Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

FL47CH20-Corke ARI 16 September 2014 12:22

Review in Advance first posted online


V I E W
E on September 25, 2014. (Changes may
R

still occur before final publication

S
online and in print.)

C E
I N

A
D V A

Dynamic Stall in Pitching


Airfoils: Aerodynamic Damping
and Compressibility Effects
Thomas C. Corke and Flint O. Thomas
Annu. Rev. Fluid Mech. 2015.47. Downloaded from www.annualreviews.org

Aerospace and Mechanical Engineering Department, Institute for Flow Physics and Control,
University of Notre Dame, Notre Dame, Indiana 46556; email: tcorke@nd.edu, fthomas@nd.edu
by Tulane University on 10/14/14. For personal use only.

Annu. Rev. Fluid Mech. 2015. 47:479–505 Keywords


The Annual Review of Fluid Mechanics is online at stall hysteresis, unsteady aerodynamics, stall flutter
fluid.annualreviews.org

This article’s doi: Abstract


10.1146/annurev-fluid-010814-013632
Dynamic stall is an incredibly rich fluid dynamics problem that manifests
Copyright  c 2015 by Annual Reviews. itself on an airfoil during rapid, transient motion in which the angle of in-
All rights reserved
cidence surpasses the static stall limit. It is an important element of many
manmade and natural flyers, including helicopters and supermaneuverable
aircraft, and low–Reynolds number flapping-wing birds and insects. The
fluid dynamic attributes that accompany dynamic stall include an eruption of
vorticity that organizes into a well-defined dynamic stall vortex and massive
excursions in aerodynamic loads that can couple with the airfoil structural
dynamics. The dynamic stall process is highly sensitive to surface roughness
that can influence turbulent transition and to local compressibility effects
that occur at free-stream Mach numbers that are otherwise incompressible.
Under some conditions, dynamic stall can result in negative aerodynamic
damping that leads to limit-cycle growth of structural vibrations and rapid
mechanical failure. The mechanisms leading to negative damping have been
a principal interest of recent experiments and analysis. Computational fluid
dynamic simulations and low-order models have not been good predictors
so far. Large-eddy simulation could be a viable approach although it remains
computationally intensive. The topic is technologically important owing to
the desire to develop next-generation rotorcraft that employ adaptive rotor
dynamic stall control.

479

Changes may still occur before final publication online and in print
FL47CH20-Corke ARI 16 September 2014 12:22

1. BACKGROUND
Largely because of the helicopter application, dynamic stall has been an active research topic
in fluid dynamics for more than 60 years. However, its foundation dates back nearly 90 years,
with the origins of classical unsteady aerodynamics that include Wagner (1925), Glauert & Holl
(1929), and Theodorsen (1935), who developed analytic solutions to incompressible flow around
thin airfoils undergoing time-dependent motions. Other early contributions were from Küssner
(1936), von Kármán & Sears (1938), Sears (1941), and Lomax (1953), who analyzed unsteady aero-
dynamics problems associated with oscillating wings, sinusoidal vertical gusts, and compressible
flow, respectively. Kramer (1932) was one of the first to experimentally document the augmented
lift associated with dynamic stall. Bailey & Gustafson (1939) and Gustafson & Myers (1946) used
tufts and a blade-mounted camera on an autogyro rotor blade to document dynamic stall, which
they found to occur on the retreating blade during forward flight.
Figure 1 illustrates the dynamic stall process that occurs on an airfoil oscillating about its
quarter-chord location. Figure 1a shows the aerodynamic loads and pitch moment during the
Annu. Rev. Fluid Mech. 2015.47. Downloaded from www.annualreviews.org

pitching cycle, and Figure 1b shows the flow field structure during the pitching cycle that was
visualized with particle streak lines introduced at an upstream location.
The conditions in Figure 1 exemplify deep dynamic stall, which occurs when the maximum
by Tulane University on 10/14/14. For personal use only.

angle of attack, α0 + α1 , of the pitching motion exceeds the steady airfoil stall angle of attack, α ss .
Here α0 denotes the mean angle of attack, and α1 denotes the pitch amplitude. This results in a
fully developed dynamic stall vortex-shedding phenomenon (McCroskey 1981) that produces large
peak aerodynamic loads and severe cycle hysteresis. The nomenclature is intended to differentiate
this condition from that of light dynamic stall, which occurs for lower peak angles of attack and
for which the peaks in the aerodynamic forces and pitch moment are less severe.
At the start of the pitch-up portion of the cycle (stage 1 in Figure 1), the boundary layer on
the suction surface of the airfoil (upper surface in the flow visualization) is attached, and the lift
force increases linearly with the pitch angle. This continues to the point at which the pitch angle
reaches the steady airfoil stall angle of attack, α ss . The lift then continues to increase beyond
α ss as a result of two mechanisms: (a) a delay in the boundary layer separation owing to the
pitching motion and (b) the formation of a closed separation bubble near the leading edge of the
airfoil.
The delay in the boundary layer separation is attributed to two effects. One is an increase in
the effective camber that is predicted from quasi-steady thin airfoil theory when α̇ > 0 (Leishman
2000). The other is the acceleration of the boundary layer due to the Magnus effect produced by
the motion of the leading edge (Ericsson & Reding 1984, 1988). Carta (1971) showed analytically
that the adverse pressure gradient over the suction side of a pitching airfoil was less than that of
a steady airfoil. The adverse pressure gradient is further attenuated as the pitch rate increases.
Walker et al. (1985) added that increasing the pitch rate promoted a stronger static pressure
suction peak at the leading edge that ultimately led to a more energetic dynamic stall vortex.
Stage 2 of the flow development corresponds to the first appearance and subsequent growth
of the dynamic stall vortex. This involves the spontaneous generation and ejection of vorticity
from the boundary layer into the inviscid outer flow in a process detailed by van Dommelen &
Shen (1980). The growth of the dynamic stall vortex results in additional aerodynamic loading
that exceeds that of the steady airfoil.
The onset of dynamic stall on a pitching airfoil can involve a number of mechanisms that depend
on Reynolds and Mach numbers. These include (a) the bursting or breakdown of the separation
bubble, (b) an abrupt breakdown of the boundary layer reverse flow within the separation bubble,
and (c) boundary layer–shock interaction that causes flow separation.

480 Corke · Thomas

Changes may still occur before final publication online and in print
FL47CH20-Corke ARI 16 September 2014 12:22

a b
α(t)
Unsteady 3
Steady

2
1 5

Cl 1 4

α = 5° 6°

5 12° 8°
Annu. Rev. Fluid Mech. 2015.47. Downloaded from www.annualreviews.org

1
by Tulane University on 10/14/14. For personal use only.

Cm
4

3 19° 13°

3
22° 19°
Cd 4

5
2
1 4
α (°) 25° 23°

Figure 1
(a) Illustration of dynamic stall events based on air loads and pitch moment cycle. Characteristic stages of the flow development in the
pitching cycle are denoted by the numbers next to specific flow visualization images and aerodynamic load and moment cycle maps.
(b) Visualized flow about a pitching airfoil undergoing deep dynamic stall. Panel b adapted with permission from Corke et al. (2011).

Certain stall situations, particularly on airfoils that coincide with thin airfoil theory, exhibit a
sudden breakdown of the laminar separation bubble owing to turbulent transition ( Johnson &
Ham 1972). McCroskey et al. (1976, 1981) extensively studied this process at lower Mach numbers
at which shock waves were not in play. In these experiments, the authors documented two alternate
scenarios for the onset of dynamic stall. The first involved trailing-edge stall in which separated
flow originating at the trailing edge gradually moved forward as the airfoil pitched upward. In
this case, the dynamic stall vortex formed aft of the maximum thickness point of the airfoil, never
reaching the leading edge.

www.annualreviews.org • Dynamic Stall in Pitching Airfoils 481

Changes may still occur before final publication online and in print
FL47CH20-Corke ARI 16 September 2014 12:22

The second scenario also originated at the trailing edge. It was manifest by a boundary layer
flow reversal that moved forward over the suction side of the airfoil. The occurrence of reverse
flow in the near-wall region of the boundary layer without the occurrence of large-scale separation
emphasizes the importance of not using the criterion of zero-skin friction as a separation indicator
in the pitching airfoil flow field (Reynolds & Carr 1985). The flow reversal was documented to have
little effect on the turbulent boundary layer or aerodynamic behavior until the forward motion
reached the maximum thickness point of the airfoil. Immediately thereafter, if not concurrently, the
turbulent boundary layer separated and abruptly moved upstream and downstream. This initiated
a dynamic stall vortex near the leading edge. McAlister & Carr (1979) and Lee & Gerontakos
(2004) observed a similar scenario.
Doligalski et al. (1994) comprehensively reviewed work focused on both the process leading to
the development of the dynamic stall vortex and its subsequent detachment from the airfoil surface.
The authors noted the proclivity of laminar boundary layers to develop a sharply focused eruption
of vortical fluid from the surface in the presence of an adverse pressure gradient. In pitching
Annu. Rev. Fluid Mech. 2015.47. Downloaded from www.annualreviews.org

airfoils at sufficiently high Reynolds numbers, several studies suggested that the initiation of the
dynamic stall vortex appears to be associated with a similar viscous/inviscid interaction in the form
of the sudden eruption of a thin plume of vortical fluid from the surface that subsequently deflects
by Tulane University on 10/14/14. For personal use only.

downstream and undergoes roll-up. Reynolds & Carr (1985) discussed the buildup of localized
boundary layer vorticity preceding dynamic stall in terms of a balance between the self-induced
upstream propagation of boundary layer vorticity and its downstream advection by the external
flow. The strong favorable pressure gradient near the airfoil leading edge of a pitching airfoil
introduces large amounts of vorticity into the boundary layer on the forward portion of the airfoil
concurrent with a decreasing external velocity over the aft portion. This sets the stage for a critical
localized accumulation of vorticity just aft of the suction peak.
At higher free-stream Mach numbers, the locally supercritical levels over the pitching airfoils
become the trigger for dynamic stall. This was evident from the experiments of McCroskey et al.
(1981), who found that a group of airfoils that exhibited one of the two trailing-edge dynamic
stall scenarios at lower free-stream Mach numbers showed leading-edge dynamic stall at a higher
free-stream Mach number in which supersonic flow occurred over the suction surface (Carr et al.
1982; McAlister et al. 1982; McCroskey et al. 1981, 1982). Chandrasekhara et al. (1998a) found
that complex interactions between the separation bubble and weak shock waves in the supersonic
region initiated the dynamic stall. At higher Mach numbers, M ∞ > 0.45, the extent of the
supersonic region over the pitching airfoil was found to considerably increase. Strong normal
shock waves were observed to form. These led to an abrupt thickening and subsequent separation
of the boundary layer flow. Chandrasekhara et al. (1998a) observed that an oblong vortical structure
formed after the shock. Thus, the shock-induced flow separation was the origin of the dynamic
stall vortex. The laminar separation bubble that formed near the leading edge was observed to be
present but played no role in the development of the dynamic stall vortex.
In stage 3 of the flow development shown in Figure 1, the fully formed dynamic stall vortex
convects over the suction side of the pitching airfoil. This motion of the low-pressure vortex core
shifts the airfoil center of pressure toward the trailing edge, resulting in an acute nose-down pitch
moment, referred to as moment stall. This occurs shortly after the dynamic stall vortex forms
(Harris et al. 1970). The moment stall precedes the lift stall. The time lag between the moment
stall and lift stall correlates with the time needed for the dynamic stall vortex to convect past the
trailing edge (Bousman 1998).
The low-pressure core of the dynamic stall vortex also augments the aerodynamic suction
on the airfoil, increasing the lift compared to a steady airfoil. The amount of lift enhancement,

