Vous êtes sur la page 1sur 59

Progress in Aerospace Sciences 36 (2000) 487}545

The control of #ow separation by periodic excitation


David Greenblatt*, Israel J. Wygnanski
Department of Fluid Mechanics and Heat Transfer, Faculty of Engineering, Tel Aviv University, Ramat Aviv 69978, Israel

Abstract

This paper presents a review of the control of #ow separation from solid surfaces by periodic excitation. The emphasis
is placed on experimentation relating to hydrodynamic excitation, although acoustic methods as well as traditional
boundary layer control, such as steady blowing and suction, are discussed in order to provide an appropriate historical
context for recent developments. The review examines some aspects of the excited plane mixing-layer and shows how its
development lays the foundation for a basic understanding of the problem. Flow attachment to, and separation from,
a de#ected #ap is then shown to be a paradigm for isolating controlling parameters as well as understanding the basic
mechanisms involved. Particular attention is paid to separation control on airfoils by considering controlling parameters
such as optimum reduced frequencies and excitation levels, performance enhancement, e$ciency, reduction of post-stall
unsteadiness, compressibility and other important features. Additional topics covered include excitation of separation
bubbles, control and exploitation of di!user #ows, three-dimensional e!ects, the in#uence of longitudinal curvature and
possible applications to unmanned air vehicles. The review closes with some recent developments in the control and
understanding of incompressible dynamic stall, speci"cally illustrating the control of dynamic stall on oscillating airfoils
and identifying the crucial time-scale disparity between dynamic stall and periodic excitation.  2000 Elsevier Science
Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 489
1.1. Flow separation: The problem. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 489
1.2. Traditional boundary layer control (BLC). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 490
1.3. Flow control by excitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 490
2. Objectives, scope and layout . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 491
3. Historical perspective*basic principles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 491
3.1. Boundary layer control (BLC). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 491
3.2. Airfoil separation control by acoustic excitation . . . . . . . . . . . . . . . . . . . . . . . . . 492
4. The Excited mixing-layer*basis for separation control . . . . . . . . . . . . . . . . . . . . . . . 493
5. Actuators and actuation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 495
5.1. Quantifying momentum addition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 495
5.2. Some aspects of actuator calibration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 496
6. Fundamentals of separation control*the de#ected #ap . . . . . . . . . . . . . . . . . . . . . . . 497
6.1. Governing parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 498
6.2. Basic mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 499
6.3. E!ect of net mass-#ux . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 500
7. Airfoil studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 502
7.1. Controlling factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 502
7.2. NACA 0015 experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 506

* Corresponding author. Tel.: #00972-3-640-8957; fax: #00972-3-640-7334.


E-mail address: greenb@eng.tau.ac.il (D. Greenblatt).

0376-0421/00/$ - see front matter  2000 Elsevier Science Ltd. All rights reserved.
PII: S 0 3 7 6 - 0 4 2 1 ( 0 0 ) 0 0 0 0 8 - 7
488 D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

Nomenclature dH boundary layer displacement thickness


* indicates the di!erence between two quantities
c airfoil chord
cycle phase
C drag coe$cient ("D/cq) g airfoil power e$ciency ("C /(C #C ))
" * " #
C form-drag coe$cient ("Dp/cq) j wavelength of eddies (or large coherent struc-
"N tures)
C skin-friction coe$cient ("q/cq)
D K sweep angle
C lift coe$cient ("/cq)
* h boundary layer momentum thickness
C maximum lift coe$cient
*  h> dimensionless momentum thickness ("f h/; )
C pitching moment coe$cient ("M/cq) 
+ o air density
C pressure coe$cient ("(p!p )/q)
.  k air dynamic viscosity
C input power coe$cient ("= /q; c)
# G  q dimensionless time ("t; /); skin friction
c steady momentum coe$cient ("J/cq) 
I
1c 2 oscillatory momentum coe$cient ("1J2)/cq)
I Abbreviations
C total (combined) moment coe$cient (c , 1c 2)
I I I AFC active #ow control
E allowable moment excursions
 DSV dynamic stall vortex
("C !C )
+  +  BLC boundary layer control
f excitation frequency
 FS #ap shoulder
f airfoil pitching frequency
f acoustic excitation frequency LCS large coherent structure
 LE leading edge
f natural shedding frequency
 MAV micro aerial vehicle
f> reduced forcing frequency ("f x/; )
 MEMS micro electrical-mechanical systems
F> reduced forcing frequency ("f X /; )
   PIV particle image velocimetry
FM "gure of merit ("g/(C /C ))
* " rms root mean square
G general step-height or slot-width length
g step height TE trailing edge
h slot width TS Tollmien-Schlichting
H boundary layer shape factor ("dH/h) UAV unmanned aerial vehicle
J steady jet momentum ("o;h) 1 2 rms quantity
H
1J2 oscillatory jet momentum ("ouh)
H Superscripts
k reduced airfoil frequency ("nf c/; )
 a rms of a
generic length scale a time-mean value of a
#ap length a
D phase-locked (averaged) instantaneous value a
Ma Mach Number
M #uid mass Subscripts
p local pressure 0 referring to baseline conditions
q free-stream dynamic pressure ("o; /2) 1,2 respective streams in a two-dimensional mix-

R (; !; )/(; #; ) ing layer
   
Re chord Reynolds number ("o; c/k) a #uid available for entrainment

St Strouhal number act relating to an actuator
t time; airfoil thickness att attachment
period of oscillation avg average value
u oscillatory velocity component b relating to steady blowing
; time-mean velocity component B based on bubble length
;> dimensionless velocity (";/; ) D based on cylinder diameter
O
; friction velocity ("(q/o) Df based on cylinder face
O
; LCS phase velocity e entrained #uid; excitation; boundary layer edge
(
x chordwise or surface-wise coordinate (from exc excursions
the origin) h based on step height
X normal distance from delta-wing LE to reat- i refers to upstream conditions

tachment line j refers to two-dimensional jet issuing from a slot
X recirculation bubble length f based in #ap length
X airfoil center of pressure max maximum value

X distance from slot or actuator to trailing edge mean time-mean value

y cross-stream normal coordinate min minimum value
y reference distance at ;"(; #; ) opt optimum value
    
or ;"; ro naturally occurring attachment
 
z spanwise coordinate r relating to attachment
a instantaneous incidence angle R relating to di!user ramp length
a mean incidence angle so naturally occurring separation
d, d #ap de#ection angle s relating to separation; static stall
D
d(x) boundary layer thickness (as a function of x) st relating to a step
d vorticity thickness R free-stream conditions
T
D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545 489

7.3. E$ciency considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 510


7.4. Performance enhancement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 511
7.5. Control of post-stall unsteadiness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 511
7.6. High-lift systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 512
7.7. Airfoil drag anomalies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 517
8. Di!user studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 517
9. Excitation of reattaching #ows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 518
9.1. Natural excitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 518
9.2. Imposed excitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 518
10. E!ects of curvature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 522
10.1. Maximum pressure recovery and concave curvature . . . . . . . . . . . . . . . . . . . . . . 522
10.2. Convex curvature*Coanda cylinder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 523
10.3. Constant curvature*the circular cylinder . . . . . . . . . . . . . . . . . . . . . . . . . . . . 523
11. Three-dimensional e!ects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 524
11.1. The swept back wing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 524
11.2. The delta wing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 525
11.3. Three-dimensional excitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 526
12. Engineering implementation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 527
13. Dynamic separation control. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 527
13.1. Unsteady separation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 527
13.2. Onset of dynamic stall. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 527
13.3. Forced dynamic attachment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 529
14. Control of airfoil dynamic stall . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 530
14.1. Experimental setup and objectives. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 530
14.2. Summary of important results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 531
15. Conclusions and future research . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 538
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 539
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 539

1. Introduction old tradition: the steady-yow assumption. Conventional


wisdom tells us that so-called steady two-dimensional
1.1. Flow separation: the problem separation is well understood, with a satisfactory analyti-
cal foundation (e.g. [4]). Experimental data reveals, how-
Flow separation is generally accepted to be the break- ever, that separation is only steady in a time-averaged
away or detachment of #uid from a solid surface (e.g. sense, while the rich time-dependent coherence is acknow-
[1}4]). Whether caused by a severe adverse pressure ledged, but it is certainly not exploited. For highly turbu-
gradient (e.g. [5,6]), a geometrical aberration (e.g. [7,8]) lent #ows which are on the verge of separation (e.g. [5,9]),
or by any other means, separation is generally accom- or known to have an underlying deterministic structure
panied by signi"cant thickening of the rotational #ow (e.g. turbo-machinery), the Reynolds-averaged equations
region adjacent to the surface with a marked increase in have limited the designer to basic assumptions which,
the velocity component that is normal to the surface. by their very de"nition, negate the phenomenon of time
Thus our traditional means for analysis, namely the dependence.
boundary layer equations, are summarily invalidated. Apart from the need to delay transition and control
Separation is almost always associated with losses of mixing, the familiar shapes of products which move
some kind, including loss of lift, drag increase, pressure through #uids or determine #uid motion are a direct
recovery losses, etc. Consequently, engineers have been consequence of the need to avoid separation. For
preoccupied, for almost a century, with altering its loca- example, the maximum thickness or camber of wings,
tion or avoiding it entirely. The multitude and variety of and the length and shape of di!users and inlets, evolved
hydro and aerodynamic vehicles and devices, taken for from this constraint whose norms are still dictated by the
granted today, bear testimony to the tremendous ad- steady-#ow assumption. For the better part of this cen-
vances that have been made in the development of means tury, methods for aircraft control have been rooted in the
that avoid or modify separation. limiting steady-#ow assumptions. A classical example is
Notwithstanding these advances, the development of wing warping initiated by the Wright brothers (see, e.g.
commercial and military vehicle aerodynamics has been [10]), which evolved into ailerons, #aps and slats prom-
severely stunted in recent decades. The underlying prob- inently displayed on today's aircraft (e.g. observe the
lem is a design philosophy which is steeped in a century multitude of moving surfaces on a Boeing 747 prior to
490 D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

landing). They are e!ective, but they are also complex, perturbations in a laminar boundary layer to trigger
costly, cumbersome and heavy. Consider a maneuverable a known instability, i.e. to initiate Tollmien-Schlichting
aircraft with no large moving parts or control surfaces, or (TS) waves. This breakthrough technique became a ma-
a single-rotor helicopter with no tail rotor or no heavy jor diagnostic aid for studying #ow stability but it was
NOTAR-type cylinder and compressor. Consider a coni- also considered a tool for controlling laminar separ-
cal di!user whose divergence angle exceeds 453, or thrust ation and transition (e.g. [18,19]). Since periodic per-
vectored jet without vectoring nozzles. All of these are turbations trigger a premature transition to turbulence
possible, by adding periodic motion to the #ow rather and a turbulent #ow is less susceptible to separation, #ow
than traditional boundary layer control (BLC) methods, separation could be delayed by initiating transition.
discussed below, which involve injecting or removing Sound was initially used to demonstrate these ideas on
large amounts of momentum and mass in a steady airfoils at low Reynolds numbers (see [20]). Although
fashion. In fact, periodic addition of momentum can laminar}turbulent transition could be e!ected, the
attain the same degree of control authority as is achieved manipulation of turbulent shear #ows was traditionally
by traditional BLC, with one important di!erence: the considered to be unattainable because of the belief that
cost, in terms of momentum input, is less by anything turbulence is a random process, determined largely by
from factors to orders of magnitude [11]. local #ow conditions. However, experiments such as
those of Brown and Roshko [21] and Winant and
1.2. Traditional boundary layer control (BLC) Browand [22] in the mixing layer demonstrated that
large coherent structures (LCSs) are primarily respon-
One year after the "rst powered #ight, Prandtl [12] sible for the transport of momentum across the #ow
introduced the concept of the boundary layer and pro- domain. Moreover, the introduction of periodic mechan-
posed a means (i.e. suction through a slot) to control its ical excitation at the #ow origin was shown to be an
attachment to a solid surface. The next ten years saw e$cient and convenient method for the control of mixing
little progress in boundary layer research [13], but the (e.g. [23}25]). Excitation accelerates and regulates the
race for air superiority, spawned universal research and generation of large coherent structures, particularly when
development e!orts in which BLC studies were given the mean #ow is unstable, thereby transferring high mo-
high a priority. The depth and breadth of international mentum #uid across the mixing layer.
research up to the early 1960s is apparent from the two The last decade has witnessed progress in #ow control
volumes edited by Lachmann [14,15], which provided where the fundamental studies of the 1970s and 1980s,
exhaustive treatment of theoretical, experimental as well such as those in free shear layers (see [26,27]), set the
as applied BLC methods. The contributions indicated scene for e!ective, e$cient and practical separation con-
that steady blowing or suction from various locations on trol. Speci"cally, it was shown that a turbulent mixing
the wing surface, and on various con"gurations, could layer could be attached to a de#ected surface [28,29] and
produce signi"cant increases in lift as well as reductions the ensuing attached #ow separation could be delayed
in drag. In fact, a large number of experimental aircraft [30,31] by periodic addition of momentum with or with-
were built, which convincingly demonstrated the e!ec- out the concomitant addition of mass #ux. The same
tiveness of BLC. Some of these aircraft reached mass concept was also applied to airfoils (e.g. [32,34]) and was
production (for example Lockheed's F-104, which re- further extended to other applications. This method was
mained in production from 1955 to 1983 and the MIG-21 shown to be much more e!ective than the traditional
of which thousands still #y in the former Eastern Bloc and steady blowing and at times attained a saving of two
in `Non-Aligneda countries), but BLC fell far short of the orders of magnitude in the momentum coe$cient re-
high expectations of the 1960s. There were two main quired to achieve a prescribed improvement in perfor-
reasons for this: "rstly, the plumbing systems required for mance. The actuators required may thus be autonomous,
BLC introduced excessive technical complexity and small, light, energy e$cient and decoupled from the main
weight, and secondly, the systems were not e$cient, re- propulsive systems } thereby overcoming all of the de"-
quiring auxiliary compressors or excessive compressor ciencies of traditional BLC.
bleed in order to obtain meaningful lift enhancements To date, separation control (acoustic or hydrodynamic)
(see, e.g. [16]). Some BLC systems, such as that used on has been demonstrated in a wide variety of relatively
the MIG-21, utilize engine combustion gases and rely on simple con"gurations, such as #ow over backward-facing
the thermal inertia of the plumbing to withstand the heat steps and ramps (e.g. [35,36]), on sharp leading-edge
for the short duration of application prior to landing. wedges [37], on blu! bodies (e.g. [38,39]), on various
airfoils (e.g. [40}42]), delta wings [43], circular cylinders
1.3. Flow control by excitation (see e.g. [44]) and behind two-dimensional fences (e.g.
[45]). These studies laid the groundwork for an extremely
The seeds for #ow control by excitation were sewn by wide variety of potential future applications to both "xed
Schubauer and Skramstad [17] who introduced periodic wing (e.g. [46]) as well as rotary wing (e.g. [47]) aircraft.
D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545 491

The emphasis of this paper is on experimental data The review also considers other `fundamentala con"g-
which either isolate parameters governing separation urations, such as excited separation bubbles over back-
control or elucidate underlying physical mechanisms. ward-facing steps and axially mounted blunt cylinders.
This paper does not cover other issues related to Important aspects of curvature are highlighted by con-
separation control, including theoretical (e.g. [48,49]) or sidering excitation in conjunction with cross-#ow over
computational prediction methods (e.g. [50}52,216,217]), circular cylinders, the excited `Stratford rampa and the
energy-e$cient micro-elecromechanical systems (MEMS in#uence of excitation on the wall jet #owing over a con-
} e.g. [53]) for actuation, `smarta or adaptive structures vex surface (the Coanda e!ect). Three-dimensional e!ects
(e.g. [54]), separation detection methods (e.g. [55]) and are considered by examining the e!ect of sweep and
control theory/neural-nets (e.g. [56,57]). This paper also delta-wing studies. Various applications, such as applica-
does not cover passive methods of separation control, tions of excitation to separation control on unmanned air
even those which may exploit #ow instabilities (see e.g. vehicles (UAVs), are also considered.
[4]). As the title suggests, the content of this paper is The last part of the review is dedicated to dynamic
limited to #ows where separation is controlled by excita- e!ects associated with excitation-controlled separation,
tion, i.e. an inherent instability is exploited to achieve the where once again the de#ected #ap is used as the guiding
desired results. Consequently, separation control by paradigm. An approach is adopted which emphasizes the
blowing and suction or actuation which does not clearly di!erence between the time-scales of #ows which separ-
exploit #ow instability (e.g. [58,59]), or similar methods ate or attach dynamically and time scales associated with
which create pseudo-surfaces (e.g. [47]), are not con- separation control. Furthermore, the e!ect of dynamic
sidered to be within the scope of this paper. stall control by excitation, on oscillating airfoils, is con-
In writing this article, the authors have endeavored to sidered.
incorporate enough background material for the non-
expert, while still maintaining adequate scope and depth
for practitioners of the art. The authors also wish to point 3. Historical perspective=basic principles
out that this review is unavoidably biased in favor of
their own work as well as that of their colleagues. As 3.1. Boundary layer control (BLC)
such, important studies pertaining to separation control
may have been unintentionally omitted. One positive Boundary layer (or #ow) control, as a means of preven-
aspect of this, however, is that relevant data, hitherto ting separation, has traditionally been associated with
languishing in Hebrew-language graduate theses, is now the injection of #uid to, or the removal of a #uid from,
available to the general reader. a boundary layer; or the motion of a surface in the
general direction of the stream (e.g. [60,61]). Excitation
can often be regarded as oscillatory addition of momentum
2. Objectives, scope and layout as opposed to steady addition of momentum. Moreover,
steady momentum or mass (blowing or suction) may be
The global objective of this paper is to review separ- superimposed on excitation and therefore a brief review
ation control by periodic excitation. The emphasis of the of the traditional BLC, from an excitation perspective, is
review is placed on hydrodynamic, as opposed to acous- appropriate here. Furthermore, for much of the data
tic, excitation methods since these have become domi- presented in this review oscillatory and steady addition
nant in recent years. However, acoustic excitation as well of momentum are compared in order to provide a quant-
as traditional active BLC methods, such as steady blow- itative assessment of excitation.
ing and suction, are initially discussed in order to provide Historically, suction was the "rst method ever pro-
an appropriate historical background and isolate those posed for the control of separation [12]. The basic
aspects that are important in the context of this review. principle is to remove decelerated #uid near a surface
The review continues with a summary of the excited and de#ect the high-momentum free-stream #uid to-
plane mixing layer, as well as some actuation methods, wards the surface. The remarkable e!ects of suction
and emphasizes those aspects which are important from were demonstrated on a variety of wind tunnel models
a separation control perspective. This is followed by an and even on experimental aircraft (e.g. [62,63]) but, to
examination of the generic #ap data, which illustrates date, the method has not been applied to the wings or
some of the basic principles of separation control and control surfaces of a production aircraft. Due to the
provides a context for the remainder of the review. This is mechanical complexity and additional weight, any
followed by aspects of actuators (excitation methods), aerodynamic gains made by suction are o!set by the
which relate to the content of the review. Since separ- power required to operate the suction device (e.g. [62]).
ation control on airfoils represents the area of most The e!ect of superimposing a small amount of suction
active recent research, a detailed review of these studies is on excitation, however, is considered in this review in
presented. Section 6.3.
492 D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

The purpose of steady blowing is to directly impart 3.2. Airfoil separation control by acoustic excitation
additional momentum to retarded #uid within the
boundary layer near the surface and thus delay separ- Until the 1960s, successful and practical approaches to
ation [61,64]. Blowing is usually achieved by an auxiliary separation control relied on the principle of direct mo-
blower or compressor bleed (e.g. [65]), but can also be mentum addition to the near-wall region, as discussed
achieved passively by diverting #uid from the free- in the previous section. In 1975, however, Collins and
stream to the surface. (Probably the most successful ap- Zelenevitz [20] demonstrated that acoustic excitation
plication of passive blowing is through the use of a slot can increase the momentum near the surface by transfer-
that allows air from the lower surface of an upstream ring it from the free-stream, without resorting to the
wing-element to pass over the upper surface of a down- traditional, direct injection of momentum. Following this
stream wing-element. Passive blowing in conjunction experiment, acoustic excitation was demonstrated on
with excitation for high-lift devices is considered in a variety of airfoils and experimental facilities, usually at
Section 7.6.1.) Steady blowing has been investigated low Reynolds Numbers. The technique involves intro-
for a variety of di!erent locations on wings but the most ducing sound into a wind-tunnel from the walls of the
successful application was that at the shoulder of a de- test section (e.g. [70,71]) by using speakers or acoustic
#ected #ap [16]. drivers. An example of how these disturbances are intro-
Poisson-Quinton [66] showed that separation control duced is provided by Zaman et al. [70] who showed
is governed by momentum (rather than mass) addition that excitation was e!ective whenever acoustic standing
and that a quantity, similar to the thrust coe$cient, was waves in the tunnel induced a transverse velocity com-
required in order to quantify the propulsive e!ect of ponent close to the airfoil surface. These investigations
blowing. Dimensional considerations gave rise to the provided some impetus for current advances because
so-called momentum coe$cient, c , which remains the they demonstrated that separation control can be
I
standard measure of relative momentum addition. In achieved by means of an externally supplied excitation
certain circumstances, particularly when the momentum source. (Studies involving `internal acoustic excitationa
coe$cient is `lowa, i.e. c (2%, blowing can be detri- are not quoted here since, within the framework of this
I
mental [16,67]. (Note that momentum coe$cient is article, these are considered hydrodynamic disturbances
quoted throughout in percentage terms for convenience (or oscillatory addition of momentum) and are cited in
when compared with excitation.) Thus many investiga- Section 7.)
tions of steady blowing present data for c '3%. This is A summary of representative investigations, with
I
primarily due to a local decrease in near-wall momentum important results, is presented in Table 1. The table
if the jet velocity does not exceed the local velocity illustrates both the important bene"cial results as
outside the boundary layer. Furthermore, Poisson-Quin- well as some drawbacks. Firstly, it can be seen that
ton and Lepage [68] showed that for c (5%, blowing separation control is not con"ned to a particular type
I
is most e!ective with the largest dC /dc being measured or class of airfoil. In fact, excitation can be bene"cial
* I
in this range. For larger c , boundary layer control is on a #at plate [74] as well as on a 17% thickness-to-
I
superseded by circulation control which is less e!ective, chord ratio airfoil such as Ahuja et al.'s [71] GA(W)-1.
i.e. although lift continues to increase, dC /dc is re- Furthermore, increases in C at a given incidence (but
* I *
duced. not C ) can be signi"cant } typically on the order
* 
BLC by means of surface motion generally involves of 50%.
moving the surface at or close to the free-stream velocity The drawbacks, however, outweigh these positive as-
and thereby eliminating the boundary layer. Such pects. Consider for example, the extensive investigations
methods are mostly used in wind tunnels, for studying of Ahuja et al. [71] and Zaman et al. [70]. Here, excita-
ground e!ect and aerodynamics of cars, but other than tion is facility dependent, since the acoustic drivers em-
that they are technological failures due to insurmount- ployed by Zaman et al. and Ahuja et al. e!ectively excited
able technical problems. Nevertheless, research on rotat- the respective wind tunnel's resonant modes, which in
ing cylinders located at blunt trailing edges (e.g. for truck turn excited the separating shear layers, thus leaving
applications, e.g. [69]) continues to date. In the context serious doubts concerning aircraft applications. This was
of excitation, surface motion generally applies to a #ip- re#ected, for example by signi"cant lift increases at some
eron, fence, ribbon or any other mechanical device incidence angles, with lift decrement at others. Moreover,
which oscillates in some manner, thereby introducing high levels of excitation were required, up to 156 db
oscillatory momentum into the #ow. in order to obtain some signi"cant e!ects, once again
calling into question the practical application of the
method. A further problem arises when attempting
 The term `#iperona is used to describe a small oscillating to compare the various investigations since there is no
#ap; the term &&#aperon'' generally refers to a combined #ap and uniform measurement location, with sound levels
aileron. being measured at the wall or on the airfoil surface.
D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545 493

Table 1
Summary of airfoil data for acoustic excitation

Investigator (s) Airfoil Re (;10) Optimum St Disturbance or a range Max. *C


*
range input level (deg) (%)

Collins and Zelenevitz [20] NACA 2412 0.53 24}90 100}134 dB on airfoil 20, 24 56
Ahuja et al. [71] GA(W)-1 0.6 4 156 dB on wall &18 40
Ahuja and Burrin [72] NACA 65 -213 0.25 2.7 147 on airfoil 16 50

Marchmann et al. [73] Wortmann 0.2 27 * 15 *
Zaman et al. [70] LRN(1)-1007 0.04}0.14 1}5 104 dB on wall )8 50
4}25 *18
Nishioka et al. [74] Flat Plate 0.04 3.8}27 Below airfoil 8}14 *
Zaman [75] LRN(1)-1007 0.075 2}5 u/; )3% near LE 18 50


