You are on page 1of 30

FL51CH04_Eldredge ARI 4 August 2018 12:17

Annual Review of Fluid Mechanics

Leading-Edge Vortices:
Mechanics and Modeling
Jeff D. Eldredge1 and Anya R. Jones2
Access provided by University of Central Florida on 08/18/18. For personal use only.

1
Department of Mechanical and Aerospace Engineering, University of California, Los Angeles,
Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

California 90095, USA; email: eldredge@seas.ucla.edu


2
Department of Aerospace Engineering, University of Maryland, College Park,
Maryland 20742, USA; email: arjones@umd.edu

Annu. Rev. Fluid. Mech. 2019. 51:75–104 Keywords


The Annual Review of Fluid Mechanics is online at leading-edge vortex, translating, surging, rotating, pitching, LEV, vortex
fluid.annualreviews.org
model, data-driven, unsteady aerodynamics
https://doi.org/10.1146/annurev-fluid-010518-
040334 Abstract
Copyright  c 2019 by Annual Reviews. The leading-edge vortex (LEV) is known to produce transient high lift in
All rights reserved
a wide variety of circumstances. The underlying physics of LEV formation,
growth, and shedding are explored for a set of canonical wing motions in-
cluding wing translation, rotation, and pitching. A review of the literature
reveals that, while there are many similarities in the LEV physics of these
motions, the resulting force histories can be dramatically different. In two-
dimensional motions (translation and pitch), the LEV sheds soon after its
formation; lift drops as the LEV moves away from the wing. Wing rotation,
in contrast, incites a spanwise flow that, through Coriolis tilting, balances
the streamwise vorticity fluxes to produce an LEV that remains attached to
much of the wing and thus sustains high lift. The state of the art of vortex-
based modeling to capture both the flow field and corresponding forces of
these motions is reviewed, including closure conditions at the leading edge
and approaches for data-driven strategies.

75
Review in Advance first posted on
August 15, 2018. (Changes may still
occur before final publication.)
FL51CH04_Eldredge ARI 4 August 2018 12:17

1. INTRODUCTION
Understanding and achieving high lift is a recurring theme in aerodynamics. Figure 1 gives a
Reynolds number: smattering of airfoil data for the maximum lift coefficient achieved over a wide range of Reynolds
the ratio of inertial to numbers. At the relatively high Reynolds numbers typical of manned flight, high maximum lift
viscous effects, coefficients are easily attained as flow remains attached through reasonably severe pressure gradi-
Re = Uref c/ν
ents. The maximum lift coefficient falls off quickly in transitional flows but can be recovered by
installing turbulators to ensure a turbulent boundary layer and encourage attached flow. At lower
Reynolds numbers, however, boundary layers remain laminar and are more prone to separate in
adverse pressure gradients; airfoils stall at low angles of attack and low lift values. It is in this
regime that insects fly, inviting the question of how flight is possible if, in steady flow, their wings
produce only a quarter of the lift required. In a quest to explain the fundamentals of natural flight,
biologists performed numerous studies of live insects, mounted dried wings in wind tunnels, and
built mechanical models of the insect wing stroke (Dickinson et al. 1999, Ellington et al. 1996,
Access provided by University of Central Florida on 08/18/18. For personal use only.

Usherwood & Ellington 2002, Willmott & Ellington 1997, Willmott et al. 1997). In doing this,
Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

they discovered a flow structure that arose from unsteady and revolving motions of the wing but
failed to appear in conventional steady flow over the same wing—the leading-edge vortex, or, its
briefer sobriquet, the LEV.
Interest in the LEV grew quickly, due in part to efforts to develop agile small-scale vehicles
and biomimetic fliers and swimmers. The engineering approach to the high lift problem has been
2.0

U N S T E A DY FLOW Airfoils with


1.8 Revolving turbulators
wings
1.6

Hovering
1.4
insects

1.2
Smooth airfoils
C L max
1.0

0.8

0.6

0.4
Insect wings S T E AD Y FLOW

0.2
10 2 10 3 10 4 10 5 10 6 10 7
Reynolds number

Laminar Transitional Turbulent

Figure 1
Maximum lift coefficient, C Lmax , as a function of Reynolds number in steady and unsteady flows. Figure
adapted with permission from Jones & Babinsky (2010).

