Sei sulla pagina 1di 16

Journal of Wind Engineering

and Industrial Aerodynamics 91 (2003) 139–154

Fluctuating fluid forces acting on two circular


cylinders in a tandem arrangement at a
subcritical Reynolds number
Md. Mahbub Alam*, M. Moriya, K. Takai, H. Sakamoto
Department of Mechanical Engineering, Kitami Institute of Technology, Kitami 090, Japan

Abstract

Aerodynamic characteristics of two circular cylinders in a tandem arrangement were


investigated experimentally in a uniform flow at a Reynolds number of 6.5  104. This
Reynolds number is within the range in which fluid forces acting on a single cylinder are
comparatively insensitive to change in the Reynolds number. Mutual interference between the
two cylinders at close proximity, however, caused significant change in parameters of the
aerodynamic characteristics, such as fluctuating lift and drag forces, time-averaged and
fluctuating pressure distributions, Strouhal number, and vortex-shedding patterns, when the
spacing between the cylinders was changed. The changes in these aerodynamic characteristics
were systematically analyzed in this study.
r 2002 Elsevier Science Ltd. All rights reserved.

Keywords: Aerodynamic characteristics; Tandem arrangement; Reattachment position

1. Introduction

Most structures on land and in the ocean are in multiple forms and are confronted
by a fluid flow. Vibrations of these structures due to fluid flows reduce the life of the
respective installations and must therefore be taken into account in the design of the
structure. For assessment of this vibration, it is important to understand the
interaction of multiple structures in a flow. An elementary shape of a structure or a
component of a structure is a circular cross-section, and a tandem arrangement of
two circular cylinders is a basic example of an array of multiple structures. The
common use of the cylindrical-shaped body in various fields of engineering

*Corresponding author.

0167-6105/02/$ - see front matter r 2002 Elsevier Science Ltd. All rights reserved.
PII: S 0 1 6 7 - 6 1 0 5 ( 0 2 ) 0 0 3 4 1 - 0
140 Md.M. Alam et al. / J. Wind Eng. Ind. Aerodyn. 91 (2003) 139–154

stimulated the investigation on flow around two circular cylinders in a tandem


arrangement. Previous investigations of tandem configurations by Biermann and
Herrnstein [1], Kostic and Oka [2], Zdravkovich and Pridden [3,4], Novak [5],
Okajima [6], Hiwada et al. [7], Igarashi [8,9], Arie et al. [10], and Jendrxejczk and
Chen [11] have revealed considerable complexity in fluid dynamics as the spacing
between the cylinders is changed.
There are many discrepant points in previous studies on two circular cylinders in a
tandem arrangement. For example, Arie et al. [10] pointed out that fluctuation in
drag force acting on both cylinders is weakly dependent on spacing. On the other
hand, Igarashi [8] reported that the fluctuation in pressure associated with
fluctuation in lift and drag forces acting on a downstream cylinder is strongly
dependent on spacing between the cylinders. Moreover, some discrepant points
among various studies can be found in [8]. Despite the large number of studies on
two circular cylinders in a tandem arrangement, there have been very few studies in
which fluctuating lift and drag forces acting on the cylinders have been measured,
and many unknown and discrepant points remain. Hence, the aim of this study was
to examine the characteristics of fluctuating aerodynamic forces acting on two
circular cylinders in a tandem arrangement, to elucidate the discrepant points and to
clarify the flow patterns over the cylinders.
The present study, on the relatively simple case of two stationary circular cylinders
of equal diameters in a tandem arrangement subjected to a steady cross flow, was
motivated by both fundamental and practical considerations.

