You are on page 1of 16

Journal of Wind Engineering

and Industrial Aerodynamics 88 (2000) 231–246

Free end effects on the near wake flow structure


behind a finite circular cylinder
Cheol-Woo Park, Sang-Joon Lee*
Department of Mechanical Engineering, Pohang University of Science and Technology,
Pohang 790-784, South Korea

Abstract

The free end effect on the near wake of a finite circular cylinder in a cross flow has been
investigated experimentally. Three finite cylinders with aspect ratios (L=D) of 6, 10 and 13
were tested in a subsonic wind tunnel at a Reynolds number of 20 000. A hot-wire anemometer
was employed to measure the wake velocity. Mean pressure distributions on the cylinder
surface were also measured. The flow near the free end was visualized to observe the flow
structure qualitatively in a circulating water channel. The experimental results from these finite
cylinder (FC) models were compared with those of a two-dimensional circular cylinder. The
flow past the FC free end shows a complicated three-dimensional wake structure. As the FC
aspect ratio decreases, the vortex shedding frequency is decreased and the vortex formation
region is elongated. The free end effect becomes dominant close to the FC free end. The three-
dimensionality of the FC wake may be attributed mainly to the strong entrainment of
irrotational fluids, caused by the downwash of counter-rotating vortices separated from the
FC free end. The downwash flow is concentrated in the central region of the wake. A peculiar
flow structure having a 24 Hz frequency component was observed near the free end using
spectral analysis and cross-correlation of the velocity signals. This 24 Hz frequency component
is closely related to the counter-rotating twin vortices formed near the FC free end.
# 2000 Elsevier Science Ltd. All rights reserved.

Keywords: Finite cylinder; Free end; Aspect ratio

*Corresponding author. Tel.: +82-54-279-2169; fax: +82-54-279-3199.


E-mail address: sjlee@postech.ac.kr (S.-J. Lee).

0167-6105/00/$ - see front matter # 2000 Elsevier Science Ltd. All rights reserved.
PII: S 0 1 6 7 - 6 1 0 5 ( 0 0 ) 0 0 0 5 1 - 9
232 C.-W. Park, S.-J. Lee / J. Wind Eng. Ind. Aerodyn. 88 (2000) 231–246

1. Introduction

Flow around a circular cylinder has been extensively investigated in the past due
to its simple geometry and coherent vortex structure. A circular cylinder is a
representative bluff body, showing two-dimensional (2-D) flow characteristics. At
low Reynolds numbers, three-dimensional (3-D) flow characteristics consisting of
altered flow structures in the axial direction are frequently observed in the vortices
shed from a 2-D circular cylinder.
This three dimensionality has been attributed to non-uniformities that exist in the
flow and along the body span, and to the particular end constraints imposed in the
experimental arrangement. The various end constraints on the wake structure have
been investigated by Slaouti and Gerrard [1].
Williamson [2] found a discontinuity in the Strouhal–Reynolds number relation-
ship for oblique vortex shedding at low Reynolds numbers. He suggested that the
critical Reynolds number could be affected by differences in the flow non-uniformity,
end conditions, or the free stream turbulence level.
In the strict sense, the cylinder wake has 3-D flow characteristics over all Reynolds
number ranges. In the near-wake region behind a 2-D circular cylinder, however, the
three dimensionality is weak, because vortex shedding is regular and parallel to the
cylinder axis in the vortex formation region. Therefore, the cylinder wake has been
assumed a 2-D flow in view of the coherent vortex structure. The Strouhal number
does not change along the cylinder axis if appropriate end plates and a small aspect
ratio are used (see Ref. [3]). Baban et al. [4] demonstrated that for a 2-D cylinder in a
cross-flow, vortex shedding is truly 2-D, and the strength of the roll-up vortices is
fairly uniform along the span.
The main causes for flow three dimensionality include the non-uniformity of free
stream flow, the presence of longitudinal vortices and a low aspect ratio (L=D) of the
cylinder. The aspect ratio is the ratio of cylinder height L to diameter D. The free end
of a finite cylinder (FC) is a direct and significant factor for flow three
dimensionality. Many high-rise bluff bodies and tall buildings can be simplified as
a finite cylinder with a free end. The free end changes the flow structure in the near
wake, including the vortex formation region, vortex-shedding pattern and surface
pressure distribution. Parallel vortex shedding can be introduced through modifica-
tion of the cylinder ends, as shown by Eisenlohr and Eckelmann [5].
Wieselsberger [6] investigated the 3-D flow characteristics of a finite circular
cylinder mounted on a flat plate. The drag force acting on the cylinder decreased at
small aspect ratios. Baban et al. [7] observed an increase in drag force fluctuations
due to highly turbulent re-circulation flow in the wake region, especially in the shear
layer separated from the end of the cylinder. Budair et al. [8] found that vortex
shedding disappeared at a Reynolds number Re=15 000 when the FC aspect ratio
(L=D) was lower than 7. Sakamoto and Arie [9] revealed that the vortex-shedding
pattern was largely dependent on the FC aspect ratio. Okamoto and Sunabashiri [10]
found that the wake behind finite cylinders of small aspect ratio (L=D ¼ 1–2) was
symmetric, but that the wake pattern became 3-D when the aspect ratio was larger
than L=D ¼ 4.
C.-W. Park, S.-J. Lee / J. Wind Eng. Ind. Aerodyn. 88 (2000) 231–246 233