Furthermore, the method has only been demonstrated 4. The excited mixing-layer=basis for separation control
for low Reynolds numbers, namely Re)600,000, and
therefore its e!ect on transition is intertwined with its The previous section indicated that excitation of
e!ect on separation. the separated shear layer was an important factor
An additional drawback is the large range of e!ective a!ecting separation control. In this connection, two im-
reduced frequencies St"f c/; which varies widely, by portant questions arise: Firstly, why should excitation of
 
two orders of magnitude, from O(1) [76] to O(100) [20]. the shear layer be important for understanding separ-
Certain investigations, however, indicate a pronounced ation control? Secondly, what are the reasons for the
shift to lower St when the excitation levels are high. For large range of e!ective excitation frequencies? To begin
example, Zaman [75] showed that when the excitation answering these questions, consider the statistical, time-
amplitude is increased, the optimum Strouhal number mean, description of the shear layer } where, in general,
can be orders of magnitude lower than that correspond- two streams (of velocity ; and ; , ; (; ) are pres-
   
ing to the linear inviscid stability of the separated shear ent (cf. [77,78]). Disregarding, for now, the existence of
#ow. He demonstrated that lower Strouhal numbers coherent motion in the mixing layer, the high degree of
2)St)5, combined with `largea disturbance ampli- shear at the discontinuity generates turbulence which
tudes (u/; +3%) resulted in the highest lift in- accelerates the slower (or stagnant) #uid, entraining it

crements. Furthermore, with large disturbance levels of into the mixing layer, while concurrently the faster
&156 dB, Ahuja and his co-workers [71,72] observed stream slows down.
that St+4. In addition, Nishioka et al. observed that the If we now consider the coherent motion, speci"cally
maximum rms associated with the vorticies generated by when subjected to excitation, (e.g. [23,26,27,79,209]), it is
excitation were four times larger for St"1.9 than for widely accepted that large quasi-deterministic, vortical,
St"30 when `stronga acoustic excitation was employed. spanwise structures are the essential `building blocksa of
The above observations indicate a clear trend towards the mixing layer and are responsible for virtually all of
lower e!ective frequencies St&O(1) at high excitation the momentum transfer across its extent. It has been
levels. clearly shown that the spreading rate, and hence the
Despite the above-mentioned drawbacks, acoustic ex- entrainment, can be signi"cantly altered by two-dimen-
citation illustrated the e!ectiveness of excitation per se sional, small amplitude (large, in terms of acoustic excita-
and the principle that separation delay is e!ective at tion) excitation applied at the origin of the shear layer
lower frequencies when excitation amplitudes are (see, e.g. [23]). For example, observe the smoke #ow
large. These observations form an important prerequisite visualization patterns in a mixing layer that is excited at
for the discussion that follows. It is also important to two di!erent frequencies in Figs. 1a and b.
note that excitation amplitudes for acoustic excitation Consider now the non-dimensionalized development
are typically u/; )3% (near separation), while rela- of a mixing layer presented in Fig. 2, which is periodically

tive steady blowing jet velocities are ; /; *200% excited at its origin by small amplitude perturbations,
H 
for typical blowing-slot widths. The interval between and is divided into three regions (see [23]). The portion of
these two limits is enormous but remains impractical in the mixing layer downstream of the excitation location
the context of either steady blowing or acoustic excita-
tion. This paper will show, however, that it is in this
`middle grounda that hydrodynamic excitation is most  More detailed answers to these questions are presented in
e!ective. Sections 6 and 7.
494 D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

onset is associated with `rollupa of instability waves into


discrete vorticies and if lower frequency waves are also
present in the #ow, vortex merging may be the dominant
factor further downstream. For this example, region II
extends over the approximate range 1(Rf >(2, where
f >"f x/;M and R"(; !; )/(; #; ). Beyond
    
Rf >+2, the mixing layer continues to spread linearly
(region III), although the spreading rate is somewhat
lower than in the linearly unstable region (region I).
The above description relates the frequency of the
imposed excitation to the spreading-rate of the mixing
layer, and indicates its e!ect on the momentum transfer
across it. A large spreading rate implies that a larger
volume of #uid is entrained from the slower (or stagnant)
#uid. As a simpli"ed example to illustrate the rami"ca-
tions for separation control, assume that only one stream
is present, say the top stream (; ), and that a `virtual

#apa is de#ected below the mixing layer (see Fig. 3). The
premise adopted here is that the #ow will `attacha to the
Fig. 1. Flow visualization of a two-dimensional mixing layer virtual #ap when the mass of #uid entrained into the
excited at two di!erent frequencies [(a) 40 Hz and (b) 80 Hz)] mixing layer (M ) exceeds that available for entrainment

showing large spanwise vortical structures which are responsible (M ). From Fig. 3 it can easily be shown that
for large-scale momentum transfer. Ratio of upper to lower


stream velocity: ; /; "0.4 [23]. 1 V
  M /M J d(x) dx (4.1)
 x
 
over an arbitrary streamwise distance x . Now, since

h(x)Jd(x), h>,f h(x)/; and f >,f x/; , equation
     
(4.1) can be written as


1 D>
M /M J h>( f >) df > (4.2)
 f >
 
where f >,f x /U . Numerical integration of the data
    
presented in Fig. 2, are shown in Fig. 4 and indicate that
the optimum momentum transfer is attained when
( f >) "O(1), where small variations in excitation am-
 
plitude are assumed to be responsible for the experi-
mental scatter. Notwithstanding the scatter, Fig. 4 shows
that the shear layer neutral stability is a crucial factor in
Fig. 2. Dimensionless spreading rate of the actively excited establishing the optimum reduced frequency.
mixing layer showing the division into three regions (adapted It should be noted, however, that higher excitation
from [27]). levels cause an overshoot of h> in region II, followed by
a contraction to a lower value, thus resulting in a higher
( f >) (e.g. [80]). Furthermore, halving the excitation
 
amplitude can also result in a higher ( f > ) (e.g.
 
(region I of Fig. 2) diverges, principally as a result of the [215,27]). Therefore, without precise quanti"cation of the
ampli"cation of quasi two-dimensional waves, enhanced disturbance amplitude (or periodic momentum addition),
by a Kelvin}Helmholz instability mechanism (e.g. determination of ( f >) can only be expected to yield an
 
[23,80]). At some point, the mixing layer becomes neu- order of magnitude approximation. It is therefore ex-
trally stable to the imposed disturbances and if larger pected * and indeed will be shown * that the same
scales are not present, its growth ceases over a signi"cant principle applies to attaching a free shear layer to a real
streamwise extent. This occurs at a momentum thickness- surface. Thus the above discussion partially answers the
based Strouhal number h>"hf /;M +0.075, where questions posed at the beginning of this section, while in

;M "(; #; )/2 and is referred to as region II. The Section 6 and the remaining sections, the practical im-
 
location underlying the onset of neutral stability depends plications of the above mentioned arguments will be-
upon the frequency and amplitude of excitation. This come apparent within the context of separation control.
D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545 495

Fig. 3. Mean-#ow mixing layer schematic for illustrating the mass ratio of #uid entrained (M ) to #uid available for entrainment (M ).


speakers/acoustic drivers, [81], pistons, oscillatory-#ow


valve-systems [82] or piezoelectric-based diaphragms
(see [47]), etc. When no net mass-#ux is superimposed
on these #ows, they are sometimes referred to as syn-
thetic-jets (e.g. [83]) or Zero Mass Blowing jets (ZMBs).
Furthermore, periodic #ow disturbances can be achieved
by surface mounted mechanical actuators such as rib-
bons (e.g. [33]), #iperons [81], piezo-based benders [84]
etc., where the common denominator in all cases is
the oscillatory addition of momentum to the #ow. It
should be noted that the term `microa, as in Micro-
electromechanical System (MEMS), generally `means
small compared to the scale of the overall #ow to be
controlleda [85] and thus includes the devices referred to
above.
Fig. 4. Numerical integration of the data presented in Fig.
2 according to Eq. (4.2). 5.1. Quantifying momentum addition

An attempt has been made in this paper to standardize


5. Actuators and actuation the quanti"cation of momentum addition as much as
possible. In concordance with historical and practical
This paper is concerned primarily with the e!ects of considerations (see Section 4.1), the oscillatory mo-
actuation and does not consider the speci"c character- mentum coe$cient is quoted or inferred wherever pos-
istics of individual actuators or actuation techniques. sible. For a 2-D con"guration, we de"ne the total (mean
However, some techniques are considered in as much as and oscillatory) momentum coe$cient as a ratio of the
they apply to the objectives of this paper. In this context, momentum added, to that in the free-stream, namely
actuators are considered to be devices that interact with (see [66]):
the #ow hydrodynamically to produce oscillatory addi-
tion of momentum, with or without superimposed mass o ;G
C "

(5.1)
#ux. Thus, externally mounted acoustic drivers, used for I 1/2o ;
acoustic excitation (e.g. [70,71]) perturb the entire `free  
streama in the wind tunnel and are further enhanced in where j refers to a two-dimensional jet, G is either the slot
the regions of concentrated vorticity (such as the tunnel width h or step-height g, and is usually taken as the
wall boundary layers) that are receptive to the imposed length of the body under consideration (e.g. chord c, or
excitation. Therefore, excitation of the ambient #ow does #ap length ).
D
not fall into this category. The jet velocity in this instance contains both mean
Periodic #ow oscillation can be achieved from two- and oscillatory components: ; ";M #u . Substituting


dimensional slots by means of internally mounted the mean and oscillatory components into Eq. (5.1),
496 D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

considering incompressible #uid, and time averaging, was provided by an acoustic speaker, as well as a #iperon
gives to delay #ow separation from a de#ected surface
(described in Section 6). Figs. 5a and b show the results of

   
2G ;M  2G u  excitation by both methods for 1c 2"0.02 and 0.01%,
C "
#
(5.2) I
I ; ; respectively, on the #ow separation delay angle d . The
  Q
"gures clearly demonstrate the appropriateness of the
where u is the mean square of the oscillatory jet velocity

quantity 1c 2 for quantifying oscillatory momentum
component. When steady and oscillatory momentum are I
addition as well as the equivalent net e!ect of di!erent
added simultaneously, the momentum coe$cient is ex-
actuation methods.
pressed in the shorthand form: C "(c , 1c 2), where the
I I I It should be noted that for all airfoil data presented, for
term enclosed in parentheses refers to the two right-hand
example in Section 7, C is always based on the chord of
side terms in Eq. (5.2) and the overbar is dropped for I
the airfoil. This has been done for historical reasons, but
convenience. In this paper, when no confusion arises (i.e.
may not be the most accurate measure of the e!ectiveness
we employ either steady or oscillatory addition of mo-
of excitation. For example, it may be more appropriate to
mentum), we simply use the term C .
I base C on the length measured downstream of the
In order to remain consistent with the above de"ni- I
excitation location, i.e. X , much like the de"nition of
tion, when no slot is present and excitation is achieved by 
F> (see Section 6).
a mechanical means, the oscillatory momentum coe$c-
ient is de"ned by


1  5.2. Some aspects of actuator calibration
1c 2" o u dy (5.3)
I 1/2o ; 
   The description and calibration principles presented
where u is the #uctuating component of velocity adjacent above are necessarily simpli"ed to be consistent with the
to, and downstream of, the actuator. Although basic objectives of this review. Within the ambit of #ow and
design di!erences exist between two di!erent methods of separation control, however, a wide variety of di!erent
actuation, from an excitation viewpoint we regard their actuator types have been proposed, where a simple calib-
net e!ect on the #ow at a given frequency to be virtually ration based on the above arguments is not straightfor-
identical. An example of this is provided by Nishri [81], ward (e.g. [86]). As an example, consider the schematics
who used both a two-dimensional slot, where excitation of two relatively simple surface-mounted actuator-types

Fig. 5. A comparison of two di!erent methods of excitation (periodic slot blowing/suction by means of an acoustic speaker and
excitation by means of a #iperon) on the maximum angle that a #ap can be de#ected without separating, (a) 1c 2"0.02%, (b)
I
1c 2"0.01% (adapted from [81]).
I
D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545 497

Fig. 6. An example of surface mounted elements used for separation control by excitation (a) two-dimensional surface-mounted #iperon
(or ribbon), (e.g. [33,87,88]); (b) similar to (a) but with a cavity below the actuator; (c) top view of multiple-element three-dimensional
actuation [84].

presented in Figs. 6a and b. Fig. 6a shows the side view of actuator that are adjacent to one another (three such
a two-dimensional surface-mounted #iperon (or ribbon), actuators are shown in Fig. 6c). The actuators were then
which is hinged on one side and oscillates about the hinge operated in anti-phase (1803 out-of-phase), e.g. when the
between the wall and the stream (e.g. [33,87}89]). As the center actuator shown in Fig. 6c was at its peak upper
actuator moves downwards it expels the #uid between it position, the two adjacent actuators were at their lowest
and the surface and generates a transient #ow that is position in the cavity, and vice versa. The authors refer-
similar to a wall jet. As the actuator moves upwards it red to this as a three-dimensional mode of excitation and
generates a positive (clockwise) vortex which is shed for purposes of calibration, this introduced additional
downstream at some point during the downward motion complexity. Data relating to separation control using
[87]. Thus there are at least two mechanisms by which these actuators in the two-dimensional (in-phase) and
this simple device can add momentum to the #ow. In three-dimensional (anti-phase) modes is presented in
reality, additional factors become important such as the Section 7.3.
height and nature of the upstream boundary layer rela-
tive to the actuator dimensions and bending modes of the
actuator itself. Thus the calibration and characterization 6. Fundamentals of separation control 0 the de6ected
of the #ow-"eld generated by this relatively simple device 6ap
is a challenge in itself.
The schematic con"guration presented in Fig. 6b has When considering the basic principles of separation
an additional complexity due to the existence of a cavity control, a question bound to arise is: Why use a straight
into which the actuator descends (see [84]). In this in- de#ected #ap as the quintessential model for demonstrat-
stance, as the actuator descends into the cavity, the cavity ing separation control? There are other important gen-
#uid is compressed and a jet issues from between the eral aerodynamic shapes, which have been extensively
actuator tip and the cavity trailing-edge (TE). Seifert et al. studied, such as certain airfoils or cylinders, which would
[84] introduced a further innovation by replacing the appear to take precedence. There are a number of com-
two-dimensional actuator with an array of "nite-span pelling reasons for this: Firstly, the de#ected #ap is the
498 D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

ubiquitous time-honored high-lift device of choice. Sec-


ondly, attachment of otherwise separated #ow to a #ap,
not only has profound impact on lift, but is the most
technologically successful form of BLC to date. Thirdly,
the general problem being considered is complex, with
many independent variables and no general theoretical
model. Thus the elimination of e!ects such as curvature
as well as uncertain upstream conditions (laminar,
transitional or turbulent) and moving separation loca-
tions, such as occur on airfoils, are vital. With a geomet-
rically simple con"guration, one can isolate the most
important governing parameters. Fourthly, for a basic
Fig. 7. Schematic representation of the generic #ap and shoul-
understanding, it is essential to avoid complications asso- der region (cf. [81]).
ciated with three-dimensional e!ects such as those on
a delta wing [90,91]. Fifthly, the control of separation
over a de#ected #ap illustrates the important di!erences
between forced attachment of a separated #ow and separ-
ation delay, which are entirely di!erent phenomena and
require di!erent ranges of dimensionless parameters for
e!ective control. For `simplea con"gurations such as
#ow separation from a backward-facing step (e.g. [35])
the location of separation is established by the geometry
and thus periodic excitation deals with the issue of reat-
tachment only. Similarly, forced separation on a #at sur-
face (e.g. [59]) will always involve a #ow that has already
separated and then reattaches further downstream.
A sixth, and "nal, point is that the #ap allows detailed
study of dynamic attachment as well as separation (dy-
namic stall) for a geometrically static scenario with
a minimum of additional parameters. It has recently been
Fig. 8. Minimum C required to attach an initially separated
shown that this is critical for understanding the prin- I
#ow to a de#ected #ap, as a function of F> for various values of
ciples of dynamic stall control [181], and is discussed
*d (from [31,81]).
further in Section 13. 

6.1. Governing parameters occurs in the absence of periodic excitation) or to d , the



angle at which attachment occurs. The dimensionless
Consider a generic-#ap con"guration (Fig. 7), where frequency: F>,f X /; , where X is the distance
   
the upstream boundary layer is fully turbulent and dis- between the excitation location and #ap trailing edge.
turbances are introduced either by two-dimensional jet Thus for the current #ap con"guration X , . As in
 D
excitation or a #iperon (see [31,81]). Dimensional analy- Section 5, total momentum coe$cient above is desig-
sis shows that dependent variables, such as bubble-length nated: C ,(c , 1c 2), and at times is used to denote
I I I
X or C describing the state of the #ow over the #ap, are either steady momentum c or periodic excitation 1c 2.
. I I
primarily functions of the variables: With the #ow (shear layer) initially detached from the
#ap, the introduction of two dimensional, periodic oscil-
*d, x/ , F>, C lations increase entrainment from the `dead-air regiona
D I
above the #ap and ultimately lead to #ow attachment.
The parameter *d describes a de#ection that exceeds the
Fig. 8 shows the minimum C required to attach the #ow
angles at which separation or reattachment occur nat- I
as a function of F> [31,81]. The "gure clearly indicates
urally, i.e.: *d "(d !d ), *d "(d !d ). In this way
      that that F>+1 is the optimum reduced frequency and
the initial condition of the #ow is accounted for, since d
that this result is Reynolds number independent. This
is either referred to d (the angle at which separation
 data ties in with the discussion presented in Section 4,
and shows that the determination of an optimum F> can
 Note that in this paper, attachment is used in the context of only be considered accurate if it is associated with the
a #ow which was initially separated; a reattachment location minimum excitation amplitude, Ck with that of the

implies an upstream separation location with a bubble or recir- excited shear layer presented in Section 4 (cf. 1)
culation region between the two. ( f >) )1.4) and indicates, in a clear and convincing
 
D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545 499

manner, how entrainment and shear layer stability a!ect indicative of this situation, where there is an almost
#ow attachment. constant pressure region extending up to 30% of the #ap
After attaching, the #ow encloses a large bubble whose length, followed by an increase in pressure up to the
size may be reduced by increasing F> or C or both (this trailing edge. Flow-"eld measurements with a PIV, show
I
will be discussed further in Section 9.2 in the context time-mean vorticity contours (Fig. 9b) and the mean
of unconditionally attached bubbles). Determination of streamline tM "0 indicated that the bubble dimensions
the minimum C for separation prevention can be made vary between 40}55% of depending on the value of
I D
in the same manner as that for e!ecting attachment, C used for the prescribed (d!d ). The #ow can be
I 
and is centered around the dimensionless frequencies maintained in this state as long as the C supplied pro-
I
2&(F>:4. At high excitation frequencies (F>94) the vides a su$cient margin of safety to prevent separation
#ow separates from the trailing edge since the Large by a bubble-bursting mechanism. Since the #ow bound-
Coherent Structures (LCSs) dissipate before reaching the ing the bubble is similar to the familiar mixing layer
end of the #ap. Hysteresis between attached and separ- (at least over the initial, constant pressure region), an
ated conditions is typical to this bi-stable #ow and it may increase in the excitation frequency shortens X for two
be induced not only by changing d, but also by changing reasons. Firstly, the streamwise distance between success-
some of the other parameters listed above such as F> and ive rolled-up vortices decreases and secondly these struc-
C . In general the ratio between the 1c 2 required for tures amplify over a shorter distance from the actuator or
I I P
attachment and the 1c 2 needed to prevent separation slot and thus the ampli"cation terminates closer up-
I 
may be as large as an order of magnitude. stream than for those structures generated at lower fre-
quencies. Since dissipation takes over further down-
6.2. Basic mechanisms stream, the perturbations may decay below the threshold
required for keeping the #ow attached. As a consequence,
Upon reattachment, due to periodic excitation at there is a limit to separation control by a simple increase
F>&O(1), the mean bubble dimension is commensurate in the excitation frequency. Since the e!ectiveness of the
with the length of the #ap and a similar bubble may exist momentum transfer mechanism diminishes over the in-
just prior to complete separation. The time-averaged creasing distance between the reattachment location and
pressure distribution shown in Fig. 9a for C "0.01% is the trailing edge of the #ap, separation reoccurs when the
I
local disturbance amplitude falls below the threshold
level required for that de#ection angle (provided d'd ).

Under these circumstances, separation commences near
the trailing edge where the amplitudes of the imposed
perturbations have dissipated.
The pressure distribution over the #ap, when the am-
plitude of the imposed perturbations was decreased to

Fig. 9. (a) Time-mean pressure distribution and (b) time mean


vorticity contours on the de#ected #ap, showing the approxim- Fig. 10. Time-mean pressure distribution on the #ap at two
ate bubble size (from [33,93]). reduced excitation frequencies (from [31,81]).
500 D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

Fig. 12. Phase-locked pressure distributions on the #ap at vari-


ous stages of the excitation cycle (from [33,93]). Same conditions
as Fig. 9.

ence. Insight into this process, which is independent of


Reynolds number, is provided by phase-locked PIV vor-
ticity measurements (Fig. 11) [92] and phase-locked sur-
face pressures (Fig. 12). Fig. 11 shows the high degree of
structural coherence near the origin of excitation as well
as the relatively slow moving structures in this region.
Further downstream, the structures dissipate due to tur-
bulent di!usion with a concomitant increase in their
phase velocities. Thus the apparent time-mean recircula-
tion bubble is, for the most part, a manifestation of these
traveling structures. These observations are re#ected by
phase-locked pressure data which indicate the progress-
Fig. 11. Phase-locked vorticity contours on the #ap at various ively increasing wavelength with distance along the
stages of the excitation cycle (from [33,93]). Same conditions as #ap (see the successively increasing wavelength j with
Fig. 9. distance downstream in Fig. 12). The high degree of
vorticity coherence is associated with relatively large
the minimum level necessary to keep the #ow attached surface pressure excursions and the increasing phase-
(i.e. to C +(10\%), is very sensitive to F>. Two limit- velocity j ) f .
I 
ing cases are shown in Fig. 10. At F>"0.7, the reduction
in amplitude generated a mean bubble whose dimensions 6.3. Ewect of net mass-yux
were commensurate with the length of the #ap. Thus,
a small additional increase in the bubble length caused it Consider an initially separated #ow over the de#ected
to burst. In this case the #ow is already separated over #ap, under the conditions Re"0.45;10 and
most of the #ap in spite of the large normal force it (d !d )"1.53. Initiating steady blowing, at c "0.1%,
  I
generates. At F>"7.5 the size of the bubble was insigni- did not attach the #ow to the #ap (Fig. 13). Attachment
"cant but the boundary layer at the trailing edge was was only achieved when c "0.18% and enclosed a large
I
thick and depleted of momentum. Thus a small reduction bubble that extended over most of the #ap. The bubble
in C caused separation which propagated upstream was still present at c "0.4% (see Fig. 13) in the absence
I I
from the trailing edge. of oscillatory input. Gradually reducing the blowing
Although the resulting time-mean #ow-"eld is of pri- intensity resulted in total separation when c "0.18%.
I
mary practical importance, it belies the generation and Flow attachment by periodic excitation required only
evolution of the LCSs, which are its very basis for exist- 1c 2+0.012% at a reduced frequency of F>"1.5 while
I
D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545 501

Fig. 13. Surface pressures for steady blowing, excitation and


a combination of the two (from [81]).

separation was prevented at 1c 2+0.002% at F>"3.


