Swirling Flow-Bluff Body - RANS - 24062019
Swirling Flow-Bluff Body - RANS - 24062019
Swirling Flow-Bluff Body - RANS - 24062019
in a Short Pipe
Ms Jinli Song
Mechanical Engineering Department, Khalifa University of Science and Technology, Abu Dhabi,
Email: [email protected]
Mechanical Engineering Department, Khalifa University of Science and Technology, Abu Dhabi,
Email: [email protected]
Email: [email protected]
Mechanical Engineering Department, Khalifa University of Science and Technology, Abu Dhabi,
Email: [email protected]
1
Abstract
Turbulent swirling flow inside a short pipe and interacting with a conical bluff body is simulated
using the commercial CFD code Fluent. The geometry used is akin to a liquid gas separators that
use swirl to initiate separation. Three turbulence models belonging to the Reynolds averaged
Navier-Stokes (RANS) equations frameworks are used, RNG k-ε, SST k-ω and the full Reynolds
stress model (RSM) in their steady and unsteady versions. The performance of the turbulence
models is assessed with particular focus on the RSM against experimental data and published
Large Eddy Simulation results using a full three-dimensional mesh of industrial scale. Steady and
unsteady RSM simulations show similar behavior and give comparatively the best predictions of
mean velocity profiles among the three RANS turbulence models but result in poor agreement on
The influence of Reynolds number on velocity profiles, swirl decay, and wall pressure on the bluff
body are also presented. In the case of Reynolds numbers that generate a combined Rankine
velocity profiles, the width and magnitude of flow reversal zone decreases along the pipe axis
direction and disappears downstream if the Reynolds number is small. The peak tangential
velocity increases with increasing Reynolds number. The swirl decay rate follows closely an
exponential form with decay rates that follow similar previously observed trends.
Key words: Swirling pipe flow; Conical bluff body; CFD; RANS
2
1. Introduction
Confined turbulent swirling flows are complex, highly three-dimensional, unsteady with strong
streamline curvature and turbulence anisotropy where the rotational component of the mean strain
rate plays an important. These flows take place in a broad range of industrial applications
including fluid phase separation in cyclone separators Rosa, et al. [1], heat transfer rates
enhancement in heat exchangers [2] and enhancing mixing in addition to flame stabilization in
combustion burners [3]. In view of this, confined turbulent swirling flows have received attention
and have been studied experimentally and numerically. However on the simulation side, the
calculation of such flows with turbulence modelling within the Reynolds averaged Navier-Stokes
(RANS) equations framework still remains a challenge. This work represents a contribution
towards assessing such models for a three-dimensional turbulent swirling pipe flow which
encounters a blockage of conical shape. This particular geometry possesses features of swirling
pipe flow and those that pertain to liquid-gas in-line separators [4] with a novel type of swirler
Confined turbulent swirling flows and their evolution in long pipes were studied by Kitoh [6];
Steenbergen [6]; Moene [7], Steenbergen and Voskamp [8] Pashtrapanka et al. [9] where swirl
was generated by stationary swirlers or rotating pipes. Studies of flows in the near field past
expansions include those of Dellenback et al. [10], Khezzar [11] and Mak and Balabani [12]. In
their study Pashtrapanka et al. [9] found that swirling flow in a pipe can loose its axial symmetry
as it progresses in the downstream direction and suggested that calculations for swirling pipe flow
should be done with three-dimensional model to capture the features of the flow.
3
Numerical simulations of turbulent swirling pipe flows have been conducted by several authors
using a variety of RANS turbulence models that ranged from the two-equation family to the full
Reynolds stress models. The encompassing study of Jakirlic et al. [13] used the standard k- model
and its low Reynolds number extension and three variants of the full Reynolds stress model (RSM),
the basic high Reynolds number version, the model of Speziale et al. [14] and a low Reynolds
number version of the RSM. This particular study permitted to highlight several important points:
(i) the standard k-e model fails to replicate important features of the flow resulting in solid body
rotation, this was also observed by several other authors previously such as Kobayachi and Yoda
[15] and Escue and Cui [16], (ii) The low-Reynolds number version of the RSM was superior since
it is able to mimic stress anisotropy in the near wall region, (iii) wall function approach is
unsuitable for strong swirling flows driven by near wall phenomena such as in rotating pipes or if
transition is present (iv) whereas if the extra strain rates responsible for non-equilibrium effects
arise from the inner part of the flow as in swirling flow entering pipes and if the bulk Reynolds
number is high the wall function concept can provide a reasonable alternative for modeling the
near wall region. Brennan [17] in his study of swirling flow in cyclonic separators found that the
