You are on page 1of 20

Engineering Applications of Computational Fluid

Mechanics

ISSN: 1994-2060 (Print) 1997-003X (Online) Journal homepage: http://www.tandfonline.com/loi/tcfm20

Effect of Blockage Ratio on Drag and Heat Transfer


from a Centrally Located Sphere in Pipe Flow
Suresh Krishnan & A. Kaman
To cite this article: Suresh Krishnan & A. Kaman (2010) Effect of Blockage Ratio on Drag
and Heat Transfer from a Centrally Located Sphere in Pipe Flow, Engineering Applications of
Computational Fluid Mechanics, 4:3, 396-414, DOI: 10.1080/19942060.2010.11015327
To link to this article: http://dx.doi.org/10.1080/19942060.2010.11015327

Copyright 2010 Taylor and Francis Group


LLC
Published online: 19 Nov 2014.

Submit your article to this journal

Article views: 1066

View related articles

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=tcfm20
Download by: [b-on: Biblioteca do conhecimento online UBI]

Date: 18 July 2016, At: 07:33

Engineering Applications of Computational Fluid Mechanics Vol. 4, No. 3, pp. 396414 (2010)

EFFECT OF BLOCKAGE RATIO ON DRAG AND HEAT TRANSFER


FROM A CENTRALLY LOCATED SPHERE IN PIPE FLOW
Suresh Krishnan and Kannan A.*

Downloaded by [b-on: Biblioteca do conhecimento online UBI] at 07:33 18 July 2016

Department of Chemical Engineering, Indian Institute of Technology Madras,


Chennai 600036, India
* E-Mail: kannan@iitm.ac.in (Corresponding Author)
ABSTRACT: Hydrodynamic and heat transfer analyses were carried out for laminar fluid flow past a heated
sphere placed centrally in a pipe using Computational Fluid Dynamics (CFD) simulations. Fully developed parabolic
velocity profile characteristic of laminar flow was specified at the pipe inlet. The effects of blockage ratio, defined as
the ratio of the sphere diameter to the pipe diameter, on the drag coefficient and Nusselt number are reported. The
particle Reynolds numbers were varied up to a maximum of 500 while the Prandtl number for the water system
studied was fixed at 5.12. At lower blockage ratios (<0.3), transient and three-dimensional simulations were required
for particle Reynolds numbers exceeding 270. The effect of blockage was more significant at lower particle
Reynolds numbers and influenced the drag coefficient more than the Nusselt number for this system. At high particle
Reynolds numbers, reasonable predictions of the drag coefficient could be obtained even at high blockage ratios, by
using the standard correlations available for a sphere immersed in an unbounded domain and the fluid approaching
the sphere with a uniform velocity profile. Conversely, at low blockage ratios and low Reynolds numbers,
considerable deviation from the standard drag correlation predictions was observed. Higher blockage ratios delayed
boundary layer separation. The local Nusselt number distribution trends along the sphere are also reported.
Correlations were developed for the drag coefficient, boundary layer separation angle and the average Nusselt
number as a function of the blockage ratio and particle Reynolds number.
Keywords:

blockage ratio, CFD, drag coefficient, local Nusselt number, angle of boundary layer separation

ratio defined as the ratio of the particle diameter


to that of the enclosure may also influence the
flow and heat transfer characteristics.
In the design of processes involving momentum
and heat transfer between relatively large particles
and confined tubes, important parameters such as
drag coefficient and Nusselt numbers were
unavailable for accurate modeling work. This
causes a serious constraint in the process design
of industrial applications. One example is aseptic
food processing where solid food particle is
immersed in a holding tube (Salengke and Sastry,
1996; Balasubramaniam and Sastry, 1996;
Ramaswamy et al., 1997). In this application,
suspended food particulate is thermally sterilized
due to the heating provided by the surrounding
carrier fluid. This occurs inside a holding tube. In
experimental studies of the heat transfer
phenomena in such processes, usually a particle is
fixed in the center of the tube and the fluid flows
around it. This approach has been adopted in
literature, for example by Zuritz et al. (1990),
Awuah et al. (1995), Ramaswamy et al. (1996)
and Chen et al. (1997). Correlations for drag
coefficients and Nusselt numbers at different
blockage ratios and higher particle Reynolds

1. INTRODUCTION
Movement of a fluid past a heated sphere has
been widely investigated under various conditions
due to its importance in both engineering
applications and basic research. In addition to the
hydrodynamics pertaining to the flow around the
sphere as characterized by the drag coefficient
plotted as a function of Reynolds number, the
heat transfer characteristics from the sphere to the
surrounding fluid is also of practical interest. Both
momentum and heat transfer phenomena are
complicated by various factors such as particle
size, fluid velocity, rheological properties of the
fluid and the entering fluids velocity profile.
While a sphere suspended in a uniform flow field
has been widely reported in the past (Whitaker,
1972; Seelay et al., 1975; Clift et al., 1978;
Johnson and Patel, 1999; Feng and Michaelides,
2000; Dhole et al., 2006), relatively few studies
are available on the hydrodynamic and heat
transfer characteristics around a sphere immersed
in a confined enclosure wherein the fluid velocity
profile is not necessarily flat (Oh and Lee, 1988;
Wham et al., 1996; Shahcheraghi and Dwyer,
1998; Bharti et al., 2007). Further, the blockage

Received: 21 Jan. 2010; Revised: 16 Mar. 2010; Accepted: 7 Apr. 2010


396

Downloaded by [b-on: Biblioteca do conhecimento online UBI] at 07:33 18 July 2016

Engineering Applications of Computational Fluid Mechanics Vol. 4, No. 3 (2010)

number values were not far different when


compared to those obtained with uniform flow
around the sphere. Maheshwari et al. (2006)
studied the effect of blockage on fluid flow and
heat transfer around single and three sphere
systems in the Reynolds number range of 1-100.
Blockage ratios were in the range of 0.1 to 0.5
and Prandtl numbers corresponding to those of
water and air were used. The confining walls were
assumed to move at the same velocity as that of
the fluid. They also observed that the blockage
seemed to affect the drag more than the heat
transfer.
Information is scarce on the effect of the sphere
diameter relative to the confining pipe diameter
on the hydrodynamics and heat transfer
characteristics. The drag coefficient and local
Nusselt numbers are not well established
especially at higher Reynolds numbers (>150)
when three dimensional effects are expected to
assume importance. The scope of the current
study is to investigate the effects of blockage
when a sphere is fixed at the center of a pipe and
the fluid enters it in well developed Poiseuille
flow. A broader range of pipe and particle
Reynolds numbers is considered in this work. The
hydrodynamic and heat transfer results are
compared with previous results that considered
blockage over a narrow range of Reynolds
numbers, for purpose of validation. Further the
results obtained here are also compared with the
classic results obtained for uniform flow around
an unconfined sphere. The objectives of the work
are summarized below:

numbers were not available. Also unavailable


were the extent of inaccuracy in using standard
drag correlations that are applicable for bodies
immersed in an unbounded fluid domain where
the fluid approaching the sphere has a flat
velocity profile, to situations involving a bounded
fluid domain and the fluid entering it with a
parabolic velocity profile. The scope of this work
is to understand the underlying transport
phenomena (such as flow and thermal patterns,
boundary layer separation angle) and develop
correlations for drag coefficient and Nusselt
numbers in confined flows which will lead to
improved design of processes. This work has
broad applicability not only in food processing
but also in other areas such as fluidization and
combustion (Bagchi et al., 2001; Hessel et al.,
2005).
2. BACKGROUND AND SCOPE
Early studies on the effect of the tube wall on the
particle drag were mainly confined to creeping
flows (Wakiya, 1957; Haberman and Sayre, 1958).
Fayon and Happel (1960) provided a corrected
correlation for the drag coefficient at low particle
Reynolds numbers (<30) when the sphere was
confined inside a tube. This correlation was found
to be increasingly inaccurate at higher Reynolds
numbers. Wham et al. (1996) studied the wall
effects when there was axisymmetric flow around
a stationary sphere. The particle Reynolds number
in their work was extended to 100 and the effect
of blockage ratio from 0.08 to 0.7 was studied.
However, the heat transfer analysis was not
considered in their work. Oh and Lee (1988)
studied experimentally and numerically the
Newtonian fluid flow past a sphere in a pipe. Flat
velocity profile was specified at the inlet and exit
boundaries. A restricted range of Reynolds
number (20 -130) and blockage ratio (0, 0.5 and
0.74) were considered. Increasing pipe blockages
were found to lead to higher drag coefficients.
Shahcheraghi and Dwyer (1998) investigated
hydrodynamics and heat transfer to air from an
eccentrically located sphere in laminar pipe flow.
The effects of different particle Reynolds
numbers (25 and 125), blockage ratios (0.2 and
0.4) and particle eccentricities on the drag
coefficient and Nusselt numbers were reported.
They observed that when the sphere was located
at an off-centered position, blockage ratios greater
than 0.04 began to influence the pressure drag and
delayed the boundary layer separation around the
sphere. In the range of Reynolds number
investigated in their work, area averaged Nusselt
397

Carry out time step and mesh size


independent CFD simulations on confined
flow occurring around a sphere located
centrally in a pipe.