76 Eldredge
Review in Advance first posted on
· Jones

August 15, 2018. (Changes may still


occur before final publication.)
FL51CH04_Eldredge ARI 4 August 2018 12:17

to distill the complex insect wing stroke into canonical motions in which the LEV is a prominent
feature of the resulting flow. The current review focuses primarily on these canonical flows, largely
because they are simple enough to allow us to establish the underlying physics of the LEV. As we
discuss in detail below, the net effect of the LEV is to induce a transient fluid dynamic load on
the wing. This has benefits (e.g., for dynamic soaring, energy harvesting, or initiating aggressive
vehicle maneuvers), but can also lead to undesirable fatigue loading and vibrations, and in some
cases, can induce unwanted vehicle motions (e.g., in a gust encounter or air wake). The timescale
over which the LEV forms and sheds is fast—only a few convective times. This is within an insect
wing stroke, thus making natural flight possible at these scales. It is, however, also very fast with
respect to our ability to effect control, whether that be locally using flow control or more globally
at the vehicle level. The future development of more maneuverable and gust-tolerant vehicles thus
depends on our ability to better understand and control leading-edge separation and the resulting
vortex-dominated flow.
Access provided by University of Central Florida on 08/18/18. For personal use only.

The significant enhancement of lift attributable to the LEV has motivated several questions
Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

that have guided most investigations of this topic: What parameters control the growth of the LEV
and its circulation? What mechanisms govern its eventual shedding, and importantly, under what
circumstances might it remain attached to the wing? Can the behavior of the LEV—its evolution
and the force it exerts on the wing—be predicted with a simple mathematical model? Is there
hope for the control of LEVs, either to mitigate transients in gust encounters and other unsteady
flows or to enhance high lift on maneuvering wings? The current state of the answers to these
questions is reviewed here. In the next section, some fundamentals of LEV growth and shedding
are presented. Following that, a set of canonical problems and their governing parameters are
discussed, including a review of LEV formation during these motions. We then summarize the
state of mathematical modeling of these flows and the resulting unsteady loading on the wing.
The review concludes with a brief discussion of avenues for future work.

2. FUNDAMENTAL FEATURES OF THE LEADING-EDGE VORTEX


Many of the general features of incipient leading-edge flow separation—the formation and evolu-
tion of the LEV and its resulting influence on force—can be observed in the fluid’s response
to the simplest of unsteady motions on the most basic of aerodynamic surfaces: the impul-
sive rectilinear translation of a two-dimensional (2D) flat plate at large angle of attack. An ex-
ample of this flow is illustrated in Figure 2, which depicts the streamlines and vorticity field
of the flow developing over a flat plate of chord c and thickness 0.023c traveling at 35◦ incidence
and Reynolds number Uc/ν = 500, where U is the plate’s speed and ν the kinematic viscosity of
the fluid. Time is measured in convective units, t ∗ = tU/c, with the motion initiated at t ∗ = 0.
The results in Figure 2 were obtained using high-fidelity numerical simulation with the lattice
Green’s function/immersed boundary projection method (Liska & Colonius 2017). From the start
of the motion, the thin wing profile ensures a separation point that remains nearly pinned to the
leading edge. Vorticity generated at the edge forms a shear layer that quickly rolls up into the
LEV, clearly visible at t ∗ = 0.5 (see the sidebar titled Vortex Identification). At early times, the
stagnation streamline forms a closed recirculation region on the upper side of the plate that occu-
pies only the forwardmost fraction of the chord. As the LEV grows, it induces a region of reversed
flow and secondary vorticity along the surface of the wing under the LEV, evident as early as
t ∗ = 0.5. Together with the base of the feeding shear layer, this develops into a characteristic
lambda-shaped region of counter-rotating vorticity just upstream of the LEV. As plate motion
continues, the LEV is fed by the shear layer and grows in both size and strength. The stagnation
point behind the vortex moves downstream, reaching and eventually passing the trailing edge.

www.annualreviews.org • Leading-Edge Vortices 77


Review in Advance first posted on
August 15, 2018. (Changes may still
occur before final publication.)
FL51CH04_Eldredge ARI 4 August 2018 12:17

t * = 0.5 t* = 1.0 t* = 1.5

t * = 2.0 t* = 2.5 t* = 3.0


Access provided by University of Central Florida on 08/18/18. For personal use only.
Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

Figure 2
Vorticity field (color contours) and streamlines (solid gray lines) from a high-fidelity numerical simulation of flow about a two-dimensional
plate undergoing impulsive translation at angle of incidence α = 35◦ and Reynolds number Re = 500. The stagnation streamline is the
thick black line. Vorticity contours range from −20U/c to 20U/c, where U is the plate speed and c is the chord length. Time is
measured in convective units, t ∗ = tU/c.

