1 s2.0 S0017931013005772 Main

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

International Journal of Heat and Mass Transfer 66 (2013) 200–209

Contents lists available at SciVerse ScienceDirect

International Journal of Heat and Mass Transfer


journal homepage: www.elsevier.com/locate/ijhmt

Laminar mixed convection flow and heat transfer characteristics in a lid


driven cavity with a circular cylinder
Khalil Khanafer a,b,⇑, S.M. Aithal c
a
Frankel Vascular Mechanics Laboratory, Biomedical Engineering Department, University of Michigan, Ann Arbor, MI 48109, USA
b
Section of Vascular Surgery, Samuel and Jean Frankel Cardiovascular Center, University of Michigan, Ann Arbor, MI 48109, USA
c
Computing, Environment and Life Sciences Directorate, Argonne National Laboratory, 9700 S. Cass Avenue, Argonne, IL 60439, USA

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

Article history: Mixed convection flow and heat transfer characteristics in a lid-driven cavity with a circular body inside
Received 30 March 2013 are studied numerically using a finite element formulation based on the Galerkin method of weighted
Received in revised form 1 July 2013 residuals. Comparisons of streamlines, isotherms and average Nusselt number are presented to show
Accepted 6 July 2013
the impact of the Richardson number, non-dimensional radius of the cylinder, and the location of the cyl-
Available online 1 August 2013
inder on the transport phenomena within the cavity. The results of this investigation show that the pres-
ence of the cylinder results in an increase in the average Nusselt number compared with a case with no
Keywords:
cylinder. This result is observed for both, an adiabatic and isothermal, boundary condition imposed on the
Lid-driven cavity
Mixed convection
cylinder. The average Nusselt number increases with an increase in the Richardson number for all non-
Circular cylinder dimensional radius of the cylinder studied in this work. Moreover, the optimal heat transfer results are
obtained when placing the cylinder near the bottom wall for various Richardson numbers. For dominant
natural convection (Ri P 2.5), the average Nusselt number increases with an increase in the non-dimen-
sional radius for 0.05 < ro/H < 0.2. Further increase in the non-dimensional radius does not change the
Nusselt number at a particular Ri. For dominant mixed convection, the average Nusselt number increases
with an increase in the radius of the cylinder for various Richardson numbers.
Ó 2013 Elsevier Ltd. All rights reserved.

1. Introduction in vertical pipes is of interest to sodium cooled reactors [8] or en-


hanced air-conditioning for improved indoor air quality [9].
Mixed convection in enclosures with a moving boundary is an Introduction of an obstruction in an enclosure impacts the flow
interesting problem largely owing to the enhancement in the flow field and heat transfer. Several studies dealing with enhancing nat-
mixing and recirculation on account of the shear stress generated ural convection in obstructed cavities are documented in the liter-
by the moving surface, and its overall effect on the imposed ther- ature [10–13]. Literature on the body inserted lid-driven cavity is
mal gradient. This phenomenon occurs in a number of engineering sparse [14–24]. Dagtekin and Oztop [14] inserted an isothermally
applications from solar power, microelectronics to nuclear reactors heated rectangular block in a lid-driven cavity at different posi-
[1–7]. In the mixed convection mode of heat transfer, features of tions to simulate the cooling of electronic equipments. They found
both forced and natural convection are important. In addition to that dimension of the body was the most effective parameter on
the numerous engineering applications, investigation of various mixed convection flow. MHD mixed convection in a lid-driven cav-
configurations under which mixed convection heat transfer occurs, ity along with joule heating and a centered heat conducting circu-
is of interest as fundamental problems in heat-transfer and fluid lar block was studied by Rahman et al. [15–18]. The numerical
flow. The impact of parameters such as Rayleigh number, Richard- results indicated that the Hartmann number, Reynolds number
son number and Grashof number (Ri has effect of Re on it) on the and Richardson number had strong influence on the streamlines,
overall Nusselt number can lead to better understanding and new- isotherms, average Nusselt number at the hot wall and average
er engineering applications requiring enhanced heat and mass temperature of the fluid in the enclosure. Islam et al. [22] studied
transfer. For instance, turbulent mixed-convection heat transfer mixed convection in a lid driven cavity with an isothermally
heated square blockage and isothermal cold wall temperature im-
posed on all cavity walls. The top wall of the cavity was assumed to
⇑ Corresponding author at: Frankel Vascular Mechanics Laboratory, Biomedical move rightward with a constant speed. Their results showed that
Engineering Department, University of Michigan, Ann Arbor, MI 48109, USA. Richardson number, blockage ratio and blockage placement eccen-
Tel.: +1 734 647 8033. tricities have an effect on the average Nusselt number measured
E-mail address: khanafer@umich.edu (K. Khanafer).

