Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Journal of Wind Engineering

and Industrial Aerodynamics 80 (1999) 31—46

2D LES of vortex shedding from a square cylinder


D. Bouris, G. Bergeles*
Laboratory of Aerodynamics, Department of Mechanical Engineering, National Technical University of Athens,
9 Heroon Polytechniou Avenue, 5, 15700 Zografou, Athens, Greece

Received 7 July 1997; accepted 2 July 1998

Abstract

In the high Reynolds number flow past a square cylinder, turbulent fluctuations are superim-
posed on the periodic vortex shedding motion making numerical calculation of the flow
a difficult task. In the present work a two-dimensional large eddy simulation is performed with
no-slip boundary conditions at the solid walls. A filtering procedure is introduced in frequency
space to separate the periodic from the turbulent fluctuations and the kinetic energy of both is
calculated along the centerline behind the rod. The drag coefficient, vortex shedding frequency
and spacing of the vortices in the wake are also calculated and the results are validated against
experimental measurements. It is found that the two-dimensional large eddy simulation using
a fine grid resolution, especially in the near wall region, gives a good representation of the
quasi-two-dimensional mechanisms of the flow since they are directly simulated instead of
being modeled as with statistical turbulence models.  1999 Elsevier Science Ltd. All rights
reserved.

Keywords: Large eddy simulation; Turbulence; Vortex shedding; Square cylinder

1. Introduction

Turbulent flows past bluff bodies are generally very complex due to large-scale
periodicity which is present in separation and vortex shedding from the solid body.
Traditional models that solve the Reynolds averaged Navier—Stokes equations using
a statistical turbulence model have been used in the past but they have not been able
to correctly predict both the periodic and the turbulent fluctuations present in the
flow [1,2]. Turbulence is present in a large frequency range and ideally a direct
solution of the Navier—Stokes equations would be needed. However, even if such
a solution were possible, from a computational point of view the cost would be very

* Corresponding author. E-mail: bergeles@fluid.mech.ntua.gr.

0167-6105/99/$ — see front matter  1999 Elsevier Science Ltd. All rights reserved.
PII: S 0 1 6 7 - 6 1 0 5 ( 9 8 ) 0 0 2 0 0 - 1
32 D. Bouris, G. Bergeles/J. Wind Eng. Ind. Aerodyn. 80 (1999) 31–46

high. The alternative is therefore a large eddy simulation (LES) of such a flow with as
accurate a resolution as possible. The goal of accurate resolution near solid walls is
difficult to attain and so law of the wall velocity distributions are often used leaving
a large part of the turbulent boundary layer unresolved [3,4]. Also, in the case of
quasi-two-dimensional flows, three-dimensional computations often fail to capture
details of the quasi-two-dimensional mechanisms which might be more influential to
the flow than the three-dimensional effects.
The purpose of the present paper is to provide an accurate simulation of
the quasi-two-dimensional turbulent flow past a square cylinder using a large
eddy simulation in two dimensions. It has been stated in the past [3,4] that 2D
LES calculations are clearly inferior to three-dimensional ones since certain impor-
tant features of three-dimensional turbulence are not resolved. The three dimensional-
ity of turbulence cannot be questioned, however, in the present paper, the importance
of detailed simulation of the quasi-two-dimensional mechanisms will be evaluated.
Sometimes, the discretization of the third direction is attained at the expense of
the other two while resolution of the boundary layer near solid walls is either
inadequate or is approximated through a logarithmic law of the wall assumption
which assumes that the instantaneous shear stress is always in phase with
the time averaged one [5,6]. This assumption, although shown to perform satisfactor-
ily, has no real physical basis. Such approaches usually leave too much to depend
on the subgrid scale model and, although the three-dimensional mecha-
nisms are included, important aspects of the flow are still modeled and not
simulated.
In the present paper, results will be presented from a two-dimensional large eddy
simulation of the quasi-two-dimensional turbulent flow past a square cylinder. The
momentum and continuity equations are solved on a collocated Cartesian grid using
the volume averaging approach of Schumann [5] with the standard Smagorinsky [7]
sub-grid-scale model and the SIMPLE [8] algorithm is used for pressure correction.
Care has been taken so that the spatial resolution is adequate, particularly near the
solid walls and no-slip boundary conditions have been implemented with no velocity
distribution assumed or implied. The results are compared to previous numerical
calculations and experimental measurements for the mean flow field and the fluctu-
ation energy of the periodic and turbulent motion which are separated by a filtering
procedure. The goal is to provide a fine enough grid in two-dimensional space to
resolve the near-wall regions which are most influential in this type of flow and to
directly simulate the quasi-two-dimensional mechanisms which are believed to be too
important to be left to the SGS model.

