You are on page 1of 6

Ali M.

Jawarneh
Assistant Professor
Department of Mechanical Engineering,
Hashemite University,
Zarqa 13115, Jordan
e-mail: jawarneh@hu.edu.jo
Georgios H. Vatistas
Professor
Department of Mechanical and Industrial
Engineering,
Concordia University 1455 DeMaisonneuve Blvd.
West,
Montreal, H3G 1M8 Canada
e-mail: vatistas@me.concordia.ca
Reynolds Stress Model in the
Prediction of Conned Turbulent
Swirling Flows
Strongly swirling vortex chamber ows are examined experimentally and numerically
using the Reynolds stress model (RSM). The predictions are compared against the ex-
perimental data in terms of the pressure drop across the chamber, the axial and tangen-
tial velocity components, and the radial pressure proles. The overall agreement between
the measurements and the predictions is reasonable. The predictions provided by the
numerical model show clearly the forced and free vortex modes of the tangential velocity
prole. The reverse ow (or back ow) inside the core and near the outlet, known from
experiments, is captured by the numerical simulations. The swirl number has been found
to have a measurable impact on the ow features. The vortex core size is shown to
contract with the swirl number which leads to higher pressure drop, higher peak tangen-
tial velocity, and deeper radial pressure proles near the axis of rotation. The adequate
agreement between the experimental data and the simulations using RSM turbulence
model provides a valid tool to study further these industrially important swirling
ows. DOI: 10.1115/1.2354530
1 Introduction
Swirling ow occurs in many engineering applications, such as
vortex separators, pumps, gas turbine combustors, furnaces, spray
dryer, the vortex valve, the vortex combustor, and gas-core
nuclear rocket. In modern combustors, swirl is used to produce
good mixing and to improve the ame stability. In all conned
vortex applications, it is important to understand adequately the
overall ow eld evolution as a function of both the geometrical
and ow parameters. A good knowledge of these ows will im-
prove the design and performance of a variety of vortex devices.
It is well known that the tangential velocity of the conned
uid changes from free to forced vortex as the ow approaches
the axis of rotation. The static pressure in an attempt to balance
the centrifugal force will reduce from a maximum value near the
cylindrical wall of the chamber to a minimum on the axis of
rotation. Depending on the inlet swirl intensity, the pressure inside
the core might drop below the outside ambient, thus inducing a
reverse ow. Escudier et al. 1 demonstrated experimentally the
axial and swirl velocities distributions using Laser Doppler An-
emomerty LDA measurements. The experiments were per-
formed with water for a range of exit diameters. The observation
revealed a remarkable change in the vortex structure as the exit
diameter is reduced, where the vortex core size changes from a
thick core to a thin core. In addition, the axial velocity was found
able to develop proles ranging from jetlike to wakelike shapes,
thus revealing the evolution of the reverse ow. Vatistas 2 re-
ported a model for single- or double-celled intense vortices, de-
pending on the values of scaling constants. It was shown that the
axial velocity component may attain proles ranging from jetlike
to wakelike. The last was an attempt to mathematically simulate
the reverse ow conditions. Sullivans 3 two-celled vortex
model can also approximately simulate the direction reversal of
the radial and axial velocity components near the axis of rotation.
The major obstacle in numerical modeling of complex turbulent
swirling ows is the selection of appropriate turbulence closure
models. In simple ow cases, the k model performs well. How-
ever, for strongly swirling ows that involves severe streamline
bending it fails. The last conclusion is clearly evident in a variety
of studies; see for example the work of Nallasamy 4, Nejad et al.
5, and Weber 6. A review of second-moment computations for
engineering ows has been provided by Launder 7, Leschziner
8, and Ferziger and Peric 9. The results of these computations
demonstrate the superiority of RSM over eddy-viscosity models
for curved ows, swirling ows and recirculating ows. Jones et
al. 10 have studied the performance of second moment closure
turbulence models for swirling ow in a cylindrical combustion
chamber. The models are found to predict mean and turbulent ow
quantities well. German and Mahmud 11 have shown that the
overall agreement between the measurements and the predictions
obtained with both the k and Reynolds-stress turbulence mod-
