Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

International Journal of Heat and Mass Transfer 46 (2003) 3639–3653

www.elsevier.com/locate/ijhmt

Buoyancy-driven heat transfer enhancement


in a two-dimensional enclosure utilizing nanofluids
a,b a,* b
Khalil Khanafer , Kambiz Vafai , Marilyn Lightstone
a
Mechanical Engineering Department, University of California, A363 Bourns Hall, Riverside, CA 92521-0425, USA
b
Mechanical Engineering Department, McMaster University, Hamilton, Canada L8S 4L7
Received 26 November 2002; received in revised form 8 March 2003

Abstract
Heat transfer enhancement in a two-dimensional enclosure utilizing nanofluids is investigated for various pertinent
parameters. A model is developed to analyze heat transfer performance of nanofluids inside an enclosure taking into
account the solid particle dispersion. The transport equations are solved numerically using the finite-volume approach
along with the alternating direct implicit procedure. Comparisons with previously published work on the basis of
special cases are performed and found to be in excellent agreement. The effect of suspended ultrafine metallic nano-
particles on the fluid flow and heat transfer processes within the enclosure is analyzed and effective thermal conductivity
enhancement maps are developed for various controlling parameters. In addition, an analysis of variants based on the
thermophysical properties of nanofluid is developed and presented. It is shown that the variances within different
models have substantial effects on the results. Finally, a heat transfer correlation of the average Nusselt number for
various Grashof numbers and volume fractions is presented.
Ó 2003 Elsevier Ltd. All rights reserved.

1. Introduction fluids with substantially higher conductivities to enhance


thermal characteristics. Small particles (nanoparticles)
Nanotechnology is considered by many to be one of stay suspended much longer than larger particles. If
the significant forces that drive the next major industrial particles settle rapidly (microparticles), more particles
revolution of this century. It represents the most rele- need to be added to replace the settled particles, result-
vant technological cutting edge currently being explored. ing in extra cost and degradation in the heat transfer
It aims at manipulating the structure of the matter at the enhancement. As such an innovative way in improving
molecular level with the goal for innovation in virtually thermal conductivities of a fluid is to suspend metallic
every industry and public endeavor including biological nanoparticles within it. The resulting mixture referred to
sciences, physical sciences, electronics cooling, trans- as a nanofluid possesses a substantially larger thermal
portation, the environment and national security. conductivity compared to that of traditional fluids [1].
Low thermal conductivity of conventional heat The presence of the nanoparticles in the fluids in-
transfer fluids such as water, oil, and ethylene glycol creases appreciably the effective thermal conductivity of
mixture is a primary limitation in enhancing the per- the fluid and consequently enhances the heat transfer
formance and the compactness of many engineering characteristics. Nanofluids have a distinctive character-
electronic devices. To overcome this drawback, there is a istic, which is quite different from those of traditional
strong motivation to develop advanced heat transfer solid–liquid mixtures in which millimeter and/or micro-
meter-sized particles are involved. Such particles can
clot equipment and can increase pressure drop due to
*
Corresponding author. Tel.: +1-909-787-2135; fax: +1-909- settling effects. Moreover, they settle rapidly, creating
787-2899. substantial additional pressure drop. However, nano-
E-mail address: vafai@engr.ucr.edu (K. Vafai). fluids exhibit little or no penalty in pressure drop when
0017-9310/03/$ - see front matter Ó 2003 Elsevier Ltd. All rights reserved.
doi:10.1016/S0017-9310(03)00156-X
3640 K. Khanafer et al. / International Journal of Heat and Mass Transfer 46 (2003) 3639–3653

Nomenclature

A aspect ratio, L=H bs solid expansion coefficient


cp specific heat at constant pressure / solid volume fraction
dp nanoparticle diameter mf kinematic viscosity
~
g gravitational acceleration vector h dimensionless temperature, ðT  TL Þ=
Gr Grashof number, gbf DTH 3 =m2f ðTH  TL Þ
H cavity height x vorticity
kf fluid thermal conductivity X dimensionless vorticity, pffiffiffiffiffiffiffiffiffiffiffiffi
xH ffi
gbf DTH 3
ks solid thermal conductivity w stream function
w
L cavity width W dimensionless stream function, pffiffiffiffiffiffiffiffiffiffiffiffi ffi
H gbf DTH 3
Nu average Nusselt number d variable used in Eq. (19)
Pr Prandtl number, mf =af q density pffiffiffiffiffiffiffiffiffiffiffiffiffi
t gbf DTH 3
Q total heat transfer from the left wall s dimensionless time, H
t time l dynamic viscosity
T temperature
Subscripts
U, V dimensionless interstitial velocity compo-
eff effective
nents
f fluid
u, v interstitial velocity components
H hot
x, y Cartesian coordinates
L cold
X, Y dimensionless coordinates
nf nanofluid
Greek symbols o reference value
a thermal diffusivity s solid
bf fluid thermal expansion coefficient

