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

International Communications in Heat and Mass Transfer 37 (2010) 1504–1512

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 simulation of mixed convection flows in a square lid-driven cavity


partially heated from below using nanofluid☆
M.A. Mansour a, R.A. Mohamed b, M.M. Abd-Elaziz b, Sameh E. Ahmed b,⁎
a
Department of Mathematics, Faculty of Sciences, Assuit University, Assuit, Egypt
b
Department of Mathematics, Faculty of Sciences, South Valley University, Qena, Egypt

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

Available online 14 October 2010 The present numerical study deals with mixed convection in a square lid-driven cavity partially heated from
below and filled with water-base nanofluid containing various volume fractions of Cu, Ag, Al2O3 and TiO2.
Keywords: Finite difference method was employed to solve the dimensionless governing equations of the problem. The
Nanofluid effects of governing parameters, namely, Reynolds number, solid volume fraction, different values of the heat
Mixed convection source length and different locations of the heat source on the streamlines and isotherms contours as well as
Lid-driven
Nusselt number and average Nusselt number along the heat source were considered. The present results are
Finite difference method
validated by favorable comparisons with previously published results. The results of the problem are
presented in graphical and tabular forms and discussed.
© 2010 Elsevier Ltd. All rights reserved.

1. Introduction They have also observed the stability of Al2O3–water and Cu–water
nanofluid. Experiments on heat transfer due to natural convection
Nanofluids are created by dispersing nanometer-sized particles with nanofluid have been studied by Putra et al. [6] and Wen and Ding
(b100 nm) in a base fluid such as water, ethylene glycol or propylene [7]. They have observed that heat transfer decreases with increase in
glycol. Use of high thermal conductivity metallic nanoparticles (e.g., concentration of nanoparticles. The viscosity of this nanofluid
copper, aluminum, silver and silicon) increases the thermal conduc- increases rapidly with the inclusion of nanoparticles as shear rate
tivity of such mixtures, thus enhancing their overall energy transport decreases. More applications and good understanding of the subject is
capability [1]. Nanofluids have attracted attention as a new generation given in the recent articles [8–17].
of heat transfer fluids in building heating, in heat exchangers, in plants On the other hand, fluid flow and heat transfer in a cavity filled by
and in automotive cooling applications, because of their excellent pure fluid which is driven by buoyancy and shear have been studied
thermal performance. Various benefits of the application of nanofluids extensively in literature [18–20]. The most usage of the mixed
include: improved heat transfer, heat transfer system size reduction, convection flow with lid-driven effect is to include the cooling of the
minimal clogging, micro channel cooling and miniaturization of electronic devices, lubrication technologies, drying technologies, etc.
systems [2]. Therefore, research is underway to apply nanofluids in Motivated by the investigations mentioned above, the purpose of
environments where higher heat flux is encountered and the the present work is to consider mixed convection flows of Cu–water,
conventional fluid is not capable of achieving the desired heat transfer. Ag–water, Al2O3–water and TiO2–water nanofluid in a square cavity
Xuan et al. [3] have examined the transport properties of nanofluid and with a moving lid that moves uniformly in the horizontal plane and
have expressed that thermal dispersion, which takes place due to the partially heated from below.
random movement of particles, takes a major role in increasing the
heat transfer rate between the fluid and the wall. This requires a
thermal dispersion coefficient, which is still unknown. Brownian 2. Mathematical formulation
motion of the particles, ballistic phonon transport through the
particles and nanoparticles clustering can also be the possible reason Consider a steady two-dimensional flow inside a square cavity filled
for this enhancement [4]. Das et al. [5] has observed that the thermal with nanofluid. In the present problem, the following assumptions have
conductivity for nanofluid increases with increasing temperature. been made

I. In the cavity, the left, right and top walls are kept to be cooled
☆ Communicated by W.J. Minkowycz.
II. A heat source is located on a part of the bottom wall and the
⁎ Corresponding author. other parts are thermally insulated.
E-mail address: sameh_sci_math@yahoo.com (S.E. Ahmed). III. The top wall moves from left to right with uniform velocity U0.