els are reasonably good. However, some features of the isothermal
and combusting ow elds are better predicted by the Reynolds-
stress model. Jakirlic et al. 12 have shown numerically using
three versions of the second-momentum closure and two eddy-
viscosity models that the second-momentum models are superior.
However, difculties in predicting accurately the transformation
from free- to forced-vortex modes or the determination of the
normal stress components inside the core still remain. Vortex
chamber ows at low Reynolds number via direct numerical
simulations were investigated by Orland and Fatica 13. Jones
and Pascau 14 and Hoekstra et al. 15 used the k turbulent
model and a Reynolds stress transport equation model of a strong
conned swirling ow. Once more, comparisons of the results
with measurements show the superiority of the transport equation
model, where k gave large discrepancies between the measured
and predicted velocity elds.
Since a Reynolds stress model RSM takes into consideration
the effects of severe streamline bending due to swirl in a more
appropriate way than the one- and two-equation models, it is best
suited for the present study. The aim of this paper is to study the
ow features in a vortex chamber experimentally and numerically
using FLUENT Fluent Inc., and to compare the results obtained
using a Reynolds stress model to available experimental data of
vortex chambers operating under different swirl numbers. Perfor-
mance assessment of the RSM in the predicting turbulent, strongly
swirling vortex chamber ows will also be one of the objectives.
Contributed by the Fluids Engineering Division of ASME for publication in the
JOURNAL OF FLUIDS ENGINEERING. Manuscript received February 9, 2005; nal manu-
script received March 22, 2006. Assoc. Editor: Ugo Piomelli.
Journal of Fluids Engineering NOVEMBER 2006, Vol. 128 / 1377 Copyright 2006 by ASME
2 Experimental Setup
The experiments have been conducted using a jet-driven vortex
chamber similar to the one utilized by Vatistas et al. 16. The
main difference between the two is that in the latest version,
shown schematically in Fig. 1. It has a cylindrical conguration
with constant cross-sectional area R
o
=7 cm and a central axis
outlet and circumferential inlets. Swirl is imparted to the uid via
the vortex generator shown in Figs. 1 and 2. It has four perpen-
dicular air inlets where the compressed air is induced. The re-
quired set of inlet conditions is obtained by the insertion of the
appropriate vortex generator blocks swirler into the vortex gen-
erator assembly a long the periphery of the vortex generator. A
number of openings of a circular cross section d
in
are drilled at
a specied angle =30 deg. When the air ow passes through the
swirlers, it is guided to enter the vortex chamber in the radial and
tangential directions so that swirl is formed inside the vortex
chamber. The swirler has 16 holes with diameter d
in
=1.267 cm
and inlet area A
in
=20.177 cm
2
. Chamber diameter ratio ,
which is dened as the ratio of the diameter of a vortex chamber
D
o
to the diameter of the exit hole D
e
, was varied from
=2.5, 3.33, 3.67, 4.0, 5.01, 5.29, 5.80, 6.47, 7.08 to 7.45. Chamber
aspect ratio , which is dened as the ratio of the chamber
length L to the diameter of a vortex chamber D
o
was xed at
=3.00. Area ratio , which is dened as the ratio of the total
inlet area A
in
to the cross-sectional area of the vortex chamber
A
o
was xed at =0.131.
The measurements were made at inlet air ow rate Q
in
=0.0187 m
3
/ s, which is corresponding to Reynolds number Re
o
=11,592, which is dened based on the average velocity as
Re
o
=
4Q
in
D
o
The static pressure is measured by a series of taps located ahead
of the tangential ports and is averaged by connecting in parallel all
the pressure pickup tubes into a common tube. The measurements
of the mean gage pressure p=p
in
p
a
were obtained using a
U-tube lled with Meriam oil, having a specic gravity equal to
1.00. The estimated uncertainty is less than 8% for the pressure
drop measurements. A rotameter was used to measure the volu-
metric ow rate of the inlet air. This was carefully calibrated in
standard conditions 1 atm and 200.5 %C. For the ow rate
used, the uncertainty was estimated to be 2%.
3 Computational Details
Governing Equations. In Reynolds averaging, the solution
variables in the instantaneous Navier-Stokes equations are decom-
posed into the mean and uctuating components. For the velocity
components: u
i
=u
i
+u
i
where u
i
and u
i
are the mean and uctu-
ating velocity components. Likewise, for pressure and other scalar
quantities: =

+ where denotes a scalar such as pressure.