I
These were close to the optimal frequencies for the
respective conditions (see [31]), indicating that, for
blowing, attachment required 15 times more momentum
input while 90 times more momentum was required for
separation prevention.
A combination of steady and oscillatory blowing was
not very e!ective in this case either, because the slot was
located where separation occurred naturally and there-
fore there was no need to advect the imposed oscillations
further downstream. For example, oscillatory amplitude
of 1c 2"0.02% generated a bubble that covered less
I
than 40% of the #ap chord (according to the criterion
discussed by Nishri and Wygnanski [31]), and had Fig. 14. Time-mean vorticity contours, illustrating the di!er-
a minimum C "!0.8 (Fig. 13). The steady c of 0.4% ence between excitation and excitation superimposed with
. I steady blowing (from [33,93]).
was inferior, since it contained a bubble whose length
was 60% of the chord. A combination of blowing and
excitation at a combined C ,(c , 1c 2)"(0.1, 0.02)% ately 50% lower than in the absence of steady blowing
I I I
was inferior to pure excitation at 1c 2"0.02%. A com- (see [31,92]).
I
bined C "(0.18, 0.02)% was required to achieve, in this Detailed PIV measurements were carried out using
I
case, the same results as excitation alone. both steady suction or blowing that was superimposed
The mean vorticity distribution over the #ap, mea- on the periodic excitation at F>"1.2 and
sured by PIV, with and without steady blowing is shown 1c 2"0.02%. The steady #ow passing through the slot
I
in Figs. 14a and b, respectively. The closed contours was measured in terms of c , where net suction through
I
marked by dotted lines near the surface represent nega- the slot is denoted as !c . The separating streamline
I
tive vorticity indicating the presence of reversed #ow, bounding the bubble, calculated from the PIV data for
while the rest represent positive vorticity. In the absence various values of !0.16%*c *0.16%, is plotted in
I
of steady blowing, the mean #ow reattaches to the surface Fig. 15. The "gure covers only 35% of the #ap length for
around 40% of the #ap length [this was also deduced on the sake of clarity. In the absence of steady mass #ow
the basis of the pressure distribution (Fig. 13)]. The though the slot the mean bubble boundary is initiated at
positive vorticity downstream of the mean reattachment x/ +8% while the addition of blowing not only in-
D
position indicates a high level of skin friction, suggesting creases the distance at which the bubble forms, but also
that separation is not imminent. The addition of steady increases its height relative to the #ap. Suction, on the
blowing at a combined C "(0.16, 0.02)% resulted in other hand, initiates the bubble formation closer to the
I
a longer and thicker average bubble that extended over slot and shortens its length. With c "!0.16%, the
I
80% of the #ap and a weak positive vorticity region bubble was initiated at x/ +4% and it reattached at
D
downstream of reattachment. In addition, the maximum x/ +32%. One should remember however that this is
D
vorticity contour in the vicinity of the slot was approxim- not a steady #ow (see Figs. 11 and 12) and therefore the
502 D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

momentum, was suggested by Viets et al. [94] and


Huang et al. [32] and demonstrated on speci"c airfoil
pro"les by Neuburger and Wygnanski [34] (also see
[95]). Numerous additional studies followed, e.g. Bar-
Sever [96], Hsiao et al. [97], Shepshelovich and Koss
[98], Chang [99], Seifert et al. [40}42], Hsiao et al.
[100], Darabi [101], Seifert et al. [67], Tinapp et al.
[102], Smith et al. [103], Seifert and Pack [104], Seifert
et al. [84], Naveh et al. [105] and others (see Table 2).
It will be shown in this section that, in contradistinc-
tion to acoustic excitation, these investigations are facil-
ity independent and they can signi"cantly enhance airfoil
Fig. 15. E!ect of steady blowing or suction, superimposed on
performance at relatively modest expense. The e!ective
excitation, on bubble dimensions. Prevailing conditions:
reduced frequencies, for the vast majority of investiga-
d!d "63, F>"1.2 and 1c 2"0.02% (from [33,93]).
 I tions, occupy a relatively small range. Moreover, much of
the data will be compared with steady blowing, in order
mean streamlines represent boundaries that oscillate sub- to illustrate the relative e$ciency of the method. In recent
stantially around the average. For example in the ab- years, a disproportionately large emphasis has been
sence of steady blowing and at one phase angle relative to placed on NACA 0015 airfoil studies, and thus a signi"-
the excitation (
"03 in Fig. 11), the bubble is not visible cant fraction of this section will review selected represen-
at x/ (30% while at
"903 it is initiated at the slot. tative data. This section will also deal with a host of
D
At still another
"2703, the reverse #ow region is in- additional aspects that relate to airfoil separation con-
itiated at x/ +20% and it terminates at x/ (40%. trol, including high-lift systems, high Re and high sub-
D D
In short, the travelling reverse #ow region near the solid sonic Mach numbers (Ma) #ows and reduction of un-
surface results from the clockwise eddies that are initiated steadiness associated with separation.
by the #apping motion of the mixing layer originating at
the discontinuity in the surface (the leading edge of the 7.1. Controlling factors
#ap). These eddies feed on the spanwise vorticity avail-
able in the #ow and may induce a reverse #ow region The two most important questions that arise when
near the surface if their circulation is su$ciently strong. surveying a cross-section of experimental data are (a)
A larger array of such eddies, resulting from higher which investigations yield the most signi"cant airfoil per-
F> and 1c 2, may generate a net skin-friction thrust as formance improvements and (b) what are the primary
I
a consequence of averaging the mostly negative skin controlling factors or parameters. Table 2 distils this in-
friction. This idea is examined further in the context of formation by considering airfoil type; Reynolds number;
airfoils (see Sections 7.7 and 14.2.1), where periodic exci- reduced frequency of excitation; disturbance input-level
tation resulted in form drag exceeding the total drag. and location as well as lift enhancement (relative to the
In the context of #apped airfoils, pulsed blowing baseline maximum and at a speci"c post-stall angle). Note
(i.e. a combination of steady blowing and excitation the di!erence between *C , and *C at a given angle,
*  *
discussed above) was observed to signi"cantly enhance which are de"ned in Fig. 16. Also, in order to maintain
airfoil performance (e.g. [67]). It should be appreciated, uniformity, the original data from the various investiga-
however, that the combined momentum coe$cients tions has been cast in terms of the dimensionless frequency
used were approximately 10 times larger than those F> (,f X /; ) and where possible, in terms of 1c 2.
   I
considered here (for more details see Section 7.2). The table reveals a number of important statistics. As
with acoustic excitation, studies have been performed on
a wide variety of airfoils, although here the vast majority
7. Airfoil studies have been relatively thick (t/c515%). Also, with the
exception of Seifert and Pack [104,107], all investigations
It was pointed out in Section 4.2 that, historically, have been performed at low Reynolds numbers:
acoustic excitation set the precedent for latter demon- 150,000)Re)900,000. For most investigations, excita-
strations of airfoil separation delay by employing hy- tion is applied close to the leading-edge at x/c(10%.
drodynamic excitation. Problems such as facility depend- Notable exceptions are those of Neuburger and Wyg-
ence, high acoustic excitation levels, and a bewilderingly nanski [34,108] and Seifert et al. [40,67,84].
large e!ective reduced frequency range (i.e. from O(1) to
O(100), see Section 3.2) added impetus to seek a di!erent 7.1.1. Reduced frequency ewect
approach to excitation. The method employing hy- When compared to the relatively simple de#ected #ap,
drodynamic disturbances, or oscillatory addition of separation control over airfoils presents additional
D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545 503

Table 2
Representative summary of airfoil data: separation delay by periodic excitation

Investigator (s) Airfoil Re (;10) E!ective Disturbance or Dist. x/c *C Post-stall


* 
F>,f X /; input level (%) (%) *C (%)
   *
Viets et al. [94] * 0.1 0.4, 0.5 Rotating mechanical 19.5 * *
element

Huang et al. [32] * 0.034 0.94 115 dB at slot exit 15 30 140


Neuberger and  NACA 0015 0.2}0.4 0.4}2.6 Vibrating ribbon: 10}20 42 86

Wygnanski [34] Wortmann 0.2 0.3}1.8 u/; "0.5}5% 10}60 9.4 70

Neuberger [95] FX63-137 * *
Bar Sever [96] LRN(1)-1010 0.15 0.7}2.7 Oscillating wire, !1 12 39
amplitude a function
of frequency
Hsiao et al. [97] NACA 63 -018 0.3 0.1}10 95 dB at slot exit 1.25 3 47

Chang et al. [99] NACA 63 -018 0.3 1.4 Slot C "0.02% 1.25 22 83
 I
0.3 4.0 Slot C "0.01% 1.25 22 83
I
Seifert et al. [40] IAI P255 0.2 2 Flaperon 8 8.4 140
NACA 0015 0.15 2 C "(0.8,0.8)% 75 63 68
I
(w/#ap)
Zhou et al. [37] Back-to-front 0.66 1 Oscillating #ap: 0 * 70
airfoil 0.05%}0.1%c peak-
to-peak oscillation
Hsiao et al. [100] NACA 63 -018 0.31 0.35}0.5 1.25 30

*
(high alpha)
Seifert et al. [67] NACA 0015 0.3 0.6 C "0.1% 0 21 92
I
Also see Darabi (#apped airfoils) 0.3}0.9 2 C "(3.2, 2.7)% 0 63 90
I
[101] PR8-40 (IAI) 0.14 1.75 C "0.015% 70 40 50
I
E214F 0.2 0.57 C "0.3% 70 10 10
I
SPCA-1 0.15 2 C "(0.8, 0.8)% 75 60 110
I
Erk et al. [106] Wortmann FX 0.25 54}81 160 dB at slot 2 * 40
61-184
Smith et al. [103] Cylinder#NACA 0.3 10.1 C "0.17%, double- 6 120 240
I
aft-portion slot at center 20%
span
Seifert and Pack NACA0015 (LE) 31 0.7}2.1 C "(0.27, 0.05)% 10 15 47
I
[104] NACA0015 (FS) C "0.05% 70
I
* * * *
Seifert et al. [84] PR8-40-SE (IAI) 0.5 1.2 Piezo elements 41 22 36
u/; "20%

Naveh et al. [105] NACA 0015 0.16 0.45 C "0.1% 66 30 50
I
Amitay et al. [44] Cylinder with 0.31 10, 15, 20 C "0.18% 6 120 240
I
NACA Aft-section

complications since, in general, the separation point is frequency F> is achieved for the minimum addition of
not "xed, curvature e!ects can be signi"cant and the momentum. It is the opinion of these authors that,
upstream conditions are not well de"ned (i.e. laminar, for a given airfoil, this is a prerequisite for establishing
transitional or turbulent). In some investigations, excita- F> .

tion levels (or amplitudes) are not reported and in many Table 2 also shows that, for the majority of cases, the
they are stated, but it is not clear if an optimum reduced optimum reduced frequencies F> (,f X /; ) are in
  
504 D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

Fig. 16. Schematic representation of the e!ect of excitation on


lift, de"ning the quantities used in Table 2.

the range 0.3)F>)4 which is consistent with the de-


#ected #ap study described in the previous section. Thus
the range of frequencies, when de"ned on the basis of F>,
Fig. 17. Smoke #ow visualization over the top surface of a (trun-
is signi"cantly smaller than the range e!ective for acous-
cated)  NACA 0015 airfoil (a) baseline data, (b) excitation by
tic excitation, described in Section 3.2, and provides 
means of an oscillating ribbon near the leading edge:
additional clari"cation to the second question posed at Re"40,000 (from [95]).
the beginning of Section 4. However, some investigations,
particularly those of Glezer and his co-workers
[44,103] suggest that higher reduced frequencies, namely observed the physical impact of the LCSs and showed
F>"10 or 20, can be e!ective. Thus, for certain classes their similarity to the generic #ap case (cf. Fig. 12).
of airfoils, the length-scale X may be inappropriate, The e!ect of F> on lift increment on a NACA 0015 for

and a better choice might be bubble length X (cf. a variety of Reynolds numbers, as well as two di!erent
Section 9). The majority of data, however, indicates excitation locations, is presented in Fig. 20 (see Fig. 25 for
that notwithstanding e!ects due to various airfoil geo- a schematic of the airfoil). The two cases presented in the
metries, excitation methods and upstream conditions, "gure are for a symmetric airfoil with leading-edge exci-
similar mechanisms to those observed on the generic tation in the post stall regime and a #apped airfoil at zero
#ap are responsible for observed improvements in airfoil incidence with #ap-shoulder excitation. The "gure indi-
performance. cates that the e!ective excitation range for this airfoil is
The limits of the analogy between the generic #ap approximately 0.5)F>)1.0. The "gure also shows,
discussed in Section 6 and the airfoil studies of this within the limits of experimental scatter, that excitation is
section cannot be accurately assessed due to the dearth of e!ectively independent of Reynolds number, even for
detailed studies of airfoil #ow-"elds. However, NACA Re"150,000, indicating that the state of the upstream
0015 smoke #ow visualization photographs of (amongst boundary layer does not signi"cantly a!ect the optimum
others) Neuburger [95], for F>"2.6 in Figs. 17a and b, F>. A larger reduced frequency range, 0.5)F>)10, for
clearly show the di!erence between separated and at- the same airfoil is presented in Fig. 21 (see [93]) for
tached scenarios, and a$rm the existence of LCSs above various momentum input levels. The "gure con"rms the
the airfoil surface. Moreover, Darabi [101] measured previous observations indicating that F>"O(10) is
separated pro"les of the baseline and excited mixing completely ine!ective for this airfoil. Zhou et al. [38]
layers above the airfoil (Fig. 18) and showed that the show a very similar result for a back-to-front airfoil,
mixing layer above the airfoil is similar to that above the with a peak F>&1 which gradually diminishes and be-
generic #ap. With the #ow attached to the airfoil surface, comes ine!ective at F>&10. At the larger C , the e!ec-
I
Seifert et al. [67] measured phase-locked surface pres- tive frequency range is increased to approximately
sures (see Fig. 19a with further discussion in Section 7.2), 0.5(F>(3, but the trend remains similar at high F>.
D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545 505

Fig. 19. (a) Phase-locked pressure measurement on the upper


surface of a NACA 0015 airfoil for high amplitude excitation; (b)
integrated, phase-locked pressures illustrating lift and center-
of-pressure excursions (adapted from [67]).

Fig. 18. Mixing layer velocity pro"les above the surface of


a NACA 0015 airfoil for (a) baseline data, and (b and c) mixing
In general, once an e!ective frequency has been estab-
layer subjected to excitation at di!erent amplitudes
lished at a minimum (or threshold) C , further addition
(y( ,(y!y )/d , where y
  T  
is the reference distance at I
;"; and d is the vorticity thickness). From [101]. of momentum has a relatively small e!ect on maximum
  T lift. Two examples for NACA 0015 and 0012 (symmetric)
airfoils with leading-edge excitation are shown in
It is important to note here that 6 times the momentum Figs. 22a and b, respectively [67,134]. The "gures both
input did not result in signi"cantly higher *C . Further show baseline data with excitation at respectively e!ec-
*
details, pertaining to comparative airfoil pressure distri- tive reduced frequencies, where C &0.1% and C &1%
I I
butions can be found in Greenblatt et al. [93]. are considered for both cases. Notwithstanding quantit-
ative di!erences between the two airfoils, excitation at
7.1.2. Excitation levels low C increases C , but the most signi"cant e!ects
I * 
Table 2 indicates that e!ective excitation levels are are on *C in the post stall regime (cf. Table 2 and Fig.
*
in the approximate range: 0.01%(C (3%. This 16). An order of magnitude increase in C does not have
I I
translates to typical e!ective relative peak jet velocities a signi"cant impact on lift, and only further increases
(u /; ) [with typical slot-widths h/c(0.5%], or per- C by approximately 10% for both airfoils.

 * 
turbation velocities u/; in the approximate range: Comparisons of C for leading-edge excitation and
 * 
10}300%. As pointed out in Section 4.2 these distur- steady blowing for both airfoils, are presented in
bances are large in the context of acoustic excitation, but Figs. 23a and b. The "gures con"rm the well-known
they are small when compared to e!ective blowing jet observations (e.g. [16]), that steady blowing at
velocities. It is important to note that these disturbance C (2}3% is ine!ective and even detrimental to lift. On
I
levels can be generated by two dimensional slots, but the the other hand, excitation at C &0.01}0.1% is e!ective.
I
upper limits of oscillatory (`microa) mechanical devices, In some instances, the e!ectiveness of excitation is re-
such as #iperons, Piezoelectric benders, etc., such as those duced with increasing C . In the case of the NACA 0015
I
discussed in Section 5, remains to be established. at F>"0.6 (Fig. 23a), the lift increment saturates at
506 D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

airfoils further illustrate that the e!ective C , for an


I
e!ective reduced frequency is in the approximate range
0.01}3%.

7.1.3. Excitation location and slot orientation


In general, hydrodynamic excitation must be applied
at, or close to, the separation line (point). Illustrative
examples are at the shoulder of a de#ected #ap (e.g.
[40,67]) or at the leading-edge of a sharp airfoil (e.g.
[38]). However, the precise location, relative to the separ-
ation point, i.e. slightly upstream or downstream is not
clear and may vary for various airfoils. It is clear that
excitation, close to the leading edge, i.e. x/c)10% can
be e!ective for a wide variety of airfoils (see Table 2).
Hsiao et al. [97] considered the e!ect of excitation at
Fig. 20. The in#uence of leading-edge and #ap-shoulder re- chordwise locations between 1.25%)x/c)13.75% for
duced excitation frequency on NACA 0015 lift for a relatively various post-stall incidence angles on a NACA 63 -018

small reduced frequency range: 0(F>(2. Reynolds number (see Fig. 24 with the leading edge detail inset). (It is
range for both data sets: 150,000)Re)600,000 and assumed that separation occurred close to the leading
ck"0.08% (adapted from [67]). edge due to the relatively large post-stall angles con-
sidered.) They observed maximum lift enhancement
for excitation at x/c"1.25%, which almost halved at
6.25% and completely disappeared at 13.25%. The slot
angle also varied from &1603 (almost directly towards
the free-stream at zero incidence) to &1003 (approxim-
ately vertically upward). Despite the change in excitation
angle, it appears that the primary reason for the reduc-
tion in e!ectiveness is that excitation was applied too far
from the vicinity of separation. A similar observation was
made by comparing the x/c"0% slot employed by
Greenblatt [109] resulting in a 40% C increase with
* 
the 10% slot data presented by Seifert and Pack [107]
which resulted in a modest increase of about 10%.
The investigation of Smith et al. [103], employing a
faired cylinder with NACA aft-portion, is the only invest-
igation where the excitation location was varied while the
slot angle, relative to the surface, was kept constant. They
observed that slot location has little e!ect on C /C for
* "N
the range of 1% (lower surface))x/c)10% (upper
surface). At transonic Mach numbers (M "0.55), Seifert
and Pack [107] observed that excitation has a profound
positive e!ect when introduced upstream of the shock.
Although various slot orientations have been con-
sidered by di!erent investigators, including a 03 (wall-jet)
Fig. 21. The in#uence of leading-edge reduced excitation fre- slot [67], a 453 slot [93], wall-normal slots [103] as well
quency on NACA 0015 lift for a relatively large frequency range as slots that are orientated at an angle that is upstream
0(F>(10 at two excitation amplitudes (from [93]). relative to the #ow [97], the systematic evaluation of slot
orientation for a given con"guration has not been as-
sessed. In future research the e!ects of location and
C "0.08% and in fact decreases with addition mo- orientation should be evaluated independently as
I
mentum input. There is currently no clear explanation a means of maximizing e!ectiveness.
for this, but it may be related to the slot being adjacent to
a region of extremely high curvature at the leading edge. 7.2. NACA 0015 experiments
For the NACA 0012, C increases monotonically
* 
with excitation amplitude, but its e!ectiveness tends to- Extensive experiments carried out on the NACA 0015
wards that of blowing for C '3%. Data from these airfoil section including #ap de#ection [40,67], sweep
I
D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545 507

Fig. 22. A comparison of leading-edge excitation for (a) NACA 0015 and (b) NACA 0012 airfoils at two excitation amplitudes. Insets:
detail of leading-edge slots (adapted from [101,134]).

Fig. 23. Comparison of leading-edge excitation and blowing for (a) a NACA 0015 where the incidence angle is "xed at a"223 and (b)
a NACA 0012 airfoil where the data is extracted from C versus a sweeps (adapted from [67,134]).
*

[105], high Re [104], high subsonic Ma [107], high exception of sweep and high Re studies, the above-
turbulence levels [110] as well as dynamic pitching mentioned experiments were carried out on an airfoil
[111,112], have provided basic understanding of separ- that is equipped with center-span pressure taps and in-
ation control by excitation on this airfoil. With the corporated a 25% trailing edge #ap (see Fig. 25 for
508 D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

because the base #ow over the #ap was separated even at
incidence corresponding to C "0. Replicating this
*
form-drag polar by steady blowing required C "10%,
I
i.e. a factor of 6.25 larger (Fig. 26). It is also important to
note, from Fig. 26a, that excitation from the #ap shoulder
does not signi"cantly a!ect a , while blowing reduces

a by approximately 43.

Experiments similar to the ones described above were
repeated at various #ap de#ections, Reynolds numbers
and blowing intensities. The additional observations
were qualitatively similar, with some e!ects accentuated
at the larger #ap de#ections (see [67]).
To understand and exploit the e!ects of excitation on
Fig. 24. The e!ect of excitation location on post-stall lift of airfoil performance, mean velocity pro"les were mea-
a 63 -018 airfoil. Inset airfoil leading-edge detail (adapted from
 sured in the boundary layer at three streamwise locations
[97]).
above the upper surface of the #ap and one above the
main body of the airfoil, just upstream of the slot [40].
The data shown in Fig. 28 were measured at the inci-
dence angle corresponding to C of the baseline con-
* 
"guration (i.e., a"123). The solid lines represent baseline
data, dashed lines are blowing (C "0.8%) and the dot-
I
ted line represent blowing with superimposed excitation
[C "(0.8, 0.8)%]. The hatched area corresponds ap-
I
proximately to the region of reversed #ow. The "gure
clearly shows that just downstream of the slot, blowing
accelerated the #ow near the wall but the overall e!ect on
the boundary layer was relatively small. The excess mo-
mentum of the jet dissipated rapidly with x due to en-
Fig. 25. Schematic of the NACA 0015 airfoil showing slot loca- trainment of the low momentum #uid above it. On
tions and the de#ectable #ap. the other hand, the superposition of excitation signi"-
cantly increased entrainment from free stream, all
but eliminating the upstream wake. In this instance
schematic). The main body of the airfoil is essentially a new, thick, sub-boundary layer developed which, to-
hollow with the interior serving as a plenum chamber. gether with the elimination of the wake, was responsible
Excitation was introduced by an oscillatory #ow emanat- for the lift increase and pressure drag elimination evident
ing from either the #ap shoulder or leading edge slot, or in Fig. 26.
by means of a ribbon or #iperon. A plot of the phase-locked pressure #uctuations (CI )
.
The e$ciency of #ap excitation (with simultaneous about the mean is provided (see Fig. 19) in order to show
steady blowing) from the #ap shoulder slot at 203 de#ec- more clearly the presence of at least two large coherent
tion may be deduced from Fig. 26, where the dependence structures at every phase of the imposed oscillation at
of the lift coe$cient on incidence (Fig. 26a) is plotted F>"1.1 and C "1.3%. A typical wavelength of the
I
together with the form-drag polar (Fig. 26b). Here, the large eddies (the size of their `footprinta in the x direc-
excitation amplitude (C ) is relatively large, but it is small tion) is e!ectively doubled during their travel above the
I
in the context of a blown #ap with C &10% (see [16]), upper surface of the airfoil. This can be seen by compar-
I
and the di!erence between the two cases is emphasized in ing the relative lengths of j and j . Furthermore, the
 
these "gures. The combination of excitation and blowing, amplitude of the pressure oscillations above the upper
shifted the entire lift curve upwards by *C "0.65 prior surface of the airfoil decreased with increasing stream-
*
to stall and by *C "0.9 at a of the baseline con"g- wise distance and essentially vanished near the trailing
* 
uration, i.e. an increase of 63% in C (Fig. 26a). The edge (Fig. 19a). One practical implication of the presence
* 
form drag was eliminated by excitation at incidences of two or more eddies above surface of the airfoil is that
corresponding to 0.7(C (1.5 (Fig. 26b). The intro- the integration of the instantaneous pressure distribu-
*
duction of pure excitation (without blowing) at F>"2 tions does not result in signi"cant oscillations of the
had the most profound e!ect on the total drag which was integrated forces and moments. This is true even for the
reduced by a factor of 3.6 (i.e. to 28% of its baseline value) relatively large amplitude excitation (C "1.3%) con-
I
at C "1 (Fig. 27). Comparable drag reductions were sidered here. For example, consider the oscillations in lift
*
observed at all lift coe$cients smaller than C (CI ) and center of pressure (XI ) over a cycle, plotted in
*  * 
D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545 509

Fig. 26. The e!ect of high amplitude #ap-shoulder blowing and combined blowing-excitation on the performance of a NACA 0015
airfoil. (a) C versus a depicting three regions and (b) C versus C polar plot (adapted from [67]).
* * "N

Fig. 28. Comparison of velocity pro"les, downstream of the


#ap, for steady blowing and combined blowing-excitation
(adapted from [40]; also see [113]).

for 6.7;10)Re)23.5;10, with #ap d "203 in


Fig. 27. C versus C polar plot for #ap-shoulder excitation on D
* " conjunction with #ap-shoulder excitation, particularly at
the performance of a NACA 0015 airfoil (adapted from [67]). low incidence. This was achieved at relatively low excita-
tion levels of C "0.05%, which are typical of the gen-
I
eric-#ap levels discussed in Section 6 and indicates that
Fig. 19b, where these data are obtained by integrating the separation delay can be achieved e$ciently at high
oscillatory pressures shown in Fig. 19a. Peak-to-peak Reynolds numbers.
variations in C are approximately $3% and center of Conventional steady blowing methods are known to
*
pressure peak-to-peak variations are 0.2% of the chord become less e!ective in compressible #ows (e.g. [68]). For
in the absence of steady blowing. such #ows, porous strips and wall bumps (e.g. [114]) as
The investigation of Seifert and Pack (also see [107]) is well as vortex generators (e.g. [115,116]) have been
unique in that it considered Reynolds numbers consider- shown to exert some in#uence on shock-induced #ow
ably larger than 10 (see Table 2). They observed signi"- separation. Hites et al. [117] showed that excitation
cant lift enhancement, with concomitant drag reduction, remains e$cient up to Ma)0.4 and is far more e!ective
510 D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

than steady blowing. Although C decreases with


* 
increasing Ma, with excitation at x/c"10%, Seifert and
Pack [107] showed that the percentage increase in
C at Ma"0.28 was the same as that for Ma"0.4.
* 
They also showed that unsteady shock-induced separ-
ation (or bu!et) was signi"cantly reduced with leading
edge excitation in the vicinity of the stall angle (see
Section 7.5).

