Download as pdf or txt
Download as pdf or txt
You are on page 1of 29

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/279977978

Mixed Convection in Power-Law Fluids from a Heated


Semicircular Cylinder: Effect of Aiding Buoyancy

Article in Numerical Heat Transfer Applications · February 2015


DOI: 10.1080/10407782.2014.937242

CITATIONS READS

12 1,578

2 authors:

Anurag Kumar Tiwari Raj Chhabra


National Institute of Technology Jalandhar Indian Institute of Technology Kanpur
63 PUBLICATIONS 174 CITATIONS 402 PUBLICATIONS 12,242 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Anurag Kumar Tiwari on 09 August 2015.

The user has requested enhancement of the downloaded file.


This article was downloaded by: [Indian Institute of Technology Kanpur]
On: 26 October 2014, At: 23:02
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

Numerical Heat Transfer, Part A:


Applications: An International Journal of
Computation and Methodology
Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/unht20

Mixed Convection in Power-Law Fluids


from a Heated Semicircular Cylinder:
Effect of Aiding Buoyancy
a a
Anurag Kumar Tiwari & R. P. Chhabra
a
Department of Chemical Engineering , Indian Institute of
Technology , Kanpur , India
Published online: 23 Oct 2014.

To cite this article: Anurag Kumar Tiwari & R. P. Chhabra (2015) Mixed Convection in Power-Law
Fluids from a Heated Semicircular Cylinder: Effect of Aiding Buoyancy, Numerical Heat Transfer,
Part A: Applications: An International Journal of Computation and Methodology, 67:3, 330-356, DOI:
10.1080/10407782.2014.937242

To link to this article: http://dx.doi.org/10.1080/10407782.2014.937242

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
Numerical Heat Transfer, Part A, 67: 330–356, 2015
Copyright # Taylor & Francis Group, LLC
ISSN: 1040-7782 print=1521-0634 online
DOI: 10.1080/10407782.2014.937242

MIXED CONVECTION IN POWER-LAW FLUIDS FROM A


HEATED SEMICIRCULAR CYLINDER: EFFECT OF
AIDING BUOYANCY

Anurag Kumar Tiwari and R. P. Chhabra


Downloaded by [Indian Institute of Technology Kanpur] at 23:02 26 October 2014

Department of Chemical Engineering, Indian Institute of Technology,


Kanpur, India

Laminar aiding-buoyancy mixed-convection heat transfer in power-law fluids from a heated


semicircular cylinder with its flat base facing the oncoming flow was studied in this work.
The coupled momentum and energy equations were solved numerically over the following
ranges of conditions: Richardson number (0  Ri  2), Reynolds number (0.1  Re  25),
Prandtl number (1  Pr  100), and power-law index (0.3  n  1.8). Extensive results on
the velocity and temperature fields are presented and discussed in terms of the streamline
and isotherm contours adjacent to the heated object. Similarly, the distribution of the press-
ure coefficient and the local Nusselt number along the surface of the semicircular cylinder
are analyzed to delineate the influence of the relevant dimensionless parameters listed
above. Next, the pressure and total drag coefficient and surface average values of the Nus-
selt number describe the momentum and heat transfer aspects at the gross macroscopic
level. Both the pressure and total drag coefficient exhibit the expected dependence on
Richardson number, Prandtl number, and Reynolds number. Similarly, the average Nusselt
number bears a positive dependence on each Richardson number, Prandtl number, and
Reynolds number. Broadly, under otherwise identical conditions, heat transfer is augmented
in shear-thinning fluids (n < 1) whereas it is somewhat impeded in shear-thickening fluids
(n > 1) with reference to that in Newtonian fluids. Finally, the present numerical results
of the average Nusselt number (recast in terms of the Colburn jH factor) were correlated
as functions of the modified Reynolds number, Prandtl number, and power-law index
thereby enabling its prediction in a new application.

1. INTRODUCTION
Due to the frequent occurrence of power-law-type, non-Newtonian fluid
behavior in scores of industrial settings (e.g., polymers, foods, pharmaceutical and
personal care, health-care product manufacturing sectors), significant research effort
has been devoted to the elucidation of the influence of non-Newtonian fluid charac-
teristics on convective momentum and heat transport in wide-ranging geometries of
practical interest. Typical examples include flow in ducts of circular and noncircular
cross sections [1, 2], porous media flows [3–6], and batch-mixing vessels [7]. Similarly,
in new and novel designs of heat exchangers [8], compact heat exchangers [9], and

Received 14 March 2014; accepted 28 May 2014.


Address correspondence to R. P. Chhabra, Department of Chemical Engineering, Indian Institute
of Technology, Kanpur 208016, India. E-mail: chhabra@iitk.ac.in

330
MIXED CONVECTION IN POWER-LAW FLUIDS 331

NOMENCLATURE
Br Brinkman number (dimensionless) Pr modified Prandtl number
CD total drag coefficient (dimensionless) (dimensionless)
CDP pressure drag coefficient ps local surface pressure (Pa)
(dimensionless) p1 free stream pressure (Pa)
Cp pressure coefficient (dimensionless) Re Reynolds number (dimensionless)
cp specific heat of fluid (J=kg  K) Re modified Reynolds number
D diameter of semicircular cylinder (m) (dimensionless)
FD total drag force per unit length of Ri Richardson number (dimensionless)
cylinder (N=m) T fluid temperature (K)
FDP pressure drag force per unit length of Tw temperature at the surface of
Downloaded by [Indian Institute of Technology Kanpur] at 23:02 26 October 2014

cylinder (N=m) cylinder (K)


g acceleration due to gravity (m=s2) T1 free stream fluid temperature (K)
Gr Grashof number (dimensionless) Uch characteristic velocity (m=s)
havg average heat transfer coefficient Ux x-component of the velocity
(W=m2  K) (dimensionless)
hl local heat transfer coefficient Uy y-component of the velocity
(W=m2  K) (dimensionless)
I2 second invariant of the rate of the U1 free stream velocity (m=s)
deformation tensor (s2) b coefficient of thermal volumetric
jH Colburn factor (dimensionless) expansion (K1)
k thermal conductivity of fluid e rate of strain tensor (s1)
(W=m  K) g viscosity of fluid (Pa  s)
m power-law consistency coefficient q density of fluid (kg=m3)
(Pa  sn) q1 density of fluid at the reference
n power-law index (dimensionless) temperature T1 (kg=m3)
ns unit vector normal to the surface of s extra stress tensor (Pa)
the cylinder h fluid temperature (dimensionless)
 