Substituting expressions of this form for the ow variables into
the instantaneous continuity and momentum equations and drop-
ping the over-bar on the mean velocity u yields the momentum
equations. They can be written in Cartesian tensor form as

x
i
u
i
= 0 1

x
j
u
i
u
j
=
P
x
i
+

x
j

u
i
x
j
+
u
j
x
i

2
3

ij
u
l
x
l

+

x
j
u
i
u
j
2
Equations 1 and 2 are called Reynolds-averaged Navier-Stokes
RANS equations. Additional terms now appear that represent the
effects of turbulence. These Reynolds stresses, u
i
u
j
, must be
modeled in order to close Eq. 2.
Reynolds Stress Transport Equations. The Reynolds stress
model 17 involves calculation of the individual Reynolds
stresses, u
i
u
j
, using differential transport equations. The indi-
vidual Reynolds stresses are then used to obtain closure of the
Reynolds-averaged momentum Eq. 2. The transport equations
for the transport of the Reynolds stresses, u
i
u
j
, can be written
as follows:

x
k
u
k
u
i
u
j

C
ij
=

x
k

k
u
i
u
j

x
k
+

x
k


x
k
u
i
u
j

u
i
u
k

u
j
x
k
+ u
j
u
k

u
i
x
k

P
ij
+
ij

2
3

ij
3
The term on the left-hand side of Eq. 3 represents the convec-
tion, the terms on the right-hand side represent the turbulent dif-
fusion as proposed by Lien and Leschziner 18, molecular diffu-
sion, stress production, pressure strain, and the dissipation,
respectively. The pressure strain term
ij
is simplied according
to the proposal by Gibson and Launder 19:

ij
=
ij,1
+
ij,2
+
ij,w
4

ij,1
= C
1

k
u
i
u
j

2
3

ij
k 5

ij,2
= C
2
P
ij
C
ij

1
3

ij
P
kk
C
kk

6
Fig. 1 Schematic of the vortex chamber
Fig. 2 Inlet ow boundary condition
1378 / Vol. 128, NOVEMBER 2006 Transactions of the ASME

ij,w
= C
1

k
u
k
u
m
n
k
n
m

ij

3
2
u
i
u
k
n
j
n
k

3
2
u
i
u
k
n
i
n
k

k
3/2
C
l
d
+ C
2

km,2
n
k
n
m

ij

3
2

ik,2
n
j
n
k

3
2

jk,2
n
j
n
k

k
3/2
C
l
d
7
where C
1
=1.8, C
1
=0.5, C
2
=0.3, n
k
is the x
k
component of the
unit normal to the wall, d is the normal distance to the wall, and
C
l
=C

3/4
/ , where C

=0.09 and is the von Krmn constant


=0.4187. The scalar dissipation rate is computed with a model
transport equation similar to that used in the standard k model

x
i
u
i
=

x
j
+

t


x
j

C
1
1
2
P
ii

k
C
2

2
k
8
where
k
=1.0, C
1
=1.44, and C
2
=1.92 are constants taken from
Launder and Spalding 20. When the turbulence kinetic energy is
needed for modeling a specic term, it is obtained by taking the
trace of the Reynolds stress tensor
k =
1
2
u
i
u
i
9
The turbulent viscosity
t
is computed similarly to the k model

t
= C

k
2

10
where C

=0.09. Because of severe pressure gradients, the non-


equilibrium wall functions were used near wall as proposed by
Kim and Choudhury 21.
Inlet Conditions for the Reynolds Stresses. Whenever ow
enters the domain, the values for individual Reynolds stresses,
u
i
u
j
, and for the turbulence dissipation rate can be input di-
rectly or derived from the turbulence intensity and characteristic
length. The turbulence intensity I can be estimated from the fol-
lowing formula derived from an empirical correlation for pipe
ows
I =
u
V
avg
= 0.16Re
o