482 Corke · Thomas

Changes may still occur before final publication online and in print
FL47CH20-Corke ARI 16 September 2014 12:22

however, is decreased at compressible Mach numbers. The added lift drops sharply when the
dynamic stall vortex convects off the airfoil and the flow is fully separated. This signifies stage 4
of the flow development.
In stage 5 of the flow development, the pitch-down motion of the airfoil causes the flow to
begin to reattach. This begins at the leading edge and gradually proceeds to the trailing edge. As
the flow attaches, there is a shift in the aerodynamic load distribution toward the leading edge,
which restores the positive pitch moment.
Recovery from dynamic stall is somewhat of a stochastic process in which significant differences
in aerodynamic loading can occur from one pitch cycle to the next (Green & Galbraith 1995,
Wernert et al. 1996, Young 1981). Liiva & Davenport (1969) proposed that these differences
resulted from random features of the separated shear layer. This was later supported through
particle image velocimetry measurements by Shih et al. (1992) and Wernert et al. (1996), who
observed that the separated shear layer contained nonrepeated patterns of vortical structures from
one cycle to the next. Werner et al. further remarked that the nonreproducible nature depended
Annu. Rev. Fluid Mech. 2015.47. Downloaded from www.annualreviews.org

on the size of the separated shear region and the reduced frequency of the pitching motion. The
cycle-to-cycle variations could be smoothed out through ensemble averaging correlated with the
pitching motion. McAlister et al. (1982) observed that 50 pitch cycles were sufficient to yield
by Tulane University on 10/14/14. For personal use only.

converged statistics.
Early rotor dynamic stall investigations focused on aeroelastic issues such as stall flutter (Brooks
& Baker 1958, Ham 1966, Zvara & Ham 1960). Ham & Garelick (1968) were the first to realize
the vortical nature of rotor blade dynamic stall. The signature of “very concentrated positive
vorticity” was captured in surface pressure measurements that revealed a dynamic low-pressure
region that formed near the leading edge and convected toward the trailing edge. Integrating
the pressure distribution around the rotor section to obtain the aerodynamic forces revealed a
severe hysteresis in the pitch moment that could couple with the pitching motion and result in
negative aerodynamic damping. Ham (1968) later linked the dynamic stall phenomenon with the
occurrence of stall flutter.
Following Ham (1967), investigators sought to identify the underlying mechanisms for stall
onset delay of a pitching airfoil that could lead to analytic or empirical models for predicting
the aerodynamic loads. At the same time, theories for unsteady flow separation developed that
highlighted fundamental differences between steady and unsteady boundary layers (Haller 2004,
Sears & Telionis 1975, Telionis 1970, van Dommelen & Shen 1980).
The early dynamic stall research almost exclusively focused on incompressible Mach numbers.
However, McCroskey et al. (1981, 1982) showed that the local Mach number near the leading
edge of pitching airfoils can be three to five times higher than that in the free stream. As a
result, incompressible free-stream Mach numbers as low as 0.18 could result in supercritical Mach
numbers over the airfoil. Carr & Chandrasekhara (1992) have particularly focused on the effect
of compressibility on dynamic stall.

1.1. Dynamic Stall Regimes


As mentioned above, the conditions shown in Figure 1 correspond to deep dynamic stall. This
follows the nomenclature set by McCroskey (1981), who divided dynamic stall into four categories:
no stall, stall onset, light stall, and deep stall. Although the majority of dynamic stall research has
focused on deep stall, it is avoided in normal rotorcraft operation owing to the severe forces
and vibrations that occur in this condition (Carr & Chandrasekhara 1996). McCroskey (1981)
delineated the stall regimes based on the maximum angle of attack, αmax = α0 + α1 .

www.annualreviews.org • Dynamic Stall in Pitching Airfoils 483

Changes may still occur before final publication online and in print
FL47CH20-Corke ARI 16 September 2014 12:22

1.1.1. No stall. In the no-stall regime, the airfoil trajectory remains below the static stall angle
of attack, α ss . Under these conditions, the aerodynamic loads are predicted well by quasi-steady
aerodynamic theory.

1.1.2. Stall onset. Stall onset is the regime in which the airfoil trajectory reaches the static stall
angle of attack. This condition produces the maximum useful lift without excessive drag or pitch
moment. Again, under these conditions, the aerodynamic loads are predicted well by quasi-steady
aerodynamic theory.

1.1.3. Light stall. The light-stall regime marks the first development of a dynamic stall vortex.
The onset, growth, and convection of the vortex are sensitive to the airfoil chord Reynolds number
and free-stream Mach number, as well as the unsteady parameters, including the pitching reduced
frequency, k, and α 0 and α 1 . Lift and moment stall peaks are less severe in this regime. The flow
separation region is on the order of the airfoil thickness (McCroskey 1981). The boundary between
Annu. Rev. Fluid Mech. 2015.47. Downloaded from www.annualreviews.org

stall onset and light stall is abrupt and can be identified by the first appearance of moment stall.

1.1.4. Deep stall. In the deep-stall regime, the dynamic stall vortex is strongly developed. The
by Tulane University on 10/14/14. For personal use only.

aerodynamic loads fluctuate dramatically through the pitching cycle with large peak forces and
strong hysteresis. The aerodynamic loads exhibit little sensitivity to Reynolds number, airfoil
geometry, or pitching motion. The flow separation region in this case is on the order of the airfoil
chord length (McCroskey et al. 1981).

1.2. Compressibility Effects


Compressibility affects the dynamic stall process of a pitching airfoil in a number of ways. This has
been the topic of many research efforts, including those by Carr & Chandrasekhara (1991, 1992,
1996), Carr et al. (1994), Chandrasekhara et al. (1993, 1998a), Dyken et al. (1996), and Bowles
(2012). Below we summarize the major effects compressibility has on the different elements of the
dynamic stall process.

1.2.1. Stall delay. Typically at M ∞ > 0.3, but possibly as low as M ∞ = 0.18, the locally
supersonic flow over the pitching airfoil reduces the flow acceleration around the leading edge,
thereby reducing the flow separation delay and the peak lift and pitch moment coefficients. This
is, however, a steady effect and is not related to the unsteady motion of the airfoil. For the same
pitching motion reduced frequency, the stall penetration, αds − αss , remains nearly constant with
increasing Mach number (Bowles 2012). Here αds denotes the dynamic lift stall angle.

1.2.2. Aerodynamic damping. At higher Mach numbers, the attached flow significantly lags
behind the pitching motion, so the hysteresis in the aerodynamic loads between pitch-up and pitch-
down increases with M∞ . As discussed in more detail below, this directly impacts the aerodynamic
damping. In this way, compressibility acts in a similar manner as the pitching reduced frequency.

1.2.3. Shock-induced stall. Local compressible Mach numbers over the pitching airfoil initially
promote leading-edge dynamic stall through incipient separation bubble bursting. At higher free-
stream Mach numbers, shock-induced dynamic stall becomes the overriding mechanism.

1.2.4. Stall vortex strength. Compressibility effects weaken the strength of the dynamic stall
vortex. This reduces the vortex-induced suction pressures on the airfoil. This is closely related to

484 Corke · Thomas

Changes may still occur before final publication online and in print
FL47CH20-Corke ARI 16 September 2014 12:22

a b
90° 90°
Uff U(r,ψ) = ωr + Uff sin(ψ) 120° 60°

150° 30°
ω

180° ψ = 0° 180° 0°
Reverse
flow
region
Annu. Rev. Fluid Mech. 2015.47. Downloaded from www.annualreviews.org

210° 330°
by Tulane University on 10/14/14. For personal use only.

240° 300°
270° 270°

UTTAS pull-up, revolution 22 Diving turn, revolution 20


UTTAS pull-up, revolution 14 Level flight, revolution 01

Figure 2
Helicopter rotor disk in forward flight (a) and rotor-disk locations of dynamic stall events during in-flight maneuvers of a UH-60A
helicopter (b). Figure adapted with permission from Bousman (1998).

the decrease in the gestation period and residence time of the dynamic stall vortex. Although the
magnitude of the induced pressures decreases with increasing Mach number, the dynamic range
of pressure coefficients on the airfoil, C p /C p min , increases with increasing Mach number. This
generally has a marginal impact on the lift and drag forces. However, it has a significant effect
on the pitch moment, with the outcome being an increase in the cycle-averaged aerodynamic
damping with increasing Mach number (Bowles 2012).

1.3. Helicopter Rotor Aerodynamics


The helicopter rotor flow field features complex, time-varying blade angles of attack, out-of-plane
mechanical and elastic rotor blade motion, unsteady boundary layers, massive flow separation,
vortex blade interaction, and varying inflow velocity. Each of these affects the rotor aerodynamic
performance and may contribute to excessive loading, mechanical vibration, and aerodynamic
noise generation (Conlisk 2001). To compound the flow complexity, the rotor blade may ex-
perience instantaneous local Mach numbers, Ml , that are incompressible, M l < 0.2, transonic,
0.7 ≤ M l ≤ 1, or supersonic, M l > 1, within a single rotor rotation (Leishman 2000, Liiva &
Davenport 1969, McCroskey et al. 1976).
Figure 2a illustrates the rotor disk for a helicopter in forward flight with velocity, Uff . The
rotor blade is considered to be rotating at a constant angular velocity of ω. The angular position of
the rotor is denoted by ψ. For 0◦ ≤ ψ ≤ 180◦ , the rotor blade is moving opposite to the direction
of the relative airflow caused by forward flight and therefore is experiencing local velocities that
are larger than Uff . This is referred to as the advancing blade. For the other half of the rotation,

www.annualreviews.org • Dynamic Stall in Pitching Airfoils 485

Changes may still occur before final publication online and in print
FL47CH20-Corke ARI 16 September 2014 12:22

180◦ ≤ ψ ≤ 0◦ , the motion is in the direction of the relative airflow caused by forward flight;
therefore, the blade experiences local velocities that are less than Uff . This is referred to as the
retreating blade. The lift generated on the advancing blade must be the same as on the retreating
blade; therefore, the angle of attack on the retreating blade must be larger than that on the
advancing blade. As a result, the rotor blade oscillates in pitch as it rotates from the advancing
to retreating sides of the rotor disk. This provides one major motivation for the pitching airfoil
research discussed above.
The main parameter governing the unsteady nature of the pitching rotor blade is the reduced
frequency, k = π f c /U ff , where f is the physical pitching frequency, and c is the rotor blade chord
length. The reduced frequency of the pitching motion represents the ratio of timescales associated
with the unsteady pitch motion to that of the free-stream convection (Leishman 2000). For k = 0,
the flow is steady. For 0 < k < 0.05, the unsteady effects are generally small and can be neglected,
provided the pitch amplitude is not too large. For k ≥ 0.05, the flow field is considered to be
unsteady. If k ≥ 0.2, the flow field is considered to be highly unsteady. The reduced frequency
Annu. Rev. Fluid Mech. 2015.47. Downloaded from www.annualreviews.org

can also be thought of as a phase lag parameter in which the fluid reaction lags behind the pitching
motion owing to the inertial effects. Larger reduced frequencies correspond to an increased
phase lag between the pitching and fluid motions. In general, the reduced frequency is constantly
by Tulane University on 10/14/14. For personal use only.

