Sie sind auf Seite 1von 15

Computers & Fluids 35 (2006) 1154–1168

www.elsevier.com/locate/compfluid

Flow over a bluff body from moderate to high Reynolds


numbers using large eddy simulation
A. Sohankar
Department of Mechanical Engineering, Faculty of Engineering, Yazd University, Yazd, Iran

Received 6 September 2004; received in revised form 2 March 2005; accepted 17 May 2005
Available online 3 November 2005

Abstract

Large eddy simulation (LES) is performed to study the uniform approach flow over a square-section cylinder with different Rey-
nolds numbers, ranging from 103 to 5 · 106. Two different sub-grid scale models, the Smagorinsky and a dynamic one-equation
model, are employed. An incompressible finite-volume code, based on a non-staggered grid arrangement and an implicit fractional
step method with second-order accuracy in space and time, is used.
The structure of the flow is studied with the instantaneous and the mean quantities such as pressure, turbulent stresses, turbulent
kinetic energy, vorticity, the second invariant of velocity gradient and streamlines. The Strouhal number, the mean and RMS values
of the lift and drag are computed for various Reynolds numbers, which show a good agreement with the available experimental
results. It is found that the effect of Reynolds number on the global quantities, the mean and the large scale instantaneous flow-struc-
tures is not much at the higher Reynolds numbers, i.e. Re > 2 · 104. In this range of Reynolds numbers, the small scales of the
instantaneous structures are more complex and chaotic as they compare with the larger ones.
 2005 Elsevier Ltd. All rights reserved.

1. Introduction shear layers, dominant large scale structures and high


turbulence.
The flow over objects such as high-rise buildings, In these bodies, the behaviour of flow changes as the
marine risers, bridges, cooling towers, chimneys and Reynolds number (Re) is increased. At this point, it may
tube banks in heat exchangers are encountered exten- be suitable to attend to special type of bluff bodies,
sively in engineering applications. These objects, which circular and square prisms. At the Reynolds numbers
under normal circumstances usually create a massive below about unity, the flow is fully attached with no sep-
wake region downstream, are called bluff bodies. aration. When Re increases, the flow separates and a
Motion of objects such as submarines, ships and cars pair of steady symmetric vortices forms behind the
through fluids are examples of a pivotal role of such body. At higher Reynolds numbers, the formation vor-
bluff objects in industrial applications. Thus, the knowl- tex length of the re-circulation region behind the body
edge of the flow field around these structures is of major grows with increasing Re. At a critical onset Reynolds
importance in a multitude of applications such as civil, number [1,2], the twin-vortex arrangement becomes
wind engineering and technical problems associated unstable, and a time-periodic oscillation wake and a
with energy conversion and structural design. This type staggered vortex street form. The separated vortices
of flow often contains many complex phenomena such shed alternately from the upper and the lower side of
as separation, wake flow, vortex shedding, curved free the body. The periodic phenomenon is referred to as
vortex shedding, whereas the anti-symmetric wake flow
pattern is referred to as the von Kármán vortex street.
E-mail address: asohankar@yazduni.ac.ir This is the first 2D wake transition which is called

0045-7930/$ - see front matter  2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.compfluid.2005.05.007
A. Sohankar / Computers & Fluids 35 (2006) 1154–1168 1155

Bénard–von Kármán instability [1,2]. This transition short review on LES mathematical formulation and
contains a 2D instability from a steady wake to a peri- the sub-grid scale models used in this work presents in
odic wake. By increasing the Reynolds number, a 3D Section 3. A description of the numerical solution proce-
transition develops at around Reynolds number of dure is given in Section 4. The analysis of the global,
about 200, and the three-dimensional flow effects appear time-averaged and instantaneous results is presented in
[3,4]. As Re still increases higher, the effects of flow three three subsections in Section 5. Finally, some conclusions
dimensionality and turbulence become more and more are drawn in Section 6.
pronounced.
The fundamental fluid dynamics problems of bluff
bodies such as circular and rectangular cylinders have 2. Problem description
been examined extensively in both numerical and exper-
imental, especially at low and moderate Reynolds num- The geometry of the domain of interest is shown in
bers [1–4]. In recent years, researchers attention has Fig. 1. As seen, a square cylinder with a side d is exposed
turned to the use of LES for studying the turbulent flow to a constant free stream velocity U. The flow is
around bluff bodies at higher Reynolds numbers. A LES described in a Cartesian coordinate system (x, y, z) in
Workshop was held in June 1995 in Germany, and its which the x-axis is aligned with the inlet flow direction,
results were published in Ref. [5], while they were com- the z-axis is parallel with the cylinder axis and the y-axis
pared with LDV-measurements of Lyn et al. [6]. The is perpendicular to both x and z. The origin of the coor-
same flow was considered as a test case LES2 at the First dinate system is located at the centre of the body, see
ERCOFTAC Workshop on Direct and Large Eddy Fig. 1. In order to obtain the governing equations and
Simulation in March 1994. Seven groups took part in
the LES2 exercise, and the results of this exercise were
reported by Voke [7]. Sohankar et al. [8,9] have investi-
gated flow around a rigid prism of a square cross-section H
with one side facing the oncoming flow at Re = 2.2 · 104
using LES with three different sub-grid scale models, the
Smagorinsky, the dynamic and a dynamic one-equation
model. The influence of a finer grid, shorter time step A
and the larger computational aspect ratio on their
results was also studied.
The reason for this focus on LES for the study of the
flow around bluff bodies has to do with poor results,
L2
when one uses statistical turbulence models. Most pro- U
bably, this has to do with complicating factors such d
as a strongly retarded stagnation flow, massive flow
separation, streamline curvature, transition from lami- L1
nar to turbulent flow, re-circulation, vortex shedding
and, perhaps most important, the existence of inher- 4
ent three-dimensional flow structures [3,5]. In recent
years, the most of the LES works for flow over a square
cylinder have been performed for a single Reynolds
number, i.e. Re = 2.2 · 104 and the effect of Reynolds 2
numbers on the results of this case is not studied
comprehensively.
The main purpose of the present study is to investi- y
0 x
gate the turbulent flow around a square cylinder for dif-
ferent Reynolds number ranging from 103 to 5 · 106.
The computation of the global results and the study of
the mean and instantaneous flow structures are also the -2
aims of this work. All computations were performed
using large eddy simulations (LES) with two different
sub-grid scale models, the Smagorinsky (SM) and a
-4
dynamic one-equation model (OEM). -2 0 2 4
The organisation of this paper is as follows. Section 2 Fig. 1. Flow configuration and the computational domain (top); and
is devoted to describe the flow configuration, computa- close-up of the grid near the body, the non-uniform grid 266 · 162 in
tional domain, grid and the boundary conditions. A the xy-plane.
1156 A. Sohankar / Computers & Fluids 35 (2006) 1154–1168

