance is studied. The present results are validated by favourable comparisons with
previously published results. The results of the problem are presented in graphical and tabular forms and discussed.

Keywords: nanofluid, mixed convection, heat source, finite volume method

INTRODUCTION observed that the thermal conductivity for nanofluid increases


with the increase in temperature. They have also observed the

N
anofluids are created by dispersing nanometer-sized par-
stability of Al2 O3 –water and Cu–water nanofluid. Experiments
ticles (<100 nm) in abase fluid such as water, ethylene
on heat transfer due to natural convection with nanofluid have
glycol or propylene glycol. Use of high thermal conduc-
been studied by Putra et al. (2003); and Wen and Ding (2006).
tivity metallic nanoparticles (e.g. copper, aluminium, silver and
They have observed that heat transfer decreases with the increase
silicon) increases the thermal conductivity of such mixtures, thus
in concentration of nanoparticles. The viscosity of this nanofluid
enhancing their over all energy transport capability (Xuan and Li,
increases rapidly with inclusion of nanoparticles as shear rate
2003). Nanofluids have attracted attention as a new generation of
decreases.
heat transfer fluids in building heating, in heat exchangers, in
Recently, Kumar et al. (2010) used a single phase thermal dis-
plants and in automotive cooling applications, because of their
persion model to study the flow and thermal field in nanofluid.
excellent thermal performance. Various benefits of the application
Talebi et al. (2010) investigated numerically the problem of mixed
of nanofluids include: improved heat transfer, heat transfer sys-
convection flows through a Cu–water nanofluid in a square lid-
tem size reduction, minimal clogging, microchannel cooling and
driven cavity. The problem of natural convection cooling of a
miniaturisation of systems (Choi, 1995). Therefore, research is
localised heat source at the bottom of a nanofluid-filled enclo-
underway to apply nanofluids in environments where higher heat
sure was discussed by Aminossadatia and Ghasemi (2009). Oztop
flux is encountered and the conventional fluid is not capable of
and Abu-Nada (2008) presented a numerical study for heat trans-
achieving the desired heat transfer. Xuan et al. (2005) have exam-
fer and fluid flow due to buoyancy forces in a partially heated
ined the transport properties of nanofluid and have expressed that
thermal dispersion, which takes place due to the random move-
ment of particles, takes a major role in increasing the heat transfer ∗ Author to whom correspondence may be addressed.
rate between the fluid and the wall. This requires a thermal dis- E-mail address: sameh sci math@yahoo.com
persion coefficient, which is still unknown. Brownian motion Can. J. Chem. Eng. 90:100–110, 2012
of the particles, ballistic phonon transport through the particles © 2011 Canadian Society for Chemical Engineering
DOI 10.1002/cjce.20533
and nanoparticles clustering can also be the possible reason for Published online 21 April 2011 in Wiley Online Library
this enhancement (Keblinski et al., 2002). Das et al. (2003) have (wileyonlinelibrary.com).

| 100 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 90, FEBRUARY 2012 |
enclosure using nanofluids made with different types of nanopar-
ticles. Wen and Ding (2005) found a systematic and definite
deterioration in the heat transfer for a particular range of Rayleigh
numbers and density and concentration of nanoparticles. Similar
results were also obtained by Santra et al. (2008) who modelled
the nanofluid as a non-Newtonian fluid. Ho et al. (2008) argued
that the heat transfer in a square enclosure filled nanofluids can
be enhanced or mitigated depending on the formulas used for
the estimated dynamic viscosity of the nanofluid. Hwang et al.
(2007) theoretically investigated thermal characteristics of nat-
ural convection in a rectangular cavity filled with a water-based
nanofluid containing alumina. Abu-Nada et al. (2010) investigated
the heat transfer enhancement in a differentially heated enclo-
sure using variable thermal conductivity and variable viscosity of
Al2 O3 –water and Cu–water nanofluids.
On the other hand, fluid flow and heat transfer in a cavity filled
by pure fluid which is driven by buoyancy and shear have been
studied extensively in the literature (Rubin and Khosla, 1977; Ghia
et al., 1982; Hsu et al., 1995). The most usage of the mixed con-
vection flow with lid-driven effect is to include the cooling of the
electronic devices, lubrication technologies, drying technologies,
etc.
The present work has been concerned with the laminar mixed
convection flows of nanofluids in a square cavity with a moving Figure 1. Physical model of the problem.
lid that moves uniformly in the horizontal plane while all other
walls of the cavity are fixed. The focus of the present study is
on the analysis of several pertinent parameters such as Rayleigh governing equations are (Talebi et al., 2010):
number, Reynolds number, length and location of the heat source
and the solid volume fraction. ∂u ∂v
+ =0 (1)
∂x ∂y