0735-1933/$ – see front matter © 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.icheatmasstransfer.2010.09.004
M.A. Mansour et al. / International Communications in Heat and Mass Transfer 37 (2010) 1504–1512 1505

Nomenclature
B length of the heat source (b/H)
Cp specific heat, J kg− 1 K− 1
D distance of heat source from the left wall(d/H)
Gr Grashof number(gβf H3ΔT/v2f )
g gravitational acceleration,ms− 2
H length of the cavity, m
k thermal conductivity, Wm− 1K− 1
Nus Nusselt number along the heat source
Num average Nusselt number along the heat source
p pressure, pa
P dimensionless pressure(p/pnfU20)
Pr Prandtl number (vf/αf)
q" heat generation per area, W/m2
Ra Rayleigh number (gβfH3ΔT/vfαf)
Re Reynolds number (ρfU0H/μf)
T temperature, K
u, v velocity components in x, y directions,ms− 1
U, V dimensionless velocity components (u/U0, v/U0)
x, y Cartesian coordinates, m
X, Y dimensionless coordinates (x/H, y/H)
Fig. 1. Physical model of the problem.

Greek symbols
α thermal diffusivity, m2s− 1(k/ρCp) The geometric and the Cartesian coordinate system are schematically
β thermal expansion coefficient, k− 1 shown in Fig. 1. Under the above assumptions, the governing
ΔT Ref. temperature difference (q"H/kf) equations are (see [9,10])
ϕ solid volume fraction
θ dimensionless temperature(T − Tc/ΔT)
∂u ∂v
μ dynamic viscosity. kgm− 1 s− 1 + =0 ð1Þ
υ kinematic viscosity,m2 s − 1(μ/ρ) ∂x ∂y
ρ density,kg m− 3 " !#
2 2
∂u ∂u 1 ∂p ∂ u ∂ u
u +v = − + μ nf + ð2Þ
∂x ∂y ρnf ∂x ∂x2 ∂y2
Subscripts
c cold wall
" ! #
f pure fluid ∂v ∂v 1 ∂p
2
∂ v
2
∂ v
m average u +v = − + μ nf + + ðρβÞnf g ðT−Tc Þ ð3Þ
∂x ∂y ρnf ∂y ∂x2 ∂y2
nf nanofluid
0 reference state
!
p nanoparticle ∂T ∂T ∂2 T ∂2 T
s surface of the heat source u +v = αnf + ð4Þ
∂x ∂y ∂x 2
∂y2

where, ρnf is the effective density of the nanofluid and it is given by


IV. The base fluid (water) and the solid spherical nanoparticles
(Cu, Ag, Al2O3 and TiO2) are in thermal equilibrium. ρnf = ð1−ϕÞρf + ϕρp ð5Þ
V. The thermo-physical properties of the nanofluid are assumed
constant except for a variation of the density which is
determined based on Boussinesq approximation. Table 2
Grid independency results (Cu–water, B = 0.4, Ra = 103, Re = 10, ϕ = 0.1 , D = 0.5).
VI. Table 1 presents the thermo-physical properties for the base
fluid and the nanoparticles. Grid ψmin θmax Num

31 × 31 − 0.1317955 0.1108473 13.61545


41 × 41 − 0.1311382 0.1099841 13.95723
61 × 61 − 0.1304138 0.1096524 14.25467
81 × 81 − 0.1293423 0.1151547 13.78368
101 × 101 − 0.1271867 0.1296093 12.2746

Table 1
Thermo-physical properties of water and nanoparticles [10]. Table 3
Comparisons of − ψmin for classical fluids.
Pure Copper Silver Alumina Titanium Oxide
water (Cu) (Ag) Al2O3 (TiO2) Calculation method Gr = 0, Re = 100 Gr = 104Re = 100
−3
ρ(kgm ) 997.1 8933 10500 3970 4250 Finite difference [19] 0.1013 –
Cp(Jkg− 1 K− 1) 4179 385 235 765 686.2 Multigrid [20] 0.103423 –
k(Wm− 1 K− 1) 0.613 401 429 40 8.9538 Spline [18] 0.1054 0.0934
β(K− 1) 21 × 10− 5 1.67×10− 5 1.89×10− 5 0.85×10− 5 0.9 × 10− 5 Finite difference (Present study) 0.1042828 0.0913895
1506 M.A. Mansour et al. / International Communications in Heat and Mass Transfer 37 (2010) 1504–1512

Table 4 In addition, (ρβ)nf is the thermal expansion coefficient of the


Comparisons of Num for Cu–water, Ag–water, Al2O3–water and TiO2–water nanofluid nanofluid and it can be determined by:
(Natural convection and ϕ = 0.1, B = 0.4, D = 0.5, Ra = 103).