However, there are still some contradictions about the three dimensionality of the
FC wake, especially for high Reynolds number flows. Farivar [11] investigated the
effect of the FC free end on the mean pressure, pressure fluctuation, and drag force
acting on a cylinder when exposed to a uniform flow. He observed that the vortex-
shedding frequency disappeared for aspect ratios smaller than L=D ¼ 7:5. On the
other hand, Zdravkovich et al. [12] noted vortex shedding in the FC wake around an
aspect ratio of L=D ¼ 2, although it was irregular and intermittent.
In previous studies, the flow past a finite cylinder at high Reynolds numbers,
particularly in the vicinity of the free end, was not fully investigated. Therefore, this
study investigates the effect of FC aspect ratio on near-wake flow characteristic,
especially near the free end.

2. Experimental apparatus and methods

The experiments were carried out in a closed-return-type subsonic wind tunnel,


with a test section of 0.72 m wide  0.6 m high  6 m long. Free-stream turbulence
intensity in the test section was less than 0.08%. Free-stream velocity was fixed at
10 m/s, and the corresponding Reynolds number based on the cylinder diameter
(D ¼ 30 mm) was 20 000. A schematic diagram of the experimental setup and
coordinate system is shown in Fig. 1.
In this study, three FC models with different aspect ratios (L=D ¼ 6; 10; 13) were
tested. For comparison, a 2-D circular cylinder with an aspect ratio of L=D ¼ 17:3,
with no gap between the free end and test section ceiling, was also tested. The
experimental models were made of stainless-steel rod and their surfaces were
polished smooth using sandpaper. The finite cylinder was installed vertically on a 15-
mm-thick flat plate, with a sharp-edged leading edge of angle 308. The FC model was
placed 0.5 m downstream from the leading edge of the flat plate. In order to avoid
flow-induced vibrations, the natural frequencies of the FC models were set to be
greater than 20 times the vortex-shedding frequency.
The boundary layer that developed on the sharp-edged flat plate was about
4.1 mm thick at the location of the cylinder. A horseshoe vortex could be formed at
the junction between the FC models and the ground plate. However, since the
measurements were performed in the upper half of the wake, from mid-height to the
free end of the cylinder, the effects of the wall boundary layer and horseshoe vortex
were negligible.
West and Apelt [13] found that the pressure distribution around a circular cylinder
varies only slightly with respect to blockage ratio. For blockage ratios of less than
6%, the Strouhal number is independent of the blockage ratio and the aspect ratio of
the cylinder. Since the maximum blockage ratio for the 2-D cylinder model was
4.2%, the blockage effect has not been considered in this study.
The wake velocity was measured using an I-type hot-wire probe (DANTEC
55P11) connected to a constant temperature hot-wire anemometer (TSI IFA 100).
The hot-wire probe was maneuvered to the measuring points using a 3-D traverse
system with an accuracy of 0.01 mm. At each measurement point, 32 000 velocity
234 C.-W. Park, S.-J. Lee / J. Wind Eng. Ind. Aerodyn. 88 (2000) 231–246