After three chord-lengths of travel, the LEV shown in Figure 2 breaks away from the plate and
convects into the wake. The progression of vortex development shown here is similar for other
wing sections at other Reynolds numbers, motion profiles, and angles of attack, so long as there
is large-scale flow separation at the leading edge.
The development and motion of the LEV has a large impact on the fluid dynamic force exerted
on the wing. The histories of the lift coefficient on the plate at various angles of incidence are
shown in Figure 3. Each case follows a similar trend: After the infinite force at t ∗ = 0 due to
the inertial reaction of the fluid from the impulsive start, the lift resets to a finite value due to
the release of the starting vortex at the trailing edge, consistent with the Wagner effect (Wagner
1925). The lift increases from this initial value, quickly at first and then more gradually, to a peak
near t ∗ = 2 that reaches as high as C L = 2.5, after which it drops rapidly. The rise in lift is stronger
and faster, and the subsequent drop is earlier, as the angle of incidence increases toward 45◦ . All
of these behaviors are attributable to the dynamics of the LEV, illustrated in Figure 3b. Here,
the pressure coefficient along the suction side of the wing is shown for the same case of α = 35◦
presented in Figure 2. Throughout the evolution of the flow, it is clear that there is a region of
low pressure directly under the LEV. This region grows in both magnitude and size as the LEV
develops (for t ∗ < 2.2) and then begins to attenuate as the vortex breaks off of the wing and moves

VORTEX IDENTIFICATION

In experiments and high-fidelity simulations, the boundaries of the LEV may be defined by various techniques,
including the 1 and 2 functions (Graftieaux et al. 2001), the Q criterion ( Jeong & Hussain 1995), and Lagrangian
coherent structure analysis (Haller 2001).

78 Eldredge
Review in Advance first posted on
· Jones

August 15, 2018. (Changes may still


occur before final publication.)
FL51CH04_Eldredge ARI 4 August 2018 12:17

4 4
a b
α 2
15°
3 20°
25° 0
30°
35° t * = 0.2 t * = 1.0
40° –2
CL 2 45°
U –C p
35º 4

2
1
0
t * = 1.8 t * = 2.6
0 –2
0 1 2 3 4 5
t*
Access provided by University of Central Florida on 08/18/18. For personal use only.
Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

Figure 3
(a) Lift coefficient, C L , generated by a two-dimensional flat plate undergoing an impulsive translation at various angles of incidence α at
Reynolds number 500. Time is measured in convective units, t ∗ = tU/c, where U is the plate speed and c is the chord length. The
purple circles correspond to the snapshots at 35◦ in panel b, showing the surface pressure coefficient, C p (red line), and vorticity field
(color contours).

away, leading to the drop in lift observed at t ∗ ≈ 2.5. In the physical world, where the flow is
not constrained to remain 2D, 3D instabilities generally cause the shed vortex structures to lose
their coherence; the flow proceeds toward a fully separated state with a concomitant loss of lift
and increase of drag. In summary, the LEV enhances the lift on the wing to values significantly
above what can be achieved in a fully developed separated flow at long times, but this enhancement
survives only as long as the LEV remains near the wing.

2.1. Control Volume Analysis


The discussion of LEV growth and evolution in this review is facilitated by definition of a control
volume suitable for accounting for the various fluxes into the vortex. The control volume depicted
in Figure 4 is constructed to serve multiple forms of wing geometries and kinematics in either
two or three dimensions; its contour C encloses the LEV and much of its feeding shear layer
at some instant. However, rather than evolving dynamically with the LEV, the control volume
will remain stationary relative to the wing and allow flux through any side. For analysis of the
spanwise variation of these fluxes, the control volume has infinitesimal width dy in that direction.
Conservation of mass in this volume is expressed, per unit spanwise length, as
 

− n · v dC = v y dS. 1.
C ∂ y S(y)