different quantities in their non-dimensional form, the ~ui;i ¼ 0 ð1Þ


solution variables are normalized as follows. All geo- 1
ot ~ui þ ð~ui ~uj Þ;j þ ~p;i þ sij;j ¼ Re ~ui;jj ð2Þ
metrical lengths are scaled with d. The dimensions of
the computational domain are denoted with L1, L2, d, The influence of the small scales on the large (resolved)
H and A (see Fig. 1) and their normalized values are scales takes place through the sub-grid scale stresses
selected respectively as 7.5, 16, 1, 16 and 4 for the pres- (SGS), sij ¼ r ~ij  ~ui ~ui ðrij ¼ ui uj Þ, which are unknown
ent study. Velocities are normalized with the inlet free and must be modelled with a sub-grid model. The sub-
stream velocity, U, and the physical time with d/U. Scal- grid stresses in LES are only a fraction of the total stres-
ing with d/U also applies for the shedding frequency, ses, thus, it is expected that the modelling error should
where St = f d/U is the Strouhal number. The Reynolds not affect the overall accuracy of the results as much
number is defined as Re = U d/v. In the y-direction, the as in RANS approach. The details of the LES formula-
vertical distance between the upper and the lower walls, tion are found in [10].
H, defines the solid blockage of the confined flow
(blockage parameter, b = d/H).
Based on the refinement study of the previous work 3.1. Sub-grid models
of the author [8], all simulations are performed for two
resolutions of 186 · 106 · 26 and 266 · 162 · 26 The most of the sub-grid models are based on the
(denoted as fine grid, see Fig. 1) points in x, y and z, eddy-viscosity model. They use the Boussinesq hypothe-
respectively. The grid distribution is uniform with a sis, which expresses a direct proportionality between the
constant cell size, outside a region from the body, deviatoric part of the turbulent stress and the velocity
which extended two units upstream, downstream and gradients, i.e. saij  sij  dij skk =3 ¼ vt ð~ui;j þ ~uj;i Þ.
sideways (in the x- and y-directions). The minimum The most widely used SGS model is Smagorinsky
distance from the cylinder surface to the nearest grid model [11] (SM). In this model, the eddy viscosity,
which is the product of a length (l / D ~ ¼ ðdV Þ1=3 ) and
point approximately is 0.008. The hyperbolic tangent qffiffiffiffiffiffiffiffiffiffi
function was used for stretching the cell sizes around a velocity (qsgs / l S~ij S~ij ) scales, is calculated algebrai-
the body. A uniform grid was used in the spanwise qffiffiffiffiffiffiffiffiffiffiffiffiffi
~ 2 2S~ij S~ij , where S~ij ¼ 0:5ð~
cally as: vt ¼ ðC s DÞ ui;j þ ~
uj;i Þ.
direction. A uniform flow (u = 1, v = w = 0) was
prescribed at the inlet as a boundary condition. At In the above relations, S~ij ; dij ; vt ; D;
~ C s and dV are the
the outlet, the convective boundary condition ðouoti þ strain tensor, the Kronecker delta, the sub-grid eddy vis-
U ou
ox
i
¼ 0Þ [1] was applied for all velocity components. cosity, the filter width (e.g. grid spacing), the propor-
No-slip conditions were prescribed at the body sur- tionality factor, and the local cell volume, respectively.
faces. Symmetry conditions simulating a frictionless In this model, Cs is a constant value that must be
wall ðou
oy
¼ ow
oy
¼ v ¼ 0Þ were used at the upper and lower assigned a priori to simulation. In the present study,
boundaries. A periodic boundary condition was Cs was set to 0.1. This model is incapable of taking into
employed in the spanwise direction. account the reduction of the length scales near the solid
walls. A remedy is to use a damping function. In addi-
tion, the Smagorinsky model is absolutely dissipative
3. Governing equations and sub-grid model and cannot account for backscatter, but energy is always
transferred from large to small scales.
In LES, all scales larger than a selected size (e.g. Dynamic models, which are capable of removing
grid size) are computed exactly. Since the small scales some of the drawbacks of the Smagorinsky model, are
tend to be more dissipative, isotropic, short-lived, a suitable alternative. The first attempt to introduce a
homogeneous, random, universal, and less affected by dynamic SGS eddy viscosity model was developed by
the boundary conditions than the larger ones, they Germano et al. [12]. In this model, the proportionality
are not explicitly simulate. Thus, the effect of the small factor is calculated during the computation process. A
scales of turbulence in LES is modelled. Three-dimen- new filter, test filter, is used to calculate this coefficient.
sional and the time dependent solution of the In this model, in addition to the grid filter (here denoted
Navier–Stokes equations are used in LES. The LES by a tilde), a test filter (here denoted by an overbar) is
governing equations are accomplished using a general also employed, whose width is larger than the grid filter
spatial filtering of the instantaneous equations, e.g. a width (e.g. D  ¼ 2D). ~ One of the drawbacks of the
box filter, to limit the range of the scales in the flow dynamic model is the numerical instability associated
field. In this procedure, a general variable is decom- with the negative values and the large variation of the
posed into a large scale and a sub-grid scale compo- proportionality coefficient (C) [8,9]. To avoid numerical
nent, e.g. f ¼ f~ þ f sgs . For an incompressible flow of instability owing to an extensive variation of C in time
a Newtonian fluid, the filtered Navier–Stokes equations and space, spatial averaging in the homogeneous direc-
take the following form [9]: tion and additional local averaging are performed on C.
A. Sohankar / Computers & Fluids 35 (2006) 1154–1168 1157

In the present study, a one-equation dynamic sub- viscosity field including a significant partition with neg-
grid model (OEM) is also employed. This model is able ative values [9,13], which is destabilizing in numerical
to capture the backscattering phenomena and it is insen- simulations. The C Khom is determined at each time step
sitive to the existence of homogeneous directions and with requirement that the sub-grid dissipation of the
also more stable than standard dynamic model resolved kinetic energy Pksgs in the whole computational
[8,9,13]. This model was developed by Davidson [13] domain remains the same as it is with the local coeffi-
and has been successfully applied to different flow con- cient, C k , i.e.
figurations [8,9,13–16]. In this model, the eddy viscosity 1 1

~ 0:5 ) is based on a length scale (l / D ~¼ h2C k k 2Ksgs S~ij S~ij Di


~ k ~ ~ ~
xyz ¼ 2C hom hk Ksgs S ij S ij Dixyz
2
ð7Þ
(vt ¼ C K Dk ksg pffiffiffiffiffiffiffi
1=3 sgs where h ixyz denotes the space averaging over the entire
ðdV Þ ) and a new velocity scale (q / k sgs ). ksgs is
the sub-grid turbulent kinetic energy and it is deter- domain.
mined by its modelled transport equation [13] as: The idea is to include all local dynamic information
through the source terms of the transport equation for
uj k sgs Þ;j ¼ ½ðv þ C khom k 0:5
ot k sgs þ ð~ ~
sgs DÞk sgs;j ;j ksgs. This is probably physically more sound since large
~
þ P Ksgs  C k k 1:5 =D; ð3Þ local variations in the dynamic coefficients appear only
 sgs
in the source term, and the effect of the large fluctuations
P Ksgs ¼ saij ~
ui;j ;
in the dynamic coefficients will be smoothed out. In this
~ k k 0:5 S~ij ;
saij ¼ sij  dij skk =3 ¼ 2vt S~ij ¼ 2DC sgs way, it turns out that the need to average or limit the
S~ij ¼ 0:5ð~ui;j þ ~uj;i Þ ð4Þ dynamic coefficients Ck in Eq. (5) and ðC k Þnþ1 in Eq.
(6) is eliminated altogether. The coefficients in the one-
In Pksgs, the dynamic coefficient Ck is computed in a way equation model affect the stresses in only an indirect
similar to that used in the standard dynamic model [12], way, while in the standard dynamics model the C coef-
i.e. ficient is linearly proportional to the stresses, which
‘ij M ij can make it numerically unstable.
Ck ¼ ; ‘ij ¼ ð~ ui ~ ui 
uj Þ  
~ ~
uj ; K ¼ k sgs þ 0:5‘ii ;
2M ij M ij
 0:5 S
M ij ¼ DK ~ 0:5 S~ij ;
~ij þ Dk 4. Numerical solution procedure
sgs

