You are on page 1of 20

This article was downloaded by: [Statsbiblioteket Tidsskriftafdeling]

On: 07 May 2014, At: 01:25


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Chemical Engineering Communications


Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/gcec20

CFD ANALYSIS OF TWO-DIMENSIONAL


NON-NEWTONIAN POWER-LAW FLOW
ACROSS A CIRCULAR CYLINDER CONFINED
IN A CHANNEL
a a
Sudheer Bijjam & Amit Kumar Dhiman
a
Department of Chemical Engineering , Indian Institute of
Technology Roorkee , India
Published online: 08 Mar 2012.

To cite this article: Sudheer Bijjam & Amit Kumar Dhiman (2012) CFD ANALYSIS OF TWO-DIMENSIONAL
NON-NEWTONIAN POWER-LAW FLOW ACROSS A CIRCULAR CYLINDER CONFINED IN A CHANNEL,
Chemical Engineering Communications, 199:6, 767-785, DOI: 10.1080/00986445.2011.625064

To link to this article: http://dx.doi.org/10.1080/00986445.2011.625064

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the
“Content”) contained in the publications on our platform. However, Taylor & Francis,
our agents, and our licensors make no representations or warranties whatsoever as to
the accuracy, completeness, or suitability for any purpose of the Content. Any opinions
and views expressed in this publication are the opinions and views of the authors,
and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content
should not be relied upon and should be independently verified with primary sources
of information. Taylor and Francis shall not be liable for any losses, actions, claims,
proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or
howsoever caused arising directly or indirectly in connection with, in relation to or arising
out of the use of the Content.

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &
Conditions of access and use can be found at http://www.tandfonline.com/page/terms-
and-conditions
Chem. Eng. Comm., 199:767–785, 2012
Copyright # Taylor & Francis Group, LLC
ISSN: 0098-6445 print=1563-5201 online
DOI: 10.1080/00986445.2011.625064

CFD Analysis of Two-Dimensional Non-Newtonian


Power-Law Flow across a Circular Cylinder
Confined in a Channel

SUDHEER BIJJAM AND AMIT KUMAR DHIMAN


Department of Chemical Engineering, Indian Institute of Technology
Downloaded by [Statsbiblioteket Tidsskriftafdeling] at 01:25 07 May 2014

Roorkee, India

Flow of non-Newtonian power-law fluids across a long circular bluff body confined
symmetrically between infinitely long two parallel plane walls is investigated numeri-
cally by solving the continuity and momentum equations using the finite volume
method–based solver Fluent. The numerical calculations are performed in the full
computational domain for the following ranges: Reynolds number ¼ 50–150 and
power-law index ¼ 0.4–1.8 (covering shear-thinning, Newtonian, and shear-
thickening behaviors) for the blockage ratio of 0.25. Global characteristics such
as drag and lift coefficients, and Strouhal number and derived variables such as
stream function are calculated for the above range of conditions. It is observed that
the shear-thinning behavior yields a lower value of the time-averaged drag coef-
ficient than the corresponding Newtonian value; however, an opposite trend is
observed in the shear-thickening behavior. The Strouhal number increases with
increasing Reynolds number for the fixed value of the power-law index.

Keywords Computational fluid dynamics (CFD); Cylinders; Fluid mechanics;


Laminar flow; Rheology; Vortex flow

Introduction
A variety of non-Newtonian fluids encountered in chemical and related process
industries such as dilute polymer solutions with smaller molecules, syrups, starch
solutions, pulp and paper suspensions, and some emulsions exhibit either shear thinn-
ing and=or shear thickening behavior. Flow of these kind of fluids around a long cir-
cular obstacle has been a research topic of much interest for years because this flow
situation represents an idealization of many flows of practical importance such as
flow meters, tubular and pin-type heat exchangers, filtration screens, and aerosol fil-
ters and instrumentation where small cylindrical wire probes (Zovatto and
Pedrizzetti, 2001) or sensors are made to obstruct the flow. Flow hindrance from
these cylindrical-shape bodies creates vortex-shedding phenomena; this causes press-
ure disturbances around the cylinder as well as vibration. The vortices emanating
from the rear part of the cylinder in the form of wakes persisting up to a certain
length in the downstream enhance mixing and effect the heat and mass transfer

Address correspondence to Amit Kumar Dhiman, Department of Chemical Engineering,


Indian Institute of Technology Roorkee, 247 667, India. E-mail: dhimuamit@rediffmail.com;
amitdfch@iitr.ernet.in

767
768 S. Bijjam and A. K. Dhiman

(Baek and Sung, 1998). Flow-related aspects of such bodies are vital for hydrody-
namic and aerodynamic design and control. The numerical results obtained in the
present study can be used in model equations for designing chemical process equip-
ment to estimate the force exerted by drag and to study the flow pattern behavior of
non-Newtonian power-law fluids.
Owing to its fundamental and practical relevance, there is extensive literature
available on the flow behavior associated with an infinitely long circular cylinder
in an unconfined configuration for both Newtonian (Zdravkovich, 1997, 2003;
Ahmad, 1996; Lange et al., 1998; Norberg, 2001) and non-Newtonian power-law
(Chhabra, 2006; Chhabra et al., 2004; D’Alessio and Pascal, 1996; D’Alessio and
Finlay, 2004; Soares et al., 2005; Whitney and Rodin, 2001; Khan et al., 2006; Bharti
et al., 2006; Sivakumar et al., 2006; Shah et al., 1962; Coelho and Pinho, 2003a, b,
Downloaded by [Statsbiblioteket Tidsskriftafdeling] at 01:25 07 May 2014

2004; Patnana et al., 2009) fluids.


