Вы находитесь на странице: 1из 15

XML Template (2013)

K:/PIC/PIC 474211.3d

[16.1.20132:14pm]
(PIC)

[115]
[PREPRINTER stage]

Original Article

Large eddy simulation study of turbulent


flow around smooth and rough domes

Proc IMechE Part C:


J Mechanical Engineering Science
0(0) 115
! IMechE 2013
Reprints and permissions:
sagepub.co.uk/journalsPermissions.nav
DOI: 10.1177/0954406212474211
pic.sagepub.com

N Kharoua and L Khezzar


Date received: 4 September 2012; accepted: 17 December 2012

Abstract
Large eddy simulation of turbulent flow around smooth and rough hemispherical domes was conducted. The roughness
of the rough dome was generated by a special approach using quadrilateral solid blocks placed alternately on the dome
surface. It was shown that this approach is capable of generating the roughness effect with a relative success. The
subgrid-scale model based on the transport of the subgrid turbulent kinetic energy was used to account for the small
scales effect not resolved by large eddy simulation. The turbulent flow was simulated at a subcritical Reynolds number
based on the approach free stream velocity, air properties, and dome diameter of 1.4  105. Profiles of mean pressure
coefficient, mean velocity, and its root mean square were predicted with good accuracy. The comparison between the
two domes showed different flow behavior around them. A flattened horseshoe vortex was observed to develop around
the rough dome at larger distance compared with the smooth dome. The separation phenomenon occurs before the
apex of the rough dome while for the smooth dome it is shifted forward. The turbulence-affected region in the wake was
larger for the rough dome.
Keywords
Computational fluid dynamics, large eddy simulation, domes, wind load

Introduction
Flow around hemispherical domed structures is relevant to a variety of practical engineering applications
notably domed buildings, roofs, and sub-ocean structures. Domes were also used in hydraulic channels to
study the shedding of hairpin vortices in their wake in
a vein to understand the development of the near-wall
region of turbulent boundary layers.1 For robust
engineering design of such structures, detailed understanding and knowledge of the ow structure around
domes is necessary.
The ow around domes is three-dimensional and
contains large scale highly unsteady motions with separation. Complex vortical structures with shedding
and multiple reattachment and separation areas characterize this ow. Several parameters aect the ow
behavior around domes; they include the dome shape,
the Reynolds number (based on the approaching free
stream velocity and dome diameter), the inow conditions such as the approaching boundary layer shape
and turbulence content, the upstream-oor and dome
surface, and the surroundings topology.
A number of relevant experimental studies were
conducted on domes. Taniguchi et al.2 conducted
an experiment to identify the eects of the dome
size and the characteristics of the approaching boundary layer on the pressure coecient and integral

properties such as drag and lift coecient. The pressure distributions along the symmetry plane of the
hemisphere were found to be highly similar in the
region of 050 , but show a marked dependency on
the Reynolds number in the region from 50 to 120 .
Savory and Toy3 conducted experiments to study the
eect of three dierent approaching boundary layers
and dome surface roughness on the mean pressure
distribution and on the critical Reynolds number
beyond which the pressure distribution becomes
invariable. For the articially roughened dome subjected to a thin boundary layer, the reattachment
length of the downstream recirculation zone was
equal to 1.25, the dome diameter from its axis.
Tamai et al.4 explored the formation and shedding
of vortices from a dome in a water tunnel and determined the frequencies characterizing each phenomenon for a rather low Reynolds number of 104.
Approaching the front side, successive recirculation
zones, increasing in size and forming the well-known

Department of Mechanical Engineering, Petroleum Institute,


United Arab Emirates
Corresponding author:
L Khezzar, Department of Mechanical Engineering, Petroleum Institute,
PO Box 2533, Abu Dhabi, United Arab Emirates.
Email: lkhezzar@pi.ac.ae

Downloaded from pic.sagepub.com at King Saud University on April 13, 2013

XML Template (2013)


K:/PIC/PIC 474211.3d

[16.1.20132:14pm]
(PIC)

[115]
[PREPRINTER stage]

Proc IMechE Part C: J Mechanical Engineering Science 0(0)

horseshoe vortices, were observed. The number of


these structures increases with increasing Reynolds
number until a critical value of about 3000 beyond
which only one single recirculation zone persists
with a constant size. Cheng and Fu5 studied the
eect of Reynolds number on pressure distribution.
This study identied three modes of ow separation
in the apex region: for low Reynolds below 1.8  105,
the free shear layer which develops in the wake region
separates without reattachment on the dome surface
this corresponded to one single peak of the uctuating
pressure coecient. At a Reynolds number of
1.8  1053  105, a second peak was observed corresponding, to the reattachment point caused by a separation bubble which has formed on the dome
surface. Beyond a value of 3  105 of the Reynolds
number, only one peak was observed meaning that
the separation bubble has disappeared.
Using Reynolds averaged NavierStokes (RANS)
based turbulence models, Meroney et al.6 found that
the three turbulence models, SpalartAllmaras, k",
and Reynolds stress model, gave similar pressure distribution results with minor variations. Tavakol et al.7
used the RNG k" turbulence model to investigate the
eects of thin and thick approaching boundary layers
on the ow aerodynamics. The reattachment point
behind the obstacle which is of the order of
1.0851.17 the dome diameter was located at a shorter
distance for the thick boundary layer. All these studies
have simulated the ow using the steady-state equations of motion. In general steady and to some extent
unsteady RANS simulate such ows with acceptable
qualitative accuracy. Manhart8 used the large eddy
simulation (LES) turbulence model based on the
Smagorinsky subgrid-stress model with the law of
the wall. In this study, the dome roughness was generated by a stepwise approximation of the curved
wall. The study was based on the case reported in
Savory and Toy,3 of an articially roughened dome
at a Reynolds number of 1.4  105 and two grid sizes
of 650,000 and 1,863,680 computational cells were
used with an integration time step of 0.00250.005 s.
Two mechanisms of vortex generation are mentioned,
one from the top face due to the separation of the
free shear layer from the region surrounding the
dome apex causing symmetric vortex shedding and
an another one from the side faces engendering asymmetric vortex shedding. Tavakol et al.9 conducted a
comparison between the RNG k" and several
LES subgrid turbulence models over smooth domes.
Although their comparison was limited to few mean
streamwise velocity proles, the RNG k" model was
overestimating the streamwise velocity over the hemisphere. The results showed that LES simulation using
the single equation transport model for the subgrid
kinetic energy turbulence model (KETM) with
proper grid resolution greater than 4 million cells,
are able to capture the mean velocity proles accurately in the apex and wake regions.