MATHEMATICAL MODEL  
∂u ∂u 1 ∂p ∂2 u ∂2 u
u +v =− + nf + 2 (2)
Consider laminar, mixed convection flow inside a square cavity. In ∂x ∂y nf ∂x ∂x 2 ∂y
the present problem, the following assumptions have been made:
• The bottom horizontal wall has an embedded constant heat  
source q . ∂v ∂v 1 ∂p ∂2 v ∂2 v g
u +v =− + nf + 2 + (T−T∞ )[(ˇ)p
• The non-heated parts of the bottom wall and the right vertical ∂x ∂y nf ∂y ∂x 2 ∂y nf
wall are insulated heated.
• The left vertical wall and the horizontal top wall of the enclosure + (1−)(ˇ)f ] (3)
are maintained at a relatively low temperature (Tc ).  
• The top wall moves from left to right with uniform velocity U0 . ∂T ∂T ∂2 T ∂2 T
u +v = ˛nf + 2 (4)
• The nanofluids used in the analysis are assumed to be Newto- ∂x ∂y ∂x 2 ∂y
nian, incompressible and laminar.
• The base fluid (water) and the solid spherical nanoparticles where the effective density of the nanofluid is given as:
(Cu, Ag, Al2 O3 and TiO2 ) are in thermal equilibrium.
• The thermo-physical properties of the base fluid and the nf = (1−)f + p (5)
nanoparticles are given in Table 1.
• The thermo-physical properties of the nanofluid are assumed
constant except for the density variation, which is determined and  is the solid volume fraction of nanoparticles. Thermal dif-
based on the Boussinesq approximation. fusivity of the nanofluid is:

The geometric and the Cartesian coordinate system are schemat- Knf
˛nf = (6)
ically shown in Figure 1. Under the above assumptions, the (Cp )nf

Table 1. Thermo-physical properties of water and nanoparticles Talebi et al. (2010)

Pure water Copper (Cu) Silver (Ag) Alumina Al2 O3 Titanium oxide (TiO2 )

 (kg m−3 ) 997.1 8933 10 500 3970 4250


Cp (J kg−1 K−1 ) 4179 385 235 765 686.2
k (W m−1 K−1 ) 0.613 401 429 40 8.9538
ˇ (K−1 ) 21 × 10−5 1.67 × 10−5 1.89 × 10−5 0.85 × 10−5 0.9 × 10−5

| VOLUME 90, FEBRUARY 2012 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 101 |
where the heat capacitance of the nanofluid given is:
Table 2. Grid independency results (Cu–water: Re = 10, B = 0.4,
D = 0.5, Ra = 1.47 × 106 and  = 0.1)
(Cp )nf = (1−)(Cp )f + (Cp )p (7)
Grid 31 × 31 61 × 61 81 × 81 101 × 101
The thermal expansion coefficient of the nanofluid can be deter-
mined by: Num 21.11349 21.16497 21.33016 21.73634

(ˇ)nf = (1−)(ˇ)f + (ˇ)p (8)


