Вы находитесь на странице: 1из 18

International Journal of Thermal Sciences 88 (2015) 110e127

Contents lists available at ScienceDirect

International Journal of Thermal Sciences


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

Laminar momentum and heat transfer phenomena of power-law


dilatant uids around an asymmetrically conned cylinder
Sudheer Bijjam a, Amit Dhiman b, *, Vandana Gautam b
a
b

CYIENT Ltd., Hyderabad 500032, India


Department of Chemical Engineering, Indian Institute of Technology Roorkee, Roorkee 247667, India

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 13 September 2013
Received in revised form
15 September 2014
Accepted 23 September 2014
Available online 21 October 2014

The present study focuses on the ow across an asymmetrically conned (heated) cylinder in a channel
for uids obeying Ostwald-de Wale (power-law) equation for the settings: Reynolds number (Re) 1
e40, power-law index (n) 1e1.8, gap ratio (g) 0.375e1, blockage ratio (b) 0.2e0.5 and Prandtl
number (Pr) 1e50. Total drag coefcient and its individual components have been analyzed as a
function of Re, b, g and n. The overall drag coefcient was found to increase with blockage and behavior
of uid, while it drops gradually for increasing Re. The asymmetrical conguration is seen to mitigate the
overall as well as individual drag coefcients. The surface heat transfer coefcient in the form of average
Nusselt number and the Colburn heat transfer jh factor has been thoroughly discussed. Heat transfer rate
is found to increase with increasing Reynolds number and wall connement, while increasing dilatant
behavior impedes the same. As expected, heat transfer results have been reconciled in a single curve by
way of the Colburn jh factor. The jh factor is found higher for the symmetric case as compared to the
asymmetric case.
2014 Elsevier Masson SAS. All rights reserved.

Keywords:
Dilatant uids
Asymmetrical cylinder
Blockage ratio
Drag
Nusselt number
Colburn heat transfer factor
Streamline and isotherm contours

1. Introduction
The phenomena of ow and heat transfer around a circular
cylinder are a conventional problem in uid mechanics, and the
problem also embodies an idealization of many practical ows.
Typical examples include ow around pipes in tubular heat exchangers, instrumentation probes, in hot-wire anemometry, ow
past dividers in polymer processing, piping installations, offshore
cylindrical drilling rigs and others. Extensive investigation in the
form of experimental, analytical and different numerical methods
has been carried out in the past for Newtonian uids with respect
to various aspects of this ow conguration [1e3].
In literature, researchers give sufcient insight into the symmetric wall connement effects around a circular cylinder on momentum and heat transfer. Chen et al. [4] experimentally studied
the formation of steady twin vortices behind the conned cylinder.
It has been found that the rst appearance of the vortices is not
associated with a bifurcation of the full dynamical problem, but
probably bifurcation of a restricted kinematical problem. The

* Corresponding author. Tel.: 91 9410329605 (mobile), 91 1332 285890


(ofce).
E-mail addresses: dhimuamit@rediffmail.com, amitdfch@iitr.ernet.in, amitdfch@
iitr.ac.in (A. Dhiman).
http://dx.doi.org/10.1016/j.ijthermalsci.2014.09.015
1290-0729/ 2014 Elsevier Masson SAS. All rights reserved.

experimental study involved perturbation of steady ow to record


time dependent motions. For a xed blockage (b d/H) in the range
0.1e0.95, the perturbations were found to die away at low Reynolds
number (Re), but above critical Reynolds number disturbance
settled to periodic motion. Anagnostopoulos et al. [5] studied the
same ow conguration at a constant Re of 106 and demonstrated
the effect of b on the wake characteristics for b 0.05, 0.15 and 0.25.
After that, Sahin and Owens [6] notably studied the wall effects on
ow characteristics for a range of blockage ratios 0.1 < b < 0.9, and
0 < Re  280, mainly focusing on stability and transition from
symmetric vortex shedding to asymmetric vortex shedding. Chakraborty et al. [7] also studied the wall effects, but for a wider range
of b (0.05e0.65) and for 0.1  Re  200. This study primarily
established the effect of Re and b on drag coefcient and recirculation zones. Afterward, Ben Richou et al. [8] extended the scope of
study to a different range of b (0.01e0.6) and Re (104e1). They
calculated the drag force exerted on a circular cylinder due to
Poiseuille ow at low Re. They showed how pressure term prevails
over the viscosity term in the lubrication regime. Their ndings can
be utilized in determining the translatory velocity at which a forcefree cylindrical body would move perpendicularly to its axis
midway between two planar walls in Poiseuille ow. Later, Bharti
et al. [9] studied the wall effects on drag coefcient for the varying
ranges of 1  Re  40, 0.2  n  1.9, 0.25  b  0.91 and concluded

S. Bijjam et al. / International Journal of Thermal Sciences 88 (2015) 110e127

Nomenclature
CD
CDF
CDP
CL
Cp
d
ex, ey
f
FD
FDF

FL
h

total drag coefcient


friction drag coefcient
pressure drag coefcient
lift coefcient
heat capacity, J/kg K
diameter of a cylinder, m
unit vectors
body force, N/m3
total drag force per unit length of the cylinder, N/m
frictional component of total drag force per unit length
of cylinder, N/m
pressure component of total drag force per unit length
of cylinder, N/m
lift force per unit length of the cylinder, N/m
local heat transfer coefcient, W/m2 K

h
H
I2
jh
k
Ld
Lu
m
n
nx
ny

average heat transfer coefcient, W/m2 K


height of the channel, m
second variant of the rate of the strain tensor, s2
Colburn heat transfer factor
thermal conductivity, W/m K
downstream length, m
upstream length, m
power-law consistency index, Pa sn
ow behavior index
x-component of the direction vector
y-component of the direction vector

FDP

that, for a xed value of the blockage ratio, the drag coefcient
increases as the shear-thickening tendency (n > 1) increases, and
vice-versa for shear-thinning behavior (n < 1). Subsequently, Bharti
et al. [10] illustrated the effects of Re (1e40), n (0.2e1.8) and b (0.25
and 0.625) on the heat transfer for varying Prandtl numbers
(1  Pr  100) in the steady regime. They reported that the heat
transfer is enhanced with increasing degree of shear-thinning
behavior. Similarly, decreasing the value of the blockage ratio
further enhances the heat transfer rate as the uid behavior
changes from Newtonian to dilatant uids. Hussam et al. [11]
numerically studied the effect of wall connement (0.1e0.4) on
uid ow and heat transfer from a heated circular cylinder to liquid
metal (Pr 0.022) owing in a rectangular duct under the inuence
of a strong magnetic eld for Re ranging from 50 to 3000. They
demonstrated critical Reynolds number increasing with increasing
blockage ratio for the ow past a conned cylinder damped by a
transverse magnetic eld. Recently, Bijjam and Dhiman [12]
investigated the symmetrical connement for the momentum
transfer in the ranges 50  Re  150, 0.4  n  1.8 and b 0.25. The
shear-thinning behavior was reported to yield a decreasing value of
time-averaged drag coefcient than Newtonian uid and the
behavior was opposite for dilatant uids. The role of wall connement across a symmetrically conned cylinder between two parallel walls has also been examined by Rao et al. [13] for the Re range
40e140 for both shear-thinning and shear-thickening behaviors
(0.4  n  1.8). For shear-thickening uids, the ow remains steady
for Re < 140 at b 0.5, while for lower blockage (b) 0.25, 0.34 this
transition is observed somewhere in the range 50  Re  100. The
similar conguration has been studied analytically for different
blockage ratios (0.2  b  0.8) by Khan et al. [14] to investigate its
effect on heat transfer. The modied von KarmanePohlhausen
method is incorporated to solve integral boundary layer momentum equation. Outside the boundary layer, potential ow exists

ns
NuL
Nu
p
Pr
Re
T
Tw
T
Uavg
Ux
Uy
x
y

111

direction vector normal to the surface of the cylinder


local Nusselt number
average Nusselt number
pressure, Pa
Prandtl number
Reynolds number
temperature, K
temperature at the surface of the cylinder, K
temperature at the channel inlet, K
average velocity of the uid at inlet, m/s
x-component of the velocity, m/s
y-component of the velocity, m/s
stream-wise coordinate, m
transverse coordinate, m