Fig. 1. Experimental model and coordinate system (unit: mm).

data were acquired at a 2 kHz sampling rate after low-pass filtering at 800 Hz.
During the experiments, the temperature variation in the wind tunnel test section
was less than 0:58C. A schematic diagram of the velocity measurement system is
shown in Fig. 2.
To measure pressure distributions around the cylinder surface, pressure taps were
installed along the longitudinal axis of the cylinder at 5-mm intervals. Pressure
distributions were measured by rotating the cylinder in 108 increments. The pressure
taps were connected to a micromanometer (FCO-19), and the analog pressure signals
were digitized using a high-precision A/D converter (DT-2838). At each measure-
ment point, 16 384 pressure data were acquired using a 500 Hz sampling rate after
low-pass filtering at 200 Hz. A time delay of a few seconds was allowed, so that the
pressure could recover from the fluctuations after each channel scanning.
The pressure difference between the surface pressure, p, and the static pressure, p0 ,
was divided by the dynamic pressure to give the pressure coefficient, Cp , expressed as
p ÿ po
Cp ¼ ; ð1Þ
1=2ra U02

where U0 is the free-stream velocity and ra is the air density. In order to provide flow
visualization of the turbulence flow structures around the FC free end, a particle
tracer method was employed in a circulating water channel with a test section of
0:3 W  0:2 H  1:2 L ðm3 Þ. Polyvinylchloride particles with an average diameter of
100 mm were seeded as tracer particles. The particle path lines near the free-end were
illuminated with a thin cold light sheet, emitted from a 150 W halogen lamp. The
scattered particle images at a free-stream velocity of 12 cm/s were photographed with
a Nikon F5 camera.
C.-W. Park, S.-J. Lee / J. Wind Eng. Ind. Aerodyn. 88 (2000) 231–246 235

Fig. 2. Schematic diagram of the velocity measurement system.

3. Results and discussion

3.1. Flow visualization

The flow around the FC free end was visualized using a particle tracer method in a
circulating water channel. Fig. 3 shows the visualized flow around the finite cylinder
of L=D ¼ 5. The light sheet illuminates the vertical XZ-plane passing through the
centerline of the cylinder wake. Near the ground, the wake shows a wavy structure,
which may be induced from the horseshoe vortex that is generated from the junction
of the finite cylinder and the flat ground plate. The vortices formed at the junction
persist downstream, but their effects are limited near the wall. In addition, the size of
the horseshoe vortex relative to the cylinder height decreases as the cylinder height
increases (see Ref. [10]).
The approaching flow moves upward, accelerates near the free end, and then
separates from the cylinder circumference at the free end. In the central wake plane,
the separated shear layer is declined and downwashed up to the FC mid-span along
the cylinder axis. In the wake region behind the lower half of the finite cylinder, the
vortices that are shed from the two sides of the cylinder are not influenced much by
the downwash flow separated from the FC free end.
The downwash flow in the central plane may be caused by the counter-rotating
pairs of vortices, generated from the FC free end. Fig. 4 shows the cross-sectional
particle images taken of the flow around the free end of the L=D ¼ 5 finite cylinder.
The vertical cross sections (YZ-plane) at X=D ¼ 0 and X=D ¼ 0:6 were illuminated
236 C.-W. Park, S.-J. Lee / J. Wind Eng. Ind. Aerodyn. 88 (2000) 231–246

Fig. 3. Visualized flow around a finite circular cylinder placed vertically on a ground (L=D ¼ 5, side view).