The local Nusselt number on the heat source surface can be
The effective dynamic viscosity of the nanofluid given by defined as:
Brinkman (1952) is:
hL
Nus = (17)
␮f Kf
␮nf = (9)
(1−)2.5
where h is the convection heat transfer coefficient:
In Equation (6), Knf is the thermal conductivity of the nanofluid
q
which for spherical nanoparticles, according to Maxwell (1904), h= (18)
is: Ts −Tc
  The local Nusselt number by using the dimensionless parame-
(Kp + 2Kf )−2(Kf −Kp )
Knf = Kf (10) ters (Equation 11) is:
(Kp + 2Kf ) + (Kf −Kp )
1
where Kp is the thermal conductivity of dispersed nanoparticles Nus (X) = (19)
s (X)
and Kf is the thermal conductivity of pure fluid.
Equations (1)–(4) can be converted to non-dimensional param- where  s is the dimensionless heat source temperature. The aver-
eters: age Nusselt number is determined by:
x y u v T−Tc p  D+B/2
X= ,Y = ,U = ,V = ,  =  ,P = (11) 1
L L U0 U0 q L/Kf nf U02 Num = Nus (X) dX (20)
B D−B/2
The non-dimensional continuity, momentum and energy equa-
tions are written as follows:

∂U ∂V NUMERICAL METHOD AND VALIDATION


+ =0 (12)
∂X ∂Y The above equations have been solved numerically based on the
  finite volume method using a collocated grid system. The con-
∂U ∂U ∂P 1 f 1 ∂ U
2
∂ U 2
vection diffusion terms were treated with a power-law scheme.
U +V =− + + (13)
∂X ∂Y ∂X Re nf (1−)2.5 ∂X 2 ∂Y 2 The resulting discretised equations have been solved iteratively
through strongly implicit procedure (SIP) (Ferziger and Peric,
  1999). The SIMPLE algorithm (Patnkar, 1980) has been adopted
∂V ∂V ∂P 1 f 1 ∂2 V ∂2 V for the pressure velocity coupling. More details of the discreti-
U +V =− + +
∂X ∂Y ∂Y Re nf (1−)2.5 ∂X 2 ∂Y 2 sation and computational procedure can be found in literature
  (Ferziger and Peric, 1999). To check the convergence of the
Ra f p ˇp sequential iterative solution, the sum of the absolute differences of
+ 1− +   (14)
Re2 Pr nf f ˇf the solution variables between two successive iterations has been
calculated. When this summation falls below the convergence cri-
 
∂ ∂ Knf (Cp )f 1 ∂2  ∂2  terion, convergence is obtained, which the convergence criterion
U +V = + (15) has been chosen as 10−5 . To allow grid independent examination,
∂X ∂Y Kf (Cp )nf RePr ∂X 2 ∂Y 2
the numerical procedure has been conducted for different grid
The dimensionless forms of the boundary conditions are: resolutions. Table 2 shows an accuracy tests using the finite vol-
At X = 0, U = V =  = 0 (left wall at constant temperature, ume method using four sets of grids: 31 × 31, 61 × 61, 81 × 81 and
T = Tc ) 101 × 101. There is a good agreement was found between 61 × 61
At X = 1, U = V = 0, ((∂)/(∂X)) = 0 (right wall is thermally and 81 × 81 grids, so the numerical computations were carried
insulated).

∂ Table 3. Comparisons of – min for classical fluids


Y = 0, 0 ≤ X < (D−0.5B) : U = V = 0, =0
∂Y
Gr = 0, Gr = 104 ,
∂ kf Calculation method Re = 100 Re = 100
Y = 0, (D−0.5B) ≤ X ≤ (D + 0.5B) : U = V = 0, =
∂Y knf
Finite difference (Hsu et al., 1995) 0.1013 —
∂ Multigrid (Ghia et al., 1982) 0.1034 —
Y = 0, (D + 0.5B) < X ≤ 1 : U = V = 0, =0
∂Y Spline (Rubin and Khosla, 1977) 0.1054 0.0934
Y = 1, 0 ≤ X ≤ 1 : U = 1, V = 0,  = 0 (16) Finite difference (Present study) 0.1042 0.0914

| 102 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 90, FEBRUARY 2012 |
out for 61 × 61 and 81 × 81 grid nodal points. In order to verify values of − m . As we can see form Table 3, the results are found
the accuracy of present method, the obtained results in special in a good agreement with these results. These favourable com-
cases are compared with the results obtained by Hsu et al. (1995); parisons lend confidence in the numerical results to be reported
Rubin and Khosla (1977); and Ghia et al. (1982) in terms of the subsequently.