2. Numerical methodology

The numerical procedure employed here is based on the volume averaging


approach of Schumann [5], who proposes the spatial averaging of the Navier—Stokes
equations over volumes defined by the computational grid. Using Green’s theorem
in space this leads to the following form for the equation of continuity and the
D. Bouris, G. Bergeles/J. Wind Eng. Ind. Aerodyn. 80 (1999) 31–46 33

momentum equations for an incompressible fluid


4*(ou )
G "0NodK GuN "0,
*x G G
G
4*(ou ) 4 * 4 *p 4 *
 
*u *u *4uN
G# (ou u )"! # u G # H No G
*t *x G H *x *x *x *x *t
H G H H G
H *u H *u
H G H G  
#odK (HuN HuN )"!dK ( G p")#dK k
H  *x
G# H ,
H
*x
G
(1)

where i, j"1, 2 denote the Cartesian coordinate directions x, y and the overbar
denotes the spatial average over a grid volume (») or the surface average over a cell
face (i, j) of the volume. The Smagorinsky [7] sub-grid scale stress model (and the
modification of Lilly [9]) has been incorporated in the above equations so that
k "k#k with k being the turbulent viscosity given by the Smagorinsky—Lilly
 2 2
model as

ou u "ouN uN #2k SM #R d


G H G H 2 GH  II GH

 
1 *uN *uN
k "o(C D)"SM ""o(C D)(2SM SM ), SM " G# H , (2)
2 1 1 GH GH GH 2 *x *x
H G
where C is the sub-grid scale stress constant (taken here to be 0.1 as in Ref. [10]) and
1
D"(*x*y) defined by the grid spacing. The choice of a rather simple SGS model is
deliberate. There have been many indications that more complex models offer only
a slight improvement at the expense of computational resources. Breuer and Pourquie
[10] came to this conclusion for the type of flow being examined here when they
compared the Smagorinsky—Lilly model to a dynamic SGS model. In any case, the
fine resolution used in the present computation will leave only the very small scales to
depend on the SGS model.
dK in Eq. (1) is the finite-difference operator for the first derivative in central
H
differences in the j direction and the normal sub-grid scale stresses (R d ) have been
 II GH
incorporated into the pressure term to give pV. The final form of the equations is very
much similar to the form of the Reynolds averaged equations when finite volume
procedures are employed. Thus, the solution procedure used in those methodologies
can be adopted here. The standard SIMPLE algorithm [8] is employed for pressure
correction, the variables are collocated on the grid with the Rhie and Chow [11]
interpolation used for the cell face velocities. The discretization in time is a first-
order fully implicit Euler scheme which gives extra stability over an explicit scheme.
For upwind differencing of the convective terms, the BSOU scheme is used [12] which
is a fully bounded scheme of second-order accuracy. This scheme was chosen to
replace the hybrid central/upwind differencing scheme which introduced too much
numerical diffusion thus damping out any unsteady mechanisms that arose. The
option of using the QUICK third-order upwind differencing scheme was also ruled
out due to the unbounded nature of this scheme which often introduces over- and
under-shoots [2].
34 D. Bouris, G. Bergeles/J. Wind Eng. Ind. Aerodyn. 80 (1999) 31–46

It should be kept in mind that in large eddy simulations the calculation is


performed for the instantaneous velocity field at each time step and the sub-grid scale
model is responsible for the effects of the small scales which cannot be resolved by the
computational grid. This means that, for calculation of the time-averaged flow field,
calculations must be performed for a large number of time steps and then each
variable’s time series must be integrated over the calculated time period.