changing on the helicopter rotor because of the local inflow dependence on Uff . As a result, it is a
somewhat ambiguous parameter with regard to the unsteadiness associated with helicopter rotor
aerodynamics (Leishman 2000). However, in wind-tunnel experiments, the reduced frequency is
a constant value for a fixed combination of pitching frequency and free-stream speed.
Dynamic stall most commonly occurs on the retreating side of the rotor disk (Harris et al. 1970,
Tarzanin 1972, Young 1981), although there is some evidence of dynamic stall on the advancing
rotor (Bousman 1998). Bousman (1998) found that stall events occurred at multiple sites during
flight tests of a UH-60A helicopter. Figure 2b shows a map of these events in the rotor disk plane.
Although the majority of the rotor stall did occur on the retreating blade side, some stall events
occurred in the first quadrant (0◦ ≤ ψ ≤ 90◦ ) of the advancing blade.
The potential for dynamic stall depends on the rotor blade loading, which is the ratio of the rotor
thrust coefficient to the rotor solidity, C T /σ , and the flight speed. Increased blade loading results in
increased dynamic stall. However, at lower blade loading, increased forward flight speed promotes
an increase in dynamic stall. Helicopter performance boundaries are defined by the rotor dynamic
stall. The dependent variables are the blade loading, C T /σ , and the advance ratio, μ = U ff / R,
which is the ratio of the flight speed to the rotor blade tip speed. At high advance ratios, dynamic
stall can occur on the advancing blade owing to shock formation (Martin et al. 2008).
The practical outcome of dynamic stall on the rotor is an increase in unsteady loads on the
rotor that are subsequently transmitted to the rotor hub and control system. In some cases, the
conditions of dynamic stall can lead to negative damping that results in a limit-cycle growth of
rotor displacements. This phenomenon is referred to as stall flutter (Ham & Young 1966), which
can lead to catastrophic mechanical failure of the rotor.

2. AERODYNAMIC DAMPING AND STALL FLUTTER


Stall flutter is a single-degree-of-freedom aeroelastic motion that results from negative system
(combined aerodynamic and structural) damping. On a helicopter, it appears as a limit-cycle
growth in amplitude of rotor torsional oscillations at the torsional natural frequency (Fung 2008).
This can drive other modes of vibration of the rotor as well. For a helicopter, limit-cycle oscillations
triggered on the retreating blade side are often damped as the flow reattaches on the advancing
blade side (Ham & Young 1966). However, if the aerodynamic damping on the advancing side is

486 Corke · Thomas

Changes may still occur before final publication online and in print
FL47CH20-Corke ARI 16 September 2014 12:22

near zero, successive rotations of the rotor through negatively damped retreating blade regions
can lead to unchecked, additive limit-cycle growth to excessive amplitudes (Bowles 2012).
The stability of the air load system has been traditionally quantified through the aerodynamic
damping coefficient, cycle , derived by Carta & Niebanck (1969) and Oates (1989). The derivation
considers a single-degree-of-freedom structure subject to harmonic forcing in a uniform airstream.
The cycle aerodynamic damping is given as

1
cycle = −C W /(π α12 ) = − 2 Cmc /4 d α, (1)
π α1
where C W is the normalized energy transfer between the airstream and airfoil, and Cmc /4 is the
moment coefficient about the quarter chord.
Equation 1 can be experimentally computed from the area enclosed in the cycle variation in
the pitch moment coefficient, Cmc /4 (α), namely,
 αmax
1
cycle = (CmDc /4 − CmUc /4 )d α, (2)
Annu. Rev. Fluid Mech. 2015.47. Downloaded from www.annualreviews.org

π α12 αmin
where the superscripts U and D denote the upward and downward angle of attack motion of the
airfoil, respectively.
by Tulane University on 10/14/14. For personal use only.

For Theodorsen’s model of a thin airfoil (flat plate) undergoing pure pitch oscillations about its
quarter-chord location in a potential flow field, the aerodynamic damping is (Carta & Niebanck
1969)
cycle = π k/2. (3)
When the maximum angle of attack in the pitching cycle is below the static stall angle of attack, α ss ,
the aerodynamic loads and moments are predicted well by Theodorsen’s (1935) method. Under
these conditions, the aerodynamic damping coefficient is always positive and linearly dependent
on the reduced frequency, k, as given in Equation 3.
If the airfoil pitching range is such that its maximum angle of attack exceeds α ss by a few degrees,
then light dynamic stall ensues (McCroskey 1982). Figure 3a depicts the moment coefficient
during a pitch cycle for light stall. McCroskey et al. (1981) found that light dynamic stall had
the largest tendency to produce unstable aerodynamic loading on helicopter rotor sections. As
indicated in Figure 3, clockwise loops in the moment coefficient trajectory are associated with
negative damping. Counterclockwise loops lead to positive damping.

a b
Light stall Deep stall

Cm
c/4
α (°)
Ξ>0
Ξ>0
Ξ<0
Ξ<0
Downstroke
Upstroke
Steady Ξ>0
Ξ>0

Figure 3
Illustration of negative damping from the pitch moment coefficient, Cmc /4 (α): upstroke, downstroke, and steady.

www.annualreviews.org • Dynamic Stall in Pitching Airfoils 487

Changes may still occur before final publication online and in print
FL47CH20-Corke ARI 16 September 2014 12:22

Deep stall occurs when the maximum angle of attack during the pitching cycle far exceeds α ss
(McCroskey 1982). Figure 3b illustrates the moment coefficient during a pitch cycle undergoing
deep stall. Based on the cycle-averaged damping, deep stall is generally more aerodynamically
stable than light dynamic stall. This is apparent in Figure 3 in which the clockwise moment
coefficient loop for light stall is much larger than that for deep stall.
Bowles (2012) and Bowles et al. (2012, 2014) developed a formulation for a time-resolved
aerodynamic damping coefficient. This involved the use of the discrete Hilbert transform of the
pitch moment and incidence time series. The integral Hilbert transform of a data series y(t),
denoted here as ỹ(t), is defined as
 ∞
1 y(τ )
ỹ(t) ≡ H[y(t)] = − P dτ, (4)
π −∞ τ − t

where H is the Hilbert operator, and ℘ denotes the Cauchy principal value of the improper
integral.
The single-degree-of-freedom equation of motion (EOM) for a pitching airfoil semichord b
Annu. Rev. Fluid Mech. 2015.47. Downloaded from www.annualreviews.org

in a uniform airstream is
πρb 4 α̈(t) + h ∗ (t)α̇(t) + κ ∗ (t)α(t) = M (t), (5)
by Tulane University on 10/14/14. For personal use only.

where h ∗ (t) = h R (t) + i h I (t) = h̄(t)e iγ1 (t) is the viscous damping moment, κ ∗ (t) = κ R (t) + iκ I (t) =
κ̄(t)e iγ2 (t) is the aerodynamic stiffness, and M(t) is the pitch moment. This EOM can account for
nonlinearities inherent to the dynamic stall problem, unlike formulations that rely on constant
amplitude, constant phase damping, or stiffness (Oates 1989). Both h ∗ (t) and κ ∗ (t) contribute to
the total aerodynamic damping.
Substitution of the polar representations of h∗ and κ ∗ into Equation 5 yields
1
α̈ + 2h 0 (t)e iγ1 (t) α̇ + ω02 (t)e iγ2 (t) α = M, (6)
πρb 4
where the damping coefficient, h0 , is defined as
h 0 ≡ h̄/2πρb 4 , (7)
and the undamped natural frequency, ω0 , is

ω0 ≡ κ̄/πρb 4 . (8)
In this formulation, the instantaneous pitch angle, α, its derivatives, and the instantaneous pitch
moment, M, remain time dependent.
To employ the Hilbert transform analysis, one replaces each term in Equation 6 by its analytic
signal counterpart found through the Hilbert transform; this generates the new EOM
1
Ä + 2h 0 (t)e iγ1 (t) Ȧ + ω02 (t)e iγ2 (t) A = M, (9)
πρb 4
where A = α + i α̃ = α1 e iωt and M = M + i M̃ = A(t)e iφ(t) (Bowles 2012; Bowles et al. 2012,
2014).
Substituting A, M, and the first and second derivatives of A into Equation 9, and equating the
imaginary parts, one obtains
2h 0 ω cos γ1 + ω02 sin γ2
1
= ( M̃ cos ωt − M sin ωt)
α1 πρb 4 (10)
A(t)
= sin(φ(t) − ωt).
α1 πρb 4

488 Corke · Thomas

Changes may still occur before final publication online and in print
FL47CH20-Corke ARI 16 September 2014 12:22

As defined by Carta & Niebanck (1969), the left-hand side of Equation 10 is the total aerodynamic
damping, −ξ/πρb 4 . Therefore,
1
ξ (t) = − ( M̃ cos ωt − M sin ωt)
α1   
MI
A(t) (11)
=− sin(φ(t) − ωt ),
α1   
ψ

where MI denotes the quadrature (out-of-phase) pitch moment component.


Nondimensionalizing Equation 11 provides the aerodynamic damping coefficient,
ACm (t)
(t) = ξ/q c 2 = − sin ψ(t), (12)
α1

where ACm (t) = Cm2 + C̃m2 , and ψ(t) is the phase lead or lag between the pitch moment and the
Annu. Rev. Fluid Mech. 2015.47. Downloaded from www.annualreviews.org

time-dependent angle of attack. Equation 12 illustrates how the phase lead or lag between the
pitch moment and the prescribed angle of attack determines the sign of the aerodynamic damping;
its form is identical to previous derivations of the aerodynamic damping (Bisplinghoff et al. 1955,
by Tulane University on 10/14/14. For personal use only.

Oates 1989, Rainey 1957). Furthermore, it remains that a phase lag, ψ(t) < 0, produces positive
damping, whereas ψ(t) > 0 indicates negative damping or a local exchange of energy from the
airstream to the airfoil. reflects the stability of the pitch moment, which in turn is a manifestation
of the developing pressure field. An unstable pitch moment, < 0, implies unstable pressure
loading and can be described as a necessary, aerodynamic condition to excite the stall flutter of an
elastic body.
The novelty of the approach, unlike previous investigations of aerodynamic damping that rely
on cycle , is that Equation 12 allows one to calculate the intracycle aerodynamic damping. The

T
average damping, avg , is then avg = T1 0 (t)d t, where T is the period of a single oscillation.