Figure 2. Streamlines (left) and isotherms (right) for the enclosures filled with Cu–water nanofluid at  = 0.0, 0.05, 0.1 and 0.2 (increasing from top
towards bottom). The referenced case is Ra = 1.47 × 105 , Re = 10, B = 1/3 and D = 0.5.

| VOLUME 90, FEBRUARY 2012 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 103 |
RESULTS AND DISCUSSION
Two-dimensional mixed convection is studied for nanofluid in
a square lid-driven cavity for Rayleigh number (1.47 × 103 ≤
Ra ≤ 1.47 × 106 ), solid volume fraction (0 ≤  ≤ 0.2), heat source
lengths (0.2 ≤ B ≤ 0.8), heat source locations (0.2 ≤ D ≤ 0.5)
and a choice of nanoparticles (Cu, Ag, Al2 O3 and TiO2 ). For
all simulations, pure water is considered as the base fluid with
Pr = 6.2.

The Effects of Solid Volume Fraction


Figure 2 shows the effects of solid volume fraction ( =
0.0, 0.05, 0.1 and 0.2) on the streamlines and isotherms contours
for cavity filled with Cu–water nanofluid. As seen for each case,
there is main recirculation flow cell, which is formed inside the
enclosure. This cell was generated by the lid and the presence of
the heat source at the bottom wall of the cavity helps the cell to
stretch horizontally beside position of the heat source. Increase in
solid volume fraction causes a decrease in the intensity of buoy-
ancy and hence the flow intensity. For more understanding for Figure 4. Profiles of the local Nusselt number along the heat source for
this influence on the fluid flow, the vertical velocity component at variations of solid volume fractions.
the enclosure mid-section is plotted in Figure 3. It is found that, as
expected, increase in the solid volume fraction leads to decrease
the vertical velocity component. Regarding isotherms contours, the buoyancy forces generated due to the fluid temperature dif-
when there is no concentration of nanoparticles in the base fluid ferences and activity of the fluid motion. On the other hand, as
(i.e.  = 0) the isotherms lines distribute along the whole cav- length of the heat source increases, the generation rates of temper-
ity and get a pushed towards the right wall (adiabatic wall). By ature increase. So the higher temperature patterns corresponds the
adding the nanoparticles, this causes in the reduction of the fluid taller heat source and the opposite is valid. In addition, in order to
temperature and increasing the corresponding local Nusselt num- understand the flow behaviour in this situation, the vertical veloc-
ber along the heat source Nus . This is clearly shown in Figure 4, ity profiles at the enclosure mid-section are presented in Figure
which displays the variation of average Nusselt number along the 6. It is clear that the vertical velocity component increases with
heat source with solid volume fraction. the increase in the heat source length. This is because of stronger
buoyant flow for higher heat generation rates. On the contrary, as
The Effects of Heat Source Length we can see from Figure 7, at high values of the heat source length,
Figure 5 displays the effects of heat source length (B = the corresponding Nusselt number along the heat source reaches
0.2, 0.4, 0.6 and 0.8) on the streamlines and isotherms contours its minimum values.
for cavity filled with Cu–water nanofluid. It is found that, increase
in the heat source length B leads to increase in activity of the The Effects of Different Locations of the Heat Source
fluid motion and there are a clockwise circular cell was formed
inside the cavity. This increase relates to the association between Figure 8 depicts the effects of different locations of the heat source
(D = 0.2, 0.3, 0.4 and 0.5) on the streamlines and isotherms con-
tours for cavity filled with Cu–water nanofluid. It is observed that,
the fluid follows geometry of the cavity by forming one clockwise
circular cell. As the heat source moves away from the left wall of
the cavity, the intensity of the fluid motion increases. In addition,
the effect of changing locations of the heat source on the fluid flow
also, can be observed from Figure 9. It can be seen from this figure
that the vertical velocity component increases with the increase in
the distance between the left wall and the centre of the heat source
which represented by parameter D. On the other hand, the fluid
temperature is influenced by changing the heat source location. It
follows the movement of the heat source. Moreover the lower gen-
eration rates of the temperature can be gotten by putting the heat
source beside the left wall (cold wall). Whereas, Figure 10 shows
the opposite behaviour for the rate of heat transfer represented by
local Nusselt number along the heat source Nus .