7.3. Ezciency considerations

Possibly the main failing of both steady blowing and


acoustic excitation techniques is that they require an
inordinately large input in order to be e!ective (i.e.
C &10% for blowing and acoustic sound levels
I
'150 dB). In the context of excitation e$ciency, there
are two principal ways to quantify system e$ciency. If
the system is de"ned as the airfoil only, then the appropri-
ate quantity for comparison purposes is the momentum
coe$cient C and as such we rede"ne the conventional
I
airfoil or wing e$ciency (C /C ) as C /(C #C ). On
* " * " I
the other hand, if the system includes the external means
for ewecting excitation, then the power input = must be
G
taken into account in the form of the input power coe$c-
ient (see [84]), de"ned as:
Fig. 29. A comparison of blowing and excitation e$ciencies by
=
C , G considering an `airfoil onlya system (adapted from [67]).
# 1/2oS;

and analogously, the airfoil power e$ciency is then
indicated comparable performance in both excitation
C modes for C (1.1, while for C '1.1 the 2-D mode
g, * * *
C #C was superior. The FM versus C data in Fig. 30b,
" # *
however, indicated that the ratio of power to aero-
Finally, the Figure-of-Merit (FM), which is the ratio of dynamic e$ciency is almost twice as large when the
power and aerodynamic e$ciencies: elements operate in the 3-D mode for C (1.1. For
*
g C '1.2, however, the ratio for both becomes compara-
FM, *
ble. Depending on what is most important, consideration
C /C
* " of the data in this manner allows informed decision-
can be used as a criterion for operating the separation making for an airfoil (or aircraft). For example, if /D is
control device and determine the optimal modes to be of overriding importance and ample power is available,
used for a speci"c #ight condition. then the 2-D excitation mode is appropriate, particularly
For illustrative purposes, consider two airfoil only sys- for C '1.1. On the other hand, if power is in short
*
tems, such as those presented in Figs. 26b and 27. Fig. 27 supply, then the 3-D excitation mode is the obvious
contains data plotted as C versus C #C and Fig. 29 choice, particularly for C (1.1.
* " I *
contains the data from Fig. 26b plotted as C versus The investigation cited above, was not conducted in
*
C #C . For both cases considered (i.e. Figs. 27 and su$cient detail to determine if the streamwise vortices
"N I
29), both blowing and excitation curves move to the generated by the 3-D excitation mode were bene"cial or
right. Note, however, that throughout the a range the detrimental. However, for the two modes, the authors
net ewect of excitation remains bene"cial for both cases, noted a factor of about four di!erence in the power
while blowing is highly ine$cient throughout. required to generate a given actuator oscillation ampli-
An example of accounting for overall e$ciency which tude (namely 2.4W for the 2-D mode versus 0.65W for the
includes the excitation means, is presented in Figs. 30a 3-D mode). The authors presumed that this was due to
and b, where composite piezoceramic-based elements the bene"cial interaction between adjacent actuators (see
are operated in a 2-D mode and a 3-D mode on a Section 5 and Figs. 6b and c) as the downward motion of
thick single-element airfoil (see [84] and Section 5). The alternate actuators assists the upward motion of their
traditional aerodynamic e$ciency, presented in Fig. 30a, immediate neighbors.
D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545 511

Fig. 31. NACA 0015 airfoil moment data for two leading-edge
excitation amplitudes; corresponding to the lift data of Fig. 22a
(adapted from [101]).

be aerodynamically ine$cient. For example, Amitay


et al. [44] showed that at *C "120%, their airfoil
* 
was essentially ine$cient with /D &3.
N
An important feature of all the studies involving lead-
ing-edge excitation is that *C in the post-stall regime
*
can be signi"cantly improved } thus indicating that the
most pronounced e!ects are to be found in the post-stall
regime. This aspect is particularly important for dynamic
stall control, since the airfoil spends an appreciable frac-
Fig. 30. A comparison of di!erent excitation modes by consid-
ering the airfoil as well as the external means for e!ecting tion of its cycle in the post-stall regime. Related to this,
excitation (from [84]). the control of the airfoil pitching moment is of crucial
importance in the context of dynamic stall, but few of the
above investigations reported this. A notable exception is
7.4. Performance enhancement Darabi [101] who showed that static moment stall can
be delayed by approximately 33 on the NACA 0015
The quantity *C is the simplest indication of the under the excitation conditions F>"1.1 and C "0.1%
*  I
e!ectiveness of an excitation method. Table 2 shows that, and up to 83 for C "1.3% (see Fig. 31). In accordance
I
in contrast to acoustic excitation, signi"cant increases in with lift data (cf. Fig. 23), steady blowing has an in-
C can be obtained by hydrodynamic excitation, for consequential e!ect on pitching moment for C )2%
*  I
a wide variety of airfoils. Of particular note is the large (not shown here).
maximum lift increment of Seifert et al. [40,67] cited
above where excitation is at the shoulder of de#ected 7.5. Control of post-stall unsteadiness
#aps. Large increases in C , e.g. on a shortened 
*  
NACA 0015 [33] and a cylinder with NACA aft portion Airfoils in the deep post-stall regime exhibit well
[44], for example, indicate the enormous potential for known `blu!-body sheddinga of vortices at St+0.2
separation control on non-aerodynamically shaped bo- (e.g. [61,118]). However, close to the onset of static stall,
dies, such as stores, undercarriages as well as rotorcraft Zaman et al. [76] and others have observed low-fre-
fuselages. It should be noted, however, that although quency periodic oscillations with St&0.01. Zaman et al.
improvements in C are signi"cant the body may still showed that these oscillations are not associated with the
* 
512 D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

magnitude less than that observed by Zaman et al. [76].


Similar stall behavior was observed on a NACA 0012
airfoil at Re"4;10 [211].
Greenblatt et al. employed two-dimensional excitation
at 140 Hz (F>"1.5) and 320 Hz (F>"3.5) to control
these oscillations, which cannot be detected on this time-
scale. The "gure indicates that excitation eliminates
C and C oscillations almost entirely at both reduced
* +
frequencies. Hot-wire measurements in the upper surface
boundary layer at x/c"0.3 indicated that the boundary
layer was fully turbulent and that the perturbations re-
sulting from the relatively small values of 1c 2 required
I
to eliminate the large amplitude oscillations (1c 2"
I
0.075%) completely dissipated by x/c"0.3. Increasing
the incidence angle well into the post-stall regime showed
that the peak oscillations in lift between 123)a)163
could be more than halved, while moment oscillations
were reduced by two-thirds [109].
Seifert and Pack [107] showed that considerable con-
trol could be exerted over bu!et, even with excitation
introduced upstream of the shock wave. This was in-
dicated by steadier wake #uctuations throughout the
measured frequency and particularly at low frequencies
which are more likely to interact with an airframe.

7.6. High-lift systems

Over the last few decades, multi-element airfoils have


become established as the main means for achieving high
lift in large-scale transport aircraft. In the context of
#ow control, two aspects appear to be of primary import-
ance: Firstly, can #ow control produce signi"cant gains
when applied to existing high-lift systems? And, secondly,
can #ow control reduce the number of wing elements,
Fig. 32. Control of large low frequency oscillations in lift and thereby reducing cost, weight and complexity? A recent
moment near static stall by means of low amplitude excitation study by McLean et al. [119] shows that the potential
(from [93]).
bene"ts of applying #ow control to high-lift systems are
signi"cant.

Karman vortex street, but rather with periodic switching 7.6.1. A multi-element airfoil
between attached and separated states. Although the The question of how excitation can be used to enhance
mechanism is not understood, these oscillations could be lift on a multi-element con"guration was investigated
modi"ed by acoustic `trippinga. Greenblatt et al. [93] by Seifert et al. [67] on a cambered 19% thick PR8-40
considered a NACA 0012 with an e!ective leading-edge airfoil with a 30% chord slotted #ap, at low Reynolds
trip and observed a similar phenomenon at one degree numbers (1.4;10)Re)6;10). Excitation ema-
beyond the static stall angle (a"123). Instantaneous nated from either of two locations: one from the trailing
surface pressure measurements were integrated to yield edge of the main element (ME) and the other from the
lift and pitching-moment time histories (see Fig. 32). The leading edge of the #ap (see Fig. 33a). This airfoil illus-
data con"rmed the "ndings of Zaman et al. and indicated trated the e!ect of the passive blowing, occurring
the abrupt nature of separation and attachment, which between the main element and the #ap, on the e!ective-
induced signi"cant peak-to-peak C excursions, up to ness of the location from which the perturbations are
*
*C "0.3 about a mean value of C "0.8, while the introduced.
* *
moment excursions even exceeded their associated mean The e$ciency of thick airfoils is signi"cantly reduced
value. Hot-wire measurements in the upper surface at low Re due to #ow separation. The separation creeps
boundary layer at x/c"0.3 indicated a prefered reduced up from the trailing edge on the upper surface with
switching frequency St&0.001 which is an order of increasing incidence. To arrest this process, slots are
D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545 513

Fig. 34. Surface pressure coe$cients on the two-element PR8-


40 airfoil comparing the introduction of excitation from the
surface of the #ap with excitation from the trailing edge of the
main element (from [67]).

between the wake of the main element and the potential


#ow above it, while the perturbations introduced from
the #ap interacted "rst with the new boundary layer
initiated by the #ow through the slot.
Fig. 33. (a) Schematic of the two-element PR8-40 airfoil, show- The above discussion can be further elucidated by
ing the main-element and #ap slot locations; (b) comparison of considering pressure distributions measured on the sur-
lift on the two-element PR8-40 airfoil for steady blowing and face of the airfoil at a"83, with and without excitation at
excitation (d "303, Re"140,000; from [67]).
D 1c 2"0.02% for the two excitation locations described
I
above (see Fig. 34). The #ow over the #ap was completely
separated without excitation and was only partially at-
introduced in the aft section of the airfoil. The high tached with excitation emanating from the main element.
momentum #ow emerging from the slot mixes with the Signi"cant #ow attachment was achieved with identical
sluggish boundary layer #ow above it, making it more 1c 2 applied along the #ap surface: the minimum pres-
I
resistant to separation. The amount of momentum sure coe$cient, C , near the #ap leading edge was re-
.
needed to overcome the strong adverse pressure gradient duced to !2, with pressure recovery over most of the
with increasing #ap de#ections is ultimately not su$- #ap chord. Pressure distributions over the entire airfoil
cient, and this problem is exacerbated by viscous e!ects indicate an overall performance improvement resulting
at low Re. from the #ow attachment over the #ap, and further
An important observation made by Seifert et al. [67] explain the signi"cant lift increments evident in Fig. 33b.
on multi-element airfoils is that it is more e!ective to
introduce excitation from the surface of the #ap than 7.6.2. Single element airfoils with large yap deyections
from the trailing edge of the main element. A similar Of great practical importance, particularly with regard
observation was made by Lin [120] using micro-vortex to a commercial aircraft, is the viability of reducing the
generators on a similar con"guration. Fig. 33b shows cost, weight and complexity by eliminating one or more
that the lift coe$cient increased from 1.9 to 2.5 when wing elements. This involves investigating the viability of
1c 2"0.015% was introduced on the #ap. Seifert et al. excitation at large #ap de#ections (i.e. d '403), typical
I D
[67] observed that a momentum input of nearly "ve of landing con"gurations [119]. Such an investigation
times this value (not shown) was required to generate was carried out by Greenblatt et al. [93] on a NACA
a comparable C when the perturbations were introduc- 0015 (see Fig. 25) at a"03 where excitation was applied
*
ed from the trailing edge of the main element. Further- at the #ap shoulder.
more, with #ap perturbations the corresponding drag For this airfoil, a #ap de#ection of d "203, with
D
coe$cient (not shown) decreased from 0.125 to 0.06 while a"03, generates a large lift increment (*C "0.82) in
*
a factor of 15 increase in main element perturbations was the absence of #ap-shoulder excitation (Fig. 35), but the
required to generate a comparable glide ratio (C /C ). e$cacy of the simple #ap wanes at higher #ap de#ections.
* "
In the latter case the oscillations enhanced the mixing This can be seen by increasing the #ap de#ection by
514 D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

Fig. 35. Lift-enhancement with large #ap de#ections on the NACA 0015 airfoil (from [93]).

*d "203 beyond d "503, which results a much small- approaches 0.1 at C "1.5. Increasing the momentum
D D *
er *C "0.1. Steady blowing at C "1.2% increased coe$cient by a factor of 205 at F>"1.1 has a negligible
* I
the lift generated at d "203 from *C "0.82 to 0.95 e!ect on C (Fig. 35c), while low frequency excitation
D * * 
[i.e. (*C ) "0.13], but was totally ine!ective at (F>"0.3) is e!ective at low #ap de#ection angles
*
d *403. Periodic excitation at identical C was more (Figs 35c and d).
D I
e!ective than steady blowing at all #ap de#ections. Spe-
ci"cally, with F>"2.5, (*C ) "0.23 even at d "653 7.6.3. The highly curved yap
* D
(due to the particular construction of airfoil d could not To assess the e!ects of curvature, the regular NACA
D
exceed 653). Excitation at F>"2.5 was more e!ective 0015 #ap was replaced by a circular arc #ap (see Fig. 36a)
than F>"1.1 for all values of d * particularly with described by Greenblatt et al. [93]. At an incidence of
D
regard to form-drag reduction (Fig. 35b) where the di!er- a"!83 the airfoil provides no lift and the #ow is
ence in form drag at C "1.4 is C +0.03. When com- attached to its upper surface. Flow separation initiates
* "N
pared to baseline or steady blowing data, the reduction in near the trailing edge and progresses upstream with in-
C due to excitation (F>"2.5) at C "1.4 is 0.07 and it creasing a; thus the lift slope reduces in a continuous
"N *
D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545 515

Fig. 36. (a) a schematic of a NACA 0015 airfoil with a highly curved #ap; (b) and (c) performance of the airfoil subjected to #ap-shoulder
excitation (adapted from [93]).

manner with increasing a (Fig. 36b). The maximum lift however, suggested that the actuator should be located in
due to the modi"cation of the airfoil increased from the vicinity of the separation point, e.g. at the shoulder of
C +1 to C "2.2, although this was accom- a de#ected #ap [31] or at the leading-edge of an airfoil
*  * 
panied by large form drag, C , yielding C /C +18 at [67]. This clearly was not the case here, since the location
"N * "N
C "1.5 (Fig. 36c). Steady blowing, from the slot loca- of the actuator remained "xed at X "75% chord as it
* 
tion shown in Fig. 36a, at C "0.8% did not increase was for the straight #ap. The authors assumed that the
I
C , it in fact decreased C at all pre-stall incidence actuator was too far upstream of the separation location
*  *
angles (a(153). Its e!ect on form drag was even more and thus excitation was ine!ective since the amplitude
detrimental increasing it by approximately 30% at low decayed between the actuator location X and the sep-

C (Fig. 36c). Periodic excitation, at a reduced frequency aration point X . This problem was partially solved by
* 
of F>"2.9, reduced the form drag signi"cantly thus de#ecting the #ap to d "203, as this brought the natural
D
increasing C /(C #C ) to 27 at C "1.3. For this separation location closer to the actuator and generated
* "N I *
con"guration, to increase C by a mere 10% required a larger *C than was achieved previously (Fig. 37a).
*  *
a periodic perturbation of C "2% (not shown) while This di!erence was particularly obvious for a(33, but
I
a similar C on a straight #ap increased the C by C did not increase above its naturally attained base-
I *  * 
63% (see Fig. 26 and Table 2). line value of 2.3. The form drag (C ) was reduced to  of
"N 
Based on the above data, it appeared that convex its natural value over the entire range of C 's (Fig. 37b).
*
curvature has a negative impact on lift enhancement in Lower frequency excitation at F>"1.6 was as e!ective
the presence of periodic excitation. Previous experience in reducing C up to a"53 while steady blowing did
"N
516 D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

Fig. 37. Performance of the highly curved #ap airfoil subjected to #ap-shoulder excitation with #ap de#ection (adapted from [93]).

not increase the lift nor did it generate a large reduction It should be appreciated that these studies are in their
in C . The quantity C /(C #C ) for this case (not infancy and that the application of periodic excitation to
"N * "N I
shown explicitly) increased almost linearly with increas- high-lift systems has not yet seriously begun; the preced-
ing C and dropped sharply at stall. The maximum value ing examples present only an indication of what is pos-
*
of C /(C #C ) was almost doubled due to excitation sible. For multi-element airfoils, e!ective control will
* "N I
but even steady blowing increased C /(C #C ) require extensive optimization to apply the system to
* "N I 
from 11 to 14. Increasing C to 2% at F>'1 did not a complex airfoil. This will involve tying together the
I
improve the performance of the airfoil except near detailed geometry of the airfoil as well as Reynolds num-
C (Fig. 37c), however when the same amplitude is ber to parameters such as amplitude, frequency and the
* 
applied at a lower frequency of F>"0.33 the ensuing location from which the oscillations emanate. In terms of
*C is greatly increased but at the cost of increasing reducing complexity, the e!ects of curvature, #ap angle,
*
C (Fig. 37d). actuator location, etc. will demand extensive study.
"N
The preceding sections illustrated separation control Indeed, McLean et al. [119], identi"ed the acquisition of
on slotted airfoils as well as the e!ect of excitation com- data to de"ne basic aerodynamic performance as their
bined with large #ap de#ections and a highly curved #ap. highest-priority item.
D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545 517

7.7. Airfoil drag anomalies C visible on this scale. At d "153 the #ow separated
"N D
from the #ap and thereafter there was a rapid increase in
The observations made at large #ap de#ections (Sec- drag mostly attributed to C . However, since the #ow
"N
tion 7.6.2) and on the curved #ap (Section 7.6.3) suggest over the main body was not separated C was always
"
that frequency and amplitude are not decoupled at high- larger than C as expected.
"N
er levels of C and this coupling may be exacerbated by With excitation initiated at the #ap-shoulder, however,
I
curvature and the addition of steady blowing. A similar Greenblatt et al. observed an apparently anomalous re-
observation was "rst made on the generic #ap [31,81] sult, namely C was consistently larger than C for all
"N "
where an increase in C caused the #ow to reattach at #ap de#ections considered. Initially this was thought
I
lower values of F>. Pressure distributions on the curved to be due to the C `thrust e!ecta, but this could only
I
airfoil suggest that the low frequency excitation generates account for part of the di!erence as seen by the magni-
a form of time-dependent supercirculation resulting from tude of C indicated at the top of Fig. 38b. There are at
I
the strong, large vortices that are continuously present least two additional possible explanations for the mea-
over the upper surface of the #ap. The results at large sured further di!erence in drag. One possible explanation
#ap de#ections over the NACA 0015 show similar is that eddies, generated by high amplitude excitation,
trends when C is large and F> is low. In both cases the were strong enough to turn the external #ow around the
I
downward pointing surface over which control is exerted, #ap shoulder while concomitantly generating strong re-
generates a low pressure region, thus causing a large C . verse #ow at the surface. A second possible explanation is
"N
However, since the mean-#ow is not separated from the that the eddies interact with the wake in a manner not yet
surface the total drag might in fact be smaller. properly understood. Regarding the former explanation,
This phenomenon was indeed observed on various the average skin friction resulting from the reverse #ow
airfoils at large #ap de#ection or high angle of incidence on the surface would be negative and thus provide thrust!
(e.g. [93,109]). Typical results, for the airfoil pro"le of Due to the di$culty of measuring the magnitude and
Fig. 38a, are presented in Fig. 38b and illustrate the total direction of skin friction, no de"nitive statements can be
drag and the form drag polars corresponding to a"03 made, although skin friction thrust would need to ac-
for a case with a 30% #ap de#ected at 03(d (503. count for approximately C !C +0.025 at C "1.1
D "N " *
In the absence of excitation and prior to separation the for example (Fig. 38). Pressure and PIV measurements
drag was low with negligible di!erence between C and have shown (see Section 6) that high-frequency excitation
"
would increase the duty cycle of the skin-friction thrust
and thus provide larger drag reduction. On the other
hand, low frequencies generate fewer but stronger eddies
that generate additional lift.
Notwithstanding a proper understanding, the reduc-
tion in total drag can be enormous. For example, at
C "1.1, C is reduced from 0.17 to 0.02. Thus the large
* "
drag reduction, coupled with the form-drag}total-drag
anomaly renders this problem fascinating from both
basic as well as applied perspectives, and therefore merits
further investigation. It should also be noted that in the
above-mentioned experiment C was measured by a rake
"
of Pitot tubes and was therefore susceptible to errors
resulting from high turbulence levels. Although these
errors were estimated at less than 4%, an independent
measurement method, such as a hot-wire anemometer,
should be used to con"rm these results.

8. Di4user studies

Tollmien apparently related (see [14,15]) that


Prandtl's interest in hydrodynamics was initiated when
he failed to achieve the expected recovery in a di!user.
Since Prandtl went on to invent boundary layer control
Fig. 38. Illustration of the total-drag}form-drag anomaly on an (BLC), a review of separation control in di!users is, at
airfoil at zero inidence, with #ap-shoulder excitation. Flap the very least, of historical signi"cance. Possibly the
de#ection angle increases from d "03 to d "503 (from [93]). simplest di!user-type con"guration is a two-dimensional
D D
518 D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

backward-facing ramp within a wind tunnel. Viets et al. end, research carried out in actively excited separation
[36] examined such a con"guration, where disturbances bubbles is reviewed within the context and theme of this
were introduced upstream of a ramp discontinuity at article. In addition, although the main thrust of this
reduced frequencies in the range 0.2)F>)0.8, where review is on actively excited #ows, due attention is paid
0
the ramp length is used in the de"nition of to some features of naturally excited separation bubbles,
0
F>,f /; . They found that pressure recovery on since they are an important prerequisite for understanding
0  0 
the ramp could be improved to some degree throughout bubbles subjected to active excitation. Finally, de#ected
the F> range tested, but 0.7)F>)0.8 was the most #ap data is included at the end of this section, in an
0 0
e!ective. Nishri [121] (also see [28,29]) considered a sim- attempt to synthesize naturally and actively excited bubble
ilar con"guration, where disturbances were supplied by data, as well as to explain certain anomalies present.
means of a #iperon at the ramp discontinuity for
1.5)F>)6. Nishri ascertained that, although signi"- 9.1. Natural excitation
0
cant pressure recovery could be attained throughout the
range of reduced frequencies, there was no clearly dis- Mabey [126] analyzed low-speed reattachment data
cernable optimum F>. Both investigations revealed that from a wide variety of natural bubbles including airfoils,
0
the most e!ective reduced frequencies were F>"O(1). steps, cavities, expansions and fences and found that all
0
Recently, Pack and Seifert [122] studied #ow attach- exhibit broadly similar pressure #uctuation peaks near
ment to the walls of a short, wide-angle di!user, which the reattachment point. Mabey observed that the fre-
was attached to the nozzle of a circular jet exit by means quency associated with pressure #uctuation peaks,
of excitation introduced at the junction between the could be used to construct a frequency parameter, based
nozzle exit and the di!user inlet. Jet vectoring was on bubble length, namely St ,f X /; , where
6  
achieved by applying the excitation over a fraction of the 0.5(St (0.8 for all cases considered. Sigurdson and
6
circumference of the circular jet, thereby enhancing its Roshko (see e.g. [127,128]) further analyzed di!erent
spreading rate on the excited side and therefore its tend- separating #ows and established that a Strouhal number,
ency to reattach to that side. In a similar manner to based on bubble-height h and velocity at separation ; ,
Q 
previous di!user studies, cited above, jet de#ection angles namely St "f h /; , has a universal value of approx-
F  
did not show a marked sensitivity to excitation fre- imately 0.08. This value was found to be consistent with
quency, but rather to relative direction between the exci- Levi's [129] `Universal Stouhal Lawa. They suggested
tation and the jet #ow. that the #ow is characterized by both the Kelvin}Hel-
moltz instability of the free shear layer as well as a shed-
ding-type instability of the entire bubble. The former
9. Excitation of reattaching 6ows interacts with itself and the latter with its images result-
ing from the presence of the wall. Kiya et al. [135]
In many instances of practical importance, a separated proposed the separation bubble is a self-excited #ow
#ow reattaches downstream of the separation point, maintained by a feed back loop. On the basis of this
without active excitation. Thus there is no breakaway or argument, they showed that St +0.5, which is consis-
6
detachment of the #ow and a natural recirculating region tent with Mabey [126] as well as other investigations.
or separation bubble is formed. Due to the prevalence of
these #ows, for example on wings (e.g. [123,124]), many 9.2. Imposed excitation
investigations have been conducted which seek to under-
stand their mean structure (e.g. [125]) and time-depen- With imposed excitation, analysis of separation bub-
dent features (e.g. [126]). Separation bubbles, at the point bles is further complicated due to a number of additional
of separation, can be both laminar or turbulent. In the factors which are not universally addressed. Firstly, in
former case, for practical Reynolds numbers, transition addition to Kelvin}Helmholtz and shedding type in-
evolves in the free shear layer downstream of separation, stabilities an additional excitation frequency f is intro-

and the #ow reattaches as a turbulent boundary layer (see duced. Secondly, excitation level, measured in terms of
e.g. [124]). In the latter case, #ow is turbulent throughout. percentage of free-steam or relative momentum addition,
In contradistinction to #ows presented in Sections 6}8, can have a signi"cant impact on certain characteristics of
the present section deals mainly with actively forced or the #ow, such as bubble length and height (c.f. Section 6).
excited separation bubbles, which reattach uncondi- Thirdly, the method, and particularly the location, of
tionally. Thus, these #ows involve the control of a separ- excitation can profoundly a!ect the #ow. A comparison
ated region, rather than the control of separation, per se. of various con"gurations, with various methods and
Their inclusion in this article is based on the fact that levels of excitation are presented in Table 3. In addition,
excited separation bubbles can constitute a signi"cant each investigator's original frequency parameter is
fraction of a #ow that would otherwise be separated, such presented in the table, along with a frequency parameter
as that on #aps (Section 6) or airfoils (Section 7). To this based on bubble length, i.e. F>,f X /; (cf. [126]).
 