Study the effect of blockage ratio, defined as


the ratio of the sphere diameter to the pipe
diameter, on the drag coefficient and Nusselt
number in the particle Reynolds number range
of 0.1 Rep 500. The blockage ratios varied
from 0.02 (corresponding to negligible
blockage) to 0.5 (maximum blockage).

Compare the drag coefficients and Nusselt


numbers obtained in the present simulations
(which account for blockage) with the
classical solutions obtained for unconfined
sphere immersed in a uniform fluid flow. The
unconfined case involving uniform fluid flow
is termed the classical case and the results on
the associated drag coefficients and Nusselt
numbers are termed classical solutions.

Engineering Applications of Computational Fluid Mechanics Vol. 4, No. 3 (2010)

approaching the sphere is uniform. When the


sphere is located at the center line of the pipe, the
particle Reynolds number is expressed in terms of
the maximum velocity of the laminar flow (Vch =
2 Vavg). When a very small sphere is immersed in
a pipe, i.e. the blockage ratio is small, it is
exposed to essentially fluid velocities that are
very close to that of the maximum fluid velocity.
Even for reasonable blockage ratios (BR<0.3), the
sphere is exposed to a velocity profile closer to
the maximum velocity rather than the average
velocity. Hence for consistency, the maximum
velocity is used to define Rep when the particle is
located at the centre of the sphere. Further, the
simple relation of maximum velocity being twice
the average velocity in laminar pipe flow enables
correlations developed with maximum velocity
based Reynolds number be simply adapted to
cases where average velocity is used to define the
Reynolds number. The fully developed laminar
flow velocity profile as function of radial distance
is given by

Explain the role of the confining wall on the


drag and heat transfer.

3. PROBLEM FORMULATION

Downloaded by [b-on: Biblioteca do conhecimento online UBI] at 07:33 18 July 2016

Incompressible laminar flow of a Newtonian


liquid past a sphere placed in a pipe is simulated
numerically. The thermophysical properties of the
fluid are assumed constant with respect to
temperature and given in Table 1. To prevent
possible loss of information due to transition from
two dimensional to three dimensional flow at
different blockage ratios, all the simulations were
carried out in 3-D.
3.1

Blockage ratio

Different diameters of the sphere were used in the


simulation while the pipe diameter was
maintained constant (at 50.8 mm). The ratio of the
sphere diameter to that of the pipe diameter is
defined as the blockage ratio (BR). Blockage
ratios in the range of 0.1-0.5 were investigated.
Results from blockage ratio of 0.02 corresponded
to classical solution results.
3.2

R 2
Vx Vmax 1

Rmax

Particle and Pipe Reynolds numbers

The assumption of laminar flow for Reynolds


numbers up to 500 was checked. A laminar
boundary layer developed which rapidly
thickened as the separation point was reached.
Seeley et al. (1975) observed based on
experimental results that there were no turbulent
disturbances either in the boundary layer or in the
free stream at particle Reynolds numbers of 760.
Further, the presence of solid wall in confined
flows will further contribute towards flow
stabilization thereby ensuring laminar flow
regime.

Two Reynolds numbers that characterize the flow


in the pipe as well as flow around the sphere are:
Particle Reynolds number (Rep):

Rep =

Dp Vch

(1)

where Vch is the characteristic velocity and is


taken to be the maximum velocity of the
incoming fully developed laminar flow.
Pipe Reynolds number (ReT):

ReT =

DT Vavg

(3)

(2)

Table 1 Physical properties of water at 30C.

The drag coefficient and Nusselt numbers are


expressed as functions of Rep, while ReT is used to
ensure that the flow entering the pipe is laminar at
different flow velocities. In the particle Reynolds
number definition, the characteristic velocity is
taken to be that of the entering fluid at a position
corresponding to the centerline of the sphere
(Shahcherghi and Dwyer, 1998). For a centrally
located sphere, this will correspond to the
maximum velocity of the incoming fluid. In case
of laminar pipe flow, the sphere, depending on its
diameter, is exposed to a part of the parabolic
velocity profile of the fluid. This is in contrast to
classical situations where the fluid velocity
398

SI. No.

Variable/parameter

Value

Density

997 kg/m3

Specific heat capacity

4174 J/kg K

Thermal conductivity

0.623 W/m K

Viscosity

7.64 x 10-4 Pa s

Prandtl number

5.12

Engineering Applications of Computational Fluid Mechanics Vol. 4, No. 3 (2010)

3.4

3.3 Estimation of drag coefficient and Nusselt


number

3.4.1 Governing equations and numerical


solution methodology

The drag coefficient is defined according to Eq. 4.


FD

Cd

1
Vmax 2 As
2

The numerical studies were carried using the


commercial ANSYS CFX 11 CFD software,
which is based on the finite volume method. The
governing equations solved are listed below:

(4)

(V ) 0
t

(5)

qw Dp
k (Tw -Tin )

Tw -T
Tw -Tin

Energy equation
(Vh

total

(K T) (12)

In Eq. 12 htotal is defined as the specific total


enthalpy. It is assumed in this work that the flow
involves an incompressible liquid and viscous
dissipation effects are negligible.

(6)

3.4.2 Initial and boundary conditions

Initially, the fluid is considered to be at rest inside


the pipe. The initial temperature of the fluid in the
pipe was specified at 302 K. The sphere is
maintained isothermal at 305 K during the
simulations. At inlet, the fluid flow is assumed to
be fully developed and with the axial velocity
given by Eq. 3. The other two velocity
components are zero:

(7)

(8)

The dimensionless temperature is defined as


follows:

-p V ( V )

htotal Dp

t
Dt

The average Nusselt number ( Nu ) was also


calculated after area averaging the wall heat flux
by integrating over the sphere surface.
Nu =

(11)

k (Tw Tin )

Momentum equation

Nu

qw Dp

(10)

( V )

t (V V )

P is a reference pressure and the corresponding


upstream velocity V is so chosen near the sphere
such that Cp at the front stagnation point is unity.
For local Nusselt number estimation, the local
wall heat fluxes at the surface of the sphere were
estimated using the ANSYS CFX v11 post
processor. The local wall heat flux is related to
the temperature gradient normal to the surface by
the following equation:

T
qw k
R R Rs

Continuity equation

Downloaded by [b-on: Biblioteca do conhecimento online UBI] at 07:33 18 July 2016

FD is the force along the flow direction and is


obtained by integration of the local pressure and
viscous forces along the sphere surface. Here, As
is the projected area of the sphere. The
dimensionless sphere surface pressure is defined
by

P - P
Cp = s
0.5 V2

Details on numerical simulations

Vy = 0 m/s

(13)

Vz = 0 m/s

(14)

At the sphere surface and pipe walls no slip


boundary conditions are imposed. The pipe wall
is assumed to be adiabatic and the heat flux qT at
the wall is specified to be zero,

(9)

As specified by Bird et al. (1960) for flow past an


immersed sphere, the temperature driving force
based on the sphere surface temperature and the
fluid inlet (approach) temperature may be used to
define the heat transfer coefficient and this was
adopted in the present study.

qT 0

(15)

At the exit boundary, the default outflow


boundary condition option in ANSYS CFX 11,
zero-diffusion flux for all flow variables was used.
From the numerical solution of the governing
transport equations with the associated initial and
boundary conditions, information on the velocity,
399

Engineering Applications of Computational Fluid Mechanics Vol. 4, No. 3 (2010)

the direction was 144. Similarly, the number of


grid points on the circumference of the circle in
the y-z plane and in the direction was 40. The
grid points were placed at equidistant intervals on
these circumferences in the and directions.
The grid sizes along the direction had a spacing
of 0.14 Dp and along the direction had a spacing
of 0.504 Dp.
A fluid-fluid interface was generated between the
upstream sub domain and the sphere sub domain.
The interface between the sphere sub domain and
the downstream sub domain was also defined as a
fluid-fluid interface. In the upstream and
downstream cylindrical sub domains, an O-Grid
was generated in the radial direction. A nonuniform grid was generated along the axialdirection with an exponential reduction factor of
1.2 from fluid inlet to the upstream sub domainsphere sub domain interface. Similarly grid was
generated in the downstream section of the pipe
from the sphere sub domain-downstream sub
domain interface to the pipe exit. Structured
meshes involving hexahedral elements could be
generated efficiently for sphere located at the pipe
centre as shown in Figure 1.

pressure and temperature fields were obtained.