Cu Ag Al2O3 TiO2 ðρβÞnf = ð1−ϕÞðρβÞf + ϕðρβÞp : ð8Þ


Aminossadatia and Ghasemi [10] 5.451 5.451 5.391 5.189
Moreover, μ nf is the effective dynamic viscosity of the nanofluid
Present study 5.459225 5.459674 5.398933 5.197152
which given by Brinkman [21]:
μf
and ϕ is the solid volume fraction of nanoparticles. Also, αnf is the μ nf = : ð9Þ
ð1−ϕÞ2:5
thermal diffusivity of the nanofluid and it is given by:
 
Following Ho et al. [15], Oztop and Abu-nada [11] and Aminossadati
αnf = knf = ρCp ð6Þ
nf and Ghasemi [10], the thermal conductivity of the nanofluid knf which
where, (ρCp)nf is the heat capacitance of the nanofluid and it can be appears in Eq. (6) can be expressed as:
expressed as: 2   3
kp + 2kf −2ϕ kf −kp
      4
knf = kf    5 ð10Þ
ρCp = ð1−ϕÞ ρCp + ϕ ρCp : ð7Þ kp + 2kf + ϕ kf −kp
nf f p

1 ψmin = -0.3135537 1
0.9 0.9
θ max = 0.059702
0.8 0.8
0.7 0.7
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0.2 0.2
0.1 0.1
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

1 ψmin = -0.2865524 1
0.9 0.9
θ max = 0.0584258
0.8 0.8
0.7 0.7
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0.2 0.2
0.1 0.1
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

1 ψmin = -0.2561652 1
0.9 0.9
θ max = 0.0580423
0.8 0.8
0.7 0.7
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0.2 0.2
0.1 0.1
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Fig. 2. The contours of streamlines (left) and isotherms (right) for cavity filled with Cu–water nanofluid at φ = 0, 0.1, 0.2. (increasing from top towards bottom). The referenced case
is Ra = 104, Re = 10, B = 0.4, D = 0.5.
M.A. Mansour et al. / International Communications in Heat and Mass Transfer 37 (2010) 1504–1512 1507

1 ψmin = -0.2378142 1
0.9 0.9
θ max = 0.0396407
0.8 0.8
0.7 0.7
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0.2 0.2
0.1 0.1
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

1 ψmin = -0.2946414 1
0.9 0.9
θ max = 0.0584516
0.8 0.8
0.7 0.7
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0.2 0.2
0.1 0.1
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

1 ψmin = -0.3327797 1
0.9 0.9
θ max = 0.1051025
0.8 0.8
0.7 0.7
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0.2 0.2
0.1 0.1
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Fig. 3. The contours of streamlines (left) and isotherms (right) for cavity filled with Al2O3–water nanofluid at B = 0.2, 0.4, 0.8 (increasing from top towards bottom). The referenced
case is Ra = 104, Re = 10, D = 0.5, ϕ = 0.1.

