You are on page 1of 15

Journal of Fluids and Structures 27 (2011) 903–917

Contents lists available at ScienceDirect

Journal of Fluids and Structures


journal homepage: www.elsevier.com/locate/jfs

On the study of vortex-induced vibration of a cylinder


with helical strakes
T. Zhou a,n, S.F. Mohd. Razali b, Z. Hao c, L. Cheng a
a
School of Civil and Resource Engineering, The University of Western Australia, 35 Stirling Highway, WA 6009, Australia
b
Faculty of Engineering & Built Environment, Universiti Kebangsaan Malaysia, 43600 Bangi, Selangor, Malaysia
c
College of Logistics Engineering, Shanghai Maritime University, Shanghai 200135, China

a r t i c l e i n f o abstract

Article history: While the effect of helical strakes on suppression of Vortex-Induced Vibrations (VIV) has
Received 17 November 2009 been studied extensively, the mechanism of VIV mitigation using helical strakes is much
Accepted 21 April 2011 less well documented in the literature. In the present study, a rigid circular cylinder of
diameter d¼80 mm attached with three-strand helical strakes of dimensions of 10d in
Keywords: pitch and 0.12d in height was tested in a wind tunnel. It was found that the helical strakes
Vortex shedding can reduce VIV by about 98%. Unlike the bare cylinder, which experiences lock-in over the
Vortex-induced vibration reduced velocity in the range of 5–8.5, the straked cylinder does not show any lock-in
Helical strakes region. In exploring the mechanism of VIV reduction by helical strakes, measurements in
Correlation length
stationary bare and straked cylinder wakes using both a single X-probe at four different
Reynolds numbers, i.e. Re¼ 10 240, 20 430, 30 610 and 40 800, and two X-probes with
variable separations in the spanwise direction at Re¼ 20 430 were conducted. It was found
that vortices shed from the straked cylinder are weakened significantly. The dominate
frequency varies by about 30% over the range of x/d¼ 10–40 in the streamwise direction
while that differs by about 37.2% of the averaged peak frequency over a length of 3.125d
in the spanwise direction. The latter is supported by the phase difference between the
velocity signals measured at two locations separated in the spanwise direction. The
correlation length of the vortex structures in the bare cylinder wake is much larger than
that obtained in the straked cylinder wake. As a result, the straked cylinder wake agrees
more closely with isotropy than the bare cylinder wake. Flow visualization on the plane
perpendicular to the cylinder axis at Reynolds number of about 300 reveals small-scale
vortices in the shear layers of the straked cylinder wake. However, these vortices do not roll
up and interact with each other to form the well-organized Karman-type vortices. Flow
visualization on the plane parallel to the cylinder axis shows vortex dislocation and
swirling flow, which should be responsible for the variations of the peak frequency in the
streamwise as well as spanwise directions.
& 2011 Elsevier Ltd. All rights reserved.

1. Introduction

When vortices are shed from a bluff body, the latter is subjected to time-dependent drag and lift forces. The lift
force oscillates at the vortex shedding frequency while the drag force oscillates at twice the vortex shedding frequency

n
Corresponding author. Fax: þ61864881018.
E-mail address: tzhou@civil.uwa.edu.au (T. Zhou).

0889-9746/$ - see front matter & 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jfluidstructs.2011.04.014
904 T. Zhou et al. / Journal of Fluids and Structures 27 (2011) 903–917

(Sumer and Fredsøe, 1997). If the cylinder is flexibly-mounted, these forces may induce vibration of the cylinder. The lift
force may induce cross-flow vibrations and the drag force may induce in-line vibrations. This phenomenon is called
vortex-induced vibration (VIV) (Blevins, 2001). Vortex-induced vibration of bluff structures is one of the key issues in riser
and pipeline designs. This is because VIV will increase not only the dynamic load to the structures but will also influence
the structural stability. The vibrations may cause structural failure or accelerate the fatigue failure. The above factors may
result in an increase in capital investment of the structures and the expenses for maintenance and replacement.
The VIV of a cylindrical structure depends on various parameters, such as Reynolds number, Re, Strouhal number,
St (  f0d/UN, where f0 in the vortex shedding frequency) and the reduced velocity, Vr (  UN/fnd, where fn is the natural
frequency of the structure), etc. The mechanism of vortex shedding of a circular cylinder and the oscillation related to it
have been studied for more than a century and yet our understanding is still far from complete even though much progress
has been made during the past several decades, both numerically and experimentally due to the advances of the
technologies in experiments and computing powers (Bearman, 1969; Carberry et al., 2005; Elston et al., 2006; Roshko,
1954; Sarpkaya, 1979; Williamson and Govardhan, 2004; Zdravkovich, 1981). The ultimate objective of studies on VIV is
the understanding, prediction and prevention of VIV preferably without drag penalty. Since VIV phenomenon is very
complex to model, its prediction is challenging because commercially available predictive tools have not been extensively
calibrated with valid experimental data. In many situations, these tools provide different answers for the same problem. As
a result, the level of confidence in VIV analyses is low and conservative approaches using high factors of safety are used in
design of the pipelines and risers. With offshore facilities being installed in increasingly deeper waters, such conservative
designs have also become increasingly economically untenable. VIV prevention or mitigation is therefore critically
important, as it will yield solutions that enable safe, cost-effective and reliable deepwater project developments.
Suppression of vortex shedding and VIV of a bluff body has been one of the most active topics of research and patenting
in fluid dynamics for many decades due to its significance in engineering applications. Previous investigations on passive
control devices to suppress VIV have contributed significantly to the fields of buildings, bridges and marine engineering
and yet there are difficulties in achieving an optimal balance between performance, cost and simplicity. To prevent VIV
lock-in, one can either change the natural frequency of the structure by structural modification or to inhibit the formation
of vortices or to disrupt their structural formation, through the application of the suppression devices. Suppression of
vortex shedding and hence VIV can be achieved by active methods (where external energy is supplied), passive methods
(where no external energy is supplied to control the flow), or a combination of the above two methods. Helical strake is
one of the most proven and widely used control measures in various industrial and offshore applications to suppress VIV.
It normally contains three-start strakes helically wound on the surface of the structural member in a definite fashion. The
efficiency of VIV suppression using helical strakes depends on the dimensions of the strakes, i.e. the height of the
protrusion and the pitch of the helical. Previous studies have shown that the optimal height of strakes for VIV suppression
is about 0.05 0.2d and a pitch of about 3.6  5d has generally been accepted as optimal, though recent tests have shown
that a pitch of 15d may be just as effective (Jones and Lamb, 1992). Basically, it is believed that they adversely affect the
shear layer to roll up and to disrupt the spanwise vortex formation and shedding process (Scruton and Walshe, 1963). As
fluid flows past a cylinder with helical strakes, the strake chops up the flow and creates vortices at various places along the
cylinder. These vortices are out of phase with one another and produce destructive interference to the dominant vortex
shedding. Due to partial cancellation of the out-of-phase lift forces at different spanwise positions, the lift coefficient for
the straked cylinder is much smaller than that of a bare cylinder, resulting in a significant reduction in the amplitude of
VIV. Even though helical strakes are found effective in suppressing VIV for high damping values, the effectiveness of VIV
suppression using helical strakes for low mass-damping values is reduced if the strake height is not big enough. Although
Bearman and Branković (2004) did not find evidence of regular Karman-type vortex shedding close to the stationary
cylinder attached with helical strakes of 0.12d in height and 5d in pitch, an undulation of the wake was observed at about 2
cylinder diameters downstream. However, when the straked cylinder is free to respond for Vr larger than about 5 and the
mass and damping are low, the flow pattern changes significantly. The straked cylinder not only experienced the lock-in
but also exhibited 2S and 2P vortex shedding modes, which is similar to that for a bare cylinder. These results are different
from those reported by Constantinides and Oakley (2006) for strakes of 0.25d in height and 15d in pitch, due mainly to the
difference in strake height. The latter authors showed that the helical strakes can suppress VIV nearly completely over the
lock-in range of the bare cylinder. They found that the vibration amplitude of the straked cylinder increased monotonically
with the increase in reduced velocity, which was attributed to the effect of wake galloping. There were no organised
vortices observed in the straked cylinder wake. Their flow visualization suggested that the separation point was at the tip
of the strake for most of the coverage. The helical separation point induces a three-dimensional flow behind the cylinder
breaking the vortex coherence. In certain locations where the strake tip is aligned with the flow, either upstream or
downstream, separation is partially controlled by the cylinder surface. The other mechanism that helical strakes use to
mitigate VIV is by limiting the interaction between the two shear layers that are formed due to separation (Constantinides
and Oakley, 2006). The helical strakes act as obstacles preventing the shear layers to communicate and form the typical
vortices we find in bare cylinder wakes.
Even though helical strakes are proven effective for suppressing VIV especially for low mass-damping values, the
mechanism of VIV suppression has not yet been fully understood (Constantinides and Oakley, 2006). Therefore, the present
study aims to shed some light on the vortex characteristics and to enhance our understanding on the mechanisms of
helical strakes have on VIV suppression using hot wire measurements and smoke wire flow visualization in a wind tunnel.
T. Zhou et al. / Journal of Fluids and Structures 27 (2011) 903–917 905

