Journal of Modern Physics, 2013, 4, 89-95
doi:10.4236/jmp.2013.45B015 Published Online May 2013 (http://www.scirp.org/journal/jmp)
On the Sub-Critical Bifurcation of Anti-Phase and
In-Phase Synchronized Vortex Shedding Forms
Yih Ferng Peng
Department of Civil Engineering, National Chi Nan University, Puli 545, Taiwan, China.
Email: yfpeng@ncnu.edu.tw
Received 2013
ABSTRACT
Transition of flows past a pair of side-by-side circular cylinders are investigated by numerical simulations and the bi-
furcation analysis of the numerical results. Various flow patterns behind the cylinder-pair have been identified by the
gap ratio (G) and Reynolds number (Re). This study focus on transition of in-phase and anti-phase vortex shedding
synchronized forms. A nested Cartesian-grid formulation, in combination with an effective immersed boundary method
and a two-step fractional-step procedure, has been adopted to simulate the flows. Numerical results reveal that the
in-phase and anti-phase vortex shedding flows at Re = 100 can co-exist at 2.08 2.58G
. Hysteresis loop with in-
creasing/decreasing G at constant Reynolds number Re = 100 is reported.
Keywords: In-phase Vortex Shedding; Anti-phase Vortex Shedding; Hysteresis
1. Introduction
Because of its fundamental importance and engineering
significance, unstable flow interferences across bluff
bodies have been investigated extensively. Flow inter-
ference with a pair of cylinders is complex in which both
relative-gap between the cylinders and arrangements of
relative position (in tandem, side-by-side, or in staggered)
play crucial roles in the physical transition. Among flows
around a pair of cylinders in various configurations, those
behind circular cylinders in a side-by-side arrangement
have been mostly extensively studied [1-11]. It is now
well established that the flow patterns behind a pair of
side-by-side cylinders can be classified [3,12] by the gap-
ratio G (surface-to-surface distance divided by cylinder-
diameter) and the Reynolds number (Reynolds number,
Re, is defined as/
o
ReU D
, where o
U is the free-
stream velocity, D is the cylinder diameter, and
is the
kinematic viscosity). At very small Gap ratios, the two
cylinders may behave in a similar fashion to a single bluff-
body [1]. At intermediate Gap ratios, the two side-by-
side cylinders is known [2] to exhibit a deflected or bi-
ased flow patterns, which are bi-stable in nature. The
deflected flow pattern is characterized by a gap flow bi-
ased towards one of the two cylinders. The gap flow in
this regime switches spontaneously from one side to the
other and thus corresponds to the flip-flopped regime [5].
For weak coupling (with relatively large G, 1
G
5)
many of the past researchers report occurrences of both
anti-phase and in-phase synchronized vortex shedding
behind a pair of circular cylinders. Findings of William-
son [4] confirm that the shedding vortices (for 40
Re
160) remain synchronized either in phase or in anti-
phase. Anti-phase streets often preserve the phase-locked
identity of the vortices even at a far downstream location.
However, for strong coupling (with relatively lower G)
only in-phase vortex shedding has been reported, and the
associated physical process eventually leads to formation
of a complex asymmetrically evolving Benard von Kar-
man streets.
Although there are varieties of flow patterns behind a
pair of side-by-side cylinders have been identified (which
include semi-single and twin vortex street formations,
symmetric and deflected flows, stationary, biased and
flip-flopped-type vortex shedding, periodic, quasi-peri-
odic, and weakly-chaotic flows, and in-phase and anti-
phase vortex synchronizations), very restriction results
have been reported on transition of those various flow
patterns. Examples can be found in the studies of Mizu-
shima & Ino [13] and Peng et al. [14]. In [13], parameter
space of the gap ratio and the critical Reynolds number
of symmetry/deflected vortex shedding flows and in-
phase/anti-phase vortex shedding flows behind a pair of
circular cylinders at low Reynolds number have been
reported by results of numerical simulations and linear
stability analyses. While in the study [14], transition from
semi-single symmetry vortex shedding flow to semi-
single deflected vortex shedding flow as well as transi-
tion from twin symmetry vortex shedding to twin de-
flected vortex shedding and then transition to the flipped-
Copyright © 2013 SciRes. JMP
Y. F. PENG
90
flopped vortex shedding regime behind a pair of elliptical
cylinders at low Reynolds number are shown by results
of direct numerical simulations. Although there are a lot
of experimental and numerical studies on the side-by-
side cylinders, the important transition phenomenon from
in-phase to anti-phase vortex shedding synchronization
and vs. are still unclear. Purpose of this study is then fo-
cus on transition of in-phase/anti-phase vortex shedding
synchronized forms behind a pair of side-by-side circular
cylinders. High resolution numerical methods based on a
nested Cartesian grid formulation, in combination with
an effective immersed boundary method and a two-step
fractional-step procedure, have been adopted to simulate
flows past a pair of side-by-side circular cylinders. Hys-
teresis loop with increasing/decreasing Re at constant gap
ratio describing the hysteresis phenomenon of in-phase/
anti-phase vortex shedding synchronized forms are re-
ported.
