Anda di halaman 1dari 29

Annu. Rev. Fluid Mech. 2008.40:113-139. Downloaded from www.annualreviews.

org
Access provided by Indian Institute of Technology - Kanpur on 10/18/16. For personal use only.

ANRV332-FL40-06

ARI

10 November 2007

16:9

Control of Flow Over


a Bluff Body
Haecheon Choi,1,2 Woo-Pyung Jeon,1
and Jinsung Kim1
1

School of Mechanical and Aerospace Engineering and 2 Center for Turbulence and
Flow Control Research, Institute of Advanced Machinery and Design, Seoul National
University, Seoul 151-744, Korea; email: choi@snu.ac.kr, wpjeon@snu.ac.kr,
jkim@lscable.com

Annu. Rev. Fluid Mech. 2008. 40:11339

Key Words

The Annual Review of Fluid Mechanics is online at


uid.annualreviews.org

wake, vortex shedding, separation, drag, lift

This articles doi:


10.1146/annurev.uid.39.050905.110149

Abstract

c 2008 by Annual Reviews.


Copyright 
All rights reserved
0066-4189/08/0115-0113$20.00

In this review, we present control methods for ow over a bluff body


such as a circular cylinder, a 2D bluff body with a blunt trailing edge,
and a sphere. We introduce recent major achievements in bluff-body
ow controls such as 3D forcing, active feedback control, control
based on local and global instability, and control with a synthetic
jet. We then classify the controls as boundary-layer controls and
direct-wake modications and discuss important features associated
with these controls. Finally, we discuss some other issues such as
Reynolds-number dependence, the lowest possible drag by control,
and control efciency.

113

ANRV332-FL40-06

ARI

10 November 2007

16:9

1. INTRODUCTION

Annu. Rev. Fluid Mech. 2008.40:113-139. Downloaded from www.annualreviews.org


Access provided by Indian Institute of Technology - Kanpur on 10/18/16. For personal use only.

Flow over a bluff body is a common occurrence associated with uid owing over an
obstacle or with the movement of a natural or articial body. Evident examples are
the ows past an airplane, a submarine and an automobile, and wind blowing past a
bridge and a high-rise building. At much lower Reynolds numbers, the ow over a
bluff body is highly viscous, and the force exerted on the body is mainly attributed
to skin friction. However, when the Reynolds number exceeds a critical value, vortex
shedding occurs in the wake, resulting in a signicant pressure drop on the rear
surface of the body. This vortex shedding occurs over a wide range of Reynolds
numbers (e.g., see Figures 1 and 2), causing serious structural vibrations, acoustic
noise and resonance, enhanced mixing, and signicant increases in the mean drag
and lift uctuations. Therefore, the effective control of vortex shedding is important
in engineering applications.
Since Roshko (1955) rst measured the period of Karman vortex shedding behind
a bluff body, researchers have been steadily investigating vortex shedding and nearwake ow as reported in the reviews by Berger & Wille (1972), Lin & Pao (1979),
Bearman (1984), Oertel (1990), Grifn & Hall (1991), Coutanceau & Defaye (1991),
and Williamson (1996, 2004). Furthermore, efforts have been made to alter and
suppress vortex shedding. In the past, many passive and active open-loop control
methods were introduced to control vortex shedding behind a 2D bluff body, such
as a circular cylinder and a 2D blunt-based bluff body. Examples include end plates,
splitter plates, geometric modication in the trailing edge, base bleed, oscillation in
line with the incident ow, and rotary oscillation. These control methods are passive
and active open loop in the sense that there is no power input and no feedback sensor,
respectively. [See the reviews by Zdravkovich (1981), Oertel (1990), and Grifn &
Hall (1991) for more information about these research activities.]

Figure 1
Visualizations of vortex shedding (a) behind a circular cylinder at Re = 140 and (b) behind Mt.
Halla, Jeju Island, Korea at Re = O(109 ). Note that vortex shedding is clearly observed at a
high Reynolds number, O(109 ). Figure 1a taken from Van Dyke 1982; Figure 1b taken from
NOAA-11 Satellite on November 24, 1991.

114

Choi


Jeon

Kim

ANRV332-FL40-06

ARI

10 November 2007

16:9

Figure 2
Visualizations of vortex
shedding behind a sphere:
(a) Re = 425; (b) 3700; and
(c) 100,000.

Annu. Rev. Fluid Mech. 2008.40:113-139. Downloaded from www.annualreviews.org


Access provided by Indian Institute of Technology - Kanpur on 10/18/16. For personal use only.

During the past 15 years, passive and active open-loop control methods have been
further and newly developed because they are easier to implement than the active
feedback control method, which requires sensing as well as actuation. Nevertheless,
many attempts have been also made to improve control efciency and effectiveness
using active feedback control methods with the advent of the microelectromechanical
system, the development of control theory, and the fast growth of computer power.
The merit of this approach is that it obtains information of the ow systems response
to the actuation, and it obtains better control performance than the passive or active
open-loop control method. We briey introduce these recent activities in the development of passive- and active-control methods in Section 2, and we describe some
of the new passive and active controls in Section 3.
Although many different shapes of bluff bodies exist, we consider three representative shapes in the present review: the circular cylinder, a 2D bluff body with a blunt
trailing edge, and the sphere. The circular cylinder and sphere are the representative
2D and 3D bluff bodies, respectively, and their separation points change depending on the Reynolds number. Conversely, in the case of the 2D blunt-based bluff
body, the separation is xed at the trailing edge, and hence the ow suddenly changes
at the trailing edge from a boundary-layer ow to a wake, which is quite different
from the case of both the circular cylinder and sphere. Therefore, different control
strategies may have to be developed when the body shapes are different. We compare
and analyze control strategies for bluff bodies with xed and movable separations
in Section 4. Some other important issues such as Reynolds-number dependence,
the lowest possible drag, and control efciency are further discussed in Section 5,
followed by conclusions in Section 6.

www.annualreviews.org Control of Flow Over a Bluff Body

115

ANRV332-FL40-06

ARI

10 November 2007

16:9

2. CLASSIFICATION OF CONTROL METHODS

Annu. Rev. Fluid Mech. 2008.40:113-139. Downloaded from www.annualreviews.org


Access provided by Indian Institute of Technology - Kanpur on 10/18/16. For personal use only.

Much research has been conducted on the control of ow over a bluff body in the past
15 years. We rst classify recent bluff-body ow controls into three groups (passive,
active open-loop, and active closed-loop controls, respectively) for cases of actuator(s)
without power input, actuator(s) with power input but no sensor, and sensor(s) and
actuator(s) with power input (Gad-el-Hak 2000).
The reduction of drag on a bluff body has been achieved by means of various
passive devices, such as surface modications with roughness (Shih et al. 1993), dimple (Bearman & Harvey 1993, Choi et al. 2006a), helical wire (Lee & Kim 1997),
longitudinal groove (Lim & Lee 2002), splitter plate (Anderson & Szewczyk 1997,
Hwang et al. 2003, Kwon & Choi 1996, Ozono 1999), and small secondary control
cylinder (Dalton et al. 2001, Sakamoto & Haniu 1994, Strykowski & Sreenivasan
1990). Among them, geometric modication in the spanwise direction near the separation point, referred to as 3D forcing in this review, has been recognized as an
effective control method for the reduction of drag on a 2D body, and therefore, the
corresponding mechanism has received much attention.
Many active open-loop controls using various forcing devices have been also applied to bluff-body ow, such as rotary (Baek & Sung 1998; Choi et al. 2002; Dennis
et al. 2000; Filler et al. 1991; Poncet 2002, 2004; Shiels & Leonard 2001; Tokumaru &
Dimotakis 1991), streamwise (Cetiner & Rockwell 2001), and transverse (Blackburn
& Henderson 1999; Carberry et al. 2003; Dennis et al. 2000; Nakano & Rockwell
1993, 1994) oscillations of a bluff body; inow oscillation (Konstantinidis et al. 2005,
Nehari et al. 2004); electromagnetic forcing (Artana et al. 2003, Kim & Lee 2000,
Mutschke et al. 2001); steady blowing/suction (Arcas & Redekopp 2004, Delaunay &
Kaiktsis 2001, Leu & Ho 2000, Lin et al. 1995, Sevilla & Martinez-Bazan 2004, Yao
& Sandham 2002); time-periodic blowing/suction (Fujisawa et al. 2004, Jeon et al.
2004, Lin et al. 1995, Williams et al. 1992); distributed forcing (Kim & Choi 2005,
Kim et al. 2004); and a synthetic jet (Amitay et al. 1998). In particular, many experimental and numerical studies on the rotary oscillation in a circular-cylinder ow
have been conducted at various Reynolds numbers to nd its inuence on the ow
and the mechanism of drag reduction since Tokumaru & Dimotakis (1991) experimentally obtained approximately 80% drag reduction at Re = 15,000. However, the
efciency of this control method may not be high enough to implement it into real
situations.
Researchers have proposed that optimal feedback control algorithms based on
mathematical analysis, such as optimal control theory, effectively control strong nonlinear ow (He et al. 2000, Homescu et al. 2002, Li et al. 2003, Min & Choi 1999,
Protas & Styczek 2002) and have applied them to ow past a bluff body for drag reduction. Further simplied feedback controls have been also suggested using reducedorder models (ROMs) (Bergmann et al. 2005, Cortelezzi 1996, Cortelezzi et al. 1997,
Gillies 1998, Graham et al. 1999, Li & Aubry 2003, Protas 2004).
Among these recent passive- and active-control methods, we select four important
ones (3D forcing for 2D bluff bodies, feedback control, control based on local/global
instability, and control using a synthetic jet) and describe them in detail in Section 3.

116

Choi


Jeon

Kim

ANRV332-FL40-06

ARI

10 November 2007

16:9

Annu. Rev. Fluid Mech. 2008.40:113-139. Downloaded from www.annualreviews.org


Access provided by Indian Institute of Technology - Kanpur on 10/18/16. For personal use only.