Folowing is a summary of the literature on the flow of non-Newtonian power-
law fluids around a circular cylinder in the unconfined regime (Chhabra et al., 2004;
D’Alessio and Pascal, 1996; D’Alessio and Finlay, 2004; Soares et al., 2005; Whitney
and Rodin, 2001; Khan et al., 2006; Bharti et al., 2006; Sivakumar et al., 2006). In
fact, the results reported elsewhere (D’Alessio and Pascal, 1996) are known to be
erroneous due to the omission of a factor of 2 in one of the authors’ equations. Sub-
sequently, Chhabra et al. (2004) reproduced their study (D’Alessio and Pascal, 1996)
for the identical range of parameters (Re ¼ 1, 20, 40 and n ¼ 0.2–1.4) for drag on the
cylinder. Khan et al. (2006) analytically investigated the unconfined fluid flow and
heat transfer from an isolated circular cylinder submerged in power-law fluids. In
the unsteady regime, Shah et al. (1962) investigated experimentally the cross-flow
of polymer solutions for Reynolds numbers up to about 12000 and found the
variation of pressure along the laminar boundary layer was independent of both
the consistency and power-law indices. Coelho and Pinho (2003a, 2003b, 2004)
experimentally studied the flow of shear-thinning fluids past a circular cylinder in
the Reynolds number range 50 < Re < 9000. They delineated the various vortex-
shedding regimes as a function of the Reynolds and elasticity numbers (Coelho
and Pinho, 2003a). The flow characteristics within each of these flow regimes were
then analyzed in detail (Coelho and Pinho, 2003b). Subsequently, Coelho and
Pinho (2004) also carried out measurements of pressure on the cylinder surface for
Newtonian and non-Newtonian fluids, and the values of the pressure drag
coefficient, pressure rise coefficient, and wake angle were reported.
Considering wall confinement effect on a long circular obstacle for Newtonian
fluids, Chen et al. (1995) carried out numerical experiments to analyze stability as
a function of Reynolds number for steady flow past a circular cylinder symmetrically
placed between parallel planes (0.1  b  0.95). Ben Richou et al. (2004) reported the
flow over a circular cylinder at very low Reynolds numbers (Re  1) and calculated
numerically the drag force exerted by a fluid with a Poiseuille flow profile across a
cylinder placed between two parallel plane walls. Subsequently, Ben Richou et al.
(2005) calculated numerically and asymptotically the wall correction of the drag
force exerted on a circular cylinder moving uniformly midway between two parallel
plane walls at very low Reynolds numbers. Numerical computations have also been
carried out by Chakraborty et al. (2004) for the flow past a circular cylinder for the
blockage ratio range 0.05  b  0.6494, but their results are based on the a priori
assumption of the flow being steady for the Reynolds number range 0.1–200. Khan
et al. (2004) found that the blockage controls the fluid flow and heat transfer and
Non-Newtonian Flow across a Confined Cylinder 769

delays the flow separation from the cylinder placed in the channel. Anagnostopoulos
et al. (1996) investigated the steady and unsteady flow around a circular cylinder by
using the finite element solution for three blockage ratios (b ¼ 0.05, 0.15, and 0.25) at
the fixed Reynolds number of 106 and presented values of hydrodynamic forces on
the cylinder and the Strouhal number. Subsequently, Anagnostopoulos and Minear
(2004) examined the viscous oscillatory flow over a circular cylinder for blockage
ratios in the range 0.10–0.50. Sahin and Owens (2004) showed that for Re  280
and b  0.9, there are at least three separate curves of neutral stability and an
additional increase in the blockage value leads to restabilization to a steady asym-
metric solution. Mettu et al. (2006) reported an unsteady numerical solution for both
momentum and heat transfer from an asymmetrically confined circular cylinder for
the Reynolds number range 10–500 and blockage ratio range 0.1–0.4 for varying gap
Downloaded by [Statsbiblioteket Tidsskriftafdeling] at 01:25 07 May 2014

ratios (0.125–1) and reported parameters, principally drag and Strouhal number.
Kumar and Mittal (2006) investigated the effect of location of the lateral boundaries
(0.005–0.125) on the critical parameters for the instability of the flow past a circular
cylinder. As the blockage ratio increases, the critical Reynolds number for the onset
of the instability first decreases and then increases.
Even though from the above literature considerable information is available for
the flow over a circular cylinder between walls for Newtonian and non-Newtonian
fluids, a gap in information still remained for non-Newtonian power-law fluids in
the confined domain. To fill this gap, Bharti et al. (2007) carried out 2-D numerical
simulation over a confined cylinder for the range 1  Re  40 and 0.2  n  1.9 for
blockage ratios of 0.25, 0.5, 0.625, and 0.9091 in the steady regime and presented
extensive results on drag coefficients and surface pressure coefficients and their
values at stagnation points and flow patterns. More recently, Patnana et al. (2009)
studied the flow over a circular cylinder for higher Reynolds numbers (Re ¼ 40–
140) for power-law fluids in the unconfined unsteady regime, but they did not study
the effect of wall confinement. Thus, based upon the above discussion, it can be con-
cluded here that no one has investigated the flow of non-Newtonian power-law fluids
across a confined circular cylinder in a channel in the unsteady flow regime (i.e., for
the ranges 50 < Re  150 and 0.4  n  1.8).
In the case of Newtonian fluids, Jackson (1987) showed that the vortex-shedding
behavior takes place as the Reynolds number is increased beyond Re47. On the
other hand, the nonlinearity of power-law fluids and the additional frictional losses
at the confinement zone can alter the stability and the nature of the flow in the
Reynolds number range Re ¼ 50–150 and thereby effect the shedding frequency
and the drag behavior compared to Newtonian fluids. On further increasing the
Reynolds number, i.e., Re  150  170, the flow is expected to be three-dimensional.
Thus, the present numerical study is carried out to investigate the forementioned
phenomena and simultaneously to fill the gap in the literature for the range
50  Re  150 and 0.4  n  1.8. A variety of engineering parameters such as
drag and lift coefficients and Strouhal number is reported for the above range of
conditions.