ð5Þ
An incompressible 3D finite volume code with a collo-
where ‘ij denotes the dynamic Leonard stresses. cated grid arrangement is used for solving the filtered
K  0.5Tii and Tij are the sub-grid kinetic energy and Navier–Stokes equations. Eqs. (1) and (2) are discretized
stresses on the test level, respectively (test filter opera- and they are solved with an implicit fractional two-step
tion denoted with an overbar in the relations). The S~ij method. Convective, viscous and sub-grid fluxes are
is denoted as the strain rate tensor on the grid level. approximated by central differences of the second-order
The Smagorinsky model is based on the assumption accuracy. All terms in the momentum equations are
of local equilibrium of SGS turbulent energy, i.e. advanced in time using the Crank–Nicolson scheme. The
Pksgs  eksgs = 0, where eksgs is the dissipation of ksgs. momentum equations are solved with the Gauss–Seidel,
A slightly better assumption for estimating the coeffi- whereas a multigrid V-cycle is used for the acceleration
cient ck in the dissipation term would be to assume that of the convergence, when solving the pressure equations.
the filtered right-hand side of the ksgs equation is equal The time-marching calculations were started with the fluid
to that of the K equation, i.e., at rest, and a constant time step, Dt = 0.02 was selected,
k nþ1
which was smaller than the chosen value for the most of
P ksgs  C k ðk 1:5 ~ k 1:5 
sgs =DÞ ¼ P K  C  ðK =DÞ ) ðC  Þ the cases in the previous work of author [8].
¼ ½P K  P ksgs þ ðC k Þn ðk 1:5 ~  1:5 The filtered Navier–Stokes equations are integrated
sgs =DÞðD=K Þ
over a time interval and over the control volumes, thus
ð6Þ the discrete form of Eq. (2) can be written as:
 0:5 
where P K ¼ 2C DK S ij S~ij . The above relation gives the
k ~
~unþ1 ¼ ~uni þ DtF ð~uinþ1 ; ~uni Þ  aDtP~ ;i
nþ1 n
 ð1  aÞDtP~ ;i ð8Þ
i
coefficient C k in the dissipation term for time step n + 1.
Note that ðC k Þn in Eq. (6) has been kept inside the filter- where F ð~unþ1
i ; ~uni Þ includes convection and viscous terms
ing process. and also the sub-grid stresses [9].
The dissipation cannot be negative, which requires The numerical procedure for obtaining ~ui at each time
that we limit C k to positive values, i.e. C k P 0. To step can be summarized as follows:
ensure numerical stability, a homogenous value of Ck
in space (C Khom ) is used in the momentum equations ~inþ1 .
1. Solve the discretized equation (8) for obtaining u
and in the diffusion term in Eq. (3). The reason is that At this stage, the most recent values of the pressure
the local coefficient Ck yields a highly oscillating eddy and velocities are used.
1158 A. Sohankar / Computers & Fluids 35 (2006) 1154–1168

2. Intermediate velocities are defined as: length was studied in the previous work of the present
n author [8] for this flow geometry. In Ref. [8], it was
~u#
i ¼ uni
~ þ DtF ð~uni ; ~
unþ1
i Þ  ð1  aÞDtP~ ;i ð9Þ
shown that using the finer resolutions in the space and
By using Eq. (8), u#
i can be calculated as:
time and also employing a larger spanwise dimension
nþ1 can improve the agreement between the predictions
u#
~ uinþ1 þ aDtP~ ;i .
i ¼ ~ ð10Þ and the experimental results, especially when also taking
The intermediate velocity field (~ u# into account the blockage effects. Based on the results of
i ) does not satisfy
the continuity equation. Velocities at control volume that work, the suitable resolutions are chosen for the
faces (~u# present study, see also Section 2.
face ) are calculated by interpolating from
neighbouring intermediate velocities at the nodes.
These interpolated cell face velocities are used to cal- 5.1. Global quantities
culate the mass fluxes and convection terms.
3. To satisfy continuity for ~ nþ1
uiðfaceÞ , take the divergence of A comparison between the data on various quantities
Eq. (10) and insert ~ nþ1
ui;iðfaceÞ ¼ 0, thus of the present LES simulations (1000 6 Re 6 5 · 106)
and the published experimental results [6,18–24] are pro-
nþ1 1
P~ ;ii ¼ ðaDtÞ ~u#
iðfaceÞ;i ð11Þ vided in Figs. 2 and 3. The 2D/3D numerical results of
the previous works of the present author (45 6 Re 6
This is the Poisson equation for obtaining pressure on 500) [4,17] are also included in this figure. It is important
the level (n + 1). This equation is solved with a multi- to mention that the RMS and mean drag (CD, CD 0 ) and
grid Poisson solver. A Neumann boundary condition also the RMS lift (CL 0 ) coefficients were computed from
was applied on the Poisson equation for the pressure. the time-dependent, spanwise-averaged forces on the
4. Compute the cell face velocities (which satisfy conti- cylinder for all simulations. The Strouhal numbers were
nuity) from the pressure intermediate velocity as calculated from the fluctuating lift signal.
nþ1 The variations of the Strouhal number and drag coef-
unþ1 u# ~
~ i;face ¼ ~ i;face  aDtðP ;i Þface ð12Þ
ficient with Reynolds number are shown in Fig. 2. As
These velocities are used to calculate the mass fluxes seen, there is an increase in the Strouhal number with
at the faces of the control volume. increasing the Reynolds number from about 50 to 200,
whereas the drag coefficient continuously decreases in
Steps 1–4 are carried out for each iteration until a this range of the Reynolds numbers, see Fig. 2. In the
global convergence criterion reaches at each time step. lower limit of this range, the phenomenon of vortex
Finally, the turbulent viscosity is computed and the next shedding results (Re  50) and a 2D instability from a
time step starts. steady wake to a periodic wake forms, which gives rise
to the fluctuating drag and lift forces [1]. Effect of three
dimensionality appears close to upper limit of this range
5. Results and discussion (150 < Re < 200 [4]), which causes to increase the St and
decrease CD with increasing the Reynolds number. The
Large eddy simulations were performed to study the St reaches to a local minimum value at around
uniform approach flow over a square-section cylinder Re = 1000, which is about 0.12 and it shows a good
for various Reynolds numbers with two different sub- agreement with the experimental result of Okajima
grid scale models, the Smagorinsky (SM) and a dynamic [18]. With increasing the Reynolds number from 103 to
one-equation model (OEM). The simulations were car- about 2 · 104, the St increases slightly, and then reaches
ried out for the selected Reynolds numbers of 103 approximately to a constant level at about 0.13 for the
(OEM), 3 · 103 (OEM), 8.7 · 103 (SM, OEM), 1.5 · higher Reynolds number ranging about 2 · 104–5 ·
104 (OEM), 2.2 · 104 (SM, OEM), 5 · 104 (SM), 105 106. This Reynolds-independent of the Strouhal number
(SM), 5 · 105 (OEM), 106 (OEM) and 5 · 106 (OEM). for Re > 2 · 104 is also observed for drag coefficient,
A series of time- and spanwise-averaged resolved veloc- RMS lift and drag in Figs. 2 and 3. As seen form Figs.
ity, pressure, turbulent stresses, and streamlines were 2 and 3, there is some difference between the results of
calculated. The time-averaged global quantities such as the present study with the experimental ones, but this
the Strouhal number, the base suction pressure, the Reynolds-independent of the results for Re > 2 · 104 is
mean and the RMS values of the drag force, the lift also seen in the experimental results. As it was shown
force and the pressure were also computed over about in the previous work of the present author [8], we can
20 shedding cycles and compared with available experi- improve the agreement between the predictions and
mental results. All simulations were carried out for a the experiments results by using finer resolutions in
solid blockage of b = 6%. space and time and also taking into account the block-
Influence of the spatial and the temporal resolutions, age effects. Some disagreements between the present
the blockage effects and the computational spanwise results and experimental ones are due to use of the
A. Sohankar / Computers & Fluids 35 (2006) 1154–1168 1159