From the review of previous work, it is clear that


RANS-based approaches to simulate ows over hemispherical domes performed with relative success and
that relatively very few LES-based numerical simulations were conducted on such ows. LES-based simulations oer the advantage of simulating explicitly the
large scales while modeling the small subgrid scales.
This approach is hence appropriate for aerodynamic
ows around domes. In this study, a numerical simulation based on the LES model using the KETM subgrid-scale model is used to calculate and benchmark
the ow over two hemispherical domes of Savory and
Toy3,10,11 and examine in detail the features of the surrounding ow eld. Available, experimental mean surface pressure coecient and velocity proles with
corresponding root mean square (RMS) are used in
the validation of the numerical results. This particular
test case consists of turbulent ow over an (a) articially roughened surface which can exist in practical
applications and, hence, presents an additional level
of complexity in modeling the surface details in contrast to ows over smooth surfaces for the LES framework and (b) over a smooth dome of the same size.
Taking advantage of the visualization of the ow
that can be provided by LES the eect of the surface
roughness is examined in detail and contrasted to the
smooth dome case. In particular, this article presents a
novel methodology of how to model and account for
surface roughness in structures such as domes within
the LES approach for industrial applications. The
standard ke model was used to generate the starting
ow for the LES simulation over the rough surface and
is also used in the discussion of the results.

Mathematical model
LES is used in this study. The uid is assumed incompressible and the ltered continuity and momentum
time dependent equations solved are given by
@U i
0
@xi

@U i @U i U j @ij @p @ij



@t
@xj
@xj @xj @xj

 reprewhere the variables with an overbar, U i and p,


sent the ltered (locally averaged) values of the velocity and pressure, respectively. The laminar stress
tensor is given by
 

@Ui @U j
ij 

3
@xj @xi
 is the molecular viscosity.
According to Tavakol et al.,9 where three subgridscale models were contrasted for ows around domes,
the subgrid-scale model based on the transport of the
subgrid-scale turbulent kinetic energy performs better
for ows of this type and it is hence used in this study.

Downloaded from pic.sagepub.com at King Saud University on April 13, 2013

XML Template (2013)


K:/PIC/PIC 474211.3d

[16.1.20132:14pm]
(PIC)

[115]
[PREPRINTER stage]

Kharoua and Khezzar

This model developed independently by a number of


authors, is not based on the local equilibrium hypothesis and belongs to the class of model based on
subgrid scales in contrast to the dynamic
SmagorinskyLilly model which is based on the
resolved scales properties.12 The grid scale cuto can
be in the inertial range and thus allowing the usage of
relatively coarse grids.13
The subgrid stress accounting for the unresolved
scales contribution dened by
ij Ui Uj  U i U j

is modeled according to the following equation


1
2
5
ij  ksgs ij 2Ck k2sgs f Sij
3
where ksgs is the subgrid-scale turbulent kinetic
energy, f V1=3 the lter size or characteristic
length, based on the average dimension of a computational nite volume, and Sij the usual resolved mean
strain rate tensor given by


1 @U i @U j

Sij

6
2 @xj @xi

and the kinetic energy of the subgrid modes is given by



1 2
Uk  U 2k
7
2
From equation (5), the subgrid-scale turbulent
eddy viscosity is then implicitly dened from kinetic
energy of the subgrid modes by
ksgs

t Ck k2sgs f

The subgrid-scale kinetic energy adds another


unknown to the problem and is evaluated by solving
a modeled transport equation which takes the form
given by equation (9).14
3

@ksgs @U j ksgs


@U i
k2sgs

ij
 C"
@t
@xj
@xj
f
1

@ k2sgs f @ksgs
Ck
@xj
k @xj

The constants Ck and C" are computed dynamically, based on the least-square method proposed by
Lilly,15 usually applied to the standard Smagorinsky
model to compute its constant dynamically, while the
Prandtl number  k is equal to unity.

Geometry and computational approach


Geometrical configuration
Figure 1 shows the geometry used for this study and
taken from Savory and Toy.3,10,11 The hemispherical

dome had a diameter of d 190 mm and was


immersed in a thin turbulent boundary layer of
86 mm thickness as in the experiment. In the experiments, two dome models were considered one with a
smooth surface and the other with an articially
roughened surface using a random coat of spherical
beads to give an equivalent roughness ratio of
k0 /d 0.01, where k0 is the diameter of the spherical
beads. The size of the computational domain, in the
spanwise y-direction, was equal to 4d. The experiments were conducted at a subcritical Reynolds
number based on the approach free stream velocity,
air properties, and dome diameter of 1.4  105.
However, the surface roughness induced ow separation resulted in an eectively supercritical ow.