Problem Statement and Governing Equations


The 2-D flow of power-law fluids across an infinitely long circular cylinder confined
in a channel is investigated here, as shown in Figure 1. The length and the width of
the computational domain are defined in terms of the axial and lateral dimensions L
770 S. Bijjam and A. K. Dhiman
Downloaded by [Statsbiblioteket Tidsskriftafdeling] at 01:25 07 May 2014

Figure 1. Schematics of the flow in a channel with a built-in circular cylinder.

and H, respectively. The cylinder is located in the middle at an upstream distance of


Xu from the inlet and at a downstream distance of Xd from the outlet. The total
length of the computational domain is L ¼ Xu þ Xd in the axial direction, and the
blockage ratio here is defined as b ¼ D=H.
The governing equations for the problem under consideration can be written as
(Bird et al., 2001):
Equation of continuity:
rV¼0 ð1Þ
Equation of momentum:
 
@V
q þ V  rV  n  r  r ¼ 0 ð2Þ
@t
where q is the density of the fluid, V is the velocity with components Vx and Vy in the
x- and y-directions, n is the body force, and r is the stress tensor.
The stress tensor r is defined as the sum of the isotropic pressure p and the
deviatoric stress tensor s and is given by
r ¼ pI þ s ð3Þ
The rheological equation of state for incompressible fluids is given as
s ¼ 2geðVÞ ð4Þ
where e(V) are the components of the rate of strain tensor and are defined by

1h i
eðVÞ ¼ ðrVÞ þ ðrVÞT ð5Þ
2
For power-law fluids the viscosity g is given by

 ðn1Þ=2
I2
g¼m ð6Þ
2
Non-Newtonian Flow across a Confined Cylinder 771

where n is the power-law index of the fluid (n < 1 implies a shear-thinning fluid; n ¼ 1
implies a Newtonian fluid; n > 1 implies a shear-thickening fluid) and I2 is the second
invariant of the rate of strain tensor e, which is given as I2 ¼ 2ðe2xx þ e2yy þ e2xy þ e2yx Þ.
The physical boundary conditions for the problem under consideration can be
written as follows (see Figure 1):
. At the inlet boundary: flow is assumed to be fully developed and is given as

  "   #
2n þ 1 
 2y ðnþ1Þ=n
Vx ¼ Uavg 1  1   for 0  y  H and Vy ¼ 0 ð7Þ
nþ1 H

. On the top and bottom boundaries: no-slip boundary condition is applied:


Downloaded by [Statsbiblioteket Tidsskriftafdeling] at 01:25 07 May 2014

Vx ¼ 0 and Vy ¼ 0 ð8Þ

. On the circular obstacle: no-slip boundary condition is used:

Vx ¼ 0 and Vy ¼ 0 ð9Þ

. At the outlet boundary: Based upon the information available in the literature
(Bharti et al., 2007; Patnana et al., 2009, and others), the default value of the
outflow option available in Fluent software is used. This condition is similar to
the homogeneous Neumann conditions:

@Vx @Vy
¼ 0 and ¼0 ð10Þ
@x @x

The numerical simulation starts with initial conditions: Vx ¼ 0, Vy ¼ 0, and p ¼ 0.


The governing equations (1)–(6) along with boundary equations (7)–(10) and
initial conditions are solved for the primitive variables, i.e., velocity (VxandVy)
and pressure (p) fields.

Numerical Details
The problem under consideration is solved by a commercial computational fluid
dynamics (CFD) solver Fluent (6.3). The computational grid is generated by using
GAMBIT, as shown in Figure 2. A very fine grid of cell size (or control volume)
of 0.01D is clustered around the circular obstacle and near the channel walls; how-
ever, the largest grid size used is 0.4D. The two-dimensional, unsteady, laminar, seg-
regated solver is employed to solve the incompressible flow on the collocated grid
arrangement. The second-order upwind scheme is used to discretize convective terms
of momentum equations, whereas the diffusive term is discretized by the central dif-
ference scheme. The first-order implicit time-integration method is used here, and the
dimensionless time step is set to 0.01, since the smaller value of the time step did not
produced any significant change in the values of the physical parameters considered
here. The SIMPLE scheme is used for solving pressure-velocity decoupling. The con-
stant density and non-Newtonian power-law viscosity models are used here. In
addition, the fully developed velocity profile at the channel inlet is represented in
Equation (7) and is incorporated by using user-defined functions available in Fluent.
772 S. Bijjam and A. K. Dhiman
Downloaded by [Statsbiblioteket Tidsskriftafdeling] at 01:25 07 May 2014

Figure 2. Schematic representation of the 2-D nonuniform computational grid structure


(magnified view close to the cylinder).