2. Experimental details

The experiments were conducted in a low speed section of the boundary layer wind tunnel of the University of Western
Australia. The dimensions of the test section used are 2230 mm high by 2870 mm wide. An aluminum cylinder with a diameter
d of 80 mm and a length L of 1600 mm was used either as a bare cylinder or after the installation of the helical strakes, as a
straked cylinder. The three-start strakes have a pitch P of 10d and height h of 0.12d (Fig. 1), which is in the range of the most
effective dimension in VIV suppression (e.g. Kumar et al., 2008). The aspect ratio of the cylinder L/d is 20. The mass of the bare
cylinder is 3.035 kg and the mass of the cylinder with strakes is 3.496 kg. At each end of the cylinder a steel ‘O’ ring fits snug
around the outer diameter and provides an attachment for two steel eye-loops from which the springs are attached. Four steel
springs, two at either end of the cylinder, suspend the cylinder approximately 1000 mm off the wind tunnel floor. The springs
were selected based on their stiffness and length. They must be stiff enough and long enough to remain in tension when
measuring the amplitude of vibrations. All four springs were of an equal stiffness k¼1680 N/m and equal original length of
250 mm. The stiffness was measured by applying a known load to the springs and measuring the subsequent deflection. Once
the springs were attached to the cylinder, the top and bottom springs stretched to a length of 340 and 330 mm, respectively.
The amplitude of the vibrations was measured using a Linear Variable Differential Transformer (LVDT) laser. The laser
was cantilevered off the frame’s vertical member and is positioned 130 mm above the cylinder. The vibration signals from
the laser were sampled into a computer in the form of voltages, which have been calibrated that 1 V¼10 mm, at a
frequency of 50 Hz using Labview 7.1. The natural frequencies of the bare cylinder and the straked cylinder are obtained
from the free-decay tests. Displacements from the cylinders after an initial displacement were recorded using the LVDT at
a frequency of 50 Hz. Spectrum was then obtained by applying fast Fourier transform (FFT) to the time series of the
measured signals. A strong peak on the distribution of the spectrum corresponds to the natural frequency, which is
identified as 6.6 and 6.37 Hz, for the bare and straked cylinders, respectively.
Velocity fluctuations u in the streamwise (x-) and n in the transverse (y-) directions are measured using X-wire probes.
For the purpose to examine the spanwise cross-correlations, two X-wire probes located at x/d¼ 5 and 10 and y/d¼0.5 were
used with one probe fixed and the other moving along the cylinder length direction. The separation between the two probes
was in the range of 10–250 mm. For the purpose to examine the evolution of the vortices in the streamwise direction,
measurements using a single X-wire probe were conducted in the range of x/d¼5 40 at four different free stream velocities
UN, i.e. 1.92, 3.83, 5.74 and 7.65 m/s, corresponding to Re¼10 240, 20 430, 30 610 and 40 800, respectively. The hot wires were
operated with in-house constant temperature circuits at an over heat ratio of 1.5. Each of the two wires in the X-wire probe had
a diameter of 5 mm. The wire separation was about 1 mm. The output signals from the anemometers were low-pass filtered
through the buck and gain circuits at a cut-off frequency fc ¼0.5–9.2 kHz, depending on the measurement location and free
stream velocity. The values of fc were determined by examining the spectrum of @u/@t and identifying the onset of electronic
noise (Antonia et al., 2002), which were normally close to the Kolmogorov frequency fK, ( U/2pZ), where Z is the Kolmogorov
length scale and can be calculated using Z  ðn3 =/eSÞ1=4 . Hereafter, the angular brackets denote time-averaging. The mean
energy dissipation rate /eS is defined as /eS¼2n/sijsijS, where sij ¼(ui,j þuj,i)/2 is the rate of strain, n is the kinematic viscosity
and ui,j ¼@ui/@xj. Strictly, the mean energy dissipation rate /eS should be obtained by measuring all the 12 terms in its
expression. However, this is not normally feasible unless using a multi-hot wire probe (e.g. Vukoslavčevié et al., 2009). On the
other hand, even though we can measure the 12 terms using the multiple hot wire probes, there also exist limitations on
spatial resolution of the probe on the measured velocity gradients, which result in an underestimation of /eS (e.g. Antonia
et al., 1998). Therefore, a very common method for experimentalists is to use /eS¼15n/(@u/@x)2S by assuming isotropy and
using Taylor’s hypothesis by replacing Dx with UDt, where Dt ( 1/fs) is the sampling time interval between two consecutive
points and fs is the sampling frequency.

y
Height
d
z

Pitch length

Fig. 1. Geometry of the cylinder with three-strand helical strakes and definition of the coordinate system.
906 T. Zhou et al. / Journal of Fluids and Structures 27 (2011) 903–917

The filtered signals were sampled at a frequency fs ¼2fc into a PC using a 16-bit A/D converter (National Instrument).
The sampling period was 60–120 s, depending on the Reynolds numbers and measurement locations. Experimental
uncertainties in U and u0 (or v0 ) were inferred from the errors in the hot-wire calibration data as well as the scatter (20 to
1 odds) observed in repeating the experiment for a number of times. The uncertainty for U was about 73%, while the
uncertainties for u0 and v0 were about 7 5% and 76%, respectively. The experimental uncertainty for St ( Uf0/d) depends
on that of the velocity and the vortex shedding frequency measurements. The vortex shedding frequency is identified
using the fast Fourier transform (FFT) algorithm with a window size of 211. Its uncertainty depends on the sampling
frequency and the window size when doing FFT (Xu and Zhou, 2004). In the present study, we estimated that the
uncertainty for St was around 3.6%.
To further examine the differences of the vortex structures in the wakes of the bare cylinder and the straked cylinder,
flow visualization in a wind tunnel of 25 cm (height)  40 cm (width) and 2 m (length) was also conducted using a smoke
wire at a Reynolds number of about 300. The diameter of the cylinder for this experiment was 12.7 mm. The dimensions of
the strakes are comparable with that used in the first experiment, i.e. 0.12d in height and 10d in pitch.