flowing through the passages. Moreover, they flow and the enhancement was proportional to the volume
smoothly through microchannels without clogging fraction of the solid phase. The dependence of the
them. Thus, nanofluids are best for applications in thermal conductivity of nanoparticle–fluid mixture on
which fluid flows through small passages because the base fluid was analyzed by Xie et al. [5].
nanoparticles are small enough to behave similar to When simulating heat transfer enhancement using
liquid molecules. Nanofluids show promise in signifi- nanofluids, modeling of the effective thermal conduc-
cantly increasing heat transfer rates in a variety of ap- tivity possesses a challenge. This can be attributed to
plications, with minimum pressure drop. Enhancements several factors such as gravity, Brownian motion, fric-
were recently reported for copper Cu nanofluids, where tion force between the fluid and the ultrafine solid
just a 0.3% volume fraction of 10 nm diameter copper particles, sedimentation, layering at the solid/liquid
Cu nanoparticles led to an increase of up to 40% in the interface, ballistic phonon transport through the parti-
thermal conductivity of ethylene glycol [2]. This can be cles and the clustering of nanoparticles. This implies that
attributed to several factors such as nanoparticle clus- the slip velocity between the fluid phase and the nano-
tering [3], ballistic phonon transport [3], layering at the particles is not zero, although the particles are ultrafine.
solid/liquid interface [3], the interaction and collision A body of theoretical work in the literature [6–8] is
among particles and surface area enhancement. In ad- available on the effective thermal conductivity of two-
dition, the suspended particles increase the surface area phase mixtures that contain powders with particle dia-
and the heat capacity of the fluid. That is, a significant meters in the order of millimeters or even micrometers
improvement in the effective thermal conductivity is since the first published theoretical work by Maxwell [9].
achieved as a result of decreasing the size of the sus- MaxwellÕs model predicted that the effective thermal
pended particles (nano-sized particle) rather than using conductivity of suspensions containing spherical parti-
larger particles (micro-sized particle). Since heat transfer cles increases with an increase in the volume fraction of
occurs on the surface of a solid, this feature greatly en- the solid particles. Hamilton and Crosser [10] investi-
hances the fluidÕs heat conduction contribution. Wang gated the possibility of increasing particle surface area
et al. [4] studied the thermal conductivity of nano-sized by controlling particle shapes to be non-spherical. Ap-
SiC suspensions using a transient hot-wire method. proximately, an order of magnitude improvement in
Their experimental results showed that the thermal surface area per particle volume was achieved experi-
conductivities of the studied suspensions were increased mentally using this approach alone. The authors devel-
K. Khanafer et al. / International Journal of Heat and Mass Transfer 46 (2003) 3639–3653 3641

oped an expression for the effective thermal conductivity the mechanisms of heat flow in suspensions of nano-
of two-component mixtures as a function of liquid and sized particles (nanofluids). Four possible explana-
solid particle thermal conductivities, particle volume tions were reported for an increase in the thermal
fraction, and an empirical scaling factor that takes into conductivity with decreasing grain size. They devel-
account the effect of different particle shapes on the ef- oped a fundamental understanding of heat transport
fective thermal conventional solid particles suffer from in solid nanoparticle colloids under stationery condi-
significant clogging problems due to their significant size tions.
conductivity. An alternative expression for calculating To the best knowledge of the authors, the problem of
the effective thermal conductivity of solid–liquid mix- buoyancy-driven heat transfer enhancement of nano-
tures with a sphericity of one was established by Wasp fluids in a two-dimensional enclosure has not been ana-
[11]. lyzed. This problem may be encountered in a number of
Two main approaches have been adopted in the lit- electronic cooling and MEMS applications. The present
erature to investigate the heat transfer enhancement by study is focused on the analysis of several pertinent
small solid particles (millimeter and/or micrometer-sized parameters on the heat transfer characteristics of nano-
particles) suspended in a fluid. The first approach is the fluids within the enclosure. The dispersion effect is
two-phase model, which enables a better understanding analyzed in the present investigation. Effective thermal
of both the fluid and the solid phases role in the heat conductivity maps will be developed in the present study
transfer process. The second approach is the single- for various pertinent parameters.
phase model in which both the fluid phase and the
particles are in thermal equilibrium state and flow with
the same local velocity. The latter approach is simpler 2. Mathematical formulation
and more computationally efficient. Several factors may
affect heat transfer enhancement using nanofluids. These Consider a two-dimensional enclosure of height H
factors include gravity, Brownian motion, layering at and width L filled with a nanofluid as shown in Fig. 1.
the solid/liquid interface, ballistic phonon transport The horizontal walls are assumed to be insulated, non-
through the particles, nanoparticles clustering, and the conducting, and impermeable to mass transfer. The
friction between the fluid and the solid particles. The nanofluid in the enclosure is Newtonian, incompressible,
phenomena of Brownian diffusion, sedimentation, and and laminar. The nanoparticles are assumed to have a
dispersion may coexist in the main flow of a nanofluid. uniform shape and size. Moreover, it is assumed that
In the absence of any experimental data and suitable both the fluid phase and nanoparticles are in thermal
theoretical studies in the literature to investigate these equilibrium state and they flow at the same velocity. The
factors, the existing macroscopic two-phase model is not left vertical wall is maintained at a high temperature
applicable for analyzing nanofluids. Accordingly the ðTH Þ while the right vertical wall is kept at a low tem-
modified single-phase, taking into the account some of perature ðTL Þ. The thermophysical properties of the
the above factors, is more convenient than the two- nanofluid are assumed to be constant except for the
phase model if the main interest is focused on the heat density variation in the buoyancy force, which is based
transfer process. Moreover, superior characteristics of on the Boussinesq approximation.
the nanofluid allow it to behave more like a fluid than The initial and boundary conditions for the present
the conventional solid–fluid mixtures. investigation are presented as
The chaotic movement of the ultrafine particles in-
creases the energy exchange rates in the fluid, i.e., ther-
mal dispersion takes place within the flow of the
nanofluid. To account for the random motion of the y
particles, dispersion model is implemented. So far, there
is a lack of theoretical and experimental works published
on the thermal diffusivity coefficients of nanofluids.
Thermal diffusivity coefficient for nanofluid can be
modeled similar to the thermal dispersion models for
g
flow through porous media. The dispersed model TL
TH H
was first applied by Taylor [12] to simulate salt diffu-
sion in water. Xuan and Li [13] presented a proce-
dure for preparing a nanofluid which is a suspension
L
consisting of nanophase powders and a base liquid. x
Later on, Xuan and Roetzel [14], analyzed theoreti-
cally the flow of a nanofluid inside a tube using a dis-
persion model. Recently, Keblinski et al. [3] investigated Fig. 1. Schematic for the physical model.
3642 K. Khanafer et al. / International Journal of Heat and Mass Transfer 46 (2003) 3639–3653