Nuavg average Nusselt number ¼ TTT 1
w T1
(dimensionless)
Nul local Nusselt number (dimensionless)
P pressure (Pa) Subscripts
Pr Prandtl number (dimensionless) i, j, x, y (Cartesian coordinates)

heat sinks employed in cooling systems for electronics applications [10], flow and
heat transfer occurs over protrusions of circular and various noncircular crosssec-
tions. Consequently, over the past 10 years or so, there has been a spurt in research
activity aimed at exploring the role of power-law rheology on convective heat trans-
port in forced-, free-, and mixed-convection regimes from a circular cylinder (e.g., see
refs [11–15], square bars [16–20], elliptic cylinders [21, 22], and triangular cylinders
[23, 24]). One such geometry which has received only limited attention is the semi-
circular cylinder. This configuration finds applications in compact heat exchangers
with grooves and humps of semicircular cross section, thermal treatment of food-
stuffs (e.g., beans and carrots), and in the electronics industry. As noted elsewhere
[25–32], this configuration has received scant attention even in Newtonian fluids
[25, 30, 31]. On the other hand, the configuration of the curved face of the semicircu-
lar cylinder oriented towards the oncoming flow of power-law fluids has been studied
in detail recently in forced- [26, 27], free- [28], and mixed-convection [29] regimes.
However, these studies are restricted to the so-called steady-flow regime. There have
been even fewer studies on the laminar vortex shedding regime in Newtonian [30, 31]
332 A. K. TIWARI AND R. P. CHHABRA

and power-law liquids [32]. The effect of orientation of the cylinder and confinement
on the fluid mechanical and thermal aspects in Newtonian media has been investi-
gated [in refs. [30] and [31], respectively]. Thus, the current state of the art in this field
can be summarized as follows: currently available numerical predictions are restric-
ted mainly to the steady-flow regime, thereby excluding the possibility of direct com-
parison to the scant experimental results in Newtonian fluids at high Reynolds
numbers and=or Grashof numbers. Irrespective of the heat transfer regime, for a
fixed configuration, shear-thinning fluid behavior (n < 1) fosters heat transfer
whereas shear-thickening fluid behavior (n > 1) impedes heat transfer with reference
to that in Newtonian fluids under otherwise identical conditions.
Downloaded by [Indian Institute of Technology Kanpur] at 23:02 26 October 2014

It hardly needs to be emphasized here that the flow patterns in the close
proximity of the semicircular cylinder are strongly influenced by its orientation
and, in turn, these modifications in the flow field directly influence the temperature
field and eventually the rate of heat transfer. This effect gets further accentuated in
the case of free- and mixed-convection regimes due to the coupled nature of the velo-
city and temperature fields. In this work, consideration is given to the aiding-
buoyancy, mixed-convection heat transfer from a semicircular cylinder with its flat
base facing in the downward direction. This study thus extends and supplements
two recent works dealing with free- [33] and forced-convection [34] heat transfer
to power-law fluids from a heated semicircular cylinder oriented with its flat base
in the downward direction identical to that studied herein.
In the mixed-convection regime, the relative strengths of forced convection
(Reynolds number, Re) and free convection (Grashof number, Gr) effects are quan-
tified in terms of the familiar Richardson number, Ri ¼ Gr=Re2. Evidently, the two
limiting cases of free convection and forced convection correspond to Ri ! 1 and
Ri ! 0, respectively. Qualitatively, the value of Ri of order unity corresponds to
the situation when the imposed flow and the buoyancy-induced flow are of compa-
rable strength and it is thus not justified to neglect either of these contributions in
process design calculations under such conditions. Besides, the transitions from
the steady- to unsteady-flow regimes are also strongly influenced not only by the
value of Richardson number but also by the type of mixed-convection region (i.e.,
aiding-, opposing-, or cross-buoyancy configurations in general [35]. Thus, for
instance, Das et al. [36] studied the mixed-convection heat transfer in air (Pr ¼ 0.71)
from a semicircular cylinder with its curved surface facing the oncoming flow in the
cross-buoyancy configuration, and found that the steady-flow regime ceases to exist
at Ri  1.25 with a Reynolds number of 30.
From the preceding discussion, it is thus fair to conclude that little is known
about aiding-buoyancy, mixed-convection heat transfer from a semicircular cylin-
der with its flat surface transverse to the imposed upward flow even in Newtonian
fluids, let alone in power-law fluids. This work aims to elucidate the role of
Prandtl number, Richardson number, and power-law index on the momentum
and heat transfer characteristics of a long, semicircular cylinder by seeking numeri-
cal solutions of the pertinent equations over the following ranges of conditions:
0  Ri  2, 1  Pr  100, 0.1  Re  25, 0.3  n  1.8. The values of the power-law
index (n) spanned here are such that it includes the limiting case of the Newtonian
fluid (n ¼ 1), shear-thinning (n < 1), and shear-thickening (n > 1) types of fluid
behavior.
MIXED CONVECTION IN POWER-LAW FLUIDS 333

2. PROBLEM STATEMENT AND MATHEMATICAL FORMULATION


Consider a semicircular cylinder of diameter D (infinitely long in the
z-direction) heated to maintain its surface at a constant temperature, Tw, which is
immersed in a uniform stream of a power-law fluid with velocity, U1 and tempera-
ture, T1 < Tw, as shown schematically in Figure 1a. Since both imposed and
buoyancy-induced flows are aligned in the same direction, heat transfer occurs from
the heated cylinder to the fluid in the aiding-buoyancy, mixed-convection regime. In
order to keep the level of complexity at a tractable level, the thermophysical proper-
ties of the fluid (specific heat, cp, thermal conductivity, k, and power-law parameters
m and n) including the fluid density (except for the body force term in the y-compo-
Downloaded by [Indian Institute of Technology Kanpur] at 23:02 26 October 2014

nent of the momentum equation) are assumed to be independent of temperature.


Also, the viscous dissipation effects are neglected (i.e., Brinkman number is small,
Br << 1). For a narrow temperature range (small value of DT ¼ Tw  T1), the vari-
ation in fluid density with temperature is approximated by the standard Boussinesq
approximation (i.e., q ¼ q1[1  b(T  T1)]). The small value of DT used in this work
also justifies the use of the fluid properties at the mean film temperature, but at the

Figure 1. Schematics of unconfined flow over a semicircular cylinder: (a) Physical model; (b) Computational
domain and boundary condition.
334 A. K. TIWARI AND R. P. CHHABRA

same time it also restricts the applicability of these results to situations wherein the
maximum temperature difference in the system is not too large. Obviously, since it
is not possible to study numerically a truly unconfined flow, this condition is approxi-
mated here by caging the semicircular cylinder in an artificial concentric cylindrical
domain of diameter D1, as shown in Figure 1b. The value of diameter D1 of the con-
fining fluid envelope is chosen to be sufficiently large in order to minimize the bound-
ary effects on the flow and heat transfer phenomena while maintaining the required
computational effort at a modest level. The computational flow domain and bound-
ary conditions used in this work are shown in Figure 1b. The governing equations
(continuity, momentum, thermal energy) in their dimensionless forms are written
Downloaded by [Indian Institute of Technology Kanpur] at 23:02 26 October 2014

as follows:

. Continuity equation

qUx qUy
þ ¼0 ð1Þ
qx qy

. x-momentum equation
   
qðUx Ux Þ q Uy Ux qP 1 qsxx qsyx
þ ¼ þ þ ð2Þ
qx qy qx Re qx qy

. y-momentum equation
     
q Ux Uy q Uy Uy qP 1 qsxy qsyy
þ ¼ þ þ þ Rih ð3Þ
qx qy qy Re qx qy

. Thermal energy equation


  !
qð U x hÞ q U y h 1 q2 h q2 h
þ ¼ þ ð4Þ
qx qy Re  Pr qx2 qy2

For a power-law fluid, the deviatoric stress components are related to the
corresponding components of the rate of deformation tensor, eij, as

sij ¼ 2ðI2 =2Þð 2 Þ eij


n1
ð5Þ

where eij is the rate of deformation tensor and I2 is the second invariant of eij for
which the relevant expressions can be found in standard texts [37].
In Eq. (5), n is the flow behavior index. For Newtonian fluids, n ¼ 1, on the
other hand, n > 1 indicates shear-thickening and n < 1 denotes shear-thinning fluid
behavior.
The problem definition and formulation is completed by specifying the bound-
ary conditions employed in this work. Broadly, these correspond to that of no-slip
MIXED CONVECTION IN POWER-LAW FLUIDS 335

(Ux ¼ 0, Uy ¼ 0) and uniform temperature (h ¼ 1) on the surface of the semicircular


cylinder. The lower half of the surrounding envelope is treated as inlet and here,
the condition of uniform flow, Uy ¼ 1, in the y-direction and h ¼ 0 are employed.
The upper half of the bounding surface is treated as the outlet and on this
surface the Neumann homogeneous boundary condition is used. In addition, over
the range of conditions spanned here, since the velocity and temperature fields are
expected to be symmetric about the y-axis, the symmetry conditions are employed
on x ¼ 0 plane to economize on the computational effort. The aforementioned
conditions are also shown in Figure 1b.
The preceding governing equations have been rendered dimensionless using the
2
Downloaded by [Indian Institute of Technology Kanpur] at 23:02 26 October 2014

following scaling variables: D for all length variables, U1 for all velocities, q1 U1
for pressure, m(U1=D)n for stress components, etc. The temperature is nondimensio-
nalized as h ¼ TTT 1
w T1
. The dimensionless groups appearing in the above equations are
defined as follows:

. Reynolds number:

q1 Dn U1
2n
Re ¼ ð6Þ
m

. Prandtl number:
 
cp m U1 n1
Pr ¼ ð7Þ
k D

. Grashof number:
(   )2
3 q1 U1 1n
Gr ¼ gbðTw  T1 ÞD ð8Þ
m D

. Richardson number:

Gr gbðTw  T0 ÞD
Ri ¼ 2
¼ 2
ð9Þ
Re U1

Evidently, this flow is governed by four dimensionless parameters, namely, Grashof


number (Gr), Prandtl number (Pr), Richardson number (Ri) or combinations thereof,
and the power-law index (n). The numerical solution of the aforementioned governing
equations along with the boundary conditions maps the flow domain in terms of the
primitive variables (i.e., velocity (Ux, Uy), temperature (h), and pressure (P)). These,
in turn, can be post-processed to evaluate the local and global momentum and heat
transfer characterizing parameters such as drag coefficient, Nusselt number, etc. as
described in our recent studies [29, 33, 34] and hence are not repeated here. In brief,
the flow and temperature fields are visualized in terms of the streamline and isotherm
336 A. K. TIWARI AND R. P. CHHABRA

contours whereas the gross engineering characteristics are expressed in terms of the
drag coefficient (CD) and its pressure component (CDP), and the Nusselt number
(Nu) defined as follows:

2FD
CD ¼ 2 D
ð10Þ
q1 U1

2FDP
CDP ¼ 2 D
ð11Þ
q1 U1
Downloaded by [Indian Institute of Technology Kanpur] at 23:02 26 October 2014

The local Nusselt number on the surface of the semicircular cylinder is evaluated
using the temperature field as
 
hl D qh
Nul ¼ ¼ ð12Þ
k qns w

The surface average values of the Nusselt number can be obtained by integrating the
local Nusselt number over the whole surface of the semicircular cylinder.
Obviously, the macroscopic characteristics like drag coefficient and Nusselt
number will only be functions of Ri, Re, Pr, and n whereas the pressure coefficient
and local Nusselt number show additional spatial dependence. This work explores
and develops these relationships.

3. NUMERICAL SOLUTION PROCEDURE AND CHOICE OF NUMERICAL


PARAMETERS
In this study, the governing equations with the appropriate boundary
conditions were solved using the finite element-based solver COMSOL Multiphysics
(Version 4.2a), both for meshing the computational domain as well as mapping the
flow domain in terms of the primitive variables Ux  Uy  P  h. Since the gradients
are expected to be steep near the fluid–solid boundary, the grid near the surface of
the semicircular cylinder was made sufficiently fine to resolve such thin boundary
layers. Furthermore, the two-dimensional, laminar, steady-flow module was used
with PARDISO solver to solve the discretized governing differential equations. A
user-defined function has been introduced to add a volumetric source term to
account for the buoyancy force, as described elsewhere [38]. The parametric solver
was used to reduce the computational effort and time to reach the desired level of
convergence and hence the Richardson number, Ri, was parameterized using the
parametric sweep with a step of 0.1. Due to the nonlinear nature of the power-law
model, the automatic highly nonlinear damping method was used to facilitate the
convergence of the numerical solution. The relative tolerance criterion of 105 for
continuity, momentum, and energy was found to be sufficient and this also ensured
that the values of the drag coefficient and Nusselt number had stabilized at least up
to four significant digits. The use of a more stringent convergence criterion resulted
in negligible changes in the values of the surface pressure, drag, and Nusselt number
over the range of conditions embraced here.
MIXED CONVECTION IN POWER-LAW FLUIDS 337

Table 1. Domain independence tests at Re ¼ 0.1 and Pr ¼ 1

Total drag (CD) Average Nusselt number (Nuavg)

D1=D n ¼ 0.3 n ¼ 1.0 n ¼ 1.8 n ¼ 0.3 n ¼ 1.0 n ¼ 1.8

Ri ¼ 0
400 493.07 112.67 31.647 0.6036 0.5572 0.5681
500 493.16 113.28 31.758 0.6043 0.5568 0.5684
600 493.29 114.12 31.885 0.6049 0.5565 0.5686
700 493.43 115.39 32.081 0.6053 0.5564 0.5689
Ri ¼ 2
400 548.86 527.98 429.82 0.6581 0.7165 0.7182
Downloaded by [Indian Institute of Technology Kanpur] at 23:02 26 October 2014

500 556.74 533.96 433.43 0.6668 0.7197 0.7204


600 562.83 538.39 436.18 0.6735 0.7220 0.7222
700 567.76 541.87 438.41 0.6787 0.7239 0.7236

Much has been written in our recent studies [33, 34, 39] about the importance
of choosing an optimum value of D1 and a numerical mesh for this geometry. By
systematically varying the value of (D1=D) as 400, 500, 600, and 700, for extreme
values of the Richardson number as Ri ¼ 0 and Ri ¼ 2 and for three values of the
power-law index as n ¼ 0.3, n ¼ 1.0, and n ¼ 1.8, the values of CD, CDP, and Nuavg
were found to change very little (<1%) as the value of (D1=D) was increased
from 600 to 700 (Table 1). Therefore, the results reported herein are based on the
value of (D1=D) ¼ 600, which is superior to most of the previous studies in this
field [33, 34].
Similarly, four computational meshes (M1, M2, M3, M4, Table 2) were created
to select an optimum grid. An examination of the results presented in Table 2 clearly
shows that grid M2 is adequate in most cases for the present results to be free from
grid effects.
Finally, another simplifying assumption made here is that of the steady and
symmetric flow regime over the range of conditions spanned here. In order to
ascertain the validity of this assumption, time-dependent governing equations were
solved using the unsteady solver over the full domain. For the extreme cases of