The resulting algebraic equations are solved by the Gauss-Siedel iterative scheme. The
scaled residuals of the continuity and x- and y-velocities used are of the order of 1020.
Due to the unsteadiness in the wake region and the nonlinearity of power-law
fluids, it is advisable to reduce the residuals to the order of 1015–1020 (Bharti
et al., 2007). The scaled residual is default option in Fluent and is defined as
(Fluent, Inc., 2005)

P P
cells;p j aim vim þ S  ap vp j
R¼ Pim
cells;p jap vp j

where ap is the center coefficient, aim is the neighboring cell coefficient, S is the source
term, and vp is the magnitude of velocities at cell P.
In order to investigate the effect of the grid on the physical parameters, three
grids of grid sizes of 94350, 70660, and 56500 cells are used, with the smallest grid
sizes of 0.008D, 0.01D, and 0.05D clustered around the circular cylinder and near
the top and bottom walls of the channel and having 500, 400, and 300 grid points
on the surface of the cylinder, respectively. The percentage relative changes in the
values of the mean drag coefficient are found to be about 2.3% (94350 cells) and less
than 1.9% (70660 cells) as compared to the value of the mean drag coefficient for the
grid size of 56500 cells for Re ¼ 150 and n ¼ 1.8. However, the corresponding differ-
ences in the values of the Strouhal number are found to be less than 1.1% and less
than 0.5%. Thus, the grid size of 70660 cells is used in this work.
Further, the influence of the upstream distance on the values of physical para-
meters is investigated for Xu ¼ 10D and 15D for Reynolds numbers of 50 and 150
for the highest value of the power-law index of 1.8 used in this work. The relative
percentage changes in the values of the mean drag coefficient and the Strouhal
Non-Newtonian Flow across a Confined Cylinder 773

number are found to be about 0.0002% and negligible, respectively, for Xu ¼ 10D and
15D at Re ¼ 150 and n ¼ 1.8. Similarly, the corresponding differences in the values of
the mean drag coefficient are found to be less than 0.003% for the two values of the
upstream distances at Re ¼ 50 and n ¼ 1.8. Thus, the upstream distance of 10D is
used here.
Similarly, the influence of the downstream distance is examined for Xd ¼ 40D
and 50D for Reynolds numbers of 50 and 150 for the power-law index of 1.8. The
percentage changes in the values of the mean drag coefficient are found to be about
0.0002% for the two values of the downstream distances for Re ¼ 50 and 150. How-
ever, negligible changes in the values of the Strouhal number are observed for
Xd ¼ 40D and 50D for Re ¼ 150 and n ¼ 1.8. Thus, the downstream distance of
40D is used here. The downstream has more length scale than the upstream because
Downloaded by [Statsbiblioteket Tidsskriftafdeling] at 01:25 07 May 2014

the disturbance in the flow in the upstream is less, but at the downstream the disturb-
ance is greater due to the separation of the flow and the vortex shedding. The extra
length in the downstream is required to capture the flow disturbances occurring at
the rear portion of the cylinder and to stabilize and reach the starting initial
conditions.
Thus, in summary, the upstream distance, downstream distance, and blockage
ratio of 10D, 40D, and 0.25 respectively are used in this study. These choices of
upstream and downstream distances and blockage ratio are also consistent with
the literature values (Bharti et al., 2007). The selection of the blockage ratio of
0.25 is also based on the results of Dhiman et al. (2008). Dhiman et al. (2008) inves-
tigated the effect of three values of blockage ratio of 0.125, 0.166, and 0.25 on the
flow and heat transfer across a square cylinder for non-Newtonian power-law fluids.
The effect of blockage ratios of 0.125 and 0.166 is found similar to that of the uncon-
fined steady regime; however, the opposite trends are reported for the blockage ratio
of 0.25. Further, sufficient numerical data is available from Bharti et al. (2007) to
validate the present results in the steady state flow regime (i.e., at Re  40).

Results and Discussion


The 2-D unsteady confined flow of non-Newtonian power-law fluids (0.4  n  1.8) is
simulated here using the full computational domain for the Reynolds number range
50–150 (in steps of 25) for the fixed blockage ratio of 0.25. Validation of the present
numerical results with literature values is reported in the following subsection.

Validation of Results
Comparison of the present confined non-Newtonian power-law results with the
results of Bharti et al. (2007) for the highest Reynolds number of 40 used by them
for varying values of the power-law index for the blockage ratio of 0.25 in the steady
flow regime is presented in Table I. This table shows the values of flow parameters
such as total drag coefficient and its individual components (skin friction and press-
ure drag coefficients) based on the Reynolds number defined on the maximum velo-
city. Excellent agreement can be seen between the present results and the values
reported in the literature (Table I). For instance, the maximum differences in the
values of friction, pressure, and mean drag coefficients are found to be less than
2.7%, 1.2%, and 0.7%, respectively for varying values of power-law indices
(0.4  n  1.8).
774 S. Bijjam and A. K. Dhiman

Table I. Comparison of present results with literature values in the steady confined
flow regime

Source n CDF CDP CD


Present 0.4 0.1891 1.0897 1.2788
Bharti et al. (2007) 0.1842 1.1024 1.2866
Present 0.8 0.3944 1.1477 1.5421
Bharti et al. (2007) 0.3908 1.151 1.5418
Present 1 0.5147 1.1892 1.7039
Bharti et al. (2007) 0.5109 1.1925 1.7034
Present 1.2 0.6436 1.2345 1.8781
Bharti et al. (2007) 0.6402 1.2391 1.8793
Downloaded by [Statsbiblioteket Tidsskriftafdeling] at 01:25 07 May 2014