RSM model with its linear and quadratic pressure strain correlations predict almost identical
velocity profiles. Erdal and Shirazi [18] simulated 3D swirling flow with different turbulent
models in a cylindrical cyclone. The simulations captured the general trend of the experimental
data. Simulation with standard k-ε model predicts higher rotational flow and RSM predicts a much
stronger decay of tangential and axial velocities. Wegner, et al. [3] evaluated the performance of
swirling flow with Precessing Vortex Core (PVC). The swirl number was 0.75 and Reynolds
number ranging from 10,000 and 42,000. The URANS results show good agreement of mean
4
velocities with experimental data. Ramirez and Cortes [19] simulated an unsteady single-phase
turbulent flow in a swirl combustor using the standard k–ε and Reynolds stresses models. Both
models predict a complex asymmetric flow with PVC and inner recirculation zone. The prediction
via RSM was more realistic, for which the flow lacks exact symmetry and periodicity, and exhibits
more stronger and persistent vortical motions. Large eddy simulation (LES) was used by Kharoua
et al. [4] for the same geometry used in the present study in the presence and absence of the bluff
cone. LES was able to predict the presence of the peak in the r.m.s of the axial and tangential
velocity components
Advanced turbulence models such as Large eddy simulation (LES) or direct numerical simulation
(DNS) can perform better than RANS model for such flows but remain prohibitively expensive
especially so for industrial type configurations which have usually flow passages of complex three-
industrial applications will be demanding in terms of meshing and the turn-around time for design
The present work involves the simulation of turbulent confined swirling flow inside a short pipe
which encounters a downstream blockage of conical shape. In contrast to many of the previous
numerical work published on swirling pipe flow, the geometry considered in the simulation is
three-dimensional so that it will be able to capture the inherent features of the flow. In order to
obtain meaningful simulations, the complex three-dimensional swirler geometry is also included
in the simulation domain. The work is an extension of Kharoua et al. [4] and aims to evaluate the
performance of mainly the RSM turbulence model with linear pressure strain correlation and non-
5
equilibrium wall function which takes into account the influence of the pressure gradients on the
distortion of the velocity distributions. While the main focus is on the RSM model, other RANS
two-equation turbulence models are also used for comparison. Thus, the RNG k-ε with its options
of eddy viscosity and swirl dominated flow and enhanced wall treatment with pressure gradient
effects and SST k-ω with low-Re and curvature corrections are used in this study. While similar
studies using versions of the k- and RSM are numerous, this is not the case for the SST k- model.
Numerical results were benchmarked with experimental data obtained using laser Doppler
Anemometry. Since the geometry is extracted from that of a two-phase separator additional
parametric simulations are conducted to study the influence of Reynolds number on flow and swirl
decay. This is meant to be a first step to evaluate the ability of CFD to mimic the flow features of
single phase swirling flow which interacts with a bluff body before extending the work later on to
two-phase flow. It is to be noted that numerical studies on confined turbulent swirling flows
interacting with a bluff body are relatively rare in the literature [20, 21]. So this study is a
contribution towards understanding of their features. when the shapes of bluff body are the disks
Section 2 of this report presents the simulation approach used, while the results are discussed in
The flow is assumed incompressible, turbulent and three-dimensional, under steady and unsteady
6
2.1. Reynolds averaged equations and turbulence models
The flow is governed by the Reynolds-averaged Navier-Stokes equations. The continuity and
𝜕
(𝜌𝑢𝑖 ) =0 (1)
𝜕𝑥𝑖
𝜕 𝜕 𝜕𝑝 𝜕 𝜕𝑢 𝜕𝑢 𝜕
(𝜌𝑢𝑖 ) + ̅̅̅̅̅̅̅
(𝜌𝑢𝑖 𝑢𝑗 ) = − 𝜕𝑥 + 𝜕𝑥 [𝜇 (𝜕𝑥 𝑖 + 𝜕𝑥𝑗 )] + 𝜕𝑥 (−𝜌𝑢 𝑖 ′𝑢𝑗 ′) (2)
𝜕𝑡 𝜕𝑥𝑗 𝑖 𝑗 𝑗 𝑖 𝑗
̅̅̅̅̅̅̅
The Reynolds stresses −𝜌𝑢 𝑖 ′𝑢𝑗 ′ require a turbulent model in order to close the above system of
equations.
The Reynolds Stress Model (RSM), solves the transport equations for each term of the Reynolds
stress tensor, it is recognized to give the best performance among RANS models for highly
swirling flow. The exact transport equations of the Reynolds stresses, ̅̅̅̅̅̅̅
𝜌𝑢𝑖′ 𝑢𝑗′ , may be written as
follows:
𝜕 ̅̅̅̅̅̅̅ 𝜕
(𝜌𝑢𝑖′ 𝑢𝑗′ ) + 𝜕𝑥 (𝜌𝑢𝑘 ̅̅̅̅̅̅
𝑢𝑖′ 𝑢𝑗′ )
𝜕 ̅̅̅̅̅̅̅̅̅
= − 𝜕𝑥 [𝜌𝑢 ′ ′ ′ ̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
′ ′ ′ 𝜕 𝜕 ̅̅̅̅̅̅
′ ′
𝜕𝑡 𝑖 𝑢𝑗 𝑢𝑘 + 𝑃 (𝛿𝑘𝑗 𝑢𝑖 + 𝛿𝑖𝑘 𝑢𝑗 )] + 𝜕𝑥 [𝜇 𝜕𝑥 (𝑢𝑖 𝑢𝑗 )] −
𝑘 𝑘 𝑘 𝑘
The term on the left side of the above equation represents the local time derivative and convective
term, and the terms on the right hand side represent the turbulent diffusion, molecular diffusion,
stress production, pressure strain and the dissipation, respectively. The stress transport equations
solved are naturally the modelled equations. In this study, we use Linear Pressure-Strain Model.