In words, any net inflow of mass through the side (C) of the control volume must be balanced by
an increase in spanwise flow. In a strictly 2D flow, which has no such spanwise outlet, the influxes
must be balanced by an equal outflow. This outflow is initially equal to the rate of spatial growth
of the LEV as it exceeds the fixed boundary C. Later, when the vortex is no longer fed by the shear
layer, the outflow is attributable to the convection of the shed LEV.
The LEV circulation—positive in the configuration depicted in Figure 4—is defined by the
total spanwise vorticity enclosed by C:

(y, t) = ω y dS. 2.
S(y)

www.annualreviews.org • Leading-Edge Vortices 79


Review in Advance first posted on
August 15, 2018. (Changes may still
occur before final publication.)
FL51CH04_Eldredge ARI 4 August 2018 12:17

Side view n

a C
3D view z

b x
ROOT

n S(y – dy)
C
Convective fluxes
S (y) dy
Diffusive fluxes c
LEV, ωy
z

TI P
y
Access provided by University of Central Florida on 08/18/18. For personal use only.

y Vortex lines vy
Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

Spanwise flow
ROOT T IP
ωˆ z

Figure 4
Illustration of control volume used for analysis of the leading-edge vortex (LEV). (a) Side (or two-dimensional) view, with convective
(blue) and diffusive (magenta) fluxes into and out of the control volume. (b) Three-dimensional view on a finite–aspect ratio wing. The
unit-outward normal vector is n, and S(y) is the cross section of the control volume at spanwise location y, bounded by contour C.
(c) Schematic of the Coriolis tilting mechanism on a rotating wing. A rear view of the wing is shown, with the control volume shown as
the red box. Spanwise and vertical vorticity components are denoted by ω y and ω̂z , respectively, and v y is the spanwise velocity.

In a 2D flow, this circulation remains invariant along the span by definition, but in three di-
mensions, it is important to remember that vorticity is divergence-free, with the consequence
that

∂(y, t)
− n · ω dC = . 3.
C ∂y
That is, any spanwise variation of circulation necessitates the emergence of nonspanwise compo-
nents of vorticity to counter this variation, a familiar result from classical lifting-line theory with
important implications for the LEV.
To develop an evolution equation for LEV circulation, we find it useful to write the vorticity
transport equation in the reference frame of the wing—which, to serve the discussion later in this
review, may be rotating with angular velocity —in a form that will emphasize fluxes:
∂ ω̂
+ ∇ · (vω̂) = ∇ · (ω̂v) + ν∇ 2 ω̂, 4.
∂t
where ω̂ = ω + 2 is the vorticity in an inertial reference frame and ω and v are the vorticity
and velocity observed in the frame corotating with the wing, respectively. The Coriolis term is
contained within the first term on the right-hand side, which also accounts for stretching and
tilting of the vorticity ω. Integrating the spanwise component of this equation over the control
volume and then applying the divergence theorem, we obtain an equation for (y, t):
  
∂(y, t) ∂ω y
= − n · vω y dC + ν dC + n · ω̂v y dC. 5.
∂t C C ∂n C

This equation is similar to one used by Wojcik & Buchholz (2014b), but by using the fact that
vorticity is divergence-free, the form used here clarifies the effect of stretching and tilting on the
LEV circulation. In particular, the terms associated with net spanwise convective flux through
S(y) and the stretching of the spanwise vorticity in this direction exactly cancel each other and are

80 Eldredge
Review in Advance first posted on
· Jones

August 15, 2018. (Changes may still


occur before final publication.)
FL51CH04_Eldredge ARI 4 August 2018 12:17

thus absent from this equation. Additionally, the variation in viscous flux in the spanwise direction
is likely negligible and has thus been omitted. The first two terms on the right-hand side describe
the convective and diffusive flux through the boundary C of the control volume, respectively, and
the third term describes the tilting of nonspanwise components of vorticity toward the direction
of the spanwise flow on C. In a strictly 2D flow, only the first two terms are nonzero. Furthermore,
the diffusive flux is only prominent along the part of C adjacent to the wing surface, denoted with
magenta arrows in Figure 4a, where it is likely small compared to the vorticity carried in via
the shear layer. Equation 5, together with the constraints given in Equations 1 and 3, provide a
framework in which to analyze LEV growth and stability; in particular, stability requires that the
right-hand side of Equation 5 is approximately zero.

2.2. Growth and Shedding in Two Dimensions


Access provided by University of Central Florida on 08/18/18. For personal use only.