The Effects of Type of Nanofluid


In this part of the study, an enclosure filled with different types of
nanofluids (Cu, Ag, Al2 O3 and TiO2 ) with a heat source is consid-
ered. Figure 11 shows the profiles of local Nusselt number along
Figure 3. Effect of the solid volume fraction on vertical velocity the heat source Nus for different nanofluids at B = 0.4, Re = 10,
component at the cavity mid-section. D = 0.5 and Ra = 1.47 × 105 . It is found that the largest rate of heat

| 104 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 90, FEBRUARY 2012 |
Figure 5. Streamlines (left) and isotherms (right) for the enclosures filled with Cu–water nanofluid at B = 0.2, 0.4, 0.6 and 0.8 (increasing from top
towards bottom). The referenced case is Ra = 1.47 × 105 , Re = 10,  = 0.1 and D = 0.5.

transfer can be obtained by adding Cu compared to other nanopar- respectively. It is observed that, for any nanoparticle, increase
ticles. On the contrary, TiO2 gives the lowest rate of heat transfer. in the solid volume fraction leads to increase the average Nus-
In addition, Figures 12 and 13 present the profiles of average selt number and decrease the maximum temperature. Moreover,
Nusselt number along the heat source Num and maximum temper- Table 4 shows the percentages of maximum temperature reduc-
ature  max with solid volume fraction for different nanoparticles, tion and average Nusselt number increasing along the heat source

| VOLUME 90, FEBRUARY 2012 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 105 |
max (%)

6.83
13.50
19.86
25.93
Table 4. The percentages for the reduction of maximum temperature and increasing of average Nusselt number along the heat source for different nanofluids comparing to pure water at

TiO2

Num (%)

4.49
10.52
17.88
26.49
max (%)

8.09
15.87
23.21
30.06
Figure 6. Effect of variations of the heat source length on vertical
velocity component at the cavity mid-section.

Al2 O3

Num (%)

5.66
13.26
22.56
33.56
max (%)

8.48
16.47
23.96
30.95
different values of solid volume fraction at B = 0.4, D = 0.5, Ra = 1.47 × 103 and Re = 10

Ag

Num (%)

6.10
13.99
23.64
35.11

Figure 7. Profiles of the local Nusselt number along the heat source for
variations of the heat source length.
max (%)

8.75
16.96
24.59
31.66

for different nanofluids respect to pure water at different values


of solid volume fraction at B = 0.4, D = 0.5, Ra = 1.47 × 103 and
Re = 10. It is clear that, in general, among all nanoparticles, Cu
Cu

provides the highest average Nusselt number increase and the


lowest maximum temperature reduction (6.51% and 8.75% for
 = 0.05) and TiO2 provides the lowest average Nusselt num-
Num (%)

6.51
14.83
24.90
36.75

ber increase and the highest temperature reduction (4.49% and


6.83% for  = 0.05). Also, Table 5 presents the effect of variations
of heat source length on the average Nusselt number for different
nanoparticles at D = 0.5, Ra = 1.47 × 105 ,  = 0.1 and Re = 10.
It is found that, for all nanoparticles, increase in the heat source
length leads to increase the average Nusselt number along the
heat source Num .
0.05

0.15
0.1

0.2


| 106 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 90, FEBRUARY 2012 |
Figure 8. Streamlines (left) and isotherms (right) for the enclosures filled with Cu–water nanofluid at D = 0.2, 0.3, 0.4 and 0.5 (increasing from top
towards bottom). The referenced case is Ra = 1.47 × 105 , Re = 10, B = 0.4 and  = 0.1.

| VOLUME 90, FEBRUARY 2012 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 107 |
Figure 9. Effect of the different locations of heat sink on vertical velocity Figure 12. Profiles of the average Nusselt number along the heat source
component at the cavity mid-section. with solid volume fraction for different nanoparticles.