The pressure-strain term, ∅𝑖𝑗 , is modeled according to the proposals by Fu, et al. [22]; Gibson and
7
2.1.2. RNG k-ε and SST k-ω Model
The two-equation turbulence models (k-ε and k-ω models) are based on the Boussinesq
𝜕𝑢𝑖 𝜕𝑢 2
̅̅̅̅̅̅̅
−𝜌𝑢 𝑖 ′𝑢𝑗 ′ = 𝜇𝑡 ( + 𝑗) − (𝜌𝑘)𝛿𝑖𝑗 (5)
𝜕𝑥𝑗 𝜕𝑥𝑖 3
This method requires low computational cost but it based on the assumption of an isotropic
turbulent viscosity 𝜇𝑡 .
In the RNG k-ε model a further two modelled transport equations for the turbulence kinetic energy
k and its rate of dissipation are solved. The model equations were derived using a statistical
technique called renormalization group theory [26]. Compared to the standard k-ε model, it has
an additional term in the dissipation rate ε equation that improves the accuracy for rapidly strained
flows and a modified turbulent viscosity to account for the influence of swirl.
The Shear-Stress Transport SST k-ω model Menter [27] accounts for the transport of the
turbulence shear stress in the definition of the turbulent viscosity. It is more accurate and reliable
for a wider class of flows (for example, adverse pressure gradient flows, airfoils, transonic shock
waves) than the standard k-ω model. Further details can be found in fluent documentation [26].
The computational geometry used in this study is relevant to the one existing in a liquid-gas
separator, a Perspex pilot model of which is shown in figure 1. The separator uses swirling flow
to separate the phases. As mentioned above the present simulation is focused on single phase
liquid flow to assess the performance of RANS turbulence models in this geometry and flow and
subsequently extend it to two-phase flow modelling. The geometry defining the flow domain is
8
presented in figure 2 with the origin of the system of coordinates adopted. It consists, see figure
2a, of a short inlet pipe of inner diameter 41 mm connected to a short cylindrical housing of length
94 mm and diameter 150 mm which houses the swirler. The swirler shown in figure 2b is made
of a stainless steel cylinder concentric to the housing with external and inner diameters of 76.2 mm
and 39 mm respectively with sixty 4 mm diameter radial holes. The swirler is axially connected
to a 610 mm long outlet pipe of 41 mm in diameter. The straight entry inflow is converted into a
swirling flow through the inclined holes inside the swirl cage as shown in figure 2a. The swirling
flow develops along the outlet pipe. At the end of the pipe a conical bluff body the details of which
are given on figure 1c. The bluff body is attached to a circular flange which contains three kidney
shaped holes through which the fluids escapes in the axial direction.
The hexahedral grids were generated using ANSYS Workbench. Different computational grids
were used as illustrated in table 1. Attention is directed towards the resulting swirling flow
which takes place between the swirler outlet and the tip of the bluff body.
Cone
Swirling flow
9
(a)
(b)
(c)
Figure 2: (a)- geometry of flow domain, (b)- Swirler, (c)- Bluff body
10
Mesh #1 #2 #3
Wall y+ ≤ 21 9 -18 ≤ 11
An inlet velocity condition was used while a constant pressure (zero gauge pressure) was
prescribed at the outlet. A no-slip condition was taken for the remaining boundaries.
The commercial software Fluent 17.1, based on the finite volume method, was used in this study
and calculations were executed under steady conditions for all turbulence models and in addition
unsteady calculations were performed with the STT k- and RSM models to investigate the effect
of steady assumption. Water was the working fluid under atmospheric pressure and ambient
temperature.
The pressure-velocity coupling is achieved by using the SIMPLE scheme. For SST k-ω and RNG
k-ε models, the pressure was discretized using second order upwind scheme while the RSM model
simulation made use of PRESTO! Scheme. The discretization of the equations for momentum,
turbulent kinetic energy, dissipation rate and Reynolds stresses is achieved by second order upwind
scheme.
During the simulation for the unsteady case, the mean axial and tangential velocity and static
pressure from two points at different axial positions within the outlet pipe are used for monitoring
purposes. Convergence is assumed when the monitoring data achieve a statistically steady state.
The transient simulation is conducted in three steps. First, a steady solution with same settings is
launched to generate a good initial instantaneous field for the transient simulation. After that, the
11
transient simulation was run to allow the instantaneous field to develop with a time step equal to
4×10-4 s. Finally, the time-averaged field was monitored at the abovementioned points until a
statistical convergence is reached. It is important to mention that the residence time of the flow,
based on the axial bulk velocity of 1 m/s and the length of the computational domain, was 0.81 s.