Fig. 2. Strouhal number (St) and mean drag coefficient (CD) versus Reynolds number, comparison with published experimental [6,18–24] and
numerical [4,17] results.

Fig. 3. RMS lift and drag versus Reynolds number, comparison with published experimental [19,20,22,23,25] and numerical [4,17] results.

different inlet turbulence intensity and different blockage sure coefficient approximately varies from 1.03 to 1.01
effects in the experiments [8,22,23]. For example, it is for Reynolds number ranging from Re = 103 to
shown that when the turbulence intensity increases or Re = 5 · 106, respectively.
the blockage decreases, the drag value reduces [25,26].
It is important to mention that the uniform flow at inlet 5.2. Mean-flow structure
and a blockage of 6% are applied for the present
investigation. Both time- and spanwise-averaged resolved and fluctu-
From simulations of the present study (1000 6 Re 6 ation quantities such as pressure, turbulent stresses, tur-
5 · 106), it is found that the ratio of the mean drag coef- bulent kinetic energy and streamlines are provided for
ficient and the base suction pressure coefficient (CD/ different Reynolds numbers (1000 6 Re 6 5 · 105) to
(CPb)) is approximately constant and equals to gain an understanding of the mean flow structure. It is
1.47 ± 0.04. This finding is in agreement with the exper- worth to mention that the time-averaged quantities were
iments and also the previous work of the present author calculated over about 20 shedding cycles. The computed
[8]. If one computes this ratio from the published exper- resolved quantities are decomposed into a time-mean
imental results, they are about 1.45 [18], 1.52 [19], 1.54 component and a fluctuation (denoted by 0 ) component.
[20], 1.53 [21]. At low Reynolds numbers, this finding For example, the resolved velocities are decomposed as,
is not valid, where it is about 3, 2.32, 1.93, 1.74 and ~ui ¼ h~uit þ u0i . Hereinafter, the tilde and bracket symbols
1.48 for Re = 45, 100, 150, 200, 500 [4,17], respectively. are dropped form the resolved and the time- and span-
The results of this study show that the stagnation pres- wise-averaged quantities for simplicity.
1160 A. Sohankar / Computers & Fluids 35 (2006) 1154–1168

The time- and spanwise-averaged pressure coeffi- value of the pressure coefficient at the centre of these
cients (Cp), turbulent shear stresses (u 0 v 0 ) and turbulent bubbles on the sides of the body are (x = 0.19,
kinetic energy (k = 0.5(u 0 u 0 + v 0 v 0 + w 0 w 0 )) contours are y = ±0.65) and 2.12, respectively. As seen from
compared for two chosen Reynolds numbers, i.e. Fig. 5, the maximum of the pressure fluctuation (p 0 p 0 )
Re = 3 · 103 and 106. One can easily observe that the is also located in the centre of these bubbles, which is
time-and spanwise-averaged quantities are symmetric about 0.22. The minimum pressure coefficient behind
and a slight asymmetry in these plots is due to the insuf- the body is 2.05, which is located at the end of the
ficient averaging time, which is taken to obtain the reverse flow region on the centreline of the cylinder at
mean-values. around x = 1. These minimum pressure coefficients
The time- and spanwise-averaged pressure coefficient and their locations on the wake centreline and on the
contours are shown in Fig. 4(top). As seen, the differ- side of the body are very close to the aforementioned
ence between these two plots is not much. For example, values for Re = 3 · 103 and 106 in Fig. 4. The length
the minimum pressure coefficient behind the body is of the time-mean recirculation zone behind the body,
located on the centreline at around x = 1 with values lr, is about 1.05 for this case, see Fig. 5. The variation
of 2.04 and 1.94 for Re = 3 · 103 and 106, respec- of lr and the maximum mean absolute velocity value in
tively. The pressure coefficient at the centre of recirculat- the backflow on the wake centreline, UBF, are also rela-
ing regions on the sides of the body are about 2.01 and tively constant at high Reynolds numbers. For example,
2.15 for Re = 3 · 103 and 106, respectively. Similar the predicted values of lr and UBF are about (1.12, 0.1),
behaviour is found for the other selected Reynolds num- (1.05, 0.11), (1.02, 0.08), (0.97, 0.08) and (1.03, 0.07) for
ber of the present study. Re = 3 · 103, Re = 22 · 103 (high resolution case),
The time-and spanwise-averaged turbulent shear Re = 5 · 105, Re = 106 and Re = 5 · 105, respectively.
stress contours (u 0 v 0 ) are compared in Fig. 4(middle). The normalized time- and spanwise-averaged of the
Two plots are very similar with a slight difference different components of the resolved normal (u 0 u 0 , v 0 v 0 ,
between the maximum values, where their absolute val- w 0 w 0 ) and shear (u 0 v 0 ) stresses are shown also in Fig. 5.
ues are about 0.20 (x = 1.37, y = 0.45) and 0.21 It can be found from this figure that the maximum of
(x = 1.28, y = 0.45) for Re = 3 · 103 and 106, respec- the streamwise (u 0 u 0 ) and cross-flow (v 0 v 0 ) normal stres-
tively. Fig. 4(bottom) shows a comparison between tur- ses are larger than the magnitude of the shear stresses
bulent kinetic energy contours (k) for two Reynolds (u 0 v 0 ). The maximum values of u 0 u 0 are located some-
numbers. Here, the similarity between these two cases where in the shear layer along the sides of the body,
is also observed and the maximum values are located which is about 0.60, while the maximum value of v 0 v 0
on the centreline at around x = 1.5 with k = 0.50 and is located on the wake centreline (x = 1.5), which is
k = 0.47 for Re = 3 · 103 and 106, respectively. about 0.80. In Fig. 5, the contours of time- and span-
In general, similar behaviours are observed for the wise-averaged shear stresses (u 0 v 0 ) are also plotted. The
other quantities such as streamlines, velocity compo- magnitude of the maximum value of u 0 v 0 is located
nents, pressure fluctuation (p 0 p 0 ) and the components somewhere in the wake (x = 1.35, y = 0.46), and its
of turbulent stresses (u 0 u 0 , v 0 v 0 , w 0 w 0 ) for different Rey- value is about 0.22, which is in a good agreement with
nolds numbers in the range of the present study, for the cases that are shown in Fig. 4, especially for the case
instance, see turbulent shear stress contours (u 0 v 0 ) in with Re = 106. In this figure, the time- and spanwise-
Fig. 5 for Re = 2.2 · 104, (this case is provided for a fine averaged normal stresses isolines (w 0 w 0 ) is also plotted.
grid, 266 · 162 · 26). The similarity between the time- As seen, the maximum value is located on the wake
and spanwise-averaged quantities becomes closer at centreline (x = l), and it is about 0.12.
the Reynolds numbers which are approximately greater In Figs. 6 and 7, the time- and spanwise-averaged
than 2 · 104, where the dependency of the results to the streamwise velocity (u), pressure coefficient (Cp), turbu-
Reynolds number is weak, see also Figs. 2 and 3. lent normal stresses (u 0 u 0 , v 0 v 0 ), turbulent shear stresses
Because of similarity of the mean quantities for various (u 0 v 0 ) and turbulent kinetic energy ðk ¼ 0:5u0i u0i Þ are com-
Reynolds numbers, the results of only one chosen case pared for different Reynolds numbers and available
(Re = 2.2 · 104) is explained here and it is quantitatively experimental results [6,27]. The results along the wake
valid for the other Reynolds numbers of this centreline, y = 0, (Fig. 6) and at the streamwise posi-
investigation. tions, x = 0, 0.5, 1.5, (Fig. 7) are provided for different
The time- and spanwise-averaged streamlines are Reynolds numbers and sub-grid models, i.e. 3 · 103
shown in Fig. 5(bottom and right-side). In this plot, it (OEM), 8.7 · 103 (SM), 2.2 · 104 (SM and fine grid
is seen that the flow separates from the upstream corners OEM), 5 · 104 (SM) and 106 (OEM).
and it is not reattached to the upper and lower faces. Turbulent normal stresses (u 0 u 0 , v 0 v 0 ) and turbulent
Consequently, two large recirculating regions, bubbles, kinetic energy (k) are shown together with the experi-
on the sides of the body and a pair of counter-rotating mental results [6] at the wake centreline for different
vortices behind the body form. The position and the Reynolds numbers in Fig. 6. As seen, the agreement is
A. Sohankar / Computers & Fluids 35 (2006) 1154–1168 1161