Greek symbols
blockage ratio (d/H)
minimum distance from the surface of the cylinder to
the nearest wall, m
g
gap ratio
h
viscosity, Pa s
q
angular displacement from the front stagnation (q 0),
degrees
r
density of the uid, kg/m3
s
stress tensor
t
shear stress, Pa
tx, ty
x- and y-components of the shear stress, Pa

strain tensor

b
D

and the velocity is obtained by the method of images. They further


extended their work into non-Newtonian uids [15]. They found
that shear-thinning uids offer less skin friction and higher heat
transfer coefcients than shear-thickening uids. Unlike above
studies, for Newtonian uids at b 0.66, Semin et al. [16] conducted experimental and numerical analysis on a tethered cylinder
for explaining the stability studies. They have found a new regime
(characteristic of a blockage effect) called conned induced vibration (CIV) which is visible prior to the regime, vertex induced vibration (VIV). The onset of this new regime has been described
using a Von der Pol model in terms of a supercritical Hopf bifurcation depicting the free oscillations. Subsequently, Fani and Gallaire [17] worked on the similar conditions at different connement
regions. They reported a CIV periodic unstable mode at higher
blockage ratios but a steady diverging instability at a moderate
connement region.
The next phase of investigation after symmetrical connement is devoted to analysis of ow and heat transfer over
asymmetric positioning of a circular cylinder in the conned
channel. Further, the kinematic ow behavior of the power-law
uids across an asymmetrically conned circular cylinder between two rectilinear parallel walls is different from that of its
symmetrical orientation due to the change of interaction between the wall boundary layer and the wake at the rear part of
the obstacle. In this eld of research, Zovatto and Pedrizzetti [18]
demonstrated the effects of placement of a cylinder in a plane
channel on the pattern of vortex shedding and showed that the
interaction between the cylinder wake and the wall boundary
layer results in delay in the onset of vortex shedding when cylinder is placed closer to one wall. The transition from a steady
ow to periodic vortex shedding regime has been analyzed for
Re  1600. They also calculated lift and drag coefcients for a
range of gap parameter (D/d) between 0 and 2; whereas, b was

112

S. Bijjam et al. / International Journal of Thermal Sciences 88 (2015) 110e127

kept xed at 0.2. Later, Mettu et al. [19] studied the ow behavior
of Newtonian uids for combined effect of blockage and asymmetrical positioning of a cylinder within the channel. The above
study is based on the belief that two-dimensional (2-D) kinematics can be considered up to Re 500 in the case of asymmetric orientation of the cylinder. Reason being, enhanced
stability of the wake owing to dissimilar interaction between
wall boundary layer and wake, thereby delaying the vortex
shedding phenomena as the circular cylinder approaches one
particular wall. Nirmalkar and Chhabra [20] extended their study
to shear-thinning uids (0.3  n  1) for 0.1  Re  100,
1  Pr  100, 0.2  b  0.4 and 0.25  g  1. They reported that
increase in asymmetry caused deteriorating effect on heat
transfer due to the ineffective recirculation mechanism and high
apparent viscosity of shear layers above of the cylinder, but this
effect was suppressed and the enhancement in heat transfer was
achieved as shear-thinning behavior increases. Recently, Hussam
and Sheard [21] studied Re, b and gap ratio effects on the heat
transfer phenomenon, when a circular cylinder placed in a
channel with an electrically conductive uid, and the heat
transfer takes place from the channel walls to the uid. They
have considered comparatively high Re (100e3000) and
b ranging from 0.1 to 0.4 and reported an optimum position (gap
ratio) at which, it enhances about 50% of heat transfer as
compared to the symmetric position of the cylinder.
Our motivations to pursue the present analysis are one; the
literature is devoid of knowledge about momentum and heat
transfer behavior for shear-thickening or dilatant uids for a
conned circular cylinder in a plane channel. Though shearthickening uids are comparatively less ubiquitous than shearthinning uids, it is important to analyze the physics of their ow
and heat transfer. The dilatant behavior (i.e., increase of viscosity by
increasing shear rate) present in mono-dispersion and polydispersion concentrated solutions which are usually processed in
food, polymer and ceramic units at low Reynolds numbers [22e27].
Second, in industrial applications due to space or design considerations it is not always possible to have cylindrical structures in
the center stage. Due to this, we have interested to know how the
asymmetric position effects the ow behavior change from Newtonian to shear-thickening uids. Similarly, it is essential to know
how this positioning of the cylinder effects drag and heat transfer
around the circular cylinder compared to symmetrical positioning
of the cylinder. This clearly justies the exploration of this area of
research. In view of above practical signicance, this work is
devoted to study the momentum and heat transfer phenomena
from an asymmetrically conned circular cylinder at low Reynolds
numbers range, 1  Re  40; ow behavior index, 1  n  1.8;
blockage ratio, 0.2  b  0.5; gap ratio, 0.375  g  1 for Prandtl
numbers (Pr) 1 and 50.

2. Problem description
The present ow system is approximated by considering the 2D, incompressible, Poiseuille ow of power-law uids over an
innitely long cylinder (of diameter d) conned asymmetrically by
two parallel plane solid adiabatic walls, as shown schematically in
Fig. 1. The position of the cylinder in the channel is denoted by the
gap ratio (g), which is dened here as D/(H/2  d/2). The value of g
is equal to 1 when the cylinder is placed symmetrically between the
plane walls, and 0 when it touches one of the walls. The cylinder is
located at 10d and 40d from inlet and outlet, respectively, which are
sufcient to obtain domain independent results [9,12,13,19,20,28].
As the length of the cylinder in the neutral direction is assumed to
be innitely long, end effects are insignicant, thereby implies that
there is no ow in the neutral direction and that no ow variables
depend upon z-coordinate. This statement can be justied as,
though three-dimensional wake instabilities can disturb 2-D ows,
they generally occur at much higher Reynolds number (Re  180)
compared to here [29]. Though observation is made for Newtonian
unbounded ow conditions, it validates the present analysis at low
Reynolds numbers (Re 1e40) from literature considering shearthickening uids and wall effects [6,9,10,19,28]. The oncoming
uid at temperature T exchanges energy with the isothermal
cylinder whose surface is maintained at a constant temperature
Tw (>T). The temperature difference between the surface of the
cylinder and the surrounding streaming liquid DT (Tw  T) is
kept low (z2 K) so that appreciable variation of the thermophysical properties of the uid (notably density, viscosity, powerlaw parameters and heat capacity) with temperature could be
neglected. Thus, the thermo-physical properties of the streaming
liquid are assumed to be independent of the temperature;
furthermore, the viscous dissipation effects are also assumed to be
negligible as Brinkman number is assumed to be very small. At
constant transport properties, these two assumptions lead to the
decoupling of the thermal energy equation from the momentum
equation, but at the same time they also conne the applicability of
the present model to the situations where the temperature difference is not too signicant and for moderate viscosity and/or
shearing rates.
The governing equations for this ow system are written as
follows [23,30]:

Continuity equation : V$U 0

(1)

Momentum equation : rU$VU  f  V$s 0

(2)

Thermal energy equation : rCp U$VT  kV2 T 0

(3)

where r, U, f, T, s are density, velocity (Ux and Uy components in


Cartesian coordinates), body force, uid temperature and stress
tensor, respectively. The stress tensor (s), which is the sum of the
isotropic pressure (p) and the deviatoric stress tensor (t), is given as
s pI t.
The rheological equation of state for incompressible uids is
given by:

t 2h U
where the components of the rate of strain tensor (U), are given
by:

U
Fig. 1. Schematic representation of a channel conned ow over a circular cylinder in
the asymmetric conguration.