2.1. Attached Flow


The cycle-integrated damping, cycle , for attached flow has been documented to increase with
increasing Mach number (Lorber et al. 1992). In fact, this increase can be modeled by typical
compressibility corrections, for example, the Prandtl-Glauert rule (Bowles 2012), up to the for-
mation of shocks on the airfoil upper surface. When shocks form, the cycle-integrated damping
rapidly increases. This cycle-averaged statistic, however, masks the complete physics that includes
negatively damped portions of the pitch cycle that could promote unstable flutter.
As an illustration of this behavior, Figure 4 depicts the Hilbert transform–based damping
analysis for two experimental runs of nearly equal stall penetration (αmax − αss ≈ −1.1◦ ) but at
low– and high–subsonic Mach number conditions. Figure 4a,d shows the pitch moment variation
with incidence. In each of these panels, Cm denotes the data series obtained by spatial integration
of the chordwise static pressures. These are ensemble-averaged quantities (≈100 cycles) chosen
in lieu of the instantaneous pitching moment to remove the cycle-to-cycle variations common
to the dynamic stall problem (Wernert et al. 1997). The Hilbert transform (C̃m ), in-phase (Cm R ),
and quadrature (Cm I ) pitch moments are also displayed. At M ∞ = 0.2, the elliptic, counterclock-
wise nature of the pitching moment is clearly seen. Figure 4b,e presents the polar representation
of the intracycle damping coefficient. In this representation, the radial component indicates the
magnitude of the damping, ACm (t)/α1 , and the vector angle is the phase difference between the
aerodynamic pitch moment and the pitch motion, ψ(t) = φ(t) − ωt. Recall from Equation 12 that
(t) ∝ − sin ψ(t); therefore, when ψ(t) is in the third and fourth polar quadrants, the damping is

www.annualreviews.org • Dynamic Stall in Pitching Airfoils 489

Changes may still occur before final publication online and in print
FL47CH20-Corke ARI 16 September 2014 12:22

π/2
0.02 3
a b c
ACm(t)/α1
0.01 2

2Ξ(t)/π k
ψ (t)
c/4

0 ±π 0 1
Cm

0 0.2 0.4

–0.01 0
UNSTABLE
–0.02 –1
0 2 4 6 8 10 12 0 π/2 π 3π/2 2π
α (°) – π /2 ωt
Cm CmR
~
Cm CmI π/2
0.02 3
d e f
Annu. Rev. Fluid Mech. 2015.47. Downloaded from www.annualreviews.org

0.01 2
2Ξcycle/π k

2Ξ(t)/π k
c/4
by Tulane University on 10/14/14. For personal use only.

0 ±π 0 1
Cm

0 0.2 0.4

–0.01 0
UNSTABLE
–0.02 –1
–4 –2 0 2 4 6 8 10 0 π/2 π 3π /2 2π
α (°) – π /2 ωt

Figure 4
Time-resolved damping analysis demonstrating unstable damping for cases in which αmax < αss . (a,d ) Data series, discrete Hilbert
transform, and in-phase and quadrature moments. Arrows indicate the direction of increasing time. (b,e) Polar representation of the
time-resolved damping coefficient. (c,f ) Transient damping coefficient over a single pitch cycle. The formation of normal shocks on the
airfoil upper surface at M ∞ = 0.6 (d–f ) results in unstable damping. (a–c) M ∞ = 0.2, k = 0.052, and α = 5.2◦ − 5.1◦ cos ωt.
(d–f ) M ∞ = 0.6, k = 0.046, and α = 3.3◦ − 5.4◦ cos ωt. Figure adapted with permission from Bowles et al. (2012, 2014).

positive, whereas in the first and second quadrants, the damping is negative. The polar represen-
tation of the intracycle damping coefficient at M ∞ = 0.2 shown in Figure 4b traces out only a
small region in the third quadrant, yielding the nearly constant (albeit noisy) positive intracycle
damping coefficient shown in Figure 4c.
At M ∞ = 0.6, the pitch moment exhibits no crossovers, indicating no observable negative
damping, although the ellipse is clearly distorted. The distortion is a consequence of shock develop-
ment and the concomitant nose-up moment. This behavior causes the phase difference (Figure 4e)
to decrease and become positive at ωt = 0.83π (α = 7.5◦ ). This leads to a brief period of negative
aerodynamic damping. This negative damping is not observed in Figure 4d because of the much
larger magnitude of the positive damping during pitch-down over the same range of angle of
attack.
The destabilizing effect of shock formation is further illustrated in Figure 5, which shows
plots of the upper-surface pressures, pitch moment Cmc /4 , instantaneous angle of attack α, and
time-resolved damping (t) at M ∞ = 0.6 and α = 4.0◦ − 4.4◦ cos ωt. For this case, as in the
M ∞ = 0.6 case examined in Figure 4, localized, negative damping occurs during a portion of
the pitching cycle. The mechanism for the negative intracycle damping stems from the growth of
a λ shock during the pitch-up portion of the cycle. The position of the shock front was inferred
from the brief increase in the pressure captured by the onboard sensors. This shock wave was

490 Corke · Thomas

Changes may still occur before final publication online and in print
FL47CH20-Corke ARI 16 September 2014 12:22

Ml < 1.0 Ml > 1.0 Ml > 1.31


10

α (°) 5

0.2
0.1
Ξ(t)
0
–0.1

0.02 Destabilizing Destabilizing

Cm 0
c/4

–0.02
x=0
Annu. Rev. Fluid Mech. 2015.47. Downloaded from www.annualreviews.org

SHO

0.005
0.022
CK
by Tulane University on 10/14/14. For personal use only.

0.049
POS

0.086
ITIO

0.134
N

0.191

0.257
0.331

0.412

0.500

0.593

0.691

–3.0 0.792

C sp 0.895
0 0.995

0 π /2 π 3π/2 2π
ωt

Figure 5
Upper-surface pressures, α = 4.0◦−4.4◦ cos ωt. The zero-axis and nondimensional chord location of each
pressure sensor is indicated on the right. Each pressure time series is offset by C pS ≈ −1.5 and color coded to
represent the local Mach number, Ml , on the airfoil surface. The position of the shock front is outlined.
Figure adapted with permission from Bowles et al. (2012, 2014).

also observed in high-speed schlieren images of the leading-edge flow field (Bowles 2012). The
destabilizing portion of the motion occurs after the streamwise propagation of the shock front
comes to a halt near the quarter-chord location. The destabilization results from a decrease in the
pressure upstream of the shock front that increases the leading-edge suction. This couples with
higher static pressure aft of the shock, which results in a nose-up pitch moment that is responsible

www.annualreviews.org • Dynamic Stall in Pitching Airfoils 491

Changes may still occur before final publication online and in print
FL47CH20-Corke ARI 16 September 2014 12:22

3 0

a 0
b
2 M∞ =
0.6
STABLE 0
1
0.5
2Ξcycle /πk

c/4
0 0.4

Cm
M∞ 0
0.2
–1 UNSTABLE
0.3 0
0.4 αss 0.3
0.5
–2
0.55 –0.1
0.6 0.2
–3 –0.2
2 6 10 14 18 22 –10 –8 –6 –4 –2 0 2 4
Maximum angle of attack, α0 + α1 (°) Stall penetration angle, αds + αss (°)
Annu. Rev. Fluid Mech. 2015.47. Downloaded from www.annualreviews.org

Figure 6
(a) Effect of dynamic stall regime, set by the maximum incidence, on the cycle damping coefficient, k ≈ 0.05, α1 ≈ 5◦ . The static stall
by Tulane University on 10/14/14. For personal use only.

angles of attack for each Mach number are indicated by the colored arrows. (b) Measured pitch moment for light stall, αds − αss ≈ 2◦ , at
M ∞ = 0.2, 0.3, 0.4, 0.5, and 0.6; k ≈ 0.05; and α1 ≈ 5◦ . The solid lines are for pitch-up and dashed lines for pitch-down. For clarity,
the moment plots at the five Mach numbers are offset by Cmc /4 = +0.1 and color coded in accordance with panel a. Figure adapted with
permission from Bowles et al. (2012).

for a decrease in the aerodynamic damping at that phase in the pitching cycle. A similar decrease
in damping occurs during pitch-down, just prior to the shock front retreating to the leading edge.
Based upon examination of the minimum intracycle aerodynamic damping coefficient, min , we
can summarize the effect of reduced frequency and the Mach number on the aerodynamic damping
characteristics of attached pitching flow as follows. (a) Regardless of the Mach number, increased
reduced frequency has a stabilizing influence (increasing positive aerodynamic damping). (b) At
any fixed reduced frequency, the effect of increasing the Mach number on the damping is minimal,
except when shocks begin to form on the suction surface. (c) Shock waves can produce negative
damping that could trigger stall flutter even in an attached pitching condition!

2.2. Light and Deep Dynamic Stall


As described above, dynamic stall occurs when the pitching motion exceeds the static stall angle
of attack, α ss . This leads to the eruption of surface vorticity and the formation and downstream
convection of a dynamic stall vortex. The process results in large peak aerodynamic loads and
pitch moments and strong cycle hysteresis that directly affects the aerodynamic damping.
Figure 6a highlights the typical behavior of the cycle-averaged damping coefficient, cycle ,
under light- and deep-stall conditions for different free-stream Mach numbers. Figure 6b singles
out five pitch moment cycles from the data set that have similar dynamic stall penetration at their
respective Mach numbers.
One can observe that the minimum damping in these cases occurred when the maximum angle
of attack, αmax = α0 + α1 , was approximately 2–3◦ past α ss . This level of stall penetration classically
falls into the light dynamic stall regime described by McCroskey (1981). As the maximum incidence
increases to enter the deep-stall regime, the cycle damping becomes more positive, which is a result
of the strongly developed dynamic stall vortex (Carta & Niebanck 1969, Fung 2008). Similar
damping and Mach trends are reported by Liiva (1969).

492 Corke · Thomas

Changes may still occur before final publication online and in print
FL47CH20-Corke ARI 16 September 2014 12:22

One can observe the trend of increased damping with increasing Mach number in the accompa-
nying pitch moment cycles. Beginning at M ∞ = 0.2, the clockwise rotation of the pitch moment
path for −2◦ ≤ (αds − αss ) ≤ 2◦ is emblematic of an unstable damping coefficient (see Figure 3).
At higher Mach numbers, the pitch moment path direction in the vicinity of α ss is decidedly coun-
terclockwise, which signifies positive damping. Moreover, the size of the moment hysteresis loop
decreases with increasing Mach number, which is consistent with an increase in positive cycle
damping with increasing Mach number, observed in Figure 6a.
Based on this assessment of the cycle-averaged damping, one might conclude that there is less
tendency for stall flutter under light and deep dynamic stall at higher Mach numbers. However,
the cycle-averaged statistics can conceal important physics that can lead to localized, unstable
damping during the dynamic stall process.
To demonstrate this, investigators analyzed three of the pitch moment cycles at M ∞ = 0.2,
0.3, and 0.6 in Figure 6 using the Hilbert transform formulation to examine intracycle damping
characteristics. Figure 7a,c,e shows the polar representation of the aerodynamic damping for
Annu. Rev. Fluid Mech. 2015.47. Downloaded from www.annualreviews.org

the pitch cycle. Figure 7b,d,f simultaneously indicates the time-resolved damping, the ensemble
pitching moment, Cmc /4 , and the position of the dynamic stall vortex based on tracking the local
minimum in the suction-side surface pressure (Lorber & Carta 1988).
by Tulane University on 10/14/14. For personal use only.