2. Numerical Methods and Validations
In this study, transitions of in-phase and anti-phase vortex
shedding flows behind a pair of circular cylinders in the
side-by-side arrangement are numerically investigated. A
nested Cartesian-grid formulation, in combination with
an effective immersed boundary method and a two-step
fractional-step procedure, has been adopted to simulate
the flows. Extensive related details of the discretization
schemes consisting of inside fine/coarse grid-areas and
the associated immersed boundary method may be found
in Peng et al. [14-15].
Here we briefly describe the implemented numerical
method. Governing equations used are unsteady incom-
pressible Navier–Stokes equations in primitive variables.
In integral forms, the dimensionless governing equations
(with lengths normalized by the diameter D of cylinders,
velocities normalized by the uniform inflow velocity U0,
and time by D /U0) appear as the following.
The mass conservation equation
d0
CS un S
 , (1)
and the momentum conservation equation
d()d
1
dd
CV CS
CS CS
uVuun S
t
pn Sun S
Re

 




Ni
n inner fine-gri
(2)
CS and CV in Eqs. (1) and (2) denote the control-sur-
face and control-volume, respectively, and is a unit
vector normal to the control-surface. While advancing in
time, a second order accurate two-step fractional-step
method is used. A second-order Adams-Bashforth scheme
is employed for discretizing the convective terms, and
diffusion terms are discretized using an implicit Crank
n
colson scheme.
In the iterations of the discretized equations, a local
grid refinement technique is adopted through the intro-
duction of two nested blocks in the computational domain.
The implemented nested-block finite-volume based Carte-
sian-grid method is noted to facilitate effective/accurate
simulation of the presently investigated unsteady viscous
incompressible flows past multiple immersed boundaries.
The procedure adopted here allows systematic simulation
of flows past the cylinder-pair, and preserves global sec-
ond-order accuracy [15]. A sketch for the computational
domain with implemented boundary conditions and a
side-by-side arrangement of cylinder-pair is provided in
Figure 1. Various domain-lengths used for simulations
under the present study are defined as: L1 = 5D, L2 = 50D,
L3 = 12D, L4 = 4D, L5 = 8D, and L6 = 5.5D, facilitating
generation of a physical domain of size 55D × 24D, con-
sisting ad (Grid 2) area 12D × 11D. Upon
taking 0.1
x
yD
the outer Grid1 (the coarse
grid), and 0.05
  for
x
yD
  for the inner Grid 2, the
total grid size became 168,000, with Grid 1 = 120,000,
an
ted body force in the Na-
vier-Stokes equation, i.e.,
d Grid 2 = 48,000.
The simple concept of immersed boundary (IB) method
adopted here helps to simulate effectively the wake evo-
lutions past the cylinders. The virtual presence of the
cylinders within the flow domain is facilitated by intro-
ducing a locally active distribu
1
()
uuup uf,
tRe
 

in which the distributed body force
f
is defined as
(( )1/)fuupReu


, and
ϕ
represents the
volume-fraction of the solid body within a computational
cell. For a cell entirely occupied by the cylinders,
ϕ
= 1
is used; and for a cell fully occupied by fluid,
ϕ
= 0 is
taken. However, for an interface-cell, partially occupied
by a cylinder and partially by fluid, 0 <
ϕ
< 1 is devoted.
Thereafter, the governing equations are solved everywhere
in the computational domain, including cells which are
ccupied by the elliptic cylinders.
o
L
6
L
D
L
5
L
2
L
1
L
3
Grid
1
Grid
2

u/ y=v=0
u/ y=v=0


u=1,
v=0
u/x = 0
v/x = 0

L
3
L
6
L
4
Figure 1. Schematic plot of the flow domain.