Table 2. Grid independence tests at Re ¼ 25 and Pr ¼ 100

Total drag (CD) Average Nusselt number (Nuavg)

Grid d=D Np n ¼ 0.3 n ¼ 1.0 n ¼ 1.8 n ¼ 0.3 n ¼ 1.0 n ¼ 1.8

Ri ¼ 0
M1 0.008 160 3.7161 3.8465 4.0001 16.382 13.359 12.335
M2 0.006 195 3.7156 3.8456 3.9984 16.371 13.365 12.343
M3 0.005 260 3.7159 3.8452 3.9975 16.342 13.363 12.348
M4 0.004 320 3.7161 3.8454 3.9988 16.344 13.363 12.348
Ri ¼ 2
M1 0.008 160 5.3950 5.7426 6.2052 19.208 14.787 12.871
M2 0.006 195 5.3894 5.7410 6.2046 19.207 14.792 12.878
M3 0.005 260 5.3951 5.7400 6.2044 19.189 14.790 12.881
M4 0.004 320 5.3858 5.7404 6.2047 19.190 14.790 12.882
338 A. K. TIWARI AND R. P. CHHABRA

Re ¼ 25, Ri ¼ 2, Pr ¼ 100, n ¼ 0.3, and n ¼ 1.8, the resulting velocity, pressure, and
temperature fields were found to be steady and also the corresponding values of
CD, CDP, and Nusselt number were found to be virtually indistinguishable from
those obtained using the assumptions of the steady symmetric flow regime. There-
fore, over the range of conditions spanned here, the flow is indeed steady and sym-
metric about the y-axis.

4. RESULTS AND DISCUSSION


Extensive new results embracing wide ranges of dimensionless parameters –
Downloaded by [Indian Institute of Technology Kanpur] at 23:02 26 October 2014

Reynolds number (0.1  Re  25), Prandtl number (1  Pr  100), Richardson num-


ber (0  Ri  2), and the power-law index (0.3  n  1.8) were obtained to elucidate
their influence on the streamline and isotherm contours as well as on the gross engin-
eering parameters for a semicircular cylinder submerged in power-law fluids. How-
ever, before undertaking a detailed discussion of the new results, a few comparisons
with prior literature results are presented in Section 4.1 to establish the precision of
our results.

4.1. Validation of Results


Since no prior numerical results are available in the literature for the present
configuration, analogous results for a semicircular cylinder in the opposite configur-
ation [29] (curved face in the upstream direction) were used for this purpose, as
shown in Table 3. The present values are seen to be well within the error band of
2% quoted by Chandra and Chhabra [29]. Further validation is provided by compar-
ing the present and literature [40, 41] values (Figure 2) of surface vorticity (dimen-
sionless) and local Nusselt number distribution over the surface of an isothermal

Table 3. Comparison of the present results for a semicircular cylinder with its curved face in the upstream
direction in terms of the total drag coefficient and average Nusselt number at Pr ¼ 1 and Pr ¼ 100

Total drag (CD) Average Nusselt number (Nuavg)

Ref. [29] Present Ref. [29] Present

Pr Re n Ri ¼ 0 Ri ¼ 2 Ri ¼ 0 Ri ¼ 2 Ri ¼ 0 Ri ¼ 2 Ri ¼ 0 Ri ¼ 2

1 30 0.4 1.485 3.852 1.493 3.822 4.058 5.215 4.067 5.203


1 1.696 4.889 1.700 4.857 3.460 4.306 3.464 4.291
1.8 1.817 5.887 1.819 5.841 3.090 3.677 3.095 3.673
5 0.4 5.026 8.585 5.057 8.519 1.890 2.346 1.898 2.413
1 3.854 11.65 3.871 11.51 1.690 2.144 1.697 2.140
1.8 3.189 13.09 3.212 12.93 1.616 1.978 1.623 1.979
100 30 0.4 1.485 2.642 1.493 2.652 23.67 26.01 23.70 26.04
1 1.696 2.863 1.700 2.873 17.52 17.49 17.46 17.37
1.8 1.817 3.114 1.819 3.125 15.09 14.27 14.92 14.23
5 0.4 5.026 6.442 5.057 6.475 8.985 9.769 9.003 9.976
1 3.854 6.084 3.871 6.081 6.786 7.619 6.791 7.603
1.8 3.189 5.637 3.212 5.632 6.211 6.838 6.220 6.835
MIXED CONVECTION IN POWER-LAW FLUIDS 339
Downloaded by [Indian Institute of Technology Kanpur] at 23:02 26 October 2014

Figure 2. Comparison between the present and literature results on the vorticity (a) (nondimensional) and
local Nusselt number distribution (b) on the surface of a circular cylinder at different values of Gr for the
case of aiding-buoyancy flow for Pr ¼ 0.7 and Re ¼ 20.
340 A. K. TIWARI AND R. P. CHHABRA

cylinder in the mixed-convection regime in air. Once again, the correspondence is


seen to be very good for the surface vorticity and quite acceptable for local Nusselt
number. Indeed, the comparison shown in Figure 2 constitutes a much more
stringent validation than that shown in Table 3. Based on the preceding comparisons
coupled with our past experience, the new results reported herein are considered to
be reliable to within 2% or so.

4.2. Streamline and Isotherm Contours


Due to the no-slip condition and constant-temperature condition applied on
Downloaded by [Indian Institute of Technology Kanpur] at 23:02 26 October 2014