3. Calculation of turbulent flow past a square cylinder

The turbulent flow past a square cylinder has been numerically studied in two-
dimensions by Refs. [2,3] with various turbulence models incorporated in the solution
of the Reynolds averaged Navier—Stokes equations. Two- and three-dimensional large
eddy simulations of the same flow have been performed by Murakami et al. [13] and
Murakami and Mochida [4] with some of their results having been presented in Ref.
[3] while Breuer and Pourquie [10] also performed three-dimensional LES computa-
tions. The computations are validated against the experimental measurements of Lyn
[14—16] at Re"22 000 and Dura o et al. [17] at Re"14 000. Some of Lyn’s measure-
ments were also presented in Ref. [2]. The characteristic of this type of flow is its
quasi-two-dimensional character and the presence of periodic vortex shedding from
the front corners of the square rod which introduces a low-frequency variation of the
velocity field behind the rod in addition to the high-frequency turbulence fluctuations.
A two-dimensional large eddy simulation of the turbulent flow past a square
cylinder at Re"22 000 will be presented here and the numerical results will be
compared to experimental measurements and other numerical calculations. Particu-
lar interest will be focused on the splitting of the periodic and turbulent fluctuations of
the flow field since it has been shown [2] that turbulence models in the Reynolds
averaged Navier—Stokes equation approach (RANS) fail to correctly predict both at
the same time. The Cartesian grid (300;350) that was used in the calculation is shown
in Fig. 1. The size of the computational domain is the same as that used by Franke
and Rodi [2] and Rodi [3] and is shown in Fig. 1 (h is the length of the side of the rod).
The densest grid used by Franke and Rodi [2] was 186;156 with the closest grid line
to the wall being dy/h"0.00125. In the three-dimensional simulations of Murakami
and Mochida [4] the grid was 104;69;10 with the computational domain being 2h
in the third direction and the closest grid line to the wall being dy/h"0.022. In Ref.
[10] the densest grid was 146;146;20 with dy/h"0.01 and the computational
domain in the third direction was 4h. In the present grid (300;350) the grid line
closest to the wall is at a distance of dy/h"4;10\, much closer than in any of the
previous calculations. It is noted that the grid used here is denser in the two-
dimensional plane than those used in all the other computations but has only half the
total number of grid points than the grid used for three-dimensional LES by Breuer
and Pourquie [10]. The computational advantages are obvious.
Franke and Rodi [2] used two turbulence models, the k—e model and the Reynolds
stress equations model while for the wall boundaries both wall functions and the two
layer approach was used. The SIMPLE algorithm was employed to the staggered
D. Bouris, G. Bergeles/J. Wind Eng. Ind. Aerodyn. 80 (1999) 31–46 35

Fig. 1. Cartesian computational grid (300;350) used in the present study for large eddy simulation of the
turbulent flow past a square rod.

layout of variables on the grid and the third-order accurate, but unbounded, QUICK
upwind scheme was used for the convection terms. Time discretization was first-order
accurate fully implicit according to the Euler scheme. For the three-dimensional LES
simulations of Murakami and Mochida [4] (and in Ref. [3]) and Ref. [10] the linear
and 1/7 power-law velocity distribution near the wall was adopted according to
Werner and Wengle [6]. Beruer and Pourquie [10] also used a collocated arrange-
ment of variables on the grid but used central differences without an upwind scheme
for the convection terms and the second-order accurate explicit Adams—Bashforth
scheme for time discretization of the momentum equations. The SIMPLE algorithm
was used for pressure correction and the whole methodology was applied using
multigrid techniques to accelerate convergence. The numerical methodologies whose
results will be compared to the present calculation have much in common with the
one presented here. Here, upwind differencing is second-order accurate but fully
bounded while time differencing is first-order accurate but fully implicit. Both these
differencing schemes allow good stability during solution. The numerical techniques
that are used are quite common with no real uncertainties arising from them. A major
difference in the numerical approaches is that in the present calculation the no-slip
boundary conditions are applied at the solid walls so the whole boundary layer will be
resolved. This is a step away from the velocity distribution assumptions that were
previously imposed in both RANS and LES. In LES the velocity distributions found
36 D. Bouris, G. Bergeles/J. Wind Eng. Ind. Aerodyn. 80 (1999) 31–46