VU VUT
2

For a power-law uid, the viscosity (h) is given by:

S. Bijjam et al. / International Journal of Thermal Sciences 88 (2015) 110e127

 
I
hm 2
2

n1
2

xx



vUy
vUx
1 vUx vUy
; yy
and xy yx

:
2 vy
vx
vy
vx

The appropriate boundary conditions for this ow system can be


written as follows (Fig. 1):
Inlet boundary at AD: The power-law uid is assumed to enter
the inlet plane with a fully developed velocity prole and at the free
stream temperature (T) [9],

"
#





2n 1
2y n1=n

Uavg 1  1  
Ux
n1
H

0  y  H;

(4)
Outow boundary at BC: The zero diffusion ux condition for all
variables is implemented at the outlet boundary. This implies that
the conditions of the out ow plane are extrapolated from within the
domain and have no impact on the upstream ow. The extrapolation
procedure used by Ansys updates the outow velocity and pressure
in a manner that is consistent with the fully developed ow
assumption, when there is no area change at the outow boundary.
The homogeneous Neumann condition is given as:

vUy
vUx
vT
0
0;
0 and
vx
vx
vx
Walls AB and CD: The usual no-slip condition is applied at the
adiabatic conning walls, i.e.,

Ux 0; Uy 0 and

vT
0
vy

On the circular cylinder surface: At the surface of the cylinder, noslip boundary condition is applied and the cylinder is at a constant
temperature of Tw,

Ux 0; Uy 0 and T Tw
The numerical computations have been carried out in the full
domain, because of the asymmetric placement of the cylinder in a
channel over the range of Re, Pr, g and n used in this work. The
numerical solution of governing equations (Eqs. (1)e(3)) in
conjunction with the above-noted boundary conditions yields the
primitive variables such as velocity (Ux and Uy), pressure (p) and
temperature (T) elds.
Some of the commonly used terms in this work and their denitions are as follows:
 Reynolds (Re) and Prandtl (Pr) numbers for non-Newtonian
power-law uids are dened here as:
2n
rdn Uavg

and Pr

The individual drag coefcients (CDF and CDP) are calculated by


using the following denitions:

CDP



C p m Uavg n1
d
k

 Total drag coefcient (CD), which is the sum of friction (CDF) and
pressure (CDP) components, is dened as:

FDP
FDF
and CDF
2 d
2 d
1=2rUavg
1=2rUavg

 The lift force exerted on the bluff body arises due to the asymmetrical orientation of the cylinder. Each wall exerts a repulsive
force on the cylinder; symmetry nullies the net lift, while
asymmetry induces a net positive lift in the y-direction (due to
greater effect of lower wall in the present case, for instance). The
lift coefcient (CL) is a direct measure of this force even if the
ow is steady and is calculated by using the following denition.

CL

Uy 0 and T T

Re

FD
CDF CDP
2 d
1=2rUavg

CD

where I2 is the second invariant of the rate of strain tensor, which is


given as I2 22xx 2yy 2xy 2yx . The components of the rate of
strain tensor are related to the velocity components in Cartesian
coordinates by the following relationships:

113

FL
2 d
1=2rUavg

 The surface averaged Nusselt number Nu hd=k is calculated as:

Nu

1
2p

Z2p
NuL dq
0

where the local Nusselt number (NuL) on the surface of the cylinder
and the unit vector normal to the surface of the cylinder (ns) are
evaluated as:

NuL

xex yey
hd
vT

and ns p nx ex ny ey
k
vns
x2 y 2

where ex and ey are the x- and y-components of the unit vector,


respectively.
The average Nusselt number can be used in process engineering
design calculations to estimate the heat transfer rate from an
isothermal cylinder. A dimensional analysis of the eld equations
and the boundary conditions suggest the average Nusselt number
to be a function of Reynolds and Prandtl numbers, power-law index, gap ratio and blockage ratio.
3. Numerical methodology
This numerical study has been carried out using a nite volume
method based commercial package Ansys Fluent. In this work, the
unstructured quadrilateral cells of non-uniform grid spacing were
generated using the commercial grid module, Gambit. A twodimensional, laminar, segregated solver was used to solve the
incompressible ow on the collocated grid arrangement. Constantdensity and non-Newtonian viscosity models were used. The
quadratic upwind differencing scheme (QUICK) has been used to
discretize the convective terms in momentum and thermal energy
equations. The semi-implicit method for the pressure linked
equations (SIMPLE) scheme was used for solving the pressureevelocity decoupling. The fully developed velocity prole for an
incompressible power-law uid has been incorporated at the inlet
of the channel by user dened functions available in Ansys. Ansys
solves the system of algebraic equations using the Gauss-Siedel
point-by-point iterative method in conjunction with the algebraic
multi-grid (AMG) method. The AMG scheme can greatly reduce the
number of iterations and thus, CPU time required to obtain a

114

S. Bijjam et al. / International Journal of Thermal Sciences 88 (2015) 110e127

converged solution, particularly when the model contains a large


number of control volumes. The stabilization method BCGSTAB was
also used to further increase the efciency of the solver. The absolute residuals of all equations are converged below 1010. A
representative time history of residuals is shown in Fig. 2 for
Re 40, n 1.8, b 0.2, g 1 and Pr 1. The unsteady simulations
were also carried out at the extreme values of Reynolds number,
power-law index for all the values of gap ratio, blockage ratio and
Prandtl number studied in this study to ensure the true solutions
were steady-state in the regime under investigation.
4. Results and discussion
The effects of Re, n, g, b and Pr have been studied on the forced
ow and heat transfer characteristics (ow and thermal patterns,
engineering parameters such as drag and Nusselt number) around a
conned symmetric/asymmetric cylinder in a channel. The entire
ranges of dimensionless control parameters covered are Re 1, 5,
10, 20, 30, 40; n 1, 1.4, 1.8 (thereby covering both Newtonian
(n 1) and dilatant (n > 1) behaviors); Pr 1 and 50 at two values
of gap ratios of 0.375 and 1. To study the wall connement effect,
blockage ratio has been varied from 0.5 to 0.2.
Based on our recent study [12] and others [9,13,19,20,28], the
present computations have been carried out using upstream and
downstream domain lengths of Lu 10d and Ld 40d, respectively.
Also, domain independence studies have been performed to choose
the appropriate values of upstream and downstream distances for
the different wall connements. These dimensions are so chosen
that they frame a domain, which computes free from domain effects and at the same time does not consume excessive computation time. The upstream length was varied from 8d to 10d and then
to 15d for the greatest value of the wall connement of 0.5 utilized.
The effect of Lu was studied at Re 40, Pr 50, g 0.375 and
n 1.8, and the results have been tabulated in Table 2. It has been
determined that Lu 10d provides the best balance between
computational time and domain sensitivity. The maximum relative
differences in CD and Nu are found to be only about 0.7%. Further,
the downstream length has been set to 40d to eliminate domain
effects at the low values of Re investigated here [9,12,13,19,20,28].
Owing to the no-slip condition employed both on the cylinder
and the channel walls; there exist sharp velocity gradients in the
vicinity of these walls. Hence, to capture the ow and temperature