Passive
control

Active
open-loop
control

Active
closed-loop
control

Twodimensional
forcing

Threedimensional
forcing

Boundarylayer control

Directwake control

Figure 3
Classication of control
methods for ow over a
bluff body.

Besides the classical categorization of control methods, we classify control methods for bluff-body ows into two different categories (Figure 3): (a) 2D versus 3D
forcing and (b) boundary-layer control versus direct-wake control. Here, 3D forcing is the control method in which the actuation property varies along the spanwise
(or azimuthal) direction, whereas it does not vary in 2D forcing. Section 3.1 describes examples of this 3D forcing. The classication of control methods as either
boundary-layer or direct-wake control depends on whether control delays separation
by modifying the boundary layer or changes directly the wake characteristics (see
Section 4 for further detail on this classication).

3. MAIN ACHIEVEMENTS FROM RECENT


FLOW-CONTROL STUDIES
In this section, we present recent successful control methods developed for ow over
a bluff body. They are 3D forcing, feedback controls based on optimal control and
ROMs, controls based on local/global instability, and a synthetic jet.

3.1. 3D Forcing for 2D Bluff Bodies


As effective control methods, passive and active open-loop controls varying along
the spanwise (or azimuthal) direction, called 3D forcing, have been applied near the
separation point to control nominally 2D wake.
Passive control. One example of a 3D geometric modication is a helical strake
(Figure 4a) (Zdravkovich 1981), which reduces force uctuations. Tanner (1972),
Rodriguez (1991), and Petrusma & Gai (1994) studied the ow behind a blunttrailing-edge body and reported signicant drag reduction with a segmented trailing
edge (Figure 4b). Tombazis & Bearman (1997) installed a wavy trailing edge on
a blunt-based model (Figure 4c) and found that the waviness of the trailing edge
produces vortex dislocation in the wake and thus increases the base pressure. Bearman
& Owen (1998) introduced a spanwise waviness to the front stagnation face of a

www.annualreviews.org Control of Flow Over a Bluff Body

117

ANRV332-FL40-06

ARI

10 November 2007

16:9

Annu. Rev. Fluid Mech. 2008.40:113-139. Downloaded from www.annualreviews.org


Access provided by Indian Institute of Technology - Kanpur on 10/18/16. For personal use only.

lz ly

Figure 4
3D forcing by passive means: (a) helical strake, (b) segmented trailing edge, (c) wavy trailing
edge, (d ) wavy stagnation face, (e) sinusoidal axis, ( f ) hemispherical bump, and
( g, h) small-size tab.

rectangular cylinderthe rear face remained at (Figure 4d )which led to the


suppression of vortex shedding due to the spanwise waviness. They also found that
the installation of the spanwise waviness on the rear surface does not reduce drag for
the rectangular cylinder because separation occurs at the front edge. Owen et al. (2000,
2001) extended this concept to a circular cylinder by changing the straight axis of the
cylinder into a sinusoidal axis along the spanwise direction (Figure 4e) and by spirally
attaching hemispherical bumps on the cylinder surface (Figure 4f ), producing an
effect similar to that observed for the ow over a rectangular cylinder. Darekar &
Sherwin (2001) performed numerical simulations to investigate the ow past a square
cylinder with a wavy stagnation face at low Reynolds numbers and showed that the
Karman vortex shedding is suppressed into a steady and symmetric structure by the
spanwise waviness (see Supplemental Figure 1; follow the Supplemental Material
link from the Annual Reviews home page at http://www.annualreviews.org). They
argued that vortex-shedding suppression is caused by the distortion of 2D shear layers
due to geometric disturbances because the three-dimensionally redistributed shear
layers become less susceptible to rolling up into a Karman vortex street. They also
found that the optimal wavelength of spanwise waviness is close to the spanwise
wavelength of mode A instability.
Park et al. (2006) proposed a small-size tab, mounted on part of the upper and lower
trailing edges of a bluff body (Figure 4g), to effectively attenuate vortex shedding
and reduce drag. In their parametric study, they varied the height (ly ) and width (lz )
of the tab and the spanwise spacing between the adjacent tabs (), and showed that
drag is decreased (or the base pressure is increased) by attaching this simple device
to the trailing edge (see Supplemental Figure 2). The optimal conguration of tabs

118

Choi


Jeon

Kim

ANRV332-FL40-06

ARI

10 November 2007

16:9

Figure 5
Instantaneous vortical
structures in a wake
(Re = u h/ = 4200):
(a) uncontrolled ow and
(b) controlled ow with tabs
of (/ h, l y / h, l z / h) =
(2, 0.2, 0.2). Shown are the
3D views of vortical
structures (left column) and
top views of isopressure
surfaces (right column).
Figure taken from Park
et al. 2006.

Annu. Rev. Fluid Mech. 2008.40:113-139. Downloaded from www.annualreviews.org


Access provided by Indian Institute of Technology - Kanpur on 10/18/16. For personal use only.

x=0

produced an approximately 33% increase in the base pressure. Owing to the tabs,
the vortices shed from the upper and lower trailing edges lost their 2D nature, and
vortex dislocation occurred. The Karman vortex shedding completely disappeared
right behind the bluff body but occurred weakly at farther downstream locations
(Figure 5). Because the main mechanism of drag reduction by the tab is to introduce
the spanwise phase mismatch in the vortex-shedding process and thus break the
nominally 2D nature of Karman vortex shedding, this passive device should work
for other 2D bluff bodies such as the circular cylinder. Yoon (2005) found that the
tab located near the separation point (Figure 4h) reduces drag on a circular cylinder
and attenuates the Karman vortex shedding in the wake, and the optimal spanwise
spacing between the adjacent tabs is similar to that in Darekar & Sherwin 2001 (see
Supplemental Figure 3).
However, for 3D bluff bodies such as a sphere and some transportation vehicles,
the vortical structures are essentially three dimensional (Figure 2) (Yun et al. 2006).
In this case, the 3D geometric modications described above may not produce any
drag reduction because they promote 3D vortical activities in the wake. Our preliminary study on ow over a 3D body with a tab did not produce any drag reduction.
Therefore, some other types of passive devices should be developed for the reduction
of drag on a 3D body.
Active open-loop control. When a time-periodic open-loop forcing is applied, vortex shedding in the wake is in general locked in phase to the forcing (Blevins 1990);
consequently, the forcing strengthens vortex shedding and increases the mean drag
and lift uctuations. There are a few successful active open-loop controls that attenuate vortex shedding and reduce drag. One example is the high-frequency rotation
of the circular cylinder by Tokumaru & Dimotakis (1991) and another is base bleed
(Bearman 1967, Wood 1964). Although these controls are effective in reducing drag,
their efciencies are not that high.

www.annualreviews.org Control of Flow Over a Bluff Body

119

ANRV332-FL40-06

ARI

10 November 2007

16:9

u
x

Annu. Rev. Fluid Mech. 2008.40:113-139. Downloaded from www.annualreviews.org


Access provided by Indian Institute of Technology - Kanpur on 10/18/16. For personal use only.

2
Figure 6
Distributed forcing: (a) schematic of the forcing, (b) Re = 100 ( = 5d ), and (c) Re = 3900
( = d ). Shown in panels b and c are the instantaneous vortical structures without (left
column) and with (right column) control. Figure taken from Kim & Choi 2005.

In the literature, only a few active-control methods employ 3D forcing. Kim &
Choi (2005) numerically investigated the effect of 3D (referred to as distributed
in that paper) forcing on the drag and lift forces on a circular cylinder at a wide
range of Reynolds numbers (Re = 47 3900). The distributed forcing considered
comprises blowing and suction from the slots located at upper and lower surfaces
of the cylinder. The blowing-and-suction prole from each slot is sinusoidal in the
spanwise direction but is steady in time (Figure 6a):
1 (z) = 2 (z) = o sin(2 z/),
where 1 and 2 are the radial velocities at the upper and lower slots, respectively;
z is the spanwise direction; o is the forcing amplitude; and is the forcing wavelength. For all Reynolds numbers larger than 46, the distributed forcing attenuates
or annihilates the Karman vortex shedding as shown in Figure 6b,c and thus significantly reduces the mean drag and the drag and lift uctuations. Owing to control,
vortex shedding completely disappears at Re = 100 (Figure 6b), nearly disappears
in the near wake, and reappears weakly in the far wake at Re = 3900 (Figure 6c).
These authors also showed that distributed forcing produces phase mismatch along
the spanwise direction in vortex shedding, weakens the strength of vortical structures
in the wake, and thus reduces drag.
Importantly, drag reduction by distributed forcing is caused by the direct interaction with vortex shedding, not by the separation delay (see Section 4.2 for details).
This fact suggests that distributed forcing also should be applicable to a body with
xed separation for drag reduction or reduction of lift uctuations. Kim et al. (2004)
applied distributed forcing to turbulent ow over a 2D model vehicle having a blunt
trailing edge and obtained a signicant amount of drag reduction. Therefore, 3D
forcing should be applicable to ow over any 2D bluff body, which contains nominally 2D Karman vortex shedding, for drag reduction at various Reynolds numbers.

120

Choi


Jeon

Kim

ANRV332-FL40-06

ARI

10 November 2007

16:9

3.2. Feedback Control

Annu. Rev. Fluid Mech. 2008.40:113-139. Downloaded from www.annualreviews.org


Access provided by Indian Institute of Technology - Kanpur on 10/18/16. For personal use only.