At the lower Mach number in Figure 7a, M ∞ = 0.2, the phase ψ(t) at the start of pitch-up
(αmin ) indicates that the aerodynamic damping is near zero or neutrally stable. As the airfoil pitches
up, ψ(t) moves into the second quadrant, indicating negative aerodynamic damping. During this
portion of the pitch-up cycle, the pitch moment amplitude, ACm (t), increases such that the polar
representation of the damping sweeps out a large area in the second quadrant corresponding to
negative damping.
The phase lead indicated by ψ(t) during pitch-up is a consequence of the induced stall delay
produced by a positive pitch rate (α̇ > 0) and an accompanying alleviation of the suction-surface
adverse pressure gradient. This allows the aerodynamic pitch moment to remain nearly constant as
the airfoil pitches up. The result is a steady increase in the negative damping throughout pitch-up,
as documented in Figure 7b.
The rate of negative damping is observed to accelerate at ωt = 3π/4. This portion of pitch-up
coincides with the initial formation of the dynamic stall vortex near the leading edge at x = 0.006.
The damping continues to become more negative up to the point at which the dynamic stall vortex
arrives at the quarter-chord (pitching) location (x = 0.25). This corresponds to ωt = 0.8π , where
(2 (t))/(π k) = −13.89.
As the dynamic stall vortex convects downstream past the quarter-chord location toward the
airfoil trailing edge, the aerodynamic damping begins to increase. The low (suction) pressure
that follows the position of the vortex loads the trailing edge and thereby produces a negative
(nose-down) pitch moment. For the M ∞ = 0.2 case, the negative pitch-down moment induced
by the vortex persists into the pitch-down portion (α̇ < 0) of the cycle so that the aerodynamic
damping continues to be negative. Positive damping does not begin until ωt = 1.04π , where the
upper-surface flow field becomes completely separated.
To summarize these observations, we note the following: Large negative aerodynamic damping
that could spawn flutter occurs in a portion of the pitching cycle. Additionally, the source of
negative aerodynamic damping stems from the stall delay that results from the α̇ > 0 pitching
motion. Negative damping is also augmented by the formation of the dynamic stall vortex. Once
the dynamic stall vortex convects past the pitching location, its effect is stabilizing. After the
dynamic stall vortex convects past the trailing edge, the fully separated flow field initiates positive
aerodynamic damping. Finally, the positive damping reaches a maximum when the flow fully
reattaches.

www.annualreviews.org • Dynamic Stall in Pitching Airfoils 493

Changes may still occur before final publication online and in print
FL47CH20-Corke ARI 16 September 2014 12:22

π/2
15
10
b
α (°)
5
a 0

ACm(t)/α1 1.0

x 0.5 Ξ(t)
ψ (t) Ξcycle
±π 0 0 Cm
0 1 2 c/4
10 0.2
5 0.1
2Ξ(t)/π k 0 Cm
0 c/4
–5
–10 –0.1
– π /2 –15 –0.2
0 π /2 π 3π /2 2π
ωt
Annu. Rev. Fluid Mech. 2015.47. Downloaded from www.annualreviews.org

π/2
20
15
d
α (°)
10
by Tulane University on 10/14/14. For personal use only.

c 5

1.0

x 0.5
±π 0 0
0 1 2
10 0.2
5 0.1
2Ξ(t)/π k 0 Cm
0 c/4
–5
–10 –0.1
– π /2 –15 –0.2
0 π /2 π 3π /2 2π
ωt

15
π/2
f
10
α (°)
5
0
e
1.0

x 0.5
±π 0 0
0 1 2
10 0.2
5 0.1
2Ξ(t)/π k 0 Cm
0 c/4
–5
–10 –0.1
– π /2 –15 –0.2
0 π /2 π 3π /2 2π
ωt

Figure 7
Time-resolved damping analysis of light dynamic stall. (a,c,e) The polar representation of the transient aerodynamic damping. The
light blue arrows indicate increasing time. (b,d,f ) The transient damping, pitch moment, and position of the dynamic stall vortex.
(a,b) M ∞ = 0.2, k = 0.051, and α = 9.8◦ − 4.9◦ cos ωt. (c,d ) M∞ = 0.3, k = 0.051, and α = 10.7◦ − 4.9◦ cos ωt. (e,f ) M ∞ = 0.6, k =
0.046, and α = 5.4◦ − 5.1◦ cos ωt. Figure adapted with permission from Bowles et al. (2012).

494 Corke · Thomas

Changes may still occur before final publication online and in print
FL47CH20-Corke ARI 16 September 2014 12:22

As evident in Figure 6, the cycle-averaged damping coefficient for the pitching airfoil at
M ∞ = 0.2 is negative. Therefore, it may not be surprising to find the large negative damping
peak in the time-resolved damping coefficient. However, in this case, the ratio min / cycle is 7.73!
Thus, there exists a significant possibility for stall-flutter oscillations.
In contrast to the lower–Mach number case, at higher Mach numbers, the cycle-integrated
damping is always positive. Therefore, one would not expect these conditions to exhibit stall flutter.
However, the flight data shown in Figure 2b suggest otherwise. Figure 7c,d for M ∞ = 0.3 and
Figure 7e,f for M ∞ = 0.6 each reveal a negative peak in the intracycle damping that accompanies
the pitch-up motion of the airfoil. The intracycle damping for M ∞ = 0.3 has all the same
characteristics as its M ∞ = 0.2 counterpart. However, the magnitude of the negative damping
peak is not as large, which is attributed to a reduction in the time that the dynamic stall vortex
resides over the airfoil, as is evident by the x trajectory in Figure 7d.
At M ∞ = 0.6, dynamic stall is triggered by shock waves on the suction surface of the airfoil.
The dynamic stall vortex in this case forms downstream of the shock front, at x = 0.134. This
Annu. Rev. Fluid Mech. 2015.47. Downloaded from www.annualreviews.org

is still upstream of the pitching location and therefore contributes a positive pitch moment that
results in a negative peak in the aerodynamic damping. The magnitude of the peak negative
damping coefficient is larger than that at M ∞ = 0.3, which is a reversal of the previous trend with
by Tulane University on 10/14/14. For personal use only.

increasing Mach number.


The x trajectory of the dynamic stall vortex in Figure 7f reveals a very gradual convection.
Chandrasekhara et al. (1994) found that the convection of the vortex in this instance is supplanted
by a near-constant-speed, low-pressure traveling gust that moves downstream. This gust fails to
induce the same positive aerodynamic damping that the dynamic stall vortex produced when it
traversed past the pitching location on the airfoil at the lower Mach numbers.

2.3. min versus cycle


For helicopter rotor operations, the magnitude of the minimum damping coefficient within the
pitch cycle, min , is potentially a more critical property than is the cycle-averaged damping coef-
ficient, cycle . Local, negative intracycle aerodynamic damping, especially of the magnitude doc-
umented by Bowles (2012) and Bowles et al. (2012, 2014), can spawn limit-cycle vibration growth
and possibly account for the stall-flutter-related divergence found during high-speed forward
flight or high g-force maneuvers (Bousman 1998). Bowles (2012) compared the cycle-averaged
and minimum intracycle aerodynamic damping coefficients for different dynamic stall regimes
over a range of Mach numbers from 0.2 to 0.6. For conditions in which the peak angle of attack
was less than the steady stall angle of attack, αmax < αss , the difference between min and cycle
was negligible across the whole Mach number regime. However, when αmax > αss , the minimum
intracycle damping was always significantly less than the cycle-averaged damping, and negative.
This was the case for the whole Mach number range.

2.4. Leading-Edge Turbulent Trip Effects


As discussed above, the pitch-up motion of the leading edge (α̇ > 0) results in an acceleration
of the boundary layer flow, which delays separation. In a similar manner, the flow acceleration
near the leading edge should make the boundary layer more stable to disturbances and therefore
suppress turbulent transition. McCroskey et al. (1981) noted that the nature of the boundary
layer separation that precedes vortex development strongly influences dynamic stall, and this is
particularly true in the light-stall regime. At least at Mach numbers below which dynamic stall is
influenced by shock waves, the transition state of the boundary layer (i.e., laminar versus turbulent)
should influence dynamic stall vortex inception and strength.

www.annualreviews.org • Dynamic Stall in Pitching Airfoils 495

Changes may still occur before final publication online and in print
FL47CH20-Corke ARI 16 September 2014 12:22

Edge Edge
Distributed roughness Distributed roughness
roughness (tape) roughness (tape)

2.0 0
a b
1.5 –2

1.0 –4

0.5 –6

Ξcycle 0 Ξ(t)min –8

–0.5 –10
Annu. Rev. Fluid Mech. 2015.47. Downloaded from www.annualreviews.org

–1.0 –12
by Tulane University on 10/14/14. For personal use only.

M∞ = 0.2
–1.5 –14
M∞ = 0.6

–2.0 –16
0 1 2 3 4 5 6 0 1 2 3 4 5 6

Effective thickness (10–3 inches)

Figure 8
Effect of leading-edge trips on (a) cycle-averaged and (b) minimum intracycle damping coefficients for two Mach numbers. Roughness
consists of masking tape and 320- and 400-grit sandpaper strips. Data adapted from Bowles (2012).

With this in mind, there have been several studies on the effect of leading-edge turbulent
trips on dynamic stall (Bowles 2012, Chandrasekhara et al. 1996). Bowles (2012) examined the
effect of three turbulent trips consisting of strips of masking tape and strips of sandpaper of two
grit sizes. The turbulent trips were placed so that they overlaid the position of the leading-edge
flow separation bubble that formed just prior to stall on the static airfoil. This corresponded to
0.005 ≤ x ≤ 0.030. The surface visualization with the different trips showed that they eliminated
the separation bubble (Bowles 2012). Figure 8 summarizes the results, documenting the effect on
the cycle-averaged and intracycle damping coefficients.
Based on the conventional cycle-averaged damping coefficient, at the lower Mach number,
M ∞ = 0.2, the leading-edge trips weakened the dynamic stall vortex (Bowles 2012), which resulted
in the cycle-integrated damping becoming positive. The distributed sand grain roughness was
found most effective. At M ∞ = 0.6, dynamic stall vortex formation occurred downstream of the
shock front, which was at x = 0.134. The turbulent trips were therefore well upstream (x ≤ 0.03)
and, based on the cycle-averaged damping coefficient, had a minimal effect.
The minimum intracycle damping coefficients for the distributed roughness trips at M ∞ = 0.2
support the cycle-averaged view, with a significant reduction in the negative damping peak that
is associated with the dynamic stall vortex formation. The edge roughness trip was found to be
much less effective. However, at M ∞ = 0.6, at which the dynamic stall vortex forms downstream
of the shock, the distributed roughness trips result in significantly larger negative damping, which
is not indicated by the conventional cycle-averaged damping coefficient. For the two sandpaper
trips, for example, min ranges from −5 to −9, and cycle is nearly neutral. This reverse trend in

496 Corke · Thomas

Changes may still occur before final publication online and in print
FL47CH20-Corke ARI 16 September 2014 12:22

TRANSONIC GLOVE

Successful control of the detrimental effects associated with dynamic stall was achieved by Martin et al. (2008)
by use of a transonic leading-edge glove and passive vortex generator combination on a VR-7 airfoil. The vortex
generators were placed at x = 0.1. It is significant to note that neither the leading-edge glove nor the passive vortex
generators alone were successful at controlling dynamic stall. Their combination, however, reduced the local Mach
number, thereby enabling shock-free operation of the passive vortex generators. In the light-stall regime, this had
the effect of virtually eliminating the adverse pitching moment associated with dynamic stall for Mach numbers
up to 0.4. In the deep-stall regime, it significantly reduced adverse pitch moments at Mach numbers up to 0.375.
These results motivate the development of flush, surface-mounted active flow control devices that can mimic the
effect of passive vortex generators without incurring the adverse effects of shock formation. One recent approach
involves the plasma vortex generator demonstrated by Schatzman & Thomas (2010).
Annu. Rev. Fluid Mech. 2015.47. Downloaded from www.annualreviews.org

aerodynamic damping with turbulent trips at higher Mach numbers associated with the advancing
rotor merits watching as a possible trigger for stall flutter. In contrast, the edge roughness trip
by Tulane University on 10/14/14. For personal use only.

had little effect on min at M ∞ = 0.6.