Fig. 4. Cross-sectional flow visualization near the free end of a finite cylinder (L=D ¼ 5).
C.-W. Park, S.-J. Lee / J. Wind Eng. Ind. Aerodyn. 88 (2000) 231–246 237

with a light sheet and the flow images were photographed from a downstream
location. A counter-rotating vortex pair is visualized above the FC free end. The left-
and right-side vortices rotate in a clockwise and counter-clockwise direction,
respectively, similar to the tip vortices formed along the leading edge of a delta wing
at incidence. The two vortices are of similar size, and are nearly symmetric with
respect to the central plane (Y=D ¼ 0) of the wake.
The counter-rotating twin-vortex is formed by the roll-up motion of the shear flow
separated from the edges of the FC free end. After separating from the free end, the
flow descends along the central section of the wake. The size of the swirling vortices
increases at the downstream location of X=D ¼ 0:6, as shown in Fig. 4(b). At
X=D ¼ 0:6, the vortices expand laterally and move slightly downwards. This is
caused by the entrainment of ambient fluids. The downwashing flow caused by the
counter-rotating vortices interacts with the vortices shed from the two sides of the
cylinder in the upper half of the near-wake region, as shown in Fig. 3. The
descending downwash flow entrains large amounts of ambient fluids.
Some typical top views of flow near the free end of the L=D ¼ 5 cylinder are
shown in Fig. 5. Flow images near the FC free end were photographed by shining the
horizontal light sheet downward from Z=L ¼ 1:066 to Z=L ¼ 0:9.
All visualized flow images are nearly symmetric with respect to the central section
(Y=D ¼ 0) of the wake. The oncoming flow ascends in front of the FC free end and
then descends downward just behind the cylinder. The shear flow separated from the
FC free end descends along the central region of the wake. Figs. 5(d)–(f ) show the
swirling vortices at the end of the vortex formation region. These are in phase (see
the discussion on the spectral analysis in Section 3.4, and Fig. 12).

3.2. Vortex shedding frequency

The velocity signals from the wake were analyzed to investigate the effect of the
FC aspect ratio (L=D) on the vortex shedding frequency. Fig. 6 shows the power
spectral density (PSD) distributions measured at X=D ¼ 3; Y=D ¼ 2 and Z=L ¼ 0:5.
The Strouhal number for the 2-D cylinder is 0.198, which corresponds well with
previous results.
Since the vortex shedding frequency was measured at the mid-height of the
cylinder, the effects of the wall boundary layer and the horseshoe vortex on the
shedding frequency can be considered to be very small.
The vortex shedding frequency decreases as the FC aspect ratio decreases;
the peak frequencies were 66 Hz (2-D), 60 Hz (L=D ¼ 13), 53 Hz (L=D ¼ 10) and
47 Hz (L=D ¼ 6). The peak amplitude of the PSD also decreases with decreasing FC
aspect ratio. The decreased vortex shedding frequency seems to be related to the
inflow of ambient inviscid flow into the near-wake region, due to the descending
downwash flow. Since the separated shear flow influences the near-wake up to a
certain length from the end of the cylinder, its effect will be enhanced at the mid-
height of cylinders with decreasing FC aspect ratio. This will be discussed later in
more detail.
238 C.-W. Park, S.-J. Lee / J. Wind Eng. Ind. Aerodyn. 88 (2000) 231–246

Fig. 5. Top views of flow near the FC free end by moving the height of horizontal light sheet (L=D ¼ 5).

3.3. Vortex formation region

The entrainment of ambient fluids into the wake region affects not only the wake
structure, but also the vortex formation behind the finite cylinder. In this study, the
flow velocity was measured along the centerline of the cylinder wake. The distance
from the cylinder to the peak location of velocity fluctuations was used to detect
where the roll-up of the shear layer took place.
A hot-wire anemometer was employed to measure the length of the vortex
formation region. Although the hot-wire technique does not provide a precise
velocity value in a highly turbulent 3-D flow, several researchers have measured the
C.-W. Park, S.-J. Lee / J. Wind Eng. Ind. Aerodyn. 88 (2000) 231–246 239

Fig. 6. Vortex shedding frequency measured at X=D ¼ 3; Y=D ¼ 2 and Z=L ¼ 0:5.