Feedback control methods are an attractive choice over passive and active open-loop
controls in that the control input is continuously modied according to the response
of the ow system. In this section, we describe successful feedback control methods
such as the single-sensor linear feedback control, optimal and suboptimal controls,
and control based on ROMs.
Single-sensor linear feedback control. Flow past a bluff body above the critical
Reynolds number possesses multiple global modes although the Karman mode is the
rst global mode to become unstable. In principle, a successful feedback control of
wake requires the suppression of all these global modes. However, at Reynolds numbers just above the critical value, the wake has typically one unstable mode resulting
from Karman vortex shedding (Huerre & Monkewitz 1990). Therefore, a linear feedback control of such ows with vortex shedding at low Reynolds numbers is possible
using a single-sensor actuator feedback loop.
Berger (1967) rst introduced the single-sensor feedback control by actuating a
bimorph cylinder with signals from a hot-wire sensor located in the wake, reporting
the possibility of suppressing vortex shedding behind a bluff body. Ffowcs Williams
& Zhao (1989) performed a feedback control of vortex shedding at Re = 400 with a
loudspeaker based on the velocity phase information measured at a point in the wake
and showed that velocity uctuations at the nominal vortex shedding frequency were
reduced. Roussopoulos (1993) conducted a similar feedback control experiment at low
Reynolds numbers and reported that control delays the onset of wake instability, but
a single-sensor actuator feedback loop does not stabilize the wake at high Reynolds
numbers. He showed an approximate map of the sensor location at which the linear
feedback control was possible at Re = 65 (see Supplemental Figure 4). In the
near wake, the presence of the sensor alone suppressed vortex shedding. Away from
the centerline of the vortex street, the unsteadiness caused by shedding is too weak
to be used as a control signal. In the far wake beyond approximately 9 diameters
downstream, the vortex street was clearly detectable, but the feedback using this
signal was unable to suppress shedding.
Park et al. (1994) conducted a numerical simulation on a similar feedback approach
by using a pair of blowing/suction slots on a cylinder and a single feedback sensor
located in the wake (Supplemental Figure 5a). The linear feedback law used was
upper (t) v(xs , y = 0, t) and lower (t) v(xs , y = 0, t), where upper and lower
denote the blowing and suction velocities at upper and lower slots, respectively, and
v(xs , y = 0) is the centerline transverse velocity at the sensing position xs . They showed
that control results are sensitive to the feedback sensor location (see, for example,
Supplemental Figure 5b for drag variation with the sensor location), and a complete
suppression of vortex shedding is obtained at Re = 60 (Supplemental Figure 5c).
However, at a higher Reynolds number, vortex shedding was not suppressed owing
to the generation of secondary vortices from control.
Huang (1996) conducted a linear feedback control for ow over a circular cylinder.
The velocity was measured at the upper shear layer and antiphase-shifted to drive

www.annualreviews.org Control of Flow Over a Bluff Body

121

ARI

10 November 2007

16:9

a loudspeaker that generated feedback sound from the slot on the upper cylinder
surface. This study demonstrated that vortex shedding in the downstream location
of the sensor position is suppressed with the feedback sound. Although only one side
of the shear ow was manipulated, vortex shedding was suppressed at both sides of
the cylinder, indicating that vortex shedding is the result of instability involving two
parallel shear ows.
Zhang et al. (2004) developed a proportional-integral-derivative (PID) controller
to suppress the in-phase vortex shedding and vortex-induced vibration on a springsupported square cylinder at the resonance condition. Three piezoelectric ceramic
actuators were used to perturb the upper surface of the cylinder. They considered
three control schemes using different feedback signals, such as the streamwise velocity
signal (PID-u control), ow-induced structural oscillation signal (PID-Y control),
and a combination of both signals (PID-Yu control). The feedback control based on
PID-Yu showed the best result and led to an almost complete suppression of vortex
shedding.

Annu. Rev. Fluid Mech. 2008.40:113-139. Downloaded from www.annualreviews.org


Access provided by Indian Institute of Technology - Kanpur on 10/18/16. For personal use only.

ANRV332-FL40-06

Optimal control theory and suboptimal control. Optimal control theory has received much attention in the eld of ow control. An example of a theoretical study
on controlling uid ow is Abergel & Temam (1990), who applied an optimal control
theory to some uid mechanics problems and derived optimality conditions for various physical situations. The procedure they developed, however, required velocity
information inside the ow to solve the adjoint equations, from which a feedback
control input was derived. Therefore, the application of their optimal control algorithm to the unsteady 3D Navier-Stokes equations is not practical in real situations.
He et al. (2000) and Protas & Styczek (2002) applied the optimal control theory to the
control of the unsteady wake of a cylinder by rotary oscillation. Li et al. (2003) applied
an adjoint-based optimal control to the problem of vortex-shedding suppression, via
blowing and suction at the cylinder surface. By minimizing the cost function dened
by the difference of the velocity eld from that of steady laminar ow, they showed
successful control of vortex shedding at low Reynolds numbers. They also pointed
out that the optimization time interval should be larger than the vortex-shedding period. However, as mentioned above, direct implementation of these optimal controls
into real situations is not possible currently because the algorithm requires the full
ow-eld information for a global time period.
To overcome the difculty of the procedure developed by Abergel & Temam
(1990), Choi et al. (1993) introduced a suboptimal feedback control algorithm, which
avoids the iterations required for a global optimal control by seeking an optimal
condition over a short time period. This control algorithm has been successfully
applied to ows behind a circular cylinder (Min & Choi 1999), a backward-facing
step (Choi et al. 1999, Kang & Choi 2002), and a sphere ( Jeon & Choi 2005).
Min & Choi (1999) applied a suboptimal feedback control to ow around a circular
cylinder at Re = 100 and 160. The location of sensors for feedback was limited
to the cylinder surface, and the control input from actuators was the blowing and
suction on the cylinder surface. The cost function to be minimized was the difference
between the real and potential pressures on the cylinder surface, and the control input

122

Choi


Jeon

Kim

ANRV332-FL40-06

ARI

10 November 2007

16:9

1
0.5

Cp

0.05

0
0.5

Real Cp

Potential Cp

0.05

1.5
0

90

Annu. Rev. Fluid Mech. 2008.40:113-139. Downloaded from www.annualreviews.org


Access provided by Indian Institute of Technology - Kanpur on 10/18/16. For personal use only.

0.1

180

0.1

90

180

Figure 7
Suboptimal control of ow over a sphere at Re = 425: (a) real and potential pressure
coefcients on the sphere surface, (b) blowing/suction prole along the polar angle,
and instantaneous vortical structures (c) without and (d ) with control.

was determined based on the measurement of instantaneous surface pressure. As a


result, vortex shedding became weak or disappeared, and the mean drag and drag/lift
uctuations signicantly decreased. Recently, Jeon & Choi (2005) applied suboptimal
control to ow over a sphere at Re = 425. The cost function to be minimized
was the difference between the real and potential pressures on the sphere surface
(Figure 7a). The control input was the blowing and suction on the sphere surface and
was determined from the measurement of instantaneous surface pressure (Figure 7b).
As a result, they obtained signicant drag reduction through the change in vortical
structures (Figure 7c,d ).
A key element of the optimal and suboptimal controls is the determination of
the cost function (e.g., drag or lift) to be minimized. One of the main difculties
in this approach is that the control result heavily depends on the choice of the cost
function. For example, drag is reduced more when the cost function is dened as
the difference between the real- and potential-ow surface pressures than when the
cost function is the drag itself (Min & Choi 1999). This is because the cost function
dened is a nonlinear function of the control input. Therefore, the physical intuition
or knowledge of the ow is required when choosing a cost function.

Control based on reduced-order models. As mentioned above, the linear feedback


control based on a single-sensor measurement is feasible at low Reynolds numbers
because the wake has typically one unstable mode resulting from Karman vortex
shedding. At high Reynolds numbers, however, the wake has multiple unstable modes
that require multiple feedback sensors to accurately estimate the ow eld. Then, an
important question is how to reduce the number of feedback sensors while keeping
dynamically important ow-eld information. One way to achieve this goal is to
develop a ROM. The basic idea of a ROM is that the spatiotemporal information of
a ow eld needed for feedback control is provided by a low-dimensional description
of the ow features based on the proper orthogonal decomposition (POD), point
vortex model, and so forth. Once the ROM is set up, one can develop a simpler and

www.annualreviews.org Control of Flow Over a Bluff Body

123

ARI

10 November 2007

16:9

computationally more feasible controller than that based on the full information of
ow eld.
In POD, more energetic modes (possibly corresponding to large-scale structures)
to less energetic modes (possibly corresponding to small-scale structures) are extracted from the ow eld (see Holmes et al. 1996), and a ROM based on POD
truncates less energetic modes for feedback. Gillies (1998) introduced a simplied
model wake and applied POD to obtain a ROM. A neural network controller was
used to determine the control input, which successfully suppressed the oscillations
corresponding to vortex shedding. Graham et al. (1999) obtained a ROM based on
POD from a controlled ow over a circular cylinder at Re = 100. The control input
was the cylinder rotation, and the controller was devised using the optimal control
theory. They showed that control reduces the level of unsteadiness in the wake, but
the prediction errors of the ROM increase during control, which degrades the control
result. Therefore, they needed periodic resetting of the ROM to maintain control
effectiveness. Bergmann et al. (2005) used a similar approach, demonstrating that the
numerical cost of using a ROM based on POD is negligible as compared with using
the full Navier-Stokes equation. Siegel et al. (2006) used only the rst POD mode
(see Supplemental Figure 6 for the rst four POD modes) to control ow over a
circular cylinder at Re = 100 using a linear proportional and differential feedback
controller through the transverse displacement of the cylinder. They also showed
that control stabilizes wake unsteadiness.
Another way of obtaining a ROM is to use a vortex model when the ow eld is
essentially governed by vortices in the wake. Cortelezzi (1996; Cortelezzi et al. 1997)
performed a nonlinear feedback control of ow past a vertical at plate with a suction
point on the downstream wall of the plate. The vortices in the separated shear layer
were modeled by a pair of point vortices having time-dependent circulation. Li &
Aubry (2003) analytically derived a feedback control strategy via a reduced model,

called Foppls
potential ow model, and showed that control successfully suppresses

nonzero lift at Re = 100 and 200. Protas (2004) also used the Foppl
point vortex
system as a ROM for the stabilization of wake at Re = 75 and discussed the possibilities
and limitations of the ow-control strategies based on the point vortex system as a
ROM. Henning et al. (2006) experimentally controlled the ow over a 2D bluff body
with a blunt trailing edge using a low-dimensional vortex model, showing delayed
vortex shedding and drag reduction by control.
One concern from the reduced-order-model approach is that control based on a
ROM is not mathematically equivalent to that based on the full Navier-Stokes equation, and thus it may yield a local optimum for the original dynamical system (see, e.g.,
Bergmann et al. 2005). Another concern is that it has not been fully studied whether
the retained modes from the ROM approach are uncontrollable or unobservable for
ow over a bluff body (see Rowley 2005 and Kim & Bewley 2007 for related issues).