for the mean velocity field are extended to the instantaneous flow field and the viscous
sublayer is not resolved by the grid. However, most major structures in the flow are
directly related to the boundary layer along the walls so the correct resolution of the
sublayer is important, especially when the boundary layer undergoes transition,
separation and is unsteady in time, as it is in the present case.
In all the above numerical calculations as well as the one presented here, inlet
conditions were a uniform, non-fluctuating velocity profile, symmetry conditions were
applied to upper and lower boundaries while at the outlet zero gradient conditions
(fully developed flow) were imposed. The time step used in the present calculation is
dt"3;10\ which in non-dimensional form is dq"dt/¹"º dt/h"0.01. In Fig. 2

the time series of the vertical velocity component at the centerline one rod length
behind the rod (x/h"2) as well as the time series of the rod lift coefficient is shown.
The calculation is found to become completely independent of the initial conditions
after about 200 non-dimensional time units. This is determined from the fact that the
dominant frequencies of the lift coefficient and the velocity behind the rod become
equal for any time series taken after q"200. The frequency spectrums for
q"230—395 are shown in Fig. 3 and give a Strouhal number of Sth"fh/º "0.134.

The time period q"230—395 is also the one over which the time-averaged quantities
are calculated from integration in time. Calculations were performed on a Hewlett-
Packard computer with an R8000 processor (peak performance 210 SPECfp) and
about 1.5 min of CPU time were required per time step. Convergence to 0.5% was
attained for all equations at each time step in under 50 iterations.
Table 1 shows the prediction of the dominant vortex shedding frequency in non-
dimensional form as predicted by the turbulence models employed by Franke and
Rodi [2] as well as the results from the present large eddy simulation. The results of
the large eddy simulation are in good agreement with the experimental measurements
while, among the other turbulence models, only the Reynolds stress equations with

Fig. 2. Time evolution of the vertical velocity component at x/h"2, y/h"0 and of the rod lift coefficient.
D. Bouris, G. Bergeles/J. Wind Eng. Ind. Aerodyn. 80 (1999) 31–46 37

Fig. 3. Frequency spectrum from (a) the vertical velocity component at x/h"2, y/h"0 and (b) from the
rod lift coefficient corresponding to q"230—395.

Table 1
Numerical calculation of the Strouhal number and the Drag coefficient from various turbulence models [2]
and present LES calculations and comparison with experimental measurements [16,17]

k—e (2 layer) Reynolds stress Reynolds stress LES (present) Experimental


equation equation
(wall functions) (2 layer)

Sth 0.124 0.136 0.159 0.134 0.132$0.004 Ref. [16]


0.139 Ref. [17]
C 1.179 2.15 2.43 2.18 2.05—2.23
"

wall functions are in agreement with the experiments. It is noted that Franke and Rodi
[2] arrived at a steady solution when the k—e model was used with wall functions with
no vortex shedding arising. As can be seen from Table 1, the resolution or modeling
technique of the near wall region was found to play an important role in the RANS
modeling while Breuer and Pourquie [10] also makes the same remark regarding the
near-wall resolution in LES. It should be noted that in the present large eddy
simulation the no-slip boundary condition was applied with the instantaneous
y>"y(q /o) /l being as low as 0.01 for the near-wall grid cells. The calculated mean

drag coefficient is also found to be in good agreement with experimental measure-
ments and it is noted that Lyn et al. [16] measured a value of C &2.1 (4% difference
"
from present calculated value of 2.18) from integration of the momentum flux in the
wake at x/h"8.5. The oscillation of the lift coefficient can be seen in Fig. 2 and the
calculated standard deviation is ( CI "1.62. Iso-vorticity regions from the present
*
38 D. Bouris, G. Bergeles/J. Wind Eng. Ind. Aerodyn. 80 (1999) 31–46

Fig. 4. Iso-vorticity (1/s) regions in the turbulent flow past a square cylinder. (Re"22 000.)