vortex formation region successfully by detecting the peak value of the fluctuating
velocity signal (see, for example, Ref. [3]). The velocity signals measured along the
centerline of the wake were band-pass-filtered at double the value of the vortex
shedding frequency. The turbulence intensity of the streamwise velocity component
was then calculated to detect the vortex formation length. In this study, the end of
the vortex formation region was defined as the point that had the maximum amount
of streamwise velocity fluctuation.
Fig. 7 shows the turbulence intensity distributions measured at the mid-height
(Z=L ¼ 0:5) of the FC models. The length of the vortex formation region increases
as the FC aspect ratio decreases. For the small aspect ratios of L=D ¼ 6 and 10, the
vortex formation length is about twice the length of the 2-D cylinder. This is a
peculiar flow characteristic of the FC wake, caused mainly by the descending shear
flow separated from the FC free end. The L=D ¼ 13 finite cylinder has nearly the
same peak value as the 2-D cylinder (L=D ¼ 17:3), but its peak location is shifted
about one cylinder diameter downstream. This indicates that the FC free end
increases the vortex formation region further downstream. With decreasing aspect
ratios (L=D), the magnitude of the turbulence intensity is decreased and the clear
peak becomes blunt. For the L=D ¼ 6 finite cylinder, the turbulence intensity
distribution around the peak point is nearly flat.
Fig. 8 shows the turbulence intensity distributions of the streamwise velocity
component measured along the central section (Y ¼ 0) of the wake for the L=D ¼ 10
cylinder. The turbulence intensities are small and the vortex formation region is
relatively large excluding the free end region, compared to the 2-D circular cylinder.
As the FC free end is approached, the turbulence intensity and the vortex
formation region are decreased. This may be attributed to the downwash flow along
the central region of the wake, induced by the twin counter-rotating vortices
240 C.-W. Park, S.-J. Lee / J. Wind Eng. Ind. Aerodyn. 88 (2000) 231–246

Fig. 7. Comparison of the vortex formation region measured at Z=L ¼ 0:5.

Fig. 8. Streamwise turbulence intensity distributions of wake behind the finite cylinder of L=D ¼ 10.
C.-W. Park, S.-J. Lee / J. Wind Eng. Ind. Aerodyn. 88 (2000) 231–246 241

separated from the FC free end, as shown in Fig. 4. In addition, the pressure behind
the FC free end is decreased, due to active interaction between the separated
downwash flow from the free end and the vortices shed from the two sides of the
cylinder. This interaction causes a pressure gradient in the longitudinal direction. As
can be seen in Fig. 8, the regular vortex shedding disappears, and the vortex
formation region is difficult to identify near the FC free end.

3.4. Spectral analysis

Fig. 9 shows the PSD distributions of the wake behind the L=D ¼ 10 finite
cylinder at downstream locations of X=D ¼ 3; 4; 5; 6 and Y=D ¼ ÿ 2. The PSD
distributions measured at mid-height (Z=L ¼ 0:5) show a clear peak at the vortex
shedding frequency, which explains the coherent vortex structure. As the FC free end
is approached, the vortex shedding frequency and the PSD value at the peak
frequency are decreased slightly. Near the free end, and only in this region, a peculiar
spectral peak occurs at 24 Hz. In the downstream direction, the PSD value at 24 Hz is
increased, as indicated in the measurements at Z=L ¼ 0:95. Fig. 9(d) shows the
intermittent frequency characteristics at a height of Z=L ¼ 0:875. This indicates that
the regular vortex shedding frequency and 24 Hz frequency components are
combined at this location. The 24 Hz frequency component is well matched with
the descending counter-rotating swirling vortices, as shown in Figs. 4 and 5. Ayoub
and Karamcheti [14] also found the 24 Hz frequency characteristics at the free end of
a finite cylinder at Re ¼ 8:5  104 .
Fig. 10 shows the PSD distributions of the streamwise velocity component
measured at Z=L ¼ 0:95 along the wake centerline (Y=D ¼ 0). The 24 Hz frequency
component is still clear at X=D ¼ 2:4 for the L=D ¼ 13 finite cylinder. Although the
24 Hz component is not as clear downstream of the L=D ¼ 6 FC model, it still exists
just behind the FC free end in the PSD distribution measured at X=D ¼ 0:6. These
results indicate that the 24 Hz frequency component is not affected by the FC aspect
ratio at all. This means that the 24 Hz component is different from the periodic
vortex shedding frequency that occurs in a 2-D circular cylinder wake. In order to
verify this, the cross-correlation function Ru1 u2 of the velocity signals of the vortices
shed from the two sides of the finite cylinder was derived. Two single hot-wire probes
were installed in parallel at Y=D ¼ 2, and velocity signals from both probes were
measured simultaneously. Fig. 11 shows the cross-correlation function Ru1 u2 for the
L=D ¼ 13 and L=D ¼ 17:3 (2-D cylinder) FC models measured at the mid-height of
the cylinder (Z=L ¼ 0:5). Since the cross-correlation function has a local minimum
negative value at t ¼ 0 (no time delay), velocity signals measured at the two sides of
the cylinder are 1808 out-of-phase. This indicates that vortices are shed alternately
from the two sides of the cylinder in turn. The correlation function for the 2-D
cylinder is slightly larger than that for the finite cylinder (L=D ¼ 13), but the time-
delay of the finite cylinder is a little longer than that of the 2-D circular cylinder. This
agrees well with the smaller vortex shedding frequency shown in Fig. 6.
Fig. 12 shows the cross-correlation function between the velocity signals measured
from the two sides (Y=D ¼ 2) of the L=D ¼ 13 FC at X=D ¼ 3. From spectral
242 C.-W. Park, S.-J. Lee / J. Wind Eng. Ind. Aerodyn. 88 (2000) 231–246