Annu. Rev. Fluid Mech. 2008.40:113-139. Downloaded from www.annualreviews.org


Access provided by Indian Institute of Technology - Kanpur on 10/18/16. For personal use only.

ANRV332-FL40-06

3.3. Control Based on Local/Global Instability


New insights into vortex shedding in a bluff-body wake have been provided from
stability theory (Chomaz 2005, Huerre & Monkewitz 1990). The concepts of local

124

Choi


Jeon

Kim

Annu. Rev. Fluid Mech. 2008.40:113-139. Downloaded from www.annualreviews.org


Access provided by Indian Institute of Technology - Kanpur on 10/18/16. For personal use only.

ANRV332-FL40-06

ARI

10 November 2007

16:9

absolute and convective instabilities were rst applied to a plane mixing layer (Huerre
& Monkewitz 1985). After that, the analysis of absolute instability was carried out in
the wake (Monkewitz 1988), and now it is well known that the existence of a region
having absolute instability is necessary for vortex shedding. There is much evidence
in the literature that global wake instability is in fact governed by absolute instability
in the near wake. This gives rise to many renewed discussions about the mechanisms
of successful controls of vortex shedding.
Mechanisms of some successful methods of controlling vortex shedding are understood in connection with changes in local absolute instability. For example, base
bleed eliminates or weakens local absolute instability in the near wake (Monkewitz
1988). A small secondary cylinder placed in the near wake for vortex-shedding suppression may be associated with changes in absolute instability (Schumm et al. 1994,
Strykowski & Sreenivasan 1990). Yu & Monkewitz (1990) investigated wake heating
and found that heating decreases local absolute-instability growth rates by reducing
the uid density of the near wake. Leu & Ho (2000) investigated the effect of base
suction on a plane wake and found that it signicantly reduces the length of the
absolutely unstable region in the near wake and suppresses global instability.
Recently, motivated by the work of Strykowski & Sreenivasan (1990), Giannetti
& Luchini (2003) and Hwang & Choi (2006) studied the effects of a small secondary
control cylinder on the global and absolute instability in the wake, respectively. They
obtained regions in which the global or absolute instability is attenuated. These
regions are the separated shear layer and the wake centerline. Figure 8 compares
2

y/d

0
Hwang &
Choi

1
Strykowski &
Sreenivasan

2
0

x/d
Figure 8
Comparison of regions of attenuating instability by a small secondary cylinder: blue lines from
Strykowski & Sreenivasan (1990), red lines from Hwang & Choi (2006). Note that the results
from Strykowski & Sreenivasan are obtained for d/ds = 10, whereas those from Hwang &
Choi are for d /d s  1. Here, ds is the diameter of the secondary cylinder. Strykowski &
Sreenivasan (1990) showed experimentally that the region of attenuating instability is
signicantly affected by the magnitude of d/ds and shrinks with increasing d/ds .

www.annualreviews.org Control of Flow Over a Bluff Body

125

ARI

10 November 2007

16:9

the results of Strykowski & Sreenivasan (1990) and Hwang & Choi (2006), showing
qualitatively good agreement between them.
The vortex-shedding attenuation from the distributed forcing by Kim & Choi
(2005) shown in Figure 6 was also explained by absolute-instability analysis (Kim
et al. 2006). Kim et al. (2006) found that the disturbance of having specic spanwise wavelengths reduces the local absolute growth rate of the model wake prole
suggested by Monkewitz (1988) and thus suppresses vortex shedding. The optimal
wavelength predicted from stability analysis was approximately 3d, which is in good
agreement with the results of 3D active and passive forcing (Darekar & Sherwin 2001,
Kim & Choi 2005).
Note that the instability analysis mentioned above is linear; hence special care
must be taken when it is applied to high-Reynolds-number ows because the linearized instability equations can be highly non-normal at high Reynolds numbers. If
the equations for linear instability are non-normal, a disturbance that stabilizes the
absolute-instability frequency may destabilize other eigenvalues by more than the
original absolute-instability frequency. Fortunately, methods for wake control such
as splitter plate, base bleed, and distributed forcing provide good results for a wide
range of Reynolds numbers. This may indicate that vortex shedding continues to be
the prime mode of wake ow up to quite high Reynolds numbers, and the related
dynamics remains similar to that of low-Reynolds-number vortex shedding.
The absolute growth rate is a function of the mean velocity prole of basic ow
(i.e., steady unstable ow), and one must modify it to control the absolute instability
and hence the ow. Therefore, to control the absolute instability, one has to supply an
adequate level of power to modify the basic wake prole, so it is inherently difcult
to devise a highly efcient control method for absolutely unstable ows such as a
wake with vortex shedding. This explains why base bleed and base suction require a
high level of input velocity, comparable with free-stream velocity, to suppress vortex
shedding: For example, bleed 0.2u for base bleed (Schumm et al. 1994) and
suction 0.46u for base suction (Leu & Ho 2000). Conversely, the high effectiveness
of 3D forcing such as distributed forcing (blowing/suction = 0.07u at Re = 100; Kim &
Choi 2005) or tab (Park et al. 2006) attracts our attention. It is not clear whether this
high effectiveness comes from an amplication of disturbances through the separated
shear layer or from the high sensitivity of the absolute instability to the spanwise
waviness.

Annu. Rev. Fluid Mech. 2008.40:113-139. Downloaded from www.annualreviews.org


Access provided by Indian Institute of Technology - Kanpur on 10/18/16. For personal use only.

ANRV332-FL40-06

3.4. Synthetic Jet


In the 1990s, Glezer and his coworkers (Glezer & Amitay 2002, Smith & Glezer
1998) invented an attractive control device called the synthetic jet. The synthetic jet
generates vortices at the edge of an orice by the time-periodic motion of a exible
diaphragm in a sealed cavity. The net mass ux of the synthetic jet is zero, but nonzero
linear momentum is transferred to the ow.
Amitay et al. (1998) and Glezer & Amitay (2002) showed that the application
of a high-frequency forcing (Std = f d /u = 6.1, 2.6, and 1.5 for Re = 31,000;
75,000; and 131,000; respectively) from a synthetic jet to ow over a circular cylinder

126

Choi


Jeon

Kim

ANRV332-FL40-06

ARI

10 November 2007

16:9

Annu. Rev. Fluid Mech. 2008.40:113-139. Downloaded from www.annualreviews.org


Access provided by Indian Institute of Technology - Kanpur on 10/18/16. For personal use only.

produced a signicant drag reduction, although the lift was increased by the forcing
because the actuator was installed at only one side of the cylinder. They also applied
another forcing frequency of Std = 4.5 for Re = 75,000 and found that the drag variation was insensitive to the forcing frequency. Therefore, they argued that global effects
such as the variation of aerodynamic forces are decoupled from the forcing frequency.
There have been many control studies using synthetic jets in the literature. Details
are found in a review by Glezer & Amitay (2002); thus we do not repeat them here.
Nevertheless, we feel that a detailed and systematic analysis of the effect of a synthetic
jet on the ow past a bluff body such as the circular cylinder is still lacking in the
literature, and thus further studies should be conducted in the near future.

4. BOUNDARY-LAYER CONTROL VERSUS DIRECT-WAKE


CONTROL
Control methods for ow over a bluff body can be classied into two groups, depending on which part of the ow the control aims to modify: the boundary-layer
control and the direct-wake control. In the rst, control changes the boundary-layer
ow characteristics from laminar to turbulent and thus delays the main separation,
resulting in drag reduction. This control method is applied to bluff bodies having a
movable separation point. Conversely, the second control directly modies the wake
characteristics, and thus it can be applied to all kinds of bluff bodies having a xed or
movable separation point.

4.1. Boundary-Layer Control


In this section, we consider bluff bodies having a movable separation point. For this
type of bluff bodies, the delay of main separation produces signicant drag reduction.
To delay separation, one should enhance the near-wall streamwise momentum near
and before the separation point, so it can overcome the adverse pressure gradient
formed in the rear part of the bluff body. The enhancement of near-wall momentum
can be realized by controls either through the direct boundary-layer transition to
turbulence or through early separation and reattachment before the main separation.
Direct transition to turbulence. The generation of strong near-wall momentum
by the direct transition to turbulence is possible at Reynolds numbers large enough
to trigger boundary-layer instability. Therefore, this type of control has been limited
to subcritical Reynolds numbers.
One example of this is surface roughness. The reduction of drag on a bluff body
by surface roughness has been achieved by quite a few people (Achenbach 1971,

1974; Choi 2006; Guven


et al. 1980). Achenbach (1971, 1974) showed that with
surface roughness the drag coefcient sharply decreases and then increases rapidly
with increasing Reynolds number, showing a local minimum at a critical Reynolds
number. This critical Reynolds number decreases with increasing roughness. Also,
this minimum drag coefcient at the critical Reynolds number increases more at
larger roughness (Figure 9). Choi (2006) conducted a detailed measurement of

www.annualreviews.org Control of Flow Over a Bluff Body

127

ANRV332-FL40-06

ARI

10 November 2007

16:9

0.6

Smooth (Achenbach 1972)


Dimple (k/d = 0.9 x 103;
Bearman & Harvey 1976)

0.5

Dimple (k/d = 0.4 x 103;


Choi et al. 2006a)
0.4

Roughness (k/d = 1.25 x 103;


Achenbach 1974)
Roughness (k/d = 0.5 x 103;
Achenbach 1974)

Annu. Rev. Fluid Mech. 2008.40:113-139. Downloaded from www.annualreviews.org


Access provided by Indian Institute of Technology - Kanpur on 10/18/16. For personal use only.