Table 3
Summary of bubble excitation data

Con"guration Investigator (s) Original parameter for Excitation method and F>"f X /; Comment
 
minimum bubble length location

Backward-facing step Reisenthel et al. [87] f h /; "0.05}0.1 Oscillating #ap; 0.2}0.4 Excitation most e!ective near
 D 
h "#ap height Variable location in separation; Decrease in bubble
D
reattachment region length with increasing #ap
amplitude
Roos and Kegelman [35] f h /; "0.29 Oscillating #ap at separation 1.3}1.8 Laminar and turbulent #ow;
  
h "step height point &30% reduction in bubble length

Huppertz and Janke [131] f h /; "0.18 Hydrodynamic, with speaker 0.44 Laminar #ow;
  
at separation point &63% reduction in bubble length
Coaxially mounted Sigurdson and Roshko [127] f h /; &0.16}0.4 Hydrodynamic, with speaker &1.3}3 Drag on cylinder face reduces with
 
ciruclar cylinder Sigurdson [128] f D/; "&2.5}4 at separation point reducing bubble size; Higher
 
h "bubble height excitation frequencies result in
; "velocity at smaller bubbles.

separation
Kiya et al. [135] f D/; &2.5}4 Hydrodynamic, with speaker &0.5 High amplitude excitation can
 
D"cylinder diameter at separation point eliminate the recirculation region
Two-dimensional Siller and Fernholz [45] f h /; "0.05 Hydrodynamic, with speaker &0.09 Reattachment length reduced by
 D 
fence h "fence height upstream of fence &50%.
D
Sound pressure in chamber"
137 dB
De#ected #ap Nishri and Wygnanski [31] F>,f /; &1 Hydrodynamic (with speaker) &0.7 Bubble length can be reduced by
 
or #aperon at #ap-shoulder increasing frequency or excitation
D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

level

Parameter for minimum drag on cylinder face.


Optimum reduced frequency for #ow attachment.
519
520 D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

For the remainder of this section, particular attention


is paid to excited separation bubbles over backward-
facing steps, blunt axially-mounted cylinder/cylinders
and the de#ected #ap.

9.2.1. The backward-facing step


The two-dimensional backward facing step is one of
the simplest con"gurations demonstrating a "xed point
of separation, followed by downstream #ow reattach-
ment and boundary layer development. Due to its im-
portance as a generic paradigm of #ow reattachment,
signi"cant e!ort has been expended in understanding its
time-mean features as well as organized structures within
the #ow (e.g. [7,8,132,133]).
Reisenthel et al. [87], Roos and Kegelman [35] and
Fig. 39. Backward-facing step data, illustrating the e!ect of
Huppertz and Janke [131] studied excitation of these
excitation on bubble length (adapted for the addition of F>;
#ows by employing a spoiler like #ap located at the base from [35]).
of the step, an oscillating #ap at the separation point and
oscillatory blowing and suction holes at the separation
point, respectively. These investigations showed that ex-
citation enhances vortical structures and increases en- tions is of the same order, and suggest that the vortical
trainment (cf. Sections 4 and 6), and in so-doing exerts structures in the shear layer are dominant and override
substantial control over the reattachment length. The the natural (i.e. unexcited) di!erences in the laminar and
traditional means for presenting backward-facing step turbulent shear layer development. They further point
data is to non-dimensionalize quantities with respect to out that when the separating #ow is transitional, single
the step height h , e.g. St "f h /; (e.g. [35]) although frequency excitation is less e!ective in regularizing the
 F   
Reisenthel et al. suggest that oscillating #ap-height development of the large-scale vortical structures and
should be used. hence less e!ective at reducing bubble length.
It is pointed out here, however, that there is no means
for objectively comparing the relative e!ectiveness of the 9.2.2. The blunt axially-mounted cylinder
above-mentioned investigations since none of them Sigurdson and Roshko [127,128] and Kiya et al. [39]
clearly presents the level of momentum added to the #ow (also see [135,136]) have both carried out extensive in-
(cf. Section 6). Related to this is the di$culty associated vestigations of excitation separation bubbles formed at
with determining an optimum reduced frequency (based the leading-edge of an axially mounted blunt-faced cylin-
on minimum bubble length for example) if excitation der. Sigurdson and Roshko (see Table 3 and Fig. 40)
levels are not explicitly accounted for. For example, in determined that e!ective reduced frequencies at a wide
the mixing layer, Cohen [80] showed that high excitation range of excitation amplitudes, for minimizing cylinder
levels cause an overshoot of h> in region II (see Section face pressure drag (C ), were in approximate the range:
"D
4), followed by a contraction to a lower value, thus 0.16)f h /; )0.4 (alternatively 2.5)F>,f D/;
  "  
resulting in a higher e!ective reduced frequencies. )4, where D is the cylinder diameter). Sigurdson further
In the context of separation control, larger excitation determined that pressure drag reduction on the blunt
levels tend to have lower e!ective frequencies (e.g. cylinder face is linearly dependent on excitation
[31,75,109,134]). These points are illustrated further by amplitude.
Table 3, which shows a range of reduced frequencies both Kiya et al. considered minimum bubble reattachment
when based on step height (0.18}0.29) or bubble length length, for a similar range of excitation frequencies and
F> (cf. [26]; 0.2 versus 1.8 [35,87]). amplitudes, and also found that the approximate opti-
It might appear that the upstream state of the bound- mum frequency range is 2.5)F>)4, for excitation
"
ary layer is responsible for the large di!erences, but for an amplitudes less than 10%. Kiya et al. showed that
upstream laminar boundary layer Huppertz and Janke bubble-length reduces logarithmically with excitation
achieve a 60% rediction in bubble length at F>"0.44 amplitude (see Fig. 41). In fact, for excitation amplitudes
while Roos and Kegelman only achieved half of this larger than about 14% the bubble is virtually eliminated.
value at F>"1.3. Plotting the data of Roos and Kegel- This should be contrasted with the above-mentioned
man to include F> (see Fig. 39) shows that the most data of Sigurdson and Roshko which shows that the
signi"cant bubble shortening is achieved in the fully maximum drag reduction is a linear function of excita-
turbulent regime. They show that the reattachment tion amplitudes up to 24%. Kiya et al. further show
length for both laminar and turbulent upstream condi- that increasing the frequency well above the optimum,
D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545 521

Fig. 42. Comparison of percentage reduction in bubble length


Fig. 40. Normalized pressure drag on the cylinder face as a func- [135] and normalized cylinder-face form-drag reduction [128]:
tion of frequency and excitation amplitude. Symbols in the key u/;+6% for both cases (from [128]).
of the "gure refer to u/; , i.e. relative rms amplitudes of

excitation (from [128]).
the #ap does not attach unconditionally. In Section 6
we considered the minimum momentum addition neces-
sary to a!ect attachment as well as prevent separation.
Here, for a given excitation frequency, we consider the
minimum momentum required to prevent the bubble from
bursting.
Consider, once again Fig. 9b, which illustrates #ow
attached to the #ap for F>"1.2 where a reduction in
C results in separation from the #ap by the mechanism
I
of bubble bursting. Therefore, just prior to separation,
F>"0.66 which is in the naturally excited bubble range
proposed by Mabey [126] and close to that of Kiya et al.
[130]. Moreover, normalizing the data with respect
Fig. 41. Minimum bubble length as a function of relative rms
amplitudes of excitation (from [135]). to bubble height and velocity at separation, yields
f h /; +0.08, which is the same as that proposed by
 
Sigurdson and Roshko [127]. Reducing f and perform-

approximately 10(F>(50, at constant excitation am- ing the same analysis at F>"0.7 (see Fig. 10), where the
"
plitude of 1%, results in a slight increase in unexcited bubble tends to the end of the #ap, results in similar
bubble size. Further increases, 50(F>(170 have values for F> and f h /; . As pointed out in Section 6, a
"  
a negligibly small e!ect on bubble length. Sigurdson small additional increase in the bubble length, due to
[128] compared his normalized form drag data, with the additional de#ection or reduced C , caused it to burst.
I
normalized bubble-length data of Shimizu et al. [136], The analysis presented here is inappropriate for high
for similar excitation amplitudes (see Fig. 42). He found values of F>. For example, at F>"7.5 (see Fig. 10),
a striking correlation between the two for a wide range of bubble size was an insigni"cant fraction of the #ap length
the reduced frequency f h /; , with its optimum value at and separation occurred as a result of momentum de-
 
around 0.35, which is more than four times larger than pletion in the boundary layer near the #ap trailing edge.
f h /; . The above analysis shows that an analogy between
 
naturally and actively excited bubbles may only be ap-
9.2.3. The deyected yap propriate if it is based on the minimum momentum
Separation control on a de#ected #ap, discussed in required to prevent bubble bursting. Fig. 43 [81] empha-
detail in Section 6, is an excellent example of the role of sizes this by showing how the bubble dimensions change
a separation bubble as it pertains to separation control. with increasing 1c 2. When viewed in this context, it
I
In the current context, this #ow is suggested as a para- appears that conditionally attached #ows are best suited
digm for resolving the di!erences in reduced frequencies for illustrating these principles. Moreover, the universal
for unforced and forced data discussed above. One of the Sigurdson}Roshko reduced frequency f h /; +0.08,
 
principal di!erences between a #ap (or airfoil), and based on the natural (shedding) frequency, is also appro-
the two cases presented above, is the fact that #ow over priate for exited bubbles in the form f h /; , providing
 
522 D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

may reduce the intensity of the streamwise vortices while


enhancing the in#ectional instability.
This section is concerned with fundamental studies
which elucidate the e!ects of curvature as well as attempt
to isolate controlling parameters when radius of curva-
ture is "xed (i.e. two-dimensional circular cylinders).
Curvature, in the context of an airfoil #ap used as a high-
lift device, is discussed in Section 7.6.

10.1. Maximum pressure recovery and concave curvature

It is generally desirable to minimize the distance over


which the deceleration of a boundary layer takes place.
Some applications include maximizing the surface area
Fig. 43. E!ect of increased excitation amplitude on the de#ec-
ted-#ap bubble length. Prevailing conditions: Re "450,000,
over which an airfoil #ow is laminar or providing greater
D useable volume per unit surface area and greater struc-
d!d "13, F>"1.6 and G/ "0.5% (from [81]).
 D tural e$ciency, as well as a reduction in length and
weight for di!users. Clearly, this desire must be tempered
the minimum momentum (or amplitude) required to prevent by the requirement to avoid separation. The competing
bubble bursting is considered. priorities have led several investigators to examine the
largest adverse pressure gradient that can be sustained
without causing massive separation from the surface. The
10. E4ects of curvature idea that the boundary condition giving the maximum
rate of pressure recovery corresponds to vanishingly
The e!ect of excitation on #ows over curved surfaces is small wall stress was "rst proposed by Stratford [9,138].
important since curvature is known to have a profound Although his derivations are based on the Reynolds
e!ect on the development of turbulent boundary layers averaged equations, which may not apply in the presence
(see e.g. [137]). For example concave curvature is known of excitation, the #ow is of considerable interest because
to initiate the generation of streamwise vortices through of its application to airfoil design (e.g. [124,139]).
a centrifugal instability of the #ow. This instability mech- Elsberry et al. [140,141] (also see [11]) considered the
anism may compete with the in#ectional instability ex- impact of excitation on #ow over a concave-shaped
ploited by excitation and limit the spanwise coherence of `Stratford Rampa (see Fig. 44). For the baseline #ow on
the eddies. Consequently, this may have a detrimental the verge of separation, it was ascertained that h de-
impact on separation control. Conversely, an adverse veloped linearly with x and that the shape factor,
pressure gradient in the presence of convex curvature H,dH/h, was approximately constant at H"2.5.

Fig. 44. Setup for the `Statford rampa experiment, intended to examine the e!ects of concave surface curvature (from Elsbery [140,141];
also see Wygnanski [11]).
D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545 523

Consideration of the two-dimensional integral momentum layer, or whether it has a detrimental e!ect due to com-
equation revealed that the experimentally observed vari- peting instability mechanisms.
ation in free stream velocity, namely ; &x\  was

consistent with vanishingly small skin friction coe$cient: 10.2. Convex curvature~Coanda cylinder
C +0 and constant H. Mean velocity pro"les, plotted
D
in wall variables (not shown), were dominated by the To separate the e!ects of curvature from those of
wake component and well exceeded the value obtained pressure gradient, Neuendorf and Wygnanski [143] (also
by Spangenberg and Roland [142] who experimented see [144]) considered the evolution of a two-dimensional
with the Stratford concept in the absence of curvature. wall jet #owing over the exterior of a circular cylinder
Excitation at an imposed amplitude of u/; "0.2% (see Fig. 46). For such a #ow, the wall jet spreads outward

was tested for its ability to increase C when the geo- to a far greater extent than that on a corresponding #at
D
metry is prescribed. Two frequencies were used: one at surface, indicating that the centrifugal instability intro-
f h /; "0.016 and the other at 0.032, corresponding to duced by the curvature contributes to the transport of
 G G
40 and 80 Hz for the upstream free stream reference momentum across the #ow. Wygnanski [11] suggests
velocity ; "15 m/s. (Note that a reduced frequency F>, that the addition of periodic excitation, may interact with
G
based on a streamwise length-scale with either free steam both instabilities (in#ectional and centrifugal) and am-
or local velocity, is inappropriate here.) The lower excita- plify the perturbations more e!ectively. Wygnanski [11]
tion frequency oscillations underwent large ampli"cation presented a comparison of two case with no free-stream
yet they did not result in a notable increase in the skin is present: with and without excitation. It was shown that
friction. Nevertheless, the mean velocity pro"les became excitation induces separation by enhancing the transport
fuller near the surface (Fig. 45) suggesting a transfer of of momentum away from the surface at a far smaller
momentum toward the wall. The higher frequency excita- angle (i.e. approximately 303 less) than when no excita-
tion, on the other hand, resulted in a signi"cant increase tion is imposed. The opposite may occur when the cylin-
in the skin friction despite the fact that its integrated der is placed in a free stream where the direction of
intensity decayed by an order of magnitude over the momentum transport will be reversed. In such a case, the
distance considered. Elsberry et al. further showed that nature of the stability will change, unless a strong wall jet
the maximum value of ;>,;/; reduced from 80 to is employed. One possible application of such a scenario
O
60 (also see [11]) and the shape factor H decreased from is a slotted #ap, where the #ow emerging from the slot
2.5 to 2.2 indicating that a boundary layer forced in such e!ectively constitutes a wall jet.
a manner may well sustain a larger pressure gradient
without separating. 10.3. Constant curvature~the circular cylinder
The e!ectiveness of the present method of separation
control was attributed to the rapid ampli"cation of the Cross-#ow over a two-dimensional circular cylinder is
forced perturbations by the preexisting inviscid instabil- a problem of signi"cant practical and theoretical interest.
ity of the mean #ow. From the above data, however, it
is not clear whether concave curvature (or increasing
"d; /dx") enhances this instability by providing addi-

tional transport of momentum across the boundary

Fig. 45. Near-wall velocity pro"les on the Straford ramp down-


stream of the excitation location, for baseline and excited #ow Fig. 46. A schematic representation of the the Coanda cylinder
(from [140,141]). experiment (see [144]; also see [11]).
524 D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

Amongst other problems of importance, drag reduction, the "rst demonstration of the potential for periodic
control of vortex shedding as well as the generation of lift excitation in the context of three dimensional, and in
are the three most intimately connected to separation particular swept back, con"gurations.
control. Initially sound was used to demonstrate the
e!ect on separation (e.g. [145,146]), but in recent years, 11.1. The swept back wing
hydrodynamic excitation has been applied to the prob-
lem (e.g. [44,97,147}151]). The application of excitation to three-dimensional
The above investigations determined profound e!ects #ows requires, "rst and foremost, an understanding of
of excitation on lift, separation location and vortex shed- the simplest possible con"guration } namely the yawed
ding frequency at sub-critical Reynolds numbers. For cylinder of in"nite span, i.e. the e!ects of sweep
example Hsiao et al. [97] used asymmetric excitation and [105,152]. Furthermore, if the method is to be used as
showed that a suction peak could be generated on one a design tool, a set of transformations is needed to
side resulting in C "0.6 with the excitation location at convert data obtained in two dimensions to the cases
*
1003 and F>,f D/; &1. Similar observations were involving sweep (cf. [153]). To this end, experiments
"  
made by Amitay et al. [44] for F>"1.5. Furthermore, were conducted on an 18% thick, airfoil where the pro-
"
Amitay et al. showed that the separation point could be "le became equivalent to that of a NACA 0015 at
moved by approximately 603 resulting in a 25% reduc- sweep-back of 303. The airfoil was equipped with a 1/3
tion in form drag. While exciting the #ow at F>"2, Liu chord #ap and with an excitation slot located at the #ap
"
and Brodie [151] observed wake momentum de"cit re- shoulder.
ductions associated with delay of separation as well as By assuming that the swept-back con"guration is of
the location of vortex shedding. It should be noted that, in"nite span and the sweep is attained by rotating the
for the above-mentioned cases, excitation frequencies basic two-dimensional con"guration, Reynolds number
were typically an order of magnitude larger than the is de"ned as (see [105] and Fig. 47):
natural shedding frequency. Heine et al. [150] observed
a reduction in the turbulent coherent energy with excita- Re "Re cos K (11.1)
tion at a subharmonic of the natural shedding frequency " "
with F>"0.1. Similarly, since the pressure distribution and thus the
"
It is di$cult to provide precise values for separation line are mainly governed by the chordwise
F>,f X /; or F>,f X /; in the current invest- component of velocity (; cosK): C "C cosK
      N" N"
igations, since attachment length, bubble length and exci- and C "C cos K (cf. [153]). Following similar
tation amplitude are not accurately known. However, *" *"
arguments Naveh et al. [105] de"ned
for the vast majority of investigations cited above, excita-
tion was found to be e!ective when applied near the F> "F> /cos (K) (11.2)
laminar separation point at &903. Thus, it seems that " "
neither curvature nor the state of the separating bound- and
ary layer has a signi"cant e!ect on optimum reduced
frequencies, since F>"O(1) and F>:1. The investiga- C "C /cos(K) (11.3)
I" I"
tions also con"rm the observations of the previous
sections that separation control is most e!ective when Lift and pressure distributions on the airfoil, with the
the excitation location is in the vicinity of the separation #ap de#ected at 203, were compared for K"03 and
point. This provides further evidence that excitation K"303 (see Figs. 48 and 49). The two data sets essential-
should not be applied to the stable #ow near the leading ly collapsed on one another when the transformation was
edge, as the disturbances decay before becoming used, for the baseline case as well as for the case with
unstable. active excitation. This demonstrates the validity of the
above transformation and facilitates the comparison of
pressure distributions at various sweep angles, K. The
11. Three-dimensional e4ects validity of this transformation can be further assessed by
comparing the C measured near the front stagnation
.
In 1987, Gad-el-Hak and Blackwelder observed that points where C
1 regardless of K. The comparison is
.
excitation along the leading edge of a 603 delta wing satisfactory, except in the reattachment zone of the lead-
signi"cantly a!ected the evolution of the bound shear ing edge bubble, which occurs around x/c
0.3. For the
layer originating from the leading edge. They ascertained baseline case, #ow over the #ap was separated, while
that the maximum changes to the vortices occurred when excitation forced the #ow to reattach over the #ap, result-
the excitation frequency was a subharmonic of the natu- ing in the reinstitution of the adverse pressure gradient at
ral shedding frequency corresponding to F> ,f c/; x/c'0.7 and an increase in circulation over the entire
"  
"5.5 and C "0.42% (c is the root chord). This was airfoil.
I
D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545 525

Fig. 47. Experimental setup and de"nitions for the swept NACA 0015 experiment (from [105,152]).

Fig. 48. Pressure distributions on the NACA 0015 airfoil, with Fig. 49. Lift versus incidence angle for the NACA 0015 airfoil,
the #ap de#ected at 203 for K"03 and 303 illustrating the with the #ap de#ected at 203 at K"03 and 303 illustrating the
transformation between 2D and swept con"gurations (from
transformation between 2D and swept con"gurations (from
[105,152]). [105,152]).

#ow is separated should cause reattachment, scaling with


11.2. The Delta wing the dimensionless frequency:
f X
It was suggested by Wygnanski [154] that the results F> "   +1 (11.3)
"6 ; cos K
obtained for yawed cylinder of in"nite span in the pre- 
vious section (i.e. the swept back wing) may be extrapo- where X is the distance between the leading edge and

lated to a "nite wing planform. This suggests that peri- the reattachment location normal to the leading edge,
odic excitation applied to a delta wing over which the and ; cos K is the normal component of velocity as

526 D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

above. This assumption appears to be tenuous because it


is based on extrapolating a transformation that relies on
in"nitely yawed wing section to a "nite delta wing. Thus
it would not be expected to apply in the vicinity of the
apex or the wing-tip.
An experiment, reported by Wygnanski [11] was con-
ducted on a 603 sharp-leading edge delta wing and in-
dicated that changing the frequencies of excitation
altered X in accordance with Eq. (11.3). This result is

illustrated in Figs. 50, which shows adapted results of
oil-surface #ow visualization for baseline and excited
#ow over the delta wing. The #ow visualization indicates
that, with excitation, the reattachment line on the wing
surface is approximately parallel to the leading-edge,
apart from the region near the apex of the wing. The
Fig. 51. E!ect of excitation on the `lift indexa of a 703 delta wing
result was similar to #ow visualization (not shown) car-
(from [43]).
ried out by Naveh [152] on the yawed wing discussed in
the previous section, and is attributed to the presence the
tunnel wall (modeled as an vortex image) that represents "rstly 2-D disturbances are often easier to introduce and
the plane of symmetry at the apex. quantify; and secondly, the (possibly mistaken) notion
Investigations carried out by Guy et al. [43] deter- has its historical basis in Squire's [156] Theorem which
mined that vortex breakdown over a 703 delta wing could implies that states that 2-D disturbances are `more
be extended by about 25% of the chord by introducing dangerousa in the context of laminar-turbulent transition
periodic perturbations at the leading edge. The delay in and, therefore, `more e!ectivea in the context of #ow
the breakdown resulted in a high velocity "eld on the control (see e.g. [61]).
upper surface of the wing that in turn increased the It is apparent however, that three-dimensional distur-
negative pressure there. In an associated study, Guy et al. bances such as those produced by vortex generators, and
[155] observed a signi"cant increase in maximum introduce streamwise counter-rotating vortices, can sig-
lift under the conditions F> "1.4 and C "0.4% (see ni"cantly delay airfoil stall, resulting in up to 40% in-
" I
Fig. 51). crease in maximum lift (see e.g. [2,120]). Such streamwise
vortices can also be produced by steady jets (e.g. [157]).
11.3. Three-dimensional excitation McManus and co-workers (e.g. [158}160]) suggest using
pulsed jets in an e!ort to combine the bene"cial e!ects of
Within the context of the control of 2-D #ows, it is streamwise vortices with the LCSs that are generated by
generally assumed that 2-D excitation is more e!ective pulsating the #ow. The concept has been demonstrated
than 3-D excitation. There are two main reasons for this: on airfoils [158,159] as well as on a generic "ghter
aircraft con"guration. Airfoil data show a 50% increase
in post-stall lift and a 30% increase in C . A generic
* 
"ghter aircraft con"guration data exhibited a 7% in-
crease with reduced frequency f c/; "1.
 