1/8
11
The turbulence length scale l is a physical quantity related to the
size of the large eddies that contain the energy in turbulent ows.
An approximate relationship between l and the physical size of
the vortex chamber diameter D
o
is
l = 0.07D
o
12
The relationship between the turbulent kinetic energy k and tur-
bulence intensity I is
k =
3
2
V
avg
I
2
13
where V
avg
is the average axial velocity. The turbulence dissipa-
tion rate can be determined as
= C

3/4
k
3/2
l
14
The values of the Reynolds stresses explicitly at the inlet are
given by
u
i
u
j
= 0 and u
i

2
=
2
3
k 15
4 Turbulence Modeling in Swirling Flows
The problem is considered to be an incompressible, steady, axi-
symmetric, and turbulent swirling ow. In this case, we can model
the ow in two-dimensional 2D i.e., solve the axisymmetric
swirl problem and incorporate the prediction of the swirl veloc-
ity; see Fig. 3.
The difculties associated with the solution of strongly swirling
ows can be attributed to high degree of coupling in the momen-
tum equations. High uid rotation gives rise to large radial pres-
sure gradient, which drives the ow in the meridional plane. This,
in turn, determines the distribution of the swirl in the eld. Nu-
merical instabilities that are attributed to momentum coupling re-
quire special solution techniques in order to obtain a converged
solution. Hence, a segregated, implicit solver, which is wellsuited
for the sharp pressure, and velocity gradients are more appropriate
for the ow under consideration. The mesh is sufciently must be
also sufciently rened in order to resolve the expected large ow
parameter gradients. The under-relaxation parameters on the ve-
locities were selected 0.30.5 for the radial and axial, and 0.9 for
the azimuthal velocity components.
There is a signicant amount of swirl in the chamber. The ap-
propriate choice depends on the strength of the swirl, which can
be gaged by the swirl number. To characterize the degree of swirl-
ing ow in a vortex chamber, a swirl number S is introduced.
Based on Gupta et al. 22 denition,
S =
G

G
z
Re
16
where G

is the axial ux of swirl momentum,


G

0
R
o
V
z
V

r
2
dr 17
G
z
is the axial ux of axial momentum,
G
z
=

0
R
o
V
z
2
rdr 18
To simplify the calculation of swirl number, the free-swirl velocity
prole and the average axial velocity are assumed inside the vor-
tex chamber.
V

=
V
in
R
o
r
, V
z
= V
avg
19
The ultimate form of the swirl number S can be determined as
S =
V
in
V
avg
20
Then, the simulations were performed for different swirl numbers
varied from S=12.520.
Grid Generation. A major challenge in calculating the ow
inside the vortex chamber is providing an adequate description of
the geometry. Because of the complex geometry of the vortex
generator, control over the grid is limited, making it difcult to
reduce the size of mesh without losing accuracy in the results.
Also, the grid size is limited by the computer memory available.
This leads to use axisymmetric problem, the formulation of 2D
grid generation are shown in Fig. 4. Triangular mesh elements and
an unstructured grid were used. A grid independent solution study
was made by performing the simulations for three different grids
consisting of 30,000, 43,000, and 50,000 nodes. The mean swirl
velocity for the three different grid sizes is shown in the Fig. 5.
Fig. 3 Computational domain
Journal of Fluids Engineering NOVEMBER 2006, Vol. 128 / 1379
Boundary Conditions. Boundary conditions have to be speci-
ed in order to solve the governing equations; see Fig. 3. At the
inlet, the values can be calculated from the given conditions at the
inlet, boundary; see Fig. 2. The total inlet velocity vector V
in
has
two components V
r,in
and V
,in
and they are related to each other
by:
V
,in
= V
in
cos , V
r,in
= V
in
sin , V
in
=
Q
in
A
in
21
However, at the outlet boundary there is no information about the
variables and some assumptions have to be made. The diffusion
uxes in the direction normal to the exit plane are assumed to be
zero. The pressure at the outlet boundary is calculated from the
assumption that radial velocity at the exit is neglected since it
does not have the space to develop, so that the pressure gradient
from r momentum is given by
p
r
=
V