u ¼ v ¼ T ¼ 0 for t ¼ 0 ð1Þ The effective thermal conductivity of the nanofluid


3 may take the following form
u ¼ v ¼ oT
oy
¼0 at y ¼ 0; H and 0 6 x 6 L
T ¼ TH ; u ¼ v ¼ 0 at x ¼ 0; 0 6 y 6 H
7
5 keff ¼ ðkeff Þstagnant þ kd ð10Þ
T ¼ TL ; u ¼ v ¼ 0 at x ¼ L; 0 6 y 6 H
Therefore, the enhancement in the thermal conductivity
for t > 0 ð2Þ
due to the thermal dispersion is given as [16]
The governing equations for the present study taking
kd ¼ Cðqcp Þnf jV j/dp ð11Þ
into the account the above mentioned assumptions are
written in dimensional form as pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
where jV j ¼ u2 þ v2 and C is an unknown constant
Vorticity equation
which should be determined by matching experimental
ox ox ox l 1 data. The above equations can be cast in non-dimen-
þu þv ¼ eff r2 x þ ½/qs;o bs
ot ox oy qnf;o qnf;o sional form by incorporating the following dimension-
oT less parameters
þ ð1  /Þqf;o bf g ð3Þ
ox X ¼ Hx ; Y ¼ Hy ; U ¼ pffiffiffiffiffiffiffiffiffiffiffiffi u ffi;
9
gbf DTH 3
>
>
pffiffiffiffiffiffiffiffiffiffiffiffiffi
>
Energy equation 3
>
>
=
t gbf DTH
oT oT oT o

kd
 
oT V ¼ pffiffiffiffiffiffiffiffiffiffiffiffi3ffi ; s ¼
v
H
; X ¼ p xH
ffiffiffiffiffiffiffiffiffiffiffiffi
ffi ; ð12Þ
gbf DTH gbf DTH 3 >
þu þv ¼ anf þ >
ot ox oy ox ðqcp Þnf ox
>
w
W ¼ pffiffiffiffiffiffiffiffiffiffiffiffi ffi ; h ¼ TT T L
;
>
>
H TL
3
;
   H gbf DTH
o kd oT
þ anf þ ð4Þ
oy ðqcp Þnf oy oX oX oX
þU þV
os oX oY
Kinematics equation r2 X oh
¼ ipffiffiffiffiffiffi þ k ð13Þ
o2 w o2 w
h
2:5 qs;o
ð1  /Þ / qf;o þ ð1  /Þ Gr oX
þ ¼ x ð5Þ
ox2 oy 2
    
oh oh oh 1 o oh o oh
where anf ¼ ðkeff Þstagnant =ðqcp Þnf . þU þV ¼ pffiffiffiffiffiffi v þ v
os oX oY Pr Gr oX oX oY oY
The effective density of a fluid containing suspended
particles at a reference temperature is given by ð14Þ

qnf;o ¼ ð1  /Þqf;o þ /qs;o ð6Þ o2 W o2 W


þ ¼ X ð15Þ
oX 2 oY 2
where qf;o , qs;o , and / are the density of clear fluid,
density of the particles, and the volume fraction of the where
nanoparticles, respectively. The effective viscosity of a h
ðkeff Þstagnant
i
fluid of viscosity lf containing a dilute suspension of kf dp pffiffiffiffiffiffipffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
v¼ ðqc Þ
þ C/ Pr Gr U 2 þ V 2 ð16Þ
small rigid spherical particles is given by Brinkman [15] ð1  /Þ þ / ðqcpp Þs H
f
as
lf In the above equations, Gr ¼ gbf DTH 3 =v2f is the Grashof
leff ¼ ð7Þ number, Pr ¼ vf =af is the Prandtl number and / is the
ð1  /Þ2:5
volume fraction of the nanoparticles. The aspect ratio is
The heat capacitance of the nanofluid can be presented defined as A ¼ L=H and is assumed unity in this inves-
as tigation. The diameter of the nanoparticle dp is taken as
10 nm in the present study. The physical dimension of
ðqcp Þnf ¼ ð1  /Þðqcp Þf þ /ðqcp Þs ð8Þ
the enclosure H is chosen to be 1 cm.
The effective stagnant thermal conductivity of the solid– The coefficient k that appears next to the buoyancy
liquid mixture was introduced by Wasp [11] as follows term is given as