3. DYNAMIC STALL CONTROL


The sensitivity of dynamic stall vortex formation and development to leading-edge turbulent trips
suggests that passive and active flow control could be an effective approach toward mitigating the
detrimental effect of dynamic stall. There have been several investigations of dynamic stall control
by both active and passive means, as described in the survey by Lorber et al. (2000). Candidate
control schemes include variable rotor geometry, leading-edge blowing (Greenblatt & Wygnanski
2001, Sun & Sheikh 1999, Weaver et al. 2004), leading-edge plasma actuation (Corke et al. 2011;
Lombardi 2011; Lombardi et al. 2013; Post & Corke 2006), vortex generators (Heine et al. 2011,
Martin et al. 2008, Traub et al. 2004), synthetic jets (Ekaterinaris 2002, Florea & Wake 2003,
Traub et al. 2004), and fixed-wing devices such as slots (Carr et al. 2001), leading-edge droop
(Chandrasekhara et al. 2004, Joo et al. 2006, Martin et al. 2008), pressure-side tabs (i.e., Gurney
flaps) (Chandrasekhara et al. 2004, Joo et al. 2006), and trailing-edge flaps (Feszty et al. 2004;
Gerontakos & Lee 2006, 2007, 2008).
One particularly notable experiment by Martin et al. (2008) examined passive streamwise vortex
generators for dynamic stall control at helicopter-relevant conditions. Figure 9 shows a photo-
graph of the transonic leading-edge glove and streamwise vortex generator (see the sidebar Tran-
sonic Glove). Vane-type vortex generators were placed at x = 0.1 on a VR-7 airfoil that was also
fitted with a transonic leading-edge glove. These experiments were performed at Re c 2 × 106
and M ∞ = 0.3−0.4. The experiments showed that vortex generators on the baseline airfoil were
not effective in dynamic stall control. Instead, they caused shock-induced dynamic stall in the
locally supersonic flow. Similarly, the transonic glove alone was also ineffective for dynamic stall
control. However, the combination of the transonic glove, which lowered the local Mach num-
ber, and the vortex generators was found to be extremely effective at suppressing dynamic stall.
Measured pitch moment cycles exhibited a significant reduction or elimination of moment stall
with the transonic glove and streamwise vortex generator combination. However, the effective-
ness of the transonic glove and passive vortex generator combination still diminished at Mach
numbers above M ∞ = 0.3−0.4, most likely as a result of shocks created by the passive vortex
generator.

www.annualreviews.org • Dynamic Stall in Pitching Airfoils 497

Changes may still occur before final publication online and in print
FL47CH20-Corke ARI 16 September 2014 12:22
Annu. Rev. Fluid Mech. 2015.47. Downloaded from www.annualreviews.org

Figure 9
Photograph of transonic leading-edge glove with streamwise vortex generator used by Martin et al. (2008)
for dynamic stall control. Figure courtesy of P.B. Martin.
by Tulane University on 10/14/14. For personal use only.

At helicopter-relevant conditions, it would appear that any flow control approach that geo-
metrically extends above the airfoil surface will be a detriment, at least on the advancing blade,
where it can generate shock waves in the locally supersonic flow. To avoid this problem, Post
& Corke (2006) and Post (2004) investigated the use of surface-mounted plasma actuators for
dynamic stall control on pitching airfoils (see the sidebar Plasma Actuators). A general overview
of plasma actuators is given by Corke et al. (2010). The plasma actuator was located at the very
leading edge, where it was found to be most effective in controlling leading-edge flow separation
and extending stall on stationary airfoils. The actuator spanned the leading edge of the airfoil. It
was oriented so that when activated, it would produce a two-dimensional body force across the
span that would accelerate the boundary layer flow in the direction from the leading edge toward
the suction surface of the airfoil. In this manner, it was designed to augment the separation delay

PLASMA ACTUATORS

The term plasma actuator has now been a part of the fluid dynamics flow control vernacular for more than a
decade. A particular type of plasma actuator that has gained wide use is based on a single dielectric barrier discharge
mechanism that has desirable features for use in air over a range of static pressures. These plasma actuators most
generally consist of two electrodes, one uncoated and exposed to the air and the other encapsulated by a dielectric
material. The electrodes are typically arranged asymmetrically. The electrodes are supplied with an ac voltage that,
at high-enough levels, causes the air over the covered electrode to weakly ionize. The ionized air appears blue. For
these flow actuators, the mechanism of flow control is through a generated body force vector field that couples with
the momentum in the external flow. The body force can be derived from first principles, and the plasma actuator
effect can be easily incorporated into flow solvers so that their placement and operation can be optimized. They
have been used in a wide range of internal and external flow applications. Although initially considered to be useful
only at low speeds, they have been shown to be effective in a number of applications at high subsonic, transonic,
and supersonic Mach numbers. This has largely come from more optimized actuator designs that were developed
through better understanding and modeling of the actuator physics. For further information, readers are referred
to Corke et al. (2010).

498 Corke · Thomas

Changes may still occur before final publication online and in print
FL47CH20-Corke ARI 16 September 2014 12:22

mechanism produced by the pitching motion with α̇ > 0. Post & Corke investigated both steady
and unsteady open-loop forcing. In both cases, the actuator operated continuously throughout
the pitching cycle. In the unsteady operation, the plasma actuator was cycled on and off at a fre-
quency, f, that was found to be effective in controlling the separated shear layer. This utilized a
dimensionless frequency, F + = f c /U ∞ = 1, that was found to be most effective in maintaining
attached leading-edge flow on stationary airfoils (Post & Corke 2006). Assuming a convection
speed of vortices generated by the actuator pulsing of 0.5 U∞ , F + = 1, results in two vortices
spaced over the chord of the airfoil. This was verified by flow visualization (Post 2004, Post &
Corke 2006). For the pitching airfoil, the unsteady plasma actuator frequency was 20 times higher
than the pitching frequency. For conditions of deep dynamic stall, the steady actuation produced
a 4.7% improvement in the cycle-integrated lift. The unsteady actuation at F + = 1 was slightly
better, with a 5.4% improvement in the cycle-integrated lift. Although the improvement in the
lift was small between the unsteady and steady actuation, the improvement in the pitch stability
was substantially better with the unsteady actuation.
Annu. Rev. Fluid Mech. 2015.47. Downloaded from www.annualreviews.org

Corke et al. (2011) proposed a method to use an unsteady plasma actuator, along with a high-
frequency response pressure sensor placed near the leading edge, to detect the onset of dynamic
stall, and thereby disrupt the development of the dynamic stall vortex. The method relied upon
by Tulane University on 10/14/14. For personal use only.

the findings of Haddad et al. (2005), who showed that the receptivity of the boundary layer over
a parabolic leading edge to an unsteady disturbance grows on the order of 100 times prior to the
formation of a separation bubble. At pitch angles below the static stall angle, α ss , the attached
boundary layer is not receptive to unsteady disturbances produced by a plasma actuator located
at the leading edge. As the airfoil pitch angle approaches α ss , the separation bubble moves toward
the leading edge. The separation bubble is receptive to the low-amplitude unsteady disturbances
produced by the plasma actuator. These are transmitted to the pressure sensor when the separation
bubble extends from the leading-edge site of the disturbances to the pressure sensor location.
This typically occurs at an angle of attack just below α ss . At this angle of attack, the unsteady
disturbances produced by the plasma actuator are easily detectable by the pressure sensor. The
unsteady disturbances remain detectable as long as the flow remains separated. However, they are
no longer detectable if the flow would naturally remain attached (Corke et al. 2011).
Based on this, Lombardi et al. (2013) devised a close-loop dynamic stall control approach that
introduced a low-amplitude periodic disturbance at F + = 1 using a plasma actuator at the leading
edge that was operated in a low-power sense state. With this approach, the authors sought to detect
the periodic disturbance with a flush-mounted pressure sensor located on the airfoil that would
indicate the onset of leading-edge flow separation. If the disturbance was sensed, the plasma actu-
ator was changed to a higher-powered control state that would reattach the flow. Figure 10 shows
an example of the outcome of the closed-loop control for conditions that correspond to light dy-
namic stall. Figure 10a illustrates the flow visualization and corresponding suction-side pressure
distribution at the maximum angle of attack during pitch-up, and at αmax + 4◦ during pitch-down,
without and with closed-loop flow control. With the closed-loop control, the flow remains at-
tached, with a commensurate increase in the suction pressure peak at the leading edge. Figure 10b
shows the total effect of the closed-loop control on the lift and pitch moment coefficient cycles.
The closed-loop control logic for the determination of the onset of flow separation utilized a
threshold on the level of the periodic disturbance sensed by the pressure sensor. For the left pair
of plots, the threshold was such that the control was operating approximately 47% of the pitching
cycle, starting during pitch-up at 4.5◦ before αmax and ending during pitch-down at α = 8.0◦ . This
is the same threshold condition for the flow visualization and pressure distributions in Figure 10a.
As evident by the lift and moment cycles, this resulted in a substantial reduction in the lift and
moment hysteresis. For the pair of plots on the right, the control operated over only 11.2% of

www.annualreviews.org • Dynamic Stall in Pitching Airfoils 499

Changes may still occur before final publication online and in print
FL47CH20-Corke ARI 16 September 2014 12:22

αmax α = 18° Pitch-down α = 14° Vthresh = 0.6 V Vthresh = 1.5 V


1.2
a b
0.9

0.6

Cr
0.3

10 + 8 sin(2π ft)

Control state Control state 0.05


Annu. Rev. Fluid Mech. 2015.47. Downloaded from www.annualreviews.org

3 0
No control c/4
Vthresh = 0.6 V –0.05
Cm

2
by Tulane University on 10/14/14. For personal use only.

–Cp

–0.10
Control state
1
–0.15 Sense state
Baseline
0 –0.20
0 0.2 0.4 0.6 0.8 1.0 0 0.2 0.4 0.6 0.8 1.0 0 5 10 15 20 25 0 5 10 15 20 25
x/c α (°)

Figure 10
Flow visualization and corresponding suction-side pressure distribution for a pitching airfoil without and with closed-loop control cases
for light dynamic stall with α0 = 10◦ , and α1 = 8◦ (a), and lift and moment coefficient cycles for two stall-detection thresholds (b). In
panel b, the dashed curves correspond to the cycle without flow control, and the solid curve to that with closed-loop control. With the
closed-loop control, the portion of the cycle over which the actuator was operating to control the flow separation is indicated by the red
circles that overlay the solid curve lift and moment cycle curves. Figure adapted with permission from Lombardi et al. (2013).

the pitching cycle starting 2.2◦ before αmax and ending just 0.1◦ past αmax during pitch-down. This
control was essentially only operating during the formation of the dynamic stall vortex. With even
that short duration, lift and moment cycles reveal the same fundamental improvement in the stall
hysteresis that occurred with the significantly longer actuator operation. Quantitatively, the cycle-
averaged lift increased by 12%, and cycle went from slightly negative (−0.007) to positive (0.007)
as a result of the closed-loop control. From an active flow control input energy perspective, this
targeted control provided an 88% improvement over the open-loop control by which the actuator
operated throughout the pitching cycle.