For stability purposes, the Courant Number was kept smaller than unity. The time integration was
based on the second order implicit scheme. After 3 s of physical flow time, data collection, was
The base case flow corresponds to a Reynolds number based on the bulk velocity and pipe diameter
of 41,000. The test fluid was water at room temperature. Grid independency tests are considered
this study is conducted primarily against experimental data of Zhang [28] and accessorily with
LES computational results of Kharoua et al. [4] which, it has to be stressed, were obtained with a
much larger grid and involved considerably more usage of computational time. The quantitative
comparison uses the mean axial and tangential velocity components, their corresponding normal
turbulent single point correlations and the cross correlation between the axial and tangential
components of the fluctuating velocities and the swirl number axial evolution. The third
subsection examines the effect of varying the Reynolds number on the mean flow quantities such
Grid independency tests were conducted based on the use of the RSM model since this is the model
at the center of this study and the pipe Reynolds number value of 41,000 and are also gaged with
12
experimental results. Three meshes of 3,631,633, 6,441,292 and 11,236,287 elements were used.
Figure 3 compares the mean axial and tangential velocity profiles for the three meshes at an axial
location x/D=10. It can be seen that the differences between mesh 2 and 3 are not large while the
results for mesh 1 are quite different and fail to a large extent to replicate certain important flow
features such as the central reverse flow region, and the Rankine vortex flow structure. The
estimated wall y+ values inside the outlet pipe and around the bluff body surface for the different
meshes are illustrated in table 1. As a compromise between accuracy and computational cost,
Mesh #2 is used for the simulation of the cases considered in this work and it is inherently assumed
that the use of the RSM as the main model used for grid convergence test will not invalidate this
conclusion when the other models are used. It is worth mentioning here that the mesh attempts to
capture all the details of the three-dimensional geometry. It therefore includes the swirler with its
intricate flow passages in the solution domain and the resulting three-dimensional structure of the
flow and hence tries to conduct a simulation and an evaluation at industrial scale under RANS
framework of modelling.
2
Mesh #1
Mesh #2
1,5
Mesh #3
Experiment
1
U/Um
0,5
-0,5
x / D = 10
-1
-1 -0,5 0 0,5 1
r/R
13
2,5
Mesh #1
2 Mesh #2
1,5 Mesh #3
1 Experiment
0,5
W/Um
0
-0,5
-1
-1,5
-2 x / D = 10
-2,5
-1 -0,5 0 0,5 1
r/R
Figure 3: Radial profiles of mean axial and tangential velocity for different meshes at x/ D = 10.
The radial profiles of the mean axial and tangential velocity, at three representative axial positions,
are compared with experimental and LES data as shown in Figure 4. The simulation using RNG
k-ε model is performed in a steady state while the simulations using other models under both
steady and unsteady states. The LES profiles of reference [4] are considered for comparison.
14
2
SST k-ω steady
SST k-ω unsteady
1,5 RNG k-ε
RSM steady
RSM unsteady
1 LES
Experiment
0,5
U/Um
-0,5
-1
x/D=5
-1,5
-1 -0,5 0 0,5 1
r/R
2
SST k-ω steady
SST k-ω unsteady
1,5 RNG k-ε
RSM steady
RSM unsteady
1 LES
Experiment
0,5
U/Um
-0,5
-1
x / D = 10
-1,5
-1 -0,5 0 0,5 1
r/R
15
2
SST k-ω steady
SST k-ω unsteady
RNG k-ε
1,5 RSM steady
RSM unsteady
LES
1 Experiment
0,5
U/Um
-0,5
x / D = 14
-1
-1 -0,5 0 0,5 1
r/R
2,5
SST k-ω steady
2 SST k-ω unsteady
RNG k-ε
1,5 RSM steady
1 RSM unsteady
LES
0,5 Experiment
0
W/Um
-0,5
-1
-1,5
-2
x/D=5
-2,5
-1 -0,5 0 0,5 1
r/R
16
2,5
SST k-ω steady
2 SST k-ω unsteady
RNG k-ε
1,5 RSM steady
RSM unsteady
1
LES
W/Um 0,5 Experiment
0
-0,5
-1
-1,5
-2
x / D = 10
-2,5
-1 -0,5 0 0,5 1
r/R
2,5
SST k-ω steady
2 SST k-ω unsteady
RNG k-ε
1,5 RSM steady
RSM unsteady
1
LES
0,5 Experiment
0
W/Um
-0,5
-1
-1,5
-2
x / D = 14
-2,5
-1 -0,5 0 0,5 1
r/R
Steady and unsteady RSM simulations predict similar results suggesting that a steady RSM
simulation is sufficient to avoid a costly unsteady RSM simulation if not specifically needed. The
mean axial velocity profiles show that the central core flow reversal region, which plays an
important role in a phase separation, is predicted well with RSM and persists from the exit of the
17
swirl cage till the bluff body. The SST k- and RNG k- under steady conditions fail to capture
the mean axial velocity shape and loose symmetry starting at x/D=4 of the mean axial velocity
profile shape. The reason behind this remains unclear. The mean axial velocity remains under-
predicted by RSM in the annular region near the wall, by as much as 50%.