0017-9310/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.ijheatmasstransfer.2013.07.023
K. Khanafer, S.M. Aithal / International Journal of Heat and Mass Transfer 66 (2013) 200–209 201

Nomenclature

g gravitational acceleration x, y cartesian coordinates


Gr Grashof Number = gb(Th  TC)H3/m2 X, Y dimensionless cartesian coordinates, (x, y)/H
H cavity side length
Nu Nusselt number Greek symbols
P dimensionless fluid pressure ¼ p=ðqu2o Þ a thermal diffusivity of the fluid
Pr Prandtl number = m/a b coefficient of thermal expansion of fluid
Re Reynolds number = uoH/m dx, dy dimensionless coordinates of the porous block
Ri Richardson number = Gr/Re2 m kinematic viscosity
ro radius of the cylinder h dimensionless temperature = (T  TC)/(Th  TC)
ro/H non-dimensional radius of the cylinder q density
T temperature W dimensionless stream function
t time
u velocity in x-direction Subscripts
uo lid speed C cold
U dimensionless horizontal velocity = u/uo
h hot
v velocity in y-direction
V dimensionless vertical velocity = m/uo

along the solid block. Oztop et al. [23] studied mixed convection and placed at nine prescribed dimensionless locations (dx, dy) with
heat transfer characteristics in a lid-driven differentially heated the default case being at the center of the cavity (0.5H, 0.5H). No
cavity having a circular body. In that study, flow was driven by slip boundary condition is imposed on all cavity walls and cylinder
the left hot wall moving upward with a constant speed while the surfaces. The top wall is moving towards the right with a constant
other walls were assumed stationary. The horizontal walls were speed uo. The vertical walls of the cavity are considered adiabatic
assumed adiabatic while the right wall was maintained at a lower while the top wall is maintained at a colder temperature TC than
temperature than the left wall. Three different boundary condi- the bottom wall, which is maintained at Th. Different thermal
tions were imposed on the inner cylinder, namely, adiabatic, iso- boundary conditions for the circular body are imposed, namely,
thermal or conductive. Their results showed that the direction of adiabatic or isothermal (TC). Moreover, the cavity is filled with
the moving lid was found to be the most important parameter air (Pr = 0.7) as the working fluid and the thermophysical proper-
on flow and temperature characteristics within the cavity. Mixed ties of the working fluid are taken to be constant except for the
convection heat transfer in a lid-driven cavity along with a heated density variation, which is assumed to vary linearly with temper-
circular hollow cylinder positioned at the center of the cavity was ature according to the Boussinesq approximation.
analyzed numerically by Billah et al. [24]. The left cold wall was as- Based on the above considerations, the following variables are
sumed to move upward with a constant speed. The horizontal introduced to render the analysis dimensionless
walls of the cavity were adiabatic while the right wall was main- x y u v
tained at a cold temperature. This study showed that the flow field X¼ ; Y¼ ; U¼ ; V¼
H H uo uo
and temperature distribution were strongly depending on the cyl-
ðT  T C Þ p
inder diameter and the solid–fluid thermal conductivity ratio. h¼ ; P¼ 2 ð1Þ
Earlier studies on mixed convection in lid-driven enclosures
ðT h  T C Þ quo
with an obstacle considered either heated obstacles or a differen- uo H gbðT h  T C ÞH3 m
Re ¼ ; Gr ¼ ; Pr ¼
tially heated cavity. Many engineering applications such as chem- m m2 a
ical reactors, mixing tanks and boilers have obstructions (such as
where P is the dimensionless pressure, Re the Reynolds number, Gr
stirrers with blades) wherein the lower surface of the reactor/boi-
the Grashof number and Pr the Prandtl number. According to the
ler is at a higher temperature as compared to the upper surface.
above assumptions and dimensionless variables, the normalized
The heat and mass transfer characteristics of such engineering
applications are difficult to evaluate from a numerical standpoint.
However, simplified configurations can provide useful engineering
insights and scaling information. With this goal in mind, the cur-
rent numerical investigation aims at exploring the effects of the
Richardson number, size and location of a circular obstruction on
the momentum and energy transport processes in the lid-driven
cavity heated from below. Two temperature boundary conditions
on the cylinder were investigated, namely, adiabatic and constant
temperature. The results of this investigation are presented in
terms of streamlines, isotherms, and average Nusselt number at
the hot surface of the cavity.