Copyright © 2013 SciRes. JMP
Y. F. PENG 91
It is noted that an extensive validation of the underly-
ing method has been well-documented in [14-15]. In [14],
computations of important critical Reynolds numbers
(Recr,v) that correspond to onsets of vortex shedding for
uniform flows past a circular cylinder, an elliptical cyl-
inder, and two side-by-side attached elliptical cylinders
were performed. As listed on Table 1, for a circular cyl-
inder, our previous study observed this Recr,v to be 47.2,
which compares quite well with the experimentally pre-
dicted values 46.9 - 47.9 [16-18], and the theoretical val-
ues 46.1 - 47.3, as obtained by those linear stability
analysis [19-20] and the bifurcation analysis of Dusek et
al. [21].
3.1. Overview of Flows behind a Pair of
are
3. Results
Side-by-Side Circular Cylinders
This study begins with the investigation of critical transi-
tion characteristics in the narrow gap range, and extracts
the underlying bifurcation patterns. For this, first, flow
properties past two side-by-side circular cylinders in the
gap-ratio range 0.2 G 3.0 and 40 Re 100 are ex-
tensively simulated. The observed distinctive physical
properties of these flows sequentially characterized in
Table 2, in which ,1.5Cx
V denotes the teraged
transverse velocity at (x, y) = (1.5, 0.0), ,1 2L
C
ime-av
is the
temporal combined lift-coefficient (subscripe- t 1 + 2 d
mu-
ted data, the readers may note that at all Reynolds.
f vortex shed-
ding for flow pass a circular cylinder.
Source
notes upper plus lower cylinder effects, i.e., ,12,1,2LLL
CCC
),
and T is the period of vortex shedding. “Flow types” (last
column, Table 2) are classified based on the following
three special characteristics. The first letter in the abbre-
viated flow-type, “S,” corresponds to the semi-single flow,
and “T” corresponds to twin flow. For the second letter,
S” represents symmetric flow, and “D” denotes the de-
flected flow. In the third and fourth places, “SS” indicates
steady-state flow, “VP” represents periodic vortex shed-
ding flow, “VQ” stands for quasi- periodic vortex shed-
ding flow, and “VC” denotes the chaotic vortex sh edd ing.
Various vortex shedding regimes, including flip-flopped,
in-phase, and anti-phase vortex shedding flows are also
denoted by superscripts in last column. From the si
la
Table 1. Critical Reynolds number of onset o
Recr,v Analytic Method
Present 47.2 bifis urcation analys
Provansal et al. [16]
7]
47.45)
0] li
47.0 experiments
Williamson [147.9 experiments
Norberg [18] ( ± 0. experiments
Jackson [19] 46.2 linear stability analysis
Kumar & Mittal [247.3 near stability analysis
Dusek et al. [21] 46.1 bifurcation analysis
Table 2. Simulated results of flows past two side-by-side
circular cylinders at 40 Re 100 and 0.2 G 3.0.
No. G Re,1.5Cx
V F
,1 2L
C T low type
1 0.2400.0 0.593 10.80 S,S,VP
2 0.2600.0 0.921 9.75 S,S,VP
3 0.2800.0 0.786 9.20 S,S,VP
4 0.21000.0 0.258 9.08 S,S,VP
5 0.4400.0 0.005 13.61 S,S,VP
6 0.4600
NA S
T
NA T,
NA T,
NA T,
0 5.23
0 5.19
0 5.13
0 5.15
0
5.00 T
0
5.04 T
0
5.10 T
0
5.1 T
60 3.01000.0 0.0 5.17 T,S,VP A
.1110.00112.76 S,D,VP
7 0.4800.0 NA NA S,S,VQ F
8 0.41000.0 NA ,S,VC F
9 0.6400.0 0.0 T,S,SS
10 0.6600.0 0.0 T,S,SS
11 0.6800.0 NA T, S, VC F
12 0.61000.0 NA ,S,VC F
13 0.8400.0 0.0 T,S,SS
14 0.8600.0 NA NA T,S,VQ F
15 0.8800.0 NA NA T,S,VC F
16 0.81000.0 NA S,VC F
17 1.0400.0 0.0 T,S,SS
18 1.0600.0 NA NA T,S,VQ F
19 1.0800.0 NA NA T,S ,VQ F
20 1.01000.0 NA S,VC F
21 1.2400.0 0.0 T,S,SS
22 1.2600.0 NA NA T,S,VQ F
23 1.2800.0 NA NA T,S,VQ F
24 1.21000.0 NA S,VQ F
25 1.4400.0 0.0 T,S,SS
26 1.4600.0 NA NA T,S,VQ F
27 1.4800.0 0.439 5.53 T,S,VP I
28 1.41000.0 .622T,S,VP I