Seifert et al. [84] employed spanwise, surface mounted
actuators on an airfoil that could be operated in a 2-D
mode (all actuators in-phase), or a 3-D mode (adjacent
actuators 1803 out-of-phase). They found that the same
lift increment could be achieved with both 2-D and 3-D
actuation at F>"1.2 (also see Section 5, Figs. 6b, c and
30). However, 3-D actuation required approximately  of

the power required by 2-D actuation to achieve the same
lift increment. Kiya et al. [130] demonstrated stall
alleviation by introducing upstream-generated three-
dimensional vortex rings into the separated layer. They

Fig. 50. Adaptation of oil-surface #ow visualization on the  Squire actually proved that a two-dimensional #ow becomes
surface of a 603 delta wing for baseline and excited con"gura- unstable at a higher Reynolds number when the disturbance is
tions (adapted from [11]). assumed to be 3-D than when it is assumed to be 2-D.
D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545 527

ascertained that the minimum wake de"cit occurs at ation. The reasons is that for steady, two-dimensional,
f c/; "4, where f is the frequency at which the up- separation the point of zero skin friction is coincidental
  
stream vortex rings were introduced. with the point of separation, while in unsteady #ows skin
friction can be zero, or even negative, with no apparent
breakaway of #ow from the surface (e.g. [4]). Due to the
12. Engineering implementation complexity of unsteady separation, some controversy still
exists over its precise de"nition. While analyzing revers-
Unmanned aerial vehicles (UAVs) present a logical ing unsteady #ow in the vicinity of a stagnation point
starting point for the application of separation control Rott [167] observed no singularity nor invalidation of
concepts. This is primarily due to cost considerations as the boundary layer assumptions. Further investigations
well as the elimination of exposing a pilot and passengers by Sears [168] and Moore [169] gave rise to the notion
to danger. Moreover, in recent years, the use of UAVs of simultaneous vanishing of velocity and shear, either at
has increased substantially and unmanned combat aerial a point above a moving wall or in a reference frame
vehicles (UCAVs) are gaining ground as alternatives to moving with the separation point on a "xed wall. This
conventional combat aircraft [161]. Recently attention well-known Moore}Rott}Sears (MRS) criterion has ser-
has focussed on Micro Aerial Vehicles (MAVs), which are ved as a basis for the analysis of various cases of unsteady
signi"cantly smaller than conventional UAVs } typically separation (e.g. [170,171]). Sears and Telionis [172,173]
from 150 mm down to insect size (e.g. [162,213]). These showed that boundary layer separation singularities ac-
aircraft bring with them a host of new challenges relating company #ow detachment, while more recent investiga-
to #ight and separation control, since conventional #ight tions have found di!erent dominant mechanisms, such as
theory does not apply at Re(70,000, which is typical for van Dommelen}Shen interaction (e.g. [174}176,214]).
these aircraft. Particularly, at low Reynolds numbers Despard and Miller [177], however, were the "rst to
wings stall at small incidence angles and hysteresis can be provide experimental evidence that reversed #ow can
signi"cant [153]. Amongst other methods, conventional exist within an otherwise attached unsteady boundary
steady BLC (e.g. [163,164]) as well as #apping and #ex- layer. Moreover, they made the important observation,
ible wings have been proposed [162,165] in the academic which will be expanded upon in the sections that follow,
and popular press. that #ow oscillation frequency has a profound e!ect on
Seifert et al. [166] wind-tunnel tested a UAV incorpor- the reversed #ow region.
ating an Eppler 214 wing section with a simply-hinged The viewpoint adopted here (see Section 1), is that #ow
30% #ap at Re"270,000. Excitation was introduced by breakaway (or detachment) is the key feature of both
means of an axial fan and rotating valve which delivered steady or unsteady separation (e.g. [3]). Thus any excita-
C "(0.18, 0.03)% in the reduced frequency range
I tion method that controls dynamic detachment or reat-
0.5)F>)0.7 at the #ap shoulder. The results in- tachment is considered to be within the scope presented
dicated, for the entire vehicle, increases in C and
*  here. A problem immediately apparent here is that both
(/D) of 0.15 and 2 respectively. Although these results
 separation and the means to e!ect its control are un-
seem modest, one should bear in mind that the fuselage steady! In an attempt to shed light on this problem,
parasite drag is large. More importantly, and with pos- dynamic separation from the generic #ap (cf. Section 6) is
sible application to MAVs, separation control by #ap- used once again as the paradigm for studying the mecha-
shoulder excitation on one wing was able to generate nisms of dynamic separation control. In particular, em-
a rolling moment comparable to that of a conventional phasis is placed on the large time-scale disparity between
aileron. Moreover, the substitution of a conventional the time required for the #ow to separate (or stall) and the
aileron with an excitation device, does not have the characteristic time associated with the excitation used
familiar adverse yaw e!ects. in separation control. The principles embodied in this
Greenblatt and Wygnanski [108,154] recently showed section are then used for analyzing data on airfoils
that periodic excitation on airfoils at very low Reynolds presented in Section 14.
numbers (30,000)Re)50,000) can increase lift and
glide ratio by a factor of four and an order of magnitude, 13.2. Onset of dynamic stall
respectively, thereby restoring conventional low
Reynolds number performance. 13.2.1. Dynamic stall simulation on a stationary yap
The onset of separation and formation of the DSV can
13. Dynamic separation control be simulated on a stationary #ap (see Section 6) in the
following manner. Consider a boundary layer that is
13.1. Unsteady separation forced to remain attached to a #ap under the conditions:
F>"1.2, (d!d )"43 and C "0.06%. At time t"0,
 I
Historically, the notion of unsteady separation has C is reduced abruptly from its prevailing value to zero
I
confounded traditional notions of so-called steady separ- thereby precipitating total separation of the #ow from
528 D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

Fig. 52. Surface pressures at various locations along the #ap as Fig. 53. Spatial surface pressure distributions over the #ap,
a function of dimensionless time after cessation of excitation measured at various dimensionless times after cessation of exci-
(from [33]; also see [93]). tation (from [92]; also see [93]).

the surface. Instantaneous pressure data is then sampled During this time interval the DSV increased in size and at
and phase averaged to determine dynamic changes that q+5 its center was located around x/ +50%. At this
D
occur during the separation process as a function of time point in time the DSV tore itself loose from the mixing
(Fig. 52). It is expedient to de"ne the dimensionless time layer (i.e. from the continuous vortical sheet) and was
q"t; / and note that the large coherent structures convected downstream. The downstream motion of the
 D
in a shear layer (or a mixing layer) propagate over the DSV was accompanied by an increase in pressure over
#ap at an approximate phase velocity ; +; . Thus, the taps located upstream of the DSV. The generation
(  
the time required for the information to travel the length and advection of the DSV is a slow process that requires
of the #ap is /; +2 /; and corresponds to q+25 for completion (i.e. for the DSV to leave the
D ( D 
q+2, which should not be confused with the period of trailing edge of the #ap). This is indicated by the coales-
the excitation, 1/f that is also marked on the "gure for cence of all the pressure tap signatures into a single curve

reference. Note that for the special case F>"1, q"t ) f . at q+25.

Surface pressure data, described below, is based mainly The spatial pressure distributions over the #ap, mea-
on the work of Darabi [33]; also see [93]. sured at various times after the cessation of excitation are
The time records provided by the various pressure taps plotted in Fig. 53. The adverse pressure gradient existing
located along the #ap re#ect the delay in the arrival of over the #ap at q+0 represents the forcefully attached
the information to the given tap (Fig. 52). Thus at #ow. The onset of separation is marked by the creation of
x/ "96% the information about the cessation of actu- a DSV (or bubble) that increases in size with time. Since
D
ation arrived at q+2. The diagonal hatched line in the the areas under these curves represent the normal force
range 0(q(2, shown in Fig. 52, represents the approx- experienced by the #ap during the separation process one
imate onset of separation at the prescribed streamwise may assess what happens during the creation and advec-
location. At small x/ the separating streamline (not tion of the DSV. Consider the pressure distribution mea-
D
shown), that initially follows the sharp bend in the sur- sured at q+4 and compare it to the pressure distribution
face, straightens up and its radius of curvature increases prevailing at the initiation of the process. As the size of
with time. Flow separation results in a decrease in the the DSV increased over the time interval considered, the
pressure that existed near the #ap leading edge suction peak near the leading-edge decreased, while
(see the time history of the pressure tap located at !C increased simultaneously for approximately
.
x/ "9%). In viscous #ow, the free streamline repres- x/ '22. Nevertheless, the integrated normal force,
D D
ents a mixing layer that is very unstable and rolls up into calculated from these pressures, increased during this
discrete vortices; therefore where the rollup occurs, time interval due to the increasing size of the DSV. This
a bubble is formed and the pressure in the bubble de- is a well known characteristic of airfoil dynamic stall
creases. This can be observed by the pressure tap located (e.g. [178]) and is discussed further in Section 14.
at x/ "29% after a time delay q+0.7. As the DSV
D
builds up and increases in size, the pressure on the solid 13.2.2. Static-yap~dynamic-airfoil analogy
surface that bounds it, decreases. Consequently at An analogous scenario to the process described above,
x/ "29% the pressure dropped until q+2 and there- is that of a thin airfoil pitching past the static stall angle
D
after remained approximately constant up to q+5. (see [109]) } thus in both cases, at an instant marked by
D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545 529

airfoil half-cycle time (q+30). Of crucial importance here


is that the DSV and LCS time-scales (or frequencies)
di!er by a factor of about 30, while the detailed compari-
son is of secondary importance. This result forms the
basis of the stark contrast between the destructive nature
of the DSV and separation-suppression attribute of exci-
tation. The experiment described above was repeated at
di!erent free-stream velocities and #ap lengths with the
data scaled up at the appropriate ratios thereby a$rming
e!ective Reynolds number independence. It should also
be noted that a pitch-and-hold motion executed by the
airfoil, from a pre-stall to a post-stall incidence, would
also have provided a suitable analogy. It should be further
appreciated that dynamic stall on a stationary #ap can be
controlled by abruptly changing the reduced frequency
F> at constant C . This form of dynamic separation
I
without the dynamic stall vortex is discussed by Green-
blatt et al. [179] and expanded upon by Darabi [33].

13.3. Forced dynamic attachment

Starting with a separated #ow over the #ap one may


turn the excitation `ona at a prescribed time q"0 and
force the reattachment of the mixing layer to the surface.
A similar process occurs on a pitching airfoil when the
instantaneous incidence angle falls below an angle, typi-
Fig. 54. Comparison of dynamic stall on a pitching airfoil with cally less than the static stall angle a , which depends on

dynamic stall simulation on the de#ected #ap (from [93]; also the frequency of the pitching motion, k. In order to
see [109]). analyze the coherent mechanisms of the attachment pro-
cess, the procedure was repeated 200 times and the data
ensemble averaged. During the transient process, care
q"0, we have attached #ow in a natural (unforced) was taken to repeat the experiment at precisely the same
post-stall regime. Figs. 54a and b compare the static #ap phase of the excitation in order to assess its coherence
with airfoil dynamic stall for C and C , on the basis of
* + over the reattachment period. As with dynamic separ-
q, where q"t; /c for a NACA 0012 airfoil with
 ation control, generation of the normal force on
a"93#53sin(2nf t!903), k"0.05 and Re"0.24 the #ap takes a long time relative to the interval of the
;10 (see Section 14 for more details). Note that the #ap
data is plotted from the cessation of forcing while the
airfoil data is plotted for a*a during one-half of the

pitch cycle. Notwithstanding the quantitative di!erences
between the #ap and airfoil, the overall trends are mark-
edly similar. The vortex formation on the pitching airfoil
is apparent from the increase in lift with simultaneous
drop in the  chord pitching moment. The same phenom-

enon is apparent on the stationary #ap discussed in the
previous section. Following lift-stall, the moments for
both airfoil and #ap increase, since the #ow tends toward
reattachment in the former case, while total separation
prevails in the latter case. During dynamic stall, the
pressure distributions over the airfoil and #ap exhibit the
same basic trend for both cases, namely: a sharp rise in
C near the leading edges, accompanied by a more mod-
.
erate decrease in C over the large remainder of the
.
chord [93]. Fig. 55. Phase-averaged surface pressures on the #ap for various
Fig. 54 also indicates the extent of one half of the LCS streamwise locations. All measurements are referenced to the
dimensionless time (q"1) and contrasts this with the excitation-initiation time (from [92]; also see [93]).
530 D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

increase in C . Darabi et al. [92] identi"ed a minimum


.
attachment time q +16 at a reduced frequency in

the vicinity of F>"1, which did not reduce signi"cantly
beyond a threshold C . This result was used by
I
Greenblatt et al. [182] for intermittent excitation
on oscillating airfoils and is described further in
Section 14.2.4.

14. Control of airfoil dynamic stall

Dynamic stall is often the primary factor limiting heli-


copter (e.g. [178,207]), "ghter aircraft (e.g. [183]) and
wind turbine [184] performance and as a consequence
much basic research has focussed on controlling (manag-
ing) or eliminating the phenomenon. BLC (see [185]) has
been proposed for rotorcraft applications and includes
methods such as steady blowing [178,186] suction
[187,188] and pulsed blowing [189,190,212] as well as
Fig. 56. Initial stage of the phase-locked surface pressures on the
control of boundary layer transition [191}193]. Modi"-
#ap, for various streamwise locations (from [92]; also see [93]).
cations to airfoil geometry, such as leading-edge slats
[194}196] leading-edge droop [195], rotation of the
leading-edge [197] as well as dynamically deforming
excitation. This interval may be considered complete geometries [198,208] have been investigated. In all cases,
when the last pressure tap located near the trailing these changes are geared speci"cally to the leading-edge
edge of the #ap responds to the amplitude of the forcing region where the dynamic-stall vortex originates. Re-
frequency in a stationary, but periodic, manner (see cently, periodic excitation has been shown as an e!ective
Fig. 55). The "gure indicates that all pressure taps further method for delaying incompressible dynamic stall (e.g.
downstream responded while the overall pressure distri- [11,93,109,111,112,179}182,199,210]). A brief summary
bution over the #ap was still changing. The "gure indi- of important "ndings, all of which involve incompressible
cates that the total dimensionless time required for re- wind tunnel testing, is presented below.
attachment is q+20 and, within the time resolution of
the measurement, is of the same order of magnitude as 14.1. Experimental setup and objectives
the time required for separation.
There is a short transient, however, that results in an Dynamic stall control in incompressible #ows was
increase in a normal force before the generation of lift is performed on a NACA 0012 airfoil as well as the NACA
initiated (Fig. 55). Expanding the time scale of the signal 0015 airfoil described in Sections 7.1 and 7.2. The NACA
to 0(q(5 reveals that, in the vicinity of x/ "30%, 0012 was of similar construction to the 0015, but was
D
the separated mixing layer attaches to the surface "rst equipped with only one slot at x/c"5%, angled at 453 to
(Fig. 56). This attachment results in a positive C (i.e. the chord-line, and was not equipped with a #ap. Both
.
a transient stagnation point) and distorts the wave form airfoils were tested in a closed loop wind-tunnel (see e.g.
of the excitation. Once again the enhanced rollup of the [40]) and oscillated by either $53 or $103 in pitch
free mixing layer is detected by the surface pressure taps about their 1/4 chord positions at typical rotorcraft
for x/ '30%. The increased pressure region is con- reduced frequencies k. By measuring dynamic surface
D
vected downstream at a representative ; +; . By pressure and wake velocities, two interrelated objectives
(  
the time this pressure disturbance reached the trailing were pursued: The "rst objective was to maximize
edge, the `pumping actiona (due to the entrainment of dynamic airfoil performance by means of periodic exci-
the mixing layer) reduced the C near the leading edge of tation (e.g. increase C and/or decrease C ), provid-
. *  "
the #ap to approximately !0.2 or by approximately ing that the moment coe$cient excursions (C ) were
+ 
25% of its asymptotic, steady state, attached value. within bounds typical of pre-stall values (C +0.04
+ 
The attachment therefore is initiated at the location and 0.06 for NACA 0012 and 0015 respectively. For
corresponding to the "rst rollup of the mixing layer and the remainder of this section, the symbol E denotes

it proceeds downstream. By initiating the excitation at `allowable moment excursionsa). The second objec-
random phases of the perturbation signal the forcing tive was to explain and analyze the "ndings on
frequency is smeared out by the averaging pro- the basis of the dynamic #ap data presented in
cedure resulting in a clearer observation of the transient Section 13.
D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545 531

Fig. 57. The e!ect of leading-edge blowing and oscillatory excitation on `lighta dynamic stall of a NACA 0012 airfoil [109];
a"93#53 sin(2pf t!903); E refers to allowable moment excursions.


14.2. Summary of important results consistent with both static data presented in Sections
6 and 7 as well as dynamic data on the NACA 0015. Both
14.2.1. Leading-edge excitation excitation frequencies, on the other hand, were e!ective
Figs. 57a and b illustrate the e!ect of oscillating the in attenuating the DSV. In particular, for F>"3.5 at
NACA 0012 by $503 such that the static stall angle is C "0.02%, drag was reduced by approximately 20%,
I
exceeded by 33 with k"0.05. Baseline, excitation and resulting in an exceedingly low relative momentum input:
blowing data are presented in the "gures, where solid C /*C "0.1%. Note that steady blowing must supply
I "
lines indicate the upstroke while hatched lines indicate a full two-orders of magnitude more momentum in order
the downstroke. As is well known, dynamically pitching to achieve comparable drag reductions. The 0012 data
an airfoil past the static stall angle results in higher lift presented here indicated that, in general, larger C results
I
than could be obtained statically and this is clearly evi- in greater gains, e.g. slightly higher lift, lower drag or
dent from Fig. 57a. However, the higher lift associated smaller moment excursions. This was somewhat di!erent
with the baseline data is accompanied by destructive to the 0015 data, where optimum momentum inputs were
nose-down pitching moment excursion shown in Fig. observed for various reduced excitation frequencies.
57b, which is associated with the DSV. When excitation In accordance with PIV measurements presented in
is employed (currently F>"1.5 and C "0.5%), the Section 6 and airfoil data in Section 7.7, Fig. 58c shows
I
same lift is achieved, but the DSV is eliminated, as can be that time-mean total drag is at times less than the form
seen from the moment excursions that are commensurate drag, speci"cally when F>"3.5. As in Section 7.7, it is
with pre-stall values, which were &0.04 for the NACA surmised here that, on average, the skin friction on the
0012. As expected from static #ap and airfoil studies (see upper surface may be negative, i.e. a separated #ow
Sections 6 and 7), steady blowing at the same C had region exists but it is controlled and orderly and its net
I
a negative impact on airfoil performance. e!ect is positive. This reiterates the importance of the
A summary of data for C , C as well as time-scale disparity introduced in Section 13 and shows
*  + 
C and C , under the conditions of Figs. 57a how control of a separated region can be further used in
"   "N  
and b, is presented for a range of momentum inputs at a positive manner.
two e!ective reduced frequencies in Figs. 58a}c. The Figs. 59a}d illustrate the e!ect of both increasing the
"gure shows that steady blowing at low momentum maximum incidence to 63 beyond the static stall angle,
input coe$cients had an adverse e!ect on the baseline as well the e!ect of airfoil oscillation rate on excitation
data and only starts being e!ective at C +2%; this is performance. For this `deepa stall case the baseline data
I
532 D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

It should be noted that the e!ects of small amplitude


leading-edge excitation (C &0.1%) on the NACA 0015
I
(e.g. [93,112]) were signi"cant at angles very much larger
than a (&a #103), whereas for the 0012 small ampli-
 
tude excitation was e!ective up to a #33. This was

attributed to the observation that stall begins in the
trailing-edge region on the NACA 0015 and moves grad-
ually upstream with increasing incidence, with full lead-
ing edge stall at &a #103. For the 0012, stall occurs

abruptly from the leading edge, due to the high leading
edge radius-of curvature (cf. Fig. 32 of Section 7.5). Con-
sequently, in the deep stall regime, the 0012 required
a signi"cantly larger C in order to e!ect meaningful
I
control.
McCroskey et al. [200] (also see [178]) showed that
the incidence angle rate-of-change in the vicinity of stall
is of paramount importance for characterizing dynamic
stall. Speci"cally, a so-called `matched pitch-ratea can be
obtained by increasing incidence angle excursions while
concomitantly reducing the airfoil pitch-rate. To deter-
mine if this principle applies to dynamic stall control as
well, the following two cases were considered:

Case I: a"123#53 sin(2pf t!903) for k"0.15,


Case II: a"73#103 sin(2pf t!1803) for k"0.1.

Plotting the lift and moment coe$cients for the two cases
illustrated the similarity that exists in the vicinity of
maximum incidence for both baseline and controlled
data (Figs. 60a}d). Small di!erences in airfoil acceler-
ation near maximum incidence were considered to be
responsible for slight disparity between the two cases, for
both baseline and controlled scenarios.

14.2.2. Flap-shoulder excitation


Before considering dynamic stall control by excitation
at the #ap shoulder, consider again the NACA 0015 static
case presented in Section 7.2 (see Fig. 26a). The "gure
illustrates typical baseline C vs. a characteristics for
*
a #apped or aft-loaded airfoil, where three distinct re-
gions may be discerned [I } pre-stall a(03; II } partial-
stall emanating from the trailing-edge region
Fig. 58. Summary of important aerodynamic indicators for (03(a(123); III } leading-edge stall (a"123)]. Note
`lighta dynamic stall control by leading-edge blowing and exci- that, although excitation increases C , it is too far aft
tation on a NACA 0012 airfoil [109]. * 
to signi"cantly a!ect the static-stall angle. Steady blow-
ing, on the other hand, at more than six times the mo-
exhibits signi"cant lift hysteresis, coupled with large mo- mentum coe$cient, decreases a for the same C .
 * 
ment excursions, which is typical of deep stall. Excitation To illustrate a dynamic case, consider the represen-
at F>"1.4 with C "2% succeeds in increasing tative example presented in Figs. 61a and b, for
I
C by up to 35% (relative to baseline) and reduces d "203, Re"0.3;10, k"0.1 and a"103#53
*  D
moment excursions, to approximately one-third of their sin(2pf t!903). At low incidence (region II) both blow-
baseline value, resulting in moment excursions that are ing and excitation increase lift, much like in the static
commensurate with pre-stall values. An important e!ect case. However, as the airfoil approaches a , surface

of excitation is that it almost entirely eliminates the e!ect pressures (not shown here) reveal that the yow over the
of k on airfoil performance. Similar observations were yap separates for blowing, thereby increasing the entire
made on the NACA 0015 airfoil. upper surface pressure, but #ow over the #ap remains
D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545 533

Fig. 59. The e!ect of leading-edge excitation on a NACA 0012 airfoil pitching at various rotorcraft reduced frequencies [109].

attached during excitation } in accordance with static somewhat aft of the  chord position in order to produce

data. As the airfoil pitches down, #ow over the blown #ap a small mean moment.
remains separated and signi"cantly increases upper sur- The impact on total drag, for the above-mentioned
face pressure, resulting in large lift hysteresis and unac- case, can be assessed from phase-locked and time-mean
ceptably large moment excursions. For the actively excit- velocity measurements in the wake, which are represent-
ed case, the #ow over the #ap remains attached and, ed here as momentum de"cit (Figs. 62a}d and 63). At
as a consequence, lift hysteresis is eliminated and mo- a"53 the form of the pro"les is similar, but the large lift
ment excursions are well within acceptable limits increment caused by blowing and excitation is evident
(C +0.06). It should be noted from Fig. 61b that from the location of the momentum de"cit peaks, as well
+ 
excitation introduces a mean nose-down pitching mo- as the de"cit disparity on the top and bottom parts of the
ment as a result of attaching the #ow to the #ap. Al- wake. As the airfoil pitches up to 103, the lift di!erence
though elimination of this feature was not considered becomes more clearly manifested from the relative loca-
central to the objectives of the investigations cited above, tions of the momentum de"cit peaks. At a"153, the
it is undesirable, and would require an airfoil pitch axis e!ect of dynamic yap stall for the blown-#ap case is
534 D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

Fig. 60. The principle of `matched pitch-ratea on a NACA 0012 airfoil applied to both baseline and leading-edge controlled scenarios
[109].

characterized by a large momentum de"cit in the wake commences in the trailing-edge region. For example, on
above the airfoil surface. During the pitch-down motion, a symmetric NACA 0015, #ap-shoulder excitation (i.e.
at a"103, the e!ectiveness of excitation is clearly evident when the #ap was not de#ected) increases dC /da and
*
by comparing 14.6b and 14.6d, where the data associated consequently the C excursion (C ), in addition to
* * 
with excitation are virtual mirror images of one another. increasing in C , which is equally important for an
* 
The time-mean momentum de"cit data (Fig. 63) provides advancing rotorcraft (see Section 14.2.4). It is important
an overall picture of this process and indicates that the to note that a further increase in incidence, for the above
mean drag is more than halved with excitation, while example, resulted in leading-edge stall, which could not
blowing only achieves an 11% reduction. be controlled by #ap-shoulder excitation. This highlights
The method of excitation presented above is not neces- an important di!erence between leading-edge and
sarily intended for #apped airfoils, but rather for aft- #ap-shoulder dynamic stall control: for leading-edge
loaded airfoils or even symmetric airfoils where stall excitation the airfoil pitch-up past the static stall angle
D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545 535

Fig. 61. The e!ect of high-amplitude #ap-shoulder blowing and excitation on a pitching NACA 0015 airfoil [109,112].

provides the additional lift and the DSV is eliminated by put forward some `informed speculationa. Seifert and
excitation, while #ap-shoulder excitation eliminates dy- Pack investigated the e!ect of periodic excitation on a
namic #ap stall and in that manner increases C with static NACA 0015 airfoil under the conditions
* 
simultaneous control of C excursions. Finally, it is 0.28)Ma)0.55 and Re&O(10). For all cases, excita-
+
pointed out that for the symmetry NACA 0015 airfoil, for tion had a bene"cial e!ect on lift and drag, with the most
a wide range of F> and C , #ap-shoulder excitation was signi"cant e!ects in the post-stall regime. In fact, Seifert
I
more e!ective than leading-edge excitation for increasing and Pack showed that upper surface suction that is
C and reducing mean drag. attenuated as a result of an increase in Ma from 0.28 to
* 
The authors wish to emphasize that #ap-shoulder exci- 0.4, can be restored by excitation at extremely low per-
tation cannot control the traditional DSV that is gener- turbation amplitudes. Greenblatt et al. [93], on the other
ated in the leading-edge region of the airfoil. Rather, hand, showed that the net result of excitation is essential-
#ap-shoulder excitation controls dynamic separation over ly insensitive to whether the airfoil is stationary or oscil-
the yap (or aft-region) and is therefore benexcial only when lating in pitch. As pointed out in detail in Section 13, this
the traditional leading-edge DSV is not present. is due to the large disparity between the time-scales
characterizing the airfoil pitch oscillations and those
14.2.3. The projected impact of compressibility characterizing the excitation-generated LCSs. It is thus
Carr et al. [201,206] show that the main features of tentatively speculated that periodic excitation will also be
compressibility during dynamic stall are that the suction e!ective in controlling dynamic stall under compressible
peak on the upper surface is attenuated due to local conditions. It is the opinion of the authors that evalu-
supersonic velocity in this region, and compressibility ation of the excitation under compressible conditions
e!ects can completely change the #ow behavior com- should be singled out as a future research priority in the
pared to that observed at low Mach numbers (Ma(0.3). context of dynamic stall control.
In future, these e!ects must be seriously considered be-
cause typical full-scale Mach numbers on a rotorcraft 14.2.4. Intermittent excitation
retreating blade in the vicinity of dynamic stall are in the While performing dynamic stall control with excita-
range 0.3}0.5. tion from the leading edge and #ap shoulder on the
Although the investigations cited above do not directly NACA 0015, a number of problems were identi"ed. First-
address the issue of compressibility, the static data of ly, excitation from the LE, while the airfoil was in the
Seifert and Pack [107], together with the discussions pre-stall regime, caused an increase in form drag due to
presented in this section and Section 13, can be used to a pressure increase in that region. It was ascertained that
536 D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

Fig. 62. Phase-locked (to the airfoil pitch phase) momentum de"cit measurements in the wake of a pitching NACA 0015 airfoil for
high-amplitude #ap-shoulder blowing and excitation, corresponding to Fig. 61 [109,112].

if the excitation was initiated and terminated near the while, simultaneously, containing the moment excur-
static stall angle, such that excitation was active only in sions. However, continuous excitation is clearly not e!ec-
the post-stall regime, increases in pre-stall form drag could tive for this incidence angle range because lift excursions
be eliminated. Secondly, with #ap-shoulder excitation or C , are small. The "gures clearly indicate that intermit-
* 
LE excitation entirely in the post-stall regime, lift excur- tent excitation, for this case, enables airfoil oscillation
sions C , which are of crucial in determining the max- between baseline and continuous excitation limits, while
* 
imum speed of a rotorcraft, could not be signi"cantly maintaining acceptable moment excursions. Excitation
controlled. It was ascertained that signi"cant control over initiation and termination e!ectively causes controlled
the lift excursions could be achieved, without diminishing dynamic stall and attachment, during the airfoil pitch
C , by employing excitation in an intermittent man- cycle. For this case, excitation is applied for approxim-
* 
ner [182], and two examples of this are presented below. ately 80% of the total pitching cycle. It was noted in
NACA 0015 lift and moment data for a deep stall Section 13 that the typical dimensionless time for dy-
scenario, where the mean incidence angle is a full 83 namic separation and attachment was q+20. Notwith-
above the static-stall angle, is shown is Figs. 64a and b. standing the added complexity of the pitch oscillations,
The graphs shows the dramatic e!ect of excitation in the these time-scales, now based on the airfoil chord, are
post-stall regime, where post-stall lift is increased by 50% consistent with the data presented in Fig. 64.
D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545 537

Fig. 63. Time-mean momentum de"cit measurements in the


wake of a pitching NACA 0015 airfoil for high-amplitude #ap-
shoulder blowing and excitation, corresponding to Figs. 61 and
62 [109,112].