2. Mathematical formulation

The physical system under study with the system of coordinates


is depicted in Fig. 1. The flow inside the cavity with height H is con-
sidered steady, laminar, incompressible, and two-dimensional. A
circular cylinder of radius ro (=0.2H) is inserted within the cavity Fig. 1. Schematic of the model with coordinates and boundary conditions.
202 K. Khanafer, S.M. Aithal / International Journal of Heat and Mass Transfer 66 (2013) 200–209

governing equations for flow and temperature within the cavity are
expressed in the following format
@U @V
þ ¼0 ð2Þ
@X @Y
!
@U @U @P 1 @2U @2U
U þV ¼ þ þ ð3Þ
@X @Y @X Re @X 2 @Y 2
!
@V @V @P 1 @2V @2V Gr
U þV ¼ þ þ þ 2h ð4Þ
@X @Y @Y Re @X 2 @Y 2 Re

@h @h r2 h
U þV ¼ ð5Þ
@X @Y Pr Re
The appropriate forms of the dimensionless boundary conditions
used to solve the governing equations (2)–(5) along the cavity
boundaries are
U ¼ 0;
V ¼ 0; h¼1 Y ¼ 0; 0<X<1
U ¼ 1;
V ¼ 0; h¼0 Y ¼ 1; 0<X<1 ð6Þ
@h
U¼V¼ ¼0 X ¼ 0; 1; 06Y 61
@x Fig. 2. Grid distribution.
The boundary conditions used in the present problem for the cylin-
der are:
Table 1
For isothermal : h ¼ 0 ð7aÞ Grid convergence study for various Richardson numbers (Re = 100, Pr = 0.7, ro/
@h H = 0.2).
For adiabatic : ¼0 ð7bÞ
@n Ri Nu
2820 nodes 19,520 nodes 43,680 nodes
2.1. Evaluation of Nusselt number
0.01 2.92 2.92 2.93
1 3.45 3.50 3.50
The local Nusselt number along the heated bottom wall of the 5 4.71 4.70 4.70
cavity is calculated based on a TC as a reference temperature: 10 5.16 5.04 5.06
 
@h
Nuy ¼  ð8aÞ
@Y y¼0

The above equation defines convective heat flux along the cavity. Table 2
The average Nusselt number is calculated by integrating as Comparison of the average Nusselt number at the top wall between various studies at

Z 1 Z 1 Gr = 104 and 106 in the absence of any obstruction.
@h
Nu ¼  Nux dX ¼  dX ð8bÞ Re Gr = 104 Gr = 106
0 0 @Y
Present Ref. [1] Ref. [2] Present Ref. [1] Ref. [2]
100 1.38 1.34 – 1.02 1.02 –
3. Numerical scheme 400 3.76 3.62 3.82 1.17 1.22 1.17
1000 6.56 6.29 6.50 1.72 1.77 1.81
A finite element formulation based on the Galerkin method was
implemented in the present investigation (ADINA 8.8.3, ADINA R&D,
Inc., Watertown, MA) to solve the governing equations in a lid-dri-
Table 3
ven cavity heated from below with a circular cylinder and subject to
Comparison of the average Nusselt number at the lower hot surface of a lid driven
the boundary conditions described above. The finite element code cavity with a circular cylinder between the present results and that of Nek5000 CFD
used in the present work is well documented [25] and the scheme solver (Re = 100, Pr = 0.7, ro/H = 0.2).
is briefly explained here. The finite element equations are obtained
Ri Isothermal cold cylinder Adiabatic cylinder
by establishing a weak form of the governing equations using the
Presenta Nek5000b %Relative Presenta Nek5000b %Relative
Galerkin procedure. The continuity equation, the momentum equa-
error error
tion and the energy equation are weighted with the virtual quanti-
0.01 2.92 2.91 0.34 2.24 2.21 1.34
ties of pressure, velocities and temperature. The governing
1 3.50 3.47 0.86 3.33 3.32 0.30
equations are integrated over the computational domain. The con- 5 4.69 4.67 0.43 4.06 4.06 0
tinuum domain is first divided into a set of simply shaped, non- 10 5.04 4.99 1.0 4.50 4.40 2.22
overlapping regions called elements, within each of which the un- a
ADINA 8.8.3 CFD.
known variables ux, uy, uz, p, and T are approximated by using the b
Open-source CFD solver based on the spectral element method (Argonne
following equations National laboratory).

ux  uhx ¼ /T ½U x 
uy  uhy ¼ /T ½U y  where /, w, and H are the interpolation functions for velocity, pres-
uz  uhz ¼ /T ½U z  ð9Þ sure, and temperature, respectively, and are local functions of the
nodal coordinates for that element as well as the independent vari-
p  ph ¼ wT ½P ables. The vectors [Ux], [Uy], [Uz], [P], and [T] consist of the values of
T  T h ¼ HT ½T the respective variables at the nodes of the element. Substituting
K. Khanafer, S.M. Aithal / International Journal of Heat and Mass Transfer 66 (2013) 200–209 203

weighted sense by making the residuals orthogonal to the interpo-


lation functions of each element as follows:
R R 9
w  E1 dV ¼ V w  f1 dV ¼ 0 >
=
RV R
V
/  E2 dV ¼ V /  f2 dV ¼ 0 ð11Þ
R R >
;
V
H  E3 dV ¼ V H  f3 dV ¼ 0