large eddy simulation are shown in Fig. 4. A vortex pairing can be observed at x/h&9
while near the rod a vortex structure is just being shed. As seen in Fig. 3 there are
frequency variations of significant amplitude near the dominant vortex shedding
frequency and these will interact with the shedding frequency to produce the vortex
pattern seen in Fig. 4.
Time-averaged results for q"230—395 are obtained from integrating the data
obtained over this time period and the mean centerline velocity is presented in Fig. 5.
It can be seen that the k—e predictions show an unrealistically large recirculation zone
while the Reynolds stress models show better behavior. However, the free stream
velocity is predicted to be attained too fast by the Reynolds stress models as well as the
large eddy simulations. In the near-wake region, the 3D LES of Murakami et al. (in
Ref. [3]) seems to agree with measurements slightly better than the present 2D
calculation but further downstream the present calculation is closer to the experi-
mental measurements, probably due to the denser grid. Altogether though, the
differences between the two calculations are minimal for the mean centerline velocity
variation. The 3D LES calculations of Breuer and Pourquie [10] for the mean velocity
on the centerline behind the cylinder show very good agreement with experiments up
to x/d"2.5. However, they also predict too slow recovery of the flow resulting in an
overprediction of the mean velocity from x/d"3 and on (for the Re"22 000 which is
the flow being examined). Breuer and Pourquie [10] stress the importance of their fine
resolution near the wall and when comparing their results to others’ conclude that this
D. Bouris, G. Bergeles/J. Wind Eng. Ind. Aerodyn. 80 (1999) 31–46 39

Fig. 5. Time averaged (mean) velocity along centerline behind rod. Comparison with other numerical and
experimental results.

is the most important factor. If one observes the results of the other 3D LES
calculations [4] which use a coarser grid, these are at the same level of accuracy as the
2D calculation presented here but, as will be shown later, they underestimate the
fluctuating energy. One can come to the conclusion that a 3D LES on its own does not
guarantee good results, fine resolution in the two-dimensional plane is still very
important.
By subtracting the mean velocity from the instantaneous one the total (periodic
# turbulent) fluctuation is obtained and from this, the total two-dimensional
fluctuation kinetic energy can be calculated from the two velocity components
k"(u#v)/2 (Fig. 6). As previously mentioned the standard k—e model (with wall
functions) yielded a steady-state solution in Ref. [2] and therefore comparisons are
made with the other turbulence model variants. Surprisingly, the 2D LES calculation
shows a closer agreement to the experimental measurements than the 3D calculation
of Murakami et al., [3]. The Reynolds stress models show an even better agreement
but this will be further investigated from the centerline variation of the turbulent
fluctuations in Fig. 7. Unfortunately, Breuer and Pouquie [10] did not present the
centerline distribution of this quantity and no comparison can be made. In order to
calculate the turbulent kinetic energy from a large eddy simulation of a flow with large
scale periodicity it is necessary to remove the periodic component from the velocity
time series. This is accomplished from filtering the velocity in frequency space.
A Fourier transform is performed, the dominant frequency is determined and then
40 D. Bouris, G. Bergeles/J. Wind Eng. Ind. Aerodyn. 80 (1999) 31–46

Fig. 6. Total fluctuation energy (periodic#turbulent) of velocity along centerline behind square rod.
Comparison with other numerical and experimental results.

filtered out of the frequency spectrum. An inverse Fourier transform will give the
velocity time series with only the turbulent fluctuations present. The filter that is used
here is the notch filter [18] which removes all frequencies in a narrow region u around
the main frequency f :

w!w
H ( f )"  , w"tan (pfdt), (3)
, (w!iuw )!w
 
where u is defined as a percentage of the main frequency w . Multiplying the Fourier

transform of the velocity time series by H ( f ), the filtered frequency spectrum
,
results.
The turbulence kinetic energy in Fig. 7 is best predicted by the 2D LES presented
here. It is important to observe that the Reynolds stress models significantly under-
estimate the turbulence kinetic energy which when combined with the good results
shown in Fig. 6 for the total fluctuation energy means that they must overpredict the
energy of the periodic motion of the flow behind the rod.
In Fig. 8 the prediction of the fluctuation energy (total, periodic and turbulent) by
the 2D LES presented here is compared with the experimental measurements of Lyn
(in Ref. [2]). The overall agreement is good but the turbulent kinetic energy seems to
be underpredicted in the region near the back of the rod, while the total fluctuation
energy is well predicted for the same region. This could mean that the periodic motion
D. Bouris, G. Bergeles/J. Wind Eng. Ind. Aerodyn. 80 (1999) 31–46 41