2. Experimental arrangement and procedures

Experiments were conducted in a low-speed, closed-circuit wind tunnel with a test


section of 0.6 m height, 0.4 m width, and 5.4 m length. The level of turbulence in the
working section was 0.19%. The cylinders used as test models were made of brass
and were each 49 mm in diameter. The geometric blockage ratio and aspect ratio at
the test section were 8.1% and 8.2, respectively. None of the results presented were
corrected for the effects of wind-tunnel blockage.
Fluid forces acting on the cylinders were measured by using two load cells installed
inside a cylinder. The details of the load cells and measurement procedure of fluid
forces have been described by Sakamoto et al. [12]. The sensitivity of each load cells
was 11.311 mV/g. To measure the surface pressure during experiments, a
semiconductor pressure transducer (TOYODA PD104 K) with a range of 710 kPa
was used, and the transducer output was calibrated to give a reading of 6.22 V for
1 kPa of applied pressure. The pressure transducer responded to pressure fluctuation
up to 500 Hz with a gain factor of 170.06, the phase lag being negligible.
Surface oil-film techniques were used to investigate the flow patterns on the
cylinders. A mixture composed of silicone oil, titanium dioxide, oleic acid and
kerosene with a ratio of 45:3:2:2 in weight was used for surface oil-flow visualization.
A cylinder wrapped in a black film of 0.03 mm in thickness was uniformly smeared
with the mixture, and then the cylinder was placed inside the wind tunnel to obtain a
Md.M. Alam et al. / J. Wind Eng. Ind. Aerodyn. 91 (2003) 139–154 141

surface oil-flow pattern. Finally, the film was carefully unwrapped from the cylinder,
and a photograph was taken with a digital camera.

3. Results and discussion

3.1. Time-averaged drag force

The placement of the two circular cylinders in a tandem arrangement, the


coordinate system, and the definition of the symbols are shown in Fig. 1. Fig. 2
shows the effect of the tandem spacing on the time-averaged drag coefficient, CD ; of
the upstream cylinder and that of the downstream cylinder. The results obtained by
Biermann and Herrnstein [1] and Zdravkovich and Pridden [4] are also included in
the same figure for comparison. First, for L=Do3; CD of the upstream cylinder
decreases gradually with increase in spacing between the cylinders, because the
existence of the downstream cylinder behind the upstream one increases the after
body length of the upstream cylinder. On the other hand, CD of the downstream
cylinder is negative (forward thrust) for L=Do3:0 and forms a peak at L=D ¼ 1:40:
This type of peak was also found in previous studies [3,7,8]. The well-known bistable
flow occurs at L=D ¼ 3:0 (critical spacing), where two values of CD are seen for two
different flow patterns; namely, reattachments flow and jump flow.

3.2. Time-averaged pressure

Fig. 3 shows the distribution of pressure coefficient, Cp ; along the surface of the
upstream cylinder for various spacings, including the results for a single cylinder.
The magnitude of pressure becomes zero at y ¼ 341 and becomes maximum negative
at y ¼ 691 for the single cylinder. The surface oil-flow pattern shows that laminar
separation occurs at y ¼ 751 in the case of the single cylinder. This angular position
of the separation point obtained from the surface oil-flow pattern agrees well with
the position of the separation point estimated by Achenbach [13], and by Son and
Hanratty [14]. The rear parts of all of the curves are flat except for L=D ¼ 0:10 and
0.30. There are peaks at y ¼ 1601 for L=D ¼ 0:10 and at y ¼ 1791 for L=D ¼ 0:30:

Fig. 1. Two cylinders in a tandem arrangement, the coordinate system, and definitions of the symbols.
142 Md.M. Alam et al. / J. Wind Eng. Ind. Aerodyn. 91 (2003) 139–154

Fig. 2. Effects of tandem spacing on time-averaged drags of two cylinders.

Fig. 3. Time-averaged pressure distribution along the surface of the upstream cylinder.

For L=Do3:0; the shear layer separated from the upstream cylinder reattaches to the
surface of the downstream cylinder and bifurcates into two shear layers at the
reattachment point, as shown in Fig. 4. One shear layer continues in the downstream
direction (hereafter referred to as the backward shear layer), and the other shear
layer flows in the upstream direction (hereafter referred to as the forward shear
layer). The peaks indicated above are due to reattachment of the forward shear layer.
Md.M. Alam et al. / J. Wind Eng. Ind. Aerodyn. 91 (2003) 139–154 143

Fig. 4. Sketch and photographs of surface oil-flow patterns for L=D ¼ 0:10:

Fig. 5. Time-averaged pressure distribution along the surface of the downstream cylinder.