elds effectively, the grid is very ne in the wall proximity. Fig. 3


shows the computational grid for the asymmetric placement of
the cylinder for b 0.5. The grid independence test has been done
using three non-uniform unstructured grids (symbolically represented as G1, G2, and G3) for the extreme values of Reynolds number
(Re) 40 and power-law index (n) 1.8 for the computational
domain employed for the asymmetrical orientation. Table 1 shows
the grid sensitivity analysis for the greatest blockage (b 0.5) at
Re 40, Pr 50, g 0.375, n 1.8 by incorporating the relative
differences in drag coefcient and Nusselt number. Based on the
results obtained, the grid G2 is found to be the most efcient
considering the two opposing factors of increased accuracy coupled
with unnecessarily large computing time incurred. The grid sizes of
38,040 cells and 43,700 cells for b 0.5, 48,400 cells and 63,200 cells
for b 0.34, 63,920 cells and 65,605 cells for b 0.2 are found
adequate for symmetric and asymmetric cases, respectively. The
maximum relative differences in CD and Nu are found only
approximately 0.8%.
However, before proceeding with new results, necessary to rst
validate the present solution procedure with benchmarked results
available in the literature, to ascertain the precision and consistency of the results presented herein for the asymmetrically
conned cylinder ow in a channel.
4.1. Benchmarking of results
Table 3 depicts the comparison of drag values at blockage
(b) 0.2 between the present results and the values of Zovatto and
Pedrizzetti [18] as benchmarked for both symmetrical (g 1) and
asymmetrical (g 0.375) settings of a cylinder. The latter has used
the nite element method to solve NaviereStokes equations with
the number of triangular elements of 10,509 for a symmetric geometry and of 17,728 elements for an asymmetric geometry. On the
other hand, to capture Newtonian ow around an asymmetric
cylinder, Mettu et al. [19] have used a coarser grid size of 90  60
cells. Whereas the optimized grid utilized in the present work is
much ner than the above two studies consisting of 63,920
quadratic cells for symmetric geometry and of 65,605 cells for
asymmetric geometry. The tabulated results in Table 3 show that
the maximum difference between present and literature values is
about 1% for the symmetric case and less than 1.5% for the asymmetric case.

Fig. 2. Convergence history at Re 40, n 1.8, b 0.2, g 1 and Pr 1.

S. Bijjam et al. / International Journal of Thermal Sciences 88 (2015) 110e127

115

Table 1
Grid independence study at Re 40, Pr 50, n 1.8, g 0.375 and b 0.5.
S. no.

Grid

Number
of cells

Number of points
on cylinder

Nusselt
number, Nu

Drag
coefcient, CD

1
2
3

G1
G2
G3

38,357
43,700
45,097

360
400
440

16.9458
16.8256
16.9445

25.4008
25.3648
25.4035

On the other hand, Table 4 compares the values of total drag


coefcients for Newtonian uids at b 0.5 with the results of Sahin
and Owens [6]. On comparison, it validates the behavior observed
by Sahin and Owens [6] and relation (CD a 1/Re) on which the drag
coefcient varies with Reynolds number due to wall effects. Table 5
also compares the total drag values for the non-Newtonian uids
for the symmetric condition at a blockage ratio of 0.25. Here, the
present results are compared with previously available values of
Nejat et al. [28] and Bharti et al. [9] at different power-law indices.
Nejat et al. [28] used a lattice Boltzmann method to solve the
momentum equations, while Bharti et al. [9] have used the nite
volume based solver, Fluent. The comparison shows a close correspondence between the present results and those in literature
[9,28]. Furthermore, extensive benchmarking of the present numerical methodology for non-Newtonian uids in both conned
and unconned domains can be found in our recent study [12]. This
validates the numerical solution procedure used in this study and
thereby enhances the reliability of the present results.
4.2. Flow patterns
The ow behavior can be analyzed by streamline contours in the
vicinity of the cylinder. The variations emanate from symmetric
and asymmetric orientations of the cylinder within the channel,
coupled with variation in Reynolds number, power-law index and
blockage ratio. However, Figs. 4e6 present the streamline contours
for different values of power-law index (1, 1.4, 1.8) for Re 1, 20, 40
for the asymmetric conguration. In the case of shear-thickening
uid ow (n > 1) in a channel, it can be observed that for the
symmetrically oriented cylinder [9,12,13], wake formation starts at
Re 20, where wakes are found to be symmetrical to the axis of the
cylinder. Similar behavior is observed for Newtonian uids (n 1)
[9,11e13,19,20], except that wake formation starts at a lower value
of Re, i.e., at Re 10. As expected, the wake region is found to increase with increasing Reynolds number (1e40) at all values of n,
thereby implying higher uid inertia which restricts it from negotiating the shape of the obstacle and hence separating the ow. It
can be observed that for a constant value of Re the wake region
decreases when increasing the power-law index. Also, the wake is
reported to be longer in shear-thinning uids (n < 1) than that
encountered in Newtonian uids [20]. There is a small wake region
for n 1.8, even at Re 40 (Fig. 6c). There is also no wake formation
for Re ranging from 1 to 5 for both Newtonian and shear-thickening
uids. This can be explained as, when the uid passes through the
annular region its local acceleration (Re) increases thereby
increasing its shear rate. The increase in shear rate makes the
apparent viscosity of shear-thickening uids to increase as
increasing its index (n) value. In addition, viscous dissipation at the
Table 2
Upstream independence study at Re 40, Pr 50, n 1.8, g 0.375 and b 0.5.
S. no.

Lu/d

Grid size

Nusselt number, Nu

Drag coefcient, CD

1
2
3

8
10
15

40,503
43,700
45,127

16.9455
16.8256
16.9439

25.4031
25.3648
24.3315

Fig. 3. The representation of the two-dimensional non-uniform computational grid for


the asymmetric orientation.

wall boundaries, thickness of the boundary layer and interaction


between the cylinder wake to the wall vorticity causes further delay
in wake formation as increasing shear-thickening behavior. Thus, it
can be inferred that wake formation is delayed from Newtonian to
dilatant uids. The conning walls are reported to exert a stabilizing effect on the ow pattern.
When considering the wake regions, there are two different
kinds of wakes that can be observed in the diagrams (Figs. 4e6),
for instance one recirculation zone attached to the lower rear
surface of the cylinder at the rear bottom of the cylinder and
another detached recirculation zone on the near wall

Table 3
Comparison of present drag results with literature values of Ref. [18] in the steady
Newtonian ow regime at b 0.2 for different values of Re and g.
Re

20
30
40

g1

g 0.375

Zovatto and
Pedrizzetti [18]

Present
work

Zovatto and
Pedrizzetti [18]

Present
work

4.75
3.9
3.4

4.7321
3.8921
3.4359

3.7
3.0
2.6

3.6956
3.0184
2.6379

Table 4
Comparison of present drag results with literature values of Ref. [6] in the steady
Newtonian ow regime at b 0.5 for different values of Re and at g 1.

Sahin and Owens [6]


Present work

Re 1

Re 5

Re 10

Re 20

265.23
264.86

53.65
53.46

27.62
27.44

15.07
14.95

Table 5
Comparison of present drag results with literature values [9,28] in the steady ow
regime for Re 40, b 0.25 and g 1 at different values of n.
n

Nejat et al. [28]

Bharti et al. [9]

Present work

0.4
0.6
1
1.4
1.6

2.15
2.65
3.85
5.2
5.9

2.1268
2.6404
3.8327
5.1766
5.9060

2.1139
2.6420
3.8338
5.1802
5.9033

116

S. Bijjam et al. / International Journal of Thermal Sciences 88 (2015) 110e127

Fig. 4. Streamline contours for n 1 and g 0.375 (a) Re 1, (b) Re 20, (c) Re 40 for b 0.2, (d) Re 1, (e) Re 20, (f) Re 40, for b 0.34 and (g) Re 1, (h) Re 20, (i)
Re 40, for b 0.5.

downstream of the cylinder. This phenomenon is similar to ow


around bluff rings studied by Sheard et al. [31,32]. They observed
these similar wakes at low aspect ratio of 1.4 (where aspect ratio
is the ratio of mean ring diameter to cross-section diameter) at
Re 25 for Newtonian uids. When comparing our results at
n 1 for asymmetric conguration to that of torus ows, some
similarities can be found. At Re 1, it is clearly observed that
both stagnation points of pressure on the cylinder are pointed
towards the bottom cross-section. This effect is augmented for
higher blockage ratios due to high viscous forces induced by

walls; though no ow separation is observed at this stage. At


Re 20, due to high inertial forces the ow eld at the bottom of
the cylinder radially diverges by shifting rear stagnation point
upwards. Due to this, the generated adverse pressure gradient
creates a low pressure zone below the axis near the wall. The
physics of the recirculation zone created at this low pressure
region is analogous to that of bubble formation in the ow
behind rings. In both cases, this bubble size increases with an
increase in Re. Another similarity between the ow around a
torus and asymmetric cylinder can be observed by considering