29 1.6400.0 0.0 T,S,SS
30 1.6600.0 NA NA T,S,VQ F
31 1.6800.0 0.415 5.50 T,S,VP I
32 1.61000.0 .588T,S,VP I
33 1.8400.0 0.0 T,S,SS
34 1.8600.0 NA NA T,S,VQ F
35 1.8800.0 0.402 5.45 T,S,VP I
36 1.81000.0 .570T,S,VP I
37 2.0400.0 0.0 T,S,SS
38 2.0600.0 0.197 6.05 T,S,VP I
39 2.0800.0 0.398 5.47 T,S,VP I
40 2.01000.0 .564T,S,VP I
41 2.2400.0 0.0 T,S,SS
42 2.2600.0 0.192 6.04 T,S,VP I
43 2.2800.0 .3995.50 T,S,VP I
44 2.21000.0 0.0 ,S,VP A
45 2.4400.0 0.0 T,S,SS
46 2.4600.0 0.196 6.06 T,S,V P I
47 2.4800.0 .4065.56 T,S,VP A
48 2.41000.0 0.0 ,S,VP A
49 2.6400.0 0.0 T,S,SS
50 2.6600.0 .2006.11 T,S ,V P I
51 2.6800.0 0.0 5.37 T,S,VP A
52 2.61000.0 0.0 ,S,VP A
53 2.8400.0 0.0 T,S,SS
54 2.8600.0 .2086.08 T,S,VP I
55 2.8800.0 0.0 5.41 T,S,VP A
56 2.81000.0 0.0 1 ,S,VP A
57 3.0400.0 0.0 T,S,SS
58 3.0600.0 0.0 5.94 T,S,VP A
59 3.0800.0 0.0 5.49 T,S,VP A
Copyright © 2013 SciRes. JMP
Y. F. PENG
92
numbers (Re 40) the single vortex shedding street is
reached in the range G 0.4. However, for larger G (G
0.6), the approach to the vortex shedding flow remained
dependent on Re. The flip-flopped vortex-shed-ding oc-
curred in the gap-ratio range 0.4 G 1.8 with Re 60.
In-phase vortex shedding is detected at 0.4 G 1.8, and
anti-phase vortex shedding is founded at large G (G
2.2). Notably, while Table 2 presents the explicit physi-
cal details of various flows, for the sake of facilitating
immediate comprehension of a trend, the parameter
space diagram is suitably summarized in Figure 2. As
shown in Figure 2, it is clear that flows are semi-single
at G < 0.5. At intermediate Gap ratios the two
side-by-side cylinders is known to exhibit a deflected or
biased flow patterns, which are bi-stable in nature. The
deflected jets through circular gap are further affected by
the shedding vortices at higher Re and flows become
flip-flopped consequently. The range of Gap ratios where
the flip-flopped flow pattern is observed extends from
approximately Gap ratios between 0.4 - 1.8 depending on
the Reynolds number. At higher Gap ratios, i.e., the cyl-
inders are spaced sufficiently far apart, the pair of cylin-
ders may behave as two independent bluff bodies. Prox-
imity interference effects, however, lead to various
modes of synchronization, anti-phase and in-phase, in the
vortex formation and shedding processes and the result-
ing parallel vortex streets. For example, in-phase and
anti-phase vortex shedding flows at Re = 100 are clearly
revealed at G = 2.0 and G = 3.0, respectively.
3.2. Hysteresis Scenario of In-phase and
Anti-phase Synchronized Forms
To investigated the transition of in-phase and anti-phase
synchronized forms in flows past a pair of side-by-side
circular cylinders, two sets of computations including
anti-phase and in-phase branches are carried out. The
anti-phase branch started from the anti-phase vortex
20 30 40 50 60 70 80 90100110120
Re
0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
1.6
1.8
2.0
2.2
2.4
2.6
2.8
3.0
3.2
3.4
3.6
G
SSSS
SSSSSSSSS
SSS
S
S
S
SS
S
S
S
S
S
S
S
DDDDDD
D
FFF
FFFF FF
FFF
FFF
FFF
FFFF
F
F
 




A
AA
AA
AA
AAA
A
A
A
A
A
A
A
A
A
A
Twin street
Semi-single street
S: symmetric flowsD: deflected flowsF: flip-flopped flows
I: in-phase flowsA: anti-phase flows
: steady flows
: periodic flows
: quasi-periodic flows
: chaotic flows
Figure 2. Simulated different wake patterns observed be-
hind two side-by-side circular cylinders.
shedding flow at G = 3.0 and Re = 100 and are calculated
by progressively decreasing G in very small steps. In the
meanwhile, the in-phase branch started from the in-phase
vortex shedding flow at G = 2.0 and Re = 100 and are
calculated by progressively increasing G in small steps.