2
r
22
At the solids walls, the no-slip condition was applied where the
velocities at the walls were specied to be zero. The centerline
boundary was considered axis of symmetry.
Discretization Scheme. The pressure-velocity coupling is
handled by using the SIMPLE-algorithm, the pressure staggering
option scheme was used for the pressure interpolation, the rst
order upwind schemes were used for momentum, swirl velocity,
turbulence kinetic energy, turbulence dissipation rate, and Rey-
nolds stresses. Convergence was assumed when the residual of the
equations dropped more than 3 orders of magnitude.
5 Results and Discussion
Pressure Drop Coefcient. Pressure drop or loss can be re-
garded as energy loss from the point of view of energy conserva-
tion. In the vortex chamber, pressure drop occurs mainly through
the dissipation of the swirl velocity as proposed by Jawarneh et al.
23. The pressure drop coefcient is dened as
C
p
=
2p
V
in
2
The estimated uncertainty for the pressure drop coefcient C
p

has appeared at the maximum of 9%. Figure 6 compares the


present experimental data to the RSM prediction of the pressure
drop coefcient C
p
for aspect ratio =3.0 and inlet angle
=30 deg. It is clear that as the diameter ratio increases, the
pressure coefcient C
p
increases. Stronger vortices will be pro-
duced by increasing the diameter ratio, resulting in a higher tan-
gential velocity and hence a higher pressure drop. It can be seen
that the Reynolds stress model gives good agreement with the
experimental data and the percentage difference error between the
predicted and experiments is 10%.
Mean Swirl Velocities Proles. The predicted and measured
radial proles of mean tangential, axial velocities and radial pres-
Fig. 4 Computational grid near the exit
Fig. 5 Grid independent solution study
Fig. 6 Pressure drop coefcient
1380 / Vol. 128, NOVEMBER 2006 Transactions of the ASME
sure for the chamber at station H=0.8L are shown in Fig. 7, 8, and
10, respectively. Figure 7 shows the mean swirl velocities proles
for the conguration =30 deg, =3.0, =0.131 at Reynolds
number Re
o
=11,592 and the predicted swirl velocity results are
compared with available experiments data LDA from Yan et al.
24 at three diameter ratios =2.5, 3.33, 4.0. It is shown the
ability of RSM to capture the free-vortex and forced-vortex re-
gions. Because of the intense swirl by increasing the swirl number
S, a high level of swirl momentum is transported around the
centreline and the consequence is the formation of the intense
swirling vortex along the center-line. In the latter gure, the peak
tangential velocity increases with increasing the diameter ratio or
the swirl number, and the location where the tangential velocity is
a maximum moves towards the vortex chamber center.
Mean Axial Velocity Prole. The mean axial velocity compo-
nent is able to develop prole ranging from jetlike to wakelike
shape as shown in Fig. 8. The predicted axial velocity results are
compared with available experiments data LDA from Escudier
et al. 1. The reverse ow backow in the vortex core is due to
the reduction of static pressure to values that are below the ambi-
ent and the stagnation point is appeared clearly in Fig. 8. Figure 9
shows the predicted axial velocities vectors near the chamber exit,
a ow-reversal region is found in the vortex core.
Mean Radial Pressure Proles. The following analysis illus-
trates the mean pressure distribution proles , and the pre-
dicted radial pressure will be compared to available experimental
data. The mean pressure distribution proles is dened ac-
cording to the following equation:
r =
2pr pr = 1
V
in
2
, where r =
r
R
o
Figure 10 compares the predicted radial pressure coefcient
to the experimental results 25,26 at three diameter ratios
=2.5, 3.33, 4.0. A good agreement with the experiments is ob-
served especially with high-diameter ratios i.e., =4.0, where
the ow eld is under strong swirling condition. Increasing the
swirl number S leads to deeper pressure proles.
6 Conclusions
Conned vortex ow inside the vortex chamber was investi-
gated both experimentally and computationally at different swirl
ratio. The RSM model is able to predict the ow features, such as
the press drop, tangential, axial velocities, and the radial pressure
proles. The prediction shows the behavior of the mean tangential
velocity distribution where a forced-vortex inside the core and a
free-vortex outside the core are existed and agree with the experi-
mental data. The reverse-ow, which is associated with the axial
velocity prole, is captured inside the core region and close to the
chamber exit. The swirl number has sufcient impact of the ow
features, the vortex core size contracts with increasing the swirl
number leads to more pressure drop energy loss, the peak tan-
gential velocity grows up, and deeper radial pressure proles. A
comparison of the results with measurement shows clearly the
superiority of the Reynolds-stress turbulence model in capturing
the major features of a conned, strongly swirling ow.
Fig. 7 Mean swirl velocity
Fig. 8 Dimensionless mean axial velocity
Fig. 9 Predicted axial velocity vectors near the exit
Journal of Fluids Engineering NOVEMBER 2006, Vol. 128 / 1381
Nomenclature
A
o
cross-sectional area of the vortex chamber
A
in
total inlet area
C
p
pressure coefcient 2p/ V
in
2