3. Results and discussion

3.1. Vortex shedding frequency for the bare and straked cylinders

The vortex shedding frequency at various flow velocities were examined by performing FFT to the velocity signals
measured in the wake using the hot wire anemometers. Before the cylinder starts to vibrate, vortex shedding from the
cylinder is apparent with a peak frequency corresponding to a Strouhal number of 0.21, which agrees well with the
consensus values published in the literature. Fig. 2 gives an example of the spectrum obtained on the centreline of the bare
cylinder wake at x/d ¼5 for Re ¼20 430. The peak height relative to the height of the plateau at low frequency is about
22 dB. After the cylinder starts to vibrate, the energy spectrum peaks at the natural frequency of the cylinder over the
whole lock-in range. The peak frequencies on the energy spectra at different reduced velocities are replotted in the form of
the ratio f0/fn versus reduced velocities (Fig. 3). The ratio increases linearly with Vr until it reaches the lock-in regime at a
reduced velocity of about 5. Over the lock-in regime, the ratio f0/fn keeps a constant value of 1. The straight line in the
figure has a slope of 0.21, which corresponds to the Strouhal number. In contrast to the bare cylinder, there is only a minor
peak on the spectrum for the straked cylinder (Fig. 2). Various locations in the streamwise (x) direction as well as in the
transverse (y) direction were tested and yet, no apparent peak on the energy spectra, which is as apparent as that for the
bare cylinder was identified. The peak height of the straked cylinder relative to the height of the plateau at low frequency
is 4 dB, which is only about 1/6 of that in the bare cylinder wake. It can also be seen that the peak region on the spectrum
of the straked cylinder is broadened, suggesting that the intensity of the vortices shed from the straked cylinder is
significantly reduced compared with that from the bare cylinder.

3.2. Vortex-induced vibration for the bare and straked cylinders

In order to validate the current experimental setup, the dynamic response of a single bare cylinder is tested first. The
response of the cylinder can be described in terms of the vibration amplitude A/d versus the reduced velocity Vr. The
cylinder was free to oscillate in y-direction. It was found that for each reduced velocity, the vibrations take about 50 s to

40

20
bare cylinder

0
φv(f) (dB)

straked cylinder

-20

-40

-60
0.001 0.01 0.1 1 10 100
fd/U∞

Fig. 2. Velocity spectra in the wake of the stationary bare and straked cylinders on the centreline at x/d ¼5 for Re ¼ 20 430.
T. Zhou et al. / Journal of Fluids and Structures 27 (2011) 903–917 907

f0/fn
2

Slope = 0.21

0
0 5 10 15
Vr

Fig. 3. Vortex shedding frequency from the bare cylinder normalized by the natural frequency. The straight line has a slope of 0.21, which corresponds to
the Strouhal number, St.

1.0

Upper branch
0.8

Lower branch
0.6
A /d

0.4

0.2

0
0 5 10 15
Vr

Fig. 4. Comparison of vibration response of a bare cylinder. .: Present; &: Khalak and Williamson (1996); J: Franzini et al. (2009); n: Feng (1968).

stabilize, after which the amplitude of vibration does not change. A test of about 60 s was recorded after stable patterns of
vibrations were observed. The amplitudes of vibration were determined by averaging the 10% highest peaks recorded on
the time history of the displacement after the vibration becomes stable (Franzini et al., 2009; Hover and Triantafyllou,
2001). The vibration response at different reduced velocities were compiled and displayed in Fig. 4 for the bare cylinder. It
can be seen that the maximum peak amplitude is about 0.51, occuring at a reduce velocity of about 6.5, which is within the
consensus range found in the literature (Sumer and Fredsøe, 1997). Experimental results by Feng (1968) in air and Khalak
and Williamson (1996) and Franzini et al. (2009) (L/d¼ 24) in water are also included for comparison. It can be seen that
the response amplitudes in the present study follow that of Feng’s results closely, especially in the upper branch. The lock-
in region in the two studies is also very similar. This is reasonable as the mass damping parameter mnzs in the two studies
are comparable, with mnzs being about 0.26 in the present study and 0.255 in Feng (1968), where mn (  m/mf, with m being
the structural mass including the enclosed fluid mass and mf being the displaced fluid mass) is the mass ratio and zs is the
structural damping factor, which can be obtained using zs ¼ ð1=2pÞlnðyn þ 1 =yn Þ from a free decay test. The mass damping
parameters mnzs in Khalak and Williamson (1996) and Franzini et al. (2009) are 0.013 and 0.0092, respectively, which are
much smaller than that in the present study and in Feng (1968). Correspondingly, the maximum vibration amplitudes in
Khalak and Williamson (1996) and Franzini et al. (2009) are much larger than that in the latter two studies. There exist
two apparent branches, i.e. the upper branch and the lower branch. Also in the former studies with low mass damping
parameters, the onset of cylinder vibration occurs earlier and diminishes later in amplitude than the latter two studies.
908 T. Zhou et al. / Journal of Fluids and Structures 27 (2011) 903–917

The maximum vibration amplitude Amax/d obtained in the present study can be compared with that obtained using the
empirical relations (e.g. Sumer and Fredsøe, 1997; Sarpkaya, 2004). Sumer and Fredsøe (1997) proposed that for stability
parameter Ks larger than 2, where Ks can be quantified using Ks ¼4pm*zs, the maximum vibration amplitude can be
estimated using

Amax 1:7
¼ for Ks Z 2: ð1Þ
d Ks

Based on the mass ratio mn and the structural damping factor zs, the stability parameter Ks in the present study
is about 3.27. This value results in a maximum amplitude Amax/d of 0.52 for the bare cylinder, which agrees very well with
the value obtained in the experiment. By introducing another parameter, the so-called response parameter SG
(  2p3 St 2 m zs ), Sarpkaya (2004) proposed another relationship for quantifying the maximum vibration amplitude
Amax/d, viz.