the surface of the cylinder, the velocity and temperature gradients would be steepest
near the surface of the submerged semicircular cylinder, and it is therefore instructive
to examine the streamline and temperature contours in close proximity of the semi-
circular cylinder as functions of the pertinent dimensionless groups. Representative
streamline patterns close to the surface of the heated semicircular cylinder for a
range of combinations of the values of Reynolds number, Re ¼ 0.1, 15, and 25,
of Richardson number Ri ¼ 0 and 2, of Prandtl number Pr ¼ 1 and 100, and
power-law index n ¼ 0.3, 1.0, and 1.8 are shown in Figures 3–5. At low Peclet
numbers, Pe ¼ Re  Pr, and in the absence of buoyancy effects (Ri ¼ 0), both the
streamline and isotherm contours closely follow the body contour as there is little
advection under these conditions. The boundary layer is seen to be slightly thinner
in shear-thinning fluids (n ¼ 0.3) and slightly thicker in shear-thickening fluids
(n ¼ 1.8) than that in Newtonian fluids under otherwise identical conditions.
Irrespective of the value of power-law index, some thinning of boundary layer is
evident with the increasing Prandtl number. At such low Reynolds numbers, how-
ever, the flow remains attached to the surface of the semicircular cylinder. Also, in
the absence of buoyancy, flow separation occurs and a pair of counterrotating vor-
tices is formed at the rear of the cylinder at Re ¼ 15 (Figure 4) and Re ¼ 25 (Figure 5).
The length of the recirculation region is seen to increase with increasing Reynolds
number, albeit for a given Reynolds number the recirculation length is seen to be
somewhat shorter in shear-thinning fluids (n < 1) and longer in shear-thickening
fluids (n > 1) with reference to that in Newtonian fluids, consistent with the previous
studies with circular and square bars [13, 16]. This trend can be explained, at least
qualitatively, in terms of the effective fluid viscosity which will be minimum
(maximum) for a shear-thinning (shear-thickening) fluid on=near the surface of the
cylinder and it progressively increases (decreases) spatially away from the cylinder.
This is tantamount to imposing a wall in terms of high-viscosity fluid in the case
of shear-thinning fluids, and such a confinement is known to shorten the wake
length. Also, with the introduction of buoyancy, due to the plume formation above
the semicircular cylinder, the recirculation zone is shortened irrespective of the type
of the fluid behavior and this is due to the opposing effects of the forced flow and
plume formation.
Finally, this section is concluded by noting that due to progressive thinning of
the thermal boundary layer (sharpening of the temperature gradient), it stands to
reason that heat transfer should bear a positive correlation with each of these para-
meters, namely, Re, Pr, and Ri. Indeed the results presented in one of the subsequent
sections corroborate these conjectures.
MIXED CONVECTION IN POWER-LAW FLUIDS 341
Downloaded by [Indian Institute of Technology Kanpur] at 23:02 26 October 2014

Figure 3. Representative streamline (right-half) and isotherm (left-half) contours at Re ¼ 0.1.


342 A. K. TIWARI AND R. P. CHHABRA
Downloaded by [Indian Institute of Technology Kanpur] at 23:02 26 October 2014

Figure 4. Representative streamline (right-half) and isotherm (left-half) contours at Re ¼ 15.


MIXED CONVECTION IN POWER-LAW FLUIDS 343
Downloaded by [Indian Institute of Technology Kanpur] at 23:02 26 October 2014

Figure 5. Representative streamline (right-half) and isotherm (left-half) contours at Re ¼ 25.

4.3. Surface Pressure Coefficient


Some further insights into the structure of the flow field and the resulting
hydrodynamic forces can be gained by examining the pressure distribution on the
surface of the semicircular cylinder (Figure 6). The dimensionless surface pressure
344 A. K. TIWARI AND R. P. CHHABRA
Downloaded by [Indian Institute of Technology Kanpur] at 23:02 26 October 2014

Figure 6. Variation of the pressure coefficient on the surface of a semicircular cylinder.


MIXED CONVECTION IN POWER-LAW FLUIDS 345

2
is expressed in the form of a coefficient, Cp ð¼ 2ðps  p1 Þ=q1 U1 Þ. In the absence of
the buoyancy effect (Ri ¼ 0), at low Reynolds numbers (Re ¼ 0.1) the pressure coef-
ficient decreases with power-law index on the front flat surface (A–B) and attains its
maximum value at the corner point B and then decreases along the rear curved sur-
face of the cylinder. However, the role of power-law index is seen to be reversed in
the rear of the cylinder. As the value of the Reynolds number is gradually increased,
the pressure coefficient decreases and the effect of power-law index also gradually
diminishes. This is due to the increasing role of inertial forces at the expense of
the viscous forces. The extent of pressure recovery is seen to increase with increasing
Reynolds number. For fixed values of Reynolds number, Prandtl number, and power-
Downloaded by [Indian Institute of Technology Kanpur] at 23:02 26 October 2014

law index, the pressure coefficient is seen to decrease with increasing the values of
Richardson number (Ri).

4.4. Pressure and Total Drag Coefficients


Naturally, the surface pressure distribution seen in the preceding section
directly influences the pressure component of the total drag force exerted on the
semicircular cylinder. Figures 7 and 8 show the combined effects of Reynolds num-
ber, Prandtl number, power-law index, and Richardson number on the pressure and
total drag coefficients. Broadly, both CDP and CD exhibit qualitatively similar func-
tional dependence on the pertinent parameters, albeit the pressure component is seen
to dominate the overall drag force under most conditions encompassed here. Indeed,
the value of CDP=CD ratio ranges from 0.61 to 1 over the range of the conditions
spanned here. At low Reynolds and Prandtl numbers, the influence of Richardson
number is seen to diminish with decreasing the values of the power-law index. This
is due to the fact that the velocity field decays very quickly in highly shear-thinning
fluids (e.g., n ¼ 0.3) under otherwise identical conditions. However, as the value of
the Reynolds number is gradually increased, buoyancy-driven flow is seen to increas-
ingly contribute to velocity gradients and hence the normal and tangential stresses
acting on the solid object are also enhanced. While the trends seen in Figures 7
and 8 are consistent with those observed for other shapes of cylinder such as circular
[13, 15], square [16], and semicircular cylinder [29], it needs to be recognized here that
the role of power-law index (n) is entwined due to not only its appearance in both
Reynolds number and Prandtl number, but also due to it being an independent
dimensionless parameter on its own. For a simple Newtonian fluid, the Grashof
number is proportional to the reciprocal of viscosity squared which, in the present
case, translates to m(U1=D)2(n  1) thereby yielding a dependence on the free stream
velocity which varies with the value of the power-law index, n (i.e., n > 1 or n < 1).
Similar additional difficulties arise here due to the dependence of Prandtl number
on the power-law index. Suffice to add here that the changeover of the curves corre-
sponding to different values of the power-law index seen in some cases (e.g., see
Figure 8 for Pr ¼ 1, Re ¼ 25, and Figure 7 for Re ¼ 10, Pr ¼ 1, and Re ¼ 25,
Pr ¼ 1) stems from the interplay between the viscous, inertial, and buoyancy forces,
all of which scale differently with velocity and power-law index. At low Reynolds
numbers when the forced convection is weak, as expected buoyancy effects are seen
to be pronounced which progressively diminish at high Reynolds numbers.
346 A. K. TIWARI AND R. P. CHHABRA
Downloaded by [Indian Institute of Technology Kanpur] at 23:02 26 October 2014

Figure 7. Dependence of the pressure drag coefficient on Re, Pr, Ri, and n.