Figure 13. Profiles of the maximum temperature with solid volume


Figure 10. Profiles of the local Nusselt number along the heat source for
fraction for different nanoparticles.
variations of the heat source locations.

Table 5. Effect of variations of the heat source length on average


Nusselt number for different nanoparticles at D = 0.5,
Ra = 1.47 × 105 ,  = 0.1 and Re = 10

B Cu Ag Al2 O3 TiO2

0.2 22.28229 22.01756 22.14328 21.86614


0.4 16.49364 16.2714 16.52369 16.24692
0.6 12.8699 12.6993 12.90113 12.67631
0.8 10.09051 9.958565 10.13802 9.957543

CONCLUSION
A numerical simulation of mixed convection flows in a square
lid-driven cavity partially heated from below using Cu–water,
Ag–water, Al2 O3 –water and TiO2 –water nanofluid was studied.
Figure 11. Profiles of the local Nusselt number along the heat source for The finite volume method was employed for the solution of the
different nanoparticles. present problem. Comparisons with previously published work

| 108 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 90, FEBRUARY 2012 |
on special cases of the problem were performed and found to REFERENCES
be in good agreement. Graphical and tabular results for various
Abu-Nada, E., Z. Masoud, H. Oztopd and A. Campo, “Effect of
parametric conditions were presented and discussed. From this
Nanofluid Variable Properties on Natural Convection in
investigation, we can draw the following conclusions:
Enclosures,” Int. J. Ther. Sci. 49, 479–491 (2010).
1. Increase in solid volume fraction leads to decrease of both Aminossadatia, S. M. and B. Ghasemi, “Natural Convection
activity of the fluid motion and the fluid temperature; how- Cooling of a Localised Heat Source at the Bottom of a
ever, it increases the corresponding average Nusselt number. Nanofluid-Filled Enclosure,” Eur J. Mech. B 28, 630–640
2. As length of the heat source increases, this leads to not only (2009).
increasing the flow intensity but also increasing the fluid Brinkman, H. C., “The Viscosity of Concentrated Suspensions
temperature however it decreases the corresponding average and Solution,” J. Chem. Phys. 20, 571–581 (1952).
Nusselt number. Choi, S., “Enhancing Thermal Conductivity of Fluids With
3. Putting heat source at the centre of bottom wall gives large Nanoparticles,” ASME 66, 99–105 (1995).
rates of the heat transfer but small rates of the heat transfer Das, S. K., N. Putra, P. Thiesen and W. Roetzel, “Temperature
can be obtained by putting heat source beside the left wall. Dependence of Thermal Conductivity Enhancement for
4. By adding Cu nanoparticles to the base fluid this gives large Nanofluids,” J. Heat Transfer 125, 567–574 (2003).
values of Nusselt number on the contrary, by adding TiO2 Ferziger, J. H. and M. Peric, “Computational Method for Fluid
nanoparticles to the base fluid this gives small values of Nus- Dynamic,” Springer-Verlag, New York (1999).
selt number. Ghia, U., K. N. Ghia and C. T. Shin, “High-Re Solutions for
Incompressible Flow Using the Navier–Stokes Equations and
a Multigrid Method,” J. Comp. Phys. 48, 387–411 (1982).
NOMENCLATURE Ho, C. J., M. W. Chen and Z. W. Li, “Numerical Simulation of
B length of the heat source (b/L) Natural Convection of Nanofluid in a Square Enclosure:
Cp specific heat (J kg−1 K−1 ) Effects Due To Uncertainties of Viscosity and Thermal
D distance of heat source from the left wall (d/L) Conductivity,” Int. J. Heat Mass Transfer 51, 4506–4516
Gr Grashof number (gˇf L3 T/vf2 ) (2008).