Amax =d ¼ 1:12g  0:35SG : ð2Þ

For a rigid circular cylinder, the dimensionless mode factor g ¼1.0. With SG ¼0.712 in the present study, the maximum
amplitude Amax/d using Eq. (2) is estimated to be about 0.53, which also agrees very well with the present experimental
value (0.51). The above results for the single cylinder have justified the present experimental arrangement for the single
cylinder-spring system.
The vibration response of the cylinder attached with helical strakes is then tested. It is found that the strakes can
suppress the vibration significantly. With the strakes attached, the cylinder does not vibrate apparently over the same
reduced velocity range of the bare cylinder. In contrary to the response of the bare cylinder, which vibrates severely over
the lock-in range, it seems that the cylinder attached with strakes does not experience a lock-in range. Actually, the peak
on the energy spectra measured using hot wire probes downstream of the straked cylinder does not coincide with the
natural frequency of the straked cylinder ( ¼6.4 Hz), indicating that lock-in phenomena does not occur. The vibration
amplitudes increase with Vr, even though with very small magnitude compared with that for the bare cylinder. The
vibration amplitudes of the cylinder with strakes are shown in Fig. 5. Even at a reduced velocity of 15, the cylinder does not
show a reduction in vibration amplitude, as that observed for a bare cylinder. This trend is consistent with that reported by
Constantinides and Oakley (2006) and Ding et al. (2004). They suggested that the increased vibration amplitude with the
increase in reduced velocity might be attributed to wake galloping rather than the classical VIV. Also shown in the figure is
the vibration amplitude of the bare cylinder. Apparently, the strakes can reduce the vibration amplitude by 98%. The
present result on the vibration amplitude for the straked cylinder is in contrast with that reported by Bearman and
Branković (2004) for a cylinder attached with helical strakes, probably due to the low damping ratios used in the latter
study. These authors found that VIV occurs over a narrower range of reduced velocity. The response characteristics,
especially for the lowest mass ratio, are surprisingly similar to those for a plain cylinder with what may be a lower branch
appearing. They also found that for the straked cylinder, lock-in phenomena occurs and the vibration exhibited 2S and 2P
modes with a behavior similar to that for the plain cylinder, although the amplitudes were smaller.

0.6

0.5

0.4
A/d

0.3

0.2

0.1

0
0 5 10 15 20 25
Vr

Fig. 5. Comparison of vibration response between bare cylinder and straked cylinder. ., present bare cylinder; þ , present straked cylinder; &, straked
cylinder by Ding et al. (2004); J, straked cylinder by Constantinides and Oakley (2006).
T. Zhou et al. / Journal of Fluids and Structures 27 (2011) 903–917 909

3.3. Energy spectra in the streamwise and spanwise directions of the stationary cylinders

In order to understand the physical mechanism of how the strakes mitigate VIV on a bare cylinder, energy spectra
measured using hot wire probes are analyzed. In Fig. 6, the energy spectra obtained in the stationary bare and straked
cylinder wakes at x/d¼5, 10, 20 and 30 on the centreline for Re ¼20 430 are compared in order to examine the evolution of
the vortex structures in the streamwise direction. It can be seen that for the bare cylinder (Fig. 6(a)), the peaks of the
spectra are sharp and occur at approximately a fixed frequency in the streamwise direction, which, after normalization by
UN and d, gives a Strouhal number St of 0.21, consistent with the consensus values published in the literature. This result
indicates that the vortex shed from the bare cylinder is regular. Zhou et al. (2003) showed that the vortex structures
generated from the bare cylinder broke down and the intensity of the vortices reduced significantly in the streamwise
direction. At about x/d¼40, the vortex structures nearly disappear. The present result agrees with this. At x/d¼ 40, there is
no peak discernable (spectrum for this location is not shown to avoid crowding the figure) and at x/d ¼30, only a very
minor peak exists. The peak frequency variation at different locations, as indicated by Df in the figure, is about 0.2 Hz,
which is only about 2% of the averaged vortex shedding frequency f0. This value should be within the experimental
uncertainty for the detection of f0 and therefore f0 can be regarded as a constant value. However, on the spectra of the
straked cylinder wake (Fig. 6(b)), the peak location is not a constant value for different downstream locations. The Strouhal
number, St, based on the averaged value of f0 at different locations, is about 0.14. This value is much smaller than that for a
circular cylinder (E0.21) at the tested Reynolds number range but closer to the value of St found in a flat plate wake
(Castro and Rogers, 2002; Roshko, 1954). Since the strakes resemble flat plates perpendicular to the flow, they increase the
effective diameter of the cylinder. The peak varies at different downstream locations. The maximum difference of the peak

0.7

(a)
0.6
Δf=0.2Hz

0.5

0.4
φv(f)

0.3 x/d = 5
10
20
0.2 30

0.1

0
0.01 0.1 1 10
fd/U∞

0.15

(b)

Δf=2.2Hz

0.10
φv(f)

x/d = 5
10
0.05 20
30

0
0.01 0.1 1 10
fd/U∞

Fig. 6. Energy spectra obtained on the centreline at different streamwise locations for Re ¼ 20 430 in the stationary bare and straked cylinder wakes.
(a) Bare cylinder; (b) straked cylinder.
910 T. Zhou et al. / Journal of Fluids and Structures 27 (2011) 903–917

frequency Df between x/d ¼5 and 30 is about 2.2 Hz. This value corresponds to a variation of 30% of the averaged peak
frequency, which is too big to be attributed to the experimental uncertainty in measuring fs. Rather, it may indicate a
dislocation, which may be caused by the spanwise motion of the fluid downstream of the straked cylinder since part of the
wake will be transported laterally and arrive in the neighboring region. This effect will also enhance the three-
dimensionality of the flow. Flow visualization using computational fluid dynamics (CFD) (Constantinides and Oakley,
2006) showed significant spanwise motion and swirling in the near wake of the strake. These authors found that the
spanwise motion in the wake is caused by the fluid being channeled by the strakes and also by the separated fluid that
induces a spanwise flow.
The variation of vortex shedding frequency along the spanwise direction can also be examined. Fig. 7 shows the spectra
measured in the stationary bare and straked cylinder wakes at different locations along the cylinder length. For the bare
cylinder, the peaks are sharp and occur at a fixed frequency along the cylinder length, i.e. f0 E10 Hz, or with a wavelength
l/d ¼4.8. This result indicates that vortices are shed from the bare cylinder at a fixed frequency even though they may be
shed in cells along the cylinder length at slightly shifted phase (see discussion for Fig. 8). In contrast, the peaks on the
energy spectra of the straked cylinder wake do not occur at a fixed frequency (Fig. 7(b)). The averaged frequency f0 is about
8.6 Hz, which corresponds to a wavelength l/d of 5.55. The maximum difference among the peaks along the cylinder
length direction is about 3.2 Hz, which corresponds to a variation of 37.2% of the averaged vortex shedding frequency. The
variation of the vortex shedding frequency in the energy spectra was also shown in the wake of wavy plate model by
Bearman and Owen (1998). The variation of near-wake width (Owen and Bearman, 2001) and the separation points
(Constantinides and Oakley, 2006) along the cylinder axis may contribute to the inconsistency of the peaks in the energy
spectra along the straked cylinder axis. For the straked cylinder, the near-wake width and the separation points depend on