An example of intermittent #ap-shoulder excitation,


with d "203, is presented in Figs. 65a and b. As before
D
(see Figs. 61), even though the maximum incidence angle
is 43 above the static stall angle, dynamic stall is light for
the baseline case, as can be best seen from the moment
excursions (Fig. 65b), due to the relatively high reduced Fig. 64. Intermittent leading-edge excitation on a NACA
pitch-rate (k"0.1) considered here. As in the leading- 0015 airfoil with the airfoil entirely in the post stall regime [182].
edge excitation case demonstrated above, continuous
excitation brings about signi"cant lift increases which are
maintained throughout the entire cycle. In contrast, how-
ever, instantaneous dC /da is not signi"cantly a!ected. 14.2.5. Post-stall unsteadiness
*
As before (Fig. 61), the moment excursions are not signi"- Finally, we conclude with a comment on unsteady
cantly a!ected, but excitation induces a further mean loads during dynamic stall. Greenblatt et al. [181] com-
nose-down pitching moment of C &0.05. Al- pared the phase-averaged aerodynamic quantities during
+  
though intrinsically di!erent to the previous case, the pitch-cycle to those measured instantaneously on
a similar control strategy was applied for #ap-shoulder a NACA 0015 airfoil for a"203#53sin(2pf t!903)
excitation. Here, excitation is initiated soon after the subjected to a variety of pitching frequencies for both
commencement of pitch-up, and terminated in the region baseline and leading-edge control scenarios. In general,
where the continuous excitation down-stroke lift data during the upstroke, the di!erences between phase-aver-
begins to increase. As indicated on the "gure, this control aged and instantaneous data were relatively small for
strategy brought about the maximum possible lift excur- both baseline and controlled cases. Near maximum inci-
sions, namely between the minimum baseline and dence and during the downstroke, however, large excur-
maximum excitation limits (C "0.86). It should be sions in baseline instantaneous data were present. This
* 
appreciated that the time-scales characterizing the airfoil was particularly evident from the lift data, which exhib-
pitch-rate are associated with the airfoil chord, whereas ited large excursions during the downstroke. Flow visual-
those governing separation and attachment are asso- ization data of Piziali [202] shows that dynamic stall has
ciated with the airfoil #ap-length. This di!erence should a strong three-dimensional character and this may
be accounted for when estimating the fraction of the cycle explain the large excursions evident here. Excitation, on
during which separation and attachment take place. the other hand, signi"cantly reduced the high level of
538 D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

illustrate important di!erences, as well as similarities,


between both BLC and acoustic excitation, and hydro-
dynamic excitation. The de#ected #ap was then further
used as a paradigm to isolate governing parameters
and illustrate the principles and mechanisms of separ-
ation control. Amongst the most important conclusions
was the identi"cation of the reduced frequency
F>,f X /; "O(1) as optimum to cause attachment
  
of an otherwise separated #ow to the #ap and the ap-
proximate range 2(F>(4 as optimum to prevent sep-
aration, providing that the upstream boundary layer was
fully turbulent. The generation and advection of LCSs,
transferring high momentum #uid to the #ap surface was
found to be the principle mechanism responsible for the
control of separation. A small amount of steady blowing,
superimposed on the excitation was found to have a det-
rimental e!ect on separation control, while the opposite
was true for a small amount of superimposed steady
suction.
A detailed analysis of two-dimensional airfoil data
showed that, for the overwhelming majority of investiga-
tions, the reduced frequency: 0.3)F>)4 was e!ective,
while the state of the upstream boundary layer (i.e.
laminar, transitional or turbulent) did not have a signi"-
cant e!ect. Unlike either steady blowing or acoustic
excitation, the e!ective momentum addition was in the
approximate range 0.01%(C (3%. Experimental
I
data showed that excitation at the shoulder of a de#ected
#ap was particularly e!ective, even up to large #ap defec-
tion angles d &603 and was e!ectively independent of
D
Reynolds number. A wide range of data showed conclus-
ively that excitation is much more e!ective and e$cient
Fig. 65. Intermittent #ap-shoulder excitation a NACA 0015 than steady blowing by anything up to two orders of
airfoil illustrating lift excursion control in the pre-stall regime magnitude. Moreover, for aerodynamically ine$cient
[180}182]. bodies, excitation could bring about large increases in lift
and/or reductions in drag. Additional important con-
clusions were that excitation e!ectiveness decreases as
unsteadiness associated with the downstroke data. These the distance between the actuator (or slot) and the separ-
observations highlight the di$culties associated with ation location increases; and signi"cant reduction of
two-dimensional modeling of baseline deep dynamic post-stall vortex-shedding and bu!et can be achieved.
stall (e.g. [203}205]). On the other hand, 2-D slot Three-dimensional e!ects were considered by accounting
excitation increases the spanwise coherence of the result- for the e!ect of pressure gradient, and thus boundary
ing LCSs, which signi"cantly reduces cycle-to-cycle layer development, on an in"nite swept wing. This result-
aberrations and validates the use of two-dimensional ed in a transformation from two-dimensional to three-
modeling. dimensional #ows and assisted in the analysis of #ow
control over delta wings.
The e!ects of surface curvature were investigated by
15. Conclusions and future research considering the #ow over a curved #ap, a `Stratford-
rampa, a Coanda cylinder and circular cylinders. Flow
The review presented in this paper outlined some of over the curved #ap emphasized the importance of the
the recent developments in separation control by peri- excitation location and showed that excitation too far
odic excitation, which is generated by hydrodynamic upstream was ine$cient for performance enhancement.
means. As a basis for the paper, the authors endeavoured Excitation of the #ow over the Stratford ramp, with
to illustrate the importance of the excited mixing layer for a boundary layer continually on the verge of separation,
understanding separation control, by showing how it revealed an e$cient transfer of momentum toward the
attaches to a de#ected #ap. An attempt was also made to wall. This resulted in signi"cant increases of the skin
D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545 539

friction accompanied by substantial decreases in the should be researched vigorously is the e!ect of curvature.
wake component of the boundary layer. On the Coanda For example, a combination in#ectional and centrifugal
cylinder, provisional data suggests that excitation inter- instability may embody performance enhancements
acts with both in#ectional and centrifugal instabilities, hitherto unexploited. A further, and possibly related, area
thus amplifying the perturbations more e!ectively. Here is that of examining large #ap de#ection. This may have
too, the location of excitation has a profound e!ect on profound implications for commercial aircraft. An addi-
the #ow. For #ow over circular cylinders, excitation had tional area of fundamental importance, that has not
profound e!ects of on lift, separation location and vortex received appropriate attention, is the e!ect of excitation
shedding frequency at sub-critical Reynolds numbers. location and orientation on separation control.
In this context, apparently curvature did not have a In closing it should be noted that the Wright brothers
signi"cant e!ect on optimum reduced frequencies, with #ew the "rst powered aircraft a year before Prandtl
F>"O(1). uncovered the secrets of the boundary layer. The prin-
"
An analysis of unexcited #ows with recirculation bub- ciple basis for design was their own empirical data. To be
bles, particularly for backward facing steps and axially successful, at least in the short term, control of separation
mounted blunt-faced cylinders, showed broad agreement by periodic excitation may follow a similar route. Theor-
of the Strouhal numbers based on shedding frequency: etical understanding and numerical prediction methods
0.5(f X /; (0.8 [126] and f h /; +0.08 [127]. will, hopefully, follow in due course.
   
Similar values for F>,f X /; and f h /; were as-
   
certained on the de#ected #ap providing that the min-
imum momentum (or amplitude) required to prevent Acknowledgements
bubble bursting, was supplied.
Dynamic attachment and separation (stall) was The authors are deeply indebted to Dr. A. Seifert for
studied by terminating or initiating excitation supplied to critically reviewing the draft and providing numerous
#ow over a static defected #ap. It was ascertained that insightful suggestions. The authors also wish to thank
dynamic stall from a static de#ected #ap bears signi"cant Dr. B. Nishri, Messrs. A. Darabi, D. Neuburger and
similarities to well-known rotorcraft-blade dynamic stall. T. Naveh for actively assisting in the preparation of much
The analysis further emphasized the principle of time- of the material presented here.
scale disparity between the dynamic stall vortex and the
structures which a!ect its control. Parametric studies,
aimed at controlling dynamic stall from airfoils oscillat-
References
ing at rotorcraft reduced frequencies, showed signi"cant
lift increases while simultaneously containing drag and [1] Maskell EC. Flow separation in three dimensions. RAE
moment excursions. Moreover, the generation and ad- Report Aero 2565, 1955.
vection of LCSs were not signi"cantly a!ected by the [2] Chang PK. Control of separation. New York: McGraw-
periodic pitching motion and thus reduced frequencies Hill, 1976.
F>, observed to be e!ective on static airfoils, were also [3] Telionis DP. Review } Unsteady boundary layers, separ-
e!ective for dynamic stall control. On the basis of the ated and attached. ASME J Fluids Eng 1979;101:
time-scale disparity, intermittent excitation was used to 29}43.
further improve performance. [4] Gad-el-Hak M, Bushnell DM. Separation control:
From an applications perspective, separation control review. J Fluid Engng 1991;113:5}30.
[5] Simpson RL. Turbulent boundary layer separation.
by excitation, is in its infancy. Provisional data presented
Annu Rev Fluid Mech 1989;21:205}34.
in this paper shows a wide range of possible applications, [6] Simpson RL. Aspects of turbulent boundary layer separ-
from high-lift devices to the elimination of control sur- ation. Prog Aerospace Sci 1996;32(5):457}521.
faces. To this end, areas which will demand attention in [7] Bradshaw P, Wong FYF. The reattachment and relax-
the near future include applications to UAVs, including ation of a turbulent shear layer. J Fluid Mech Part
low Reynolds number problems associated with MAVs. 1 1972;52:113}35.
Aircraft, or components of aircraft, which include in- [8] Kim J, Kline SJ, Johnston JP. Investigation of a reattach-
herent aerodynamic ine$ciencies should be targeted as ing turbulent shear layer: #ow over a backward-facing
areas of immediate research as the returns could be step. J Fluids Eng 1980;102 (9).
[9] Stratford BS. The prediction of separation of the turbu-
considerable. In the context of dynamic stall, future re-
lent boundary layer. J Fluid Mech 1959;5:1}16.
search should address the issue of compressibility.
[10] Anderson Jr JD. Introduction to #ight, 3rd ed. New
From a fundamental perspective is strongly recom- York: McGraw-Hill, 1989.
mended that experimental data be reported with a calib- [11] Wygnanski I. Boundary layer #ow control by periodic
ration of disturbance amplitude as well as frequency, so addition of momentum (Invited). AIAA Paper 97-2117
as to facilitate comparison of relative performance for 4th AIAA Shear Flow Conference, Snowmass Village,
various generic con"gurations. A priority area that CO, June 29}July 2, 1997.
540 D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

[12] Prandtl L. UG ber FluK ssigkeitsbeweung bei sehr kleiner- [33] Darabi A. On the mechanism of forced #ow reattach-
Reibung. Proceedings of Third International Mathemat- ment. PhD Thesis, Tel Aviv University, 2000.
ical Congress, Heidelberg, 1904, p. 484}91. [34] Neuburger D, Wygnanski I. The use of a vibrating ribbon
[13] Dryden HL. Fifty years of boundary layer theory and to delay separation on two-dimensional airfoils: somep-
experiment. Sciences 1955;121:375}80. reliminary observations. Presented at the Workshop on
[14] Lachmann GV. Boundary layer and #ow control. Its Unsteady Separated Flow, Air Force Academy, July 1987.
principles and application, vol. 1. New York: Pergamon [35] Roos FW, Kegelman JT. Control of coherent structures
Press, 1961. in reattaching laminar and turbulent shear layers. AIAA
[15] Lachmann GV. Boundary layer and #ow control. Its J 1986;24(12):1956}63.
principles and application, vol. 2. New York: Pergamon [36] Viets H, Piatt M, Ball M. Boundary layer control by
Press, 1961. unsteady vortex generation. J Wind Eng Ind Aerodyn
[16] Attinello JS. Design and engineering features of #ap 1987;7:135}44.
blowing installations. In: Lachmann GV, editor. Bound- [37] Zhou MD, Fernholz HH, Ma HY, Wu JZ, Wu, JM.
ary layer and #ow control. Its principles and application, Votrex capture by a two-dimensional airfoil with a small
vol. 1. New York: Pergamon Press, 1961. p. 463}515. iscillating leading-edge #ap, AIAA Paper 93-3266, 1993.
[17] Schubauer GB, Skramstad HK. Laminar boundary layer [38] Jacot D, Mabe J. Boeing active #ow control system.
oscillations and transition on a #at pate. NACA Rep 909, AIAA Paper 2000-2473, Fluids 2000, Denver CO, June
1948. 19}22, 2000.
[18] Mangiarotty RA. Control of laminar #ow in #uids by [39] Kiya M, Shimizu M, Mochizuki O, Ido Y, Ogura Y. Active
means of acoustic energy. US Patent 4,802,642, 1989. control of a turbulent separated #ow zone. XVIIIth
[19] Meier HU, Zhou MD. Method and apparatus for in- ICTAM, pp. 1}13, Haifa, Israel, 22}28 August 1992.
#uencing a laminar turbulent boundary layer transition [40] Seifert A, Bachar T, Koss T, Shepshelovich M, Wyg-
on bodies in #ow. US Patent 4,989,810, 1991. nanski I. Oscillatory blowing, a tool to delay boundary
[20] Collins FG, Zelenevitz J. In#uence of sound upon separ- layer separation. AIAA J 1993;31(11):2052}60.
ated #ow over wings. AIAA J 1975;13(3):408}10. [41] Seifert A, Bachar T, Wygnanski I, Koss D, Shepshelovich
[21] Brown GL, Roshko A. On density e!ects and large M. Oscillatory blowing, a tool to delay boundary layer
structure in turbulent mixing layers. J Fluid Mech separation. AIAA Paper 93-0440, 31st AIAA Aerospace
1974;64:775}816. Sciences Meeting, January 1993.
[22] Winant CD, Browand FK. Vortex pairing: the mecha- [42] Seifert A, Darabi A, Nishri B, Wygnanski I. The e!ects of
nism of turbulent mixing layer growth at moderate forced oscillations on the performance of airfoils. AIAA
Reynolds number. J Fluid Mech Part 2 1974;63: Paper 93-3264, AIAA Shear Flow Conference, July 1993.
237}56. [43] Guy Y, Morrow J, Mclaughlin T. Control of vortex
[23] Oster D, Wygnanski I, Dziomba B, Fiedler H. The e!ect breakdown on a delta wing by periodic blowing and
of initial conditions on the two-dimensional, turbulent suction. AIAA Paper 99-132, 37th Aerospace Sciences
mixing layer. In: Fiedler H, editor. Structure and mech- Meeting and Exhibit, Reno, NV, January 12}15, 1999.
anics of turbulence. Lecture Notes in Physics, vol. 75. [44] Amitay M, Smith BL, Glezer A. Aerodynamic #ow con-
Berlin: Springer, 1978. p. 48}64. trol using synthetic jet technology. AIAA Paper 98-0208,
[24] Ho CM, Huang LS. Subharmonics and vortex merging in 36th Aerospace Sciences Meeting and Exhibit, Reno, NV,
mixing layers. J Fluid Mech 1982;119:119}42. January 12}15, 1998.
[25] Browand FK, Ho CM. The mixing layer: an example [45] Siller HA, Fernholz H-H. Control of the separated #ow
of quasi two-dimensional turbulence. Journal de downstream of a two-dimensional fence. Euromech Col-
Mecanique, 1983. loquium 361 * Active Control of Turbulent Shear
[26] Ho CM, Huerre P. Perturbed free shear layers. Annu Rev Flows, Berlin, 17}19 March 1997.
Fluid Mech 1984;16:365}424. [46] Wlezien RW, Horner GC, McGowan AR, Padula SL,
[27] Wygnanski I, Petersen RA. Coherent motion in excited Scott MA, Silcox RJ, Simpson JO. Aircraft morphing
free shear #ows. AIAA J 1987;25(2):201}13. program. SPIE Proceedings, vol. 3326. Paper C: 3326-
[28] Katz Y, Nishri B, Wygnanski IJ. The delay of turbulent 20, 1998, p. 176}87.
boundary layer separation by oscillatory active control. [47] Hassan AA. Numerical simulations and potential ap-
AIAA Paper 98-0975, 1989. plications of zero-mass jets for enhanced rotorcraft
[29] Katz Y, Nishri B, Wygnanski I. The delay of turbulent aerodynamic performance. AIAA Paper 98-0211, 36th
boundary layer separation by oscillatory active control. Aerospace Sciences Meeting and Exhibit, Reno, NV,
Phys Fluids 1989;1:179}81. January 12}15, 1998.
[30] Nishri B, Wygnanski I. On #ow separation and its con- [48] Chernyshenko SI. Stabilization of trapped vortices by
trol. Computational methods in applied sciences. New alternating blowing suction. Phys Fluids 1995;4:802}7.
York: Wiley, 1996. p. 471}82. [49] Goldstein ME. Generation of instability waves in #ows
[31] Nishri B, Wygnanski I. E!ects of periodic excitation on separating from smooth surfaces. J Fluid Mech
turbulent separation from a #ap. AIAA J 1998;36(4): 1984;145:71}94.
547}56. [50] Donovan JF, Kral LD, Cary AW. Active #ow control
[32] Huang LS, Maestrello L, Bryant TD, Separation control applied to an airfoil. AIAA Paper 98-0210, 36th Aero-
over airfoils at high angles of attack by sound emanating space Sciences Meeting and Exhibit, Reno, NV, January
from the surface. AIAA Paper 87-1261, 1987. 12}15, 1998.
D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545 541

[51] Wu JZ, Vakili AD, Wu JM. Review of the physics of [69] Modi VJ, Hill SS, Yokomizo T. Drag reduction of trucks
enhancing vortex lift by unsteady excitation. Prog Aero- through boundary-layer control. J Wind Eng Ind Aero-
space Sci 1998;28:73}131. dyn 1995;54:583}94.
[52] Wu JZ, Lu XY, Denny AG, Fan MF, Wu JM. Post-stall [70] Zaman KBMQ, Bar-Sever A, Mangalam SM. E!ect of
#ow control on an airfoil by local unsteady forcing. acoustic excitation on the #ow over a low-Re airfoil.
J Fluid Mech 1998;371:21}58. J Fluid Mech 1987;182:127}48.
[53] McMichael JM. MEMS and challenges of #ow control. [71] Ahuja KK, Whipkey RR, Jones GS. Control of turbulent
Invited oral presentation, AIAA 3rd Shear #ow control boundary layer #ow by sound. AIAA Paper 83-0726,
conference, Orlando, FL, July 1993. 1983.
[54] Friedman PP, Millott T. Vibration reduction in rotor- [72] Ahuja KK, Burrin RH. Control of #ow separation by
craft using active control } A comparison of various sound. AIAA Paper 84-2298, 1984.
approaches. AIAA J Guidance Control Dyn 1995;18(4): [73] Marchmann JF, Sumantran V, Shaefer CG. Acoustic
664}73. and turbulence in#uences on stall hysteresis. AIAA
[55] Ramiz MA, Acharya M. Detection of #ow state in an J 1987;25(1):50}1.
unsteady separating #ow. AIAA J 1992;30(1):117}23. [74] Nishioka M, Asai M, Yoshida S. Control of separation by
[56] Ha CM. Neural networks approach to AIAA Aircraft acoustic excitation. AIAA J 1990;28(11):1909}15.
Control design challenge. AIAA Paper 91-2672, 1991. [75] Zaman KBMQ. E!ect of acoustic excitation on stalled
[57] Faller WE, Schreck SJ. Neural networks: applications #ows over an airfoil. AIAA J 1992;30(6).
and opportunities in aeronautics. Prog Aerospace Sci [76] Zaman KBMQ, McKinzie DJ, Rumsey CL. A natu-
1996;32(5):433}56. ral low-frequency oscillation of #ow over an airfoil
[58] D'Andrea R, Behnken RL, Murray RM. Active control of near stalling conditions. J Fluid Mech 1989;202:
an axial #ow compressor via pulsed air injection. J Tur- 403}42.
bomachinery 1997;119(4):742}52. [77] Liepmann HW, Laufer J. Investigation of free turbulent
[59] Nelson C, Koga D, Eaton J. Unsteady separated #ow mixing. NACA TN 1257, 1947.
behind an oscillating two-dimensional #ap. AIAA Paper [78] Rajaratnam N. Turbulent jets. Amsterdam: Elsevier Sci-
89-0288, 27th Aerospace Sciences Meeting, Reno, NV, enti"c Publishing Company, 1976.
January 9}12, 1989. [79] Oster D, Wygnanski I. The forced mixing layerbetween
[60] Flatt J. The history of boundary layer control research in parallel streams. J Fluid Mech 1982;123:91.
the United States of America. In: Lachmann GV, editor. [80] Cohen J. Instabilities and resonances in turbulent free
Boundary layer and #ow control. Its principles and ap- shear #ows. PhD thesis, University of Arizona, Tucson,
plication, vol. 1. New York: Pergamon Press, 1961. 1985.
p. 122}43. [81] Nishri B. On the dominant mechanisms governing active
[61] Schlichting H. Boundary layer theory. New York: control of separation. PhD thesis, Tel Aviv University,
McGraw-Hill, 1979. 1995 [in Hebrew].
[62] Betz A. History of boundary layer control in Germany. [82] Bachar T. Alternating #ow generator. Patent Pending
In: Lachmann GV, editor. Boundary layer and #ow con- 1998.
trol. Its principles and application, vol. 1. New York: [83] Smith BL, Glezer A. The formation and evolution of
Pergamon Press, 1961. p. 1}20. synthetic jets. Phys Fluids 1998;10(9):2289}97.
[63] Head MR. History of research on boundary layer control [84] Seifert A, Eliahu S, Greenblatt D, Wygnanski I. Use of
for low drag in UK. In: Lachmann GV, editor. Boundary piezoelectric actuators for airfoil separation control.
layer and #ow control. Its principles and application, vol. AIAA J 1998;36(8):1535}7.
1. New York: Pergamon Press, 1961. p. 104}21. [85] DARPA. Proposer Information Pamphlet for BAA 98-
[64] Carriere P, Eichelbrenner EA. Theory of #ow reattach- 19: Enabling thechnologies for micro adaptive #ow con-
ment by a tangential jet discharging against a strong trol, 1998. http://www.darpa.mil/tto/mafc/pip..html,
adverse pressure gradient. In: Lachmann GV, editor. Accessed on 17/03/99.
Boundary layer and #ow control. Its principles and ap- [86] LoK fdahl L, Gad-el-Hak M. MEMS applications in turbu-
plication, vol. 1. New York: Pergamon Press, 1961. lence and #ow control. Prog Aerospace Sci
p. 209}31. 1999;35(2):101}203.
[65] Attinello JS. Flow control } the integration of power [87] Reisenthel PH, Nagib HM, Koga DJ. Control of separ-
plant and airframe for future aircraft. Bureau of Aero- ated #ows using forced unsteadiness. AIAA Paper 85-
nautics Research Division, Report No. DR-1745, 1952. 0556, AIAA Shear Flow Control Conference, Boulder,
[66] Poisson-Quinton Ph. Recherches theH oriques et expeH ri- Colorado, 12}14 March 1985.
mentales sur le contro( l de couche limits. 7th Congress of [88] Wygnanski I. Method and apparatus for delaying the
Applied Mechanics, London, September 1948. separation of #ow from a solid surface. US Patent
[67] Seifert A, Darabi A, Wygnanski I. Delay of airfoil stallby 5,209,438,1993.
periodic excitation. AIAA J Aircraft 1996;33(4):691}8. [89] Nagib H, Reisenthel P, Koga D. On the dynamical scal-
[68] Poisson-Quinton Ph, Lepage L. Survey of French re- ing of forced unsteady separated #ows. AIAA Paper
search on the control of boundary layer and circulation. 85-0553, 1985.
In: Lachmann GV, editor. Boundary layer and #ow con- [90] Gad-el-Hak M, Blackwelder RF. Control of the dis-
trol. Its principles and application, vol. 1. New York: crete vortices from a delta wing. AIAA J 1987;25(8):
Pergamon Press, 1961. p. 21}73. 1042}9.
542 D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