The profiles of the mean tangential velocity correspond to a Rankine vortex. As the flow proceeds
downstream, the mean axial and tangential velocities decay due to viscous dissipation and the peak
tangential velocity shifts towards the pipe axis which is consistent with previous findings
Ahmadvand, et al. [30]. The RNG k-ε and steady SST k-ω simulations exhibit a strong asymmetry
in the profile of the velocity at x/D= 10 and 14. Similar to the axial velocity the results of unsteady
SST k-ω simulation show great improvement compared with steady SST k-ω simulation. All the
RANS and URANS simulations fail to capture the mean tangential velocity peaks in strong
contrast with LES simulation results of [4] which were much closer to the experimental ones. In
particular all RSM calculations display a lower solid body rotation than the experiments inside the
core and hence suggest that the calculated swirl intensity is decaying faster that the experimental
one.
In swirling pipe flow, the profiles of the normal turbulent stresses differ from those inside uni-
directional conventional pipe flow by the fact that they exhibit larger values, peaks at the center
and are highly anisotropic. For this reason, the radial profiles of the axial and tangential Reynolds
stress components at three representative axial positions are shown in figure 5 and comparison is
limited to RSM, experiments and LES of reference [4]. The RSM model fails to capture the peak
along the axis of the pipe, in the core region. Similar behavior was also obtained in the
computational work of highly swirling flow in combustors investigated by Leschziner and Hogg
[30]. They concluded that the inadequate modeling of pressure strain, especially its rapid part, is
18
responsible for the inaccurate prediction of Reynolds stresses. It is well known that the model
coefficient could be adjusted to improve the results but the approach remains customized and not
universal. The LES simulation predicts the turbulence, at the core, in a better way although still
exhibiting important discrepancies suggesting that model and mesh-resolution tunings are required
in the core region due to this complex flow behavior. Outside the core region the levels of the
normal stresses remain in qualitative agreement. The same can be said for the shear stress
0,5
x/D=5 Experiment
0,45
LES
0,4
RSM unsteady
0,35
RSM steady
0,3
0,25
uu
0,2
0,15
0,1
0,05
0
-1 -0,5 0 0,5 1
r/R
0,5
x / D = 10 Experiment
0,45
LES
0,4
RSM unsteady
0,35
RSM steady
0,3
0,25
uu
0,2
0,15
0,1
0,05
0
-1 -0,5 0 0,5 1
r/R
19
0,5
x / D = 14 Experiment
0,45 LES
0,4 RSM unsteady
0,35 RSM steady
0,3
0,25
uu
0,2
0,15
0,1
0,05
0
-1 -0,5 0 0,5 1
r/R
0,4
x/D=5 Experiment
0,35 LES
RSM unsteady
0,3
RSM steady
0,25
0,2
ww
0,15
0,1
0,05
0
-1 -0,5 0 0,5 1
r/R
20
0,4
x / D = 10 Experiment
0,35 LES
RSM unsteady
0,3
RSM steady
0,25
0,2
ww
0,15
0,1
0,05
0
-1 -0,5 0 0,5 1
r/R
0,4
x / D = 14 Experiment
0,35 LES
RSM unsteady
0,3
RSM steady
0,25
0,2
ww
0,15
0,1
0,05
0
-1 -0,5 0 0,5 1
r/R
21
0,3
x/D=5 Experiment
LES
0,25
RSM unsteady
uw 0,2 RSM steady
0,15
0,1
0,05
0
-1 -0,5 0 0,5 1
r/R
0,3
x / D = 10 Experiment
LES
0,25
RSM unsteady
0,2 RSM steady
0,15
uw
0,1
0,05
0
-1 -0,5 0 0,5 1
r/R
22
0,3
x / D = 14 Experiment
LES
0,25
RSM unsteady
0,2 RSM steady
0,15
uw
0,1
0,05
0
-1 -0,5 0 0,5 1
r/R
The swirl number is a non-dimensional number that is used to characterize intensity of swirling
flows, which for a pipe of radius R can be expressed as, see Kitoh [5]:
R
2 r 2UWdr
S 0 (6)
R 3U b2
where 𝑈𝑏 is the bulk velocity inside the pipe. To calculate S a numerical integration is necessary
to evaluate the numerator and it is assumed that the velocities vary linearly between the closest
measured velocity to the wall and the zero value at the wall itself.
The swirl number variation along the length of the pipe is shown for the RSM model and
experimental data in figure 6. Results from the two-equation turbulence models are not shown
since they failed to capture accurately the variation of the fluid velocity along the pipe at least with
the present mesh used. The RSM variation shows a decay of swirl with downstream distance and
rate of decay almost equal to the experimental one but lower by about 10%. This is due to the fact
23
that swirl intensity was under predicted by RSM as shown with the mean tangential velocity.