4. SUMMARY
This review highlights the rich flow physics that underlies dynamic stall on a pitching airfoil.
Although it is relevant to many manmade and natural flyers, we emphasize helicopter rotor aero-
dynamics because dynamic stall ultimately limits their maneuverability and agility, as well as speed
and payload.
We place particular focus above on aerodynamic damping because of the tendency for dynamic
stall to lead to flutter, which is an aeroelastic coupling that can result in limit-cycle growth of vibra-
tion amplitudes and eventual structural failure. The recently introduced intracycle aerodynamic

500 Corke · Thomas

Changes may still occur before final publication online and in print
FL47CH20-Corke ARI 16 September 2014 12:22

damping formulation of Bowles (2012) has provided new insight into this process that was pre-
viously masked by the traditional cycle-averaged view. Particularly significant is the observation
that the intracycle damping can exhibit relatively large negative peaks over a wide range of Mach
numbers, even in cases with positive cycle-averaged damping. In this manner, successive rotations
of the rotor through negatively damped retreating blade regions can lead to additive limit-cycle
growth, even though the cycle-averaged damping remains positive.
The phase relation between the airfoil pitching motion and the aerodynamic pitch moment
determines the sign of the aerodynamic damping. The intracycle damping analysis makes it clear
that the initial formation and convection of the organized dynamic stall vortex prior to reaching the
maximum angle of attack are the sources of strong negative aerodynamic damping. This contrasts
with earlier interpretations (Liiva 1969) that attributed negative aerodynamic damping to the
sudden flow separation at stall and the combination of the nose-down pitch moment and α̇ < 0
at αmax . This has important implications for damping control schemes, which should properly be
focused on weakening the organized vorticity.
Annu. Rev. Fluid Mech. 2015.47. Downloaded from www.annualreviews.org

The dynamic stall process is highly sensitive to surface roughness that can influence turbulent
transition in the separating boundary layer and to local compressibility effects that occur at free-
stream Mach numbers that are otherwise incompressible for the stationary airfoil. In lieu of shock
by Tulane University on 10/14/14. For personal use only.

formation at higher Mach numbers, dynamic stall originates with a sudden eruption of vorticity
from the boundary layer near the leading edge, which leads to a strong viscous/inviscid interaction
with the outer flow. This gives rise to a well-defined dynamic stall vortex, which then convects aft
over the airfoil. It is the formation and subsequent evolution of the dynamic stall vortex that causes
massive excursions of the aerodynamic loads and pitch moment from linear thin airfoil theory.
Compressibility has been shown to alter the onset mechanism for dynamic stall, promoting
leading-edge stall through separation bubble bursting. In addition, compressibility effects gener-
ally weaken the dynamic stall vortex strength. However, if the Mach number is high enough to
form shocks downstream of the leading edge, the dynamic stall vortex forms downstream of the
shock front, and its topology differs from that at lower Mach numbers. With such shock-induced
dynamic stall, any flow separation at the leading edge has little effect on the aerodynamic loads
and moments.
Numerous studies cited in this review have provided descriptions of the processes that char-
acterize both light and deep dynamic stall regimes, as well as the influence of parameters such as
airfoil geometry, reduced frequency, stall penetration angle, and Reynolds and Mach numbers.
However, a unified understanding of the physics surrounding the pitching airfoil flow field still re-
mains elusive. In fact, many aspects of dynamic stall inception in even incompressible flow remain
incomplete. Although compressibility effects have been well characterized, the physical mecha-
nisms associated with them are, in many cases, not fully understood. There remains a particular
need for better understanding regarding the physics of shock-induced dynamic stall that can occur
at higher subsonic Mach numbers.
At Mach numbers below which shocks occur, turbulent trips in the form of distributed rough-
ness near the leading edge are able to weaken the dynamic stall vortex and, as a result, positively
increase the aerodynamic damping. However, at higher Mach numbers associated with shock-
induced dynamic stall, the turbulent trips produced large negative intracycle damping. Because
helicopter rotor leading-edge surfaces are generally rough, this could be the source of unexpected
aeroelastic flutter on the advancing rotor. Ultimately, the sensitivity of dynamic stall to leading-
edge roughness bodes well for the possibility of dynamic stall control with minimal power input.
Above we cite several studies that have investigated dynamic stall control by both active
and passive means. Regardless of the approach, the focus is to (a) increase cycle-integrated lift,
(b) reduce impulsive blade pitching moments, and (c) assure a positive cycle-averaged aerodynamic

www.annualreviews.org • Dynamic Stall in Pitching Airfoils 501

Changes may still occur before final publication online and in print
FL47CH20-Corke ARI 16 September 2014 12:22

damping and minimize the intracycle negative damping that can occur during portions of the
pitching cycle. If the dynamic stall boundary can be extended via a suitable control scheme,
a revolutionary stall-free rotor could become a reality. However, despite the demonstrated
successes reported in marginalizing some detrimental effects of dynamic stall, few of the flow
control approaches have been demonstrated at the combination of Rec and M∞ that is applicable
to the full rotor environment. It is recommended that this be a focus of future dynamic stall flow
control studies. Of particular interest are smart active control schemes that can sense dynamic
stall onset and then exploit a shear layer receptivity mechanism in such a manner as to mitigate
the detrimental effect of dynamic stall with minimal power input. Particularly attractive are flush
surface actuator designs that can operate at high subsonic Mach numbers without giving rise to
the detrimental effects associated with shock formation.

DISCLOSURE STATEMENT
Annu. Rev. Fluid Mech. 2015.47. Downloaded from www.annualreviews.org

The authors are not aware of any biases that might be perceived as affecting the objectivity of this
review.
by Tulane University on 10/14/14. For personal use only.

ACKNOWLEDGMENTS
The authors would like to acknowledge Dr. Patrick Bowles, Mr. Anthony Lombardi, and Mr.
Dustin Coleman at the University of Notre Dame; Dr. Mark Wasikowsi and Mr. Thomas Wood
at Bell Helicopter; and Dr. Preston Martin at NASA Langley Research Center for insightful
discussions related to this article.

LITERATURE CITED
Bailey FJ, Gustafson FB. 1939. Observations in flight of the region of stalled flow over the blades of an autogiro rotor.
Tech. Rep. 741, Natl. Advis. Comm. Aeronaut., Washington, DC
Bisplinghoff RL, Ashley H, Halfman RL. 1955. Aeroelasticity. Reading, MA: Addison-Wesley
Bousman WG. 1998. A qualitative examination of dynamic stall from flight test data. J. Am. Helicopter Soc.
43:279–95
Bowles P. 2012. Wind tunnel experiments on the effect of compressibility on the attributes of dynamic stall. PhD
Thesis, Univ. Notre Dame
Bowles P, Coleman DG, Corke TC, Thomas FO, Wasikowski M. 2012. Compressibility effects on aerodynamic
damping during dynamic stall events. Presented at Annu. Meet. Am. Helicopter Soc., 68th, Fort Worth,
TX
Bowles P, Corke T, Coleman D, Thomas F. 2014. Improved understanding of aerodynamic damping through
Hilbert transform. AIAA J. In press
Brooks GW, Baker JE. 1958. An experimental investigation of the effect of various parameters including tip Mach
number on the flutter of some model helicopter rotor blades. Tech. Rep. TN 4005, Natl. Advis. Comm.
Aeronaut., Washington, DC
Carr LW, Chandrasekhara MS. 1991. A study of compressibility effects on dynamic stall of rapidly pitching
airfoils. Comput. Phys. Commun. 65:62–68
Carr LW, Chandrasekhara MS. 1992. Design and development of a compressible dynamic stall facility.
J. Aircr. 29:314–18
Carr LW, Chandrasekhara MS. 1996. Compressibility effects on dynamic stall. Prog. Aerosp. Sci. 32:523–73
Carr LW, Chandrasekhara MS, Brock NJ. 1994. Quantitative study of unsteady compressible flow on an
oscillating airfoil. J. Aircr. 31:892–98
Carr LW, Chandrasekhara MS, Wilder MC, Noonan KW. 2001. Effect of compressibility on suppression of
dynamic stall using a slotted airfoil. J. Aircr. 38:296–309

502 Corke · Thomas

Changes may still occur before final publication online and in print
FL47CH20-Corke ARI 16 September 2014 12:22

Carr LW, McCroskey WJ, McAlister KE, Pucci SL, Lambert O. 1982. An experimental study of dynamic stall
on advanced airfoil sections. Volume 3: Hot-wire and hot-film measurements. Tech. Rep. USAAVRADCOM
TR-82-A-8, Natl. Aeronaut. Space Admin., Washington, DC
Carta FO. 1971. Effect of unsteady pressure gradient reduction on dynamic stall delay. J. Aircr. 8:839–41
Carta FO, Niebanck CF. 1969. Prediction of rotor instability at high forward speeds. Tech. Rep. 44-177-AMC-
332(T), US Army Aviation Mater. Lab., Fort Eustis, VA
Chandrasekhara MS, Ahmed S, Carr LW. 1993. Schlieren studies of compressibility effects on dynamic stall
of transiently pitching airfoils. J. Aircr. 30:213–20
Chandrasekhara MS, Carr LW, Wilder MC. 1994. Interferometric investigations of compressible dynamic
stall over a transiently pitching airfoil. AIAA J. 32:586–93
Chandrasekhara MS, Tung C, Martin PB. 2004. Aerodynamic flow control using a variable droop leading edge airfoil.
Presented at AVT Specialists’ Meet. Enhanc. NATO Mil. Flight Perform., Pap. RTO-MP-AVT-111
Chandrasekhara MS, Wilder MC, Carr LW. 1996. Boundary-layer-tripping studies of compressible dynamic
stall flow. AIAA J. 34:96–103
Chandrasekhara MS, Wilder MC, Carr LW. 1998a. Competing mechanisms of compressible dynamic stall.
Annu. Rev. Fluid Mech. 2015.47. Downloaded from www.annualreviews.org

AIAA J. 36:383–93
Chandrasekhara MS, Wilder MC, Carr LW. 1998b. Unsteady stall control using dynamically deforming
airfoils. AIAA J. 36:1792–800
by Tulane University on 10/14/14. For personal use only.