g gravitational acceleration (m s−2 ) Hsu, T.-H., P.-T. Hsu and C.-K. Chen, “Thermal Convection of
k thermal conductivity (W m−1 K−1 ) Micropolar Fluids in a Lid-Driven Cavity,” Int. Commun. Heat
L length of the cavity (m) Mass Transfer 22, 189–200 (1995).
Nus Nusselt number along the heat source Hwang, K. S., J. H. Lee and S. P. Jang, “Buoyancy-Driven Heat
Num average Nusselt number along the heat source Transfer of Water-Based Al2 O3 Nanofluids in a Rectangular
p pressure (pa) Cavity,” Int. J. Heat Mass Transfer 50, 4003–4010 (2007).
P dimensionless pressure (p/pnf U02 ) Keblinski, P., S. R. Phillpot, S. U. S. Choi and J. A. Eastman,
Pr Prandtl number (f /˛f ) “Mechanisms of Heat Flow in Suspensions of Nano-Sized
q heat generation per area (W/m2 ) Particles (Nanofluids),” Int. J. Heat Mass Transfer 45,
Ra Rayleigh number (gˇf L3 T/vf ˛f ) 855–863 (2002).
Re Reynolds number (f U0 L/ ␮f ) Kumar, S., S. K. Prasad and J. Banerjee, “Analysis of Flow and
T temperature (K) Thermal Field in Nanofluid Using a Single Phase Thermal
u,v velocity components in x,y directions (m s−1 ) Dispersion Model,” Appl. Math. Model. 34, 573–592 (2010).
U,V dimensionless velocity components (u/U0 , v/U0 ) Maxwell, J., “A Treatise on Electricity and Magnetism,” 2nd ed.,
x,y Cartesian coordinates (m) Oxford University Press, Cambridge, UK (1904).
X,Y dimensionless coordinates (x/L,y/L) Oztop, H. F. 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
Greek Symbols (2008).
Patnkar, S. V., “Numerical Heat Transfer and Fluid Flow,”
˛ thermal diffusivity (m2 s−1 (k/Cp ))
Hemisphere, New York (1980).
ˇ thermal expansion coefficient (k−1 )
Putra, N., W. Roetzel and S. K. Das, “Natural Convection of
T Ref. temperature difference (q L/kf )
Nanofluids,” Heat Mass Transfer 39, 775–784 (2003).
 solid volume fraction
Rubin, S. G. and P. K. Khosla, “Polynomial Interpolation
 dimensionless temperature (T − Tc /T)
Methods for Viscous Flow Calculations,” J. Comp. Phys. 24,
␮ dynamic viscosity (kg m−1 s−1 )
217–244 (1977).
 kinematic viscosity (m2 s−1 (␮/))
Santra, A. K., S. Sen and N. Chakraborty, “Study of Heat Transfer
 density (kg m−3 )
Characteristics of Copper–Water Nanofluid in a Differentially
Heated Square Cavity With Different Viscosity Models,” J.
Subscripts Enhanced Heat Transfer 15, 273–287 (2008).
Talebi, F., A. H. Mahmoudi and M. Shahi, “Numerical Study of
c cold wall
Mixed Convection Flows in a Square Lid-Driven Cavity
f pure fluid
Utilising Nanofluid,” Int. Commun. Heat Mass Transfer 37,
m average
79–90 (2010).
nf nanofluid
Wen, D. and Y. Ding, “Formulation of Nanofluids for Natural
0 reference state
Convective heat Transfer Applications,” Int. J. Heat Fluid
p nanoparticle
Flow 26, 855–864 (2005).
s surface of the heat source

| VOLUME 90, FEBRUARY 2012 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 109 |
Wen, D. and Y. Ding, “Natural Convective Heat Transfer of
Suspensions of Titanium Dioxide Nanoparticles
(Nanofluids),” IEEE Trans. Nanotechnol. 5, 220–227 (2006).
Xuan, Y. and Q. Li, “Investigation on Convective Heat Transfer
and Flow Features of Nanofluids,” J. Heat Transfer 125,
151–155 (2003).
Xuan, Y., K. Yu and Q. Li, “Investigation on Flow and Heat
Transfer of Nanofluids by the Thermal Lattice Boltzmann
Model,” Prog. Comput. Fluid Dyn. 5, 13–19 (2005).

Manuscript received August 5, 2010; revised manuscript


received October 16, 2010; accepted for publication November 14,
2010.

| 110 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 90, FEBRUARY 2012 |

You might also like