" #
where, kp is the thermal conductivity of dispersed nanoparticles and ∂U ∂U ∂P 1 ρf 1 θ2 U θ2 U
U +V =− + + ð13Þ
kf is the thermal conductivity of pure fluid. ∂X ∂Y ∂X Re ρnf ð1−ϕÞ2:5 ∂X 2 ∂Y 2
Introducing the following dimensionless parameters:

x y u v T−Tc p " #
X= ;Y = ;U = ;V = ;θ = ;P = ;
1 ρf
2 2
H H U0 U0 ΔT ρnf U02 ∂V ∂V ∂P 1 ∂ V ∂ V
U +V =− + +
ð11Þ ∂X ∂Y ∂Y Re ρnf ð1−ϕÞ2:5 ∂X 2 ∂Y 2
q″ H gβf H 3 ΔT ρf U0 H υf
ΔT = ; Ra = ; Re = ; Pr = : ð14Þ
kf υf α f μf αf
ðρβÞnf RaPr
+ θ
ρnf βf Re2
By using Eq. (11), Eqs.(1)–(4) are converted to the following
dimensionless form:
" #
∂θ ∂θ αnf 1 ∂2 θ ∂2 θ
∂U ∂V U +V = + : ð15Þ
+ =0 ð12Þ ∂X ∂Y αf RePr ∂X 2 ∂Y 2
∂X ∂Y
1508 M.A. Mansour et al. / International Communications in Heat and Mass Transfer 37 (2010) 1504–1512

1 ψmin = -0.217599 1
0.9 0.9
θ max = 0.0588509
0.8 0.8
0.7 0.7
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0.2 0.2
0.1 0.1
0 1 0 1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

1 ψmin = -0.27639 1
0.9 0.9
θ max = 0.0372559
0.8 0.8
0.7 0.7
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0.2 0.2
0.1 0.1
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

1 ψmax = 0.1497968 1
0.9 0.9
ψmin = -0.1120793
0.8 0.8
θ max = 0.0622308
0.7 0.7
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0.2 0.2
0.1 0.1
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Fig. 4. The contours of streamlines (left) and isotherms (right) for cavity filled with Ag–water nanofluid at D = 0.2, 0.5, 0.8 (increasing from top towards bottom). The referenced case
is Ra = 1.4 × 104, Re = 10, B = 0.2, ϕ = 0.05.

The dimensionless form of the boundary conditions is given by: The heat transfer coefficients in terms of the local Nusselt number
along the heat source surface is defined by

X = 0; 0≤Y≤1 : U = V = 0; θ = 0
1
Nus ð X Þ = ð17Þ
θs ð X Þ
X = 1; 0≤Y≤1 : U = V = 0; θ = 0

∂θ where, θs is the dimensionless heat source temperature. The average


Y = 0; 0≤XbðD−0:5BÞ : U = V = 0; = 0;
∂Y Nusselt number (Num) is determined by integrating Nus along the heat
ð16Þ
∂θ kf source.
Y = 0;ðD−0:5BÞ≤X≤ðD + 0:5BÞ : U = V = 0; = ;
∂Y knf
∂θ
Y = 0;ðD + 0:5BÞbX≤1 : U = V = 0; = 0;
∂Y D + 0:5B
1
Num = ∫ Nus ð X Þdx ð18Þ
Y = 1; 0≤X≤1 : U = 1; V = 0; θ = 0: B D−0:5B
M.A. Mansour et al. / International Communications in Heat and Mass Transfer 37 (2010) 1504–1512 1509

1 ψmin = -0.8159643 1
0.9 0.9
θ max = 0.1364358
0.8 0.8
0.7 0.7
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0.2 0.2
0.1 0.1
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

1 ψmin = -0.1311382 1
0.9 0.9
θ max = 0.1099841
0.8 0.8
0.7 0.7
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0.2 0.2
0.1 0.1
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

1 ψmin = -0.1084029 1
0.9 0.9
θ max = 0.0859788
0.8 0.8
0.7 0.7
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0.2 0.2
0.1 0.1
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Fig. 5. The contours of streamlines (left) and isotherms (right) for cavity filled with Cu–water nanofluid at Re = 1, 10, 100 (increasing from top towards bottom). The referenced case
is Ra = 103, ϕ = 0.1, B = 0.4, D = 0.5.