4.5. Local Nusselt Number


Variation in local Nusselt number at any point on the surface of a semicircular
cylinder is determined by the local temperature gradient normal to the surface. In the
case of a Newtonian fluid characterized by a constant value of viscosity, each point
MIXED CONVECTION IN POWER-LAW FLUIDS 347
Downloaded by [Indian Institute of Technology Kanpur] at 23:02 26 October 2014

Figure 8. Dependence of the total drag coefficient on Re, Pr, Ri, and n.

on the surface of a semicircular cylinder thus corresponds to the same values of Re,
Pr, and Ri. In contrast, in the present case, since the effective shear rate (hence fluid
viscosity) varies from one point to another on the surface of the semicircular cylin-
der, the local values of these parameters (Re, Pr, and Ri) not only differ from the
values based on free stream values but also vary along the surface of the semicircular
cylinder and these, in turn, exert varying levels of influence on the value of the local
Nusselt number. Figure 9 shows representative results on the distribution of the local
Nusselt number along the surface of the heated cylinder for a range of combinations
of the values of Re, Pr, n, and Ri. Detailed examination of these results reveals the
348 A. K. TIWARI AND R. P. CHHABRA
Downloaded by [Indian Institute of Technology Kanpur] at 23:02 26 October 2014

Figure 9. Variation of the local Nusselt number on the surface of the semicircular cylinder.
MIXED CONVECTION IN POWER-LAW FLUIDS 349

following overall trends: at low Reynolds and Prandtl numbers (weak convection),
the conduction mechanism dominates the overall heat transfer and, as expected,
the power-law index (n) exerts very little influence on heat transfer under these
conditions (e.g., see results for Re ¼ 0.1, Pr ¼ 1) and even the superimposition of
buoyancy flow helps only a little. The effect of power-law index begins to manifest
increasingly as the extent of convective transport increases with increasing Reynolds
number, Prandtl number, Richardson number, or all these three. As expected, shear-
thinning fluid behavior (n < 1) promotes heat transfer whereas shear-thickening
(n > 1) fluid behavior is seen to impede it, but only when there is a fair degree of con-
vection. Apart from these trends, the effect of power-law index is significant along
Downloaded by [Indian Institute of Technology Kanpur] at 23:02 26 October 2014

the entire surface at low Reynolds numbers (viscous forces dominate here) whereas
it is limited mainly to the flat surface at high Reynolds numbers. The maximum value
of the local Nusselt number at the corner point B is due to the sudden change in body
contour. At low Reynolds numbers, due to negligible fluid inertia, a fluid element is
able to negotiate this abrupt change whereas at high Reynolds numbers (Re ¼ 25),
due to the formation of the recirculation region which extends almost up to the cor-
ner point, there is a weak minimum also at the point of separation. This effect gets
accentuated when buoyancy-induced flow is superimposed on forced flow (e.g., see
the results in Figure 9 for Re ¼ 25, Pr ¼ 100, and Ri ¼ 0, Ri ¼ 2). Strictly speaking,
the temperature gradient is indeterminate at the corner point B and, since the com-
putational mesh used in this work is quite fine, such a discontinuity is not evident in
these results. On the other hand, such high values of the local Nusselt number at the
corner points (or their uncertainty) are of no practical significance as the correspond-
ing heat transfer area is identically zero. In summary, in view of the preceding trends
seen in isotherm contours (Figures 3–5) and local Nusselt number distribution
(Figure 9), it stands to reason that heat transfer should bear a positive dependence
on each of Re, Pr, and Ri for a given value of power-law index (n). Similarly, for
fixed values of Re, Pr, and Ri, heat transfer is facilitated in shear-thinning fluids
(n < 1) and it is impeded in shear-thickening fluids (n > 1) with reference to that in
Newtonian fluids, as will be seen in the next section.

4.6. Average Nusselt Number


From a practical standpoint, reliable values of the surface average Nusselt
number are frequently required in thermal process design calculations. Naturally,
as postulated above, the average Nusselt number would be a function of the four
dimensionless parameters, namely, Re, Pr, Ri, and n. Figure 10 shows representative
results elucidating this functional relationship. The average Nusselt number shows a
positive dependence on Reynolds number, Prandtl number, and Richardson num-
ber, respectively, which is simply due to the progressive thinning of the thermal
boundary layer. In the pure forced convection limit (Ri ¼ 0), the average Nusselt
number shows a positive dependence on Prandtl number and Reynolds number
for a given value of power-law index (n). On the other hand, at extremely small
values of Peclet number (Re ¼ 0.1, Pr ¼ 1), it is very difficult to induce convection
in highly shear-thinning fluids (n ¼ 0.3 and n ¼ 0.5) due to a rapid increase in their
effective viscosity, as can be seen in Figure 10a. Similarly, an unusual effect is seen
at about Ri ¼ 0.2 or so for Re ¼ 10 and Re ¼ 25. It is useful to recall here that at
350 A. K. TIWARI AND R. P. CHHABRA
Downloaded by [Indian Institute of Technology Kanpur] at 23:02 26 October 2014

Figure 10. Effects of Reynolds number (Re), Prandtl number (Pr), Richardson number (Ri), and
power-law index (n) on average Nusselt number.

Ri ¼ 0, there would be a well-developed recirculation zone at the rear of the cylinder


both at Re ¼ 10 and Re ¼ 25 [34]. The gradual introduction of buoyancy-induced
flow (Ri increasing) counters this propensity for flow separation. The minimum in
the Nusselt number corresponds to where these two effects, namely, growth in the
MIXED CONVECTION IN POWER-LAW FLUIDS 351

recirculation region due to Reynolds number is in equilibrium with the suppression


brought about by the buoyancy-induced flow (value of Ri). However, the decrease in
the value of the average Nusselt number is only of the order of 10% at most in an
extremely shear-thickening fluid (n ¼ 1.8). Such high values of the power-law index
are not at all common in process engineering applications [2].
Finally, it is useful to correlate the present numerical results on average Nusselt
number in terms of the influencing parameters, namely, Re, Pr, Ri, and n. Thus, one
can express this functional relationship as follows:

Nuavg ¼ f ðPr; Re; Ri; nÞ ð13Þ


Downloaded by [Indian Institute of Technology Kanpur] at 23:02 26 October 2014

As noted elsewhere [38, 39], the functional relationship shown in Eq. (13) can be sim-
plified by introducing an effective velocity, Uch, simply as a sum of the external and
buoyancy-induced contributions as

Uch ¼ U1 þ ðgbDDT Þ1=2 ð14Þ

This choice of velocity scaling yields the following modified definitions of the
Reynolds number (Re ) and Prandtl number (Pr ) and it also eliminates the
Richardson number (Ri) from the set of parameters in Eq. (13) as
 pffiffiffiffiffi2n
Re ¼ Re 1 þ Ri ð15Þ

 pffiffiffiffiffin1
Pr ¼ Pr 1 þ Ri ð16Þ

In other words, Eq. (13) can now be rewritten as


Nuavg ¼ f ðPr ; Re ; nÞ ð17Þ

This approach will still yield a family of curves corresponding to individual


values of the modified Prandtl number, Pr , and n (shown in Figure 11). Further
consolidation of the present results can be achieved by introducing the usual
Colburn jH factor defined as
Nuavg
jH ¼ ð18Þ
Re Pr1=3

The present heat transfer results are thus correlated by the following simple
expression:
 
0:72 3n þ 1 0:55
jH ¼ 2=3 ð19Þ
Re 4n

Equation (19) reproduces the present numerical results (1,008 data points) with an
average error of 9.26%, which rises to a maximum of 23.71%, without any
discernable trends. Furthermore, Eq. (19) embodies the widely accepted scaling of
 
the average Nusselt number (Nu  Pr 1=3) as well as that of jH factor (jH  Re 2=3)
352 A. K. TIWARI AND R. P. CHHABRA
Downloaded by [Indian Institute of Technology Kanpur] at 23:02 26 October 2014