However, the rate of decay predicted by the RSM is similar to the experimental one. The swirl
intensity itself will depend on accurate prediction of the tangential mean velocity profile which
itself will depend on accurate prediction of the flow inside the intricate swirler passages which are
directly responsible for swirl generation and this could be the source of the discrepancy.
1,6
1,5
1,4
Swirl number
1,3
1,2
1,1 Experiment
1 LES
RSM steady
0,9
RSM unsteady
0,8
0 3 6 9 12 15
x/D
Numerical simulations were conducted for a range of Reynolds numbers from 4,100 to 82,000 to
examine its effect on the velocity distribution, swirl number decay and pressure on the conical
Figure 7 depicts the mean axial and tangential velocity profiles for the cases with different inlet
Reynolds number at three representative axial positions between the exit from the swirler and
the tip of the conical bluff body. The numerical results and discussion are confined the the RSM
24
simulations. Each velocity profile is normalized by the bulk axial velocity for that profile. For
Re=4,100, the reversal flow disappears at x/D=6 and the size of forced vortex is smallest among
all the cases. The flow then develops with a positive axil velocity similar to unidirectional pipe
flow. This type of flow indicates possibly vortex breakdown. For Re=20,500, the width and
magnitude of reversal flow zone decrease along the pipe axis direction and disappear at x / D =
14 just ahead of the tip of the bluff body. The velocity profiles show slight deviation for
Re=41,000 and 82,000. The maximum reversal flow region is obtained when Re=41,000. The
peak tangential velocity in the annular region increases with Reynolds number. It is to be noted
that for Re=4,100 after the disappearance of the core reverse flow region, the flow is trying to
adjust to solid body rotation distribution, while the positive axial velocity increases rapidly in the
core of the flow. Parchen and Steenbergen [31] have demonstrated how the axial velocity
distribution is swirling pipe flow can be influenced by the swirl generator geometry so that in
some cases the central reverse flow region can be eliminated and in the work of Pashtrapanska et
1,5
0,5
U/Um
Re=4,100
-0,5 Re=20,500
Re=41,000
Re=82,000 x/D=5
-1
-1 -0,5 0 0,5 1
r/R
25
1,5
0,5
U/Um
0
Re=4,100
-0,5 Re=20,500
Re=41,000
Re=82,000 x / D = 10
-1
-1 -0,5 0 0,5 1
r/R
1,5
0,5
U/Um
Re=4,100
-0,5 Re=20,500
Re=41,000
x / D = 14
Re=82,000
-1
-1 -0,5 0 0,5 1
r/R
26
2,5
Re=4,100
2
Re=20,500
1,5
Re=41,000
1
Re=82,000
0,5
W/Um
0
-0,5
-1
-1,5
-2 x/D=5
-2,5
-1 -0,5 0 0,5 1
r/R
2,5
Re=4,100
2
Re=20,500
1,5
Re=41,000
1
Re=82,000
0,5
W/Um
0
-0,5
-1
-1,5
-2
x / D = 10
-2,5
-1 -0,5 0 0,5 1
r/R
27
2,5
Re=4,100
2
Re=20,500
1,5
Re=41,000
1
Re=82,000
0,5
W/Um
0
-0,5
-1
-1,5
-2
x / D = 14
-2,5
-1 -0,5 0 0,5 1
r/R
Figure 7: Radial profiles of mean axial and tangential velocity for different Reynolds Numbers.