D
e
diameter of the exit port 2Re
D
in
diameter of the inlet port
D
o
chamber diameter 2R
o

k turbulent kinetic energy


L chamber length
p static pressure
p
a
ambient static pressure
p
in
static pressure at the inlet
Q
in
inlet volumetric ow rate
r, , z radial, tangential and axial coordinate
respectively
r normalized radius r/ R
o

R
e
radius of exit port
Re
o
Reynolds number Re
o
=4Q
in
/ D
o

R
o
radius of the chamber
S swirl number
u
i
, u
j
, u
k
velocity components in Cartesian coordinates
V

, V
z
mean tangential and axial velocity components
V
in
total velocity vector at the inlet
V
avg
average axial velocity
V
,in
inlet tangential velocity component
V
r,in
inlet radial velocity component
Greek Symbols
area ratio A
in
/ A
o

p static pressure difference P


in
P
a

radial pressure 2pr pr =1 / V


in
2

ij
Kronecker delta
turbulence dissipation rate
kinematics viscosity
dynamic viscosity

t
eddy or turbulent viscosity
density of the uid
inlet angle
aspect ratio L/ D
o

diameter ratio D
o
/ D
e

References
1 Escudier, M. P., Bornstein, J., and Zehender, N., 1980, Observations and
LDA Measurements of Conned Turbulent Vortex Flow, J. Fluid Mech., 98,
pp. 4963.
2 Vatistas, G. H., 1998, New Model for Intense Self-Similar Vortices, J. Pro-
pul. Power, 144, pp. 462469.
3 Sullivan, R. D., 1959, A Two-Cell Vortex Solution of the Navier-Stokes
Equations, J. Aerosp. Sci., 2611, pp. 767768.
4 Nallasamy, M., 1987, Turbulence Models and Their Applications to the Pre-
diction of Internal Flows, Comput. Fluids, 152, pp. 151194.
5 Nejad, A. S., Vanka, S. P., Favaloro, S. C., Samimy, M., and Langenfeld, C.,
1989, Application of Laser Velocimetry for Characterization of Conned
Swirling Flow, Trans. ASME: J. Eng. Gas Turbines Power, 111, pp. 3645.
6 Weber, R., 1990, Assessment of Turbulence Modeling for Engineering Pre-
diction of Swirling Vortices in the Near Burner Zone, Int. J. Heat Fluid Flow,
113, pp. 225235.
7 Launder, B. E., 1989, Second-Moment Closure and Its Use in Modelling
Turbulent Industrial Flows, Int. J. Numer. Methods Fluids, 98, pp. 963
985.
8 Leschziner, M. A., 1990, Modelling Engineering Flows With Reynolds Stress
Turbulence Closure, J. Wind. Eng. Ind. Aerodyn., 351, pp. 2147.
9 Ferziger, J. H., and Peric, M., 1999, Computational Methods for Fluid Dynam-
ics, 2nd ed., Springer-Verlag, Berlin.
10 Jones, L. N., Gaskell, P. H., Thompson, H. M., Gu, X. J., and Emerson, D. R.,
2005, Anisotropic, Isothermal, Turbulent Swirling Flow in a Complex Com-
bustor Geometry, Int. J. Numer. Methods Fluids, 47, pp. 10531059.