Fig. 7. Turbulent kinetic energy along centerline behind square rod. Comparison with other numerical and
experimental results.

is overpredicted but it should be kept in mind that the sub-grid scale energy has not
been added here and this is expected to slightly raise the turbulence energy level.
Generally though, in the present computation, the SGS model has only a minor
influence on the mean flow and statistical turbulence quantities since the turbulence
viscosity is of the order of 3—4 times the fluid viscosity value. In comparison, for RANS
calculations, depending on the flow, the turbulent viscosity could be two or more
orders of magnitude larger than the fluid viscosity. This indicates that most of the
structures present in the flow are simulated leaving little to be modeled by the SGS
model. A more advanced dynamic sub-grid scale model has been considered by
Breuer and Pourquie [10] which incorporates a sub-grid scale coefficient defined by
the local flow instead of a constant one all over the flow. However, although they used
a coarser grid, no clear superiority of either of the two could be discerned for the flow
past a surface mounted cube.
Since most of the structures are simulated there could be a problem regarding the
three-dimensional effects, which are missing. Actually, in the immediate vicinity
&0.25h behind the rod there is an overprediction of total fluctuation energy which
could be interpreted as an overprediction of the periodic motion and might be
attributed to the absence of the simulation of three-dimensional effects. This causes all
mechanisms to appear in the two resolved dimensions as two-dimensional structures
which add to the periodic motion. This has also been noted by Tamura et al. [19] who
performed two- and three-dimensional simulations of the turbulent flow past circular
42 D. Bouris, G. Bergeles/J. Wind Eng. Ind. Aerodyn. 80 (1999) 31–46

Fig. 8. Kinetic energy (total, periodic and turbulent) along centerline behind square rod. Comparison with
experimental results.

cylinders. When performing two-dimensional calculations, the flow behaved like it


was “confined” and the three-dimensional structures appeared as strong secondary
vortices in the other two-dimensions. The three-dimensional effects seem to be most
important in the recirculation region behind the cylinder where the 3D calculations
with relatively fine resolution [10] gave the better prediction of the near-wake region
(Fig. 5).
The non-dimensional autocorrelation function of the fluctuating u velocity com-
ponent at a given point in the flow is defined as the correlation between the velocity at
two different instants in time:

u(xo ,t)u(xo ,t#q)


R(q)" . (4)
u(xo ,t)

The autocorrelation function gives important information regarding the correlation


of the fluctuating velocity component in time. The point at which the function first
reaches zero is considered to be a measure of the integral time scale of the flow at the
position where the function has been calculated. The autocorrelation function at three
different positions on the centerline behind the square rod has been calculated and is
presented in Fig. 9.
D. Bouris, G. Bergeles/J. Wind Eng. Ind. Aerodyn. 80 (1999) 31–46 43

Fig. 9. Autocorrelation function R(q) of the u velocity component at three centerline positions behind the
square rod.

It is first noted that the function at position x/h"5 shows a clear periodicity which
is actually the effect of the large-scale motion of the vortex shedding. At the positions
closer to the cylinder the smaller-scale turbulent motion appears as fluctuations in the
correlation. This is expected since the kinetic energy of the fluctuating motion of the
velocity is at a peak around x/h"2—3 and has decreased by a factor of two at x/h"5
(see Fig. 6). From the periodic motion at x/h"5 one can find a mean time difference
between two peaks in the autocorrelation function at about q/¹"8. This can
be considered as the time between two consecutive passings of vortices from this
position. Given a calculated mean velocity of º"0.79º at this position (see Fig. 5)

one can calculate the distance between two consecutive vortices as ¸ /h"