Cp -0
Cp -0 .
44
Re=3000 -0 . . 5 Re=1000000
6 -0 .
1 -0 . 1 68
8
-1
-1 .
04

-2
-1 . -1 .4
9
-1 .6 -1 .4
-1 .8 -1 .6 4
-2 -1 .8 8
0
1

y
y

-1 -1

0 1 2 0 1 2
x x

u' v' u' v'


-0 .0 1 Re=1000000 -0 .0 1
Re=3000 5
-0 . 0 -0 .0 5
1 -0 .0 9 1 -0 .0 9
- 0 .1 3 -0.13
-0
-0 . .1
1 7
-0 .1 9 7 -0 .2
1

0. 03 0 .0 3
0 0
y

-1 -1

0 1 2 0 1 2
x x

k 2
Re=1000000
2
0 .0 6 k 0 .0 0 .0 6
Re=3000 0 .0
0 .1 2 0 .1 4
1 1
0.2 0 .2
0 .3 0 .3
0 .3 0 .3
0. 6
8 42
0.
44
0 .4 6
0.26
0 .2 6

0 0
y

y
0 .4 8

-1 -1

0 1 2 0 1 2
x x

Fig. 4. Time- and spanwise-averaged pressure coefficients (Cp), turbulent shear stresses (u 0 v 0 ) and turbulent kinetic energy ðk ¼ 0:5u0i u0i Þ contours for
Re = 3 · 103 and Re = 106.

good between the results of this study and the experi- these values, especially k, for Re > 2 · 104 is not much.
mental one, especially for Re = 2.2 · 104. In general, This means that although the values of u 0 u 0 , v 0 v 0 and also
the results are qualitatively similar for all cases. For w 0 w 0 change slightly with Reynolds number, but the sum
example, the maximum values of u 0 u 0 , v 0 v 0 and k are of these values at the high Reynolds numbers is approx-
about (0.212, 0.719, 0.495), (0.14, 0.755, 0.494), (0.138, imately constant. Thus, the dependency of the time- and
0.736, 0.479), (0.147, 0.716, 0.478) and (0.151, 0.699, spanwise-averaged quantities with Re are weak at the
0.470) for Reynolds numbers of 3 · 103 (OEM), high Reynolds numbers as it was mentioned before.
8.7 · 103 (SM), 2.2 · 104 (SM), 5 · 104 (SM) and 106 Pressure coefficient (Cp) at the wake centreline
(OEM), respectively. As seen, the difference between together with experimental results [27] are shown in
1162 A. Sohankar / Computers & Fluids 35 (2006) 1154–1168

u' u' v' v'


03 Re= 22000 0 .0
5
Re= 22000 0. 0 .0 9
1 0 .1 8 1
0 0 .1 5
0. 3 6 . 3 0 0 .3 0
0.40

0 .6
5

0.

0.
15

75
0 0
y

y
-1 -1

0 1 2 0 1 2
x x

w' w' u' v' 1


Re= 22000 1 Re= 22000 -0 .0
1 0 .0 1 -0 .0 7
0 .0 4 -0 .1 3
-0.17
0. 07 -0 .2 1
0 .0 9
0 .1 0 .0
1 3
0 0
y

y
3
-0 .0

1
0 .2
17
0.
0 .13
-1 -1 0 .0 7
0 .0 1

0 1 2 0 1 2
x x

p' p' 0.02 Re= 22000


Re= 22000 0.03
1 1
0.06

0.22
0 .0 9

1
0 .1
0 0
y
y

0 .0
6
-1 -1

0 1 2 0 1 2
x x
Fig. 5. Time- and spanwise-averaged streamlines, pressure fluctuation (p 0 p 0 ), the components of turbulent normal stresses (u 0 u 0 , v 0 v 0 , w 0 w 0 ) and
turbulent shear stresses (u 0 v 0 ) contours for Re = 2.2 · 104.