Computational approach
In the present LES approach, the near-wall ow eld
is resolved rather than modeled. To simulate the surface roughness of the rough dome of Savory and
Toy,11 Manhart and Wengle16 adopted a technique
of blocking out the body-lled grid cells of the
dome within a Cartesian grid. In this study, a rather
dierent and more elaborate approach is adopted.
Alternate rectangular solid blocks (Figure 2) were
extruded from the curved wall of the dome. These
had a perpendicular depth to the dome surface
equal to 1.5 mm and edge sides less than 6 mm. In
this way, the articial roughness introduced by
Savory and Toy11 which corresponds to an average
bead size of 1.5 mm was modeled. The ideal representation of the dome roughness would use surface protrusions based on solid cubic cells with an average size
of 1.5 mm or even smaller. However, the generation
of the corresponding mesh, and the implementation of
the boundary conditions would, in this case, be highly
prohibitive. The approach used represents an acceptable compromise between a reasonable representation
of the surface details and time resources. The distribution of the solid blocks in Figure 2 is uniform in the
azimuthal direction but is in some places non-uniform
in the latitude direction. These would have a minor
impact since real surface roughness non-uniformity is
random. This simple modeling strategy will be shown
to account for the main ow features.
A hybrid mesh (Figure 3) of hexahedral and tetrahedral cells was generated. The computational
domain was decomposed into several blocks. A rst
block of clustered grids having the shape of a hemisphere surrounding the dome was created to facilitate
the generation of a structured hexahedral mesh in the
zones where most of the relevant ow phenomena
were expected to occur namely; the front side, the
apex region, and the wake. Further out, this hemispherical mesh was surrounded by a cube of tetrahedral cells due to the diculty of generating a
hexahedral mesh of good quality within the outer
cube truncated by the hemisphere surrounding

Downloaded from pic.sagepub.com at King Saud University on April 13, 2013

XML Template (2013)


K:/PIC/PIC 474211.3d

[16.1.20132:14pm]
(PIC)

[115]
[PREPRINTER stage]

Proc IMechE Part C: J Mechanical Engineering Science 0(0)

Wall
Inlet

Outlet

3d

Symmetry

z
yx
4d

8d

Figure 1. Boundary conditions and dimensions of the domain (flow from left to right).

Figure 2. Roughened dome model (the blue or dark blocks


represent the solid protrusions).

microscales. Based on an a priori RANS simulation,


the corresponding scale for this problem, was,
approximately, equal to 0.2 mm. The cell size in this
problem was chosen, due to the computational
resources limitations, to be 0.4 mm inside the region
surrounding the dome where smaller scales are
expected to reside. The values of y near the wall of
the dome was in the range 114. This mesh resolution
necessitated the use of a near-wall treatment model
which is discussed below.
Air is then working uid at ambient temperature,
having a density of 1.225 kg/m3 and dynamic viscosity
of 1.7894  105 kg/ms.
The boundary conditions were imposed in such
way to mimic the conditions under which the experiments of Savory and Toy3,10,11 were conducted. At the
inlet, measured proles of the mean streamwise velocity component in addition to the turbulent parameters required (k, ", and Reynolds stresses) were
imposed. The turbulent parameters were calculated
based on the prole of the streamwise turbulence
intensity measured in Savory and Toy.3 The spectral
synthesizer technique18,19 was employed for the generation of the uctuating velocity components.
In this method, uctuating velocity components
are computed by synthesizing a divergence-free velocity-vector eld from the summation of Fourier harmonics. In ANSYS FLUENT, the number of Fourier
harmonics is xed to 100.
The turbulent kinetic energy and its dissipation
rate proles were calculated from the measured turbulence intensity proles.

Figure 3. A view of the computational mesh.

k
the dome. Finally, a relatively coarse mesh was generated in the rest of the domain. The rened-mesh size
in the hemispherical block surrounding the dome is
1.5 mm and approximately the same size was kept for
the tetrahedral mesh. The computational domain was
divided into about 9 million cells for the two cases
studied. Studies on optimization of computational
resources based on LES requirements, such as
Addad et al.17 for example, showed that an ecient
grid should have a cell size around the Taylor

2
3
Uavg It
2

" C3=4


k3=2
l

10

11

where It is the turbulence intensity, C the constant


equal to 0.09, and l the turbulence length scale, and is
equal to 0.07 the hydraulic diameter based on fully
developed duct ow empirical results.
A zero gradient boundary condition was imposed
at the outlet of the domain relying on the fact that it

Downloaded from pic.sagepub.com at King Saud University on April 13, 2013

XML Template (2013)


K:/PIC/PIC 474211.3d

[16.1.20132:14pm]
(PIC)

[115]
[PREPRINTER stage]

Kharoua and Khezzar

was suciently far from the dome. At the top of the


domain, a symmetry condition was used at a location
where no eect of the dome on the ow is noticeable.
Manhart8 invoked the fact that, with such boundary
conditions, the mass ux arising from the growth of
the approaching boundary layer is omitted and might
result in some discrepancies compared with the
experiments which were conducted inside a wallbounded test section. In this study, the computational
domain was truncated in the free stream, at z/d 3,
relying on preliminary simulations conducted with the
real test section dimensions and the truncated
domain. The two lateral sides were taken as periodic
boundaries, placed at 2d from the central vertical
plane of symmetry. No-slip wall boundary conditions
were imposed at solid surfaces. The Werner and
Wengle20 wall function was used in regions where
the mesh was not ne enough to resolve the viscous
sub-layer. Fluent documentation states that there are
no restrictions on the grid size near the wall for LES
since the wall boundary conditions were implemented
using the law of the wall approach although it is
advisable to rene the mesh near the wall to a value
of y 1 for a better accuracy of the results.
The nite volume technique implemented in the
uent code version 12.1 is used to integrate the conservation equations. The computations were performed using the unsteady segregated solver. A
bounded central dierencing scheme was used to discretize the convective term in the ltered Navier
Stokes and subgrid-scale turbulent kinetic energy
equations. It is a composition of the pure secondorder central dierencing scheme, the second-order
upwind scheme and the rst-order upwind scheme,
see the documentation of Fluent version 12.121 referring to the Normalized Variable Diagram approach
proposed by Leonard.22 The pressure and velocity
were decoupled using the SIMPLE algorithm. The
pressure interpolation was performed by the
PRESTO scheme.
The implicit temporal discretization was of secondorder accuracy. The steady mean ow was computed
using the k" model to provide reasonable initial conditions for the LES simulation. Subsequently, an LES
simulation was conducted during about 1 s corresponding roughly to ve times the residence time of
the ow based on an average velocity of 10.76 m/s
and the domain length. When the ow developed,
the simulation was run during another 1 s for the collection of statistics. The local time-averaged variables
were calculated using
;