Conlisk A. 2001. Modern helicopter rotor aerodynamics. Prog. Aerosp. Sci. 37:419–76
Corke TC, Bowles PO, He C, Matlis EH. 2011. Sensing and control of flow separation using plasma actuators.
Philos. Trans. R. Soc. A 369:1459–75
Corke TC, Enloe CL, Wilkinson SP. 2010. Dielectric barrier discharge plasma actuators for flow control.
Annu. Rev. Fluid Mech. 42:505–29
Doligalski T, Smith C, Walker J. 1994. Vortex interactions with walls. Annu. Rev. Fluid Mech. 26:573–616
Dyken RV, Ekaterinaris JA, Chandrasekhara MS, Platzer MF. 1996. Analysis of compressible light dynamic
stall flow at transitional Reynolds numbers. AIAA J. 34:1420–27
Ekaterinaris JA. 2002. Numerical investigations of dynamic stall active control for incompressible and com-
pressible flows. J. Aircr. 39:71–78
Ericsson LE, Reding JP. 1984. Shock-induced dynamic stall. J. Aircr. 21:316–21
Ericsson LE, Reding JP. 1988. Moving wall effects in unsteady flow. J. Aircr. 25:977–90
Feszty D, Gillies EA, Vezza M. 2004. Alleviation of airfoil dynamic stall moments via trailing-edge-flap flow
control. AIAA J. 42:17–25
Florea R, Wake BE. 2003. Parametric analysis of directed-synthetic jets for improved dynamic-stall performance.
Presented at AIAA Aerosp. Sci. Meet., 41st, Reno, NV, AIAA Pap. 2003-0216
Fung YC. 2008. An Introduction to the Theory of Aeroelasticity. Toronto: Dover
Gerontakos P, Lee T. 2006. Dynamic stall flow control via a trailing-edge flap. AIAA J. 44:469–80
Gerontakos P, Lee T. 2007. Trailing-edge flap control of dynamic pitching moment. AIAA J. 45:1688–94
Gerontakos P, Lee T. 2008. PIV study of flow around unsteady airfoil with trailing edge flap deflection. Exp.
Fluids 45:955–72
Glauert H, Holl H. 1929. Die Grundlagen der Tragflügel- und Luftschraubentheorie. Berlin: Springer
Green RB, Galbraith RAM. 1995. Dynamic recovery to fully attached aerofoil flow from deep stall. AIAA J.
33:1433–40
Greenblatt D, Wygnanski I. 2001. Dynamic stall control by periodic excitation, part 1: NACA 0015 parametric
study. J. Aircr. 38:430–38
Gustafson FB, Myers GC Jr. 1946. Stalling of helicopter blades. Tech. Rep. TN 1083, Natl. Advis. Comm.
Aeronaut., Washington, DC
Haddad O, Erturk E, Corke T. 2005. Acoustic receptivity of boundary layer over parabolic bodies at angles
of attack. J. Fluid Mech. 536:377–400
Haller G. 2004. Exact theory of unsteady separation for two-dimensional flows. J. Fluid Mech. 512:257–311
Ham ND. 1966. Torsional oscillation of helicopter blades due to stall. J. Aircr. 3:218–24
Ham ND. 1967. Stall flutter of helicopter rotor blades: a special case of the dynamic stall phenomenon.
J. Am. Helicopter Soc. 12:19–21

www.annualreviews.org • Dynamic Stall in Pitching Airfoils 503

Changes may still occur before final publication online and in print
FL47CH20-Corke ARI 16 September 2014 12:22

Ham ND. 1968. Aerodynamic loading on a two-dimensional airfoil during dynamic stall. AIAA J. 6:1927–34
Ham ND, Garelick MS. 1968. Dynamic stall considerations in helicopter rotors. J. Am. Helicopter Soc. 13:49–55
Ham ND, Young MI. 1966. Limit cycle torsional motion of helicopter blades due to stall. J. Sound Vib.
4:431–44
Harris FD Jr, Tarzanin FJ Jr, Fisher RK. 1970. Rotor high speed performance, theory versus test. J. Am.
Helicopter Soc. 15:35–42
Heine B, Mulleneres K, Joubert G, Raffel M. 2011. Dynamic stall control by passive disturbance generators.
Presented at AIAA Appl. Aerodyn. Conf., 29th, Honolulu, AIAA Pap. 2011-3371
Johnson W, Ham ND. 1972. On the mechanism of dynamic stall. J. Am. Helicopter Soc. 17:36–45
Joo W, Lee B-S, Yee K, Lee D-H. 2006. Combining passive control method for dynamic stall control.
J. Aircr. 43:1120–28
Kramer M. 1932. Die Zunahme des maximalauftriebes von tragflugeln bei plotzlicher anstellwinkelver-
grosserung (Boeneffekt). Z. Flugtech. Motorluftschiff. 23:185–89
Küssner H. 1936. Zusammenfassender Bericht über den instationären auftrieb von flügeln. Luftfahrt-Forschung
13:410–24
Annu. Rev. Fluid Mech. 2015.47. Downloaded from www.annualreviews.org

Lee T, Gerontakos P. 2004. Investigation of flow over an oscillating airfoil. J. Fluid Mech. 512:313–41
Leishman JG. 2000. Principles of Helicopter Aerodynamics. Cambridge, UK: Cambridge Univ. Press
Liiva J. 1969. Unsteady aerodynamic and stall effects on helicopter rotor blade airfoil sections. J. Aircr. 6:46–51
by Tulane University on 10/14/14. For personal use only.

Liiva J, Davenport F. 1969. Dynamic stall of airfoil sections for high speed rotors. J. Am. Helicopter Soc.
14:26–33
Lomax H. 1953. Lift development on unrestrained rectangular wings entering gusts and subsonic and supersonic speeds.
Tech. Rep. 2925, Natl. Advis. Comm. Aeronaut., Washington, DC
Lombardi AJ. 2011. Closed-loop dynamic stall control using a plasma actuator. MS Thesis, Univ. Notre Dame
Lombardi AJ, Bowles PO, Corke TC. 2013. Closed-loop dynamic stall control using a plasma actuator. AIAA
J. 51:1130–41
Lorber PF, Carta FO. 1988. Airfoil dynamic stall at constant pitch rate and high Reynolds number. J. Aircr.
25:548–56
Lorber PF, Carta FO, Covino AF. 1992. An oscillating three-dimensional wing experiment: compressibility, sweep,
rate, waveform, and geometry effects on unsteady separation and dynamic stall. Tech. Rep. R92-958325-6,
United Technol. Res. Cent., East Hartford, CT
Lorber PF, McCormick DC, Anderson TJ, Wake BE, MacMartin GG, et al. 2000. Rotorcraft retreating blade
stall control. Presented at FLUIDS 2000 Conf. Exhib., Denver, AIAA Pap. 2000-2475
Martin PB, Wilson JS, Berry JD, Wong T-C, Moultron M, McVeigh MA. 2008. Passive control of compressible
dynamic stall. Presented at AIAA Appl. Aerodyn. Conf., 26th, Honolulu, AIAA Pap. 2008-7506
McAlister KW, Carr LW. 1979. Water tunnel visualizations of dynamic stall. J. Fluids Eng. 101:376–80
McAlister KW, Pucci SL, McCroskey WJ, Carr LW. 1982. An experimental study of dynamic stall on advanced air-
foil sections. Volume 2: Pressure and force data. Tech. Rep. USAAVRADCOM TR-82-A-8, Natl. Aeronaut.
Space Admin., Washington, DC
McCroskey WJ. 1981. The phenomenon of dynamic stall. Tech. Rep., Natl. Aeronaut. Space Admin., Washington,
DC
McCroskey WJ. 1982. Unsteady airfoils. Annu. Rev. Fluid Mech. 14:285–311
McCroskey WJ, Carr LW, McAlister KW. 1976. Dynamic stall experiments on oscillating airfoils. AIAA J.
14:57–63
McCroskey WJ, McAlister KW, Carr LW, Pucci SL. 1982. An experimental study of dynamic stall on advanced
airfoil section. Volume 1: Summary of the experiment. Tech. Rep. USAAVRADCOM R-82-A-8, Natl.
Aeronaut. Space Admin., Washington, DC
McCroskey WJ, McAlister KW, Carr LW, Pucci SL, Lambert O, Indergrand RF. 1981. Dynamic stall on
advanced airfoil sections. J. Am. Helicopter Soc. 26:40–50
Oates GC, ed. 1989. Aircraft Propulsion Systems Technology and Design. Reston, VA: Am. Inst. Aeronaut.
Astronaut.
Post ML. 2004. Plasma actuators for separation control on stationary and oscillating airfoils. PhD Thesis, Univ.
Notre Dame

504 Corke · Thomas

Changes may still occur before final publication online and in print
FL47CH20-Corke ARI 16 September 2014 12:22

Post ML, Corke TC. 2006. Separation control using plasma actuators: dynamic stall vortex control on oscil-
lating airfoil. AIAA J. 44:3125–35
Rainey AG. 1957. Measurement of aerodynamic forces for various mean angles of attack on an airfoil oscillating in
pitch and on two finite-span wings oscillating in bending with emphasis on damping in the stall. Tech. Rep. 1305,
Natl. Advis. Comm. Aeronaut., Washington, DC
Reynolds WC, Carr LW. 1985. Review of unsteady, driven, separated flows. Presented at AIAA Shear Flow Conf.,
Boulder, CO, AIAA Pap. 85-0527
Schatzman DM, Thomas FO. 2010. Turbulent boundary-layer separation control with a single dielectric
barrier discharge plasma actuators. AIAA J. 48:1621–34
Sears WR. 1941. Some aspects of non-stationary airfoil theory and its practical application. J. Aeronaut. Sci.
8:104–8
Sears WR, Telionis DP. 1975. Boundary-layer separation in unsteady flow. SIAM J. Appl. Math. 28:215–35
Shih C, Lourenco LM, van Dommelen LL, Krothapalli A. 1992. Unsteady flow past an airfoil pitching at
constant rate. AIAA J. 30:1153–61
Sun M, Sheikh S. 1999. Dynamic stall suppression on an oscillating airfoil by steady and unsteady blowing.
Annu. Rev. Fluid Mech. 2015.47. Downloaded from www.annualreviews.org

Aerosp. Sci. Technol. 6:355–66


Tarzanin FJ Jr. 1972. Prediction of control loads due to blade stall. J. Am. Helicopter Soc. 17:33–46
Telionis DP. 1970. Boundary-layer separation. PhD Thesis, Cornell Univ., Ithaca, NY
by Tulane University on 10/14/14. For personal use only.

Theodorsen T. 1935. General theory of aerodynamic instability and the mechanism of flutter. Tech. Rep. 496, Natl.
Advis. Comm. Aeronaut., Washington, DC
Traub LW, Miller A, Rediniotis O. 2004. Effects of active and passive flow control on dynamic-stall vortex
formation. J. Aircr. 41:405–8
van Dommelen LL, Shen SF. 1980. The spontaneous generation of the singularity in a separating laminar
boundary layer. J. Comput. Phys. 38:125–40
von Kármán T, Sears WR. 1938. Airfoil theory for non-uniform motion. J. Aeronaut. Sci. 5:379–90
Wagner H. 1925. Über die Entstehung des dynamischen Äuftriebes von Tragflügeln. Z. Angew. Math. Mech.
5:17–35
Walker JM, Helin HE, Strickland JH. 1985. An experimental investigation of an airfoil undergoing large-
amplitude pitching motions. AIAA J. 23:1141–42
Weaver D, McAlister KW, Tso J. 2004. Control of VR7 dynamic stall by strong steady blowing. J. Aircr.
41:1404–13
Wernert P, Geissler W, Raffel M, Kompenhans J. 1996. Experimental and numerical investigations of dynamic
stall on a pitching airfoil. AIAA J. 34:982–89
Wernert P, Koerber G, Wietrich F, Raffel M, Kompenhans J. 1997. Demonstration by PIV of the non-
reproducibility of the flow field around an airfoil pitching under deep dynamic stall conditions and
consequences thereof. Aerosp. Sci. Technol. 2:125–35
Young WH Jr. 1981. Fluid mechanics mechanisms in the stall process for helicopters. Tech. Rep. NASA-TM-81956,
Natl. Aeronaut. Space Admin., Washington, DC
Zvara J, Ham ND. 1960. Helicopter rotor model research at Massachusetts Institute of Technology. J. Am.
Helicopter Soc. 5:24–30

www.annualreviews.org • Dynamic Stall in Pitching Airfoils 505

Changes may still occur before final publication online and in print

You might also like