ðkeff Þstagnant ks þ 2kf  2/ðkf  ks Þ


2 3
¼ ð9Þ 1 bs 1 5 ¼ bnf
kf ks þ 2kf þ /ðkf  ks Þ k¼4 q þ / qs;o ð17Þ
1 þ ð1/Þ f;o bf 1 þ ð1/Þ
/ qs;o q
bf
f;o
This equation is applicable for the two-phase mixture
containing micro-sized particles. In the absence of any The Nusselt number of the nanofluids is expected to
convenient formula for the calculations of the stagnant depend on a number of factors such as thermal con-
thermal conductivity of nanofluids, Eq. (9) may ap- ductivity and heat capacitance of both the pure fluid and
proximately apply to obtain a reasonable estimation. ultrafine particles, the volume fraction of the suspended
K. Khanafer et al. / International Journal of Heat and Mass Transfer 46 (2003) 3639–3653 3643

particles, the dimensions of these particles, flow struc- 3. Numerical method


ture, and the viscosity of the nanofluid. The local vari-
ation of the Nusselt number of the nanofluid can be The governing equations (13)–(15) were discretized
expressed as using a finite volume approach [17]. A brief description
ðkeff Þstagnant oh of the numerical approach is presented here. The gov-
Q
Nu ¼ ¼ ð18Þ erning equations can be represented by a general dif-
Qcond;fluid kf oX ferential equation as follows
where
   
ou o ou o ou
du þ U u  Cu þ V u  Cu ¼ Su
oT os oX oX oY oY
Q ¼ ðkeff Þstagnant A j
ox x¼0 ð19Þ

where u stands for either X or h with

1
dX ¼ 1;
31× 31 1 oh
41× 41 CX ¼ h ipffiffiffiffiffiffi ; SX ¼ k
0.8 2:5qs;o
ð1  /Þ / qf;o þ ð1  /Þ Gr oX
61× 61
81× 81
ð20Þ
Temperature

0.6

v
0.4
dh ¼ 1; Ch ¼ pffiffiffiffiffiffi ;
Pr Gr
 
1 ov oh ov oh
0.2 Sh ¼ pffiffiffiffiffiffi þ ð21Þ
Pr Gr oX oX oY oY
0 The transient finite difference equations, Eqs. (13)
0 0.2 0.4 0.6 0.8 1
X
and (14), were solved using an alternating direct implicit
(ADI) algorithm in conjunction with the power-law
1

0.8

41 × 41
0.6
Y

0.4
31 ×31
0.2
61× 61, 81 ×81
0
-0.05 -0.04 -0.03 -0.02 -0.01 0 0.01 0.02 0.03 0.04 0.05
U-Velocity

0.1

0.06 41× 41
61 ×61
81 × 81
V-Velocity

0.02

-0.02

31× 31
-0.06

-0.1
0 0.2 0.4 0.6 0.8 1
X
Fig. 3. Comparison of the streamlines and the isotherms be-
Fig. 2. Velocity and temperature profiles at mid-sections of the tween the present work and that of Fidap [18] (Pr ¼ 0:7,
cavity for various mesh sizes (Gr ¼ 105 , Pr ¼ 6:2, / ¼ 5%). Ra ¼ 103 ).
3644 K. Khanafer et al. / International Journal of Heat and Mass Transfer 46 (2003) 3639–3653

technique [17]. In addition, false transient accelerator for various Rayleigh numbers. Comparison of the so-
was implemented to expedite the convergence rate of the lution with previous works for different Rayleigh num-
solution towards steady state condition. Furthermore, bers is shown in Table 1. The comparison is concerned
successive over relaxation (SOR) method was applied to with the average Nusselt number along the hot wall,
solve for flow kinematics, as described by Eq. (15). maximum and minimum velocity values and their cor-
The vorticity on the boundaries is presented from its responding locations. This table shows an excellent
definition in terms of the primitive velocity variables as agreement between the present results and other bench-
mark solutions. Moreover, the present numerical code
ð4Ui;2 þ Ui;3 Þ ð4Ui;N 1  Ui;N 2 Þ >
9
was also validated against the experimental results of
Xi;1 ¼ ; Xi;N ¼ ;=
2DY 2DY Krane and Jessee [23] for natural convection in an en-
ð4V2;j  V3;j Þ ð4VM1;j þ VM2;j Þ > closure filled with air as shown in Fig. 7. It can be seen
X1;j ¼ ; XM ;j ¼ ;
2DX 2DX from the comparison that both solutions are in a very
ð22Þ good agreement.

To test and assess grid independence of the solution


scheme, numerical experiments were performed as 4. Discussion
shown in Fig. 2. These experiments show that an equally
spaced grid mesh of 61 61 is adequate to describe the The numerical code developed in the present inves-
flow and heat and mass transfer processes correctly. tigation is used to carry out a number of simulations for
Further increase in the number of grid points produced a wide range of controlling parameters such as Grashof
essentially the same results. The validation of our in- number and the volume fraction of particles. The range
house numerical code was performed against the results of the Grashof number Gr for this investigation is varied
generated by a commercial package [18] for pure fluid as between 103 6 Gr 6 105 . The range of the volume frac-
shown in Figs. 3–6. It can be seen from these figures that tion / used in this study is varied between 0 6 / 6 25%.
the solution of the present numerical code is in excellent The thermophysical properties of fluid and the solid
agreement with the numerical results from FIDAP [18] phases are shown in Table 2.