The observed distinctive physical properties of these
flows are sequentially characterized in Table 3. It is
noted that in anti-phase branch, the solution at a higher G
is used as the initial condition for the next lower G.
Similarly, in in-phase branch, the solution at a lower G is
used as the initial condition for the next higher G.
The flow past cylinder-pair at constant Reynolds num-
ber (Re = 100) retains anti-phase vortex shedding syn-
chronized forms at high gap ratio (), as
indicated by our extracted data on Table 3(a). Once G is
decreased to G = 2.06, the anti-phase vortex shedding
flow transits to the in-phase synchronized vortex shed-
ding form. For clarity, the existence of the anti-phase
vortex shedding flow pattern at Re = 100, and G = 2.32 is
exhibited in Figure 3. It reveals clearly the anti-phase
vortex shedding behavior of the simulated flow in the
sub-domain x = [5, 50], y = [9, 9]. The continuation of
zero central-line velocity (Figure 3(d)), and perfectly
anti-phase synchronized growth of the lift (CL,1, CL,2)
coefficients (Figures 3(b) and (c)) ensure the inherent
anti-phase characteristic of vortex shedding (Figure 3(a))
in the wake. Note that, unlike for in-phase vortex shed-
ding flow (Figure 4), the equal but opposite natured
variations of CL,1 and CL,2 in the present case contribute
to the continued vanishing of CL,1+2 (Figure 3(d)) during
the entire time-evolution.
2.08 3.0G
Upon maintaining anti-phase synchronized vortex shed-
ding forms with the past findings related to side-by-side
cylinder-pair (Re = 100, ), once the gap-
ratio was subsequently increased (from
2.08 3.0G
2.0G
), the
in-phase synchronized vortex shedding were encountered
behind the cylinder-pair (Re = 100, ). The
distinguishable physical characteristics associated with
the in-phase vortex shedding flows past the cylinder-pair
at Re = 100 and again G = 2.32 is extracted in Figure 4.
It can be noted from the figure that the gap-flow quickly
lost stability; however, the shedding vortices appeared
clearly in-phase synchronized at least up to x 12.
Thereafter, instability is seen to quickly grow, leading to
the development of a combined binary vortex street
within 12 < x <26, and beyond that there occurred an
irregular flow pattern over a conversion point. Physical
details of this in-phase flow in terms of enhanced CL,1+2,
and significantly modulated transient evolution (in-phase)
of individual lift (CL,1 and CL,2) coefficients are extracted
in Figures 4(d), (b) and (c).
2.0 2.58G
3.3. Hysteresis Loop
From the simulated data on Table 3, the reader may note
that at constant Reynolds number (Re = 100), the anti-
Copyright © 2013 SciRes. JMP
Y. F. PENG 93
phase branch ranges between , and the
in-phase branch ranges between . In the
other word, the anti-phase and in-phase synchronized forms
behind a pair of circular cylinders (Re = 100) can
co-exist at . The co-existences of the
anti-phase and in-phase vortex shedding flow patterns (at
Re = 100 and G = 2.32) have been shown in Figures 3
and 4, respectively. Figures 3(a) and 4(a) reveal the
anti-phase and in-phase synchronized vortex shedding
forms, respectively, by iso-vorticity plots of the flows. The
persistence of symmetric flow nature (having zero time
mean central-line transverse velocity and lift-coefficient)
in both anti-phase and in-phase vortex shedding flows
are clearly reflected by the time histories of ,1 2L
2.08 3.00G
2.0 2.58G
2.08 2.58G
C
(Figures 3(d) and 4(d)). However, an enlarged view of
individual lift- coefficients (, ) for the upper
and the lower
,1L
C,2L
C
Table 3. (a). Numerical results of flow past a pair of side-
by-side circular cylinders by decreasing G slowly (Re = 100).
* denotes the computed case where transition is happen; (b)
Numerical results of flow past a pair of side-by-side circular
cylinders by increasing G slowly (Re = 100). * denotes the
computed case where transition is happen.