Swirl decay prediction for swirling pipe flow is relevant to several engineering applications
(Steenbergen and Voskamp [8] including flow inside two-phase separation equipment, it is
therefore important to consider the influence of the Reynolds number on it. At this point it is fit
to point out that several researchers have obtained an expression for the variation of swirl number
with axial distance in a pipe Steebergen and Voskamp [8] and Najafi et al. [32]. Thus by
integrating the momentum equation in cylindrical coordinates and assuming slowly developing
flow along the axial direction, neglecting turbulent shear stress and assuming relatively low swirl
𝑆 (𝑥−𝑥𝑟 )
= 𝑒 −𝛽 𝐷 (7)
𝑆𝑟
28
1,2
0,8 =0.027
=0.033
S/Sr
0,6
Re=4,100
=0.058
0,4 Re=20,500
Re=41,000
0,2 Re=82,000
Experiments =0.184
0
0 5 10 15
x/D
The present discussion will be limited to cases when Re>4,100 where the flow exhibits a tangential
velocity profile of the Rankine compound type. The calculated values of were 0.058, 0.033 and
0.027 for Reynolds number values of 20,500, 41,000 and 82,000 respectively. These values are
in agreement with the findings of Najafi et al. [32], so that as decreases almost linearly with
Reynolds number. In addition, Steenbergen and Voskamp [8] have generated a comprehensive
graph encapsulating the variation of the decay rate of swirl with Reynolds number for a number
of previous experimental studies performed during the 1960’s till 1998. The results cover a wide
range of swirl generation methods and swirl intensities. The graph confirms the exponential nature
of the decay rate and despite a relatively wide scatter depending on the swirl number ranges and
exact experimental conditions remains however a very useful reference for smooth pipes. The
present results agree very well with the values reported in that graph. Thus the value of for
Reynolds of 20,500, 41,000 and 82,000 agree very well with the findings of Nissan and Bresan
29
[33], Bake [34] and Kitoh [5] respectively. The value of b for Reynolds of 41,100 of 0,033
compares favourably with the experimental value of Steenbergen and Voskamp [8] for a near
This information is crucial when considering the optimum axial position of the bluff body inside
Figure 9 depicts the pressure distributions on the bluff body surface for different inlet Reynolds
numbers. It illustrates the effect of the swirling flow, with different swirl intensities (shown in
figure 8), on the loads experienced by the bluff body. For the case with Re=4,100, the pressure
peak is observed on the bluff body apex. It means the stagnation point is at the cone apex for
low swirl since the flow for this Reynolds number is flowing the downstream direction without
flow reversal on the axis. Increasing the Reynolds number (swirl intensity) causes flow reversal
on the axis and the apex to become gradually a region with minimum pressure and the first
cylindrical surface after the cone to become the maximum pressure region. The implication for
flow separation is that if the Reynolds number is low enough, the central recirculation zone does
not extend to the tip of the cone and hence the swirl strength is reduced which would adversely
affect separation. Thus, it is very important to consider swirling flow characteristics when
investigating the optimal design of the bluff body in terms of size, shape and location for a more
efficient separation. It is worth mentioning that the hollow cone would undergo a negative drag
force, for high swirl numbers, tending to dismantle it from its supporting hollow tail pipe.
30
Re=4100
Re=20,500
Re=41,000
31
Re=82,000
4. Conclusions
A numerical study on turbulent swirling flow interacting with a conical bluff body inside a short
pipe was conducted. The simulations in contrast to previous ones on swirling pipe flows which
were done using two-dimensional grids, include the full complex three-dimensional geometry of
the swirler in the solution domain. The following conclusions can be made:
The RSM turbulence model presents the best performance in contrast to the two-equation
turbulence models such as RNG k-e and k-w. The RNG k-e and SST k-w in steady mode fail to
predict important features of the flow and can be considered as the worst of the models. The
SST k-w model in unsteady mode performs better than the first two but remains inferior to the
RSM.
The RSM in both steady and unsteady modes performed equally well and provided the best
performance in the sense of being able to capture the variation of the mean axial and tangential
32
The two-equation models and the steady RSM are not able to capture the correct variation of the
Reynolds stresses. Especially the peak in the central core zone. Only the unsteady RSM is able
The effect of the Reynolds number on the flow calculated with the unsteady RSM indicated that
for Reynolds greater than 4,100, the flow is of the Rankine compound type and that the rate of
decay of swirl is exponential. The rate of decay was found to decrease almost linearly with
increase in Reynolds number. The obtained values of the rate of decay are in agreement with
5. Acknowledgement
The authors are grateful to ADNOC Onshore Company. (ADCO) for the financial support of
this research project and for Khalifa University of Science and Technology for granting Jinli
6. References
[1] E. S. Rosa, F. A. França, and G. S. Ribeiro, "The cyclone gas–liquid separator: operation
and mechanistic modeling," Journal of Petroleum Science and Engineering, vol. 32, pp. 87-101,
2001.
[2] S. Martemianov and V. L. Okulov, "On heat transfer enhancement in swirl pipe flows,"
International Journal of Heat and Mass Transfer, vol. 47, pp. 2379-2393, 2004.
[3] B. Wegner, A. Maltsev, C. Schneider, A. Sadiki, A. Dreizler, and J. Janicka, "Assessment
of unsteady RANS in predicting swirl flow instability based on LES and experiments,"
International Journal of Heat and Fluid Flow, vol. 25, pp. 528-536, 2004.
[4] N. Kharoua, L. Khezzar, and M. Alshehhi, "The interaction of confined swirling flow
with a conical bluff body: Numerical simulation," Chemical Engineering Research and Design,
vol. 136, pp. 207-218, 2018.
[5] O. Kitoh, "Experimental study of turbulent swirling flow in a straight pipe," Journal of
Fluid Mechanics, vol. 225, pp. 445-479, 1991.
33
[6] W. Steenbergen, Turbulent pipe flow with swirl: Citeseer, 1995.
[7] A. F. Moene, Swirling pipe flow with axial strain: experiment and large eddy simulation:
Technische Universiteit Eindhoven, 2003.
[8] W. Steenbergen and J. Voskamp, "The rate of decay of swirl in turbulent pipe flow,"
Flow Measurement and Instrumentation, vol. 9, pp. 67-78, 1998.
[9] M. Pashtrapanska, J. Jovanović, H. Lienhart, and F. Durst, “Turbulence measurements in
a swirling pipe flow”,. Exp. Fluids, 41, 813-827, 2006.