Fig. 9. Power spectral density distributions of wake behind the finite cylinder of L=D ¼ 10 measured at
Y=D ¼ ÿ2.

analysis of the velocity signals measured near the FC free end, the 24 Hz frequency
component appears dominant compared with other frequency components in this
region. Therefore, the velocity signals at two locations, Y=D ¼ 2 and Z=L ¼ 0:95,
were measured simultaneously, very close to the free end. The velocity signals were
band-pass-filtered at 24 Hz to observe the characteristics of the 24 Hz component in
more detail. The cross-correlation function measured in the region near the FC free
end is quite different from those of Fig. 11. The cross-correlation has a maximum
C.-W. Park, S.-J. Lee / J. Wind Eng. Ind. Aerodyn. 88 (2000) 231–246 243

Fig. 10. Power spectral density distribution along the wake centerline ðY=D ¼ 0Þ at Z=L ¼ 0:95.
244 C.-W. Park, S.-J. Lee / J. Wind Eng. Ind. Aerodyn. 88 (2000) 231–246

Fig. 11. Cross-correlation function of velocity signals measured from both sides ðY=D ¼ 2Þ of the finite
cylinder at X=D ¼ 3; Z=L ¼ 0:5.

positive value at no time delay (t ¼ 0). This means that the vortices on the two sides
of the cylinder with the 24 Hz frequency component have no phase difference (in-
phase). The downwashing counter-rotating vortices shown in Fig. 4 are in-phase in
the near wake. From these results, we can conjecture that the 24 Hz frequency
component is attributed to the downwash swirling vortices separated from the top of
the finite cylinder. This peculiar frequency is dominant near the FC free end and
independent of the periodic vortex shedding frequency.

3.5. Surface pressure

The mean surface pressure distributions near the free end of the L=D ¼ 10 finite
cylinder are shown in Fig. 13. Moving downward along the cylinder axis, the surface
pressure gradually approaches that of the two-dimensional cylinder. At
Z=L ¼ 0:883, the pressure distribution has nearly recovered to that of a 2-D
cylinder.
The mean pressures on the leeward surface of the finite cylinder decreases as the
free end is approached. The shear flow separated from the FC free end declines
toward the downward direction due to this low pressure. Moving downward along
the finite cylinder, however, the mean pressures increase, because the vortices shed
from the two sides of the cylinder become dominant.
C.-W. Park, S.-J. Lee / J. Wind Eng. Ind. Aerodyn. 88 (2000) 231–246 245

Fig. 12. Cross-correlation function of velocity signals having the 24 Hz frequency component at
Y=D ¼  2 for the finite cylinder of L=D ¼ 13 ðX=D ¼ 3; Z=L ¼ 0:95Þ.