(a)
No. G Re ,1.5Cx
V ,12L
C T Flow type
1 3.00 100 0.0 0.0 5.54 T,S,VP A
2 2.60 100 0.0 0.0 5.13 T,S,VP A
3 2.56 100 0.0 0.0 5.13 T,S,VP A
4 2.54 100 0.0 0.0 5.13 T,S,VP A
5 2.52 100 0.0 0.0 5.13 T,S,VP A
6 2.50 100 0.0 0.0 5.12 T,S,VP A
7 2.48 100 0.0 0.0 5.12 T,S,VP A
8 2.46 100 0.0 0.0 5.11 T,S,VP A
9 2.44 100 0.0 0.0 5.11 T,S,VP A
10 2.42 100 0.0 0.0 5.09 T,S,VP A
11 2.40 100 0.0 0.0 5.06 T,S,VP A
12 2.38 100 0.0 0.0 5.09 T,S,VP A
13 2.36 100 0.0 0.0 5.09 T,S,VP A
14 2.34 100 0.0 0.0 5.10 T,S,VP A
15 2.32 100 0.0 0.0 5.08 T,S,VP A
16 2.30 100 0.0 0.0 5.04 T,S,VP A
17 2.28 100 0.0 0.0 5.08 T,S,VP A
18 2.26 100 0.0 0.0 5.06 T,S,VP A
19 2.24 100 0.0 0.0 5.09 T,S,VP A
20 2.22 100 0.0 0.0 5.04 T,S,VP A
21 2.20 100 0.0 0.0 5.03 T,S,VP A
22 2.18 100 0.0 0.0 5.05 T,S,VP A
23 2.16 100 0.0 0.0 5.04 T,S,VP A
24 2.14 100 0.0 0.0 5.08 T,S,VP A
25 2.12 100 0.0 0.0 5.04 T,S,VP A
26 2.10 100 0.0 0.0 5.02 T,S,VP A
27 2.08 100 0.0 0.003 5.00 T,S,VP A
28* 2.06 100 0.0 0.659 5.21 T,S,VP I
29 2.04 100 0.0 0.665 5.21 T,S,VP I
(b)
No. G Re ,1.5Cx
V ,12L
C T Flow type
30 2.02100 0.0 0.665 5.18 T,S,VP I
31 2.04100 0.0 0.665 5.21 T,S,VP I
32 2.06100 0.0 0.659 5.21 T,S,VP I
33 2.08100 0.0 0.651 5.19 T,S,VP I
34 2.10100 0.0 0.563 5.15 T,S,VP I
35 2.12100 0.0 0.664 5.20 T,S,VP I
36 2.14100 0.0 0.664 5.22 T,S,VP I
37 2.16100 0.0 0.659 5.21 T,S,VP I
38 2.18100 0.0 0.651 5.19 T,S,VP I
39 2.20100 0.0 0.564 5.15 T,S,VP I
40 2.22100 0.0 0.664 5.20 T,S,VP I
41 2.24100 0.0 0.665 5.19 T,S,VP I
42 2.26100 0.0 0.660 5.20 T,S,VP I
43 2.28100 0.0 0.653 5.20 T,S,VP I
44 2.30100 0.0 0.566 5.15 T,S,VP I
45 2.32100 0.0 0.665 5.20 T,S,VP I
46 2.34100 0.0 0.666 5.20 T,S,VP I
47 2.36100 0.0 0.662 5.20 T,S,VP I
48 2.38100 0.0 0.655 5.20 T,S,VP I
49 2.40100 0.0 0.569 5.16 T,S,VP I
50 2.42100 0.0 0.667 5.20 T,S,VP I
51 2.44100 0.0 0.669 5.20 T,S,VP I
52 2.46100 0.0 0.665 5.20 T,S,VP I
53 2.48100 0.0 0.658 5.20 T,S,VP I
54 2.50100 0.0 0.572 5.13 T,S,VP I
55 2.52100 0.0 0.670 5.20 T,S,VP I
56 2.54100 0.0 0.672 5.20 T,S,VP I
57 2.56100 0.0 0.668 5.20 T,S,VP I
58 2.58100 0.0 0.662 5.20 T,S,VP I
59* 2.60100 0.0 0.0 5.13 T,S,VP A
0 10203040
x
-9
-6
-3
0
3
6
9
y
G = 2.32, Re = 100
(a)
-0.40
0.00
0.40
0.80
CL,1
(b)
-0.80
-0.40
0.00
0.40
CL,2
(c)
7005 7020 7035 7050
t
-0.15
-0.10
-0.05
0.00
0.05
0.10
0.15
CL,1+2
(d)
Figure 3. The simulated anti-phase flow at G = 2.32 and Re
= 100. (a) Observed anti-phase flow pattern. Time-histories
of: (b) CL,1(t); and (c) CL,2(t). (d) Vanishing of the combined
lift- coefficient CL,1+2.