[10] P. A. Dellenback, D. E. Metzger, D.E. and G. Neitzel," Measurements in turbulent swirling
flow through an abrupt axisymmetric expansion," AIAA J., 26(6), pp. 669-681, 1988.
[11] L. Khezzar, "Velocity measurements in the near field of a radial swirler," Exp. Therm.
Fluid Sci. , 16(3), pp. 230-236, 1998.
[12] H. Mak, and S. Balabani, "Near field characteristics of swirling flow past a sudden
expansion," Chem. Eng. Sci., 62(23), pp. 6726-6746, 2007.
[13] S. Jakirlic, K. Hanjalic, and C. Tropea, “Modeling rotating and swirling turbulent flows:
A perpetual challenge”, AIAA J., 40, 10, pp. 1984-1996, 2002.
[14] C. G. Speziale, S. Sarkar, and T. B. Gatski, “Modelling the pressure-strain correlation of
turbulence: An invaraint dynamical systems approach”, Journal of Fluid Mechanics, 227, pp.
245-272, 1991.
[15] T. Kobayachi and M. Yoda, “Modified k-e model for turbulent swirling flow in a pipe”,
JSME International Journal, 30, 259, pp. 66-71, 1987.
[16] A. Escue and J. Cui, "Comparison of turbulence models in simulating swirling pipe
flows," Applied Mathematical Modelling, vol. 34, pp. 2840-2849, 2010.
[17] M. Brennan, "CFD Simulations of Hydrocyclones with an Air Core: Comparison
Between Large Eddy Simulations and a Second Moment Closure," Chemical Engineering
Research and Design, vol. 84, pp. 495-505, 2006.
[18] F. M. Erdal and S. A. Shirazi, "Local velocity measurements and computational fluid
dynamics (CFD) simulations of swirling flow in a cylindrical cyclone separator," Journal of
Energy Resources Technology, vol. 126, pp. 326-333, 2004.
[19] J. A. Ramirez and C. Cortes, "Comparison of different URANS schemes for the
simulation of complex swirling flows," Numerical Heat Transfer, Part B: Fundamentals, vol. 58,
pp. 98-120, 2010.
[20] R. Huang and F. Tsai, "Observations of swirling flows behind circular disks," AIAA
journal, vol. 39, pp. 1106-1112, 2001.
[21] K. Atvars, M. Thompson, and K. Hourigan, "Modification of the flow structures in a
swirling jet," in IUTAM Symposium on Unsteady Separated Flows and their Control, pp. 243-
253, 2009.
[22] S. Fu, B. Launder, and M. Leschziner, "Modelling strongly swirling recirculating jet flow
with Reynolds-stress transport closures," in 6th Symposium on Turbulent Shear Flows, pp. 17-6,
1987.
[23] M. Gibson and B. Launder, "Ground effects on pressure fluctuations in the atmospheric
boundary layer," Journal of Fluid Mechanics, vol. 86, pp. 491-511, 1978.
[24] B. E. Launder, "Second‐moment closure and its use in modelling turbulent industrial
flows," International Journal for Numerical Methods in Fluids, vol. 9, pp. 963-985, 1989.
[25] B. E. Launder, "Second-moment closure: present… and future?," International Journal of
Heat and Fluid Flow, vol. 10, pp. 282-300, 1989.
[26] A. Fluent, "Theory Guide 17.2," Ansys Inc. USA, 2016.
34
[27] F. R. Menter, "Two-equation eddy-viscosity turbulence models for engineering
applications," AIAA journal, vol. 32, pp. 1598-1605, 1994.
[28] T. Zhang, " Experimental investigation of swirling flow interaction with a bluff body,"
Master, Khalifa University of Science and Technology, Abu Dhabi, UAE, 2018.
[29] M. Ahmadvand, A. Najafi, and S. Shahidinejad, "An experimental study and CFD
analysis towards heat transfer and fluid flow characteristics of decaying swirl pipe flow
generated by axial vanes," Meccanica, vol. 45, pp. 111-129, 2010.
[30] M. Leschziner and S. Hogg, "Computation of highly swirling confined flow with a
Reynolds stress turbulence model," AIAA journal, vol. 27, pp. 57-63, 1989.
[31] R. R. Parchen & W. Steenbergen, “An experiemnatl and numerical study of turbulent
swirling pipe flows”, Journal of Fluis Engineering, 120, 54-61, 1998.
[32] A. F. Najafi, S. M. Mousavian, and K. Amini, "Numerical investigations on swirl
intensity decay rate for turbulent swirling flow in a fixed pipe," International Journal of
Mechanical Sciences, vol. 53, pp. 801-811, 2011.
[33] A. H. Nissan, & V. P. Bresan, “Swirling flow in cylindres”, A.I.Ch.E. Journal, 7, 4, 543-
547, 1961.
[34] D. W. Baker, “Decay of swirling turbulent flow of incompressible fluids in long pipes”,
PhD thesis, University of Maryland, 1967.
35