Fig. 4. Comparison of the streamlines and the isotherms be- Fig. 5. Comparison of the streamlines and the isotherms be-
tween the present work and that of Fidap [18] (Pr ¼ 0:7, tween the present work and that of Fidap [18] (Pr ¼ 0:7,
Ra ¼ 104 ). Ra ¼ 105 ).
K. Khanafer et al. / International Journal of Heat and Mass Transfer 46 (2003) 3639–3653 3645

1 1

__ Present result Ra=10 3 Ra=103 __ Present result


0.8 _ _ Fidap [18] 0.6
4
_ _ Fidap [18]
Ra=10
Ra=105

Temperature
0.6 0.2

Ra=104
Y

0.4 -0.2
Ra=105
0.2 -0.6

X=0.5 Y=0.5
0 -1
-0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8 1
Temperature X

1
Ra=104 0.5
__ Present result __ Present result
_ _ Fidap [18] Ra=104
0.8 _ _ Fidap [18]
Ra=105 0.25
Ra=103
0.6
Ra=105

V-Velocity
Y

3 0
Ra=10
0.4

-0.25
0.2

X=0.5 Y=0.5
0 -0.5
-0.3 -0.2 -0.1 0 0.1 0.2 0.3 0 0.2 0.4 0.6 0.8 1
U-Velocity X

Fig. 6. Comparison of the temperature and velocity profiles at the mid-sections of the cavity between the present results and that of
Fidap [18] ðPr ¼ 0:7Þ.

Table 1
Comparison of laminar solution with previous works for different Ra-values
Present Barakos and Markatos and De Vahl Davis [21] Fusegi et al. [22]
Mitsoulis [19] Pericleous [20]
Ra ¼ 103
Nu 1.118 1.114 1.108 1.118 1.105
Umax (at y=H) 0.137 (0.812) 0.153 (0.806) – (0.832) 0.136 (0.813) 0.132 (0.833)
Vmax (at x=H ) 0.139 (0.173) 0.155 (0.181) – (0.168) 0.138 (0.178) 0.131 (0.200)
Ra ¼ 104
Nu 2.245 2.245 2.201 2.243 2.302
Umax (at y=H) 0.192 (0.827) 0.193 (0.818) – (0.832) 0.192 (0.823) 0.201 (0.817)
Vmax (at x=H ) 0.233 (0.123) 0.234 (0.119) – (0.113) 0.234 (0.119) 0.225 (0.117)
Ra ¼ 105
Nu 4.522 4.510 4.430 4.519 4.646
Umax (at y=H) 0.131 (0.854) 0.132 (0.859) – (0.857) 0.153 (0.855) 0.147 (0.855)
Vmax (at x=H ) 0.258 (0.065) 0.258 (0.066) – (0.067) 0.261 (0.066) 0.247 (0.065)
Ra ¼ 106
Nu 8.826 8.806 8.754 8.799 9.012
Umax (at y=H) 0.077 (0.854) 0.077 (0.859) – (0.872) 0.079 (0.850) 0.084 (0.856)
Vmax (at x=H ) 0.262 (0.039) 0.262 (0.039) – (0.038) 0.262 (0.038) 0.259 (0.033)

To show that nanofluids behave more like a fluid various Grashof numbers and volume fractions as
than the conventional solid–fluid mixture, a comparison shown in Fig. 8. This figure shows that the nanofluid
of the temperature and the velocity profiles is conducted behaves more like a fluid than the conventional solid–
inside a thermal cavity with isothermal vertical walls at fluid mixtures in which relatively larger particles with
3646 K. Khanafer et al. / International Journal of Heat and Mass Transfer 46 (2003) 3639–3653

1 millimeter or micrometer orders are suspended for var-


ious Grashof number. Fig. 8 illustrates the effect of
0.8 Grashof number and the volume fraction on the tem-
perature and the velocity profiles at the mid-sections of
the cavity for water with a Prandtl number of 6.2. The
Temperature

0.6

numerical results of the present study indicate that the


0.4 heat transfer feature of a nanofluid increases remarkably
with the volume fraction of nanoparticles. As the vol-
0.2 ume fraction increases, irregular and random move-
ments of particles increases energy exchange rates in the
0
0 0.2 0.4 0.6 0.8 1
X 1
__ nanofluid
Gr =104 _ _ pure fluid
1 0.8
φ =0, 0.1, 0.2
Gr =105

Temperature
0.8 0.6

0.6 0.4
φ =0.2, 0.1,0
Y

0.4 0.2

0.2 0
0 0.2 0.4 0.6 0.8 1
X
0
-0.2 -0.15 -0.1 -0.05 0 0.05 0.1 0.15 0.2
1
U-Velocity __ nanofluid
_ _ pure fluid φ =0.2, 0.1, 0
0.8
0.4
Experimental data [23]
0.3
0.6
φ =0, 0.1, 0.2
Gr =104
0.2
Y

Present numerical result


0.1 0.4
V-Velocity

0
0.2
-0.1 Gr =105

-0.2 0
-0.08 -0.06 -0.04 -0.02 0 0.02 0.04 0.06 0.08
-0.3
U-Velocity

-0.4
0 0.2 0.4 0.6 0.8 1 0.15
__ nanofluid
X Gr =104 _ _ pure fluid
0.1
Fig. 7. Comparison of the temperature and the velocity profiles φ =0, 0.1, 0.2
inside a thermal cavity with isothermal vertical walls between 0.05
V-Velocity

the present results and the experimental results by Krane and Gr =105
Jessee [23] (Ra ¼ 1:89 105 , Pr ¼ 0:71). 0