This procedure yields a system of equations for each element, which


can be written as

@u
M þ KðuÞ  u ¼ F ð12Þ
@t
Where u is the column vector of the unknown variables, F is the
force vector (incorporating the boundary conditions), M is the mass
matrix, K is the stiffness matrix (representing the diffusion and con-
vection of energy). The representation of the entire continuum re-
Fig. 3. Comparison of the variation of the local Nusselt number along the heated gion of interest is obtained through an assemblage of elements.
bottom wall (isothermal cylinder) between present results and that of Nek5000 for
A non-uniform grid was used to capture the rapid changes in
different Richardson numbers (Re = 100, Pr = 0.7, ro/H = 0.2).
the dependent variables especially in the vicinity of the walls
where major gradients occur inside the boundary layer as depicted
these basis functions into the governing equations and boundary in Fig. 2. The Newton–Raphson method was used to solve the dis-
conditions yields a residual (error) in each of the equations as cretized equations in the fluid region. The solution was assumed to
follows: have converged when the relative change in the dependent vari-
able between two successive iterations was <104 as shown below
Continuity : f 1 ð/; U x ; U y ; U z Þ ¼ E1 ð10aÞ X .X
cþ1
jki;j  kci;j j cþ1
jki;j j 6 104 ð13Þ
Momentum : f 2 ð/; w; H; U x ; U y ; U z ; P; TÞ ¼ E2 ð10bÞ
Energy : f 3 ð/; H; U x ; U y ; U z ; TÞ ¼ E3 ð10cÞ where kci;j represents any particular dependent variable at iteration
c. Different numerical experiments with various mesh sizes were
where E1, E2, and E3 are the residuals resulting from the use of the performed to achieve grid-independent results. Three different
finite element approximations. The Galerkin form of the method of non-uniform grid systems with the following number of nodes:
weighted residuals seeks to reduce these errors to zero in a 2820, 19,520 and 43,680 were examined to calculate the average

Present Streamlines Islam et al. [22]

Isotherms

Fig. 4. Comparison of the streamlines and isotherms between the present results and that of Islam et al. [22] (h/H = 0.25, Re = 100 and Ri = 0.1).
204 K. Khanafer, S.M. Aithal / International Journal of Heat and Mass Transfer 66 (2013) 200–209

Nusselt number along the heated bottom wall. Table 1 shows that a heated from below with a circular cylinder. Moreover, Fig. 3 shows
non-uniform grid system of 19,520 nodes gave grid-independent excellent comparison of the local Nusselt number variation along
results and hence was used for all the results presented in this the bottom heated surface obtained using ADINA and Nek5000.
paper. As an additional check on the accuracy of the present scheme,
Fig. 4 illustrates a comparison of streamlines and isotherms be-
tween the present results and that of Islam et al. [22] in a square
4. Code validation cavity with an isothermally heated square blockage at Reynolds
number Re = 100 and a Richardson number Ri = 0.1.The results
The present numerical scheme was validated against many shown in Tables 2 and 3 along with Figs. 3 and 4 provide sufficient
numerical results available in the literature. Table 2 demonstrates confidence in the present numerical scheme.
an excellent comparison of the average Nusselt number measured
at the top wall of a lid-driven cavity between various studies at
Gr = 104 and 106 in the absence of any obstruction. Table 3 illus- 5. Results and discussion
trates an excellent comparison of the average Nusselt number be-
tween the present results and that of Nek5000 CFD solver (Open- Fig. 5 shows the effect of the presence of the circular cylinder in
source solver; Argonne National Laboratory) in a lid-driven cavity the lid-driven cavity for various Richardson numbers and cylinder

Fig. 5. Effect of the presence of the circular cylinder on the streamlines for various boundary conditions and Richardson numbers (Re = 100, Pr = 0.7, ro/H = 0.2).
K. Khanafer, S.M. Aithal / International Journal of Heat and Mass Transfer 66 (2013) 200–209 205

wall boundary conditions, namely, adiabatic and isothermal. For Dw is about 33%. This implies that introduction of the blockage
the isothermal cylinder wall case, the wall temperature is main- has a significant impact on the flow characteristics. It also signifies
tained at Tc. Re, Pr and ro/H were kept constant at 100, 0.7 and that the thermal boundary condition on the cylinder surface has a
0.2, respectively for these simulations. The results were compared negligible effect.
to a case without a cylinder in the cavity. It is instructive to study Fig. 6 shows the effect of the cylinder on the isotherms for var-
Dw (difference between the maximum and minimum value of the ious Ri. Two cases, namely, isothermal cylinder wall (at Tc) and adi-
stream function) as a function of Ri. For all the cases shown (no cyl- abatic cylinder wall are shown (along with a case with no
inder and the two wall BCs) Dw increases monotonically with Ri. A cylinder). As expected, a higher Ri implies a higher natural convec-
higher Ri represents a higher influence of natural convection and tion, leading to a greater ‘‘homogenization’’ of the temperature
thus a higher value of buoyancy induced flow rate. For every value field in the cavity. The presence of the isothermal cylinder leads
of Ri studied, the Dw for the two different boundary conditions, to a ‘‘thermal shielding’’ effect which results in the fluid on the
namely, isothermal and adiabatic differ by less than 1% while the right side of the cavity being at a lower temperature as compared
Dw for the case with no cylinder is about 33% higher. It must be to the no cylinder and adiabatic cylinder case. This is an important
noted that while the overall blockage of the cylinder is only about effect to consider when using a blockage in engineering applica-
12.5% (p  0.22 = 0.125) of the cavity area, the drop in the overall tions using mixed-convection effects. Figs. 7 and 8 show the effect