0.7
(a)
0.6

0.5

0.4
Δz/d = 0.125
1.25
0.3
φv(f)

2.5
3.125
0.2

0.1

0
0.01 0.1 1 10
fd/U∞

0.15
(b)
Δf=3.20Hz

0.10
φv(f)

Δz/d = 0.125
1.25
2.5
3.125
0.05

0
0.01 0.1 1 10
fd/U∞

Fig. 7. Energy spectra obtained at different spanwise locations in the stationary bare and straked cylinder wakes at x/d ¼5 for Re ¼20 430. (a) Bare
cylinder wake; (b) straked cylinder wake.
T. Zhou et al. / Journal of Fluids and Structures 27 (2011) 903–917 911

1.0

0.8 (a)

0.6 Δz/d = 0.125 Δz/d = 1.25

0.4

0.2

Rv ,v (τ)
0

-0.2

-0.4
Δz/d = 3.125
-0.6
Δz/d = 2.5
-0.8

-1.0
-10 -5 0 5 10
τU∞/d

1.0

0.8 (b)

0.6

0.4 Δz/d = 0.125

0.2 Δz/d = 2.5


Rv ,v (τ)

-0.2
Δz/d = 3.125
Δz/d = 1.25
-0.4

-0.6

-0.8

-1.0
-10 -5 0 5 10
τU∞/d

Fig. 8. Cross-correlation coefficients at x/d¼ 5 for Re ¼ 20 430 as a function of time delay in the stationary bare and straked cylinder wakes, respectively.
(a) Bare cylinder wake; (b) straked cylinder wake.

the cylinder surface, i.e. the tip of the strakes and the alignment of the strake tip with the flow (Constantinides and Oakley,
2006). The peak in Fig. 7(b) at fd/UN ¼0.21(for Dz/d¼ 1.25) corresponds to the separation at the cylinder surface while that
at fd/UN ¼0.14 (for Dz/d ¼0.125 and 3.125) correspond to separation from the tips of the strakes, which resemble the wake
of a flat plate. This mechanism of the straked cylinder is in contrast to the bare cylinder for which the flow separation
points along the span are more or less fixed, resulting in a coherent vortex shedding (Constantinides and Oakley, 2006).
Using the spanwise vorticity iso-surface, Constantinides and Oakley (2006) showed the separation lines and the disorgan-
ized vorticity in the straked cylinder wake. The wake has a strong three dimensional character without any coherent
vortex structures. This result may indicate that the strakes have disrupted the large scale structures from the cylinder.
Bearman and Branković (2004) suggested that the three-dimensionality of the separating flow introduced by the strakes
could destroy the regular vortex shedding and hence suppress the VIV of the cylinder.

3.4. Cross-correlations in the stationary cylinder wakes

Previous studies have shown that there is a phase shift among the vortex structures shed from a bare cylinder
(e.g. Alam et al., 2002; Fujisawa et al., 2004; Roshko, 1954; Szepessy, 1994; So et al., 2005; Xu et al., 2008). To examine the
phase shift between the vortex structures, time series of the fluctuating quantities such as the pressure, velocity or
vorticity at a single point may not show the phase variation along the cylinder axis clearly. For this purpose, the cross-
correlation coefficient of velocity or pressure measured at two points along the cylinder axis as a function of time delay can
be used. This method has been used previously by a number of researchers (e.g. Fujisawa et al., 2004; Szepessy, 1994;
912 T. Zhou et al. / Journal of Fluids and Structures 27 (2011) 903–917

So et al., 2005; Xu et al., 2008). The cross-correlation between two velocity signals measured at two points separated by a
distance Dz along the cylinder axial direction as a function of time delay t is defined as
/v1 ðz,tÞv2 ðz þ Dz,t þ tÞS
Rv1 ,v2 ðtÞ ¼ , ð3Þ
sv1 sv2
where sv1 and sv2 are the standard deviations of the velocity components v1 and v2, respectively. The cross-correlation
coefficients measured in the stationary bare and straked cylinder wakes for Re ¼20 430 as a function of time delay are
shown in Fig. 8. It can be seen that there is a very small phase difference among the dominate vortex structures for the
bare cylinder (Fig. 8(a)). The maximum phase difference is about 0.25d over a spanwise length of 3.125d, which
corresponds to 0.1p. This behavior of the vortical structures in the bare cylinder wake may imply a large dynamic force
induced on the cylinder and hence a large response in VIV. In contrast, the phase variation of the velocity signals in the
straked cylinder wake is apparent at all probe separations (Fig. 8(b)). This variation of the phase in the spanwise direction
might be caused by the oblique shedding in the presence of helical strakes. The maximum phase variation is about 1.8d,
which corresponds to 0.65p. This result implies a significant phase variation of the vortical structures along the spanwise
direction in the vortex shedding process. It may suggest that helical stakes prevent the vortex from being correlated along
the spanwise direction, in agreement with Bearman and Branković (2004). The variation in phase is because of the
different geometry of the helical strakes along the cylinder spanwise direction (i.e. position of the helical strakes and the
tip location of the strakes). The above results confirm that the strakes do not necessarily suppress vortex shedding from
the cylinder but strongly disorganize the vortices, resulting in an apparent reduction in VIV.
The cross-correlation can be used to quantify the three-dimensional flow characteristics of the vortical structures in the
wake of the cylinder in the spanwise direction. Using two hot wire probes, with one probe being stationary at location 1
and the other being movable at location 2 separated in the spanwise direction by Dz, the cross-correlation coefficient
between the velocity signals of the two probes can be calculated. It has been shown clearly in Fig. 8(b) that there exists
apparent phase shift for the vortical structures in the straked cylinder wake. To account for the phase shift in evaluating
the cross-correlation coefficient, the following definition is used:
 
 o v1 ðz,tÞv2 ðz þ Dz,t þ tÞ 4 
rv1 ,v2 ðDzÞ ¼ max , ð4Þ
ð1=2f Þ r t r ð1=2f Þ sv1 sv2
where f is the vortex shedding frequency. Using this definition, the influence of phase shift, which may be caused, for
example, by oblique shedding, or vortex dislocation, can be avoided. The cross-correlation coefficients between v1 and v2,
as defined by Eq. (4) are shown in Fig. 9. It can be seen that all the cross-correlation coefficients decrease with the increase
in probe separation. The magnitude of rv1,v2(Dz) in the bare cylinder wake is much larger than that in the straked cylinder
wake, indicating that the vortical structures in the wake of the bare cylinder are much larger than that in the straked
cylinder wake. Indeed, if the correlation length Lv1,v2 is evaluated, it is found that the correlation length in the bare cylinder
wake is much larger than that in the straked cylinder wake. There are a few definitions for the correlation length. The most
common one is using the following integral (Norberg, 2003):
Z L0
Lv1 ,v2 ¼ rv1 ,v2 ðDzÞdz: ð5Þ
0

1.0

bare cylinder

0.5
ρv ,v (Δz)

straked cylinder

0 1 2 3
Δz/D

Fig. 9. Cross-correlation coefficients of the velocity components in the stationary bare and straked cylinder wakes obtained at x/d ¼5 for Re ¼20 430.
T. Zhou et al. / Journal of Fluids and Structures 27 (2011) 903–917 913