CD 0.3

Roughness (k/d = 0.67 x 103;


Choi 2006)

0.2

0.1

1e + 5

1e + 6

Re
Figure 9
Variations of the drag coefcient for smooth, dimpled, and roughened spheres with Reynolds
number. Here, k is the dimple depth or the roughness height.

streamwise velocity above the roughened surface and conrmed that the boundarylayer transition to turbulence occurs earlier at higher Reynolds number after the
critical Reynolds number.
Another popular device for producing turbulence before the main separation is
a vortex generator located at the leading edge of a bluff body [see the review by
Lin (2002)]. This vortex generator produces streamwise vortices in a boundary layer,
and thus strong near-wall momentum is generated through mixing enhancement,
resulting in the delay of main separation.
Early separation and reattachment before main separation. In uncontrolled
ows over a circular cylinder and sphere, the drag coefcients rapidly decrease down
to approximately 0.25 and 0.07, respectively, and this phenomenon has been called
the drag crisis (Achenbach 1972, Bearman 1969, Fage 1936, Farell & Blessmann
1983) (see Figure 9). The cause of this rapid drag-coefcient reduction is the existence of small separation bubble(s) above the surface (Supplemental Figure 7)
(Suryanarayana & Prabhu 2000). At the critical Reynolds number, disturbances
existing in the boundary layer rapidly grow along the separated shear layer, and highmomentum uids in the free stream are entrained toward the bluff-body surface.
This causes the reattachment of the ow (thus forming a separation bubble above the
surface) and generates strong near-wall momentum, resulting in the delay of main
separation.
The generation of strong near-wall momentum by early separation and reattachment is possible at the critical Reynolds number in uncontrolled ow (Supplemental

128

Choi


Jeon

Kim

ANRV332-FL40-06

ARI

10 November 2007

a
110

90

16:9

Figure 10

b
Main separation

Annu. Rev. Fluid Mech. 2008.40:113-139. Downloaded from www.annualreviews.org


Access provided by Indian Institute of Technology - Kanpur on 10/18/16. For personal use only.

110

90

Oil (top panels) and


smoke-wire (bottom panels)
ow visualizations from
high-frequency forcing
(a) without control and
(b) with control at
Std = f d /u = 4.95. Here,
the ow goes from right to
left. Figure taken from Jeon
et al. 2004.

Separation bubble

Figure 7). In the subcritical Reynolds-number range, one must provide a control
input to generate this early separation and reattachment that produces strong nearwall momentum and delays main separation.
Jeon et al. (2004) conducted an active control of ow over a sphere for drag
reduction using a local time-periodic blowing and suction at subcritical Reynolds
numbers (Re = 6 104 2 105 ) and obtained signicant reductions of the drag
coefcient at forcing frequencies much higher than the vortex-shedding frequency
(Supplemental Figure 8). They showed that disturbances from high-frequency
forcing rapidly grow along the separated shear layer, and high momentum in the free
stream is entrained toward the sphere surface, resulting in the reattachment of the
ow, thus forming a separation bubble above the sphere surface (see Figure 10), and
the delay of main separation. This mechanism is nearly identical to that observed in
the drag crisis.
Choi et al. (2006a) presented a mechanism of drag reduction by dimples on a
sphere, such as golf-ball dimples, by measuring the streamwise velocity above the
dimpled surface. They showed that dimples cause local ow separation and trigger
shear layer instability along the separated shear layer, resulting in the generation of
large turbulent intensity, reattachment to the sphere surface with high momentum
near the wall, and delay of main separation (Supplemental Figure 9). Again, this
mechanism is not very different from that of the drag crisis.
Nearly the same drag-reduction mechanisms can be found from controls using
a trip wire located at the front surface of a bluff body (Maxworthy 1969), transverse groove (Kimura & Tsutahara 1991), and free-stream turbulence (Blackburn &
Melbourne 1996, Kiya et al. 1982) [see Choi et al. (2006b) and Son et al. (2006) for
the detailed mechanisms of drag reduction by trip and free-stream turbulence, respectively]. Therefore, the generation of a separation bubble before main separation

www.annualreviews.org Control of Flow Over a Bluff Body

129

ANRV332-FL40-06

ARI

10 November 2007

16:9

(i.e., a closed-loop streamline consisting of early separation and reattachment) may


be an important ow-control strategy for drag reduction on a bluff body having a
movable separation point such as a sphere and cylinder.

4.2. Direct-Wake Control

Annu. Rev. Fluid Mech. 2008.40:113-139. Downloaded from www.annualreviews.org


Access provided by Indian Institute of Technology - Kanpur on 10/18/16. For personal use only.

In this section, we explain control methods that perform direct-wake modication


not caused by separation delay through the boundary-layer transition. We classify
these control methods into two groups depending on whether control generates 2D
or 3D disturbances to the base ow.
2D forcing. Typical 2D forcings for direct-wake control are the splitter plate located
at the base surface or downstream of a bluff body (Anderson & Szewczyk 1997,
Bearman 1965, Hwang et al. 2003, Kwon & Choi 1996, Ozono 1999, Roshko 1955)
and base bleed (Arcas & Redekopp 2004, Bearman 1967, Delaunay & Kaiktsis 2001,
Wood 1964, Yao & Sandham 2002). The splitter plate and base bleed stabilize the
near wake by delaying the interaction of top and bottom shear layers in the wake and
hence suppress vortex shedding. The attenuation of vortex strength in the near-wake
region leads to an increase in the base pressure, resulting in drag reduction. However,
the splitter plate may not be so practical in certain situations because a relatively large
splitter plate (at least the length of cylinder diameter) is required for an effective drag
reduction, and also it is sensitive to the ow direction. The base bleed requires a
relatively high power input for an effective drag reduction because it is applied inside
the recirculation region.
Most feedback controls (described above in Sections 3.2 and 3.3) based on the
linear feedback controller, ROM, and local/global instability modications belong to
this direct-wake-modication group.
3D forcing. Controls providing 3D disturbances to the base ow are explained above
in Section 3.1, and thus we do not repeat them here. We do emphasize here that
controls change the wake eld directly, not through the separation delay. It is manifest
that, in the case of bluff bodies with a blunt trailing edge, successful controls such as 3D
geometric modications have no relation with the boundary-layer modication and
thus directly modify the wake (Darekar & Sherwin 2001, Park et al. 2006, Petrusma &
Gai 1994, Rodriguez 1991, Tanner 1972, Tombazis & Bearman 1997). In the case of a
circular cylinder having a movable separation point, Kim & Chois (2005) distributed
forcing also directly modies the wake eld: The separation angle is delayed at suction
locations but signicantly advances at blowing locations (Supplemental Figure 10).
For this reason, distributed forcing was also successful for a 2D model vehicle with a
blunt trailing edge (Kim et al. 2004).
Unlike controls based on the boundary-layer transition to turbulence, these controls performing direct-wake modication can be applied to bluff-body types both
with and without xed separation.

130

Choi


Jeon

Kim

ANRV332-FL40-06

ARI

10 November 2007

16:9

5. FURTHER DISCUSSION

Annu. Rev. Fluid Mech. 2008.40:113-139. Downloaded from www.annualreviews.org


Access provided by Indian Institute of Technology - Kanpur on 10/18/16. For personal use only.

5.1. Reynolds-Number Dependence


It is well known that ow over a bluff body exhibits successive bifurcations with
increasing Reynolds number and shows different characteristics at different Reynoldsnumber ranges. Therefore, it is questionable that a successful control method in a
certain Reynolds-number range works for other Reynolds-number ranges.
Clearly, some control methods work for various ow regimes but some do not,
depending on how the control objective, such as drag reduction, is achieved. For
example, Park et al.s (1994) linear feedback control showed complete suppression of
vortex shedding at a low Reynolds number of 60, but it did not for a higher Reynolds
number. This is because ow at a higher Reynolds number contains global oscillations
with frequencies different from that of the original vortex shedding, and thus a simple feedback control law based on a single-sensor measurement cannot completely
stabilize the wake. In the case of surface roughness (Achenbach 1971, 1974), drag
decreases signicantly at subcritical Reynolds numbers (104 < Re < critical Reynolds
number) but does not change at a lower Reynolds number because the boundary-layer
ow is not destabilized at those numbers. Conversely, control methods that change
local/global instability modes seem to work for various Reynolds numbers, such as
base bleed (Schumm et al. 1994), some 3D geometric modications (Park et al. 2006),
and distributed forcing (Kim & Choi 2005, Kim et al. 2004).
So far, the Reynolds number has been dened based on the free-stream velocity
and body height h. Another important length scale may be the boundary-layer thickness (or momentum thickness, ) at or just before separation. Petrusma & Gai (1994)
showed from their and other experimental data that the base-pressure coefcient,
which is directly related to the drag and vortex strength in the wake, for 2D bluff bodies
with a blunt trailing edge depends on h/. Although it is almost constant for sufciently large h/ (C p b 0.5 0.6 for h/ > 40), the base-pressure coefcient
rapidly increases as h/ decreases (Supplemental Figure 11). This result indicates
that the vortex dynamics in the wake is also affected by the characteristics of boundarylayer ow before separation. Therefore, the momentum thickness before separation
may also be an important parameter to consider in future control development.

5.2. Lowest Possible Drag


What is the lowest possible drag achievable from control in the case of bluff bodies?
This is an important question because knowing the lowest possible drag not only sets
the limit of control performance, but also provides direction for developing efcient
and effective control methods.
It is well known that vortex shedding signicantly increases the mean drag and the
lift uctuations as compared to those without vortex shedding. Thus, one normally
assumes that ow without vortex shedding has the lowest possible drag achievable
by control. However, this ow still has quite large form drag. Thus one may pursue
achieving a controlled ow in which the surface pressure distribution is similar to
that of the potential ow having zero form drag. Min & Chois (1999) control target

www.annualreviews.org Control of Flow Over a Bluff Body

131

ANRV332-FL40-06

ARI

10 November 2007

16:9

was to reduce the difference between the real and potential pressure distributions
on the cylinder surface. Conversely, for Park et al. (1994) and Kim & Choi (2005),
the control target was to eliminate or attenuate vortex shedding. A theoretical study
regarding the lowest possible drag has not been conducted, and we believe this is an
important issue to investigate in the future.