Fig. 6. Effect of the presence of the circular cylinder on the isotherms for different boundary conditions and Richardson numbers (Re = 100, Pr = 0.7, ro/H = 0.2).
206 K. Khanafer, S.M. Aithal / International Journal of Heat and Mass Transfer 66 (2013) 200–209

Fig. 7. Effect of varying Richardson number on temperature along a horizontal


plane (Y = 0.15) of the cavity (Re = 100, Pr = 0.7, ro/H = 0.2).

of Richardson number on the variation of the temperature along


different locations namely, Y = 0.15 (below the cylinder), X = 0.15
(left of the cylinder) and X = 0.85 (right of the cylinder), respec-
tively. Fig. 7 show that for both cases (isothermal and adiabatic cyl-
inders), the higher the value of Ri, the lower the value of
temperature at any given horizontal location. It is also seen that
for any given Ri, the temperature of the adiabatic case is higher
than the isothermal case. Additionally, for both cases (isothermal
and adiabatic cylinders), the difference between the maximum
and minimum temperatures in the horizontal direction decreases
with decreasing Ri. This is to be expected as lower Ri indicates a

Fig. 9. Effect of a fixed circular cylinder on (a) local Nusselt number along the
heated wall and (b) the average Nusselt number for various Richardson numbers
(Re = 100, Pr = 0.7, ro/H = 0.2).

Ri = 0.01
(a)
higher Re, implying greater mixing of the fluid leading to less var-
iation in temperature.
Fig. 8(a) and (b) shows the effect of Richardson number on the
θ temperature along vertical planes of the cavity, namely X = 0.15
Ri = 1 Ri = 5
(left of cylinder) and X = 0.85 (right of cylinder), respectively. The
motion of the top surface moving from left to right induces a clock-
X = 0.15 wise flow effect. This flow pattern has an interesting effect on the
temperature gradient on the left and right sides of the cylinder. The
Isothermal Cylinder Ri = 0.01 temperature gradient is affected both by the Re and boundary con-
Adiabatic Cylinder
dition imposed on the cylinder. As the Re increases (lower Ri), the
thermal gradient on the left becomes more uniform for both the
isothermal and adiabatic cases. This can be explained by the en-
hanced fluid mixing on account of the higher Re. It is seen that
Isothermal Cylinder (b) the isothermal BC leads to a near-linear dT/dy for Ri = 0.01 (high
Adiabatic Cylinder
Re). The thermal gradient for the adiabatic BC also more uniform
as compared to Ri = 1 and Ri = 5, but less pronounced as compared
to the isothermal case. The thermal shielding effect due to the iso-
thermal BC is evident in the temperature profiles shown on the
Ri = 0.01 X = 0.85
θ right side (X = 0.85) of the cylinder. It is seen that for Ri = 0.01,
there is practically no temperature gradient for about 60% of the
Ri = 1 distance from the top surface. The adiabatic case has a slight dT/
dy at Ri = 0.01 from Y = 0.5 to Y = 0 as compared to the adiabatic
case.
Ri = 5
The effect of varying Richardson number on the local Nusselt
number measured along the bottom heated surface is demon-
strated in Fig. 9(a). Local Nusselt number is found to increase with
an increase in the Richardson number. For centrally placed isother-
Fig. 8. Effect of varying Richardson on the temperature along a vertical plane of the mal cylinder, Richardson number of 10 (dominant natural convec-
cavity (Re = 100, Pr = 0.7, ro/H = 0.2). tion) exhibits the largest local Nusselt number variation along the
K. Khanafer, S.M. Aithal / International Journal of Heat and Mass Transfer 66 (2013) 200–209 207

Fig. 11(a). Effect of varying the size of the cylinder on the streamlines for various
Richardson numbers using Re = 100 and Pr = 0.7.