These were used to obtain the drag coefficients
and Nusselt number.

Downloaded by [b-on: Biblioteca do conhecimento online UBI] at 07:33 18 July 2016

3.4.3 Details on the mesh size

The 3D structural grid with non-uniform


structured mesh was generated for different
blockage ratios using the commercial ICEM CFX
11. The computational domain is divided into
upstream, sphere and downstream sub domains.
In all the sub domains, finer hexahedral mesh was
generated near the wall surfaces including the
sphere surface, in order to resolve rapid changes
in velocities and temperatures. A circular grid is
generated to stretch the grid spacing from smaller
values corresponding to finer meshes near the
walls to larger values for the coarser ones in the
bulk. An expansion factor of 1.2 was used for this
purpose. In the sphere sub domain, the grid sizes
used near the sphere surface and pipe wall had a
minimum spacing of 0.0125 times the sphere
diameter (Dp) with a bi-geometric expansion in
the radial direction. This radial grid resolution
was invariant along the and directions around
the sphere. The number of grid points on the
circumference of the circle in the x-y plane and in

(a)

x
Fig. 1

(b)

(c)

(d)

Generated mesh for a sphere located axially in a pipe (a) side view, (b) Magnified side view around the
sphere, (c) front view at BR = 0.2, (d) front view at BR = 0.4 and (e) front view at BR = 0.5.

400

(e)

Engineering Applications of Computational Fluid Mechanics Vol. 4, No. 3 (2010)

Downloaded by [b-on: Biblioteca do conhecimento online UBI] at 07:33 18 July 2016

3.4.4

needed to resolve the velocity and temperature


gradients along the radial direction rather than the
other directions.

Solution Methodology

This numerical study has been carried out using


the robust and bounded second-order backward
Euler scheme. The high-resolution scheme option,
which maintains the blend factor as close as
possible to unity without violating the
boundedness principle, was chosen (ANSYS Inc.
(2007). Convergence criterion was fixed at
residuals root mean square value lower than 10-5.
The simulations were executed in the Intel core
2 Duo with specifications of CPU E7200 @
2.53 GHz and 1 GB RAM with windows XP
platform. Transient simulations consumed
typically 70 CPU hours at the highest Reynolds
number and lower blockage ratio investigated in
this work.

4.1.1 Mesh and domain length independence


studies

First the mesh size independence in the sphere


sub domain was established. Then extensive
numerical experimentation was carried out to
identify the optimal upstream and downstream
distances Lu and Ld respectively so that accurate
numerical results could be obtained without
excessive computational effort. While mesh
independence tests were implemented and
checked for many combinations of Rep and BR,
the results for high Reynolds number and high
blockage ratios are presented in Table 2.
First the upstream and downstream domain
lengths, expressed in terms of number of particle
diameters were fixed at 12 and 16 respectively.
The length of the first mesh along the radial
direction and originating from the sphere surface
(h) was expressed as a fraction of particle
diameter. These results are given in Table 2a. In
the second row, while the mesh size in the radial
direction was decreased relative to the first row,
the total number of mesh elements were kept
constant by increasing the size of the meshes
along the and directions. Mesh independence
was not achieved indicating that both smaller
mesh sizes and increased number of meshes were
inevitable. Hence, (h/Dp) was then gradually
further reduced along the radial direction until the
combination shown in the third row led to mesh
size independence. The (h/Dp) values were
constant at 0.14 along the and directions. In
the fourth row of Table 2a, the mesh sizes were
increased along the azimuthal and axial directions
while decreasing the mesh size along the radial
direction in a manner leading to lesser number of

4. RESULTS AND DISCUSSIONS

First, mesh size and time step independence


results are given. This is followed by a validation
of the current CFD simulation results by
comparing them with the correlations available in
literature. The effects of blockage ratio on drag
coefficients and Nusselt number at both low and
high Reynolds numbers are presented. The effects
of blockage ratio on hydrodynamics and heat
transfer are discussed.
4.1

Mesh size and time step independence

Since there are three regions in the pipe domain,


viz. upstream and downstream of the sphere and
the region around the sphere itself, the mesh sizes
were resolved iteratively. Preliminary studies
indicated the upper limit of the upstream and
downstream distances (expressed as multiples of
the sphere diameter) to achieve mesh
independence as well as the relative sensitivity of
the results to the mesh size changes in different
directions. More grid points per unit length were
Table 2a

Table 2b

Mesh independence test in the sphere


sub-domain (time step =0.2 S, Rep=500,
BR=0.5, Lu=12 and Ld=16).

Item
no.

h/Dp

Number of
meshes

Cd

Nu

0.1

64540

0.61

51.93

0.05

64540

0.65

44.32

0.0125

105528

0.62

27.72

0.004

97000

0.62

27.77
401

Mesh independence tests in the upstream


and downstream domains (time step
=0.2S, Rep =500, BR=0.5 and
h/Dp=0.0125).

Item no.

Lu

Ld

Cd

Nu

0.62

27.69

0.62

27.87

12

0.62

27.87

10

15

0.62

27.71

12

14

0.62

27.71

12

16

0.62

27.73

Engineering Applications of Computational Fluid Mechanics Vol. 4, No. 3 (2010)

4.1.2 Time step independence

70

When the sphere was located at the centre of the


pipe, the simulations were able to attain steady
state until Rep=270. The drag coefficient and
Nusselt number were independent of time until
Rep<270 (Bagchi et al., 2001). However since this
study involves different blockage ratios, it was
decided to carry out all the simulations in the
transient mode. For the transient computations,
the time step was chosen to be in the range of 0.1
to 0.5s. This depended on the Peclet number (Pe=
Rep Pr) with larger values indicating more
convection and requiring smaller time steps (Feng
et al., 2000). For the centrally located sphere, a
time step of 0.2s was sufficient to ensure time
step independence at the maximum Reynolds
number of 500. Halving this value did not
produce any significant change in the drag
coefficient and average Nusselt number results.
At lower Reynolds number, where steady state
conditions could be attained, a time step of 0.5s
proved to be sufficient. However, all transient
simulations were carried out at 0.2s. Local
Nusselt numbers at different angular positions
were plotted for time steps of 0.1, 0.2 and 1 s for
a total time of 250 sec. There is no significant
difference between the time steps of 0.1s and 0.2s
as shown in Figure 2. Therefore the results
presented herein are believed to be free from
upstream-downstream distances, mesh sizes and
time step effects.

60

Nu (-)

50
40
30
20
10
0
0

30

60

90

120

150

180

Angle (degree)
Time Step = 0.2

Downloaded by [b-on: Biblioteca do conhecimento online UBI] at 07:33 18 July 2016

Fig. 2

Time Step = 0.1

Time Step = 1

Effect of different time steps on the local


Nusselt numbers estimated at the end of 250 s.
(Rep=500, BR=0.5).

meshes than when compared to the third row. It


may be seen that the mesh arrangement
corresponding to row 3 of Table 2a is mesh size
independent.
When the blockage ratio was very low resulting in
higher gap width between the sphere and the wall,
the same number of mesh elements used at higher
blockage ratio did not lead to mesh size
independence. More mesh elements had to be
used from the sphere to the wall while
maintaining the same (h/Dp) value at the first
grid point from the sphere. As the blockage ratio
reduced by a factor of 2.5 (from 0.5 to 0.2), the
number of mesh elements in the gap between the
sphere and the wall had to be increased by a
factor of two to ensure mesh size independence.
Next the upstream and downstream distances
were adjusted to obtain mesh size independent
results (Table 2b). From these studies, it was
found that in order to reach mesh independence,
Lu needs to be a minimum of 10 and Ld needs to
be a minimum of 12. The values actually used in
the present work were 12 and 14 for Lu and Ld
respectively.
Bharti et al. (2007) suggested using at least 10
pipe diameters upstream and 40 pipe diameters
downstream in their simulations involving heat
transfer from a confined sphere to power law
fluids. We compared our optimized grid results
with this suggestion and found negligible
difference (<1%) in the Nusselt number and drag
coefficients. We also tested on values such as
local Nusselt number, angle of separation of
boundary layer and found negligible differences.
Hence the optimized grid size scheme was
adopted in all further simulations.