3. Numerical method and validation 41 × 41, 61 × 61, 81 × 81, 101 × 101. There is a good agreement found
between 41 × 41 and 61 × 61 grids, so the numerical computations
The numerical algorithm used to solve the dimensionless governing were carried out for (41 × 41) and (61 × 61) grid nodal points. This
Eqs. (12)–(15) with the boundary conditions (Eq. (16)) is based on the method was found to be suitable and gave results that are very close
finite difference methodology. Central difference quotients were used to to the numerical results obtained by Hsu et al. [18], Rubin and Khosla
approximate the second derivatives in both the X- and Y-directions. [19], and Ghia et al. [20] for classical fluid and Aminossadatia and
Then, By following Grosan et al. [22], the obtained discretized equations Ghasemi [10] for nanofluid. As we can see from Tables 3 and 4 the
are solved using a Gauss–Seidel iteration technique. The solution present results give a good agreement with Hsu et al. [18], Rubin and
procedure is iterated until the following convergence criterion is Khosla [19] and Ghia et al. [20] and a very good agreement with
satisfied: Aminossadatia and Ghasemi [10]. These favorable comparisons lend
confidence in the numerical results to be reported subsequently.
∑jχi;j −χi;j j≤10
new old −7
ð19Þ
i; j 4. Results and discussion

where χ is the general dependent variable. The numerical method In this section, numerical computations are carried out and a
was implemented in a FORTRAN software. Table 2 shows an accuracy parametric study is performed to illustrate the influence of the physical
test using the finite diference method using five sets of grids: 31 × 31, parameters on the resulting streamlines and isotherms contours as well
1510 M.A. Mansour et al. / International Communications in Heat and Mass Transfer 37 (2010) 1504–1512

55
1.0 50 Cu-water
B = 0.4
45 D = 0.5
0.5 Ra = 104
40
φ = 0, 0.05, 0.1, 0.2 Re = 10
35

Nus
0.0
V 30
Cu-water φ=0, 0.05, 0.1, 0.2
B = 0.4 25
-0.5
D = 0.5
20
Ra = 104
-1.0 Re = 10 15

10
0.0 0.2 0.4 0.6 0.8 1.0 0.3 0.4 0.5 0.6 0.7
X X

1.0 55 B = 0.2
0.8 Al2O3-water
50 B = 0.4
0.6 D = 0.5 B = 0.6
45
0.4 B = 0.2, 0.4, 0.6, 0.8 Ra = 104
0.2 40 B = 0.8
Re = 10
0.0 35 φ = 0.1

Nus
V

-0.2
30
-0.4 Al2O3-water
25
-0.6 D = 0.5
-0.8 Ra = 104 20
-1.0 Re = 10 15
-1.2 φ = 0.1
10
-1.4
0.0 0.2 0.4 0.6 0.8 1.0 0.2 0.4 0.6 0.8 1.0
X X

1.0
55 D = 0.2
0.8 D = 0.5
Ag-water
0.6 D=0.2, 0.5 50 B = 0.4
0.4 45 Ra = 104
D=0.8 Re = 10 D = 0.8
0.2 40
φ = 0.1
Nus

0.0
V

35
-0.2
Ag-water 30
-0.4
B = 0.4
25
-0.6 Ra = 104
-0.8 Re = 10 20
φ = 0.1 15
-1.0
0.0 0.2 0.4 0.6 0.8 1.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
X X

Fig. 6. Effect of the solid volume fraction (top), different values of the heat source length (center) and different positions of the heat source (bottom) on the vertical velocity
component at the cavity mid-section and Nusselt number along the heat source.