Copyright © 2013 SciRes. JMP
Y. F. PENG
94
0 10203040
x
-9
-6
-3
0
3
6
9
y
G = 2.32, Re = 100
(a)
11440 11456 11472 11488
t
-0.80
-0.40
0.00
0.40
0.80
C
L,1+2
(d)
-0.40
-0.20
0.00
0.20
0.40
0.60
C
L,1
(b)
-0.60
-0.40
-0.20
0.00
0.20
0.40
C
L,2
(c)
Figure 4. The simulated periodic in-phase flow at G = 2.32
and Re = 100. (a) In-phase vortex shedding behavior.
Time-histories of: (b) CL,1(t); and (c) CL,2(t). (d) Enhance-
ment of the combined lift-coefficient CL,1+2.
1.8 2.0 2.2 2.4 2.6 2.8 3.0
G
-0.2
0.0
0.2
0.4
0.6
0.8
1.0
1.2
C
L,1+2
: byincreasingG gradually
: by decreasing G gradually
the in-phase branch
the anti-phase branch
Figure 5. The hysteresis region with G of in-phase and anti-
hase synchronized vortex shedding forms for flows past a
pair of circular cylinders at Re = 100.
cylinders (Figures 3(b) and (c)/Figures 4(b) and (c))
reveals occurrence of anti-phase/in-phase vortex shed-
ding vortices of the upper and lower vortex streets.
Readers may note that the corresponding lift-amplitudes
(,12 ) in anti-phase vortex shedding flows become
neutralized as shown in Figure 3(d), on the contrary, the
corresponding lift-amplitudes () in in-phase vortex
shedding flows become enhanced as shown in Figure
4(d), respectively. Since the value ofstands as a
characteristic value of anti-phase/in-phase synchronized
vortex shedding forms, it is worthy to show the
L
C
,1 2L
C
,1 2L
C
,1 2L
C
distributions. As shown in Figure 5, the ,1 2L
C
distri-
bution along the anti-phase branch (,1 2
C) combines
the in-phase branch (
0
L
,1 2L
C0
) becomes a hysteresis
loop. Particularly, the anti-phase branch starting from G
= 3.0 trace along a straight segment, ending at G = 2.08,
and then merges to the in-phase branch. While the in-
phase branch starting from G = 2.0 trace along a wavy
line, ending at G = 2.58, and then merges to the
anti-phase branch.
4. Conclusions
Numerical results have been presented for the in-phase
and anti-phase vortex shedding synchronized forms of
flows behind a pair of side-by-side circular cylinders.
Flows are restricted in low-Reynolds-number (Re
100)
laminar regime for various small/middle gap ratio
(0.23.0G
). The computations have been carried out
in two-dimensional, using a high resolution numerical
method based on a nested Cartesian grid formulation, in
combination with an effective immersed boundary method
and a two-step fractional-step procedure.
Hysteresis phenomenon of the in-phase/anti-phase
vortex shedding synchronized forms of flows has been
studied in detail. For flows behind a pair of side-by-side
circular cylinders at Re = 100, the hysteresis loop with
width ranges between is clearly found.
2.08 2.58G
5. Acknowledgements
This work was supported in part by the National Science
Council of the Republic of China (Taiwan) under Con-
tract No. NSC 101-2221-E-260-038-.
REFERENCES
[1] H. M. Spivac, “Vortex Frequency and Flow Pattern in the
Wake of Two Parallel Cylinders at Varied Spacing Nor-
mal to An Airstream,” Journal of the Aeronautical Sci-
ences, Vol. 13, 1946, pp. 289-301.
[2] P. W. Bearman and A. J. Wadcock, “The Interference
between A Pair of Circular Cylinders Normal to A
Stream,” Journal of Fluid Mechanics, Vol. 61. No. 3,
1973, pp. 499-511. doi:10.1017/S0022112073000832
[3] M. M. Zdravkovich, “Review of Flow Interference Be-
tween Two Circular Cylinders in Various Arrangements,”
ASME, Vol. 99, No. 4, 1977, pp. 618-633.
doi:10.1115/1.3448871
[4] C. H. K. Williamson, “Evolution of a Single Wake behind
A Pair of Bluff Bodies,” Journal of Fluid Mechanics, Vol.
159, 1985, pp. 1-18. doi:10.1017/S002211208500307X
[5] H. J. Kim and P. A. Durbin, “Investigation of the Flow
between a Pair of Circular Cylinders in the Flopping Re-
gime,” Journal of Fluid Mechanics, Vol. 196, 1988, pp.