φ =0.2, 0.1, 0
-0.05

Table 2
-0.1
Thermophysical properties of different phases
Property Fluid phase Solid phase -0.15
(water) (copper) 0 0.2 0.4 0.6 0.8 1

cp (J/kg K) 4179 383 X

q (kg/m3 ) 997.1 8954


Fig. 8. Comparison of the temperature and velocity profiles
k (W/m K) 0.6 400 between nanofluid and pure fluid for various Grashof numbers
b (K1 ) 2.1 104 1.67 105 (Pr ¼ 6:2, / ¼ 10% and 20%).
K. Khanafer et al. / International Journal of Heat and Mass Transfer 46 (2003) 3639–3653 3647

fluid and consequently enhances the thermal dispersion havior is also present for a single-phase flow. As the
in the flow of nanofluid. In addition, the velocities at the volume fraction increases, the velocity components of
center of the cavity for higher values of Grashof number nanofluid increase as a result of an increase in the energy
are very small compared with those at the boundaries transport through the fluid. High velocity peaks of the
where the fluid is moving at higher velocities. This be- vertical velocity component are shown in this figure at

Fig. 9. Streamlines contours and isotherms at various void fractions (Gr ¼ 104 , Pr ¼ 6:2).
3648 K. Khanafer et al. / International Journal of Heat and Mass Transfer 46 (2003) 3639–3653

high volume fractions. The effect of an increase in the The effect of the volume fraction on the streamlines
volume fraction on the velocity and temperature gradi- and isotherms of nanofluid for various Grashof numbers
ents along the centerline of the cavity is shown in Fig. 8. is shown in Figs. 9 and 10. In the absence of nanopar-

Fig. 10. Streamlines contours and isotherms at various void fractions (Gr ¼ 105 , Pr ¼ 6:2).
K. Khanafer et al. / International Journal of Heat and Mass Transfer 46 (2003) 3639–3653 3649

ticles and for a low Grashof number ðGr ¼ 103 Þ, a the fluid flow. As the Grashof number increases, as
central vortex appears as a dominant characteristic of shown in Figs. 9 and 10 ð/ ¼ 0Þ, the central vortex tends

Fig. 11. Comparison of the streamlines and isotherms contours between nanofluid (––) and pure fluid (- - -) at various Grashof
numbers ð/ ¼ 10%Þ.
3650 K. Khanafer et al. / International Journal of Heat and Mass Transfer 46 (2003) 3639–3653

to become elliptic for Gr ¼ 104 and eventually breaks up increase with an increase in the volume fraction as a
into three vortices for a Grashof number of Gr ¼ 105 . result of high-energy transport through the flow asso-
Figs. 9 and 10 show that the intensity of the streamlines ciated with the irregular motion of the ultrafine particles.

Fig. 12. Effective thermal conductivity enhancement contours ( 102 ) for various volume fractions ðPr ¼ 6:2Þ.
K. Khanafer et al. / International Journal of Heat and Mass Transfer 46 (2003) 3639–3653 3651

In addition, for a Grashof number of 104 , the central 12


__ Numerical
elliptic vortex of the streamline rotates clockwise as the Correlation Gr =105
volume fraction increases. This is associated with higher
9
velocities along the centerline of the enclosure.
As the volume fraction increases, the velocities at the