S. Bijjam et al. / International Journal of Thermal Sciences 88 (2015) 110e127

117

Fig. 5. Streamline contours for n 1.4 and g 0.375 (a) Re 1, (b) Re 20, (c) Re 40 for b 0.2, (d) Re 1, (e) Re 20, (f) Re 40 for b 0.34 and (g) Re 1, (h) Re 20, (i)
Re 40 for b 0.5.

the formation of a small recirculation zone near the rear bottom


of the cylinder on increase in Reynolds number further (at
Re 40 for b 0.2), due to the formation of an additional
stagnation point at the rear bottom of the cylinder.
The commonalities of these two different kinds of wake
formation are not only found in the ow around bluff rings.
This kind of analogous ows has been visible in ow through
180 sharp bends [33] and in ow over backward-facing step
[34,35]. In laminar ow around a 180 bend at opening ratios
(it is the ratio of bend opening to the entrance width) greater

than 0.3, the recirculation region near the outlet corner is


similar to the wake formation at the wall in Figs. 4e6. Similarly, ow behind a backward-facing step exhibits a common
feature of diversion of streamlines sticking to the wall
boundary layer on increase of Reynolds number. This phenomenon creates a low pressure inection point causing an
adverse pressure gradient causing the ow to return back to
form a recirculation region. The size of these wakes formed
near to the wall, in all cases, is observed to be increased with
increase in Re in the laminar regime.

118

S. Bijjam et al. / International Journal of Thermal Sciences 88 (2015) 110e127

Fig. 6. Streamline contours for n 1.8 and g 0.375 (a) Re 1, (b) Re 20, (c) Re 40 for b 0.2, (d) Re 1, (e) Re 20, (f) Re 40 for b 0.34 and (g) Re 1, (h) Re 20, (i)
Re 40 for b 0.5.

When considering the shear-thickening effect, as the power-law


index increases, the wake formation at both regions, including at
the rear of the circular cylinder and at the wall boundary layer, is
apparently delayed. This can be attributed to apparent high viscous
forces of the shear-thickening uids dampening the streamlines
divergence from the regular pathway thereby avoiding pressure
formation gradients. One attribute observed here that as the shearthickening nature increases the wake near the wall approaches
towards the circular cylinder and on the other hand irrespective of
the nature of uid the wake size increases with increasing Reynolds

number. This can be explained as the Reynolds number increases


the high velocity streamlines augmenting the pressure gradients at
the wall boundary layer thereby lengthening size of the wake.
4.3. Thermal patterns
To study heat transfer characteristics between the owing uid
and the heated bluff body, temperature proles have to be examined closely. For a symmetric orientation of a cylinder, at constant
values of Prandtl number and power-law index, the isotherm

S. Bijjam et al. / International Journal of Thermal Sciences 88 (2015) 110e127

contours start narrowing down as Reynolds number increases


[10,20]. This has also been observed for all the cases studied here.
Similarly, as the Prandtl number is increased from Pr 1 to 50, for
Newtonian and dilatant uids and for all Re studied, the thermal
contours narrow down along the downstream of the channel as a
result of the thinning of thermal boundary layer, and hence higher
heat transfer rate is achieved (Figs. 7e10). There is slight turning of
isotherm contours for Re  30 for n 1.4 and there is no turning for
n 1.8 (Figs. 9 and 10). This effect is seen to aggravate with increase
in Prandtl number. The asymmetrical orientation of the cylinder
brings in certain complex characteristic features of isotherms. It can
be observed that for Pr 1, the direction of the isotherm pattern
changes from facing the lower wall surface to the downstream
direction as we move to higher Re (Re  1), for both Newtonian and
shear-thickening uids (Figs. 7 and 9). This arises due to the higher
mass ow rate at increasing Re. As we move to Pr 50, the isotherm
contours are concentrated more in the downstream region of the
cylinder and are seemingly more restricted for higher Re values
(Re > 10) (Figs. 8 and 10). This can be attributed to the decay of
temperature eld as a result of thinning of thermal boundary layer
with increase in Pr. It would be important to note that due to the

119

low rate of thermal diffusivity relative to molecular diffusivity at


Pr 50, the hot uid tightly follows the departing streamline
behind the cylinder. Isotherms are also inclined towards the lower
channel wall at these conditions (for n  1).
4.4. Friction drag coefcient
It is notable from Fig. 11aec that blockage parameter (b) lays a
great impact on friction drag coefcient (CDF) as the wall connement effect increases (increasing b from 0.2 to 0.5). This is attributed to the increased shear stress on the surface of the cylinder as it
approaches a wall, for both gap ratios and at a particular ow
behavior index. The symmetrical orientation (g 1) of the circular
cylinder experiences more CDF than its counterpart (g 0.375) for
all Reynolds numbers and at similar conditions. This can be
explained as, when the circular cylinder approaches one of the
walls (g 0.375), the gap between cylinder and its adjacent wall is
lled with viscous boundary layer which offers great resistance to
the upcoming liquid, and this in turn causes a greater proportion of
the uid to ow above the cylinder. Therefore, the drag force due to
skin friction caused by the uid ow on the cylinder is less

Fig. 7. Thermal contours for n 1, Pr 1 and g 0.375 (a) Re 1, (b) Re 20, (c) Re 40 for b 0.2, (d) Re 1, (e) Re 20, (f) Re 40 for b 0.34, (g) Re 1, (h) Re 20, (i) Re 40
for b 0.5.

120

S. Bijjam et al. / International Journal of Thermal Sciences 88 (2015) 110e127

Fig. 8. Thermal contours for n 1, Pr 50 and g 0.375 (a) Re 1, (b) Re 20, (c) Re 40 for b 0.2, (d) Re 1, (e) Re 20, (f) Re 40 for b 0.34, (g) Re 1, (h) Re 20, (i)
Re 40 for b 0.5.

compared to the symmetrically placed cylinder. Also, the blockage


effect is more prominent on CDF as shear-thickening nature increases, thereby enhancing wall effects. Hence, CDF varies in direct
proportion with power-law index, symmetry in orientation and
blockage ratio, while it varies inversely with Reynolds number.
4.5. Pressure drag coefcient
It can also be observed from Fig. 11aec that the wall connement has a signicant role on pressure drag coefcient (CDP) similar
to as observed in CDF. As the blockage percentage increases
(increasing b from 0.2 to 0.5) the CDP also increases for both congurations (g 1 and 0.375) at all values of n considered. An increase in power-law index causes CDP to rise, but this inuence
gradually diminishes with increasing Re. With increasing blockage,
CDP value rises with increase in Re. Again, the effect of n on CDP
decreases as the wall connement decreases and more of wall effects come into play. The symmetrical orientation (g 1) of the

circular cylinder exhibits higher values of CDP than the asymmetrical orientation (g 0.375). This phenomenon can be explained as
the cylinder approaches the wall its wake decreases due to wall
repulsive forces and on the other side the wake increases this in
turn is responsible for uneven pressure distribution over the cylinder, this difference can be observed much higher at low Reynolds
numbers (Re  5) and attenuates the same slowly as Reynolds
number increases. On the other hand, the wall effects become more
signicant on the CDP value as the shear-thickening effect increases.
It is evident from the gure that for all blockage ratios the CDP
values increase with ow behavior index (shear-thickening nature).
This observation is further strengthened by the ndings of Nirmalkar and Chhabra [20]. They reported pressure drag coefcient to
increase as shear-thinning behavior transits to Newtonian behavior.
Similarly, at a xed connement ratio, the symmetrical orientation
(g 1) of the cylinder gives higher values of CDP compared to its
counterpart (g 0.375), and this difference is aggravated at low
Reynolds numbers (Re  5), for all n values investigated here.