For L=D ¼ 0:10; the shear layer separates from the upstream cylinder at y ¼ 741 and
attaches to the downstream cylinder at y ¼ 751: The backward shear layer separates
at y ¼ 1091 and the forward shear layer separates at y ¼ 351: The forward shear
layer separating from the downstream cylinder reattaches again to the rear surface of
the upstream cylinder at y ¼ 1591 and bifurcates. One part of the bifurcated shear
layer separates at y ¼ 1361 and the other part separates at y ¼ 1741: Reattachment
of the forward shear layer behind the upstream cylinder occurs up to a spacing ratio
of L=D ¼ 0:50:
The distributions of pressure coefficients along the surface of the downstream
cylinder are shown in Fig. 5. For L=Dp3:50; the pressure on the whole surface of the
144 Md.M. Alam et al. / J. Wind Eng. Ind. Aerodyn. 91 (2003) 139–154

downstream cylinder is negative, and for L=Do3:0; a peak on the front side of the
cylinder is apparent in each of the pressure distributions. This peak represents the
reattachment of the shear layer that separates from the upstream cylinder.

3.3. Fluctuating pressure

In order to obtain a clearer understanding of the flow patterns around the


cylinders, rms pressure coefficient, Cpf ; on the surface of each cylinder was measured
for certain spacings, and the results are shown in Fig. 6. In the case of a single
cylinder, there is a sharp peak at y ¼ 751: As has already been presented in the

Fig. 6. Distribution of fluctuating pressure along the surface of the cylinders: (a) upstream cylinder,
(b) downstream cylinder.
Md.M. Alam et al. / J. Wind Eng. Ind. Aerodyn. 91 (2003) 139–154 145

previous section that the separation of the shear layer on a single cylinder occurs at
y ¼ 751: Therefore, the peak in the Cpf distribution occurs at or very near the
separation point of the shear layer in the case of a laminar separation. The results for
other spacings also concur well with the above conclusion. Batham [15] also
compared the separation position measured by Son and Hanratty [14] with the
position of the peak in the fluctuating pressure distribution and noted that the
separation position coincides within experimental accuracy with the position of
the peak in the fluctuating pressure distribution. For L=D ¼ 2:50; where the shear
layers separated from the upstream cylinder reattach steadily to the surface of the
downstream cylinder, the peaks disappear and the profile becomes flat.
The pressure fluctuation on the surface of the downstream cylinder is very large
compared with that for a single cylinder or corresponding upstream cylinder. In the
case of L=D ¼ 0:30; three peaks are self-evident in the fluctuating pressure
distribution. The middle peak is near the reattachment of the shear layer that
separates from the upstream cylinder, and the first peak is due to the separation of
the forward shear layer. The backward shear layer separation causes the third peak
on the rear surface. The fluctuating pressure on the whole surface of the cylinder
increases with increase in spacing up to L=D ¼ 1:40: At this spacing, Cpf on the
whole surface becomes large (compared with that at L=Do3) and decreases again
with increase in spacing. Beyond the critical spacing where two vortex streets are
developed, say, at L=D ¼ 3:50 and 6.0, the pressure fluctuation has fairly large
values in the range of y ¼ 201B601 and even exceeds that of the rear surface on
which pressure fluctuation is usually larger than that of the front surface due to
periodic vortices. On seeing Cpf distribution for other spacing also (not shown), it
can be decided that the buffet of the incoming periodic vortex from the upstream cylinder
to the downstream cylinder causes a higher Cpf value in the range of y ¼ 201B601:
The peaks in the Cpf distribution are sharp for L=D in the range of 0oL=Do2:0;
but not so sharp for L=D in the range of 2:0oL=Do3:0 (not shown in all of the Cpf
distributions). This is due to the fact that the shear layer separated from the
upstream cylinder reattaches alternately for L=D in the range of 0oL=Do2:0 and
steadily for L=D in the range of 2:0oL=Do3:0: Sketches for L=D ¼ 1:40 and L=D ¼
2:50 are shown in Fig. 7 as representative sketches of the ranges of 0oL=Do2:0 and
2:0oL=Do3:0; respectively. Photographs of flow visualization performed by
Lakshmana and Prabhu [16] also support the existence of the alternate reattachment
and steady reattachment flows in those ranges of spacing. The alternate and steady
reattachments of the shear layers can also be corroborated from the power spectrum
distribution of fluctuating pressure on the surface of the downstream cylinder
(Fig. 8). In the case of alternate reattachment, a superharmonic frequency of twice
the Strouhal frequency is observed on the side surface. In the case of steady
reattachment, only Strouhal frequency appears.