as the vertical velocity component at the enclosure mid-section, Nusselt − 0.2561652 at ϕ = 0.1 and ϕ = 0.2, respectively. Regarding isotherms
number and average Nusselt number along the heat source. The results contours, when there is no concentration of nanoparticles in the base
of this parametric study are shown in Figs. 2–6 and Tables 2–6 for a fluid (i.e. ϕ = 0) the isotherms lines distribute along the whole cavity
range of Reynolds number (1 ≤ Re ≤ 100), solid volume fraction and take the maximum value 0.059702. By adding the nanoparticles,
(0 ≤ ϕ ≤ 0.2), the different values of the heat source length this causes in the reduction of the fluid temperature and the maximum
(0.2 ≤ B ≤ 0.8) and the different positions of the heat source value of the temperature becomes 0.0580423 at ϕ = 0.2.
(0.2 ≤ D ≤ 0.8). Fig. 3 displays effect of the variations of heat source length
Fig. 2 shows effect of the solid volume fraction ϕ on the contours of represented by parameter B on the contours of streamlines and
streamlines and isotherms for cavity filled with Cu–water nanofluid. It is isotherms for cavity filled with Al2O3–water nanofluid. It is clear that,
found that, the fluid moves from bottom towards top and forms a increasing the heat source length, leads to increasing activity of the fluid
clockwise circular cell with ψmin = − 0.3135537 at ϕ = 0. This cell was motion. This increasing relates to the association between the buoyancy
generated by the lid. Also, presence of a heat source at the bottom wall of forces generated due to the fluid temperature differences and activity of
the cavity helps the cell to stretch horizontally beside position of the the fluid motion. On the other hand, as the length of the heat source
heat source. Increasing in solid volume fraction causes in a decreasing increases, the generation rates of the temperature increase. So the
the intensity of buoyancy and hence the flow intensity so that, the higher temperature patterns correspond to the taller heat source and
minimum value of stream function changes from − 0.2865524 to the opposite is valid.
M.A. Mansour et al. / International Communications in Heat and Mass Transfer 37 (2010) 1504–1512 1511