S. Bijjam et al. / International Journal of Thermal Sciences 88 (2015) 110e127

121

Fig. 9. Thermal contours for n 1.8, Pr 1 and g 0.375 (a) Re 1, (b) Re 20, (c) Re 40 for b 0.2, (d) Re 1, (e) Re 20, (f) Re 40 for b 0.34, (g) Re 1, (h) Re 20, (i)
Re 40 for b 0.5.

4.6. Drag ratio CDP/CDF


To understand the physics of the problem considered, it is not
only required to study the individual drag coefcients, but also to
understand the relative contributions of friction and pressure drag
components. Fig. 11aec depicts the relative contribution of pressure and friction drag coefcients (i.e. CDP/CDF) on the blockage
effects at different values of ow behavior indices, gap ratios and
Reynolds numbers. For a particular ow behavior, the ratio of
pressure drag to friction drag increases with blockage ratio for the
entire range of Re covered in this study. This behavior is found to
derogate with increasing shear-thickening nature of uid ow
owing to the relative increase in uid viscosity, hence total drag
deriving primarily from friction drag. As expected, this is in contrast
with increasing drag ratio for shear-thinning uids [20]. This
further suggests a decreasing impact of n as ow transits from
Newtonian to shear-thinning nature. The orientation of cylinder
(symmetric g 1/asymmetric g 0.375) is seen to have a small

effect on CDP/CDF, at low connement ratio (b 0.2). This is due to


the wall effects having a minimal inuence at low blockage ratios.
Contrary to this, as we increase the connement ratio the nature of
contribution of these two drag coefcients differs greatly for the
orientation changes, i.e. the pressure drag contribution has a major
role compared to friction drag for the asymmetric orientation
(g 0.375) compared with the symmetric orientation (g 1). The
same phenomenon is found to aggravate as Reynolds number increases from 1 to 40. The relative inuence of pressure drag forces
compared to friction drag forces decreases either approaching
symmetry of the cylinder (i.e. g 1) or increasing the shearthickening behavior of uids.
4.7. Overall drag coefcient
The ow dynamics differ for an asymmetrically conned cylinder from a symmetric orientation owing to proximity of the stationary wall to the surface of the cylinder. This leads to a signicant

122

S. Bijjam et al. / International Journal of Thermal Sciences 88 (2015) 110e127

Fig. 10. Thermal contours for n 1.8, Pr 50 and g 0.375 (a) Re 1, (b) Re 20, (c) Re 40 for b 0.2, (d) Re 1, (e) Re 20, (f) Re 40 for b 0.34, (g) Re 1, (h) Re 20, (i)
Re 40 for b 0.5.

change in the value of total drag coefcient (CD). Fig. 12aec illustrates the effect of wall connement and power-law index on the
total drag coefcient for both symmetrical and asymmetrical positions of the circular cylinder in the channel. The overall observation of total drag coefcient on the ow and geometrical
parameters (Re, b, n and g) of Fig. 12 has been presented in equation
(5) in the form of multi parameter power-law regression t to
elucidate the effect of these parameters and for greater understanding of the effect of these parameters on ow behavior. For this
analysis, the Reynolds number range Re  10 has been considered
as the dependency on parameters is higher at lower Reynolds
numbers.

CD

Kna bb
qRec gd

(5)

The values of exponents a, b, c and d on the respective parameters have been given in Table 6 at low Reynolds numbers (10). In
the above power-law expression, K is proportionality constant and

its value in this expression is 822.87. Similarly, q is an error


parameter and is a function of b and Re. The value of q has to be
calculated using equation (6)

q b; Re xn2 yn z

(6)

The coefcient values x, y and z at respective b and Re have been


presented in Table 7. The value of q takes into consideration nonlinearities associated in equation (5).
From equation (5) and Fig. 12, it is clear that Reynolds number
and gap ratio are inversely proportional, but power-law index and
blockage ratio are directly proportional to the drag coefcient.
When considering the impact of Reynolds number on drag coefcient in the unconned cylinder, it is found to be inversely proportional with an exponent of around 0.56 [7,19] for Re  10. On the
other hand, in the conned situation, walls augmented dependency
on Reynolds number with an exponent value equal to nearly one
(CD a 1/Re) is evident from Table 6 and the results of Sahin and
Owens [6]. Considering the wall effects on drag coefcient at xed

S. Bijjam et al. / International Journal of Thermal Sciences 88 (2015) 110e127

123

Fig. 11. Effects of Reynolds number (Re), power-law index (n), blockage ratio (b) and gap ratio (g) on drag ratio (CDP/CDF).

Re, drag coefcient increases with increasing blockage ratio owing


to the sharpening of the velocity gradient and energy dissipation at
the conning walls. The effect is observed for both shear-thinning
[20] and shear-thickening uids. The change in drag coefcient is
more prominent at low Re. This is attributed to thicker boundary
layer at low Re [19]. But the effect of b on it slowly decreases once Re
increases. From Table 6 and Fig. 12, it can be observed that the
connement effect is more prominent in the symmetric position of
the cylinder than asymmetric position. Further, analyzing the effect
of shear-thickening behavior, at xed Reynolds number and at
higher blockage ratios the impact of power-law index on drag is
dominant compared to all other parameters. However, at low
blockage ratios the dominating element for drag contribution is b.
In other words, it can be stated that for both orientations of the
cylinder at xed Re, at higher blockages, shear-thickening phenomenon has the major contribution of drag, conversely, at lower
blockage ratios wall effects play the major dominating factor. This
can be attributed to shear-thickening (high apparent viscosity)
nature, which enhances the wall effects thereby further sharpening
of velocity gradients and dissipation.
Furthermore, CD is found to vary inversely with g. It is found that
the total drag values are higher for g 1 compared to g 0.375 at a
xed blockage ratio. This can be explained as both CDF and CDP have
a higher value for symmetric orientation than that of asymmetric
case and CD is derived from these two components. Overall, every
parameters of equation (5) at a given Reynolds number has higher
impact on drag at symmetrical position than asymmetric.

4.8. Lift coefcient


The ow being asymmetric about the mid plane for g 0.375,
incorporates the effect of asymmetric variation of normal and
tangential forces exerted on the surface of cylinder. Hence, a net lift
force is brought into effect by the increasing degree of asymmetry.
Fig. 13aec represents the dependence of lift coefcient (CL) on
different ow variables such as Reynolds number, blockage ratio and
power-law index. It is notable from the gures that lift coefcient
increases (positive dependency) as the blockage ratio increases from
0.2 to 0.5. This is attributed to the connement of the walls, i.e.
repulsive force exerted by the walls. The repulsive force from the wall
tends to force the cylinder away from the wall and this repulsion
force increases as the wall proximity increases. Due to asymmetric
position of a cylinder in the channel, the unbalanced repulsive forces
on circular cylinder produce higher values of lift coefcients with
increasing blockage ratio. From the gure, it can be observed that as
the shear-thickening nature causes higher value of lift coefcient
when compared to Newtonian uids. The effect is aggravated as
blockage ratio increases and in the same way attenuated as blockage
ratio decreases. This can be ascribed to the fact that shear-thickening
uids intensify the wall effect of repulsion and this effect increases
with shear-thickening behavior. Shear-thinning uids, on the other
hand, experience lesser lift than Newtonian/shear-thickening uids
due to reduction in effective viscosity [20]. Further, lift coefcients
show a negative dependency on the Reynolds number, irrespective of
blockage ratio and ow behavior index. In the time-periodic regime,

124

S. Bijjam et al. / International Journal of Thermal Sciences 88 (2015) 110e127

Fig. 12. Effects of Reynolds number (Re), power-law index (n), blockage ratio (b) and gap ratio (g) on total drag coefcient (CD).