3.4. Fluctuating lift and drag forces

Variations in the fluctuating lift coefficient, CLf ; and the fluctuating drag
coefficient, CDf ; plotted against the spacing ratio, L=D; are shown in Fig. 9, together
146 Md.M. Alam et al. / J. Wind Eng. Ind. Aerodyn. 91 (2003) 139–154

Fig. 7. Photographs of the surface oil-flow and sketch of flow pattern: (a) L=D ¼ 1:40; (b) L=D ¼ 2:50:

with the values measured by Arie et al. [10]. Before the critical spacing, CLf and CDf
of the upstream cylinder are very small compared with those for the downstream
cylinder or those for a single cylinder. As the shear layer that separates from the
upstream cylinder reattaches to the surface of the downstream cylinder and finally, a
Karman-type vortex is formed only behind the downstream cylinder, CLf and CDf of
the upstream cylinder are very small. The figures show that CLf and CDf of the
downstream cylinder are very sensitive to the spacing between the cylinders,
especially before the critical spacing. The figures also show that there exist two
noteworthy peaks in each of the figures, at L=D ¼ 0:4 and 1.40. In this regard, it
should be mentioned that Zdravkovich and Pridden [3] performed an experiment on
two flexible cylinders arranged in tandem and observed that both cylinders vibrated
with a large amplitude in the transverse direction when they were placed at L=D ¼
0:25: They also found that there was a mode of large amplitude vibrations for L=D in
the range of 1B2. Comparison of the present results with the results of Arie et al.
[10] shows that there is a basic agreement, but, unfortunately, the data of Arie et al.
are too sparse to show the actual characteristics of CLf and CDf variations.
It is interesting that at and near L=D of 1.40, the values of CLf and CDf are very
large, and in the case of CDf ; the value even exceeds that of the jump flow that occurs
at the critical spacing. The values of CLf and CDf of the downstream cylinder at
Md.M. Alam et al. / J. Wind Eng. Ind. Aerodyn. 91 (2003) 139–154 147

Fig. 8. Power spectrum distribution obtained from fluctuation in pressure on the surface of the
downstream cylinder: (a) L=D ¼ 0:40; (b) L=D ¼ 2:50:

L=D ¼ 1:40; where the second peak occurs, are 2 and 2.8 times larger, respectively,
than those for the single cylinder. The value of CLf of the single cylinder for the pre-
sent case is 0.45, which is equal to the value obtained by Lesage and Gartshore [17]
148 Md.M. Alam et al. / J. Wind Eng. Ind. Aerodyn. 91 (2003) 139–154

Fig. 9. Variation in fluctuating fluid forces with increase in spacing ratio L=D: (a) fluctuating lift
coefficient, CLf ; (b) fluctuating drag coefficient, CDf :

at the same Reynolds number. However, the values obtained by Arie et al. [10] and
Batham [15] are different from the value of the present case. This may be due to the
fact that the fluctuating fluid force acting on a circular cylinder strongly depends on
the Reynolds number [18] and on turbulent intensity. Beyond the critical spacing
where two cylinders form vortices individually, CLf and CDf of the downstream
cylinder decrease as the spacing increases. However, in the case of the upstream
cylinder, the variation in the CLf distribution is undulating, and the amplitude of the
undulated distribution is in both sides of the value of fluctuating lift coefficient of the
single cylinder. CLf reaches maximum at L=D ¼ 3:0 and 6.25, and reaches minimum
Md.M. Alam et al. / J. Wind Eng. Ind. Aerodyn. 91 (2003) 139–154 149

at L=D ¼ 4:5 and 8.0. The occurrence of the maximum and minimum values of CLf
at these spacings must be, as will be discussed later, due to synchronization of the
wakes behind the cylinders with ‘in-phase’ and ‘out-of-phase’ conditions, respec-
tively. It should be mentioned that Sakamoto et al. [19] noted that, in the case of two
square cylinders, CLf becomes maximum at the critical spacing due to synchroniza-
tion of the cylinders with phase a lag of 2p (in phase).