Fig. 6, where the minimum pressure coefficient is located mental results [6] at one streamwise position (x = 1.5)
at x = l with values of 2.04, 1.95, 1.86, 1.95 and the streamwise velocity (u) at three positions
1.84, 1.94 for Reynolds number of 3 · 103(OEM), (x = 0, 0.5, 1.5) are shown for different Reynolds num-
8.7 · 103 (SM), 2.2 · 104 (SM), 2.2 · 104 (OEM), bers. A relatively sufficient spatial resolution is used
5 · 104 (SM) and 106 (OEM), respectively. close the body. For example, the maximum negative
Time- and spanwise-averaged pressure coefficient velocity (u = 0.26) occurs at the 7th node from the
(Cp), turbulent shear stresses (u 0 v 0 ) and turbulent kinetic body at around y = 0.57 for Re = 106 for the coarser
energy ðk ¼ 0:5u0i u0i Þ versus y together with the experi- grid while this point is located at the 9th node for the
A. Sohankar / Computers & Fluids 35 (2006) 1154–1168 1163

Re
0. 8 3000, OEM
8700, SM Re
22000, SM 3000 , OEM
22000, OEM 8700, SM
50000, SM 22000, SM
1000000, OEM 22000, OEM
21400, Lyn et al. 50000, SM
v' v'

1000000, OEM
0. 5

k
0. 4

0 0
5 10 15 5 10 15
x x

0 0. 5

-0 .5
Re
3000, OEM
8700, SM
22000, SM
Re 22000, OEM
-1
50000, SM
3000, OEM
u'u'
Cp

8700, SM 0. 25 1000000, OEM


22000, SM 21400, Lyn et al.
-1 .5 22000, OEM
50000, SM
1000000, OEM
67000, Nakamura & Ohya
-2

-2 .5 0
5 10 15 5 10 15
x x
Fig. 6. Time- and spanwise-averaged pressure coefficient (Cp), turbulent normal stresses (u 0 u 0 , v 0 v 0 ) and turbulent kinetic energy ðk ¼ 0:5u0i u0i Þ versus x
at the centreline (y = 0) for different Reynolds numbers. Experiments: Lyn et al. [6], Nakamura and Ohya [27].

finer grid. As seen, the agreement between the present sional upstream corner separation, predominantly
results and the experiments is good and there are not from upstream corners at Re = 150 and finally, at all
large differences between the results for various Rey- times, from upstream corners for Re P 175. The results
nolds numbers, especially at higher Reynolds numbers. of this study also verify that the separation points
It is important to mention that some of the differences remains at the upstream corners at all times for the
between the results for different Reynolds numbers are higher Reynolds numbers. Thus the mean-structure of
due to the effect of sub-gird models, for example, see this the flow is not much sensitive to where the boundary
difference between the values of minimum pressure coef- layer is laminar or turbulent. In general, the sharp-edged
ficient for case Re = 2.2 · 104. The effect of sub-grid bodies, which tend to cause flow separation regardless of
model on the results was also reported in the previous the character of the boundary layer, are insensitive to
work of the present author [8]. In general, it is seen that the Reynolds number. This point was also mentioned
the effect of Reynolds number on the time- and span- in Ref. [28], where a relatively constant value for the
wise-averaged quantities is not much at high Reynolds mean drag coefficient (CD  2.1 for Re > 104) was
number, Re > 2 · 104, especially when one employs the reported for flow over a square cylinder.
same grid and sub-grid model.
It is worth mentioning that the previous numerical 5.3. Instantaneous-flow structure
work by the author [17] and later analysis [1] show that
the separation for Re 6 100, at all times in the fully To gain further insight into the similarity and dissim-
saturated state, occurred from the rear corners, predom- ilarity of the flow structure for different Reynolds num-
inantly from the rear corners at Re = 125 with occa- bers, the instantaneous wake flow structures were
1164 A. Sohankar / Computers & Fluids 35 (2006) 1154–1168

4 4 4
x= 0 x= 0.5 x= 1.5

3 3 3

2 2 2
y

y
1 1 1

0 0 0
-0.5 0 0.5 1 1.5 -0.5 0 0.5 1 1.5 0 0.5 1 1.5
u u u

4 4 4
Re Re
3000, OEM 3000, OEM
8700, SM 8700, SM
22000, SM 22000, SM
22000F, OEM 22000F, OEM
3 3 3
50000, SM 50000, SM
1000000, OEM 1000000, OEM
21400, Exp.

2 2 2
y

1 1 1

0 0 0
-0.2 0 0.2 -1.5 -1 -0.5 0 0 0.2 0.4 0.6
u'v' Cp k

Fig. 7. Time- and spanwise-averaged streamwise velocity (u at x = 0, 0.5, 1.5, see u 0 v 0 -plot for legend), pressure coefficient (Cp), turbulent shear
stresses (u 0 v 0 ) and turbulent kinetic energy ðk ¼ 0:5u0i u0i Þ versus y at x = 1.5 for different Reynolds numbers, Exp.: Lyn et al. [6].

provided for a number of selected Reynolds numbers in should have a vortex core with a net vorticity and the
the range of 103 6 Re 6 5 · 106. Due to the lack of geometrical characteristics of the identified vortex must
space, the results of only some cases with the same grid be Galilean invariant [29]. Because an agreed definition
and subgrid model (OEM) are compared here, i.e. of a vortex does not yet exit, thus it is difficult to recog-
Re = 3 · 103 and Re = 5 · 106 in Figs. 8 and 9, and nize the general vortical structure for all flow cases with
Re = 103 and Re = 106 in Fig. 10. Please note that the only one method, where these requirements for the vor-
x-axis in Figs. 8–10 is originated from the inlet of the tex core may not yield with a single scheme. Therefore,
computational domain (the body centre is located at several methods such as using streamlines, pathlines,
x = 7.85), while it is measured from the body centre in vorticity, pressure, the rate of strain tensor, etc. are pro-
Figs. 4–7, see also Fig. 1. pounded for identification and visualization of vortical
To study the instantaneous-flow structure, it is neces- structures [29]. In the present work, the instantaneous
sary to recognize the significant vortical regions in the flow structures at the mid-span and wake centre planes
flow field. It is important to mention that a vortex are studied using vorticity, pressure and the second
A. Sohankar / Computers & Fluids 35 (2006) 1154–1168 1165

Fig. 8. Instantaneous pressure coefficient (top), spanwise vorticity (middle) and the second invariant of velocity gradient (bottom) contours in the
streamwise direction at the mid-span (z = 0). Left: Re = 3 · 103 (OEM), right: Re = 5 · 106 (OEM), here the x-coordinate is measured from the inlet.