case is higher than the adiabatic case and no cylinder case. For
example, the centrally placed isothermal cylinder causes an in-
crease in the velocity in the region between the bottom heated sur-
face and the cylinder and consequently increases heat transfer as
depicted in Fig. 10. The increase in the velocity is due to a small
gap between the bottom wall and the cylinder. Fig. 10 illustrates
the effect of varying the Richardson number on the velocity and
temperature profiles along a horizontal plane placed at Y = 0.15
from the bottom surface. The velocity components were found to
increase with an increase in the Richardson number while the tem-
perature decreases. Fig. 10 shows that the horizontal velocity com-
ponent reaches a maximum value in the region below the center of
the cylinder while the vertical velocity component is nearly zero
Fig. 10. Effect of varying Richardson number on the velocity (U and V) and below the center of the cylinder. This is to be expected since the
temperature along a horizontal plane of the cavity (Re = 100, Pr = 0.7, ro/H = 0.2). motion of the lid causes a clockwise flow. On the right, the flow
vectors are directed towards the bottom plate while towards the
left they are directed towards the top plate. At Y = 0.5 (mid-plane)
bottom surface compared to other configurations. Fig. 9(a) illus- the vertical component of velocity goes to zero indicating a near
trates that no cylinder configuration exhibits the lowest local Nus- horizontal flow in the negative X-direction. The increased velocity
selt number variation along the bottom surface compared to with increasing Ri causes a greater convective flow of the heat from
isothermal and adiabatic centrally placed cylinders for various the bottom surface leading to higher Nusselts number.
Richardson numbers. This can be attributed to a fact that the cen- The effect of changing the size of the cylinder on the stream-
trally placed cylinder causes an increase in the flow velocity be- lines and isotherms for various Richardson numbers (isothermal
tween the cylinder and the bottom surface and consequently cylinder wall) are depicted in Figs. 11(a) and 11(b) respectively.
increases the temperature gradient in the vertical direction. Fig. 11(a) shows that as the cylinder radius increases, the passage
Fig. 9(b) shows the average Nu as a function of Ri at the heated width between the cylinder and the top wall of the cavity deceases
surface of the cavity. For each of the cases considered, namely, no and accordingly the flow behavior changes. For dominant forced
cylinder, adiabatic cylinder and isothermal cylinder – the Nu in- convection mechanism (Ri = 0.01), the main recirculation vortex
creases monotonically with Ri but tends to an asymptotic value starts to elongate horizontally as the non-dimensional radius of
for higher Ri. For a given Ri, the average Nu for the isothermal wall the cylinder increases (ro/H = 0.2) and eventually breaks up into
208 K. Khanafer, S.M. Aithal / International Journal of Heat and Mass Transfer 66 (2013) 200–209

Fig. 12. Effect of varying the non-dimensional radius on the average Nusselt
number for various Richardson numbers (Re = 100 and Pr = 0.7).

Table 4
Effect of varying the location of the cylinder on the average Nusselt number using
Ri = 0.01, Re = 100, Pr = 0.7, and ro/H = 0.2.

Vertical placement (dx, dy) Nu
Top (0.3, 0.7) 1.83
(0.5, 0.7) 2.26
(0.7, 0.7) 1.87
Middle (0.3, 0.5) 2.43
(0.5, 0.5) 2.92
(0.7, 0.5) 2.42
Bottom (0.3, 0.3) 5.03
(0.5, 0.3) 5.19
(0.7, 0.3) 4.99
Fig. 11(b). Effect of varying the size of the cylinder on the isotherms for various
Richardson numbers (Re = 100 and Pr = 0.7).

Table 5
two small vortices between the cylinder and the top wall for ro/ Effect of varying the location of the cylinder on the average Nusselt number using
Ri = 5, Re = 100, Pr = 0.7, and ro/H = 0.2.
H = 0.3. As the Richardson number increases, Fig. 11(a) shows an

absence of vortices within the cavity for Richardson number of 5 Vertical placement (dx, dy) Nu
for various non-dimensional radius of the cylinder. This is to be ex-
Top (0.3, 0.7) 3.17
pected since the natural convection mechanism becomes domi- (0.5, 0.7) 2.89
nant with increasing Ri and there is not much flow induced (0.7, 0.7) 3.03
effect. Fig. 11(b) illustrates the influence of changing the radius Middle (0.3, 0.5) 3.50
of the cylinder on the isotherms for various Richardson numbers. (0.5, 0.5) 4.70
Fig. 11(b) shows that the isotherms are more clustered near the (0.7, 0.5) 3.45
bottom wall as the radius of the cylinder increases, indicating the Bottom (0.3, 0.3) 5.81
development of thin boundary layers with large normal tempera- (0.5, 0.3) 5.52
ture gradients along the bottom wall. (0.7, 0.3) 5.83