Fig. 13. Comparison of mean pressure coefficient distributions around the free end of L=D ¼ 10 cylinder.
246 C.-W. Park, S.-J. Lee / J. Wind Eng. Ind. Aerodyn. 88 (2000) 231–246

4. Conclusion

The effect of a finite cylinder (FC) free end in a cross flow on near-wake flow
characteristics was investigated experimentally. The three dimensionality of the FC is
attributed to the counter-rotating twin vortices separated from the FC free end. The
downwash flow is concentrated on the central region of the wake. From spectral
analysis and cross-correlation of the velocity signals, a peculiar flow structure having
a 24 Hz frequency component was observed near the free end. This 24 Hz frequency
component seems to be closely related to the counter-rotating twin vortices formed
near the FC free end.
The vortex shedding frequency decreases and the vortex formation region
increases as the FC aspect ratio decreases. The vortex formation region and periodic
vortex shedding disappear very close to the free end. This is caused by the twin
vortices descending from the free end, which interact with the regular vortices shed
from the two sides of the cylinder.

Acknowledgements

The authors thank the BK21 project for its contribution.

References

[1] A. Slaouti, J.H. Gerrard, An experimental investigation of the end effects on the wake of a circular
cylinder towed through water at low Reynolds numbers, J. Fluid Mech. 112 (1981) 297–314.
[2] C.H.K. Williamson, Oblique and parallel modes of vortex shedding in the wake of a circular cylinder
at low Reynolds numbers, J. Fluid Mech. 206 (1989) 579–627.
[3] S. Szepessy, P.W. Bearman, Aspect ratio and end plate effects on vortex shedding from a circular
cylinder, J. Fluid Mech. 234 (1992) 191–217.
[4] F. Baban, R.M.C. So, M.V. Otugen, Unsteady forces on circular cylinders in a cross-flow, Exp.
Fluids. 7 (1989) 293–302.
[5] H. Eisenlohr, H. Eckelmann, Vortex splitter and its consequences in the vortex street wake of
cylinders at low Reynolds number, Phys. Fluids: Part A 1 (2) (1989) 189–192.
[6] C. Wieselsberger, Versuche uber den luftwiderstand gerundeter und kantiger korper, in: L. Prandtl
(Ed.), Ergebnisse Aerodyn. Versuchsanstalt Gottingen, Vol. II. Lifeerung, 1923, p. 23.
[7] F. Baban, R.M.C. So, Aspect ratio effect on flow-induced forces on circular cylinders in a cross-flow,
Exp. Fluids. 10 (1991) 313–321.
[8] M. Budair, A. Ayoub, K. Karamcheti, Frequency measurements in a finite cylinder wake at a
subcritical Reynolds number, J. AIAA 29 (1991) 2163–2168.
[9] H. Sakamoto, M. Arie, Vortex shedding from a rectangular prism and a circular cylinder placed
vertically in a turbulent boundary layer, J. Fluid Mech. 126 (1983) 147–165.
[10] S. Okamoto, Y. Sunabashiri, Vortex shedding from a circular cylinder of finite length placed on a
ground plane, J. Fluids Eng. 114 (1992) 512–521.
[11] D. Farivar, Turbulent uniform flow around cylinders of finite length, J. AIAA 19 (3) (1981) 275–281.
[12] M.M. Zdravkovich, V.P. Brand, G. Mathew, A. Weston, Flow past short circular cylinders with two
free ends, J. Fluid Mech. 203 (1989) 557–575.
[13] G.S. West, C.J. Apelt, The effects of tunnel blockage and aspect ratio on the mean flow past a circular
cylinder with Reynolds numbers between 104 and 105 , J. Fluid Mech. 114 (1982) 361–377.
[14] A. Ayoub, K. Karamcheti, An experiment on the flow past a finite circular cylinder at high subcritical
and supercritical Reynolds numbers, J. Fluid Mech. 118 (1982) 1–26.

You might also like