5.3. Control Efficiency

Annu. Rev. Fluid Mech. 2008.40:113-139. Downloaded from www.annualreviews.org


Access provided by Indian Institute of Technology - Kanpur on 10/18/16. For personal use only.

In the case of active control, efciency is an important element and highly depends
on the control method applied. A rigorous estimation of the efciency is not easy
because it depends on how the mechanism of the control device is set up to realize
the control input in practical situations. Thus, we do not consider any additional
losses occurring inside the mechanism of the control device in this section.
In the case of drag reduction, control efciency is dened as the ratio of the save
power to the control input power. For example, when the control input is blowing
and suction distributed over the bluff-body surface, the ideal control efciency may
be dened as
(Du Dc )u
1 = 
,
3 + p ) dA
(0.5
w
A
where Du and Dc are the drags on a bluff body without and with control, respectively;
A is the body surface where the control is applied; is the uid density; pw is the surface
pressure; and is the control velocity. The rst term in the control input power is
the energy convection, and the second one is the pressure work. This denition of
control input power is ideal in the sense that control may completely utilize available
power sources, which is not possible in practical situations. Conversely, the following
denition of control efciency assumes that there is no chance of using an available
power source existing in the ow system:
2 = 

(Du Dc )u
.
(|0.5 3 | + | p w |) dA
A

It can be argued that the control method is sufciently efcient when 2  1


even if one considers additional losses occurring inside the mechanism of the control
device. The actual efciency of a control method may be between 1 and 2 . For Kim
& Chois (2005) distributed forcing, 1 150 and 2 30. Because 2  1, this
control is efcient. Conversely, for base bleed, 1 becomes negative because p w < 0
and | p w | > |0.5 3 |, indicating that 1 may not be a good measure of the efciency,
and 2 3.8. The base bleed is efcient but much less efcient than the distributed
forcing.

6. CONCLUSIONS
In this review, we classify existing control methods in three different ways: (a) passive
versus active open-loop versus active closed-loop controls, (b) 2D versus 3D forcing, and (c) boundary-layer control before the main separation versus direct-wake

132

Choi


Jeon

Kim

Annu. Rev. Fluid Mech. 2008.40:113-139. Downloaded from www.annualreviews.org


Access provided by Indian Institute of Technology - Kanpur on 10/18/16. For personal use only.

ANRV332-FL40-06

ARI

10 November 2007

16:9

modication. Then we discuss recent major achievements in the control of bluff-body


ows, namely 3D forcing, active feedback control, control based on local/global instability, and control with a synthetic jet. Some other important control issues include
Reynolds-number dependence, the lowest possible drag, and control efciency.
As mentioned above, control methods can be categorized into three different
groups: passive, active open-loop, and active closed-loop controls. The active and passive controls differ in that one provides power and the other does not, respectively.
Among the three different control groups, the passive control is easiest to implement in experiments, but this control normally requires a parametric study, which is
sometimes tedious, for successful control output. Conversely, the active closed-loop
control provides the best results in general, once a proper control algorithm (i.e.,
adequate actuation and sensor combination and controller) is set up. However, this
control method is not easy to implement in practical situations even with the rapid
developments in computer power and micromachines, and thus it has been adopted
mainly in numerical studies. It is important to note that, once a successful result is
obtained from an active closed-loop control, the next task should be the further development of easier control methods (active open-loop or passive control) for practical
implementation. This procedure may signicantly reduce the effort spent during the
parametric study required for passive and active open-loop controls. One may have to
ultimately establish a passive device for practical applications, considering its simplicity and lack of input-power requirement. Toward this ultimate goal, future research
should focus signicantly on developing successful active-control methods.
Finally, we stress the importance of developing control methods for ow over a
3D body such as a truck and bus. In this body type, vortical structures in the wake
signicantly change depending on the streamwise length of the body (which may determine the momentum thickness at separation), aspect ratio (i.e., height to width),
body height from the ground, and so forth. Many control studies have been conducted for drag reduction using geometric modications of the body and have shown
successful drag reductions (see, for example, Khalighi et al. 2002, McCallen et al.
2004, Verzicco et al. 2002). However, the detailed drag-reduction mechanism due
to control has not been well understood. Furthermore, the sizes of those geometric
modications are relatively large, O(h) or O(0.1h), to implement in practical situations. Therefore, more studies should be conducted in these ows using both passive
and active controls to obtain realistic drag reduction.

DISCLOSURE STATEMENT
The authors are not aware of any biases that might be perceived as affecting the
objectivity of this review.

ACKNOWLEDGMENT
We gratefully acknowledge nancial support from the National Research Laboratory
Program through the Korean Ministry of Science and Technology.

www.annualreviews.org Control of Flow Over a Bluff Body

133

ANRV332-FL40-06

ARI

10 November 2007

16:9

LITERATURE CITED

Annu. Rev. Fluid Mech. 2008.40:113-139. Downloaded from www.annualreviews.org


Access provided by Indian Institute of Technology - Kanpur on 10/18/16. For personal use only.

Abergel F, Temam R. 1990. On some control problems in uid mechanics. Theor.


Comput. Fluid Dyn. 1:30325
Achenbach E. 1971. Inuence of surface roughness on the cross-ow around a circular
cylinder. J. Fluid Mech. 46:32135
Achenbach E. 1972. Experiments on the ow past spheres at very high Reynolds
numbers. J. Fluid Mech. 54:56575
Achenbach E. 1974. The effect of surface roughness and tunnel blockage on the ow
past spheres. J. Fluid Mech. 65:11325
Amitay M, Smith BL, Glezer A. 1998. Aerodynamic ow control using synthetic jet
technology. 36th AIAA Aerosp. Sci. Meet., Reno, Nev., AIAA Pap. No. 980208
Anderson E, Szewczyk A. 1997. Effects of a splitter plate on the near wake of a circular
cylinder in 2 and 3-dimensional ow congurations. Exp. Fluids 23:16174
Arcas D, Redekopp L. 2004. Aspects of wake vortex control through base blowing/suction. Phys. Fluids 16:45256
Artana G, Sosa R, Moreau E, Touchard G. 2003. Control of the near-wake ow
around a circular cylinder with electrohydrodynamic actuators. Exp. Fluids
35:58088
Baek S, Sung HJ. 1998. Numerical simulation of the ow behind a rotary oscillating
circular cylinder. Phys. Fluids 10:86976
Bearman PW. 1965. Investigation of the ow behind a two-dimensional model with
a blunt trailing edge and tted with splitter plates. J. Fluid Mech. 21:24155
Bearman PW. 1967. The effect of base bleed on the ow behind a two-dimensional
model with a blunt trailing edge. Aeronaut. Q. 18:20724
Bearman PW. 1969. On vortex shedding from a circular cylinder in the critical
Reynolds number region. J. Fluid Mech. 37:57787
Bearman PW. 1984. Vortex shedding from oscillation bluff bodies. Annu. Rev. Fluid
Mech. 16:195222
Bearman PW, Harvey JK. 1976. Golf ball aerodynamics. Aeronaut. Q. 22:11222
Bearman PW, Harvey JK. 1993. Control of circular cylinder ow by the use of
dimples. AIAA J. 31:175356
Bearman PW, Owen JC. 1998. Reduction of bluff-body drag and suppression of
vortex shedding by the introduction of wavy separation lines. J. Fluids Struct.
12:12330
Berger E. 1967. Suppression of vortex shedding and turbulence behind oscillating
cylinders. Phys. Fluids 10:S19193
Berger E, Wille R. 1972. Periodic ow phenomena. Annu. Rev. Fluid Mech. 11:31340
Bergmann M, Cordier L, Brancher J. 2005. Optimal rotary control of the cylinder
wake using proper orthogonal decomposition reduced-order model. Phys. Fluids
17:097101
Blackburn H, Henderson R. 1999. A study of two-dimensional ow past an oscillating
cylinder. J. Fluid Mech. 385:25586
Blackburn H, Melbourne W. 1996. The effect of free-stream turbulence on sectional
lift forces on a circular cylinder. J. Fluid Mech. 306:26792
Blevins R. 1990. Flow-Induced Vibrations. New York: Van Nostrand Reinhold. 2nd ed.

134

Choi


Jeon

Kim

Annu. Rev. Fluid Mech. 2008.40:113-139. Downloaded from www.annualreviews.org


Access provided by Indian Institute of Technology - Kanpur on 10/18/16. For personal use only.