The integration upper limit L0 can be chosen either as infinity or a finite value, such as the full length of the cylinder.
However, at large separations and for turbulent shedding conditions, the cross-correlation coefficient is expected to vanish
(Norberg, 2003). There are also some other definitions about the correlation length, such as by Xu et al. (2008), who
defined it as the separation where the cross-correlation coefficient is equal to 0.5. In the present study, we use the latter
definition for the purpose to compare the correlation length of the vortical structures in the two wakes. In Fig. 9,
corresponding to rv1,v2(Dz)¼0.5, the correlation Lv1,v2 for the bare cylinder is about 2.4d. In contrast, the correlation length
Lv1,v2 for the straked cylinder is only about 0.3d, which is about 1/8 of that obtained in the bare cylinder wake. This result
suggests that the strakes have successfully disrupted the vortical structures in the spanwise direction and thus enhancing
the three-dimensionality of the flow. This result is in agreement with that proposed by Bearman and Branković (2004),
who suggested that the strakes do not necessarily suppress vortex shedding but they prevent the shedding from becoming
correlated along the span.

3.5. Flow visualization of the stationary cylinder wakes

To further examine the differences of the large scale structures in the wakes of the bare cylinder and the straked
cylinder, flow visualization in a wind tunnel was conducted using a smoke wire at a Reynolds number of about 300. For the
bare cylinder wake (Fig. 10(a)), the large organized vortices can be seen clearly with two rows of counter-rotating vortex
structures connected by the streamwise rib-like structures. The two rows of counter-rotating vortices interact with each
other when they evolve downstream. The formation length of the wake is about 2d, in agreement with that proposed by
Unal and Rockwell (1988). Flow visualization on the x–z plane (Fig. 10(b)) shows approximately parallel shedding even
though vortex dislocation may occur. The spanwise vertical structures persist a long distance in the streamwise direction.
For the straked cylinder wake, there were a few flow visualization results reported previously by Bearman and Branković
(2004), Constantinides and Oakley (2006) and Korkischko and Meneghini (2010). Bearman and Branković (2004) did not
find any evidence in their study about Kármán vortices but they claimed undulations of the wake at about 2d downstream.
Therefore, they concluded that there was not regular vortex shedding in their study of the straked cylinder wake (strake
height 0.12d). Using numerical simulations, Constantinides and Oakley (2006) (strake height 0.25d) showed that in the
lock-in region, the VIV is nearly completely suppressed by the strakes while in the higher reduced velocity region, the
response of the straked cylinder starts to increase in magnitude even though remains much smaller than the bare cylinder
one. Flow visualization showed significant spanwise motion and swirling in the near wake, which might be caused by the
fluid being channeled by the strakes but did not show any organized vortices even at high reduced velocity. They proposed
that the strakes control the separation points, which occur either at the tip of the strakes or on the surface of the cylinder.
The experimental results of Korkischko and Meneghini (2010) showed that for strake cylinder wake, the two shear layers
do not interact with each other, resulting in the absence of the oscillating wake. What is more, the helical configuration
disrupts the regular shedding of vortices along the span. The present visualization results for the straked cylinder wake in

(a) Bare cylinder x-y plane (c) Straked cylinder x-y plane

(d) Straked cylinder x-z plane


(b) Bare cylinder x-z plane

Fig. 10. Flow visualization of the bare and straked cylinder wakes on different planes. The arrows indicate the direction of the swirling flow.
914 T. Zhou et al. / Journal of Fluids and Structures 27 (2011) 903–917

the x–y plane (Fig. 10(c)) show clearly the two shear layers where smaller-scale vortices are found. These vortex structures
are much smaller than the Karman vortices as shown in Fig. 10(a). They interact much farther downstream from the
cylinder, implying that the strakes prevent the interaction between the two shear layers. They do not roll up as occurring
for the bare cylinder to form Karman-like structures. Instead, they decay quickly as evolving downstream. While the flow
visualization in the x–z plane (Fig. 10(d)) also shows that vortices are generated initially, they are broken down and are
dislocated quickly. At the same time, the vortices also swirl as they evolve downstream. These results are in agreement
with those reported by Constantinides and Oakley (2006) based on numerical simulations.

3.6. Isotropy assessments in the wake of stationary cylinders

As the strakes have effectively disrupted the large organised structures in the wake of the straked cylinder, it is
therefore expected that the energy containing structures in the straked cylinder wake should be smaller than that in the

Table 1
Velocity ratio /v2 S=/u2 S at different downstream locations in the stationary bare and straked cylinder wakes.

/v2 S=/u2 S x/d

5 10 20 30 40

Straked cylinder 0.96 1.33 1.23 1.07 0.99


Bare cylinder 1.89 1.4 0.78 0.71 0.7

6 6
10 10
(a) x/d = 10 (b) x/d = 20
m
4
m
φv 4
φv
10 10
cal

cal

cal
φv cal
φv
φu(k1), φv(k1), φv /φv

φu(k1), φv(k1), φv /φv


m

Re = 10240 Re = 10240
20430
*

2 2 20430
10 10
30610 30160
40800 40800
*

0 0
10 10

-2 -2
10 10
-4 -3 -2 -1 0 -4 -3 -2 -1 0
10 10 10 10 10 10 10 10 10 10
* *
k1 k1

5 5
10 10
m
φv (c) x/d = 30 m (d) x/d = 40
φv
cal
3 φv 3
cal
φv
10 10
cal

cal

Re = 10240
φu(k1), φv(k1), φv /φv

φu(k1), φv(k1), φv /φv

Re = 10240
20430
m

20430
30610 30610
*

1 40800 1
10 10 40800
*

-1 -1
10 10

-3 -3
10 10
-4 -3 -2 -1 0 -4 -3 -2 -1 0
10 10 10 10 10 10 10 10 10 10
k*1 k*1
cal m m cal
Fig. 11. Velocity spectra fv and fv and the ratio fv =fv between the measured and calculated transverse velocity spectra in the stationary straked
cylinder wake at different downstream locations and Reynolds numbers.
T. Zhou et al. / Journal of Fluids and Structures 27 (2011) 903–917 915

bare cylinder wake. As a result, the flow field of the former should agree more favorably with isotropy than the latter.
This is true as the ratio /v2 S=/u2 S in the former wake is closer to 1 than the latter, irrespective of the downstream
locations. Isotropy requires that /v2 S=/u2 S be 1. The ratio at different downstream locations for stationary body of bare
cylinder and straked cylinder wakes are given in Table 1. It can be seen clearly that for the bare cylinder wake, /v2 S=/u2 S
is far from 1. Even in the far wake, Browne et al. (1987) (at x/d¼420) and Zhu and Antonia (1998) (at x/d¼240) found that
the ratio /v2 S=/u2 S is only around 0.73, indicating that the effect of the large organised structures persists for a long
time. A measure to check the departure from isotropy for various turbulent scales is provided by comparing the measured
spectrum of transverse or spanwise velocity component with that calculated based on the isotropic assumption. For
isotropic turbulence, the relationship between the streamwise and transverse (or spanwise) velocity spectra is (Zhu and
Antonia, 1995)
 