N
1X
;x, y, z, ti
N i1

12

where ;x, y, z, ti is the local and instantaneous variable at the position (x, y, z) of the computational
domain at time ti and N the number of samples

collected to compute the statistical averages, the sampling interval time was chosen equal to the integration
time step.
The integration time step was 104 s and insured a
cell Courant number in the domain less than 2. Recent
studies on optimization of LES computational eort
(e.g. Kornhaas et al.23) showed that a Courant
number equal to 2 would be sucient for an acceptable solution, using an implicit discretization scheme,
in terms of accuracy and stability for shear ows. The
calculations were conducted with uent running in
parallel mode using 32 processors of a high performance cluster HPC during about 60 days of continuous
run characterized by a total wall-clock time/time step
equal to 296 s and a total CPU time per time step
equal to 9480 s with a total of 19,597 time steps.

Results
The results of the simulations are presented in this
section for the smooth and rough domes of Savory
and Toy.3,10,11 The mean ow eld, represented by the
pressure coecient, the average streamwise velocity
component, and the dividing streamline showing the
size of the wake recirculation zone, are discussed rst.
A qualitative and quantitative assessment of the ow
features, in dierent regions surrounding the dome, is
then presented. Finally, the development of the
Reynolds stress proles, in the wake region are
discussed.

Mean flow field


Figure 4 illustrates the distribution of the pressure
coecient on the dome centerline, in the streamwise
x-direction, for the smooth and rough domes subjected to an approaching thin boundary layer.
Savory and Toy3 stated that for the Reynolds
number considered, 1.4  105, the ow regime over
the rough model was supercritical in which case the
pressure distribution over the dome surface becomes
independent of the Reynolds number while it was still
subcritical for the smooth dome. The calculated and
experimental values of the mean pressure coecient
for the smooth dome are shown on Figure 4(a). Both
LES and k" models predictions are good on the front
side until the apex region at about 85 , where the peak
corresponds to the separation of the boundary layer
developing on the dome surface. A slight decrease is
observed between 0 and 7 . On the front side, a small
recirculation zone was generated in the corner formed
by the dome curved surface and the bottom wall. The
second peak at about 21 corresponds to the stagnation point caused by the impact of the approaching
ow on the dome surface. The incoming ow impacts
rst on a recirculation zone generated upstream of the
dome which acts as an obstacle. The approaching ow
is, then, deviated upward impacting the dome surface
at the position of 21 . Discrepancies are observed

Downloaded from pic.sagepub.com at King Saud University on April 13, 2013

XML Template (2013)


K:/PIC/PIC 474211.3d

[16.1.20132:14pm]
(PIC)

[115]
[PREPRINTER stage]

Proc IMechE Part C: J Mechanical Engineering Science 0(0)

(a) 0.8
[3]

0.6

LES
Mean pressure coefficient

0.4
k
0.2
0.0
0.2
0.4
0.6
0.8
1.0
0

20

40

60

80

100

120

140

160

180

Angle ()
(b) 0.8
[3]

0.6

LES
Mean pressure coefficient

0.4

0.2
0.0
0.2
0.4
0.6
0.8
1.0
0

20

40

60

80

100

120

140

160

180

Angle ()

Figure 4. Mean pressure coefficient distribution on the dome centerline: (a) smooth dome and (b) rough dome.

starting from 93 for the LES model and 100 for the
k" model. The LES exhibits a local jump in the pressure coecient prole due probably to a reattachment
of an existing separation bubble causing the second
negative peak at about 97 but during recovery
matches the experimental values. In the wake
region, the LES model slightly under-predicts the
pressure coecient, whereas the k" model shows
better agreement with experiments. For the rough
dome and in the front stagnation region, the maximum value of the pressure coecient is slightly
over-predicted by the ke and LES calculations until
40 . This behavior might be due to a lack of resolution
of the ow in this region. This was also observed by
Manhart8 who attributed it to the dierent shapes of
the oncoming experimental and calculated boundary
layers. The decrease from the maximum value is captured correctly by the calculations until the peak suction value around 80 . Both the LES and the ke
models exhibit a rst local negative maximum at 65
followed by a decreasing trend until 75 , where a
second sharp increase is observed. The roughness
eect has moved the separation point backward

compared with the smooth dome, i.e. there is an earlier separation. The separation and reattachment phenomena are more pronounced on the surface of the
roughened dome due to the nature of the surface. The
recovery of the ke model appears to be more rapid
than the LES results.
Savory and Toy3 referred to the local centerline
drag coecient as a useful parameter to predict the
trend of the pressure forces acting on the dome surface for dierent Reynolds numbers. The local drag
coecient exhibits a minimum which designates the
critical ow regime (critical Reynolds number). In this
study, the local drag centerline drag coecient was
equal to 0.5 for the rough dome and 0.63 for the
smooth dome with a deviation from the experimental
results equal to 4.6% and 40%, respectively.
The uctuating pressure coecient is presented in
Figure 5 along the centerline of the dome in the symmetry plane y/d 0. The distribution along the rough
dome is characterized by three peaks at 10 , 90 , and
109 , respectively, while, for the smooth dome, the
peaks are observed at 2 , 14 , and 112 , respectively.
Cheng and Fu5 explained that the uctuating pressure

Downloaded from pic.sagepub.com at King Saud University on April 13, 2013

XML Template (2013)


K:/PIC/PIC 474211.3d

[16.1.20132:14pm]
(PIC)

[115]
[PREPRINTER stage]

Kharoua and Khezzar

0.16
LES rough

rms pressure coefficient

0.12

LES smooth

0.08

0.04

0.04
0

20

40

60

80

100

120

140

160

180

Angle ()

Figure 5. RMS surface pressure coefficient distribution on the dome centerline.


RMS: root mean square.