4.2

Validation of CFD simulations

To simulate the classical situation involving fluid


flow at uniform velocity past an unconfined
sphere, simulations were carried out in the
following manner. The blockage ratio of the
sphere was chosen at a very small value of 0.02.
Free slip conditions were given at the pipe wall
and no slip boundary conditions were given at the
sphere surface. Uniform velocity profile was
given at the pipe inlet. The sphere surface
temperature is maintained at 305 K and incoming
fluid temperature is 302 K.
The drag coefficient values obtained under these
conditions are within 2% of the values given by
the standard Cd curve for a sphere (Clift et al.,
1978) as shown in Figure 7. For these conditions,
average Nusselt numbers were also computed and
compared with the correlation due to Kramers
(1946) as given below:
hDp
k

= Nuo = 0.97+0.68Rep 0.5 Pr 3

1< Rep <2000

(16)
402

Engineering Applications of Computational Fluid Mechanics Vol. 4, No. 3 (2010)

the velocity profile of the fluid entering the pipe


is parabolic rather than uniform.
When inspecting the drag coefficients at initial
times, high values were obtained. After the initial
transients, the drag coefficient tends to decrease
and approach the classical solution.

1000
Re p = 0.1

100

Cd (-)

Re p = 2

10

Re p = 10

4.3.2 Effect of blockage ratio

At very low Rep and BR value of 0.1, the steady


state drag coefficient value exceeds the
predictions of Clift et al. (1978) significantly by
about 23% as given in Table 3. However for this
blockage ratio, as Rep increased, the steady state
solution nearly merged with the classical case
result. Upon increasing the blockage ratio at a
given Reynolds number, the deviation from the
classical solution increased as shown in Table 3.
The deviations are found to be quite severe at
higher blockages and very low particle Reynolds
numbers.
The percentage deviation used in Table 3 and the
following ones is defined as follows:

0.1
10

100

1000

10000

Time (sec)
--- Clift et al., 1978

Downloaded by [b-on: Biblioteca do conhecimento online UBI] at 07:33 18 July 2016

Fig. 3

BR = 0.02

Variation of drag coefficient with time at


different Reynolds numbers (BR=0.02, Clift et
al., 1978).

The Nusselt number values from CFD simulations


were within 3% of those corresponding to
Kramers (1946) correlation as shown in Figure 16.
4.3 Drag coefficients at low particle Reynolds
numbers (10)

Percentage deviation

4.3.1 Attainment of steady state

CFD Prediction Classical value


x 100
Classical value

(17)

All the simulations were carried out in the


transient mode. At lower Reynolds numbers, it
took more time for the drag coefficient to reach
steady state values. This is because of molecular
flux dominating over the convective flux of
momentum. The CFD simulation results for
transient drag coefficient at BR = 0.02 are
compared with the classical solutions of Clift et al.
(1978) in Figure 3. The classical solutions
referred to in this work correspond to a uniform
flow over an unconfined sphere. In contrast to this
situation, results are presented in this work for a
sphere with various degrees of confinement in a
pipe depending on the spheres diameter. Further,

4.3.3 Comparison with literature correlations


for the effect of blockage ratio on drag
coefficient in the low Reynolds number limit

The actual drag coefficients computed in confined


flow were compared to those predicted from
Stokes law applicable in the creeping flow regime
(Cds = 24/Rep). The effect of blockage ratio on G
( Cd ) at a Reynolds number of 1 is shown in
Cds

Figure 4. The different correlations given in


literature and the range of applicability of these
are summarized below:

Table 3 Effect of blockage ratio on deviation of steady state drag coefficient from classical solution at different Rep
(The percentage deviation from the classical solution is given in parenthesis).
Rep

Clift et al., 1978

0.1

244.25

27.15

10

4.26

100
350
500

1.087
0.63
0.555

BR = 0.1
301.06
(23.3)
30.59
(12.6)
4.33
(1.6)
1.081
(-0.55)
0.63
(0)
0.56
(1.81)

BR = 0.2
402.69
(64.9)
39.57
(45.7)
4.61
(8.2)
1.06
(-2.48)
0.62
(-1.58)
0.558
(0.54)
403

BR = 0.4
787.41
(222.4)
75.99
(179.8)
8.03
(88.5)
1.215
(11.77)
0.63
(0)
0.554
(-0.18)

BR = 0.5
1214.6
(397.3)
122.33
(350.5)
12.31
(189)
1.556
(43.14)
0.722
(14.60)
0.620
(11.71)

Engineering Applications of Computational Fluid Mechanics Vol. 4, No. 3 (2010)

14
12

G (-)

10
8
6
4
2
0
0.00

0.10

0.20

0.30

0.40

0.50

0.60

0.70

BR (-)
Wakiya, 1957
Present correlation at Rep = 100
Present Simulation

Downloaded by [b-on: Biblioteca do conhecimento online UBI] at 07:33 18 July 2016

Fig. 4

G=

Haberman and Sayre, 1958


Wham et al., 1996

Present correlation

Wham et al., at Rep = 100

Present correlation at Rep = 100

Present simulation at Rep = 100

Drag correction for a sphere in creeping flow.

Fig. 5

occurring at higher G values where the deviations


were much greater with respect to Stokes solution.
The R2 for the fit was 0.97.

Wakiya (1957)
2
1- BR 2
3

(18)

1- 2.105BR + 2.09 BR 3 -1.11BR 5

4.4 Drag coefficients at higher particle


Reynolds numbers

This correlation is applicable for Rep 1

Once the particle Reynolds number exceeds 300,


Cd values did not attain steady state (Lee, 2000).
The transients in the drag coefficient are indicated
in Figure 5 for a particle Reynolds number of 500
and different blockage ratios. With increasing BR,
the time period of oscillation increased and the
amplitudes declined, indicating the tendency
towards stabilization in the flow fields around the
particle. The deviation of time averaged drag
coefficient from the classical results of Clift et al.
(1978) is shown in parenthesis in Table 3. Further,
at higher particle Reynolds numbers the flows
around the sphere became asymmetric. It has been
well established in literature (Magarvey and
Bishop, 1961) that unsteadiness and asymmetry is
created by supercritical regular and Hopf
bifurcations. Sakamoto and Haniu (1990) have
shown that with increasing Reynolds numbers,
the vortex shedding location changes irregularly
along the azimuthal direction leading to threedimensional flow at Rep = 420.

Haberman and Sayre (1958)

2
1- BR 2 - 0.20217 BR 5
3
G=
1- 2.150BR + 2.0865BR 3 -1.7068BR 5 + 0.72603BR 6

(19)

This correlation is applicable for Rep 1.

Wham et al. (1996)

G = (1+0.05488 Rep

(1.2905-0.07529ln Rep )

2
1- BR 2 - 0.20217 BR 5
3
1- K BR + 2.0865BR 3 -1.7068BR 5 + 0.72603BR 6

(20)
where
(-0.07836 Rep )
K =1.358 + 0.7486 e

(21)

This correlation is applicable for Rep 100.


4.3.4 Correlation developed in the present
work (Rep500) with 0.1 BR 0.5
G = (1+ 0.033 Rep0.888 )

2
1- BR 2 - 0.20217 BR 5
3
1- K BR + 2.0865BR 3 -1.7068BR 5 + 0.72603BR 6

4.5 Explanation for the effect of blockage


ratio on drag coefficient

(22)

The effect of blockage ratio on the terms


contributing to the drag coefficient was analyzed.
The contributing terms to the drag due to
pressure (Cdp) and friction (Cdf) are plotted with
respect to Rep for different blockage ratios in the
bar diagrams shown in Figure 6. The overall
effect of blockage ratio on drag coefficient is
shown in Figure 7. At lower Rep the effect of

where
(-0.009435 Rep )

K =1.3 1+ 0.61224 e

Effect of blockage ratio on the transients in


drag coefficients at higher particle Reynolds
number of 500.