3.5. Reattachment position

As stated above, before the bistable flow spacing, the shear layer that separates
from the upstream cylinder reattaches somewhere to the surface of the downstream
cylinder. The reattachment position, yR ; obtained from surface oil-flow patterns in
which a clear reattachment line was found, is shown in Fig. 10. The figure also shows
the positions of maximum pressure (peak) due to reattachment in the time-averaged
pressure distribution, and the peak due to reattachment in the fluctuating pressure
distributions. The results of Hiwada et al. [7] included in the figure were also taken
from the peak of the time-averaged pressure distribution. From the figure, it is clear
that the maximum pressure due to reattachment of a shear layer in the time-averaged
pressure distribution occurs at the reattachment position or slightly (0B21)
upstream. However, the peak in the fluctuating pressure distribution due to
reattachment does not maintain constant remoteness from the reattachment
position. Also, it is notable that when the shear layers reattach steadily to the
downstream cylinder (i.e., at L=D ¼ 2:0B3:0), the position of maximum pressure
and the position of the peak due to reattachment in the fluctuating pressure
distribution coincide with the reattachment position. By comparing Figs. 9(a), (b)
and 10, it is evident that there is a dormant relation of the values of CLf and CDf
of the downstream cylinder to the reattachment position of the shear layer. For

Fig. 10. Variation in reattachment position with increase in L=D:


150 Md.M. Alam et al. / J. Wind Eng. Ind. Aerodyn. 91 (2003) 139–154

L=D ¼ 0:10; the reattachment position of the shear layers is 751. As the spacing
increases, yR precedes forward sharply and CLf and CDf increase sharply. When the
distribution of yR generates a valley at L=D ¼ 0:4; the CLf and CDf distributions
generate peaks at the same spacings. Again, when yR recedes backward for the
spacings 0:4oL=Do0:8; CLf and CDf decrease, and when yR again precedes forward
for the spacings 0:8oL=Do1:40; CLf and CDf greatly increase. At L=D ¼ 1:40; CLf
and CDf reach maximum values of 0.86 and 0.28, respectively, and the yR reaches a
minimum value of 551. Beyond L=D ¼ 1:40; the reattachment position recedes
backward again. Therefore, it can be concluded that the fluctuating fluid forces
acting on the downstream cylinder strictly depend on the reattachment position and
increase when the reattachment position of the shear layer precedes forward and vice
versa.

3.6. Time-averaged pressure at the reattachment point

Fig. 11 shows the relationship between the magnitude of time-averaged pressure,


CPR ; at the reattachment position and the reattachment position, yR ; of the shear
layers when the spacing between the cylinders is changed. The figure shows that CPR
has an inverse relationship with the position of reattachment; i.e., CPR increases
when yR precedes forward and decreases when yR recedes backward. At L=D ¼ 0:4
and 1.40, where the yR reaches a minimum value, CPR reaches a maximum value. It is
understandable that when a shear layer attaches to a surface with a smaller incidence
angle (the angle between the approaching shear layer and a normal to the surface at
the reattachment point), the pressure at the reattachment point becomes higher.
Thus, when the reattachment position precedes forward, the incidence angle of the
shear layers becomes smaller and vice versa. Also, when a flow approaches a body
with a smaller angle of incidence, the flow experiences the body as a more bluff one,

Fig. 11. Relationship between time-averaged pressure at the reattachment point and position of
reattachment.
Md.M. Alam et al. / J. Wind Eng. Ind. Aerodyn. 91 (2003) 139–154 151

Fig. 12. Variation in phase lag of fluctuating lift forces between the cylinders.

and then the flow behind the body rolls strongly. As CPR reaches a maximum value
at L=D ¼ 1:40; the incidence angle of the shear layers on the surface seems to be a
minimum. Therefore, at and near L=D ¼ 1:40; the alternate buffet of the shear layers
with a smaller angle of incidence and strong rolling of the shear layers cause higher
fluctuating fluid forces.