Average Nu
center of the cavity increase as a result of higher solid– 6
fluid transportation of heat. Moreover, the velocities Gr =104
along the vertical walls of the cavity show a higher level
of activity as predicted by thin hydrodynamic boundary 3
layers. This is illustrated in the vertical velocity com- Gr =103
ponent variation along the horizontal centerline of the
0
cavity for various volume fractions. The isotherms in
0 0.04 0.08 0.12 0.16 0.2
Figs. 9 and 10 show that the vertical stratification of the
Volume Fraction
isotherms breaks down with an increase in the volume
fraction for higher Grashof numbers. This is due to a Fig. 13. Comparison of the average Nusselt number between
number of effects such as gravity, Brownian motion, the numerical results and that obtained by the correlation
ballistic phonon transport, layering at the solid/liquid ðPr ¼ 6:2Þ.
interface, clustering of nanoparticles, and dispersion
effect. In this study we considered only the effect of
dispersion that may coexist in the main flow of a along the hot wall from the correlation and the numer-
nanofluid. ical results for various Grashof numbers and volume
A comparison of the streamlines and isotherms fractions is shown in Fig. 13. This figure shows a linear
contours between nanofluid and the conventional fluid is variation of the average Nusselt number with the void
conducted for various Grashof numbers and a volume fraction. It should be noted that the trend in Fig. 13 for
fraction of / ¼ 10% as shown in Fig. 11. This figure the average Nusselt number versus the volume fraction
clearly shows the impact of the presence of nanoparticles would be downward if the Nusselt number is based on
on the isotherms for a low Grashof number. For a clear the effective thermal conductivity, keff , instead of the
fluid, the isotherms at the center of the cavity are hori- fluid thermal conductivity, kf . The presence of nano-
zontal (stratification in the vertical direction) and be- particles in the fluid enhances the Nusselt number by
come vertical only inside the thermal boundary layers at about 25% for Gr ¼ 104 and Gr ¼ 105 at volume frac-
the vertical walls. The streamlines of a clear fluid show tion of / ¼ 0:2. This increase in the average Nusselt
that the central vortex occupies a larger zone than that number plays a significant role in engineering applica-
for nanofluid at a Grashof number of 104 . For a Gras- tions such as in electronic cooling.
hof number of 105 , the central vortex does not breakup
into three vortices as in the case of a clear fluid. This is
associated with the dispersion effect. 6. Sensitivity to model properties
The effective thermal conductivity enhancement
contours of the nanofluid within the enclosure at dif- Different models based on the physical properties of
ferent Grashof numbers and volume fractions are shown nanofluid as displayed in Table 3 are examined with
in Fig. 12. This figure shows a significant enhancement respect to variations of the average Nusselt number as a
in the effective thermal conductivity of nanofluid com- function of the volume fraction. These variations are
pared to the thermal conductivity of a clear fluid based on different scenarios for the density, viscosity,
ððkeff;nf  kf Þ=kf Þ. and thermal expansion of nanofluid as shown in Table 2
and displayed in Fig. 14. All models used the effective
thermal conductivity of nanoparticles in the present
5. Heat transfer correlation simulations. Fig. 14 gives the upper and lower bounds
for the average nanofluid Nusselt number variations for
The average Nusselt number along the hot vertical different values of volume fractions. It can be seen that
wall is correlated in terms of the Grashof number modeling of the density, viscosity and the thermal ex-
ð103 6 Gr 6 105 Þ and the particles volume fraction pansion coefficient of nanofluid play a central role in
ð0 6 / 6 0:25Þ. Using the results from the present sim- heat transfer enhancement. Model III has the highest
ulations, the correlation can be expressed as average Nusselt number among other models due to a
Nu ¼ 0:5163ð0:4436 þ /1:0809 ÞGr0:3123 ð23Þ higher thermal expansion and density, which results in a
higher convection heat transfer. It should be noted that
where the confidence coefficient of the above equation is model IV has a lower average Nusselt number than
determined as R2 ¼ 99:9%. The average Nusselt number model III due to a lower thermal expansion coefficient.
3652 K. Khanafer et al. / International Journal of Heat and Mass Transfer 46 (2003) 3639–3653

qf qs , / 6 5%
16

General model
Physical basis
_ __ ρeff , µeff , β f , k eff ( III )
(III)

Clear fluid
___ ρ f , µ f , β f , k eff ( I ) (IV)

/ 6 0:5%
12
(I)

Average Nu
(II)
8

3
5
5 (IV)
Gr =10 (III)
Non-dimensional buoyancy term

qf;o
/ qs;o
(I)

1 þ ð1/Þ
(II)

1
4
_ . _ ρeff , µeff , βnf , k eff ( IV )
4
Gr =10

þ
ρ , µ
f eff , β f , k eff ( II )

bs
/ qs;o f
0

qf;o
b
0 0.05 0.1 0.15 0.2

1 þ ð1/Þ
Volume Fraction
1

Fig. 14. Average nanofluid Nusselt number variations for dif-


2
4

ferent models.

1

Both models have the same nanofluid density and ef-


ipffiffiffiffiffiffi

ipffiffiffiffiffiffi
Non-dimensional viscous term

þ ð1  /Þ Gr

þ ð1  /Þ Gr

fective viscosity except for the thermal expansion coef-


ficient which is lower for model IV. As such, the
resulting convection heat transfer for model IV is less
than the one for model III. Model I has a higher average
Nusselt number than model II due to a larger effective
1

1
ð1  /Þ2:5 / qf;o

ð1  /Þ2:5 / qf;o
qs;o

qs;o

viscosity for model II resulting in a thicker momentum


pffiffiffiffiffiffi
ð1  /Þ2:5 Gr

boundary layer and an increase in the shear stress be-


h

tween the fluid layers. As such, the average Nusselt


1

number for model II is lower than model I. Models II


pffiffiffiffiffiffi
Gr
1

and III have the same effective viscosity and thermal


expansion coefficient except that model III has a higher
Thermal expansion coefficient

nanofluid density than model II. Higher effective density


Different models of nanofluid density, viscosity, and thermal expansion coefficient

indicates higher momentum and consequently more heat


transfer enhancement. Model IV is the one that is used
as the default model earlier in the paper. It should be
noted again that the trend in Fig. 14 for the average
Nusselt number versus the volume fraction would be
beff ¼ kbf

downward if the Nusselt number is based on the effective


thermal conductivity, keff , instead of the fluid thermal
bf

bf

bf

conductivity, kf .
ð1  /Þ2:5

ð1  /Þ2:5

ð1  /Þ2:5
lf

lf

lf

7. Conclusions
Viscosity

leff ¼

leff ¼

leff ¼

Heat transfer enhancement in a two-dimensional


lf

enclosure is studied numerically for a range of Grashof


numbers and volume fractions. The present results il-
lustrate that the suspended nanoparticles substantially
qeff ¼ /qs þ ð1  /Þqf

qeff ¼ /qs þ ð1  /Þqf

increase the heat transfer rate at any given Grashof


number. In addition, the results illustrate that the
nanofluid heat transfer rate increases with an increase
in the nanoparticles volume fraction. The presence of
Density