(23)

This correlation fitted the data at different


blockage ratios and Reynolds numbers within a
maximum of 15%, the higher % deviations
404

Engineering Applications of Computational Fluid Mechanics Vol. 4, No. 3 (2010)

1000.00

1000.00

100.00

100.00
10.00
Cdf

Cdp

10.00
1.00

1.00

0.10

0.10

0.01

0.01
0.1

10

40

100

270

350

500

0.1

10

40

Re p
BR = 0.02

BR = 0.2

BR = 0.4

BR = 0.02

BR = 0.5

Fig. 6b

270

350

500

BR = 0.2

BR = 0.4

BR = 0.5

Effect of blockage ratio on viscous drag


variation with Rep.

1.20

10000

(a) Rep = 0.1

1.00
V/Vmax (-)

1000
C d (-)

100
10
1

0.80
0.60
0.40
0.20

0.1
0.01

0.1

10

100

0.00
0.00

1000

0.50

1.00

1.50

Rep (-)
BR = 0.02

BR=0.2

BR = 0.4

2.00

2.50

3.00

3.50

4.00

R/R s (-)
BR = 0.5

Clift et al. (1978)

BR = 0.2

Drag coefficients at different blockage


ratios compared with the classical solution
given by Clift et al. (1978) for a sphere.

BR = 0.4

BR = 0.5

1.20

(b) Rep = 500

1.00
V/Vmax (-)

Downloaded by [b-on: Biblioteca do conhecimento online UBI] at 07:33 18 July 2016

Fig. 6a Effect of blockage ratio on form drag variation


with Rep.

Fig. 7

100
Re p

blockage is felt more on Cdp than on Cdf. The Cdf


values were found to merge faster, i.e. at a lower
Rep, with the classical results (BR = 0.02) than Cdp
values.
At low Rep, both Cdf and Cdp increase with
increasing blockage. The fluid undergoes
convective acceleration in the region between the
sphere surface and the pipe wall with the
accompanying loss in pressure. The velocity
gradient also increases at the sphere surface when
the velocity gradient at the pipe wall boundary
increases at increasing blockage ratios (Figure 8a).
Due to this, there is an increase in the viscous
force acting at the sphere wall. These factors
cause the delay in the boundary layer separation
and also a higher drag coefficient.
The adverse pressure gradient required for the
boundary layer separation, where the x-directional
velocity gradient becomes zero, occurs at an angle
further downstream on the sphere surface. The
momentum of the fluid near the surface is not
sufficiently high to overcome the increase in

0.80
0.60
0.40
0.20
0.00
0.00

0.50

1.00

1.50

2.00

2.50

3.00

3.50

4.00

R/R s (-)
BR = 0.2

Fig. 8

BR = 0.4

BR = 0.5

Effect of blockage ratio on the velocity


profiles taken at the north pole of the
sphere.

pressure. When the velocity gradient in y


direction at the sphere surface becomes zero, the
flow is said to have reached a separation point
(Holman, 1997; Dhole et al., 2006; Antar and EIShaarawi, 2009). Reverse flow begins to occur
near the sphere surface beyond this point.
At separation point the following equation holds.
V x
R

405

R Rs

=0

(24)

Engineering Applications of Computational Fluid Mechanics Vol. 4, No. 3 (2010)

The separation angles, estimated using the CFX


post processor, were compared with literature
correlation of Kalra and Uhlherr (1971) as shown
in Figure 9. The estimated values at a BR of 0.02
were within 2% of the predictions from the
correlation of Kalra and Uhlherr (1971) that was
applicable for 30 < Rep < 750.
s - 83 262 Rep -0.372

increasing blockage ratio i.e. the separation gets


delayed due to the stabilizing influence of the
pipe wall. At a given BR, s decreases with
increasing Rep. Due to possible asymmetry in the
flow field especially at higher Rep and lower BR,
the s was estimated at both the top and bottom
hemispheres. Only the case involving the
blockage ratio of 0.2 and particle Reynolds
number of 500 indicated a significant difference
in the separation angles. The angle of separation
when estimated at the bottom hemisphere is 121o
in contrast to that observed at the top hemisphere
(111o). The following correlation is developed for
the boundary layer separation angle ( s)
measured along the upper hemisphere, which fits
the CFD data to within 4%.

(25)

s= (83 + 262 Rep-0.37 ) (1+1.224 BR1.35 Rep-0.24 )

s
(deg)

Fig. 9

(26)
The delay in the boundary layer separation at
higher Rep shrinks the vortex region behind the
sphere. However, the narrowing gap width
between the pipe wall and sphere surface at
higher blockage ratios and the resulting
acceleration of the fluid in the gap also leads to a

Comparison of CFD simulation predicted


separation angles with the correlation of Kalra
et al., 1971 at different Rep.

1.00
0.80
0.60

1.00
0.40

0.80
Cp (-)

0.60
0.40
0.20
Cp (-)

Downloaded by [b-on: Biblioteca do conhecimento online UBI] at 07:33 18 July 2016

The separation angles at different particle


Reynolds numbers and blockage ratios are shown
in Table 4. At a fixed value of particle Reynolds
number, the angle of separation increases with

0.20
0.00
-0.20

0.00
-0.20 0

20

40

60

80

100

120

140

160

180

20

40

60

80

100

120

140

160

180

-0.40

-0.40

-0.60

-0.60

-0.80

-0.80

Angle (degree)

-1.00
-1.20
Angle (degree)
BR = 0.2

BR = 0.4

BR = 0.2 (upper)

BR = 0.4 (upper)

BR = 0.5 (upper)

BR = 0.2 (lower)

BR = 0.4 (lower)

BR = 0.5 (lower)

Fig. 10b Variation of Cp with angle at different


blockage ratios at Rep = 500 and calculated
along the upper and lower hemispheres.

BR = 0.5

Fig. 10a Variation of Cp with angle at different


blockage ratios at Rep = 100.

Table 4 Effect of blockage ratio and particle Reynolds number on angle of separation calculated along the upper
hemisphere. NS indicates no separation. (The % deviation from the classical solution is given in
parenthesis.)
Rep

Kalra et al., 1971

40

151.42

100

132.24

200

121.50

350

114.64

500

110.96

BR = 0.2
157
(3.68)
142
(7.38)
126
(3.70)
121
(5.54)
111
(0.04)

406

BR = 0.4

BR = 0.5

NS

NS

147.7
(11.69)
133
(9.46)
124
(8.16)
119.5
(7.69)

152.5
(15.32)
136.5
(12.34)
129.5
(12.96)
124
(11.75)

Downloaded by [b-on: Biblioteca do conhecimento online UBI] at 07:33 18 July 2016

Engineering Applications of Computational Fluid Mechanics Vol. 4, No. 3 (2010)

minimum was not observed at lower Reynolds


numbers. At the maximum Rep of 500
investigated in this work, it may be seen from
Figure 11 that the flow pattern has changed
considerably. Here the flow paths were traced
using the CFX post processor and represent the
path taken by a particle of zero mass driven by the
vector field (Vinoj et al., 2009). At the lower BR
of 0.2 there is considerable mixing especially in
the upper wake region. The fluid in the bottom
portion of the wake reaches the upper wake
region. The flow from the bottom to the top is
driven by the pressure differences between the
bulk fluid region and the region immediately
behind the sphere (Figure 12). However, the
pressure near the sphere surface is lower at the
bottom hemisphere than at the top (Figure 10b).
The hydrodynamic boundary layer separation
occurs earlier at the top hemisphere while it is
relatively delayed at the bottom hemisphere. The
separation angle s is 111o in the upper
hemisphere while it is 121o in the bottom
hemisphere. The pressure recovery in the upper
hemisphere which occurred until about 120o is
arrested and the pressure again begins to decline
due to intrusion from the fluid in the upper wake
region (Figure 10b). As shown in Figure 10b,
only the case involving the lower BR of 0.2 and
Rep of 500 exhibits asymmetry leading to

greater difference in pressure between the front


and rear regions of the sphere (Figure 10a). At
moderate Rep of 100, it may be seen that while the
boundary layer separation shifts to higher angles
with increase in blockage, the form drag still
increases (Figure 6a). Here, the effect of the pipe
wall on the velocity profile and the consequent
pressure loss outweighs the gain due to the effect
of the shrinking of the wake region behind the
sphere. The shrinking of the wake region behind
the sphere is shown in Figure 11.
The velocity gradients are not significantly
different even at higher Rep (Figure 8b). As
mentioned in section 3.2., laminar velocity profile
was specified at the domain inlet and Vmax is the
maximum velocity at inlet pipe centre. V is the
velocity of the fluid taken at the gap between the
pipe wall and north pole location on the sphere
surface. The fluid undergoes convective
acceleration in this region leading to even greater
velocity in this narrow region when compared to
Vmax. Hence this leads to V/Vmax being greater than
1.
At higher Reynolds numbers, the blockage ratio
has a relatively smaller effect on Cd (Figure 7).
The drag coefficient showed a minimum at an
intermediate BR of 0.4 at Rep of 500. This was
previously observed in the time averaged values
of drag coefficient in Figure 5. However, this