1 @f
fcal cal
v ðk1 Þ ¼ fw ðk1 Þ ¼ fu k1 u , ð6Þ
2 @k1

where the superscript cal represents ‘‘calculation’’ and k1 is the wavenumber. In Eq. (6), the longitudinal velocity spectra
fu measured on the centreline at different streamwise locations are used as the input for the calculation of fv or fw. The
m cal cal m
distributions of fv , fv and the ratio fv =fv at different streamwise locations and different Reynolds numbers in the
straked and bare cylinder wakes are shown in Figs. 11 and 12, respectively, where the superscript m represents
‘‘measured’’ and the superscript asterisk denotes normalization by Kolmogorov length scale Z and/or velocity scale UK ( n/Z).
cal m
When isotropy is satisfied, fv =fv should be 1 at all scales. At x/d ¼10 of the straked cylinder wake (Fig. 11(a)), the
departure from isotropy for scales k1 o 0:01 is apparent. The peak at k1 ¼ 0:004 is due to the vortex shedding. With the
increase in Re, the peak decreases and shifts towards the lower wavenumber region. For wavenumbers higher than 0.01,
there is satisfactory agreement between measurement and calculation, indicating satisfactory agreement with isotropy at

6 5
10 10
cal
φv
(a) x/d = 10 m
φv (b) x/d = 20
cal
φv m
4
φv 3
10 10
cal

cal

Re = 10240
φu(k1), φv(k1), φv /φv

φu(k1), φv(k1), φv /φv

Re = 10240 20430
m

20430 30610
30610 40800
*

2 1
10 10
40800
*

0 -1
10 10

-2 -3
10 10
-4 -3 -2 -1 0 -4 -3 -2 -1 0
10 10 10 10 10 10 10 10 10 10
* *
k1 k1

5 5
10 10
cal cal
φv (c) x/d = 30 φv (d) x/d = 40
m m
3
φv 3 φv
10 10
cal

cal
φu(k ), φv(k ), φv /φv

φu(k1), φv(k1), φv /φv

Re = 10240 Re = 10240
m

20430 20430
30610 30610
1
*

1
*

1
10 40800 10 40800
1
*

-1 -1
10 10

-3 -3
10 10
-4 -3 -2 -1 0 -4 -3 -2 -1 0
10 10 10 10 10 10 10 10 10 10
* *
k1 k1

cal m m cal
Fig. 12. Velocity spectra fv and fv and the ratio fv =fv between the measured and calculated transverse velocity spectra in the stationary bare
cylinder wake at different downstream locations and Reynolds numbers.
916 T. Zhou et al. / Journal of Fluids and Structures 27 (2011) 903–917

these scales. The upturn at around k1 ¼ 1 is caused by the imperfect spatial resolution of the X-wire probe (Zhu and
Antonia, 1995). At x/d ¼20 (Fig. 11(b)), the dependence of the ratio on Reynolds number is comparable to that at x/d ¼10.
The magnitude of the ratio at the vortex shedding frequency is still very apparent. Similar to that at x/d ¼10, the peak shifts
to lower wavenumbers as Re increases. At x/d ¼30 and 40, vortex shedding is not as apparent as that at x/d¼ 10 and 20. The
cal m
peaks in Fig. 11(c) and (d) are much less apparent. The distribution of fv =fv is quite flat and close to the horizontal line
cal m
of fv =fv ¼ 1 at all wavenumbers. The results shown in Fig. 11 indicate that the turbulent structures in the straked
cylinder wakes agree well with isotropy except in the region x/dr20, which is consistent with the results shown in
Table 1.
The vortex shedding phenomena in the bare cylinder wake at x/d r20 is very apparent (Fig. 12(a) and (b)). As a result,
cal m
the ratio fv =fv at these two locations departs significantly from the isotropic value for scales larger than the most energy
containing structures. It is true that at these two locations, the values of the ratio /v2 S=/u2 S are 1.89 and 1.4,
respectively (Table 1). With the increase in Reynolds number, the peak reduces and shifts to the lower values of the
wavenumbers, indicating that the influence from the large organised structures to the smaller scale turbulent structures
become less dominated as Reynolds number increases. With the evolution of the downstream locations, the influence from
cal m
the large scale structures persists, as reflected by the apparent departure from fv =fv ¼ 1 at the small wavenumber
region (corresponding to the large scale structures). Even at x/d ¼40 where the large scale structures disappear due to
cal m
vortex break down (e.g. Zhou et al., 2003), the ratio fv =fv is far from 1 for k1 o0:004, reflecting the departure from
isotropy even when there is no apparent large organised structures detected at this location. From Figs. 11 and 12, it can be
inferred that the turbulent structures in the straked cylinder comply more with isotropy than that in the bare cylinder
cal m
wake. With the evolution of the downstream locations, vortex structures break down and the ratio fv =fv in the straked
cylinder wake complies more favorably with isotropy than in the bare cylinder wake.

4. Conclusions

In the present study, the effect of three-strand helical strakes with a dimension of 10d in pitch and 0.12d in height on
VIV suppression are studied experimentally in a wind tunnel. For the bare cylinder, lock-in occurs over the range of
reduced velocity of 5–8.5 with the maximum vibration amplitude of 0.51. This value agrees well with that estimated using
the empirical relation. However, after three-start helical strakes are attached to the cylinder, vortex shedding is not
apparent and lock-in phenomena of the cylinder does not occur. The vibration amplitude of the cylinder is suppressed
by 98%.
In exploring the mechanism of helical strakes have on VIV suppression, measurements were conducted using both a
single X-probe in the streamwise direction and two X-probes separated in the spanwise direction in the stationary bare
and straked cylinder wakes. The peak frequency on the energy spectra is very stable for the bare cylinder both in the
streamwise and in the spanwise directions. In contrast, the peak frequency on the spectra of the straked cylinder wake
varies significantly in both directions. Even though flow visualization shows smaller-scale vortices (compared with the
Karman-type vortices in the bare cylinder) are generated in the shear layers of the straked cylinder wake, they do not roll
up and interact with each other to form the well-organised Karman-type vortices. Instead, they decay quickly in the
streamwise direction. At the same time, vortex dislocation and swirling are found, which should be responsible for the
variations of the peak frequency in the streamwise as well as spanwise directions.
The cross-correlation coefficient in the straked cylinder wake is much smaller than that in the bare cylinder wake. The
correlation length of the vortical structures in the former is only about 1/8 of the latter. As a result, the straked cylinder
wake agrees better with isotropy than that in the bare cylinder wake. This is verified by the comparison of the measured
transverse velocity spectra with the calculation based on isotropy.

Acknowledgments

T. Zhou would like to thank the Australian Research Council for support under a discovery project (DP110105171).
Z. Hao acknowledges the financial supported by Science and Technology Commission of Shanghai Municipality (Pujiang
Programme, Grant no. 10PJ1404700) and Shanghai Maritime University (Science & Technology Program, Grant no.
20100089). The authors are grateful to the suggestion of Eq. (4) by one of the anonymous reviewers.