(a)

(b)

Figure 6. Mean velocity vectors (plane of symmetry) superimposed with the contours of RMS of Cp (dome surface).
Thick lines indicate the maximum RMS of Cp. (a) Smooth and (b) rough.
RMS: root mean square.

coecient peaks are related to transition of separation


bubbles and reattachment phenomena. For
Re < 1.8  105, as in the present case, they observed
a rst peak near the apex region (8090 ) and a
broader lump at 140 . In the present simulation with
a smooth dome, only one peak, near the separation
region, is observed at about 112 , the second peak is
not detected probably because the separation bubble
is not resolved. Two other peaks are generated in the
front side under the eect of the impingement of the
incoming ow on the dome surface at 14 and the
reattachment of a small recirculation zone, generated
in the corner formed by the dome and the oor, at 2 .
The peak at 112 corresponds approximately to the
separation of the boundary layer developing along the
upper surface of the dome. The related ow behavior
is illustrated in Figure 6(a) on which was superimposed the surface contours of the RMS of the pressure
coecient, and the mean velocity vectors plotted in
the plane of symmetry. The position of the RMS

Cp peak is indicated by the thick horizontal line.


The abovementioned small recirculation zone and
impact point are clearly seen. The situation is dierent
for the rough dome case where only one peak is
observed on the front side due to the absence of the
small recirculation zone (Figure 6(b)). Near the apex,
a rst peak is observed at 90 and second one at 109
which, according to Cheng and Fu,5 are due to separation and reattachment phenomena, respectively.
Figure 7 shows the dividing streamline in the wake
region. Taking the axis of the dome as a reference in
the streamwise x-direction, the position of the mean
reattachment is 0.8d and 1d for the smooth and rough
domes, respectively. An underestimation of the
reattachment distance is clearly seen compared with
the values measured by Savory and Toy3 for the same
case or the value of Tavakol et al.7 which was 1.17 for
a Reynolds number equal to 6.4  104, based on a free
stream velocity of 8.5 m/s and a diameter of the dome
of 0.12 m.

Downloaded from pic.sagepub.com at King Saud University on April 13, 2013

XML Template (2013)


K:/PIC/PIC 474211.3d

[16.1.20132:15pm]
(PIC)

[115]
[PREPRINTER stage]

Proc IMechE Part C: J Mechanical Engineering Science 0(0)

0.8
[3]: rough
0.7

LES rough
LES smooth

0.6

z/d

0.5
0.4
0.3
0.2

Dome
Surface

0.1
0
0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.1

1.2

1.3

x/d
Figure 7. Dividing stream line.

0.8

x/d=0.4

x/d=1.2

x/d=1

x/d=0.8

x/d=0.6

0.7
0.6

z/d

0.5
0.4
0.3
0.2
0.1
0
0.5 0.5

1.5

0.5 0.5

1.5

0.5

0.5

1.5

0.5

0.5

1.5

0.5

0.5

1.5

Figure 8. Profiles of the mean streamwise velocity component in the vertical direction at different x/d positions: , Savory;24 , LES
rough; and - - -, LES smooth.
LES: large eddy simulation.

To investigate the causes of the underestimation of


the reattachment point, proles of the average streamwise velocity component, obtained by LES for both
rough and smooth domes and experiments of
Savory24 for the rough dome, are plotted at dierent
longitudinal positions in the vertical z-direction in
Figure 8. At x/d 0.8 and approaching the oor surface, the reversed ow starts to become underestimated
causing the location of zero-streamwise velocity to
occur at a lower position z/d. At x/d 1, the simulated

ow has already reattached while the experiments still


show a small backow in the region below z/d 0.33.
At x/d 1.2, which is approximately the measured
reattachment position, the streamwise velocity component, predicted by LES, is positive everywhere in the
vertical direction while it is equal to zero below
z/d 0.3. From the mean velocity proles, the main
discrepancy between a rough and smooth dome exist
within the separating shear layer, with the rough dome
proles showing a higher momentum decit.

Downloaded from pic.sagepub.com at King Saud University on April 13, 2013

XML Template (2013)


K:/PIC/PIC 474211.3d

[16.1.20132:15pm]
(PIC)

[115]
[PREPRINTER stage]

Kharoua and Khezzar

(a1) smooth

(b1) rough

(a2) smooth

(b2) rough

Figure 9. Isosurface vorticity contours from side and isometric views: (a) smooth (1000 s1) and (b) rough (1200 s1).

Flow structures and patterns


Isosurfaces of vorticity and mean path lines are used
to illustrate features of the ow. In addition, stream
traces obtained from the mean velocity eld are also
used. A comparison between the smooth and rough
domes is conducted in the light of the qualitative and
quantitative results presented.
Figure 9 shows isosurface vorticity contours which
illustrate the main ow features previously described.
An intense turbulent activity close to the oor is
observed in the front side of the dome and develops
around it. This turbulent behavior corresponds to the
horseshoe vortex and is more pronounced for the
smooth dome. For the rough dome, it is not clearly
shown due to the value of the vorticity, which was
chosen to capture the wake turbulence rather. The
ow separation occurs earlier for the rough than for
the smooth dome as shown before with the wall pressure coecient results. The turbulent structures generated by the separated boundary layer from the top
face of the smooth dome, tend to be entrained in a
direction slopping downward, while for the rough
model, they move parallel to the longitudinal
direction.
Figure 10 shows the streamlines projected on several horizontal planes. Close to the bottom at
z/d 0.005, the horseshoe vortex is present with a lateral axis at about x/d 0.9 for the smooth dome and
x/d 1.1 for the rough dome. Two counter-rotating
recirculation zones, generated in the wake region,
extend to a longitudinal position of x/d 0.8 for the
smooth dome and x/d 1 for the rough dome. At
z/d 0.05, the horseshoe vortex for the rough dome,
disappears while, in the wake, the recirculation zones
are still present with a larger size for the rough dome.