Rep = 500

Rep = 100

BR = 0.2

BR = 0.4

Fig. 11

Velocity labeled particle path lines in the fluid domain at Rep= 100 and Rep=500. Also shown on the
sphere are the local stresses at BR = 0.2 and BR = 0.4.

407

Engineering Applications of Computational Fluid Mechanics Vol. 4, No. 3 (2010)

BR = 0.2

BR = 0.4

Fig. 12

Pressure distribution in the fluid domain, shown at Rep = 500 at BR = 0.2 and BR = 0.4.

4.6 Nusselt number at low Reynolds number


( 10)

differences in the trends (boundary layer


separation, flow patterns, Cp) between the top and
bottom hemispheres. The Cp trends for the bottom
hemisphere were quite similar to those obtained
for the upper hemisphere at BR of 0.4 and 0.5.
A higher BR of 0.4 stabilizes the wake region
even more behind the sphere and it is relatively
more axisymmetric when compared to the lower
BR case (Figure 11). There is no mixing in the
upper wake region and the usual pressure
recovery occurs (Figure 10b). When blockage
ratio increased to 0.5, the wall effect became
more significant.
In Figure 11, the fluid path lines and the sum of
local shear stress and pressure on the sphere are
given. The dimensionless local stresses (Slocal)
acting on the sphere surface in the direction of
fluid flow is defined as shown below:

l
a
c
o

Sl

Px xx yx zx
0.5 Vmax 2

4.6.1 Attainment of steady state

At low Reynolds number, the Nusselt number


took longer to reach steady state when compared
to the drag coefficient. The thermal diffusivity is
less when compared with the momentum
diffusivity for the water system (Pr = 5.12). CFD
predicted transient Nusselt numbers at BR of 0.02
are compared with the steady classical solution of
Kramers (1946) in Figure 13.
4.6.2 Effect of blockage ratio

The effect of blockage ratio on average Nusselt


number is weaker than that on the drag coefficient.
This may be seen by comparing the results in
tables 3 and 5. The average Nusselt number
deviation from the classical solution was not as
high as the deviation noticed in the case of the
drag coefficient even at a low Reynolds number
of 0.1 and high blockage ratios.

(27)

At Rep = 100, the flow is axisymmetric and the


wake region formed due to flow separation is
evident. The sum of the local shear and pressure
are the highest at the front stagnation point. They
decline once the flow separation has occurred.
The wake region behind the sphere has shrunk at
a higher BR of 0.4 and Rep = 100 when compared
to BR = 0.2. It is because of the higher
(normalized) local stress and delay in the
boundary layer separation at BR = 0.4. At Rep =
500, the wake structure at BR = 0.4 is not as
irregular and asymmetric as that encountered at
BR = 0.2.

10
9
8
7
6
Nu (-)

Downloaded by [b-on: Biblioteca do conhecimento online UBI] at 07:33 18 July 2016

Rep = 500

Pe = 51.2

5
4

Pe = 10.24

Pe = 0.51

2
1
0
0

1000

2000

3000

4000

5000

6000

Time (sec)
---- Kramers, 1946

BR = 0.02

Fig. 13 Variation of average Nusselt number with


time at different Peclet numbers (BR = 0.02).
408

Engineering Applications of Computational Fluid Mechanics Vol. 4, No. 3 (2010)

Rep

Kramers, 1946

0.1

2.04

2.88

10

5.37

100

13.39

350

23.59

500

27.87

BR=0.1

BR = 0.2

BR = 0.4

BR = 0.5

2.08
(1.96)
2.98
(3.47)
5.41
(0.74)
12.97
(-3.13)
23.97
(1.61)
28.89
(3.6)

2.08
(1.96)
3.03
(5.21)
5.48
(2.05)
12.94
(-3.3)
23.94
(1.48)
28.79
(3.3)

2.72
(33.33)
3.30
(14.58)
5.99
(11.54)
13.07
(-2.39)
23.42
(-0.72)
27.96
(0.32)

2.70
(32.35)
3.37
(17.01)
6.39
(18.99)
13.68
(2.17)
23.54
(-0.21)
27.68
(-0.68)

towards stabilization in the thermal fields around


the particle. The deviation of time averaged
Nusselt number from the classical results of
Kramers (1946) is shown in parenthesis in Table
5. The percentage deviations are not significant at
higher particle Reynolds numbers and lower
blockage ratios.

30

Nu (-)

29.5
29

BR = 0.2

28.5
28

BR = 0.4
BR = 0.5

27.5
150

160

170

180

190

200

210

4.8 Explanation for the effect of blockage


ratio on Nusselt number

Time (sec)

Fig. 14 Effect of blockage ratio on the transients


shown by average Nusselt numbers at a
higher particle Reynolds number of 500.

As the thermal boundary layer around the sphere


is thin, the effect of the pipe wall on Nusselt
number is smaller when compared to the drag
coefficient. The average Nusselt number is close
to 2 for a sphere at low Reynolds number (i.e. in
the creeping flow region). But the average Nusselt
number at higher blockage ratios shows
significant enhancement from this limiting value
(Table 5). The dimensionless temperature
variation is affected by the wall at the high
blockage ratio as shown in Figure 15a and there is
an increase in the gradient.
At low Reynolds number, the molecular flux
dominates the convective thermal flux. This leads
to the thermal boundary layer to be much thicker
at particle Reynolds number of 0.1. As the gap
width between the sphere surface and the pipe
wall decreases with increasing blockage ratio, and
the thermal boundary layer is quite thick in this
region at the low particle Reynolds number, the
dimensionless temperature never reaches unity at
Rep = 0.1 and BR = 0.5 as shown in Figure 15a.
But when the particle Reynolds number is greater
than or equal to 10, the thickness of the thermal
boundary layer is considerably reduced and is
much thinner than the gap width. Hence the
dimensionless temperature can reach the unity
value in the higher Reynolds number cases.

0.90
0.80
0.70
0.60
0.50
(-)

Downloaded by [b-on: Biblioteca do conhecimento online UBI] at 07:33 18 July 2016

Table 5 Effect of blockage ratio on deviation of steady state Nusselt number from classical solution at different Rep
(The % deviation from the classical solution is given in parenthesis).

0.40
0.30
0.20
0.10
0.00
0.00

0.50

1.00

1.50

2.00

2.50

3.00

3.50

4.00

R/R s (-)
BR = 0.2

BR = 0.4

BR = 0.5

Fig. 15a Effect of blockage ratio on the thermal


profile taken at the north pole of the sphere
at different blockage ratios (Rep = 0.1).

4.7 Nusselt number at higher particle


Reynolds numbers (Rep<500)

Figure 14 shows the oscillations in area averaged


Nusselt numbers as a function of time at Rep =
500 and different blockage ratios. The period of
oscillation is shorter at lower blockage ratios the
amplitudes are higher. With increasing BR, the
time period of oscillations increased while the
amplitudes declined, indicating the tendency
409

1.20

30

1.00

25

0.80

20
Nu (-)

(-)

Engineering Applications of Computational Fluid Mechanics Vol. 4, No. 3 (2010)

0.60

15

0.40

10

0.20

0.00
0.00

0.10

0.20

0.30

0.40

0.50

0.60

0.70

0.80

0.90

1.00

100

200

R/R s (-)

Downloaded by [b-on: Biblioteca do conhecimento online UBI] at 07:33 18 July 2016

BR = 0.2

BR = 0.4

BR = 0.5

BR = 0.02

Fig. 15b Effect of blockage ratio on the thermal profile


taken at the north pole of the sphere at
different blockage ratios (Rep = 500).
Rep = 100

500

BR = 0.2

BR = 0.4

BR = 0.5

Kramers (1946)

Rep = 500

0.77

0.89

0.8

Nu

Nu
0.88

0.88

0.55

BR = 0.4

400

Fig. 16 Effect of blockage ratio on average Nusselt


number with Rep.