The early development of the LEV is characterized by monotonic growth in both its size, charac-
Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

terized by a length, D(t), and its strength, quantified by its circulation, (t). As the control volume
analysis of the previous section shows, the rate at which the LEV grows spatially in a strictly 2D
setting is determined by the influx of mass from the shear layer at the leading edge and from the
flow induced between the vortex and the wing, as illustrated in Figure 4a. The first of these is
a function of the shear layer’s thickness and is thus entirely attributable to viscosity. The second
form of influx, in contrast, represents an entrainment of mass into the vortex core and is present
even if the vortex develops in a purely inviscid context, where the shear layer reduces to a spiraling
vortex sheet (Kaden 1931, Pullin & Wang 2004). Although this entrainment in a viscous flow
is somewhat mitigated by the induced boundary layer, it still represents the dominant path for
mass influx (Huang & Green 2015). As indicated by Equation 5, the growth of LEV circulation
is primarily determined by the convective fluxes of vorticity into the LEV. The vorticity in the
shear layer is the primary contributor to this growth, whereas the ingestion of secondary vorticity
in the boundary layer under the vortex—sometimes called “annihilation” (Wojcik & Buchholz
2014b)—tends to slow the growth.
The overall development of the LEV is a complex interplay of viscous and inviscid effects.
At early times, the vortex development is independent of the chord length and, with no imposed
length scale, is therefore necessarily self-similar. Xu & Nitsche (2014) have found that, for an
infinitely thin flat plate moving at normal incidence across a wide range of Reynolds numbers
(larger than 250), the circulation and centroid position of the vortex emerging from either edge
are well predicted at early times by the inviscid solution—the self-similar roll-up of a vortex sheet
from a semi-infinite plate (Pullin 1978). This inviscid solution is an extension of the classical result
by Kaden (1931) and leads to the conclusions that /(Uc) ∼ t ∗1/3 and D/c ∼ t ∗2/3 . To adapt these
results to the LEV, the proportionality constant in these relations must depend on the angle of
incidence (Pullin & Wang 2004).
As the observations suggest, the dependence of LEV development on viscosity is second order
and difficult to succinctly describe in scaling arguments. Furthermore, in the latter stages of its
evolution, various influences appear to be in conflict with each other: At lower Reynolds number,
the shear layer will be thicker, increasing the flux of mass that this shear layer carries into the
LEV, but simultaneously, the induced flow ingested into the vortex will be relatively weaker.
Finally, with viscous influence comes a dependence on the geometry of the leading edge, since the
shape of this edge—most simply measured by a characteristic dimension, LLE —will determine the
thickness of the feeding shear layer. Since LEV development remains essentially independent of
chord length until late in the process, the viscous influence on vortex properties must be through
a leading-edge Reynolds number, ReLE = ULLE /ν. Through differences in this parameter, two

www.annualreviews.org • Leading-Edge Vortices 81


Review in Advance first posted on
August 15, 2018. (Changes may still
occur before final publication.)
FL51CH04_Eldredge ARI 4 August 2018 12:17

wings in the same motion at the same chord Reynolds number but with different nose shapes may
exhibit different behaviors in vortex development.
With a finite length of chord, the LEV cannot continue to grow indefinitely. Inevitably, the
supply of circulation from the shear layer is severed, the vortex becomes a topologically distinct
structure, and it is shed into the wake of the wing. This chord-limiting mechanism for shedding,
sometimes called “bluff-body shedding” (Rival et al. 2014, Widmann & Tropea 2015) is intuitive
and is clearly triggered when the extent of the vortex approaches that of the chord, D/c ≈ 1. It
is identifiable by the bursting of the closed recirculation region, as seen in the results depicted
in Figure 2. As the example in Figure 3 shows, this process is reflected in a dramatic change
in lift. Widmann & Tropea (2015) have postulated a second form of vortex shedding called
“boundary layer eruption”: The ingestion of secondary vorticity accumulates into the lambda-
shaped structure and eventually chokes the feed of vorticity from the shear layer. The appearance
of this mechanism, which is independent of chord length, would necessarily depend on ReLE .
Access provided by University of Central Florida on 08/18/18. For personal use only.