3.7. Phase lag of the fluctuating lift between two cylinders

Fig. 12 shows the trend in variation of the phase lag of the fluctuating lift between
two cylinders with variation in L=D: The phase lag is calculated from the cross
correlation between the fluctuating lift forces of the upstream and the downstream
cylinders. Though the data are somewhat staggered, a straight line is drawn. The
equation given below fits the experimental results:
fðradÞ ¼ 0:508pðL=DÞ þ 0:85p ðL=DX3Þ:
From the figure, it can be seen that the two cylinders are ‘in phase’ (f ¼ 2np;
n ¼ 1; 2y) with phase lags of 2p and 4p near L=D ¼ 3:0 and 6.25, where the
fluctuating lift force acting on the upstream cylinder is maximum, and the cylinders
are ‘out-of-phase’ ff ¼ ð2n þ 1Þpg with phase lags 3p and 5p at L=D ¼ 4:5 and 8.0,
where the fluctuating lift force is minimum (Section 3.4). Therefore, the phase of the
flow pattern of the downstream cylinder strongly influences the fluctuating lift force
acting on the upstream cylinder, and fluctuating lift force of the upstream cylinder
becomes maximum when the flow patterns of the two cylinders are in phase and
becomes minimum when the flow patterns of the two cylinders are out-of-phase.

3.8. Vortex-shedding frequency from two cylinders

Strouhal numbers of the two cylinders were calculated from a spectral analysis of
the fluctuating lift force acting on the cylinders, and the results are shown in Fig. 13.
152 Md.M. Alam et al. / J. Wind Eng. Ind. Aerodyn. 91 (2003) 139–154

Fig. 13. Variation in Strouhal number with increase in spacing ratio L=D:

Though there was no distinct vortex shedding detectable behind the upstream
cylinder up to L=D ¼ 3; the lift forces acting on the upstream cylinder fluctuate with
the frequency of the alternate shear layer that separates from the upstream cylinder
and reattaches to the downstream cylinder. As has already been stated, the shear
layer reattaches alternately to the downstream cylinder at L=Do2 and reattaches
steadily to the downstream cylinder at 2oL=Do3: Therefore, at 2oL=Do3; there is
no distinct frequency of the oscillating lift of the upstream cylinder.

4. Conclusions

The main results of this investigation may be summarized as follows:


(1) The fluctuating lift and drag forces acting on a downstream cylinder are very
sensitive to the spacing between the cylinders, especially before the critical
spacing. Each of the fluctuating lift and drag coefficient distributions of the
downstream cylinder shows two peaks: one is at L=D ¼ 0:4 and the other is at
L=D ¼ 1:40: The values of CLf and CDf at L=D ¼ 1:40 are remarkable and are
about 2 and 2.8 times higher, respectively, than those for a single cylinder.
(2) The well-known bistable flow occurs at L=D ¼ 3:0 (critical spacing). Before the
critical spacing, the shear layer that seperates from the upstream cylinder
reattaches to the surface of the downstream cylinder. For the present case, the
shear layer reattaches alternately for L=Do2:0 and steadily for 2:0oL=Do3:0:
In the case of alternate reattachment, a superharmonic frequency of twice the
Strouhal frequency can be observed in the power spectrum distribution obtained
from the fluctuating pressure on the side surface.
(3) There exists a clear relationship between the fluctuating fluid forces acting on the
downstream cylinder and the reattachment position, yR ; of the shear layer. The
fluctuating lift and drag forces acting on the downstream cylinder increase when
the reattachment position of shear layer precedes forward and vice versa.
Md.M. Alam et al. / J. Wind Eng. Ind. Aerodyn. 91 (2003) 139–154 153

(4) Beyond the critical spacing, the fluctuating lift force acting on the upstream
cylinder is strictly influenced by the phase of the flow pattern of the downstream
cylinder. When the phase of the flow pattern of the downstream cylinder
coincides with the phase of the flow pattern of the upstream cylinder, the
fluctuating lift force acting on the upstream cylinder becomes maximum, and
when the phase of the flow pattern of the downstream is out-of-phase with the
downstream cylinder’s flow pattern, the fluctuating lift force becomes minimum.