0.88
0.55

BR = 0.2

300
Re p (-)

0.33

Nu

Nu

Fig. 17 Dimensionless Temperature () distribution in the fluid domain at (a) Rep= 100 and (b) Rep=500. Also
shown are the Nusselt number distributions at the sphere surface.

There is about 19% enhancement in the value of


the Nusselt number for a sphere at particle
Reynolds number of 10 when the blockage ratio is
0.5. However the influence of blockage ratio
diminishes with increase in the particle Reynolds
number (Table 5). At higher Rep, the variation in
dimensionless temperature from the wall of the
sphere is confined to an even smaller region
(Figure 15b). Hence, the pipe wall does not have
much influence on the Nusselt number. Figure 16
summaries the effect of BR on the average
Nusselt number for Rep up to 500.

The Nusselt number correlation that accounts for


the effect of blockage at different particle
Reynolds numbers is shown below. This
correlation fitted the data (Rep 500, 0.02 BR
0.5) within 5%.
0.919 BR1.47
Nu = Nuo 1
0.67

Rep

Nuo refers to Kramers solution (Eq. 16).

410

(28)

Engineering Applications of Computational Fluid Mechanics Vol. 4, No. 3 (2010)

This trend persists up to flow separation point,


and then local Nu started to increase for a particle
Reynolds number of 100 and blockage ratio of 0.2
(Figure 18). This is due to recirculation in the
wake bringing in colder fluid and increasing the
temperature gradients as shown in Figure 17. This
results in a local maximum in the local Nusselt
number at the rear stagnation point as shown in
Figure 18. The local Nusselt number attains a
minimum at the separation angle and since s is
lower at higher Rep, the minimum Nu occurs
earlier for Rep=500. Further, the local Nusselt
number trends are different at higher particle
Reynolds numbers.

Figure 17 shows the dimensionless temperature


distribution in the fluid flowing past the sphere
and local Nusselt number distribution along the
sphere surface at two different particle Reynolds
numbers and blockage ratios.
The variations of local Nusselt number with two
different Rep are illustrated for two different
blockage ratios, in Figure 18 and 19. The local
Nusselt number has a maximum at the front
stagnation point and decreases along the stream
wise direction due to increase in thermal
boundary layer thickness (Bagchi et al., 2001).
80

70

70

60

60

50
Nu (-)

50

Nu (-)

Downloaded by [b-on: Biblioteca do conhecimento online UBI] at 07:33 18 July 2016

4.9 Effect of blockage ratio on temperature


pattern and local Nusselt number

40
30

40
30

20

20

10

10

0
0

20

40

60

80

100

120

140

160

180

20

40

60

80

100

120

140

160

180

Angle (degree)

Angle (degree)
Along lower hemisphere Rep = 500

Along upper hemisphere Rep = 100

Along upper hemisphere Rep = 100

Alone upper hemisphere Rep = 500

Along lower hemisphere Rep = 100

Along upper hemisphere Rep = 500

Along lower hemisphere Rep = 500

Along lower hemisphere Rep = 100

Fig. 18 Variation of local Nusselt number around the


sphere at two different Reynolds numbers of
100 and 500 at BR = 0.2.

Fig. 19 Variation of local Nusselt number around the


sphere at two different Reynolds numbers of
100 and 500 at BR = 0.4.
(b) Rep = 500

(a) Rep = 100

BR = 0.2

BR = 0.4

Fig. 20 Dimensionless convective heat fluxes in the fluid domain at (a) Rep= 100 and (b) Rep=500. Also shown are
wall heat fluxes at the sphere surface at two different blockage ratios of 0.2 and 0.4.

411

Engineering Applications of Computational Fluid Mechanics Vol. 4, No. 3 (2010)

ratio indicating flow stabilization. The form drag


was affected more than the frictional drag by the
wall of the pipe. Higher blockage ratios delayed
the onset of adverse pressure gradient necessary
for boundary layer separation and consequently
the boundary layer separation was delayed.
The effect of blockage ratio on Nusselt number is
weaker than on the drag coefficient. It is because
the thermal boundary layer thickness is smaller
than that of the momentum boundary layer for the
water system investigated. The local Nusselt
number distribution along the upper and lower
hemispheres of the sphere was significantly
different at blockage ratio of 0.2 and Reynolds
number of 500. However, the differences declined
at higher blockage ratios. Unusual local patterns
in flows were observed at low blockage ratio of
0.2 and particle Reynolds number of 500. There
was considerable mixing in the wake region and
the flow and thermal fields behind the sphere
became considerably asymmetric. The cause for
these phenomena needs further study.
Correlations were proposed to predict drag
coefficient, boundary layer separation angle and
average Nusselt numbers as a function of
blockage ratio and particle Reynolds numbers.
This study indicates that there may be
considerable error in using classical correlations
that are applicable for sphere located in an
unbounded flow with a flat velocity profile to
situations where the flow is bounded and has a
velocity profile. However, for those cases, when
particle Reynolds numbers increased and
blockage ratios decreased, the errors decreased.
For Rep > 300 and small blockage ratios (<0.2)
the errors incurred from using classical solutions
for predicting CD and Nu were <2%. Transient
three-dimensional simulations are recommended
for Rep > 270 and low blockage ratios.

In Figure 20, the velocity streamlines are labeled


in terms of the convective heat flux which is
defined as follows:

Downloaded by [b-on: Biblioteca do conhecimento online UBI] at 07:33 18 July 2016

CV (T - Tin )
qw

(29)

It may be seen that the convective heat fluxes are


higher in the boundary layer than in the wake
region as the velocities are higher in the former.
The local Nusselt numbers and temperature
distribution around the sphere are axisymmetric
for Reynolds of 100 as shown in Figure 17. This
was applicable for both blockage ratios (0.2 and
0.4) illustrated in these figures. However, when
Rep was increased to 500, the flow became
asymmetric at the lower blockage ratio of 0.2.
This is reflected in the local Nusselt numbers
(Figure 18) as well as in the convection current
patterns depicted in Figure 20. There is a
significant variation in the local Nusselt numbers
traced along the upper hemisphere when
compared to those traced along the lower
hemisphere.
After attaining a minimum at the separation point,
the Nusselt number values began to increase and
attained a local maximum near an angle of 140o.
This may be attributed to the mixing in the upper
wake at higher Rep leading to colder fluid brought
into this wake region (Figure 20). Hence there is
an appreciable increase in the local Nusselt
number behind the upper hemisphere of the
sphere when compared to those obtained behind
the lower hemisphere. The boundary layer
separation occurs at a later angle along the lower
hemisphere and the local Nusselt number reduced
until this point was reached. When blockage
ratios increased, the stabilization effect of the pipe
wall led to axisymmetry even at the higher Rep
(Figures 17-20). At lower Rep and higher
blockage ratios, the Nusselt numbers continued to
decrease as the boundary layer separation was
considerably delayed.