nanoparticles in the fluid is found to alter the structure


of the fluid flow. A comparative study of different
qf

qf

models based on the physical properties of nanofluid is


Model

analyzed in detail. The variances among these models


Table 3

III

IV

are analyzed in the present study. The variants among


II
I

models for the nanofluid density are found to be sub-


K. Khanafer et al. / International Journal of Heat and Mass Transfer 46 (2003) 3639–3653 3653

stantial. Model III has the highest average Nusselt [7] R.H. Davis, The effective thermal conductivity of a
number. The variants among models for the effective composite material with spherical inclusion, Int. J. Ther-
viscosity are found to be more pronounced. Model I is mophys. 7 (1986) 609–620.
found to have a higher average Nusselt number than [8] D.J. Jeffrey, Conduction through a random suspensions of
spheres, Proc. R. Soc. Lond. A 335c (1973) 355–367.
model II. Finally, the variants among models for ther-
[9] J.C. Maxwell, A Treatise on Electricity and Magnetism,
mal expansion coefficient are found to be significant. A second ed., Oxford University Press, Cambridge, 1904,
heat transfer correlation for the nanofluid is obtained pp. 435–441.
and verified for various Grashof numbers and volume [10] R.L. Hamilton, O.K. Crosser, Thermal conductivity of
fractions. This work paves the way for a well-described heterogeneous two-component systems, I & EC Fundam.
systematic experimental investigation to better model 1 (1962) 182–191.
nanofluids. [11] F.J. Wasp, Solid–liquid slurry pipeline transportation,
Trans. Tech. Berlin, 1977.
[12] G.I. Taylor, Dispersion of soluble matter in solvent flowing
Acknowledgements through a tube, Proc. R. Soc. Col. A 219 (1953) 186–203.
[13] Y. Xuan, Q. Li, Heat transfer enhancement of nanofluids,
We acknowledge support of this work by DOD/ Int. J. Heat Fluid Flow 21 (2000) 58–64.
[14] Y. Xuan, W. Roetzel, Conceptions for heat transfer
DARPA/DMEA under grant number DMEA 90-02-2-
correlation of nanofluids, Int. J. Heat Mass Transfer 43
0216. The grant from National Sciences and Engineering (2000) 3701–3707.
Research Council of Canada (NSERC-2002) is ac- [15] H.C. Brinkman, The viscosity of concentrated suspensions
knowledged and appreciated. and solutions, J. Chem. Phys. 20 (1952) 571–581.
[16] A. Amiri, K. Vafai, Analysis of dispersion effects and non-
thermal equilibrium, non-Darcian, variable porosity, in-
References compressible flow through porous media, Int. J. Heat Mass
Transfer 37 (1994) 939–954.
[1] J.A. Eastman, S.U.S. Choi, S. Li, W. Yu, L.J. Thompson, [17] S.V. Patankar, Numerical Heat Transfer and Fluid Flow,
Anomalously increased effective thermal conductivities of Hemisphere, Washington, DC, 1980.
ethylene glycol-based nanofluids containing copper nano- [18] FIDAP Theoretical Manual, Fluid Dynamics Interna-
particles, Appl. Phys. Lett. 78 (2001) 718–720. tional, Evanston, IL, 1990.
[2] S.U.S. Choi, Enhancing thermal conductivity of fluids with [19] G. Barakos, E. Mitsoulis, Natural convection flow in a
nanoparticles, in: D.A. Siginer, H.P. Wang (Eds.), Devel- square cavity revisited: laminar and turbulent models with
opments and Applications of Non-Newtonian Flows, wall functions, Int. J. Numer. Methods Fluids 18 (1994)
FED-vol. 231/MD-vol. 66, ASME, New York, 1995, pp. 695–719.
99–105. [20] N.C. Markatos, K.A. Pericleous, Laminar and turbulent
[3] P. Keblinski, S.R. Phillpot, S.U.S. Choi, J.A. Eastman, natural convection in an enclosed cavity, Int. J. Heat Mass
Mechanisms of heat flow in suspensions of nano-sized Transfer 27 (1984) 772–775.
particles (nanofluids), Int. J. Heat Mass Transfer 45 (2002) [21] G. De Vahl Davis, Natural convection of air in a square
855–863. cavity, a benchmark numerical solution, Int. J. Numer.
[4] X.W. Wang, X.F. Xu, S.U.S. Choi, Thermal conductivity Methods Fluids 3 (1962) 249–264.
of nanoparticle–fluid mixture, J. Thermophys. Heat Trans- [22] T. Fusegi, J.M. Hyun, K. Kuwahara, B. Farouk, A
fer 13 (1999) 474–480. numerical study of three-dimensional natural convection
[5] H.Q. Xie, J.C. Wang, T.G. Xi, Y. Li, F. Ai, Dependence of in a differentially heated cubical enclosure, Int. J. Heat
the thermal conductivity of nanoparticle–fluid mixture on Mass Transfer 34 (1991) 1543–1557.
the base fluid, J. Mater. Sci. Lett. 21 (2002) 1469–1471. [23] R.J. Krane, J. Jessee, Some detailed field measurements for
[6] Z. Hashin, S. Shtrikman, A variational approach to the a natural convection flow in a vertical square enclosure,
theory of the effective magnetic permeability of multiphase Proceedings of the First ASME-JSME Thermal Engineer-
materials, J. Appl. Phys. 33 (1962) 3125–3131. ing Joint Conference, vol. 1, 1983, pp. 323–329.

You might also like