Present 1.8 1.0761 1.4009 2.477


Bharti et al. (2007) 1.0711 1.4054 2.4765

Since no work is available for the flow of non-Newtonian power-law fluids


across a circular cylinder for the range of conditions studied here in the unsteady
confined flow regime, additional computations were made in the unsteady uncon-
fined regime for Re ¼ 100 and n ¼ 0.4–1.8. For this, numerical calculations were
carried out by using the same computational domain as Patnana et al. (2009) in
the unsteady unconfined flow regime. Validation of the present unconfined unsteady
results is reported in Table II with the only non-Newtonian power-law results avail-
able, from Patnana et al. (2009), in the unconfined unsteady regime along with the
results of Sivakumar et al. (2006), Mettu et al. (2006), Clift and Weber (1978),
and Mittal (2005) for Newtonian values. Table II compares the present values of
the time-averaged drag coefficient and the Strouhal number for the Reynolds num-
ber of 100 for varying values of the power-law index. It can be seen from this table
that excellent agreement exists between the present and literature values. For
instance, the maximum relative changes in the values of the time-averaged drag coef-
ficient and the Strouhal number are found to be less than 4.1% and 3.65% respect-
ively with the results of Patnana et al. (2009) for the Reynolds number of 100 for

Table II. Comparison of present results with literature values in the unsteady
unconfined flow regime

Source n CD St
Present 0.4 1.0902 0.1992
Patnana et al. (2009) 1.1345 0.2067
Present 1 1.3063 0.1626
Patnana et al. (2009) 1.3409 0.1657
Clift and Weber (1978) 1.3300 0.167
Mettu et al. (2006) 1.3020 0.1600
Sivakumar et al. (2006) 1.3250 0.1641
Mittal (2005) 1.3220 0.1644
Present 1.8 1.5627 0.1344
Patnana et al. (2009) 1.6294 0.1392
Non-Newtonian Flow across a Confined Cylinder 775

n ¼ 0.4, 1, and 1.8. This further validates the present numerical solution procedure
and assures us that the grid and numerical parameters reported here are trustworthy.

Flow Patterns
Figures 3–6 represent the streamline profiles in the vicinity of the confined circular cy-
linder for the flow of power-law fluids for varying Reynolds numbers and power-law
indices. The flow is found to be steady, as two symmetric vortices are formed behind
the cylinder for the entire range of the power-law index considered here (i.e., n ¼ 0.4–
1.8) for the fixed Reynolds number of 50, as shown in Figure 3. In other words, this is
steady state due to the reduction of disturbances and the increase in flow stability due
Downloaded by [Statsbiblioteket Tidsskriftafdeling] at 01:25 07 May 2014

Figure 3. Streamline contours for Re ¼ 50 at various values of the power-law index.


776 S. Bijjam and A. K. Dhiman
Downloaded by [Statsbiblioteket Tidsskriftafdeling] at 01:25 07 May 2014

Figure 4. Streamline contours representing steady state for Re ¼ 75 (n ¼ 1.4–1.8), 100


(n ¼ 1.6–1.8), and 125 (n ¼ 1.8).

the confined walls and is unlike the unconfined flow situation where for Re ¼ 50 the
unsteady state persists (Patnana et al., 2009). The size of these symmetric vortices
decreases with increasing value of the power-law index. On gradually increasing the
value of the Reynolds number (Re > 50), the flow is found to be unsteady periodic
for the power-law index range 0.4  n  1.2 and for the range 1.2 < n  1.8
(Figures 4(a)–(c)), the flow becomes steady at the Reynolds number of 75. This is
due to the existence of stationary channel walls and the higher damping nature of
the effective viscosity of the shear-thickening fluids; it is increased with the increasing
value of the power-law index and reduces with increasing Reynolds number. On the
other hand, the flow is found to be unsteady periodic in the unconfined regime for
the above range of conditions (Patnana et al., 2009). Figure 5 presents the instan-
taneous streamline profiles for four successive moments of time that span over the
whole period for 0.4  n  1.4 at Re ¼ 100. At this Reynolds number of 100, the flow
is found to be unsteady periodic for the range 0.4  n  1.4, and it becomes steady for
the range 1.4 < n  1.8 (Figures 4(d) and 4(e)). Similarly, the flow is found to be
unsteady periodic for 0.4  n  1.6, and it becomes steady again for the range
1.6 < n  1.8 (Figure 4(f)) at Re ¼ 125. On further increasing the value of the Reynolds
number, the flow is found to be unsteady (periodic) for all the values of power-law
indices studied here for the Reynolds number of 150 (see Figure 6).
Further, in the unsteady flow regime, as the value of the power-law index and=or
Reynolds number increases, the recirculation=wake region behind the cylinder
increases, and this is in line with the unconfined circular cylinder case (Patnana
et al., 2009). However, complex dependence of the wake phenomena is observed
for n < 1for Re > 50 in the unsteady periodic regime (see Figures 5(a)–(d) and
6(a)–(d)). For instance, an additional appearance of the small wake behind the
Non-Newtonian Flow across a Confined Cylinder 777
Downloaded by [Statsbiblioteket Tidsskriftafdeling] at 01:25 07 May 2014

Figure 5. Instantaneous streamline contours for Re ¼ 100 at various values of power-law


index.

cylinder can be seen in these figures at times t ¼ T and t ¼ 2T=4. In general, instan-
taneous flow patterns in the unsteady (periodic) confined flow regime are observed
similar to those in the unsteady unconfined flow regime.