A typical vortex shedding event is likely to be a mixture of these mechanisms. Huang & Green
Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

(2015) have illuminated the kinematics of the vortex shedding process on a pitching flat-plate wing
with Lagrangian coherent structure (LCS) analysis. They showed that the LCS saddle point at the
top of the lambda-shaped structure remains nearly stationary as long as the LEV continues to be
fed by the shear layer. This saddle, however, splits into two at the first break in the feeding shear
layer, and the subsequent motion of the first saddle point can be used to unambiguously identify
the instant at which the LEV sheds from the wing. More subtle changes in the vortex dynamics
(e.g., intermittent vorticity transfer from the shear layer) are indicated by the emergence and
rapid motion of additional saddle points. This process appears to be consistent with the boundary
layer eruption mechanism described by Widmann & Tropea (2015), but shedding occurs nearly
simultaneously with the reattachment point reaching the trailing edge.

3. KINEMATIC FAMILIES AND PARAMETERS


While biological flight has motivated some of the recent study of LEVs, the flows generated by
flapping wings cannot be easily disentangled to isolate the underlying LEV physics. As a result,
many investigators interested in these physics have decomposed flapping, as well as other complex
kinematics in which the LEV is a culprit, into a set of more basic wing motions illustrated in
Figure 5: translation, rotation, and pitch. It is important to stress that the use of the term “de-
compose” does not suggest an underlying superposition principle; it is abundantly clear that no
such principle can exist in these large-amplitude motions with highly nonlinear fluid dynamics.
Rather, this decomposition provides a foundation on which further studies of combinations of
these canonical motions can build. In similar fashion, the geometry of the wing, though assum-
ing widely disparate forms across nature and technology, has generally been studied in a simple
symmetric form: a rigid, rectangular planform, with the profile of a flat plate or a thin NACA
(National Advisory Committee for Aeronautics) section. Of course, even basic motions can take
many forms if left to the whims of individual research groups. The canonical motions and ge-
ometries established in recent years by two consortia of researchers—NATO AVT-202 (North
Atlantic Treaty Organization Applied Vehicle Technology) and the AIAA (American Institute
of Aeronautics and Astronautics) Massively Separated Flows Discussion Group (Ol & Babinsky
2016, Ol et al. 2010)—have aligned many of the LEV-focused investigations and are the target of
much of this review.
Wing translation, illustrated in subpanel iii of Figure 5a, is defined as motion in which the
wing is set at a fixed incidence and either driven in purely rectilinear motion at velocity U(t) or
subjected to a freestream flow with this velocity. (Note that the forces generated by these two

82 Eldredge
Review in Advance first posted on
· Jones

August 15, 2018. (Changes may still


occur before final publication.)
FL51CH04_Eldredge ARI 4 August 2018 12:17

a b c
i ii Ω
α˙ 0
U α˙ 0
x α

t t

K v α˙
iii iv Ω
α t t
U α
0
a α¨
Access provided by University of Central Florida on 08/18/18. For personal use only.

t t
Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

0 1/Ro

Figure 5
Overview of the canonical wing kinematics. (a) Schematics of the wing motions: (i) pitch and translation, (ii) pitch and rotation,
(iii) translation, and (iv) rotation. The vertical axis represents reduced pitch rate, K , and the horizontal axis is the inverse of the Rossby
number, Ro. (b) Position, velocity, and acceleration profiles for the translating and rotating wing. (c) Angle of incidence, pitch rate, and
angular acceleration for the pitching wing. U is the translational speed,  is the angular velocity of rotation, and α̇0 is the rate of
pitching.

different configurations will differ by an effective buoyant force, ρ U̇V , where V is the volume of
the wing.) The most basic form of U(t) is an impulsive start from rest to constant velocity, U.
A more practical form is obtained by a linear ramp to this constant velocity over some specified
distance, giving rise to the kinematics depicted in Figure 5b. Since this motion requires a sudden
change of acceleration at its start and finish, several recent studies have made use of a smoothed
(continuously differentiable) version of the ramp given by Eldredge et al. (2009). Each smooth
change in behavior from constant value to linear ramp is generated by the function
1  
Ga (t) = log (2 cosh at) + at , 6.
2a
which approaches 0 as t → −∞ and is approximately t as t → ∞. The parameter a controls
the degree of smoothing: Its inverse represents the blending time between behaviors. A smooth
increase in velocity from 0 to U between times t1 and t2 , for example, is generated by
U
[Ga (t − t1 ) − Ga (t − t2 )] . 7.
t2 − t1
In any case, U serves as the characteristic velocity, Uref , of the problem and, together with chord
length, c, leads to the dimensionless expression of time in convective units, t ∗ = tU/c. Since this
definition equivalently expresses the distance traveled by the wing in units of chord lengths, it is
easily generalized for accelerating wings (Chen et al. 2010) to