Figure 11. Dependence of average Nusselt number on modified Reynolds and Prandtl numbers at n ¼ 0.3
and n ¼ 1.8.

as shown in Figure 12. Similarly, the positive index of the term ((3n þ 1)=4n) also
captures the fact that, all else being equal, shear-thinning behavior enhances heat
transfer which can lead to energy savings in process engineering applications.
Before concluding this section, it is worthwhile to compare the present results
with those for the configuration studied by Chandra and Chhabra [29] of when the

Figure 12. Variation in jH factor with modified Reynolds number and power-law index.
MIXED CONVECTION IN POWER-LAW FLUIDS 353

curved surface is oriented towards the oncoming free stream of fluid. As expected,
both the hydrodynamic drag experienced by the semicircular cylinder and the
accompanying Nusselt number values are strongly dependent on the orientation of
the semicircular cylinder with respect to both the direction of mean flow and the
gravity vector. Intuitively, to move a cylinder with its flat base downward in a fluid
requires the fluid to be displaced across a distance of D. On the other hand, when the
cylinder with its curved surface is oriented towards the oncoming flow, it starts with
a point contact which gradually extends up to a distance D. Therefore, one would
expect the drag to be higher in the former case than in the latter. Indeed, the drag
results shown here in Figure 8 and reported in [29] substantiate this conjecture, at
least under the limit of Ri ¼ 0. The difference between the two values decreases with
Downloaded by [Indian Institute of Technology Kanpur] at 23:02 26 October 2014

the increasing buoyancy effect and eventually the role of orientation switches over.
By invoking this reasoning, the Nusselt number also exhibits similar dependence on
orientation under otherwise identical conditions. In conclusion, it is thus possible to
modulate the rate of heat transfer by manipulating the rheological properties of the
fluid, strength of buoyancy-induced flow, and the orientation of the semicircular cyl-
inder in a given situation.

5. CONCLUSIONS
In this work, extensive new results on the detailed flow and heat transfer char-
acteristics for aiding-buoyancy, steady mixed-convection heat transfer in power-law
fluids from a heated semicircular cylinder with its flat face oriented in the upstream
direction are reported over the following ranges of conditions: 0  Ri  2,
0.3  n  1.8, 0.1  Re  25, and 1  Pr  100. The structure of the flow and tempera-
ture fields is analyzed in terms of the streamline and isotherm contours in close prox-
imity of the heated semicircular cylinder to delineate the regions of high gradients
and=or secondary flows. At the next level, the fluid mechanical and heat transfer
aspects are examined in terms of the distributions of surface pressure and local
Nusselt number along the surface of the semicircular cylinder. Regardless of the flow
conditions and type of fluid, both the value of the pressure coefficient and the local
Nusselt number are found to be maximal at the corner point of the semicircular cyl-
inder. The total drag coefficient is seen to increase with the increasing value of
Richardson number and decreasing value of Reynolds number, Prandtl number,
and power-law index due to the sharpening of the velocity gradient on the surface
of the semicircular cylinder. The average Nusselt number also increases with each
of this parameter (i.e., Reynolds number, Prandtl number, and Richardson number).
All else being equal, shear-thinning fluid behavior promotes heat transfer over and
above that in Newtonian fluids, which can result in energy economy in process engin-
eering applications. Conversely, it can be used to regulate the rate of heating=cooling
in process applications, especially during the processing of temperature-sensitive
materials. Finally, the present numerical results of heat transfer were correlated in
terms of the jH factor using a simple expression thereby enabling the interpolation
of the present results for the intermediate values of the governing parameters and=or
or prediction of the heat transfer coefficient in a new application.
354 A. K. TIWARI AND R. P. CHHABRA

REFERENCES
1. A. Lawal and A. S. Mujumdar, Laminar Duct Flow and Heat Transfer to Purely Viscous
Non-Newtonian Fluids, in A. S. Mujumdar, and R. A. Mashelkar (eds.), Advances in
Transport Processes, pp. 352–443, Wiley Interscience, New York, 1989.
2. R. P. Chhabra and J. F. Richardson, Non-Newtonian Flow and Applied Rheology:
Engineering Applications, 2nd ed., Butterworth-Heinemann, Oxford, 2008.
3. R. P. Chhabra, Bubbles, Drops, and Particles in Non-Newtonian Fluids, 2nd ed., CRC
Press, Boca Raton, FL, 2006.
4. A. Banerjee, A. Haji-Sheikh, and S. Nomura, Heat Transfer with Axial Conduction in
Triangular Ducts Filled with Saturated Porous Materials, Numer. Heat Transfer A, vol.
62, pp. 1–24, 2012.
Downloaded by [Indian Institute of Technology Kanpur] at 23:02 26 October 2014

5. M. Zeng, P. Yu, F. Xu, and Q. W. Wang, Natural Convection in Triangular Attics Filled
with Porous Medium Heated from Below, Numer. Heat Transfer A, vol. 63, pp. 735–754,
2013.
6. A. J. Chamkha and M. A. Ismael, Conjugate Heat Transfer in a Porous Cavity Heated by
a Triangular Thick Wall, Numer. Heat Transfer A, vol. 63, pp. 144–158, 2013.
7. R. P. Chhabra, Fluid Mechanics and Heat Transfer with Non-Newtonian Liquids in
Mechanically Agitated Vessels, Adv. Heat Transfer, vol. 37, pp. 77–178, 2003.
8. S. Kakac and H. Liu, Heat Exchangers: Selection, Rating and Thermal Design, 2nd ed.,
CRC Press, Boca Raton, FL, 2002.
9. J. E. Hesselgreaves, Compact Heat Exchangers, Pergamon Press, Oxford, 2001.
10. D. S. Steinberg, Cooling Techniques for Electronic Equipment, 2nd ed., Wiley, New York,
1991.
11. A. A. Soares, J. M. Ferreira, and R. P. Chhabra, Flow and Forced Convection Heat
Transfer in Cross Flow of Non-Newtonian Fluids over a Circular Cylinder, Ind. Eng.
Chem. Res., vol. 44, pp. 5815–5827, 2005.
12. R. P. Bharti, R. P. Chhabra, and V. Eswaran, Steady Forced Convection Heat Transfer
from a Heated Circular Cylinder to Power-Law Fluids, Int. J. Heat Mass Transfer, vol. 5,
pp. 977–990, 2007.
13. A. T. Srinivas, R. P. Bharti, and R. P. Chhabra, Mixed Convection Heat Transfer from a
Cylinder in Power-Law Fluids: Effect of Aiding Buoyancy, Ind. Eng. Chem. Res., vol. 48,
pp. 9735–9754, 2009.
14. A. Prhashanna and R. P. Chhabra, Laminar Natural Convection from a Horizontal
Cylinder in Power-Law Fluids, Ind. Eng. Chem. Res., vol. 50, pp. 2424–2440, 2011.
15. A. A. Soares, J. Anacleto, L. Caramelo, J. M. Ferreira, and R. P. Chhabra, Mixed
Convection from a Circular Cylinder to Power-Law Fluids, Ind. Eng. Chem. Res., vol. 48,
pp. 8219–8231, 2009.
16. A. K. Dhiman, N. Anjaiah, R. P. Chhabra, and V. Eswaran, Mixed Convection from a
Heated Square Cylinder to Newtonian and Power-Law Fluids, J. Fluids Eng. (Trans
ASME), vol. 129, pp. 506–513, 2007.
17. A. K. Sahu, R. P. Chhabra, and V. Eswaran, Two-Dimensional Unsteady Laminar
Flow of a Power-Law Fluid across a Square Cylinder, J. Non-Newt. Fluid Mech., vol. 160,
pp. 157–167, 2009.
18. A. K. Sahu, R. P. Chhabra, and V. Eswaran, Forced Convection Heat Transfer from a
Heated Square Cylinder to Power-Law Fluids in the Unsteady Flow Regime, Numer. Heat
Transfer A, vol. 56, pp. 109–131, 2009.
19. A. K. Sahu, R. P. Chhabra, and V. Eswaran, Effect of Blockage on Forced Convection
Heat Transfer from a Heated Square Cylinder to Power-Law Fluids, Numer. Heat
Transfer A, vol. 58, pp. 641–659, 2010.
MIXED CONVECTION IN POWER-LAW FLUIDS 355