[91] Guy Y, Morrow J, Mclaughlin T. Pressure measurement [108] Greenblatt D, Wygnanski I. The use of periodic excita-
and #ow "eld visualization on a delta wing with periodic tion to enhance airfoil performance at low Reynolds
blowing and suction. AIAA Paper 99-4178, AIAA Atmo- numbers. AIAA J Aircraft 2001; to appear.
spheric Flight Mechanics Conference and Exhibit, Port- [109] Greenblatt D. Dynamic stall control by oscillatory exci-
land, OR, August 9}11, 1999. tation. PhD Thesis, Tel-Aviv University, November
[92] Darabi A, Lourenco L, Wygnanski I. On #ow reattach- 1999.
ment by periodic excitation. Eighth European Turbu- [110] Bachar T, Wygnanski I, Ashpis D. Active control of
lence Conference, Advances in Turbulence VIII, Bar- separation in the presence of high freestream turbulence.
celona, Spain, 2000, pp. 201}4. Abstract JK.01, Control of Turbulent Flows, APS Meet-
[93] Greenblatt D, Darabi A, Nishri B, Wygnanski I. Some ing 1998.
factors a!ecting stall control with particular emphasis [111] Greenblatt D, Wygnanski I. Dynamic stall control by
on dynamic stall. AIAA Paper 99-3504, 30th AIAA oscillatory forcing. AIAA Paper 98-0676, 36th Aerospace
Fluid Dynamics Conference, Norfolk, VA, 28 June}1 Sciences Meeting and Exhibit, Reno, NV, January 12}15,
July 1999. 1998.
[94] Viets H, Piatt M, Ball M. Unsteady wing boundary layer [112] Greenblatt D, Wygnanski I. Parameters a!ecting dy-
energization. AIAA Paper 79-1631, 1979. namic stall control by oscillatory excitation. AIAA Paper
[95] Neuburger D. An active delay of separation on two- 99-3121, 17th AIAA Applied Aerodynamics Conference,
dimensional airfoils. MSc thesis, Tel Aviv Univer- Norfolk, VA, 28 June}1 July 1999.
sity,1989 [in Hebrew]. [113] Wygnanski I. On active #ow control: a personal pespec-
[96] Bar-Sever A. Separation control on an airfoil by periodic tive, 39th Israel Annual Conference on Aerospace
forcing. AIAA J 1989;27(6):820}1. Sciences, February 1999.
[97] Hsiao F-B, Liu C-F, Shyu J-Y. Control of wall-separated [114] Delery JM. Shock wave/turbulent boundary layer inter-
#ow by internal acoustic excitation. AIAA J 1990;28(8): action and its control. Prog Aerospace Sci 1985;22:209}80.
1440}6. [115] Wallis RA, Stuart CM. On the control of shock-induced
[98] Shepshelovich M, Koss T, Wygnanski I, Seifert A. Active boundary layer separation with discrete air jets. ARC CP
#ow control on low Re airfoils. AIAA Paper 89-0538, No. 595, 1962.
January 1989. [116] McCormick DC. Shock-boundary layer interaction with
[99] Chang RC, Hsiao F-B, Shyu RN. Forcing level e!ects of low pro"le vortex generators and passive cavity. AIAA
internal acoustic excitation on the improvement of airfoil Paper 92-0064, 30th Aerospace Sciences Meeting, Reno,
performance. AIAA J 1992;29(5):823}9. NV, 6}9 January 1992.
[100] Hsiao F-B, Shyu R-N, Chang RC. High angle of attack [117] Hites M, Nagib H, Sytsma B, Wygnanski I, Seifert A,
airfoil performance improvement by internal acoustic Bachar T. Lift enhancement using pulsed blowing at
excitation. AIAA J 1994;32(3):655}7. compressible #ow conditions. 50th Annual Meeting of
[101] Darabi A. The e!ect of oscillatory blowing on a stalling The APS's Division of Fluid Dynamics, San Francisco,
airfoil. MSc thesis, Tel Aviv University, 1995 [in He- 23}25 November 1997.
brew]. [118] Roshko A. On the drag and shedding frequency of two-
[102] Tinapp F, Stumpf E, Nitche W. Separation control on dimensional blu! bodies. NACA TM 3169, 1954.
a high-lift con"guration. Euromech Colloquium 361 [119] McLean JD, Crouch JD, Stoner RC, Sakurai S, Seidel
} Active Control of Turbulent Shear Flows, Berlin, 17}19 GE, Feifel WM, Rush HM. Study of the application of
March 1997. separation control by unsteady excitation to civil trans-
[103] Smith D, Amitay M, Kibens V, Parekh D, Glezer A. port aircraft. NASA/CR-1999-209338, 1999.
Modi"cation of lifting body aerodynamics using syn- [120] Lin JC. Control of turbulent boundary-layer separations
thetic jet actuators. AIAA Paper 98-0209, 36th Aerospace using micro-vortex generators. AIAA Paper 99-3404,
Sciences Meeting Exhibit, Reno, NV, January 12}15, 30th AIAA Fluid Dynamics Conference, Norfolk, VA, 28
1998. June}1 July 1999.
[104] Seifert A, Pack LG. Oscillatory control of separation at [121] Nishri B. The delay of turbulent boundary layer separ-
high Reynolds numbers. AIAA Paper 98-0214, 36th ation by oscillatory active control. MSc thesis, Tel Aviv
Aerospace Sciences Meeting and Exhibit, Reno, NV, University, Israel, 1989 [in Hebrew].
January 12}15, 1998. [122] Pack LG, Seifert A. Periodic excitation for jet vectoring
[105] Naveh T, Seifert A, Tumin A, Wygnanski I. Sweep e!ect and enhanced spreading. AIAA Paper 99-0672, 37th
on parameters governing control of separation by peri- AIAA Aerospace Sciences Meeting and Exhibit, Reno,
odic excitation. AIAA J 1998;35(3):510}2. Nevada, January 11}14, 1999.
[106] Erk P, Graichen K, Fernholz H-H. Separation control on [123] Schmidt GS, Mueller TJ. Analysis of low Reynolds num-
an airfoil using strong acoustic perturbations. Euromech ber separation bubbles using semi-empirical methods.
Colloquium 361 } Active Control of Turbulent Shear AIAA J 1989;27(8):993}1001.
Flows, Berlin, 17}19 March 1997. [124] Liebeck RH. Laminar separation bubbles and airfoil de-
[107] Seifert A, Pack LG. Oscillatory excitation of unsteady sign at low Reynolds numbers. AIAA Paper 92-2735-CP,
compressible #ows over airfoils at #ight Reynolds 1992.
numbers. AIAA Paper 99-0925, 37th Aerospace [125] Horton HP, Laminar separation bubbles in two and
Sciences Meeting and Exhibit, Reno, NV, January 11}14, three dimensional incompressible #ow. PhD thesis, Uni-
1999. versity of London, London, 1968.
D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545 543

[126] Mabey DG. Analysis and correlation of data on pressure [145] Peterka JA, Richardson PD. E!ects of sound on separ-
#uctuations in separated #ow. AIAA J Aircraft 1972;9(9): ated #ow. J Fluid Mech 1969;37:265}87.
642}5. [146] Blevins RD. The e!ect of sound on vortex shedding from
[127] Sigurdson LW, Roshko A. Controlled unsteady excita- cylinders. J Fluid Mech 1985;161:217}37.
tion of a reattaching #ow. AIAA Paper 85-0552, AIAA [147] Williams D, Acharya M, Bernhardt J, Yang P. The mech-
Shear Flow Control Conference, Boulder Colorado, anism of #ow control on a cylinder with the unsteady
12}14 March 1985. bleed technique. AIAA Paper 91-0039, 29th Aerospace
[128] Sigurdson LW. The structure control of a turbulent reat- Sciences Meeting and Exhibit, Reno, NV, 1991.
taching #ow. J Fluid Mech 1995;298:139}65. [148] Schewe G. On the forced #uctuations acting on acircular
[129] Levi E. A universal stouhal law. ASCE J Eng Mech cylinder in a cross-#ow from sub-critical to trans-critical
1983;109(3):718}27. Reynolds number. J Fluid Mech 1983;133:265}85.
[130] Kiya M, Mochizuki O, Suzuzki N. Separation control by [149] Pal D, Sinha K. Controlling an unteady separating
vortex projectiles. AIAA Paper 99-3400, 30th AIAA boundary layer on a cylinder with an active compliant
Fluid Dynamics Conference, Norfolk, VA, 28 June}1 wall. AIAA Paper 97-0212, 35th Aerospace Sciences
July 1999. Meeting, Reno, NV, 1997.
[131] Huppertz A, Janke G. Some new results on the control of [150] Heine C, Spiess MC, MoK ser M, Fiedler HE. The in#uence
the #ow over a backward-facing step. Euromech Collo- of feedback control on the vortex dynamics in a turbulent
quium 361 } Active Control of Turbulent Shear Flows, cylinder wake. Euromech Colloquium 361 } Active Con-
Berlin, 17}19 March 1997. trol of Turbulent Shear Flows, Berlin, 17}19 March 1997.
[132] Eaton JK, Johnson JP. A review of research on subsonic [151] Liu WP, Brodie G. A demonstartion of active turbulence
turbulent #ow reattachment. AIAA J 1981;19(9): transition with MEMS sensors. Abstract YC06.06,
1093}100. American Physical Society Centennial Meeting Pro-
[133] Troutt TR, Scheelke B, Norman TR. Organized struc- gram, March 20}26, Atlanta, GA, 1999.
tures in a reattaching separated #ow"eld. J Fluid Mech [152] Naveh T. The e!ect of sweep on separation control
1984;143:413}27. over an airfoil. MSc thesis, Tel Aviv University, 1997
[134] Hasdai G. Delay of airfoil stall by periodic oscillatory [in Hebrew].
excitation: NACA 0012. Final Year Research Report, Tel [153] McCormick BW. Aerodynamics, aeronautics and #ight
Aviv University, 1999 (also see Hasdai G. Delay of airfoil mechanics. New York: Wiley, 1979.
stall by periodic oscillatory excitation on a NACA 0012 [154] Wygnanski I. Some new observations a!ecting the con-
Aifoil. Mechonot, Issue No. 28, pp. 32}35, March 2000.) trol of separation by periodic excitation (Invited). AIAA
[135] Kiya M, Shimizu M, Mochizuki O. Sinusoidal forcing of Paper 2000-2314, Fluids 2000, Denver CO, June 19}22,
a turbulent separation bubble. J Fluid Mech 1997;342: 2000.
119}39. [155] Guy Y, Morrow J, Mclaughlin T. Parametric investiga-
[136] Shimizu M, Kiya M, Mochizuki O, Ido Y. Response of an tion of the e!ects of active #ow control on the normal
axisymmetric separation bubble to sinosoidal forcing. force of a delta wing. AIAA Paper 2000-0549, 38th Aero-
Trans JSME 1993;B59:721}7 [in Japanese]. space Sciences Meeting and Exhibit, Reno, NV, January
[137] Patel VC, Sotiropoulos F. Longitudinal curvature e!ects 12}15, 2000.
in turbulent boundary layers. Prog Aerospace Sci [156] Squire HB. On the stability of three-dimensional distri-
1997;33(1-2):1}70. bution of viscous #uid between parallel walls. Proc Roy
[138] Stratford BS. An experimental #ow with zero skin friction Soc London A 1933;142:621}8.
throughout its region of pressure rise. J Fluid Mech [157] Johnson JP, Nishi M. Vortex generator jets * a means
1959;5:17}35. for #ow separation control. AIAA J 1990;28:989.
[139] Liebeck RH. On the design of subsonic airfoils for high [158] McManus KR, Joshi PB, Legner HH, Davis SJ. Active
lift. AIAA 9th Fluid and Plasma Dynamics Conference, control of aerodynamic stall using pulsed jet actuators.
San Diego, California, 14}16 July 1976. AIAA Paper 95-2187, 26th AIAA Fluid Dynamics Con-
[140] Elsberry K. The Stratford ramp. MSc thesis, University ference, San Diego, CA, June 19}22, 1995.
of Arizona, Tucson, 1996. [159] McManus K, Magill J. Separation control in incompress-
[141] Elsberry K, Loe%er J, Zhou MD, Wygnanski I. An ible and compressible #ows using pulsed jets. AIAA Pa-
experimental study of a boundary layer that is main- per 96-1948, 27th AIAA Fluid Dynamics Conference,
tained on the verge of separation. J Fluid Mech 2000; in New Orleans, LA, June 17}20, 1996.
press. [160] McManus K, Magill J. Airfoil performance enhancement
[142] Spangenberg WR, Roland NE. Measurements in a turbu- using pulsed jet separation control. AIAA Paper 97-1971,
lent boundary layer maintained in a nearly separating 28th AIAA Fluid Dynamics Conference, Snowmass Vill-
condition. In: Sovran G, editor. Fluid mechanics of inter- age, CO, June 29}July 2, 1997.
nal Flow. Amsterdam: Elsevier Publishing Company, [161] Cannan JW. Seeing more, and risking less, with UAVs.
1967. Aerospace America, October 1999. p. 26}34.
[143] Neuendorf R, Wygnanski I. On a turbulent wall jet [162] Shyy W, Berg M, Ljungqvist D. Flapping #exible wings
#owing over a circular cylinder. J Fluid Mech 1999;381: for biological and micro air vehicles. Prog Aerospace Sci
1}25. 1999;35(5):455}505.
[144] Neuendorf R. Turbulent wall jet along a convex curved [163] Nordwall BD. Micro air vehicles hold great promise,
surface. PhD thesis, Technical University, Berlin, 2000. challenges. Aviation Week Space Technol 1997;68}70.
544 D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545

[164] Englar RJ, Hudson GG. Development of advanced circu- [184] Wikens RH, Wind-tunnel investigation of dynamic stall
lation control wing high-lift airfoils. AIAA J Aircraft of NACA 0018 airfoil oscillating in pitch. Aero Note
1984;21:476. NAE-AN-27, NRC24262, National Research Council of
[165] Shyy W, Jenkins DA, Smith RW. Study of adaptive shape Canada, Ottawa, Canada, February 1985.
airfoils at low Reynolds number in oscillatory #ows. [185] Stepniewski WZ, Keys CN. Rotary-wing aerodynamics.
AIAA J 1997;35:1545}8. New York: Dover Publications Inc., 1984.
[166] Seifert A, Bachar T, Wygnanski I, Kariv A, Cohen H, [186] McCloud III KL, Hall LP, Brady JA. Full-scale wind
Yoeli R. Application of active separation control to a tunnel tests of blowing boundary layer control applied to
small unmanned air vehicle. AIAA J Aircraft 1999;36(2): helicopter rotor. NASA TN D-335, 1960.
474}7. [187] Addington GA, Schreck SJ, Luttges MW. Static and
[167] Rott N. Unsteady viscous #ow in the vicinity of a stagna- dynamic #ow "eld development about a porous suction
tion point. Quart Appl Math 1956;13:444}51. surface wing. AIAA Paper 92-2628-CP, 1992.
[168] Sears WR. Some recent developments in airfoil theory. [188] Karim MA, Acharya M. Suppression of dynamic-stall
J Aeronaut Sci 1956;23:490}9. vortices over pitching airfoils by leading-edge suction.
[169] Moore FK, On the separation of the unsteady laminar AIAA J 1994;32(8):1647}55.
boundary layer. In: GoK rtler H, editor. Boundary layer [189] Luttges MW, Robinson MC, Kennedy DA. Control of
research. Berlin: Springer, 1958, p. 296}310. unsteady separated #ow structures on airfoils. AIAA
[170] Williams III JC, Johnson WD. Note on unsteady bound- Paper 85-0531, AIAA Shear Flow Control Conference,
ary layer separation. AIAA J 1974;12:1427}9. 1985.
[171] Walker, JDA. Unsteady boundary layer separation [190] Weaver D, McAlister KW, Tso J. Suppression of dy-
in two dimensional #ows. Workshop on Supermaneu- namic stall by steady and pulsed upper-surface blowing.
verability: physics ofseparated #ows at high angle of AIAA Paper 98-2413,16th AIAA Applied Aerodynamics
attack, Lehigh University, Bethlehem, PA, April 9}10, Conference, Albuquerque, NM, June 15}18, 1998.
1992. [191] Green RB, Gilbraith RAMcD. An investigation of dy-
[172] Sears WR, Telionis DP. Unsteady boundary-layer separ- namic stall through the application of leading edge
ation. In: Eichelbrenner EA, editor. Recent research on roughness. Paper No. 137, 18th European Rotorcraft
unsteady boundary layers, vol. 1. Quebec, Canada: Forum, Avignon, France, 1992.
Presse de l'Universite Laval, 1972; p. 404}42. [192] Wilder MC, Chandrasekhara MS, Carr LW. Transition
[173] Sears WR, Telionis DP. Boundary layer separation of e!ects on compressible dynamic stall of transiently
unsteady #ow. J Appl Math 1975;28:215}35. pitching airfoils. AIAA Paper 93-2978, AIAA 24th Fluid
[174] Van Dommelen L, Shen SF. The spontaneous generation Dynamics Conference, Orlando, FL, July 6}9,
of the singularity in a separating laminar boundary layer. 1993.
J Comput Phys 1980;38:125}40. [193] Chandrasekhara MS, Wilder MC, Carr LW. Boundary
[175] Choudhuri PG, Knight DD, Visbal MR. Two-dimen- layer tripping studies of compressible dynamic stall #ow.
sional unsteady leading-edge separation on a pitching AIAA Paper 94-2340, 25th AIAA Fluid Dynamics Con-
airfoil. AIAA J 1994;32(4):673}81. ference, Colorado Springs, CO, June 20}23, 1994.
[176] Shih C, Lourenco LM, Krothapalli A. Investigation of [194] Carr LW, McAlister KW. The e!ect of a leading-edge slat
#ow at leading trailing edges of pitching-up airfoil. AIAA on the dynamic stall of an oscillating airfoil. AIAA Paper
J 1995;33(8):1369}76. 83-2533, AIAA/AHS Aircraft Design System and Opera-
[177] Despard RA, Miller JA. Separation in oscillating laminar tions Meeting, 1983.
boundary layer #ows. J Fluid Mech 1971;47:21}31. [195] Yu YH, Lee S, McAlister KW, Tung C, Wang C. Dy-
[178] Carr LW. Progress in the analysis and prediction of namic stall control for advanced rotorcraft application.
dynamic stall. AIAA J Aircraft 1988;25(1):6}17. AIAA J 1995;33(2):289}95.
[179] Greenblatt D, Darabi A, Nishri B, Wygnanski I. Separ- [196] Carr LW, Chandrasekhara MS, Wilder MC, Noonan
ation control by periodic addition of momentum with KW. The e!ect of compressibility on suppression of
particular emphasis on dynamic stall. American Heli- dynamic stalls using a slotted airfoil. AIAA Paper 98-
copter Society Paper T3-4, Gifu, Japan, April 21}23, 0332, 36th Aerospace Sciences Meeting and Exhibit,
1998. Reno, NV, January 12}15, 1998.
[180] Greenblatt D, Nishri B, Darabi A, Wygnanski I. Dy- [197] Freymuth P, Jackson S, Bank W. Toward dynamic separ-
namic stall control by oscillatory addition of momentum. ation without dynamic stall. Exp Fluids 1989;7:187}96.
Part I: Mechanism. AIAA J Aircraft 2000a; in press. [198] Chandrasekhara MS, Wilder MC, Carr LW. Control of
[181] Greenblatt D, Wygnanski I. Dynamic stall control by #ow separation using adaptive airfoils. AIAA Paper No
oscillatory addition of momentum. Part II: Parametric 97-0655, 1997.
study, AIAA J Aircraft 2000b; in press. [199] Greenblatt D, Seifert A, Wygnanski I. Dynamic stall
[182] Greenblatt D, Neuburger D, Wygnanski I. Dynamic stall management by oscillatory forcing. Euromech Collo-
control by intermittent excitation. AIAA J Aircraft, 2001; quium 361: Active Control of Turbulent Shear Flows,
to appear. 1997.
[183] Herbst, W. Supermaneuverability. AFOSR/FJSRL/Uni- [200] McCroskey WJ, McAlister KW, Carr LW, Pucci SL. An
versity of Colorado Workshop on unsteady Separated experimental study of dynamic stall on advanced airfoil
Flows, US Air Force Academy, Colorado Springs, CO, sections; vol. 1. Summary of the experiment. NASA TM
August 1983. 84245, July 1982.
D. Greenblatt, I.J. Wygnanski / Progress in Aerospace Sciences 36 (2000) 487}545 545

[201] Carr LW, Chandrasekhara MS. Compressibility e!ects [210] Greenblatt D, Seifert A, Wygnanski I. Dynamic stall
on dynamic stall. Prog Aerospace Sci 1996;32:523}73. control by oscillatory forcing. Patent Pending, USA,
[202] Piziali RA. 2-D and 3-D oscillating wing aerodynamics Europe & Japan, 1998.
for a range of angles of attack including stall. NASA TM [211] McAlister KW, Pucci SL, McCroskey WJ, Carr LW. An
4632, 1994. experimental study of dynamic stall on advanced airfoil
[203] Ekaterinaris JA, Menter FR. Computation of oscillating sections, vol. 2. Pressure and force data. NASA TM
airfoil #ows with one- and two-equation turbulence mod- 84245, 1982.
els. AIAA J 1994;32(12):2359}65. [212] Magill JC, McManus KR. Control of dynamic stall using
[204] Ko S, McCroskey WJ. Computations of unsteady separ- pulsed vortex generator jets. AIAA Paper 98-0675, 36th
ating #ows over an oscillating airfoil. AIAA J 1997;35(7): Aerospace Sciences Meeting and Exhibit, Reno, NV,
1235}8. Januarny 12}15, 1998.
[205] Ekaterinaris JA, Platzer MF. Computational prediction [213] Mullins J. Palmtop planes. New Scientist, April 1997.
of airfoil dynamic stall. Prog Aerospace Sci 1997;33: [214] Shih C, Lourenco L, Van Dommelen L, Krothapalli A.
759}846. Unsteady #ow past an airfoil pitching at a constant rate.
[206] Carr LW, Chandrasekhara NJ, Brock NJ. Quantitative AIAA J 1992;30(5):1153}61.
study of compressible #ow on an oscillating airfoil. AIAA [215] Weisbrot I. A mixing layer excited by two frequencies.
J 1994;31(4):892}8. MSc Thesis, Tel Aviv University, 1985.
[207] Carr LW, McCroskey WJ. A review of recent advances in [216] Wu JM, Lu XY, Denny AG, Fan MF, Wu JZ. Post-stall
computational and experimental analysis of dynamic lift enhancement on an airfoil by local unsteady control.
stall. IUTAM Symposium on Fluid Dynamics of High Part I. Drag and pressure characteristics. AIAA Paper
Angle of Attack, Tokyo, Japan, September 13}17, 1992. 97-2063, 4th AIAA Shear Flow Conference, Snowmass
[208] Chandrasekhara MS, Wilder MC, Carr LW. Unsteady Village, CO, June 29}July 2, 1977.
stall control using dynamically deforming airfoils. AIAA [217] Wu JZ, Lu XY, Wu JM. Post-stall lift enhancement on an
J 1998;36(10):1792}800. airfoil by local unsteady control. Part II. Mode competi-
[209] Fiedler HE, Fernholz H-H. On management and control tion and vortex dynamics. AIAA Paper 97-2063, 4th
of turbulent shear #ows. Prog Aerospace Sci 1990;27: AIAA Shear Flow Conference, Snowmass Village, CO,
305}87. June 29}July 2, 1977.

Vous aimerez peut-être aussi