Table 5 opposite behavior when the length of the heat source increases. On the
Effects of the different values of the heat source length and solid volume fraction on other hand, as the heat source moves away from the left wall of the
average Nusselt number Num for different nanoparticles at (Re = 10, Ra = 104 , D = 0.5).
cavity the vertical velocity component increases. However, a further
B ϕ= 0 ϕ= 0.05 ϕ= 0.1 ϕ= 0.2 increase in D leads to a decrease in the vertical velocity component.
0.2 Cu 29.3655 30.31618 31.21571 32.44886 Regarding the Nusselt number along the heat source, increasing the
Ag 29.3655 30.06877 30.76026 31.66153 concentration of the nanoparticles in the base fluid results in an increase
Al2O3 29.3655 30.61812 31.63786 32.43085 in rate of the heat transfer along the heat source. In addition, an inverse
TiO2 29.3655 30.2766 30.9994 31.47948
relation between the mean values of rate of heat transfer and length of
0.8 Cu 14.67008 15.29119 15.92657 16.91041
Ag 14.67008 15.1598 15.71438 16.60303 the heat source is observed which means that the average Nusselt
Al2O3 14.67008 15.45061 16.09925 16.87992 number decreases with the increasing length of the heat source (see
TiO2 14.67008 15.2714 15.76723 16.27051 Table 5). Other observations can be seen, the heat source gives large
rates of the heat transfer when it arises in the left half of the bottom wall.
Also, at these positions, the Nusselt number increases with the
There is an interesting behavior in Fig. 4 which displays the effect of increasing length of the heat source. On the contrary, when the heat
changing the heat source location on the streamlines and isotherms source is put in the right half of the bottom wall, it gives low values of
contours for cavity filled with Ag–water nanofluid. It is observed that, Nusselt number and the Nusselt number decreases with the increasing
the fluid follows the geometry of the cavity by forming one clockwise length of the heat source.
circular cell with ψmin = -0.217599 atD = 0.2. As the heat source moves Tables 5 and 6 shows the different values of average Nusselt
away from the left wall of the cavity, the intensity of the flow becomes number Num for different nanoparticles. It is observed that, the large
complex and two clockwise and anticlockwise circular cells were values of average Nusselt number can be obtained by adding alumina
formed inside the cavity with ψmax = 0.1497968 and ψmin = -0.1120793. (Al2O3) nanoparticles whereas; by adding titanium oxide (TiO2)
This means that, for the practical applications, position of the heat nanoparticles, this gives small values of average Nusselt number. On
source must be carefully determined because it has a slightly effect on the other hand, average Nusselt number increases by increasing either
the fluid flows. Also, the fluid temperature is influenced by changing the solid volume fraction or Reynolds number whereas it decreases by
heat source location. It follows the movement of the heat source. increasing the heat source length. All these behaviors are presented in
Moreover, the higher generation rates of the temperature can be Tables 5 and 6 with the referenced caseRa = 104, D = 0.5.
obtained by putting the heat source beside the right wall and the lower
generation rates of the temperature can be obtained by putting the heat 5. Conclusion
source in center of the bottom wall.
Fig. 5 shows the effect of Reynolds number Re on the contours of In the present paper, a numerical simulation of mixed convection
streamlines and isotherms for the cavity filled with Cu–water flows in a square lid-driven cavity partially heated from below using
nanofluid. It is found that, for small values of Reynolds number Cu–water, Ag–water, Al2O3–water and TiO2–water nanofluid was
(Re = 1) the lid-driven effect is insignificant and the intensity of the studied. The finite difference method was employed for the solution of
flow are concentrated beside the heat source position. However, as the present problem. Comparisons with previously published work on
Reynolds number increases, the lid-driven effect increases and hence special cases of the problem were performed and found to be in good
forced convection flow. In addition, for the high values of Reynolds agreement. Graphical and tabular results for various parametric
number (Re = 100), presence of the heat source has no noticeable conditions were presented and discussed. From this investigation, we
effect on the streamlines. So the fluid motion concentrated at top of can draw the following conclusions:
the cavity and the motion of the upper lid stretch the streamlines
towards the left wall of the cavity. The isothermal contours distribute I. Increasing in solid volume fraction leads to decrease both of
inside the whole cavity for low Reynolds number (Re = 1) with activity of the fluid motion and fluid temperature however, it
θmax = 0.1364358 but as the Reynolds number increases, the fluid leads to increase the corresponding average Nusselt number.
temperature decreases and isotherm contours concentrate beside the II. As length of the heat source increases, this leads to not only
left wall. These behaviors are clearly shown in Fig. 5 with Ra = 103, increasing the flow intensity but also increasing the fluid
ϕ = 0.1, B = 0.4, D = 0.5. temperature. However, it decreases the corresponding average
Fig. 6 displays the effects of solid volume fraction ϕ, different values Nusselt number.
of the heat source length B and different positions of the heat source D III. The large values of Reynolds number lead to increase the lid-
on the vertical velocity component at the enclosure mid-section V driven effect whereas the small values of Reynolds number lead
(Y = 0.5) and Nusselt number along the heat source for Cu–water, to increase effect of presence of the heat source on the flow and
Ag–water and Al2O3–water nanofluid, respectively. The results show heat characteristics.
that, in general, upward flow and downward flow are symmetric with IV. Putting heat source at the left half of the bottom wall gives
respect to the center of the cavity. Moreover, as the solid volume fraction large rates of heat transfer but when it is moved to the right half
increases the vertical velocity component decreases whereas, it takes an of the bottom wall, it gives small rates of the heat transfer.
V. By adding alumina (Al2O3) nanoparticles to the base fluid, this
Table 6 gives large values of Nusselt number. On the contrary, by
Effects of Reynolds number and the different positions of the heat source on average adding titanium oxide (TiO2) nanoparticles to the base fluid
Nusselt number Num for different nanoparticles at (B = 0.4, Ra = 104, ϕ = 0.1). this gives small values of Nusselt number.
Re D = 0.2 D = 0.5 D = 0.8

1 Cu 29.51092 19.81568 20.0475 References


Ag 29.32486 19.74269 19.87845
Al2O3 29.55065 19.5014 20.19601 [1] Y. Xuan, Q. Li, Investigation on convective heat transfer and flow features of
TiO2 28.6872 19.26302 19.66946 nanofluids, Journal of Heat Transfer 125 (2003) 151–155.
10 Cu 32.02906 23.47167 18.83937 [2] S. Choi, Enhancing thermal conductivity of fluids with nanoparticles, ASME
Ag 31.7597 23.13634 18.65713 Publications 66 (1995) 99–105.
Al2O3 32.12183 23.79216 19.0212 [3] Y. Xuan, K. Yu, Q. Li, Investigation on flow and heat transfer of nanofluids by the
TiO2 31.27511 23.29148 18.56293 thermal Lattice Boltzmann model, Progress in Computational Fluid Dynamics 5
(2005) 13–19.
1512 M.A. Mansour et al. / International Communications in Heat and Mass Transfer 37 (2010) 1504–1512