20. C. Sasmal and R. P. Chhabra, Laminar Natural Convection from a Heated Square
Cylinder Immersed in Power-Law Liquids, J. Non-Newtonian Fluid Mech., vol. 166,
pp. 811–830, 2011.
21. R. P. Bharti, P. Sivakumar, and R. P. Chhabra, Forced Convection Heat Transfer from
an Elliptic Cylinder to Power-Law Fluids, Int. J. Heat Mass Transfer, vol. 51, pp. 1838–
1853, 2008.
22. C. Sasmal and R. P. Chhabra, Effect of Aspect Ratio on Natural Convection in
Power-Law Fluids from a Heated Horizontal Elliptic Cylinder, Int. J. Heat Mass
Transfer, vol. 55, pp. 4886–4899, 2012.
23. A. Prhashanna, A. K. Sahu, and R. P. Chhabra, Flow of Power-Law Fluids Past an
Equilateral Triangular Cylinder: Momentum and Heat Transfer Characteristics, Int. J.
Downloaded by [Indian Institute of Technology Kanpur] at 23:02 26 October 2014

Therm. Sci., vol. 50, pp. 2027–2042, 2011.


24. A. K. Tiwari and R. P. Chhabra, Effect of Orientation on Steady Laminar Free Convec-
tion Heat Transfer in Power-Law Fluids from a Heated Triangular Cylinder, Numer. Heat
Transfer A, vol. 65, pp. 780–801, 2014.
25. A. Chandra and R. P. Chhabra, Flow Over and Forced Convection Heat Transfer in
Newtonian Fluids from a Semi-Circular Cylinder, Int. J. Heat Mass Transfer, vol. 54,
pp. 225–241, 2011.
26. A. Chandra and R. P. Chhabra, Momentum and Heat Transfer Characteristics of a
Semi-Circular Cylinder Immersed in Power-Law Fluids in the Steady Flow Regime, Int.
J. Heat Mass Transfer, vol. 54, pp. 2734–2750, 2011.
27. A. Chandra and R. P. Chhabra, Influence of Power-Law Index on Transitional Reynolds
Numbers for Flow over a Semi-Circular Cylinder, Appl. Math. Model., vol. 35, pp. 5766–
5785, 2011.
28. A. Chandra and R. P. Chhabra, Laminar Free Convection from a Horizontal Semicircu-
lar Cylinder to Power-Law Fluids, Int. J. Heat Mass Transfer, vol. 55, pp. 2934–2944,
2012.
29. A. Chandra and R. P. Chhabra, Mixed Convection from a Heated Semi-Circular Cylinder
to Power-Law Fluids in the Steady Flow Regime, Int. J. Heat Mass Transfer, vol. 55,
pp. 214–234, 2012.
30. D. Chatterjee, B. Mondal, and P. Halder, Unsteady Forced Convection Heat Transfer
Over a Semi-Circular Cylinder at Low Reynolds Numbers, Numer. Heat Transfer A,
vol. 63, pp. 411–429, 2013.
31. A. P. S. Bhinder, S. Sarkar, and A. Dalal, Flow Over and Forced Convection Heat
Transfer Around a Semi-Circular Cylinder at Incidence, Int. J. Heat Mass Transfer,
vol. 55, pp. 5171–5184, 2012.
32. A. Chandra and R. P. Chhabra, Momentum and Heat Transfer from a Semi-Circular
Cylinder to Power-Law Fluids in the Vortex Shedding Regime, Numer. Heat Transfer
A, vol. 63, pp. 489–510, 2013.
33. A. K. Tiwari and R. P. Chhabra, Laminar Natural Convection in Power-Law Liquids
from a Heated Semi-Circular Cylinder with its Flat Side Oriented Downward, Int. J. Heat
Mass Transfer, vol. 58, pp. 553–567, 2013.
34. A. K. Tiwari and R. P. Chhabra, Momentum and Heat Transfer Characteristics for
the Flow of Power-Law Fluids over a Semi-Circular Cylinder, Numer. Heat Transfer A,
vol. 66, pp. 1365–1388, 2014.
35. M. M. Zdravkovich, Flow Around Circular Cylinders: Fundamentals, vol 1, Oxford
University Press, New York, 1997.
36. G. Das, S. Sarkar, and A. Dalal, Mixed Convective Flow, and Heat Transfer Around a
Semi-Circular Cylinder, 22nd National and 11th ISHMT-ASME Conference on Heat
and Mass Transfer, IIT Kharagpur, pp. HMTC1300586, 2013.
356 A. K. TIWARI AND R. P. CHHABRA

37. R. B. Bird, W. E. Stewart, and E. N. Lightfoot, Transport Phenomena, 2nd ed., Wiley,
New York, 2001.
38. N. Nirmalkar and R. P. Chhabra, Mixed Convection from a Heated Sphere in Power-Law
Fluids, Chem. Eng. Sci., vol. 89, pp. 49–71, 2012.
39. A. Bose, N. Nirmalkar, and R. P. Chhabra, Forced Convection from a Heated Equilateral
Triangular Cylinder in Bingham Plastic Fluids, Numer. Heat Transfer A, vol. 66, pp. 107–
129, 2014.
40. H. M. Badr, Laminar Combined Convection from a Horizontal Cylinder – Parallel and
Contra Flow Regimes, Int. J. Heat Mass Transfer, vol. 27, pp. 15–27, 1984.
41. S. C. R. Dennis and G.-Z. Chang, Numerical Solutions for Steady Flow Past a Circular
Cylinder at Reynolds Numbers Up to 100, J. Fluid Mech., vol. 45, pp. 49–64, 1970.
Downloaded by [Indian Institute of Technology Kanpur] at 23:02 26 October 2014

View publication stats

You might also like