Fig. 9. Instantaneous pressure coefficient (bottom), cross stream vorticity (middle) and the second invariant of velocity gradient (top) contours in the
streamwise direction at the wake-centre (y = 0). Left: Re = 3 · 103 (OEM), right: Re = 5 · 106 (OEM), here the x-coordinate is measured from the
inlet.
1166 A. Sohankar / Computers & Fluids 35 (2006) 1154–1168

On the downstream part of the body the spanwise


(von Kármán) vortices of the alternating sign form,
see the corresponding black and white colours in
Fig. 8(middle). The strength of the vorticity decreases
rapidly in the streamwise direction, especially beyond
the vortex formation region (x > 11, approximately).
For example, the pressure coefficient and the magni-
tude of vorticity at the core of four vortices (Fig. 8 at
x > 11 and Re = 3000) are (2.72, 8.51), (1.9, 6.5),
(1.85, 4.53) and (1.13, 3), respectively. As seen in
Fig. 8, the three methods of vortex identification pre-
dict the same structure and vortex core at the mid-
span plane, while the similarity of the pressure and the
second invariant of velocity gradient patterns are much
closer.
It was reported in the previous work of the present
author [4] that a 2D/3D transition develops at
Re = Rec  150–200 for flow over a square cylinder
and both spanwise instability modes (A and B) are pres-
ent in the wake transitional process, similar to the case
of the flow around a circular cylinder [3]. At Re = 200
the spanwise wavelength of mode A instability associ-
ated with the initiation of transition from 2D to 3D flow
is approximately five diameters (diameter = side dimen-
sion of cylinder) [4]. The short-wavelength mode B
instability, appearing at around Re  250, has a wave-
length with the order of one diameter [4].
From the present results (e.g. Figs. 8–10), the stream-
Fig. 10. Iso-surfaces of the second invariant of velocity gradient
wise-oriented tubular flow structures similar to the mode
(Q = 0.4, 1). Top: Re = 106 (OEM), bottom: Re = 103 (OEM), here the B instability are approximately observed. The A mode
x-coordinate is measured from inlet. vortices are not seen in the present results, but close to
the cylinder they appear to have some similarities with
invariant of the velocity gradient tensor. It is worth the mode B instability as found at lower Reynolds num-
mentioning that a region with a pressure value lower ber [4] for Re P 250 (in Ref. [4] the highest Re is 500). It
than the ambient one and a positive second invariant is not possible to determine accurately the spanwise dis-
of velocity gradient tensor (Q = 0.5(ui,iuj,j  ui,juj,i) > 0) tance between the pairs of these tubular-like, streaky
is also defined as a vortex core [30]. structures from this instantaneous flow field. However
Instantaneous pressure coefficient, vorticity and the within the different planes (e.g. xz-plane in Fig. 9), the
second invariant of velocity gradient (Q) contours at number of pairs along lines, e.g. x = 12 at Re = 5 ·
the wake-centre plane (y = 0, xz-plane) and at the mid- 106, appears to be in the same order as the spanwise
span plane (z = 0, xy-plane) are provided for two differ- dimension. This implies as a rough estimate that the
ent Reynolds numbers, i.e. Re = 3 · 103 (CL = 2.39) spanwise distance of these structures are of the order
and Re = 5 · 106 (CL = 2.05) at one instant, where of one diameter. This compares with findings at high
the lift coefficient is close to its minimum value, see Figs. Reynolds numbers for the flow around a circular
8 and 9. Iso-surfaces of the second invariant of velocity cylinder [31].
gradient (Q) are shown at one instant (t = 160) for As it was shown in Sohankar et al. [4], the strength
Re = 103 (CL = 1.61) and Re = 106 (CL = 1.31) in and the spanwise coupling of the non-spanwise vortical
Fig. 10. As it is seen from these figures, although there structures in the near wake are closely related to the fluc-
are some differences in the size and the strength of the tuating drag. As can be seen in Fig. 8, there is a high
small-scale vortices, but the large scale instantaneous activity of the streamwise vorticity with high peak levels
flow-structures are approximately similar for both in the near wake but with a relatively small degree of
Reynolds numbers. These similarity and dissimilarity the spanwise coupling. The low spanwise coupling is
are observed in the instantaneous structures for the consistent with the fact that for this case the ratio
other selected Reynolds numbers of the present study between the spanwise-averaged and the spanwise-mean
(these results are not shown here due to the lack of sectional RMS drag (cD [4]) was as low as about 0.7.
space). Because the RMS drag and cD approximately constant
A. Sohankar / Computers & Fluids 35 (2006) 1154–1168 1167

at Re > 2 · 104, it is expected that the non-spanwise pared, no significant changes is observed in the global,
vortical structures become approximately similar, as it mean and instantaneous flow-structure. Thus, it is
is also seen in Figs. 3 and 9. expected that the accurate results can be obtained with
Fig. 10 shows 3D contour surfaces of (positive) Q. the fine resolution.
Only two contour surfaces are plotted in this figure. The correct prediction of the SGS stresses near the
The spanwise rollers of the von Kármán vortices are solid walls is also important to capture the real flow-
captured from this figure and also from the surfaces of structure. Based on the previous publications regarding
the constant pressure (only 2D plots for pressure are the LES with OEM model [8,9,13–17], it is believed that
shown in Fig. 9 at wake centreline plane) for different the OEM model can predict the SGS stresses near the
Reynolds numbers. The fluctuating lift is closely con- body more accurate than the Smagorinsky and the stan-
nected to the strength and near-wake development of dard dynamic models. This model is able to capture the
the spanwise (von Kármán) vortices [4]. The ratio backscattering phenomena and it is insensitive to the
between the spanwise-averaged and the spanwise-mean existence of homogeneous directions and also more sta-
sectional RMS lift (cL [4]) for this study was about ble than standard dynamic model [8,9,14]. Thus, it is
0.96. The sectional RMS lift was about 6 times larger expected that the effect of the sub-grid model, especially
than the sectional RMS drag. The near-unity value of OEM model, on the results and the flow-structure is not
cL implies, in the mean sense, a strong spanwise coupling much.
of the sectional lift forces. As can be seen in Figs. 9 and
10, there is a relatively high degree of spanwise coupling
for the developing spanwise vortex still connected to the 6. Conclusions
cylinder. For other spanwise (von Kármán) vortices,
which can be traced in Figs. 8–10, some appear to be Viscous uniform approach flow over a square-section
strong and highly correlated across the span with only bluff body is studied using large eddy simulation with
little undulation, e.g. the vortex centred in Fig. 9 at two sub-grid scale models and different Reynolds
approximately x = 8.5 (Re = 3 · 103 and Re = 5 · 106) numbers, ranging from 103 to 5 · 106. The calculation
and x = 10.5 (Re = 5 · 106), see also Fig. 10 at the posi- of global results and the study of mean and instanta-
tion of x  14 for Re = 103 and Re = 106. While some neous flow structures for different Reynolds numbers
vortices appear to be rather distorted across the span are the aims of this work. The Strouhal number, the
and with a much higher degree of spanwise undulation, mean and RMS values of lift and drag are computed
see Figs. 9 and 10. Although the large-scale structure of for various Reynolds numbers and they are com-
the spanwise vortices is similar for different Reynolds pared with the available experimental results. The global
numbers, but it seems that the spanwise undulation of results show a good agreement with experiments and
the von karman vortices slightly increases at the higher they are approximately Reynolds-independent for
Reynolds numbers, see Figs. 9 and 10. Reynolds number greater than 2 · 104. It is found that
In general, while the large scale spanwise vortical the ratio of mean drag coefficient and base suction pres-
structures become approximately similar for Re > 2 · sure is approximately constant for various Reynolds
104 due to approximately constant values of the RMS numbers.
lift and cL in this range of Reynolds numbers, see also It is also found that the effect of Reynolds number
Figs. 3 and 8–10, but there is some dissimilarity in the on the mean and the large scale instantaneous flow-
size and the strength of the small scale vortical struc- structure is not much at the high Reynolds number,
tures at the higher Reynolds numbers. Some of these i.e. Re > 2 · 104, especially when one employs the
dissimilarity may result from insufficient grid resolution same grid and sub-grid model. The results of this
and incorrect behaviour of the SGS stresses near the study also show that the separation points remains at
solid walls, especially for the higher Reynolds num- the upstream corners at all times for the various
bers. The presence of a solid wall can restrain the growth Reynolds numbers. Thus, it is expected that the flow-
of the small scales and can alter the exchange mecha- structure, especially the mean one, is not much sensitive
nisms between the resolved and unresolved scales in to where the boundary layer is laminar or turbulent. In
LES [32]. Thus, if the grid is very fine close to the solid general, the sharp-edged bodies, which tend to cause
walls, the real flow-structure will appear exclusively flow separation regardless of the character of the
through the resolved scales of LES. When the grid is boundary layer, are insensitive to the Reynolds number,
coarsened, part of the flow-structure becomes due to especially at higher Reynolds numbers. Finally, while
the unresolved scales and most of the energy production the large scale instantaneous flow-structures become
may occur at the sub-grid scale level. Based on the approximately similar for various Reynolds numbers,
refinement study of the previous work of the author but there is some dissimilarity in the size and the
[8], two resolutions are employed here for all simula- strength of the small-scale structures at higher Reynolds
tions. When the results of these resolutions are com- numbers.
1168 A. Sohankar / Computers & Fluids 35 (2006) 1154–1168