0.79*8"6.32 which can be verified from Fig. 4. Lyn et al. [16] found the separation
of successive vorticity peaks in the near-wake region to be 5.8h and thus the present
calculation is quite close to their measured value. The difference of 9% can be partly
attributed to the overprediction of the mean velocity at x/h"5 on the centerline
(Fig. 5) since the vortex shedding frequency (Strouhal) number has been well predicted
(0.134 compared to 0.132$0.004 of Lyn et al. [16]).
The overall agreement of the 2D LES performed here has proven to be much better
than any of the RANS models when the major quasi-two-dimensional mechanisms of
44 D. Bouris, G. Bergeles/J. Wind Eng. Ind. Aerodyn. 80 (1999) 31–46

the flow and the statistical turbulence quantities are examined. The most difficult
characteristics to be calculated by the RANS models were the levels of periodic and
turbulent fluctuation energy and these have been satisfactorily calculated here. There
are certain discrepancies, some of which are attributed to the absence of three-
dimensional effects (most important in the near-wake region) but it seems that the fine
resolution used in the 2D plane is much more influential than the use of a third spatial
direction with a coarser grid. In fact a 3D calculation that was performed in this way
[4] showed inferior results to the 2D LES since it underpredicted the fluctuation
energy. In any case, another 3D calculation [10] provided the best results as far as the
mean flow was concerned and this was attributed to the fine resolution in the 2D
plane, especially near the walls. It seems that a fine resolution for the quasi-two-
dimensional plane must first be satisfied since the two-dimensional mechanisms are
still more important to the major structures of the flow than the three-dimensional
effects. This is the most important point of the paper.

4. Conclusions

A large eddy simulation has been performed for the turbulent flow past a square
cylinder. The flow is characterized by quasi-two dimensionality with vortex shed-
ding from the front corners of the square rod and has been experimentally studied
in the past. Other large eddy simulations of the same flow have appeared in the
literature and there have been claims that large eddy simulations should only
be performed in three dimensions since two-dimensional computations fail to cor-
rectly represent the three-dimensional character of turbulent structures. In the
present paper, it is shown that two-dimensional large eddy simulations should
not be dismissed so easily. A two-dimensional large eddy simulation has been
performed with no slip conditions imposed at the walls. This condition is supported
by an increased spatial resolution near the wall, closer to the wall than any
of the previously published large eddy simulations for the same flow. Comparisons
of the present simulation with previously published numerical computations are
presented.
When examining the large-scale mechanisms of the quasi-two-dimensional flow, the
computations performed here show results superior to any of the turbulent models
used in the solution of the Reynolds averaged Navier—Stokes equations [2]. This is
evident both from the close agreement with experimental measurements of the
predicted shedding frequency, the drag coefficient and the distribution of the kinetic
energy of turbulent and periodic velocity fluctuations on the centerline behind the rod.
Filtering has been performed in frequency space in order to remove the periodic
velocity component and to predict the kinetic energy of the turbulent fluctuations.
Autocorrelation functions were also computed in the present work giving an indica-
tion of the time scales present along the centerline behind the rod. From these
functions, the distance between two consecutive vortices was calculated and was
found to agree within 9% with the experimentally measured value. The fine resolution
that was implemented lead to only a small part of the small scales being modeled by
D. Bouris, G. Bergeles/J. Wind Eng. Ind. Aerodyn. 80 (1999) 31–46 45

the SGS model and to a good overall representation of the mean flow and periodic
and turbulence fluctuation energy.
Large eddy simulations admittedly require denser grids and more computer time
than Reynolds averaged approaches but in certain cases, such as the one under
consideration here, it seems that the Reynolds averaged approaches fail to predict
important aspects of the flow. However, when performing a 3D LES, computer
limitations might confine the calculation to a coarse resolution in the 2D plane. The
resolution in this plane is considered to be very important and should be given
priority over the 3D effects when examining the quasi-two-dimensional flow and the
statistical properties of turbulence. Nevertheless, it should be stressed that from
a physical point of view, one should aim at a 3D LES with 2D LES being but
a compromise imposed by computer limitations.