the oscillations in CL are wiped away when cylinder is very close to


one of the conning walls as reported in Ref. [19].
4.9. Average Nusselt number
The combined effects of different controlling parameters such as
blockage ratio (b), gap ratio (g), ow behavior index (n), Reynolds
number (Re) and Prandtl number (Pr) on the average Nusselt
number (Nu) are discussed here. Fig. 14a and b shows the dependency of the average Nusselt number on power-law index as a
function of Reynolds number at Pr 1 for g 1, 0.375 and
b 0.2e0.5. At small value of Reynolds number (Re 1, for
instance), the average Nusselt number is seen not to be signicantly
Table 6
Parameter exponents of power-law model to calculate the total drag coefcient of
equation (5) for symmetric and asymmetric circular cylinder in a channel.
Re

10

0.2
0.34
0.5
0.2
0.34
0.5
0.2
0.34
0.5

Symmetrical cylinder
b

0.8708
1.9698
3.2773
0.7803
1.9116
3.2621
0.7433
1.79
3.219

1.7749

1.097

0.3967
1.4879
2.7978
0.4497
1.3353
2.7239
0.4951
1.1666
2.5426

1.7373

1.056

0.5181

1.4092

Table 7
Coefcients of equation (6) to calculate error parameter q.
Re

Asymmetrical cylinder

1.6103

inuenced by power-law index. This is due to the fact that in this


situation, advection contributes to heat transfer only slightly. As
Reynolds number is progressively increased, convection starts
playing an important role and hence this feature derives a strong
heat transfer dependency on power-law index. Heat transfer is
found to be more dependent on Re than Pr for shear-thinning uids
[20]. It is notable from Fig. 14a and c that the symmetrically positioned circular cylinder achieves a stronger heat transfer coefcient
compared to the asymmetrical one, for all blockage ratios studied.
Similarly, overall observation elucidates that the surface Nusselt
number decreases as the shear-thickening effect increases
(increasing n) for all blockage ratios. Following an opposite trendline, increasing shear-thinning behavior increases Nu for all
blockage ratios [10,20]. When considering wall connement, the

1.4825

1.3291

10

0.2
0.34
0.5
0.2
0.34
0.5
0.2
0.34
0.5

Symmetrical cylinder

Asymmetrical cylinder

0.2114
0.67
0.88
0.1653
0.7135
0.8818
0.1103
0.7524
0.9113

0.588
1.8635
2.4483
0.4597
1.9847
2.4534
0.3069
2.0929
2.5355

0.5452
0.0391
0.6998
0.6216
0.0946
0.7136
0.7191
0.1569
0.7712

0.1066
0.5288
0.7658
0.0727
0.5384
0.814
0.0604
0.4345
0.9366

0.2965
1.4706
2.1304
0.2023
1.4976
2.2647
0.168
1.2085
2.6058

0.7379
0.2149
0.4897
0.7985
0.2011
0.5794
0.8309
0.3613
0.7897

Fig. 13. Effects of Reynolds number (Re), power-law index (n), blockage ratio (b) on lift coefcient (CL) for g 0.375.

Fig. 14. Effects of Reynolds number (Re), power-law index (n), blockage ratio (b) and gap ratio (g) on average Nusselt number (Nu) at different values of Prandtl numbers (Pr).

126

S. Bijjam et al. / International Journal of Thermal Sciences 88 (2015) 110e127

average Nusselt number increases as the blockage ratio increases


from 0.2 to 0.5 for both symmetrical and asymmetrical cylinder
congurations, and is a reverse phenomenon at low Reynolds
number (Re 1). From the same gure, it can be observed that
increasing the wall connement effect over powers the effect of
ow behavior index; for example, at b 0.2 the Nu value maintains
approximately a single point up to Re 5, for b 0.34 it is up to
Re 10, for b 0.5 it is up to Re 20 for both symmetrical and
asymmetrical positions of the circular cylinder.
On the other hand, Fig. 14c and d shows the effect of the above
specied control parameters on the average Nusselt number
comparatively at higher Prandtl number (Pr 50). The values of the
average Nusselt numbers are higher compared to the corresponding values estimated at Pr 1. Nevertheless, as the blockage ratio
increases, the value of average Nusselt number also increases for
both symmetric and asymmetric positions of the circular cylinder
for all values of the Reynolds numbers. In contrast, the Nu values
estimated at Pr 1 and at low Reynolds numbers (Re 1) were
found to vary inversely with blockage ratio. The overpowering
nature of the wall connement effect on the dilatant ow behavior,
as seen above for the Pr 1 case is, to a moderate extent reduced by
increasing the Prandtl number to Pr 50. Due to this, at b 0.2, Nu
value maintains approximately constant limited to low Re 1, for
b 0.34 it is up to Re 5 and for b 0.5 it is up to Re 10 for both
asymmetrical and symmetrical positioning of the circular cylinder
at Pr 1. While at a higher Pr 50, the effect is seen at a low value
of Re 1 for all blockage ratios and both the orientations.

Further insights are provided by the Colburn heat transfer factor


(jh), which is dened as jh Nu/(RePr1/3). The Colburn factor (jh)
characterizes the analogy between heat transfer and friction
developed by uid ow. This nds application in the range of
Prandtl number investigated here. The signicance of this parameter lies in the fact that it gives the opportunity of merging the
results for a range of values of Reynolds and Prandtl numbers into a
single curve. Further, this dimensionless number facilitates calculation of heat transfer coefcient or Nusselt number at intermediate
values of Reynolds and Prandtl numbers especially in the heatexchangers design [36]. Fig. 15aed shows the Colburn heat transfer factor dependency on Reynolds numbers for three blockage
ratios (b 0.2, 0.34 and 0.5) for both symmetrical and asymmetrical positions of the circular cylinder at Pr 1 and 50. It can be
observed from the gure that results for liquids of different ow
behavior index (n 1, 1.4 and 1.8) collapse almost to a single curve
for a xed blockage ratio at both gap ratios. The jh factor increases
with increasing blockage ratio for all Reynolds and Prandtl numbers
and decreases with increasing degree of asymmetry. However, at
Re 1, Pr 1 for both gap ratios, the opposite trend of the jh factor
is observed because the conduction effects are prominent at low
values of Re and Pr. Furthermore, it is quite clear that at Re 1,
Pr 1, jh Nu and hence the jh factor follows the same variation as
Nu. At higher values of Re and Pr, it can be observed by the present
numerical values that the jh factor varies inversely with Re. Similarly, the jh factor is a monotonically decreasing function of Re for
shear-thinning behavior [10].

Fig. 15. Effects of Reynolds number (Re), power-law index (n), blockage ratio (b) and gap ratio (g) on the jh factor at different values of Prandtl numbers (Pr).

S. Bijjam et al. / International Journal of Thermal Sciences 88 (2015) 110e127

5. Conclusions
Two-dimensional numerical analyses have been carried out for
momentum and heat transfer for an asymmetric cylinder in a plane
channel for shear-thickening power-law uids. The Reynolds number is studied from 1 to 40 with varying values of n (1e1.8),
b (0.20.5), g (0.375e1) and Pr (1e50). In the preceding range of
conditions, the ow is found to be laminar and at steady state. The
present analysis enhanced the understanding of the engineering
parameters and the heat transfer characteristics with the help of ow
patterns. These ow patterns enabled to nd the distinguished
feature of two wakes one at the bottom of the cylinder and another
detached recirculation zone on the near wall downstream of the
cylinder, this is not found in the symmetrically positioned circular
cylinder. Similarly, this paper identies the similar kind of ow
behavior and wake regions occurred in other geometries and
compared the analogies. In addition, we have quantitatively presented the relationship between various ow parameters and
geometrical parameters in terms of multi parameter power-law
model using regression analysis. In future perspective, these kinds
of multi parametric models can also be developed for shear-thinning
uids as well as for non-Newtonian uids at higher Reynolds
numbers. This will enhance the deep understanding of the effects of
various control parameters on ow and thermal characteristics. It has
been found that at constant Pr and g, the average Nusselt number is
increasing with increasing Re. The values of average Nusselt numbers
are signicantly increased when Pr is increased from 1 to 50. Also, the
effect on average Nusselt number due to an increase in Prandtl
number is much more prominent than due to placement of cylinder.
The heat transfer jh factor plots for different Re, Pr, g and n reconcile
into a single plot with the jh factor being higher for symmetric case as
compared to asymmetric case for the above range of settings.
Acknowledgments

[10]
[11]

[12]

[13]
[14]

[15]

[16]

[17]

[18]
[19]

[20]

[21]

[22]
[23]
[24]

[25]
[26]

The authors would like to thank the four anonymous reviewers


for their valuable and helpful comments on this work.