ANRV332-FL40-06

ARI

10 November 2007

16:9

Carberry J, Sheridan J, Rockwell D. 2003. Controlled oscillations of a cylinder: a new


wake state. J. Fluids Struct. 17:33743
Cetiner O, Rockwell D. 2001. Streamwise oscillations of a cylinder in a steady current.
Part 1: locked-on states of vortex formation and loading. J. Fluid Mech. 427:128
Choi H, Hinze M, Kunisch K. 1999. Instantaneous control of backward-facing step
ows. Appl. Numer. Math. 31:13358
Choi H, Temam R, Moin P, Kim J. 1993. Feedback control for unsteady ow and its
application to the stochastic Burgers equation. J. Fluid Mech. 253:50943
Choi J. 2006. Mechanism of drag reduction by surface modication on a sphere: dimples,
roughness and trip wire. PhD thesis. Seoul Natl. Univ., Korea
Choi J, Jeon W-P, Choi H. 2006a. Mechanism of drag reduction by dimples on a
sphere. Phys. Fluids 18:041702
Choi J, Son KM, Jeon W-P, Choi H. 2006b. Characteristics of ow around a sphere
with a surface trip. Bull. Am. Phys. Soc. 51:76
Choi S, Choi H, Kang S. 2002. Characteristics of ow over a rotationally oscillating
cylinder at low Reynolds number. Phys. Fluids 14:276777
Chomaz J-M. 2005. Global instabilities in spatially developing ows: non-normality
and nonlinearity. Annu. Rev. Fluid Mech. 37:36792
Cortelezzi L. 1996. Nonlinear feedback control of the wake past a plate with a suction
point on the downstream wall. J. Fluid Mech. 327:30324
Cortelezzi L, Chen Y, Chang H. 1997. Nonlinear feedback control of the wake past
a plate: from a low-order model to a higher-order model. Phys. Fluids 9:200922
Coutanceau M, Defaye J. 1991. Circular cylinder wake congurations: a ow visualization survey. Appl. Mech. Rev. 44:25537
Dalton C, Xu Y, Owen JC. 2001. The suppression of lift on a circular cylinder due
to vortex shedding at moderate Reynolds numbers. J. Fluids Struct. 15:61728
Darekar RM, Sherwin SJ. 2001. Flow past a square-section cylinder with a wavy
stagnation face. J. Fluid Mech. 426:26395
Delaunay Y, Kaiktsis L. 2001. Control of circular cylinder wakes using base mass
transpiration. Phys. Fluids 13:3285302
Dennis S, Nguyen P, Kocabiyik S. 2000. The ow induced by a rotationally oscillating
and translating circular cylinder. J. Fluid Mech. 407:12344
Fage A. 1936. Experiments on a sphere at critical Reynolds numbers. Aeronaut. Res.
Counc. R & M No. 1766
Farell C, Blessmann J. 1983. On critical ow around smooth circular cylinders. J.
Fluid Mech. 136:37591
Ffowcs Williams J, Zhao B. 1989. The active control of vortex shedding. J. Fluids
Struct. 3:11522
Filler J, Marston P, Mih W. 1991. Response of the shear layers separating from a circular cylinder to small-amplitude rotational oscillations. J. Fluid Mech. 231:48199
Fujisawa N, Takeda G, Ike N. 2004. Phase-averaged characteristics of ow around a
circular cylinder under acoustic excitation control. J. Fluids Struct. 19:15970
Gad-el-Hak M. 2000. Flow Control: Passive, Active and Reactive Flow Management.
Cambridge, UK: Cambridge Univ. Press
Giannetti F, Luchini P. 2003. Receptivity of the circular cylinders rst instability.
Presented at 5th Euromech. Fluid Mech. Conf., Aug. 2428, Toulouse

www.annualreviews.org Control of Flow Over a Bluff Body

135

ARI

10 November 2007

16:9

Gillies E. 1998. Low-dimensional control of the circular cylinder wake. J. Fluid Mech.
371:15778
Glezer A, Amitay M. 2002. Synthetic jets. Annu. Rev. Fluid Mech. 34:50329
Graham WR, Peraire J, Tang KY. 1999. Optimal control of vortex shedding using
low-order models. Part II: model-based control. Int. J. Numer. Methods Eng.
44:97390
Grifn OM, Hall MS. 1991. Review: vortex shedding lock-on and ow control in
bluff body wakes. Trans. ASME J. Fluids Eng. 113:52637

Guven
O, Farell C, Patel V. 1980. Surface-roughness effects on the mean ow past
circular cylinders. J. Fluid Mech. 98:673701
He J-W, Glowinski R, Metcalfe R, Nordlander A, Periaux J. 2000. Active control
and drag optimization for ow past a circular cylinder. 1: oscillatory cylinder
rotation. J. Comput. Phys. 163:83117
Henning L, Pastoor M, King R, Noack BR, Tadmor G. 2006. Feedback control applied
to bluff body wake. Presented at Active Flow Control 2006, Sept. 2729, Berlin
Holmes P, Lumley JL, Berkooz G. 1996. Turbulence, Coherent Structures, Dynamical
Systems and Symmetry. Cambridge, UK: Cambridge Univ. Press
Homescu C, Navon I, Li Z. 2002. Suppression of vortex shedding for ow around a
circular cylinder using optimal control. Int. J. Numer. Methods Fluids 38:4369
Huang X. 1996. Feedback control of vortex shedding from a circular cylinder. Exp.
Fluids 20:21824
Huerre P, Monkewitz PA. 1985. Absolute and convective instabilities in free shear
layers. J. Fluid Mech. 159:15168
Huerre P, Monkewitz PA. 1990. Local and global instabilities in spatially developing
ows. Annu. Rev. Fluid Mech. 22:473537
Hwang J-Y, Yang K-S, Sun S-H. 2003. Reduction of ow-induced forces on a circular
cylinder using a detached splitter plate. Phys. Fluids 15:243336
Hwang Y, Choi H. 2006. Control of absolute instability by basic-ow modication
in parallel wake at low Reynolds number. J. Fluid Mech. 560:46575
Jeon S, Choi H. 2005. Suboptimal feedback control for drag reduction in ow over
a sphere. Bull. Am. Phys. Soc. 50:56
Jeon S, Choi J, Jeon W-P, Choi H, Park J. 2004. Active control of ow over a sphere
for drag reduction at a subcritical Reynolds number. J. Fluid Mech. 517:11329
Kang S, Choi H. 2002. Suboptimal feedback control of turbulent ow over a
backward-facing step. J. Fluid Mech. 463:20127
Khalighi B, Balkanyi SR, Bernal LP. 2002. Analysis of the near wake of bluff bodies in
ground proximity. Presented at ASME Annu. Winter Meet., Nov. 1722, New
Orleans
Kim J, Bewley TR. 2007. A linear systems approach to ow control. Annu. Rev. Fluid
Mech. 39:383417
Kim J, Choi H. 2005. Distributed forcing of ow over a circular cylinder. Phys. Fluids
17:033103
Kim J, Hahn S, Kim JS, Lee DK, Choi J, et al. 2004. Active control of turbulent ow
over a model vehicle for drag reduction. J. Turbul. 5:019

Annu. Rev. Fluid Mech. 2008.40:113-139. Downloaded from www.annualreviews.org


Access provided by Indian Institute of Technology - Kanpur on 10/18/16. For personal use only.

ANRV332-FL40-06

136

Choi


Jeon

Kim

Annu. Rev. Fluid Mech. 2008.40:113-139. Downloaded from www.annualreviews.org


Access provided by Indian Institute of Technology - Kanpur on 10/18/16. For personal use only.

ANRV332-FL40-06

ARI

10 November 2007

16:9

Kim J, Hwang Y, Choi H. 2006. Suppression of absolute instability in a wake by spatially varying disturbances. Presented at Whither Turbul. Predict. Control, March
2629, Seoul Natl. Univ., Seoul, Korea
Kim S, Lee C. 2000. Investigation of the ow around a circular cylinder under the
inuence of an electromagnetic force. Exp. Fluids 28:25260
Kimura T, Tsutahara M. 1991. Fluid dynamic effects of grooves on a circular cylinder
surface. AIAA J. 29:206268
Kiya M, Suzuki Y, Arie M, Hagino M. 1982. A contribution to the free-stream
turbulence effect on the ow past a circular cylinder. J. Fluid Mech. 115:15164
Konstantinidis E, Balabani S, Yianneskis M. 2005. The timing of vortex shedding in a
cylinder wake imposed by periodic inow perturbations. J. Fluid Mech. 543:45
55
Kwon K, Choi H. 1996. Control of laminar vortex shedding behind a circular cylinder
using splitter plates. Phys. Fluids 8:47986
Lee S, Kim H. 1997. The effect of surface protrusions on the near wake of a circular
cylinder. J. Wind Eng. Ind. Aerodyn. 6971:35161
Leu TS, Ho CM. 2000. Control of global instability in a nonparallel near wake. J.
Fluid Mech. 404:34578
Li F, Aubry N. 2003. Feedback control of a ow past a cylinder via transverse motion.
Phys. Fluids 15:216376
Li Z, Navon I, Hussaini M, Le Dimet F. 2003. Optimal control of cylinder wakes via
suction and blowing. Comput. Fluids 32:14971
Lim H, Lee S. 2002. Flow control of circular cylinders with longitudinal grooved
surfaces. AIAA J. 10:202736
Lin J, Pao Y. 1979. Wakes in stratied uids. Annu. Rev. Fluid Mech. 11:31738
Lin J, Towghi J, Rockwell D. 1995. Near-wake of a circular cylinder: control by
steady and unsteady surface injection. J. Fluids Struct. 9:65969
Lin JC. 2002. Review of research on low-prole vortex generators to control
boundary-layer separation. Prog. Aerosp. Sci. 38:389420
Maxworthy T. 1969. Experiments on the ow around a sphere at high Reynolds
numbers. J. Appl. Mech. E 36:598607
McCallen R, Browand F, Ross J, eds. 2004. The Aerodynamics of Heavy Vehicles: Trucks,
Buses, and Trains. Lecture Notes in Applied and Computational Mechanics. Berlin:
Springer-Verlag
Min C, Choi H. 1999. Suboptimal feedback control of vortex shedding at low
Reynolds numbers. J. Fluid Mech. 401:12356
Monkewitz PA. 1988. The absolute and convective nature of instability in twodimensional wakes at low Reynolds numbers. Phys. Fluids 31:9991006
Mutschke G, Gerbeth G, Shatrov V, Tomboulides A. 2001. The scenario of threedimensional instabilities of the cylinder wake in an external magnetic eld: a
linear stability analysis. Phys. Fluids 13:72334
Nakano M, Rockwell D. 1993. The wake from a cylinder subjected to amplitudemodulated excitation. J. Fluid Mech. 247:79110
Nakano M, Rockwell D. 1994. Flow structure in the frequency-modulated wake of a
cylinder. J. Fluid Mech. 266:93119