References [16] Schmidt S, Thiele F. Comparison of numerical methods applied to


the flow over wall-mounted cubes. Int J Heat Fluid Flow
[1] Sohankar A, Norberg C, Davidson L. Low-Reynolds number 2002;23:330–9.
flow around a square cylinder at incidence: study of blockage, [17] Sohankar A., Davidson L, Norberg L. Numerical simulation of
onset of vortex shedding and outlet boundary condition, I. J unsteady flow around a square 2D cylinder. In: Bilger RW, editor.
Numer Meth Fluids 1998;26:39–56. Proc 12th Australasian fluid mechanics conference, 1995. p. 517–
[2] Provansal M, Mathis C, Boyer L. Bénard–von Kármán instabil- 20.
ity: transient and forced regimes. J Fluid Mech 1987;182:1–22. [18] Okajima A. Strouhal numbers of rectangular cylinders. J Fluid
[3] Williamson CHK. Three-dimensional wake transition. J Fluid Mech 1982;123:379–98.
Mech 1996;328:345–407. [19] Luo SC, Yazdani MG, Chew YT, Lee TS. Effects of incidence and
[4] Sohankar A, Norberg C, Davidson L. Simulation of unsteady 3D afterbody shape on flow past bluff cylinders. J Wind Engng Ind
flow around a square cylinder at moderate Reynolds number. Aero 1994;53:375–99.
Phys Fluids 1999;11(2):288–306. [20] Vickery BJ. Fluctuating lift and drag on a long cylinder of square
[5] Rodi W, Ferziger JH, Breuer M, Pourquie M. Status of large eddy cross-section in a turbulent stream. J Fluid Mech 1966;25:481–94.
simulations: results of workshop. Trans ASME: J Fluids Eng [21] Norberg C. Flow around rectangular cylinders: pressure forces
1997;119:248–62. and wake frequencies. J Wind Engng Ind Aero 1993;49:187–96.
[6] Lyn DA, Einav S, Rodi W, Park JH. A laser-Doppler velocimetry [22] Lee BE. The effect of turbulence on the surface pressure field of a
study of ensemble-averaged characteristics of the turbulent near square prism. J Fluid Mech 1975;69:263–82.
wake of a square cylinder. J Fluid Mech 1995;304:285–319. [23] Tamura T, Miyagi T. The effect of turbulence on aerodynamic
[7] Voke PR. Flow past a square cylinder: test case Les2. In: Chollet forces on a square cylinder with various corner shapes. J Wind
JP et al., editors. Direct and large eddy simulation II. Kluwer Engng Ind Aero 1999;83:135–45.
Academic Publishers; 1997. [24] Saha AK, Muralidhar K, Biswas G. Experimental study of flow
[8] Sohankar A, Davidson L, Norbeg C. Large eddy simulation of past a square cylinder at high Reynolds numbers. Exper Fluids
flow past a square cylinder: comparison of different subgrid scale 2000;29:553–63.
models. ASME J Fluids Eng 2000;122:39–47. See also Erratum [25] McLean I, Gartshore C. Spanwise correlations of pressure on a
J Fluids Eng, vol. 122, p. 643. rigid square section cylinder. J Wind Engng Ind Aero 1992;41–
[9] Sohankar A. Numerical study of laminar, transitional and 44:779–808.
turbulent flow past rectangular cylinders. PhD thesis, Department [26] Saathoff P, Melbourne WH. Effects of free stream turbulence on
of Thermo and Fluid Dynamics, Chalmers University of Tech- streamwise pressure measured on a square-section cylinder.
nology, Gothenburg, 1998. J Wind Engng Ind Aero 1999;79:61–78.
[10] Sagaut P. Large eddy simulation for incompressible flows, an [27] Nakamura Y, Ohya Y. The effect of turbulence on the mean flow
introduction. Springer Verlag; 2001. past 2D rectangular cylinders. J Fluid Mech 1984;149:255–73.
[11] Smagorinsky J. General circulation experiments with the primitive [28] White FM. Fluid mechanics. 4th ed. McGraw-Hill; 1999.
equations. Month Weather Rev 1993;91:99–165. [29] Jeong J, Hussain F. On the identification of a vortex. J Fluid
[12] Germano M, Piomelli U, Moin P, Cabot WH. A dynamic Mech 1995;285:69–94.
subgrid-scale eddy viscosity model. Phys Fluids A 1991;3:1760–5. [30] Hunt JCR, Wray A, Moin P. Eddies, stream, and convergence
[13] Davidson L. Large eddy simulation: a dynamic one-equation sub- zones in turbulent flows. Center of Turbulence Research Report
grid model for three-dimensional re-circulating flow. In: 11th int CTR-S88, 1988. p. 193.
symp on turbulent shear flow, Grenoble, vol. 3, 1997. p. 26.1–6. [31] Chyu CK, Rockwell D. Evolution of patterns of stream-wise
[14] Sohankar A, Davidson L, Norbeg C. A dynamic one-equation vorticity in the turbulent near wake of a circular cylinder. J Fluid
sub-grid model for simulation of flow around a square cylinder. Mech 1996;320:117–37.
In: 4th international symposium on engineering turbulence [32] Piomelli U, Chasnov JR. Large eddy simulation: theory and
modeling and measurements, Corsica, France, May 24–26, 1999. applications. In: Hallback M, Henningson DS, Johansson AV,
[15] Krajnovic S, Davidson L. Large eddy simulation of the flow Alfredsson PM, editors. Turbulence and transition modelling.
around a bluff body. AIAA J 2002;40(5):927–36. Ercoftac Series. Kluwer Academic Publishers; 1996.

Das könnte Ihnen auch gefallen