Once the horseshoe, for the rough dome, disappears,


the streamlines tend to converge in a more pronounced way in the far wake compared with the
smooth dome case. For both geometries, it seems
that, in the positive lateral wake side, some of the
streamlines circumventing the recirculation zone are
entrained by the opposite recirculation zone
rather than the streamwise ow. The counter-rotating
recirculation zones vanish at a vertical position of
z/d 0.25 for the smooth dome, while for the rough
dome, they disappear at z/d 0.375.
The most important ow feature seen in the vertical planes perpendicular to the main ow direction
(Figure 11) is the rolled up vortices in the lateral
sides of the dome belonging to the horseshoe vortex.
At x/d 0 (apex), trailing vortices, with a quasi-circular cross section, are seen to be closer to the dome for
the smooth model while attened vortices are located
farther from the rough dome, at about y/d 1.25. In
the wake region at x/d 1, the trailing longitudinal
vortices are attened for both cases under the eect
of diusion and decay. In the vertical streamwise direction (Figure 12), the main observed features are the
horseshoe vortex upstream of the dome and the recirculation zone in the wake. In the plane y/d 0, the
horseshoe vortex axis is located at x/d 0.65 and
z/d 0.05 for the smooth dome. A attened thinner
horseshoe vortex is generated at x/d 0.98 and
z/d 0.025. In the wake of the smooth dome, the
center of the recirculation zone has the coordinates
x/d 0.58 and z/d 0.27 while a larger one is generated behind the rough dome at x/d 0.59 and at a
higher vertical position z/d 0.355. The center of
the recirculation zone obtained in the experiments of
Savory and Toy,3 was at approximately x/d 0.77
and z/d 0.38. The recirculation zone axis predicted

Downloaded from pic.sagepub.com at King Saud University on April 13, 2013

XML Template (2013)


K:/PIC/PIC 474211.3d

[16.1.20132:15pm]
(PIC)

[115]
[PREPRINTER stage]

10

Proc IMechE Part C: J Mechanical Engineering Science 0(0)

Streamlines in different planes xy

Streamlines at z/d =0.005


1

1
y/d

y/d

0.5
0

-0.5

-1

-1
-1

1
x/d

-1

x/d

a) smooth

b) rough

Streamlines at z/d = 0.05

0.5

0.5
y/d

y/d

-0.5

-0.5

-1

-1
-1

1
x/d

a) smooth

-1

1
x/d

b) rough

Figure 10. Time-averaged streamlines plotted in xy planes at different vertical positions z/d.

by LES is located at a closer distance to the dome and


slightly lower which explains the shorter reattachment
distance compared with the experimental results
(Figure 7). At y/d 0.5, the recirculation zone in the
wake has disappeared for the smooth dome while it is
still present for the rough dome.
The main ndings through this comparison are the
dierent shapes and positions of the features characterizing the turbulent ow around domes especially
the attened shape of the horseshoe vortex for the
rough dome and its position away from the obstacle.
Bradshaw25 explained that the pressure gradients
might play a key role in determining the position of
the characteristic phenomena. Longitudinal pressure
gradients aect the separation position along the
dome surface while lateral pressure gradients might
move the vortex shoe away from the corner formed
by the dome and the oor. Unfortunately, the literature does not contain enough experimental results for

a detailed comparison between the smooth and the


rough domes.

Turbulent flow field prediction


Figures 13 and 14 show the proles of the Reynolds
stresses <u2>, <v2>, and turbulent kinetic energy k in
the wake region in the plane of symmetry y/d 0 and
in a horizontal plane at z/d 0.158. These results refer
to the rough dome for which experimental values are
available in Savory and Toy.11
In the plane of symmetry, the calculated longitudinal correlation <u2> over-predicts the experimental
maximum peaks within the shear layers between
x/d 0.4 to 0.8. The values inside the recirculation
wake zone are, however, nicely replicated.
Nonetheless, the positions of the peaks, indicating
the approximate location of the shear layer bounding
the wake region, are well captured. In view of the

Downloaded from pic.sagepub.com at King Saud University on April 13, 2013

11

Kharoua and Khezzar

Streamlines at 1Jd = 0.1

0
x/d

x/d

a) smooth

b) rough

Stream li nes at d cl =0.25


0.5

0.5

~ 0

~ o

-0.5

-0.5

0
x/d

x/d

a) smooth

b) rough

Streaml ines at d cl =0.375


0.5

">.

~ 0

-0.5

-0.5

x/d

a) smooth

nature and structure of the wake for both rough and


smooth domes, the profiles of <u 2 > for the smooth
dome case exhibit large deviations. The region
bounded by the shear layer is remarkably larger for
the rough dome. The profiles for the <v2 > component
are, also, better predicted for the rough geometry.
They show that the velocity fluctuations are Jess
important at high z/d positions corresponding to the
shear layer separated from the top surface of the
dome. In contrast , the Reynolds stress component in
the spanwise direction exhibit a highly turbulent
behavior at z/d < 0.2 far from the dome at 0.8 :(
x/d :( 1.2. This region corresponds to the zone of
impact of the turbulent structures resulting from the
separation phenomenon from the lateral sides of the
dome, at low z/d positions, known as Von Karman
vortices. Similarly to the Reynolds stresses, the turbulent kinetic energy, shown in Figure 13(c), is well predicted at x /d = 1.2 while, for the remaining axial
positions, discrepancies, especially for the peaks, can
be observed. It is noteworthy to mention that the

x/d

b) rough

RNG k- e model, also, predicted, correctly, the location of the peaks of turbulent kinetic energy although
with a clear underestimation. This is expected since
such model is unable to simulate correctly flows
with strong mean streamline curvature and anisotropy
as exists in the wake of the dome.
Figure 14(a) shows profiles of the <u2 > component
in a horizontal plane at z/d = 0.158 and shows that the
profiles are, relatively, well predicted for the positions
x/d = I and x/d = 1.2 while remarkable overestimated
peaks are observed close to the obstacle at x /d = 0.6
and x /d = 0.8. This is, probably, due to the strong
turbulent effect induced by the technique used to represent the surface roughness of the dome which is not
entirely perfect. However, contrary to the predictions
on the plane of symmetry, the location of the peaks is
slightly shifted downward compared with the experimental results which means that a shorter shear-Jayerbounded region is being predicted by the numerical
simulation and which might , also, explain the shorter
reattachment length observed from Figure 8.