References

Alam, M.M., Moriya, M., Takai, K., Sakamoto, H., 2002. Suppression of fluid forces acting on two square prisms in a tandem arrangement by passive control
of flow. Journal of Fluids and Structures 16 (8), 1073–1092.
Antonia, R.A., Zhou, T., Romano, G.P., 2002. Small-scale turbulence characteristics of two dimensional bluff body wakes. Journal of Fluids Mechanics 459,
67–92.
Antonia, R.A., Zhou, T., Zhu, Y., 1998. Three-component vorticity measurements in a turbulent grid flow. Journal of Fluids Mechanics 374, 29–57.
Bearman, P.W., 1969. On vortex shedding from a circular cylinder in the critical Reynolds number regime. Journal of Fluids Mechanics 37, 577–586.
Bearman, P., Branković, M., 2004. Experimental studies of passive control of vortex-induced vibration. European Journal of Mechanics—Fluids 23, 9–15.
Bearman, P.W., Owen, J.C., 1998. Reproduction of bluff-body drag and suppression of vortex shedding by the introduction of wavy separation lines.
Journal of Fluids and Structures 12, 123–130.
T. Zhou et al. / Journal of Fluids and Structures 27 (2011) 903–917 917

Blevins, R.D., 2001. Flow-Induced Vibration, second ed. Krieger Publishing, Inc., Malabar/Florida, USA.
Browne, L.W.B., Antonia, R.A., Shah, D.A., 1987. Turbulent energy dissipation in a wake. Journal of Fluids Mechanics 179, 307–326.
Carberry, J., Sheridan, J., Rockwell, D., 2005. Controlled oscillations of a cylinder: forces and wake modes. Journal of Fluids Mechanics 538, 31–69.
Castro, I.P., Rogers, P., 2002. Vortex shedding from tapered plates. Experiments in Fluids 33, 66–74.
Constantinides, Y., Oakley, Jr., O.H., 2006. Numerical prediction of bare and straked cylinder VIV. In: Proceedings of 25th International Conference on
Offshore Mechanics and Arctic Engineering, June 4–9, Hamburg, Germany.
Ding, Z.J., Balasubramanian, S., Lokken, R.T., Yung, T.-W., 2004. Lift and damping characteristics of bare and straked cylinders ar riser scale Reynolds
numbers. In: Proceedings of the Offshore Technology Conference.
Elston, J.R., Blackburn, H.M., Sheridan, J., 2006. The primary and secondary instabilities of flow generated by an oscillating circular cylinder. Journal of
Fluids Mechanics 550, 359–389.
Feng, C.C., 1968. The Measurement of Vortex-Induced Effects in Flow Past Stationary and Oscillating Circular and D-Section Cylinders. MA.Sc. Thesis.
University of British Columbia.
Franzini, G.R., Fujarra, A.L.C., Meneghini, J.R., Korkischko, I., Franciss, R., 2009. Experimental investigation of vortex-induced vibration on rigid, smooth
and inclined cylinders. Journal of Fluids and Structures 25, 742–750.
Fujisawa, N., Takeda, G., Ike, N., 2004. Phase-averaged characteristics of flow around a circular cylinder under acoustic excitation control. Journal of Fluids
and Structures 19, 159–170.
Hover, F.S., Triantafyllou, M.S., 2001. Vortex-induced vibrations of a cylinder with tripping wires. Journal of Fluids Mechanics 448, 175–195.
Jones, G.S., Lamb, W.S., 1992. The use of helical strakes to suppress vortex induced vibration. In: Proceedings of the Sixth International Conference on
Behaviour of Offshore Structures, London.
Khalak, A., Williamson, C.H.K., 1996. Dynamics of a hydroelastic cylinder with very low mass and damping. Journal of Fluids and Structures 10, 455–472.
Korkischko, I., Meneghini, J.R., 2010. Experimental investigation of flow-induced vibration on isolated and tandem circular cylinders fitted with strakes.
Journal of Fluids and Structures 26, 611–625.
Kumar, R.A., Sohn, C.-H., Gowda, H.L., 2008. Passive control of vortex-induced vibration: an overview. Recent Patents on Mechanical Engineering 1, 1–11.
Norberg, C., 2003. Fluctuating lift on a circular cylinder: review and new measurements. Journal of Fluids and Structures 17, 57–96.
Owen, J.C., Bearman, P.W., 2001. Passive control of VIV with drag reduction. Journal of Fluids and Structures 15, 597–605.
Roshko, A., 1954. On the Development of Turbulent Wakes from Vortex Streets. NACA Report 1191.
Sarpkaya, T., 1979. Vortex induced oscillations: a selective review. Journal of Applied Mechanics 46 (2), 241–258.
Sarpkaya, T., 2004. A critical review of the intrinsic nature of vortex-induced vibrations. Journal of Fluids and Structures 19, 389–447.
Scruton, C., Walshe, D.E., 1963. Stabilisation of wind-excited structures. United States Patent Office.
So, R.M.C., Liu, Y., Cui, Z.X., Zhang, C.H., Wang, X.Q., 2005. Three-dimensional wake effects on flow-induced forces. Journal of Fluids and Structures 20,
373–402.
Sumer, B.M., Fredsøe, J., 1997. Hydrodynamics around Cylindrical Structures. World Scientific, Singapore.
Szepessy, S., 1994. On the spanwise correlation of vortex shedding form a circular cylinder at high subcritical Reynolds number. Physics of Fluids 6,
2406–2416.
Unal, M.F., Rockwell, D., 1988. On vortex formation from a cylinder. Part 1. The initial instability. Journal of Fluids Mechanics 190, 491–512.
Vukoslavčevié, P.V., Beratlis, N., Balaras, N., Wallace, E., Sun, O., J.M., 2009. On the spatial resolution of velocity and velocity gradient-based turbulence
statistics measured with multi-sensor hot-wire probes. Experiments in Fluids 46, 109–119.
Williamson, C.H.K., Govardhan, R., 2004. Vortex-induced vibrations. Annual Review of Fluid Mechanics 36, 413–455.
Xu, G., Zhou, Y., 2004. Strouhal numbers in the wake of two incline cylinders. Experiments in Fluids 37, 248–256.
Xu, S.J., Zhou, Y., Tu, J.Y., 2008. Two-dimensionality of a cantilevered-cylinder wake in the presence of an oscillating upstream cylinder. Journal of Fluids
and Structures 24, 467–480.
Zdravkovich, M.M., 1981. Review and classification of various aerodynamic and hydraodynamic means for suppressing vortex shedding. Journal of Wind
Engineering and Industrial Aerodynamics 7, 145–189.
Zhou, T., Zhou, Y., Yiu, M.W., Chua, L.P., 2003. Three-dimensional vorticity in a turbulent cylinder wake. Experiments in Fluids 35, 459–471.
Zhu, Y., Antonia, R.A., 1995. Effect of wire separation on X-probe measurements in a turbulent flow. Journal of Fluids Mechanics 287, 199–223.
Zhu, Y., Antonia, R.A., 1998. Performance of a three-component vorticity probe in a turbulent far-wake. Experiments in Fluids 287, 21–30.

You might also like