[27]
[28]

References
[1] M.M. Zdravkovich, Flow Around Circular Cylinders, in: Fundamentals, vol. 1,
Oxford University Press, New York, 1997.
[2] M.M. Zdravkovich, Flow Around Circular Cylinders, in: Applications, vol. 2,
Oxford University Press, New York, 2003.
[3] C. Norberg, Fluctuating lift on a circular cylinder: review and new measurements, J. Fluid Struct. 17 (2003) 57e96.
[4] J.H. Chen, W.G. Pritchard, S.J. Tavener, Bifurcation of ow past a cylinder between parallel plates, J. Fluid Mech. 284 (1995) 23e41.
[5] P. Anagnostopoulos, G. Illiadis, S. Richardson, Numerical study of the blockage
effect on viscous ow past a circular cylinder, Int. J. Numer. Method Fluids 22
(1996) 1061e1074.
[6] M. Sahin, R.G. Owens, A numerical investigation of wall effects up to high
blockage ratios on two-dimensional ow past a conned circular cylinder,
Phys. Fluids 16 (2004) 1305e1320.
[7] J. Chakraborty, N. Verma, R.P. Chhabra, Wall effects in the ow past a circular
cylinder in a plane channel: a numerical study, Chem. Eng. Process. 43 (2004)
1529e1537.
[8] A. Ben Richou, A. Ambari, J.K. Naciri, Drag force on a circular cylinder midway
between two parallel plates at very low Reynolds numbers, part 1: Poiseuille
ow (numerical), Chem. Eng. Sci. 60 (2005) 2535e2543.
[9] R.P. Bharti, R.P. Chhabra, V. Eswaran, Two-dimensional steady Poiseuille
ow of power-law uids across a circular cylinder in a plane conned

[29]
[30]
[31]

[32]

[33]

[34]

[35]
[36]

127

channel: wall effects and drag coefcients, Ind. Eng. Chem. Res. 46 (2007)
3820e3840.
R.P. Bharti, R.P. Chhabra, V. Eswaran, Effect of blockage on heat transfer from a
cylinder to power law liquids, Chem. Eng. Sci. 62 (2007) 4729e4741.
W.K. Hussam, M.C. Thompson, G.J. Sheard, Dynamics and heat transfer in a
quasi-two-dimensional MHD ow past a circular cylinder in a duct at high
Hartmann number, Int. J. Heat Mass Transfer 54 (2011) 1091e1100.
S. Bijjam, A.K. Dhiman, CFD analysis of two-dimensional non-Newtonian
power-law ow across a circular cylinder conned in a channel, Chem. Eng.
Commun. 199 (2012) 767e785.
M.K. Rao, A.K. Sahu, R.P. Chhabra, Effect of connement on power-law uid
ow past a circular cylinder, Poly. Eng. Sci. 51 (2011) 2044e2065.
W.A. Khan, J.R. Culham, M.M. Yovanovich, Fluid ow and heat transfer from a
cylinder between parallel planes, J. Thermophys. Heat Transfer 18 (2004)
395e403.
W.A. Khan, J.R. Culham, M.M. Yovanovich, Fluid ow and heat transfer in
power-law uids across circular cylinders: analytical study, J. Heat Transfer
128 (2006) 870e878.
B. Semin, A. Decoene, J.-P. Hulin, M.L.M. Franois, H. Auradou, New oscillatory
instability of a conned cylinder in a ow below the vortex shedding
threshold, J. Fluid Mech. 690 (2012) 345e365.
A. Fani, F. Gallaire, A 2D pendulum submitted to an incoming ow: drag acting
like gravity and new instabilities, in: Bulletin of the American Physical Society,
66th Annual Meeting of the APS Division of Fluid Dynamics, vol. 58, Pennsylvania, Pittsburgh, November 24e26, 2013.
L. Zovatto, G. Pedrizzetti, Flow about a circular cylinder between parallel
walls, J. Fluid Mech. 440 (2001) 1e25.
S. Mettu, N. Verma, R.P. Chhabra, Momentum and heat transfer from an
asymmetrically conned circular cylinder in a plane channel, Heat Mass
Transfer 42 (2006) 1037e1048.
N. Nirmalkar, R.P. Chhabra, Forced convection in power-law uids from an
asymmetrically conned heated circular cylinder, Int. J. Heat Mass Transfer 55
(2011) 235e250.
W.K. Hussam, G.J. Sheard, Heat transfer in a high Hartmann number MHD
duct ow with a circular cylinder placed near the heated side-wall, Int. J. Heat
Mass Transfer 67 (2013) 944e954.
R.K. Gupta, Polymer and Composites Rheology, second ed., Marcel Dekker,
New York, 2000.
R.P. Chhabra, Bubbles, Drops and Particles in Non-Newtonian Fluids, CRC
Press, Boca Raton, FL, 2006.
R.P. Chhabra, Fluid ow and heat transfer from circular and non-circular
cylinders submerged in non-Newtonian liquids, Adv. Heat Transfer 43
(2011) 289e417.
R.P. Chhabra, J.F. Richardson, Non-Newtonian Flow in the Process Industries,
Butterworth-Heinemann, Oxford, 2008.
R.P. Chhabra, J.F. Richardson, Non-Newtonian Flow and Applied Rheology,
second ed., Butterworth-Heinemann, Oxford, 2008.
W.H. Boersma, J. Laven, H.N. Stein, Shear-thickening (dilatancy) in concentrated dispersions, AIChE J. 36 (1990) 321e332.
A. Nejat, V. Abdollahi, K. Vahidkhah, Lattice Boltzmann simulation of nonNewtonian ows past conned cylinders, J. Non-Newtonian Fluid Mech. 166
(2011) 689e697.
C.H.K. Williamson, Vortex dynamics in the cylinder wake, Ann. Rev. Fluid
Mech. 28 (1996) 477e539.
R.B. Bird, W.E. Stewart, E.N. Lightfoot, Transport Phenomena, second ed.,
Wiley, New York, 2002.
G.J. Sheard, M.C. Thompson, K. Hourigan, From spheres to circular cylinders:
non-axisymmetric transitions in the ow past rings, J. Fluid Mech. 506 (2004)
45e78.
G.J. Sheard, K. Hourigan, M.C. Thompson, Computations of the drag coefcients for low-Reynolds-number ow past rings, J. Fluid Mech. 526 (2005)
257e275.
L. Zhang, A. Potherat, Inuence of the geometry on the two-and threedimensional dynamics of the ow in a 180 sharp bend, Phys. Fluids 25 (2013)
053605e053629.
B.F. Armaly, F. Durst, J.C.F. Pereira, B. Schonung, Experimental and theoretical
investigation of backward-facing step ow, J. Fluid Mech. 127 (1983)
473e496.
D. Barkley, M.G.M. Gomes, R.D. Henderson, Three-dimensional instability in
ow over a backward-facing step, J. Fluid Mech. 473 (2002) 167e190.
C.C. Wang, Y.J. Chang, Y.C. Hsieh, Y.T. Lin, Sensible heat and friction characteristics of plate n-and-tube heat exchangers having plane ns, Int. J. Refrig.
19 (1996) 223e230.

Вам также может понравиться