Temporal Drag and Lift Coefficients


Temporal variations of drag and lift coefficients for Reynolds numbers of 75, 100,
and 150 for different values of power-law indices at the fixed blockage ratio of
0.25 are shown in Figures 7–9. For the temporal variation, instantaneous values
of the drag and lift coefficients are calculated at each time step and plotted versus
time. In the present study, the previously converged velocity and pressure fields
are used for the generation of new results. Therefore, the results of oscillating drag
and lift coefficients are shown after these oscillations stabilized (Figures 7–9). In the
unsteady periodic regime, the simulation run is considered stable and converges after
the completion of at least 10 successive uniform periodic cycles for both drag and lift
coefficients. Since the flow is found to be steady at Re ¼ 50 for 0.4  n  1.8, the vor-
tex shedding (periodicity) does not take place in the flow for these ranges of con-
ditions. It can be seen from these figures that the amplitude of the lift coefficient
increases with increasing value of the Reynolds number for the fixed value of the
power-law index due to the confinement of the channel walls. However, the ampli-
tude of the lift coefficient decreases from shear-thinning behavior to Newtonian
778 S. Bijjam and A. K. Dhiman
Downloaded by [Statsbiblioteket Tidsskriftafdeling] at 01:25 07 May 2014

Figure 6. Instantaneous streamline contours for Re ¼ 150 at various values of power-law


index.

and then to shear-thickening behavior for the fixed value of the Reynolds number.
On the other hand, the amplitude of the drag coefficient increases for shear-thinning
behavior, and it decreases for Newtonian and shear-thickening behaviors with
increasing value of the Reynolds number. However, the amplitude of the drag coef-
ficient increases from shear-thinning behavior to Newtonian and then to
shear-thickening behavior for the fixed value of the Reynolds number.

Time-Averaged Drag and Lift Coefficients


The time-averaged total drag is the contribution of two components: friction drag
and pressure drag. The variation of the time-averaged drag coefficient with varying
values of Reynolds number and power-law index for a fixed blockage ratio of 0.25 is
shown in Figure 10(a). At low values of the power-law index (n  0.4), as the
Reynolds number increases the mean drag coefficient increases linearly due to the
small wake formation on the surface of the cylinder (see Flow Patterns subsection
above). However, for the power-law index of 0.6, the total drag coefficient increases
with increasing Reynolds number for the range Re > 50. Further, this phenomenon
is similar to unconfined flow over square and circular cylinders reported earlier by
Sahu et al. (2009) and Patnana et al. (2009), respectively. The increase in pressure
drag compared to friction drag with Reynolds number at fixed n for high
shear-thinning fluids is the cause of an increase in total drag with increase in
Non-Newtonian Flow across a Confined Cylinder 779
Downloaded by [Statsbiblioteket Tidsskriftafdeling] at 01:25 07 May 2014

Figure 7. Temporal variation of drag and lift coefficients with power-law index at Re ¼ 75.

Reynolds number. However, with increasing value of n (i.e., n  0.8), the pressure
drag decreases with the increase in Reynolds number. In the present study, for
n  0.8, the value of the total drag coefficient decreases from shear-thinning behavior
to Newtonian and then to shear-thickening behavior with increasing value of the
Reynolds number. It is also observed that the total drag increases with increasing
value of the power-law index for the fixed value of the Reynolds number. Similar
to the case of an unconfined circular cylinder, shear-thinning behavior always yields
a lower value of the time-averaged drag coefficient than the corresponding Newto-
nian value; however, the opposite trend is seen in shear-thickening behavior. On
the other hand, the value of the time-averaged lift coefficient remains approximately
zero (in the order of about 104–103) for the range of conditions covered here.

RMS Values of Drag and Lift Coefficients


In order to present the measure of the amplitude of the unsteady cylinder wake oscil-
lations, the root mean square (rms) values of both drag and lift coefficients for
780 S. Bijjam and A. K. Dhiman
Downloaded by [Statsbiblioteket Tidsskriftafdeling] at 01:25 07 May 2014

Figure 8. Temporal variation of drag and lift coefficients with power-law index at Re ¼ 100.

different Reynolds numbers and power-law indices are presented in Figures 10(b)
and 11(a), respectively. The rms value is the statistical measure of the coefficients
of drag and lift from the mean values. The time-averaged and the rms values of
any quantity u are calculated as follows:

vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
uN
P
N uP
ui u ðui  umean Þ2
i¼1
ti¼1
umean ¼ and urms ¼
N N

where N is the total number of time steps in the 10 cycles.


It can be observed that the rms value of the drag coefficient increases linearly
with increasing value of the Reynolds number for the fixed value of the flow beha-
vior index; however, it decreases with increasing value of the power-law index for the
fixed value of the Reynolds number. Similar to the rms value of the drag coefficient,
Non-Newtonian Flow across a Confined Cylinder 781
Downloaded by [Statsbiblioteket Tidsskriftafdeling] at 01:25 07 May 2014

Figure 9. Temporal variation of drag and lift coefficients with power-law index at Re ¼ 150.

the rms value of the lift coefficient increases with increasing value of the Reynolds
number for the fixed value of the power-law index and the rms value of the lift
coefficient decreases with increasing value of the power-law index for the fixed value
of the Reynolds number.

Strouhal Number
The Strouhal number (St) is the dimensionless parameter used to measure the
frequency of the vortex shedding and is defined as
fD
St ¼
Uavg
where f is the frequency of the vortex shedding and D is the diameter of the cylinder.
In the present study, the temporal variation of the lift coefficient is used to cal-
culate the frequency of the vortex shedding (f) by using fast Fourier transformation.
782 S. Bijjam and A. K. Dhiman
Downloaded by [Statsbiblioteket Tidsskriftafdeling] at 01:25 07 May 2014

Figure 10. Variation of (a) time-averaged drag coefficient and (b) rms value of the drag
coefficient with Reynolds number and power-law index.