Fig. 12 shows the effect of size of cylinder on the average Nus-


selt number at the heated surface (bottom) for various Richardson
numbers (isothermal cylinder wall). The higher temperature gradi- impact of changing the location of the cylinder on the average Nus-
ent for higher values of non-dimensional radius leads to increased selt number is shown in Tables 4 and 5 for different Richardson
average Nusselt number at the bottom wall (as discussed above). numbers. The cylinder is placed at three vertical positions near
The average Nusselt number increases with an increase in the the top, middle, and bottom of the cavity. In this regard, nine dif-
Richardson number for all non-dimensional radius of the cylinder ferent positions of the cylinder were obtained with their dimen-
studied in this work. Moreover, for dominant mixed convection, sionless center point (dx, dy) being given in Tables 4 and 5. Two
the average Nusselt number increases with an increase in the ra- Richardson numbers representing two limiting cases; dominant
dius of the cylinder for various Richardson numbers. Fig. 12 shows forced convection (Ri = 0.01) and dominant natural convection
an interesting result associated with the size of the cylinder. For (Ri = 5) were considered using Re = 100, Pr = 0.7, and ro/H = 0.2.
dominant natural convection (Ri P 2.5), the average Nusselt num- For both cases, it is seen that placing the isothermal cylinder near
ber increases with an increase in the non-dimensional radius for the bottom heated wall results in the highest heat transfer rate in
0.05 < ro/H < 0.2. Further increase in the non-dimensional radius the cavity. This behavior can be attributed to the development of
does not change the Nusselt number for Ri > 2.5. thin thermal boundary layer over the bottom wall. It is also seen
In addition to the size (non-dimensional cylinder radius), the that the least Nusselt number can be obtained when the cylinder
location of the cylinder also plays an important role in modifying is placed near the top wall for both Richardson numbers. Placing
the velocity and thermal fields in the lid-driven cavity. The the cylinder nearer the top wall reduces the constriction of the
K. Khanafer, S.M. Aithal / International Journal of Heat and Mass Transfer 66 (2013) 200–209 209