11 German, A. E., and Mahmud, T., 2005, Modelling of Non-Premixed Swirl
Burner Flows Using a Reynolds-Stress Turbulence Closure, Fuel, 845, pp.
583594.
12 Jakirlic, S., Hanjalic, K., and Tropea, C., 2002, Modeling Rotating and Swirl-
ing Turbulent Flows: A Perpetual Challenge, AIAA J., 4010, pp. 1984
1996.
13 Orland, P., and Fatica, M., 1997, Direct Simulation of Turbulent Flow in a
Pipe Rotating About Its Axis, J. Fluid Mech., 343, pp. 4372.
14 Jones, W. P., and Pascau, A., 1989, Calculation of Conned Swirling Flows
With a Second Momentum Closure, ASME Trans. J. Fluids Eng., 111, pp.
248255.
15 Hoekstra, A. J., Derksen, H. E. A., and Akker, V. D., 1999, An Experimental
and Numerical Study of Turbulent Swirling Flow in Gas Cyclones, Chem.
Eng. Sci., 54, pp. 20552065.
16 Vatistas, G. H., Lam, C., and Lin, S., 1989, Similarity Relationship for the
Core Radius and the Pressure Drop in Vortex Chambers, Can. J. Chem. Eng.,
67, pp. 540544.
17 Launder, B. E., Reece, G. J., and Rodi, W., 1975, Progress in the Develop-
ment of a Reynolds-Stress Turbulence Closure, J. Fluid Mech., 683, pp.
537566.
18 Lien, F. S., and Leschziner, M. A., 1994, Assessment of Turbulent Transport
Models Including Nonlinear RNG Eddy-Viscosity Formulation and Second-
Moment Closure, Comput. Fluids, 238, pp. 9831004.
19 Gibson, M., and Launder, B. E., 1978, Ground Effects on Pressure Fluctua-
tions in the Atmospheric Boundary Layer, J. Fluid Mech., 86, pp. 491511.
20 Launder, B. E., and Spalding, D. B., 1974, Numerical Computation of Tur-
bulent Flows, Comput. Methods Appl. Mech. Eng., 32, pp. 269289.
21 Kim, S. E., and Choudhury, D., 1995, A Near-Wall Treatment Using Wall
Functions Sensitized to Pressure Gradient, ASME FED, 217, Separated and
Complex Flows, pp. 273280.
22 Gupta, A. K., Lilly, D. G., and Syred, N., 1984, Swirl Flows, Abacus, Tun-
bridge Wells, England, UK.
23 Jawarneh, M. A., Vatistas, G. H., and Hong, H., 2005, On the Flow Devel-
opment in Jet-Driven Vortex Chambers, J. Propul. Power, 213, pp. 564
570.
24 Yan, L., Vatistas, G. H., and Lin, S., 2000, Experimental Studies on Turbu-
lence Kinetic Energy in Conned Vortex Flows, J. Therm. Sci., 91, pp.
1022.
25 Lam, H. C., 1993, An Experimental Investigation and Dimensional Analysis
of Conned Vortex Flows, Ph.D. thesis, Department of Mechanical and in-
dustrial Engineering, Concordia University, Montreal.
26 Alekseenko, S. V., Kuibin, P. A., Okulov, V. L., and Shtork, S. I., 1999,
Helical Vortices in Swirl Flow, J. Fluid Mech., 383, pp. 195243.
Fig. 10 Mean radial pressure
1382 / Vol. 128, NOVEMBER 2006 Transactions of the ASME

You might also like