The variation of the Strouhal number with Reynolds number at various values of the
power-law index is shown in Figure 11(b) for the fixed blockage ratio of 0.25. The
Strouhal number becomes zero at the steady state due to the attachment of vortices
symmetrically to the rear part of the cylinder about the mid-plane. As the value of
the Reynolds number gradually increases, the value of the Strouhal number increases
for the fixed value of the power-law index, while the Strouhal number has a complex
dependence on the Reynolds number in the unconfined circular cylinder case
(Patnana et al., 2009). Similar to the unconfined case, the Strouhal number increases
with decreasing value of the power-law index for the fixed value of the Reynolds
number. However, this is not valid here for Re ¼ 75 and n ¼ 0.4. For instance, the
Strouhal number for n ¼ 0.4 at Re ¼ 75 is lower than that of n ¼ 0.6 and 0.8.

Figure 11. Variation of (a) rms value of the lift coefficient and (b) Strouhal number with
Reynolds number and power-law index.
Non-Newtonian Flow across a Confined Cylinder 783

Conclusions
The 2-D flow of power-law fluids over a circular cylinder confined in a channel is
studied for a varying range of Reynolds numbers (50  Re  150) and power-law
indices (0.4  n  1.8) for the fixed blockage ratio of 0.25. Extensive numerical results
are presented such as total drag and lift coefficients and Strouhal number. The
present study clearly elucidates the role of wall confinement and the effect of
Reynolds number on power-law fluids. The important phenomena that can be
observed are, broadly, that the instantaneous flow patterns in the unsteady confined
flow regime are similar to those in the unsteady unconfined flow regime. For highly
shear-thinning fluids (n  0.6), the total drag coefficient increases with increasing
value of Reynolds number; however, for power-law fluids (n > 0.6) the drag coef-
Downloaded by [Statsbiblioteket Tidsskriftafdeling] at 01:25 07 May 2014

ficient decreases on increasing the Reynolds number. Shear-thinning behavior


always yields a lower value of the time-averaged drag coefficient than the corre-
sponding Newtonian value; however, an opposite trend is seen in shear-thickening
behavior. With an increase in the value of the Reynolds number, the Strouhal num-
ber increases for the fixed value of the power-law index; however, the Strouhal
number decreases with increasing value of the power-law index for the fixed value
of the Reynolds number.

Nomenclature
2
CD total drag coefficient, 2FD =qUavg D
CDrms rms value of the drag coefficient, dimensionless
CDF friction drag coefficient, dimensionless
CDP pressure drag coefficient, dimensionless
2
CL total lift coefficient, 2FL =qUavg D
CLrms rms value of the lift coefficient, dimensionless
D diameter of the cylinder, m
f frequency of vortex shedding, s1
FD drag force per unit length of the cylinder, N=m
FL lift force per unit length of the cylinder, N=m
H height of the computational domain, m
I2 second invariant of the rate of strain tensor, s2
L length of the computational domain, m
m power-law consistency index, Pa sn
n power-law index, dimensionless
p pressure, Pa
Re Reynolds number, qDn Uavg 2n
=m
St Strouhal number, fD=Uavg
t time, s
T time period, s
Uavg average inlet velocity of the fluid, m=s
Vx components of the velocity in the x-direction, m=s
Vy components of the velocity in the y-direction, m=s
x stream-wise coordinate, m
Xd downstream length, m
Xu upstream length, m
y transverse coordinate, m
784 S. Bijjam and A. K. Dhiman

Greek letters
b blockage ratio, D=H
g power-law viscosity, Pa s
q density of the fluid, kg=m3
s shear stress, Pa

References
Ahmad, R. A. (1996). Steady-state numerical solution of the Navier-Stokes and energy equa-
tions around a horizontal cylinder at moderate Reynolds numbers from 100 to 500, Heat
Transfer Eng., 17, 31–81.
Anagnostopoulos, P., and Minear, R. (2004). Blockage effect of oscillatory flow past a fixed
Downloaded by [Statsbiblioteket Tidsskriftafdeling] at 01:25 07 May 2014

cylinder, Appl. Ocean Res., 26, 147–153.


Anagnostopoulos, P., Iliadis, G., and Richardson, S. (1996). Numerical study of the blockage
effects on viscous flow past a circular cylinder, Int. J. Numer. Methods Fluids, 22,
1061–1074.
Baek, S. J., and Sung, H. J. (1998). Numerical simulation of the flow behind a rotary
oscillating circular cylinder, Phys. Fluids, 10, 869–876.
Ben Richou, A., Ambari, A., and Naciri, J. K. (2004). Drag force on a circular cylinder mid-
way between two parallel plates at very low Reynolds numbers. Part 1: Poiseuille flow
(numerical), Chem. Eng. Sci., 59, 3215–3222.
Ben Richou, A., Ambari, A., Lebey, M., and Naciri, J. K. (2005). Drag force on a circular
cylinder midway between two parallel plates at Re  1. Part 2: Moving uniformly
(numerical and experimental), Chem. Eng. Sci., 60, 2535–2543.
Bharti, R. P., Chhabra, R. P., and Eswaran, V. (2006). Steady flow of power-law fluids across
a circular cylinder, Can. J. Chem. Eng., 84, 406–421.
Bharti, R. P., Chhabra, R. P., and Eswaran, V. (2007). Two-dimensional steady Poiseuille flow
of power-law fluids across a circular cylinder in a plane confined channel: Wall effects and
drag coefficients, Ind. Eng. Chem. Res., 46, 3820–3840.
Bird, R. B., Stewart, W. E., and Lightfoot, E. N. (2001). Transport Phenomena, 2nd ed., Wiley,
New York.
Chakraborty, J., Verma, N., and Chhabra, R. P. (2004), Wall effects in the flow past a circular
cylinder in a plane channel: A numerical study, Chem. Eng. Process., 43, 1529–1537.
Chen, J. H., Pritchard, W. G., and Tavener, S. J. (1995). Bifurcation for flow past a cylinder
between parallel planes, J. Fluid Mech., 284, 23–41.
Chhabra, R. P. (2006), Bubbles, Drops and Particles in Non-Newtonian Fluids, 2nd ed., CRC
Press, Boca Raton, Fla.
Chhabra, R. P., Soares, A. A., and Ferreira, J. M. (2004), Steady non-Newtonian flow past a
circular cylinder: A numerical study, Acta Mech., 172, 1–16.
Clift, G. J., and Weber, M. E. (1978). Bubbles, Drops and Particles, Academic Press, New
York.
Coelho, P. M., and Pinho, F. T. (2003a). Vortex shedding in cylinder flow of shear-thinning
fluids, I. Identification and demarcation of flow regimes, J. Non-Newton. Fluid Mech.,
110, 143–176.
Coelho, P. M., and Pinho, F. T. (2003b). Vortex shedding in cylinder flow of shear-thinning
fluids, II. Flow characteristics, J. Non-Newton. Fluid Mech., 110, 177–193.
Coelho, P. M., and Pinho, F. T. (2004). Vortex shedding in cylinder flow of shear-thinning
fluids, III. Pressure measurements, J. Non-Newton. Fluid Mech., 121, 55–68.
D’Alessio, S. J. D., and Finlay, L. A. (2004). Power-law flow past a cylinder at large distances,
Ind. Eng. Chem. Res., 43, 8407–8410.
D’Alessio, S. J. D., and Pascal, J. P. (1996). Steady flow of a power-law fluid past a cylinder,
Acta Mech., 117, 87–100.
Non-Newtonian Flow across a Confined Cylinder 785