Table 6 [2] N. Alleborn, H. Raszillier, F. Durst, Lid-driven cavity with heat and mass
CPU time for various Richardson numbers transport, Int. J. Heat Mass Transfer 42 (1999) 833–853.
(Re = 100, Pr = 0.7, ro/H = 0.2). [3] K. Khanafer, K. Vafai, Double-diffusive mixed convection in a lid-driven
enclosure filled with a fluid-saturated porous medium, Numer. Heat Transfer,
Ri CPU (s) Part A 42 (2002) 465–486.
[4] K. Khanafer, A. AlAmiri, I. Pop, Numerical simulation of unsteady mixed
0.01 16.19
convection in a driven cavity using an externally excited sliding lid, Eur. J.
0.1 19.84 Mech. B Fluids 26 (2007) 669–687.
1 24.36 [5] R. Iwatsu, J.M. Hyun, K. Kuwahara, Mixed convection in a driven cavity with a
5a 62.18 stable vertical temperature gradient, Int. J. Heat Mass Transfer 36 (1993)
10a 84.39 1601–1608.
a
[6] A. Al-Amiri, K. Khanafer, J. Bull, I. Pop, Effect of sinusoidal wavy bottom surface
Results from previous Ri were used to on mixed convection heat transfer in a lid-driven cavity, Int. J. Heat Mass
obtained the results for new Ri. Transfer 50 (2007) 1771–1780.
[7] A. Al-Amiri, K. Khanafer, Fluid–structure interaction analysis of mixed
convection heat transfer in a lid-driven cavity with a flexible bottom wall,
Int. J. Heat Mass Transfer 54 (2011) 3826–3836.
flow near the bottom wall leading to reduced heat transfer rates [8] J.D. Jackson, B.P. Axcell, A. Walton, Mixed convection heat transfer to sodium in
(opposite effect as with more constriction near the bottom wall). a vertical pipe, Exp. Heat Transfer 7 (1994) 71–90.
Robustness and computational time are important consider- [9] A. Al-Amiri, K. Khanafer, Augmentation of a heat rejection mechanism in a
ventilated enclosure filled partially with a porous medium, J. Porous Media 15
ations when using simulations for design and parametric studies. (2012) 835–848.
The computation time required for each Richardson number is de- [10] B.S. Kim, D.S. Lee, M.Y. Ha, H.S. Yoon, A numerical study of natural convection
picted in Table 6 (Re = 100, Pr = 0.7, ro/H = 0.2). One can notice that in a square enclosure with a circular cylinder at different vertical locations, Int.
J. Heat Mass Transfer 51 (2008) 1888–1906.
the maximum time taken by the CFD simulations is about 84.39 s
[11] J.R. Lee, M.Y. Ha, A numerical study of natural convection in a horizontal
(Ri = 10) on a single processor (Intel Core, 4 GB RAM, 2.4 GHz) enclosure with a conducting body, Int. J. Heat Mass Transfer 48 (2005) 3308–
using a non-uniform grid. Converged results for Ri = 1, were used 3318.
as initial conditions for Ri = 5 for increasing the stability and [12] J.M. Lee, M.Y. Ha, H.S. Yoon, Natural convection in a square enclosure with a
circular cylinder at different horizontal and diagonal locations, Int. J. Heat
decreasing the computational time. Similarly, the converged re- Mass Transfer 53 (2010) 5905–5919.
sults for Ri = 5 were used as initial conditions for Ri = 10. [13] H.S. Yoon, M.Y. Ha, B.S. Kim, D.H. Yu, Effect of the position of a circular cylinder
in a square enclosure on natural convection at Rayleigh number of 107, Phys.
Fluids 21 (2009) 1–11.
6. Conclusions [14] I. Dagtekin, H.F. Oztop, Mixed convection in an enclosure with a vertical
heated block located, in: Proceedings of ESDA2002: Sixth Biennial Conference
on Engineering Systems Design and Analysis, 2002, pp. 1–8.
Two-dimensional laminar mixed convection flow and heat
[15] M.M. Rahman, M.A. Alim, MHD mixed convection flow in a vertical lid-driven
transfer characteristics in a lid-driven cavity with a circular body square enclosure including a heat conducting horizontal circular cylinder with
inside is investigated numerically for various parameters, namely, Joule heating, Nonlinear Anal. Modell. Control 15 (2) (2010) 199–211.
[16] M.M. Rahman, M.A.H. Mamun, R. Saidur, Shuichi Nagata, Effect of a heat
Rayleigh number, type of boundary condition, non-dimensional ra-
conducting horizontal circular cylinder on MHD mixed convection in a lid-
dius of the cylinder, and the location of the cylinder. The results of driven cavity along with joule heating, Int. J. Mech. Mater. Eng. 4 (3) (2009)
this study reveal that the presence of the cylinder results in an in- 256–265.
crease in the average Nusselt number compared with the case with [17] M.M. Rahman, M.A. Alim, M.M.A. Sarker, Numerical study on the conjugate
effect of joule heating and magneto-hydrodynamics mixed convection in an
no cylinder. This observation indicates that the cylinder can be obstructed lid-driven square cavity, Int. Commun. Heat Mass Transfer 37 (5)
used as a control parameter for enhancing heat transfer character- (2010) 524–534.
istics in such a configuration. Results obtained in this study indi- [18] M.M. Rahman, M.M. Billah, M.A.H. Mamun, R. Saidur, M. Hassan uzzaman,
Reynolds and Prandtl numbers effects on MHD mixed convection in a lid-
cate that the placing the cylinder near the bottom wall increased driven cavity along with joule heating and a centered heat conducting circular
the maximum average Nusselt number while moving the cylinder block, Int. J. Mech. Mater. Eng. 2 (2010) 163–170.
near the top wall resulted in the lowest average Nusselt number on [19] Y. Shih, J. Khodadadi, K. Weng, A. Ahmed, Periodic fluid flow and heat transfer
in a square cavity due to an insulated or isothermal rotating cylinder, J. Heat
the bottom wall. For dominant natural convection (Ri P 2.5), the Transfer 131 (2009) 1–11.
average Nusselt number (on the bottom wall) was found to in- [20] V.A.F. Costa, A.M. Raimundo, Steady mixed convection in a differentially
crease with an increase in the non-dimensional radius for heated square enclosure with an active rotating circular cylinder, Int. J. Heat
Mass Transfer 53 (2010) 1208–1219.
0.05 < ro/H < 0.2. Additional increase in the non-dimensional radius
[21] M.A.H. Mamun, M.M. Rahman, M.M. Billah, R. Saidur, A numerical study on the
does not change the Nusselt number at a particular Ri. For domi- effect of a heated hollow cylinder on mixed convection in a ventilated cavity,
nant mixed convection, the average Nusselt number increased Int. Commun. Heat Mass Transfer 37 (2010) 1326–1334.
[22] A.W. Islam, M.A.R. Sharif, E.S. Carlson, Mixed convection in a lid driven square
with an increase in the radius of the cylinder for various Richard-
cavity with an isothermally heated square blockage inside, Int. J. Heat Mass
son numbers. This study also showed the impact of the boundary Transfer 55 (2012) 5244–5255.
condition on the inner cylinder on the overall thermal field in [23] H.F. Oztop, Z. Zhao, B. Yu, Fluid flow due to combined convection in lid-driven
the cavity. It was seen that imposing an isothermal boundary con- enclosure having circular body, Int. J. Heat Fluid Flow 30 (2009) 886–901.
[24] M.M. Billah, M.M. Rahman, U.M. Sharif, N.A. Rahim, R. Saidur, M.
dition on the inner cylinder leads to a ‘‘thermal shielding’’ effect Hasanuzzaman, Numerical analysis of fluid flow due to mixed convection in
that results in the fluid on the right side of the cavity being at a a lid-driven cavity having a heated circular hollow cylinder, Int. Commun.
lower temperature as compared to the no cylinder and adiabatic Heat Mass Transfer 38 (2011) 1093–1103.
[25] ADINA, Theory and Modeling Guide, vol. III: ADINA CFD & FSI, 2011, pp. 265–
cylinder case. 278.

References

[1] K. Khanafer, A.J. Chamkha, Mixed convection flow in a lid-driven enclosure


filled with a fluid-saturated porous medium, Int. J. Heat Mass Transfer 42
(1998) 2465–2481.

You might also like