[4] P. Keblinski, S.R. Phillpot, S.U.S. Choi, J.A. Eastman, Mechanisms of heat flow in [14] A.K. Santra, S. Sen, N. Chakraborty, Study of heat transfer characteristics of
suspensions of nano-sized particles (nanofluids), International Journal of Heat copper–water nanofluid in a differentially heated square cavity with different
and Mass Transfer 45 (2002) 855–863. viscosity models, Journal Enhanced Heat Transfer 15 (2008) 273–287.
[5] S.K. Das, N. Putra, P. Thiesen, W. Roetzel, Temperature dependence of thermal [15] C.J. Ho, M.W. Chen, Z.W. Li, Numerical simulation of natural convection of
conductivity enhancement for nanofluids, Journal of Heat Transfer 125 (2003) 567–574. nanofluid in a square enclosure: effects due to uncertainties of viscosity and
[6] N. Putra, W. Roetzel, S.K. Das, Natural convection of nanofluids, Heat and Mass thermal conductivity, International Journal of Heat and Mass Transfer 51 (2008)
Transfer 39 (2003) 775–784. 4506–4516.
[7] D. Wen, Y. Ding, Natural convective heat transfer of suspensions of titanium [16] K.S. Hwang, J.H. Lee, S.P. Jang, Buoyancy-driven heat transfer of water-based
dioxide nanoparticles (nanofluids), IEEE Transport Nanotechnology 5 (2006) Al2O3 nanofluids in a rectangular cavity, International Journal of Heat and Mass
220–227. Transfer 50 (2007) 4003–4010.
[8] S. Kumar, S.K. Prasad, J. Banerjee, Analysis of flow and thermal field in nanofluid [17] E. Abu-Nada, Z. Masoud, H. Oztopd, A. Campo, Effect of nanofluid variable
using a single phase thermal dispersion model, Applied Mathematical Modeling. properties on natural convection in enclosures, International Journal of Thermal
34 (2010) 573–592. Sciences 49 (2010) 479–491.
[9] F. Talebi, A.H. Mahmoudi, M. Shahi, Numerical study of mixed convection flows in [18] T.-H. Hsu, P.-T. Hsu, C.-K. Chen, Thermal convection of micropolar fluids in a lid-driven
a square lid-driven cavity utilizing nanofluid, International Communications in cavity, International Communications in Heat and Mass Transfer 22 (1995) 189–200.
Heat and Mass Transfer 37 (2010) 79–90. [19] S.G. Rubin, P.K. Khosla, Polynomial interpolation methods for viscous flow
[10] S.M. Aminossadatia, B. Ghasemi, Natural convection cooling of a localized heat calculations, Journal of Computational Physics 24 (1977) 217–244.
source at the bottom of a nanofluid-filled enclosure, European Journal of [20] U. Ghia, K.N. Ghia, C.T. Shin, High-Re solutions for incompressible flow using the
Mechanics B/Fluid 28 (2009) 630–640. Navier–Stokes equations and a multigrid method, Journal of Computational
[11] H.F. Oztop, E. Abu-Nada, Numerical study of natural convection in partially heated Physics 48 (1982) 387–411.
rectangular enclosures filled with nanofluids, International Journal of Heat and [21] H.C. Brinkman, The viscosity of concentrated suspensions and solution, Journal of
Fluid Flow 29 (2008) 1326–1336. Chemical Physics 20 (1952) 571–581.
[12] N. Putra, W. Roetzel, S.K. Das, Natural convection of nanofluids, Heat and Mass [22] T. Grosan, C. Revnic, I. Pop, D.B. Ingham, Magnetic field and internal heat
Transfer 39 (2003) 775–784. generation effects on the free convection in a rectangular cavity filled with a
[13] D. Wen, Y. Ding, Formulation of nanofluids for natural convective heat transfer porous medium, International Journal of Heat and Mass Transfer 52 (2009)
applications, International Journal of Heat and Fluid Flow 26 (2005) 855–864. 1525–1533.

You might also like