XML Template (2013)


K:/PIC/PIC 474211.3d

[16.1.20132:16pm]
(PIC)

[115]
[PREPRINTER stage]

12

Proc IMechE Part C: J Mechanical Engineering Science 0(0)

Figure 11. Time-averaged streamlines plotted in yz planes at different longitudinal positions x/d.

0.5

z/d

z/d

0.5

0
-1

-0.5

0.5
x/d

1.5

-1

-0.5

0.5
x/d

a) smooth

1.5

1.5

b) rough

0.5
z/d

z/d

0.5

-1

-0.5

0.5
x/d

a) smooth

1.5

-1

-0.5

0.5
x/d

b) rough

Figure 12. Time-averaged streamlines plotted in xz planes at different lateral positions y/d.
Downloaded from pic.sagepub.com at King Saud University on April 13, 2013

XML Template (2013)


K:/PIC/PIC 474211.3d

[16.1.20132:16pm]
(PIC)

[115]
[PREPRINTER stage]

Kharoua and Khezzar

(a) 0.8

x/d=0.8

(b) 0.8

x/d=1.2

x/d=1

x/d=0.4

x/d=0.6

x/d=0.8

0.7

0.6

0.6

0.5

0.5

z/d

z/d

x/d=0.6

x/d=0.4

0.7

13

0.4

x/d=1.2

x/d=1

0.4
0.3

0.3

0.2

0.2

0.1

0.1

0
0

0
0

0.1

0.1

0.1

0
(c)

0.1
0.8

x/d=0.4

0.7

0.1

0.1

0.1

0.1

0.1

0.1

x/d=0.8

x/d=0.6

x/d=1.2

x/d=1

0.6
z/d

0.5
0.4
0.3
0.2
0.1
0
0

0.1

0.1

0.1

0.1

0.1

Figure 13. Profiles of the Reynolds stresses and the turbulent kinetic energy in the plane of symmetry y/d 0 at different longitudinal positions: , Savory and Toy;11 , LES rough; - - -, LES smooth; and , RNG k".
LES: large eddy simulation.

(a) 0.8

x/d=0.6

x/d=0.8

x/d=1

(b) 0.8

x/d=1.2

x/d=0.6

x/d=0.8

0.7

0.6

0.6

0.5

0.5
y/d

y/d

0.7

0.4

x/d=1

x/d=1.2

0.4

0.3

0.3

0.2

0.2

0.1

0.1

0
0

0.1

0.1

0.1

0.1

0.05

0.05

0.05

0.05

(c) 0.8

x/d=0.6

x/d=0.8

0.7

x/d=1

x/d=1.2

0.6
y/d

0.5
0.4
0.3
0.2

0.1
0
0

0.1

0.1

0.1

0.1

Figure 14. Profiles of the Reynolds stresses and the turbulent kinetic energy in the horizontal plane z/d 0.158 at different longitudinal positions: , Savory and Toy;11 , LES rough; - - -, LES smooth; and , RNG k".
LES: large eddy simulation.

Downloaded from pic.sagepub.com at King Saud University on April 13, 2013

XML Template (2013)


K:/PIC/PIC 474211.3d

[16.1.20132:16pm]
(PIC)

[115]
[PREPRINTER stage]

14

Proc IMechE Part C: J Mechanical Engineering Science 0(0)

The peaks are much lower for the smooth model leading to a smaller shear-layer-bounded region in the
wake and an even shorter reattachment longitudinal
location. Figure 14(b) illustrates proles of the lateral
Reynolds stress component <v2 > in the same horizontal plane close to the bottom surface. It can be
observed that, close to the dome at x/d 0.6, a
slight peak exists at approximately y/d 0.47 which
is not well predicted by LES. Beyond x/d 0.6, the
lateral stress component <v2> is overestimated and
the trend is a monotonic decay of the lateral Reynolds
stress component from the symmetry plane toward
the outer region until, approximately, y/d 0.5.
Proles of the turbulent kinetic energy are shown in
Figure 14(c). At x/d 0.6 and x/d 0.8, the peaks at
y/d 0.38 and y/d 0.32, respectively, represent the
eect of the longitudinal component while the relatively good prediction of k beyond x/d 1 is consistent with the good prediction of the stress components
at the same positions.

Conclusions
A LES of a turbulent ow around smooth and rough
domes was conducted. The subgrid-scale model based
on the transport of the subgrid-scale turbulent kinetic
energy was used. The rough dome surface modeling
presented a challenge to capture its geometrical
details. The surface roughness of the rough dome surface was incorporated into the geometry using solid
blocks extruded with an average size of the glass
beads diameter used in the experiment and represents
a reasonable approach to model surface details.
The pressure coecient distribution along the centerline of the dome was predicted with very good
accuracy although with slight discrepancies in the
recovery region. The LES under-predicted the pressure coecient slightly for the smooth dome but
over-predicted it for the rough dome although the
dierences were slight. The predicted rear reattachment point downstream of the dome was located
closer to the dome compared to the experiments.
For the rough dome, the Reynolds stresses were
well predicted in the vertical plane of symmetry
although the peaks were overestimated. In the horizontal plane close to the oor, however, the peaks
were underestimated announcing a smaller turbulence-aected region in the wake compared with the
experiments.
The LES model allowed the visualization of features of the ow that are dicult to observe experimentally at high Reynolds numbers. The ow around
the rough dome was characterized by a attened
horseshoe vortex shifted away from the obstacle compared to the smooth dome. The horseshoe vortex
develops around and behind the dome in the form
of trailing vortices, becoming progressively parallel
to the ow direction with a separating lateral distance
between either sides being larger for the rough dome.

In the wake, a larger vorticity content region bounded


by the separated shear layers was observed for the
rough dome.
Funding
This research received no specic grant from any
funding agency in the public, commercial, or not-for-prot
sectors.