Acknowledgements

The authors are indebted to Mr. Y. Obata, a technical officer, for his expertise in
precise fabrication of the experimental apparatus.

References

[1] D. Biermann, Herrnstein Jr., The interference between struts in various combinations, National
Advisory Committee for Aeronautics, Technical Report 468, 1933.
[2] Z.G. Kostic, S.N. Oka, Fluid flow and heat transfer with two circular cylinders in cross flow, Int. J.
Heat Mass Transfer 15 (1972) 279–299.
[3] M.M. Zdravkovich, D.L. Pridden, Flow around two circular cylinders, Research Report, Proceedings
of the Second US National Conference on Wind Engineering Research, Fort Collins, IV (18), 1975.
[4] M.M. Zdravkovich, D.L. Pridden, Interference between two circular cylinders; series of unexpected
discontinuities, J. Ind. Aerodyn. 2 (1977) 255–270.
[5] J. Novak, Strouhal number of a quadrangular prism, angle iron and two circular cylinders arranged
in tandem, Acta Tech. CSAV 19 (3) (1974) 361–373.
[6] A. Okajima, Flow around two tandem circular cylinders at very high Reynolds numbers, Bull. JSME
22 (166) (1979) 504–511.
[7] M. Hiwada, I. Mabuchi, H. Yanagihara, Flow and heat transfer from two same size circular cylinders
in tandem arrangement, Trans. JSME 48 (1982) 499–508 (in Japanese).
[8] T. Igarashi, Characteristics of the flow around two circular cylinders arranged in tandem (first
report), Bull. JSME 24 (188) (1981) 323–331.
[9] T. Igarashi, Characteristics of the flow around two circular cylinders arranged in tandem (second
report), Bull. JSME 27 (233) (1984) 2380–2387.
[10] M. Arie, M. Kiya, M. Moriya, H. Mori, Pressure fluctuations on the surface of two circular cylinders
in tandem arrangement, ASME J. Fluids Eng. 105 (1983) 161–167.
[11] J.A. Jendrzejczk, S.S. Chen, Fluid forces on two circular cylinders in crossflow, ASME, PVP 14
(1986) 141–143.
[12] H. Sakamoto, H. Haniu, Y. Obata, S. Matubara, Optimum suppression of fluid forces acting on a
circular cylinder and its effectiveness, JSME Int. J. Ser. B 37 (2) (1994) 369–376.
[13] E. Achenbach, Distribution of local pressure and skin friction around a circular cylinder in cross-flow
up to Re=5  106, J. Fluid Mech. 34 (Part 4) (1968) 625–639.
[14] J.S. Son, T.J. Hanratty, Velocity gradient at the wall for flow around a cylinder at Reynolds number
from 5  103 to 105, J. Fluid Mech. 35 (1969) 353–368.
[15] J.P. Batham, Pressure distribution on circular cylinders at critical reynolds numbers, J. Fluid Mech.
57 (1973) 209–228.
[16] B.H. Lakshmana Gowda, D.R. Prabhu, Interference effects on the flow-induced vibrations of a
circular cylinder, J. Sound Vib. 112 (3) (1987) 487–502.
154 Md.M. Alam et al. / J. Wind Eng. Ind. Aerodyn. 91 (2003) 139–154

[17] F. Lesage, I.S. Gartshore, A method of reducing drag and fluctuating side force on bluff bodies,
J. Wind Eng. Ind. Aerodyn. 25 (1987) 229–245.
[18] I.S. Gartshore, Some effects of upstream turbulence on the unsteady lift forces imposed on prismatic
two dimensional bodies, ASME J. Fluids Eng. 106 (1984) 418–424.
[19] H. Sakamoto, H. Haniu, Y. Obata, Fluctuating forces acting on two square prisms in a tandem
arrangement, J. Wind Eng. Ind. Aerodyn. 26 (1987) 85–103.

Potrebbero piacerti anche