NOMENCLATURE

As

5. SUMMARY AND CONCLUSIONS

BR

Three dimensional and transient CFD simulations


were carried out to investigate how the
hydrodynamics and heat transfer phenomena were
affected by the blockage due to a sphere centrally
located in a pipe. At higher Reynolds number, the
drag and Nusselt number values did not attain
steady state especially at lower blockage ratios
and particle Reynolds numbers exceeding 300.
The period of oscillation increased and the
amplitudes declined with increasing blockage

C
Cd
Cdp
Cdf
Cds
Cp
Dp
DT
412

Projected Area of the sphere (/4 Dp2)


(m2)
Diameter of particle (Dp) to diameter
of pipe (DT) (-)
Specific heat capacity (J/Kg K)
Drag coefficient (-)
Pressure coefficient (-)
Friction coefficient in x - direction (-)
Stokes drag coefficient (24/Rep) (-)
Local pressure coefficient (-)
Diameter of the particle (m)
Pipe diameter (m)

Engineering Applications of Computational Fluid Mechanics Vol. 4, No. 3 (2010)

FD

h
htotal
k
Ld
Lu
Nu
Downloaded by [b-on: Biblioteca do conhecimento online UBI] at 07:33 18 July 2016

Nu
P
Pe
Pr
Ps
Px
P
qT
qw

qw
R
Rmax
Rs
Rep
ReT
Slocal
T
Tin
Tw
V
Vch
Vavg
Vmax
Vx Vy Vz

Force along the flow direction (N)


Ratio of actual drag coefficient to
Stokes drag coefficient (-)
Average heat transfer coefficient
(W/m2 K)
Specific total enthalpy (KJ/Kg)
Thermal conductivity (W/m K)
Downstream distance expressed in
terms of sphere diameters (-)
Upstream distance in terms of sphere
diameters (-)
Local Nusselt Number (-)
Average Nusselt Number (-)

REFERENCES

1. ANSYS
Inc.
(2007).
ANSYS
CFX
documentation. Canonsburg: ANSYS Inc.
2. Antar MA, El-Shaarawi MAI (2009). The
entropy generation for a rotating sphere under
uniform heat flux boundary condition in
forced-convection flow. International Journal
of Numerical Methods for Heat and Fluid
Flow 19(3/4):396410.
3. Awuah GB, Ramaswamy HS, Simpson BK
(1995). Comparsion of two methods for
evaluating fluid-to-surface heat transfer
coefficients. Food Research International
28(3):261271.
4. Bagchi P, Ha MY, Balachandar S (2001).
Direct Numerical Simulation of flow and heat
transfer from a sphere in a uniform cross flow.
J. of Fluids Eng. 123(2):347358.
5. Balasubramaniam VM, Sastry SK (1996).
Liquid to particle heat transfer in continuous
tube flows: comparison between experimental
techniques. Int. J. Food Sci. Tech. 31(2):177
187.
6. Bharti RP, Chhabra RP, Eswaran V (2007).
Effect of blockage on heat transfer from a
cylinder to power law liquids. Chemical
Engineering Science 62(17):47294741.
7. Bird RB, Stewart WE, Lightfoot EN (1960).
Transport Phenomena. John Wiley and Sons,
New York p. 424.
8. Chen Su-Lin, Yeh An-I, Wu JS (1997).
Effects of particle radius, fluid viscosity and
relative velocity on the surface heat transfer
coefficient of spherical particles at low
Reynolds numbers. J. Food Engg.31:473484.
9. Clift R, Grace JR, Weber WR (1978).
Bubbles, Drops and Particles, Academic
Press: New York.
10. Dhole SD, Chhabra RP, Eswaran VA (2006).
A numerical study on the forced convection
heat transfer from an isothermal and isoflux
sphere in the steady symmetric flow regime.
Int. J. Heat and Mass Transf. 49(56):984
994.
11. Fayon OH, Happel J (1960). Effect of a
cylindrical boundary on a fixed rigid sphere
in a moving viscous fluid. AIChE J. 6(1):55
58.
12. Feng Zhi-Gang, Michaelides EE (2000). A
numerical study on the transient heat transfer

Pressure (N/m2)
Peclet Number (-)
Prandtl number (-)
Pressure at the sphere surface (N/m2)
Pressure acting at X-direction (N/m2)
Reference pressure (N/m2)
Heat flux at the pipe wall (W/m2)
Local wall heat flux (W/m2)
Average wall heat flux (W/m2)
Radial position (m)
Radius of the pipe (m)
Radius of the sphere (m)
Particle Reynolds number (-)
Tube Reynolds number (-)
Dimensionless local shear stresses (-)
Temperature (K)
Inlet Temperature (K)
Sphere surface Temperature (K)
Velocity (m/s)
Characteristic velocity (m/s)
Average velocity (m/s)
Maximum velocity (m/s)
x-, y- and z- components velocity
(m/s)

Greek symbols
Kronecker delta

Del operator

Size of the first mesh in the radial


h
direction (m)
Angular position on the sphere surface

in the x-y plane


Separation
angle (degree)
s

i j

Angular position on the sphere surface


in the y-z plane
Dimensionless convective heat flux (-)

Viscosity of the fluid (N s/m2)


Dimensionless temperature (-)
Density of the fluid (Kg/m3)
Shear stress (N/m2)

413

Engineering Applications of Computational Fluid Mechanics Vol. 4, No. 3 (2010)

13.

14.

15.
Downloaded by [b-on: Biblioteca do conhecimento online UBI] at 07:33 18 July 2016

16.
17.
18.
19.

20.
21.

22.
23.

24.

25.

26.

27. Seelay SE, Hummel RL, Smith JW (1975).


Experimental velocity profile in laminar flow
around sphere at intermediate Reynolds
number. J. Fluid Mech. 68(3):591608.
28. Shahcheraghi N, Dwyer HA (1998). Fluid
flow and heat transfer over a three
dimensional spherical object in a pipe. J. Heat
Transf. Trans. ASME 120(4):985990.
29. Vinoj Kurian, Mahesh NV, Kannan A (2009).
Numerical studies on laminar natural
convection inside inclined cylinders of unity
aspect ratio. Int. J. Heat and Mass Transf.
52(34):822838.
30. Wakiya S (1957). Viscous flows past a
spheroid. J. Phys. Soc. Jpn. 12:11301141.
31. Wham RM, Basaran OA, Byers CH (1996).
Wall effect on flow past solid sphere at finite
Reynolds number. Ind. Eng. Chem. Res.
35(3):864874.
32. Whitaker S (1972). Forced conversion heat
transfer correlations for flow in pipes, past
flat plates, single cylinders, single sphere and
flow in packed beds and tube bundles. AICHE
J. 18(2):361371.
33. Zuritz CA, McCoy SC, Sastry SK (1990).
Convective heat transfer coefficients for
irregular particles immersed in nonNewtonian fluid during tube flow. J. Food
Engg. 11:179194.

from a sphere at high Reynolds and Peclet


numbers. Int. J. of Heat and Mass Transf.
43(2):219229.
Haberman WL, Sayre RM (1958). Motion of
rigid and fluid spheres in stationary and
moving liquids inside cylindrical tubes. David
Taylor Model Basin Report No. 1143.
Hessel VH, Lowe H, Muller A, Kolb G
(2005). Chemical micro process engineering:
and
plants.
Wiley-VCE,
processing
Weinheim.
Holman JP (1997). Heat transfer. Singapore:
McGraw Hill.
Johnson TA, Patel VC (1999). Flow past a
sphere up to Reynolds number of 300. J.
Fluid Mech. 378(1):1970
Kalra TR and Uhlherr PHT (1971). 4th Aust.
Conf. on Hydraul. & Fluid Mech., Monash
University.
Kramers H (1946). Heat transfer from sphere
to flowing media. Physica 12:61.
Lee S (2000). A numerical study of the
unsteady wake behind a sphere in a uniform
flow at moderate Reynolds number.
Computers & Fluids 29(6):639667.
Magarvey RH, Bishop RL (1961). Transition
stages for three dimensional wakes. Canadian
Journal of Physics 39:14181422.
Maheshwari A, Chhabra RP, Biswas G
(2006). Effect of blockage on drag and heat
transfer from a single sphere and in-line array
of three spheres. Powder Technology
168(2):7483.
Oh JH, Lee SJ (1988). A Study on the
Newtonian Fluid Flow Past a Sphere in a
Tube. Korean J. of Chem. Eng. 5(2):190196.
Ramaswamy HS, Awuah GB, Simpson BK
(1996). Influence of particle characteristics on
fluid-to-particle heat transfer coefficient in a
pilot scale holding tube simulator. Food Res.
Int. 29(34):291300.
Ramaswamy HS, Awuah GB, Simpson B.K
(1997). Heat transfer and lethality
considerations in aseptic processing of
liquid/particle mixture: a review. Crit. Rev.
Food Sci. Nutrition 37(3):253286.
Sakamoto H, Haniu H (1990). A study on
vortex shedding in spheres in uniform flow.
Transactions of ASME: Journal of Fluid
Engineering 112(4):386392.
Salengke S, Sastry SK (1996). Residence
time distribution of cylindrical particles in a
curved section of a holding tube: effect of
particle concentration and bend radius of
curvature. J. Food Eng. 27(2):159176.

414

You might also like