References
1. Acarlar MS and Smith CR. A study of hairpin vortices in
a laminar boundary layer, Part1. Hairpin vortices generated by a hemisphere protuberance. J Fluid Mech 1987;
175: 141.
2. Taniguchi S, Sakamoto H, Kiya M, et al. Time-averaged
aerodynamic forces acting on a hemisphere immersed in
a turbulent boundary. J Wind Eng Ind Aerodyn 1982; 9:
257273.
3. Savory E and Toy N. Hemispheres and hemispherecylinders in turbulent boundary layers. J Wind Eng Ind
Aerodyn 1986; 23: 345364.
4. Tamai N, Asaeda T and Tanaka N. Vortex structures
around a hemispheric hump. Boundary Layer Meteorol
1987; 39: 301314.
5. Cheng CM and Fu CL. Characteristic of wind loads on a
hemispherical dome in smooth flow and turbulent
boundary layer flow. J Wind Eng Ind Aerodyn 2010;
98: 328344.
6. Meroney RN, Letchford CW and Sarkar PP. Comparison
of numerical and wind tunnel simulation of wind loads on
smooth, rough and dual domes immersed in a boundary
layer. Wind Struct 2002; 5: 347358.
7. Tavakol MM, Yaghoubi M and Masoudi MM. Air flow
aerodynamic on a wall mounted hemisphere for various
turbulent boundary layers. Exp Therm Fluid Sci 2010; 34:
538553.
8. Manhart M. Vortex shedding from a hemisphere in a
turbulent boundary layer. Theor Comput Fluid Dyn
1998; 12: 128.
9. Tavakol MM, Abouali O and Yaghoubi M. Large eddy
simulation of turbulent air flow over a surface mounted
hemisphere. In: Proceedings of the ASME 2010 3rd joint
US-European fluids engineering summer meeting and 8th
international conference on nanochannels, microchannels,
and minichannels, Montreal, QC, 15 August 2010,
Paper no. FEDSM-ICNMM2010-30742, pp.18.
10. Savory E and Toy N. The flow regime in the turbulent near wake of a hemisphere. Exp Fluids 1986; 4:
181188.
11. Savory E and Toy N. The separated shear layers
associated with hemispherical bodies in turbulent
boundary layers. J Wind Eng Ind Aerodyn 1988; 28:
291300.
12. Sagaut P. Large eddy simulation for incompressible
flows. 3rd ed. Berlin: Springer Verlag, 2006, p.173.
13. Patel N, Stone C and Menon S. Large eddy simulation
of turbulent flow over and axisymmetric hill. AIAA
paper no. 20030967, 2003.
14. Kim WW and Menon S. Application of the localized
dynamic subgrid-scale model to turbulent wall-bounded
flows. In: 35th AIAA aerospace sciences meeting, Reno,
NV, 610 January 1997, Technical report AIAA-970210.

Downloaded from pic.sagepub.com at King Saud University on April 13, 2013

XML Template (2013)


K:/PIC/PIC 474211.3d

[16.1.20132:17pm]
(PIC)

[115]
[PREPRINTER stage]

Kharoua and Khezzar

15

15. Lilly DK. A proposed modification of the Germano


subgrid-scale closure model. Phys Fluids 1992; 4:
633635.
16. Manhart M and Wengle H. Large-eddy simulation of
turbulent boundary layer flow over a hemisphere.
In: PR Voke, L Kleiser and J-P Chollet (eds)
Direct and large-eddy simulation I. Dordrecht, the
Netherlands.
ERCOFTAC,
Kluwer
Academic
Publishers, 1994, pp.299310.
17. Addad Y, Gaitonde U, Laurence D, et al. Optimal
unstructured meshing for large eddy simulations.
In: MV Salvetti, B Geurts, J Meyers, et al. (eds)
Quality and reliability of large-eddy simulations,
ERCOFTAC series 12(II). Berlin: Springer Verlag
GmbH, 2008, pp.93103.
18. Kraichnan R. Diffusion by a random velocity field.
Phys Fluids 1970; 11: 2131.
19. Smirnov R, Shi S and Celik I. Random flow generation
technique for large eddy simulations and particledynamics modeling. J Fluids Eng 2001; 123: 359371.
20. Werner H and Wengle H. Large-eddy simulation of
turbulent flow over and around a cube in a plate channel. In: Eighth symposium on turbulent shear flows,
Munich, Germany, 911 September 1991, Technical
University Munich.
21. Fluent Inc. Fluent 12.1, User Guide, February 2010.
22. Leonard BP. The ULTIMATE conservative difference
scheme applied to unsteady one-dimensional advection.
Comput Meth Appl Mech Eng 1991; 88(1): 1774.
23. Kornhaas M, Sternel DC and Schafer M. Influence of
time step size and convergence criteria on large eddy
simulations with implicit time discretization.
In: MV Salvetti, B Geurts, J Meyers, et al. (eds)

Quality and reliability of large-eddy simulations,


ERCOFTAC series 12(II). Berlin: Springer Verlag
GmbH, 2008, pp.119130.
24. Savory E. Hemispheres in turbulent boundary layers.
PhD Thesis, University of Surrey, UK, 1984.
25. Bradshaw P. Complex three-dimensional turbulent
flows. In: 8th Australasian fluid mechanics conference,
University of Newcastle, N.S.W., Newcastle, UK, 28
November to 2 December 1983, pp.k5.1k5.7.

Appendix
Notation

x, y, z

dome diameter
turbulence intensity
turbulent kinetic energy
pressure
strain rate tensor
time
fluctuating velocity components
velocity components
filtered velocity component in the xi
direction
Cartesian coordinates

f

t
 ij
ij

LES filter size


laminar dynamic viscosity
turbulent dynamic viscosity
laminar stress tensor
subgrid-scale stress tensor

d
It
k
p
Sij
t
u, v, w
U, V, W
U i

Downloaded from pic.sagepub.com at King Saud University on April 13, 2013

Вам также может понравиться