References
[1] J. Ferziger, Approaches to turbulent flow computation: applications to flow over obstacles, J. Wind
Eng. Ind. Aerdyn. 35 (1990) 1—19.
[2] R. Franke, W. Rodi, Calculation of vortex shedding past a square cylinder with various turbulence
models, in: Proc. 8th Symp. Turbulent Shear Flows, 9—11 September 1991, Tech. Univ. Munich,
Springer Berlin, 1991, pp. 189—204.
[3] W. Rodi, On the simulation of turbulent flow past bluff bodies, J. Wind Eng. Ind. Aerodyn. 46&47
(1993) 3—19.
[4] S. Murakami, A. Mochida, On turbulent vortex shedding flow past a square cylinder predicted by
CFD, J. Wind Eng. Ind. Aerodyn. 54 (1995) 191—211.
[5] U. Schumann, Subgrid scale model for finite difference simulations of turbulent flows in plane
channels and annuli, J. Comp. Phys. 18 (1975) 376—404.
[6] H. Werner, Wengle, Large eddy simulation of turbulent flow over and around a cube in a plate
channel, in: Proc. 8th Symp. Turbulent Shear Flows, Tech. Univ. Munich, 9—11 September 1991,
Springer, Berlin, 1991, pp. 155—168.
[7] J. Smagorinsky, Mon. Weather Rev. 91 (1963) 99—164.
[8] S.V. Patankar, D.B. Spalding, A calculation procedure for heat, mass and momentum transfer in three
dimensional parabolic flows, Int. J. Heat Mass Transfer 15 (10) (1972) 1787.
[9] D. Lilly, The representation of small scale turbulence in numerical simulation experiments, in: H.H.
Goldstine (Ed.), Proc. IBM Scientific Computing Symposium on Environmental Sciences, IBM Form
No. 320-1951, 1967, pp. 195—210.
[10] M. Breuer, M. Pourquie, First experiences with LES of flows past bluff bodies, in: W. Rodi, G.
Bergeles (Ed.), Proc. 3rd Int. Symp. on Engineering Turbulence Modelling and Measurements,
Heraklion-Crete, Greece, 27—29 May 1996, Elsevier, Amsterdam, 1996, pp. 177—186.
[11] C. Rhie, W. Chow, Numerical study of the turbulent flow past an airfoil with trailing edge separation,
AIAA J. 21 (1983) 1525—1532.
[12] G. Papadakis, G. Bergeles, A locally modified second order upwind scheme for convection terms
discretisation, Int. J. Num. Meth. Heat Fluid Flow 5 (1995) 49—62.
[13] S. Murakami, A. Mochida, Y. Hayashi, S. Sakamoto, Numerical study on velocity pressure field and
wind forces for bluff bodies with k—e, ASM and LES, in: Proc. 8th Int. Conf. on Wind Engineering,
London, Canada, 1991.
[14] D. Lyn, Phase averaged turbulence measurements in the separated shear layer region of flow around
a square cylinder, in: Proc. 23rd Congress of the Int. Association for Hydraulic Research, Ottawa,
Canada, 21—25 August 1989, A85—A92.
[15] D. Lyn, W. Rodi, The flapping shear layer formed by flow separation from the forward corner of
a square cylinder, J. Fluid Mech. 267 (1994) 353—376.
46 D. Bouris, G. Bergeles/J. Wind Eng. Ind. Aerodyn. 80 (1999) 31–46

[16] D. Lyn, S. Einav, W. Rodi, J. Park, A laser doppler velocimetry study of ensemble-averaged
characteristics of the turbulent near wake of a square cylinder, J. Fluid Mech. 304 (1995) 285—319.
[17] D. Dura o, M. Heitor, J. Pereira, Measurements of turbulent and periodic flows around a square
cross-section cylinder, Exp. Fluids 6 (1988) 298—304.
[18] W. Press, B. Flannery, S. Teukolsky, W. Vetterling, Numerical Recipies, Cambridge University Press,
Cambridge, 1987.
[19] T. Tamura, I. Ohta, K. Kuwahara, On the reliability of two-dimensional simulation for unsteady
flows around a cylinder-type structure, J. Wind Eng. Ind. Aerodyn. 35 (1990) 275—298.

You might also like