431-448. doi:10.1017/S0022112088002769
[6] D. Sumner, S. S. T. Wong, S. J. Price and M. P. Päidous-
sis, “Fluid Behavior of Side-by-side Circular Cylinders in
Steady Cross-flow,” Journal of Fluids and Structures,
Vol. 13. No. 3, 1999, pp. 309-339.
doi:10.1006/jfls.1999.0205
Copyright © 2013 SciRes. JMP
Y. F. PENG
Copyright © 2013 SciRes. JMP
95
[7] Y. Zhou, H. J. Zhang and M. W. Yiu, “The Turbulent
Wake of Two Side-by-side Circular Cylinders,” Journal
of Fluid MechanicsVol. 458, 2002, pp. 303-332.
doi:10.1017/S0022112002007887
[8] S. J. Xu, Y. Zhou and R. M. C. So, “Reynolds Number
Effects on the Flow Structure behind Two Side-by-side
Cylinders,” Physics of Fluids, Vol. 15. No. 5, 2003, pp.
1214-1219. doi:10.1063/1.1561614
[9] S. Kang, “Characteristics of Flow over Two Circular
Cylinders in A Side-by-side Arrangement at Low Rey-
nolds Numbers,” Physics of Fluids, Vol. 15, 2003.
[10] S. Kumar, B. Gonzalez and O. Probst, “Flow Past Two
Rotating Cylinders,” Physics of Fluids, Vol. 23, No. 1,
2011, 01402. doi:10.1063/1.3528260
[11] Md Mahbub Alam, Y. Zhou, and X. W. Wang, “The
Wake of Two Side-by-side Square Cylinders,” Journal of
Fluid Mechanics, Vol. 669, 2011, pp. 432-471.
doi:10.1017/S0022112010005288
[12] D. Sumner, “Two Circular Cylinders in Cross-flow: A
Review,” Journal of Fluids and Structures, Vol. 26. No. 6,
2010, pp. 849-899.
doi:10.1016/j.jfluidstructs.2010.07.001
[13] J. Mizushima and Y. Ino, “Stability of Flows Past A Pair
of Circular Cylinders in A Side-by-side Arrangement,”
Journal of Fluid Mechanics, Vol. 595, 2008, pp. 491-507.
doi:10.1017/S0022112007009433
[14] Y. F. Peng, A. Sau, R. R. Hwang, W. C. Yang and C. M.
Hsieh, “Criticality of Flow Transition behind Two
Side-by-side Elliptic Cylinders,” Physics of Fluids, Vol.
24. No. 3, 2012, 034102.
doi:10.1063/1.3687450
[15] Y. F. Peng, R. Mittal, A. Sau and R. Hwang, “Ested Car-
tesian Grid Method in Incompressible Viscous Fluid
Flow,”Journal of Computational Physics, Vol. 229,
No.19, 2010, pp. 7072-7101.
doi:10.1016/j.jcp.2010.05.041
[16] M. Provansal, C. Mathis and L. Boyer, “Benard-von
Karman Instability: Transient and Forced Regimes,”
Journal of Fluid Mechanics, Vol. 182, 1987, pp. 1-22.
doi:10.1017/S0022112087002222
[17] C. H. K. Williamson, “Oblique and Parallel Modes of
Vortex Shedding in the Wake of a Circular Cylinder at
Low Reynolds Numbers,” Journal of Fluid Mechanics,
Vol. 206, 1989, pp. 579-627.
doi:10.1017/S0022112089002429
[18] C. Norberg, “An Experimental Investigation of the Flow
around a Circular Cylinder: Influence of Aspect Ratio,”
Journal of Fluid Mechanics, Vol. 258, 1994, pp. 287-316.
doi:10.1017/S0022112094003332
[19] C. P. Jackson, “A Finite-element Study of the Onset of
Vortex Shedding in Flow past Variously Shaped Bodies,”
J ournal of Fluid Mechanics, Vol. 182, 1987, pp. 23-45.
doi:10.1017/S0022112087002234
[20] B, Kumar and S. Mittal, “Effect of Blockage on Critical
Parameters for Flow past A Circular Cylinder,” Interna-
tional Journal for Numerical Methods in Fluids, Vol. 50.
No. 8, 2006, pp. 987-1001. doi:10.1002/fld.1098
[21] J. Dusek, P. Le Gal and P. Fraunie, “A Numerical and
Theoretical Study of the First Hopf Bifurcation in A Cyl-
inder Wake,” Journal of Fluid Mechanics, Vol. 264, 1994,
pp. 59-80. doi:10.1017/S0022112094000583