Dhiman, A. K., Chhabra, R. P., and Eswaran, V. (2008). Steady flow across a confined square
cylinder: Effects of power-law index and of blockage ratio, J. Non-Newton. Fluid Mech.,
148, 141–150.
Fluent, Inc. (2005). User’s Manual to FLUENT 6.3, Fluent Inc., Lebanon, N.H.
Jackson, C. P. (1987). A finite-element study of the onset of vortex shedding in flow past
variously shaped bodies, J. Fluid. Mech., 182, 23–45.
Khan, W. A., Culham, J. R., and Yovanovich, M. M. (2004). Fluid flow and heat transfer
from a cylinder between parallel planes, J. Thermophys. Heat Transfer, 18, 395–403.
Khan, W. A., Culham, J. R., and Yovanovich, M. M. (2006). Fluid flow and heat transfer
in power-law fluids across circular cylinders: Analytical study, J. Heat Transfer, 128,
870–878.
Kumar, B., and Mittal, S. (2006). Effect of blockage on critical parameters for flow past a
circular cylinder, Int. J. Numer. Methods Fluids, 50, 987–1001.
Downloaded by [Statsbiblioteket Tidsskriftafdeling] at 01:25 07 May 2014

Lange, C. F., Durst, F., and Breuer, M. (1998). Momentum and heat transfer from cylinders
in laminar cross-flow at 104  Re  200, Int. J. Heat Mass Transfer, 41, 3409–3430.
Mettu, S., Verma, N., and Chhabra, R. P. (2006). Momentum and heat transfer from an
asymmetrically confined circular cylinder in a plane channel, Heat Mass Transfer, 42,
1037–1048.
Mittal, S. (2005). Excitation of shear layer instabilities in flow past a cylinder at low Reynolds
number, Int. J. Numer. Methods Fluids, 49, 1147–1167.
Norberg, C. (2001). Flow around a circular cylinder: Aspects of fluctuating lift, J. Fluids
Struct., 15, 459–469.
Patnana, V. K., Bharti, R. P., and Chhabra, R. P. (2009). Two-dimensional unsteady flow of
power-law fluids over a cylinder, Chem. Eng. Sci., 64, 2978–2999.
Sahin, M., and Owens, R. G. (2004). Numerical investigations of wall effects up to high block-
age ratios on two-dimensional flow past a confined circular cylinder, Phys. Fluids, 16,
1305–1320.
Sahu, A. K., Chhabra, R. P., and Eswaran V. (2009). Two-dimensional unsteady laminar flow
of a power-law fluid across a square cylinder, J. Non-Newton. Fluid Mech., 160, 157–167.
Shah, M. J., Petersen, E. E., and Acrivos, A. (1962). Heat transfer from a cylinder to a
power-law Non-Newtonian fluid, AIChE J., 8, 542–549.
Sivakumar, P., Bharti, R. P., and Chhabra, R. P. (2006). Effect of power-law index on critical
parameters for power-law flow across an unconfined circular cylinder, Chem. Eng. Sci.,
61, 6035–6046.
Soares, A. A., Ferreira, J. M., and Chhabra, R. P. (2005). Flow and forced convection heat
transfer in cross flow of non-Newtonian fluids over a circular cylinder, Ind. Eng. Chem.
Res., 44, 5815–5827.
Whitney, M. J., and Rodin, G. J. (2001). Force velocity relationships for rigid bodies translat-
ing through unbounded shear-thinning power-law fluids, Int. J. Non-linear Mech., 34,
947–953.
Zdravkovich, M. M. (1997). Flow around Circular Cylinders: Fundamentals, Oxford University
Press, New York.
Zdravkovich, M. M. (2003). Flow around Circular Cylinders: Applications, Oxford University
Press, New York.
Zovatto, L., and Pedrizzetti, G. (2001). Flow around a circular cylinder between parallel walls,
J. Fluid Mech., 440, 1–25.

You might also like