www.annualreviews.org Control of Flow Over a Bluff Body

137

ARI

10 November 2007

16:9

Nehari D, Armenio V, Ballio F. 2004. Three-dimensional analysis of the unidirectional oscillatory ow around a circular cylinder at low Keulegan-Carpenter and
numbers. J. Fluid Mech. 520:15786
Oertel H. 1990. Wakes behind blunt bodies. Annu. Rev. Fluid Mech. 22:53964
Owen JC, Bearman PW, Szewczyk AA. 2001. Passive control of VIV with drag
reduction. J. Fluids Struct. 15:597605
Owen JC, Szewczyk AA, Bearman PW. 2000. Suppression of Karman vortex shedding. Phys. Fluids 12:S9 (Gallery of Fluid Motion)
Ozono S. 1999. Flow control of vortex shedding by a short splitter plate asymmetrically arranged downstream of a cylinder. Phys. Fluids 11:292834
Park DS, Ladd DM, Hendricks EW. 1994. Feedback control of von Karman vortex shedding behind a circular cylinder at low Reynolds numbers. Phys. Fluids
6:2390405
Park H, Lee D, Jeon W-P, Hahn S, Kim J, et al. 2006. Drag reduction in ow over a
two-dimensional bluff body with a blunt trailing edge using a new passive device.
J. Fluid Mech. 563:389414
Petrusma MS, Gai SL. 1994. The effect of geometry on the base pressure recovery
of the segmented blunt trailing edge. Aeronaut. J. 98:26774
Poncet P. 2002. Vanishing of mode B in the wake behind a rotationally oscillating
circular cylinder. Phys. Fluids 14:202123
Poncet P. 2004. Topological aspects of three-dimensional wakes behind rotary oscillating cylinders. J. Fluid Mech. 517:2753
Protas B. 2004. Linear feedback stabilization of laminar vortex shedding based on a
point vortex model. Phys. Fluids 16:447386
Protas B, Styczek A. 2002. Optimal rotary control of the cylinder wake in the laminar
regime. Phys. Fluids 14:207387
Rodriguez O. 1991. Base drag reduction by the control of three-dimensional unsteady
vortical structures. Exp. Fluids 11:21826
Roshko A. 1955. On the wake and drag of bluff bodies. J. Aeronaut. Sci. 22:12432
Roussopoulos K. 1993. Feedback control of vortex shedding at low Reynolds numbers. J. Fluid Mech. 248:26796
Rowley CW. 2005. Model reduction for uids, using balanced proper orthogonal
decomposition. Int. J. Bifurc. Chaos 15:9971013
Sakamoto H, Haniu H. 1994. Optimum suppression of uid forces acting on a circular
cylinder. Trans. ASME J. Fluids Eng. 116:22127
Schumm M, Berger E, Monkewitz PA. 1994. Self-excited oscillations in the wake of
two-dimensional bluff bodies and their control. J. Fluid Mech. 271:1753
Sevilla A, Martinez-Bazan C. 2004. Vortex shedding in high Reynolds number axisymmetric bluff-body wakes: local linear instability and global bleed control.
Phys. Fluids 16:346069
Shiels D, Leonard A. 2001. Investigation of a drag reduction on a circular cylinder
in rotary oscillation. J. Fluid Mech. 431:297322
Shih WCL, Wang C, Coles D, Roshko A. 1993. Experiments on ow past rough
circular cylinders at large Reynolds numbers. J. Wind Eng. Ind. Aerodyn. 49:351
68

Annu. Rev. Fluid Mech. 2008.40:113-139. Downloaded from www.annualreviews.org


Access provided by Indian Institute of Technology - Kanpur on 10/18/16. For personal use only.

ANRV332-FL40-06

138

Choi


Jeon

Kim

Annu. Rev. Fluid Mech. 2008.40:113-139. Downloaded from www.annualreviews.org


Access provided by Indian Institute of Technology - Kanpur on 10/18/16. For personal use only.

ANRV332-FL40-06

ARI

10 November 2007

16:9

Siegel S, Cohen K, McLaughlin T. 2006. Numerical simulations of a feedbackcontrolled circular cylinder wake. AIAA J. 44:126676
Smith BL, Glezer A. 1998. The formation and evolution of synthetic jets. Phys. Fluids
31:228197
Son K, Choi J, Jeon W-P, Choi H. 2006. Effect of free-stream turbulence on ow
over a sphere. Bull. Am. Phys. Soc. 51:76
Strykowski PJ, Sreenivasan KR. 1990. On the formation and suppression of vortex
shedding at low Reynolds numbers. J. Fluid Mech. 218:71107
Suryanarayana GK, Prabhu A. 2000. Effect of natural ventilation on the boundary
layer separation and near-wake vortex shedding characteristics of a sphere. Exp.
Fluids 29:58291
Tanner M. 1972. A method of reducing the base drag of wings with blunt trailing
edges. Aeronaut. Q. 23:1523
Tokumaru PT, Dimotakis PE. 1991. Rotary oscillatory control of a cylinder wake. J.
Fluid Mech. 224:7790
Tombazis N, Bearman PW. 1997. A study of three-dimensional aspects of vortex
shedding from a bluff body with a mild geometric disturbance. J. Fluid Mech.
330:85112
Van Dyke M. 1982. An Album of Fluid Motion. Stanford, CA: Parabolic Press
Verzicco R, Fatica M, Iaccarino G, Moin P, Khalighi B. 2002. Large eddy simulation
of road vehicle with drag-reduction devices. AIAA J. 40:244755
Williams D, Mansy H, Amato C. 1992. The response and symmetry properties of a
cylinder wake subjected to localized surface excitation. J. Fluid Mech. 234:7196
Williamson CHK. 1996. Vortex dynamics in the cylinder wake. Annu. Rev. Fluid
Mech. 28:477539
Williamson CHK. 2004. Vortex-induced vibrations. Annu. Rev. Fluid Mech. 36:413
55
Wood CJ. 1964. The effect of base bleed on a periodic wake. J. R. Aeronaut. Soc.
68:47782
Yao YF, Sandham ND. 2002. Direct numerical simulation of turbulent trailing-edge
ow with base ow control. AIAA J. 40:170816
Yoon J. 2005. Control of ow over a circular cylinder using wake disrupter. MS thesis.
Seoul Natl. Univ., Korea
Yu MH, Monkewitz PA. 1990. The effect of nonuniform density on the absolute
instability of two-dimensional inertial jets and wakes. Phys. Fluids A 2:117581
Yun G, Kim D, Choi H. 2006. Vortical structures behind a sphere at subcritical
Reynolds numbers. Phys. Fluids 18:015102
Zdravkovich MM. 1981. Review and classication of various aerodynamic and hydrodynamic means for suppressing vortex shedding. J. Wind Eng. Ind. Aerodyn.
7:14589
Zhang M, Cheng L, Zhou Y. 2004. Closed-loop-controlled vortex shedding and
vibration of a exibly supported square cylinder under different schemes. Phys.
Fluids 16:143948

www.annualreviews.org Control of Flow Over a Bluff Body

139

AR332-FM

ARI

22 November 2007

21:57

Annu. Rev. Fluid Mech. 2008.40:113-139. Downloaded from www.annualreviews.org


Access provided by Indian Institute of Technology - Kanpur on 10/18/16. For personal use only.

Contents

Annual Review of
Fluid Mechanics
Volume 40, 2008

Flows of Dense Granular Media


Yol Forterre and Olivier Pouliquen p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 1
Magnetohydrodynamic Turbulence at Low Magnetic Reynolds
Number
Bernard Knaepen and Ren Moreau p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 25
Numerical Simulation of Dense Gas-Solid Fluidized Beds:
A Multiscale Modeling Strategy
M.A. van der Hoef, M. van Sint Annaland, N.G. Deen, and J.A.M. Kuipers p p p p p p p 47
Tsunami Simulations
Galen R. Gisler p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 71
Sea Ice Rheology
Daniel L. Feltham p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 91
Control of Flow Over a Bluff Body
Haecheon Choi, Woo-Pyung Jeon, and Jinsung Kim p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p113
Effects of Wind on Plants
Emmanuel de Langre p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p141
Density Stratication, Turbulence, but How Much Mixing?
G.N. Ivey, K.B. Winters, and J.R. Koseff p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p169
Horizontal Convection
Graham O. Hughes and Ross W. Grifths p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p185
Some Applications of Magnetic Resonance Imaging in Fluid
Mechanics: Complex Flows and Complex Fluids
Daniel Bonn, Stephane Rodts, Maarten Groenink, Salima Rafa,
Noushine Shahidzadeh-Bonn, and Philippe Coussot p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p209
Mechanics and Prediction of Turbulent Drag Reduction with
Polymer Additives
Christopher M. White and M. Godfrey Mungal p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p235
High-Speed Imaging of Drops and Bubbles
S.T. Thoroddsen, T.G. Etoh, and K. Takehara p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p257

AR332-FM

ARI

22 November 2007

21:57

Oceanic Rogue Waves


Kristian Dysthe, Harald E. Krogstad, and Peter Mller p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p287
Transport and Deposition of Particles in Turbulent and Laminar Flow
Abhijit Guha p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p311
Modeling Primary Atomization
Mikhael Gorokhovski and Marcus Herrmann p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p343

Annu. Rev. Fluid Mech. 2008.40:113-139. Downloaded from www.annualreviews.org


Access provided by Indian Institute of Technology - Kanpur on 10/18/16. For personal use only.

Blood Flow in End-to-Side Anastomoses


Francis Loth, Paul F. Fischer, and Hisham S. Bassiouny p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p367
Applications of Acoustics and Cavitation to Noninvasive Therapy and
Drug Delivery
Constantin C. Coussios and Ronald A. Roy p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p395
Indexes
Subject Index p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p421
Cumulative Index of Contributing Authors, Volumes 140 p p p p p p p p p p p p p p p p p p p p p p p p p p431
Cumulative Index of Chapter Titles, Volumes 140 p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p439
Errata
An online log of corrections to Annual Review of Fluid Mechanics articles may be
found at http://uid.annualreviews.org/errata.shtml

vi

Contents

Anda mungkin juga menyukai