1 t
t∗ = U(τ ) dτ. 8.
c 0
Aspect ratio: the ratio
In pure translation of a wing of rectangular planform with span b, in a fluid of density ρ and of wing span to chord
dynamic viscosity μ, the flow is characterized by only three dimensionless parameters: the angle length, AR = b/c
of incidence, α; the aspect ratio, AR = b/c; and the Reynolds number, Re = Uc/ν. It should
be noted that applications involving unsteady leading-edge separation are generally at low Mach

www.annualreviews.org • Leading-Edge Vortices 83


Review in Advance first posted on
August 15, 2018. (Changes may still
occur before final publication.)
FL51CH04_Eldredge ARI 4 August 2018 12:17

number, so the flow can be considered incompressible. The resulting forces on the wing are
nondimensionalized to form lift and drag coefficients, C L and C D , with the usual inertial scaling,
ρU 2 bc/2.
Rossby number: the
ratio of inertial force, In wing rotation, illustrated in subpanel iv of Figure 5a, the wing adopts a fixed angle of
ρUref
2 /c, to Coriolis incidence, α, and rotates in a propeller-like motion. The axis of rotation passes through a point
effects, ρUref , i.e., that is either at or inboard of the wing root; the distance from the axis to the root is called the root
Ro = Uref /(c ) cutout, rc . Like translation, wing rotation may start impulsively, although it is usually initiated
Reduced pitch rate: with a linear ramp, often smoothed with Equation 6, toward a constant angular velocity, .
the ratio of pitch The rotating wing problem is nondimensionalized by some combination of the collection of
velocity to
parameters , c, b, rc , α, and ν. However, in the interest of constructing a physically relevant
translational velocity,
K = α̇0 c/(2U) scale for convective and inertial effects, and perhaps facilitating a meaningful comparison with the
translating wing, a characteristic velocity Uref must be formed from among these. Because the local
Pivot axis position:
for a pitching wing, incident velocity varies linearly from the axis of rotation, every possibility for the characteristic
Access provided by University of Central Florida on 08/18/18. For personal use only.

the distance from the scale, Uref = R, requires a selection of some spanwise location R; the definitions for Reynolds
Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

leading edge to the number, Re = Rc/ν, and force coefficients [e.g., C L = 2L/(ρ2 R2 bc )] follow. Within the
pivot axis, scaled by literature, various choices of R have been made, including midspan; three-quarter span; radius of
chord length, Xp /c
gyration, rg ; and wing tip, rc + b, so that a comparison between any two rotating wing studies must
first reconcile their definitions of parameters.
In addition to Re and α, there are two ratios of length, b/c (i.e., the aspect ratio, AR) and rc /c,
that potentially influence the fluid dynamics of a rotating rectangular wing. These length ratios can
be varied independently, and it is useful to note that the translating wing represents the extreme
where rc /c → ∞. However, collectively, these two parameters have a physical role in assessing
the ratio of inertial force, ρUref
2
/c, to Coriolis effects, ρUref , measured by the Rossby number,
Ro = Uref /(c ). In fact, as the present review shows, the flow generated by a rotating wing is more
sensitive to this ratio than it is to the ratio of inertial to viscous effects (i.e., Reynolds number).
The definition of Ro requires the same arbitrary choice of reference location R, so that multiple
definitions of Rossby number have appeared in the literature, e.g., Ro = (rc + b)/c (Lentink &
Dickinson 2009) and Ro = rg /c (Wolfinger & Rockwell 2014). The first definition quantifies the
local inertia at the wing tip, while the second represents a spanwise-average influence. In cases
where rc = 0, the first definition is equivalent to the wing aspect ratio, AR, and the second differs
only by a constant factor. However, for wings with root cutout, the difference between these
definitions is more significant, and AR remains an independent parameter (Schlueter et al. 2014).
For either definition, a translating wing of finite aspect ratio represents the limit Ro → ∞ (with
the constraint that the product U = rc is kept finite in this limit). Thus, rotating wings with root
cutout fall on a spectrum of 1/Ro between translating and rotating, as depicted in Figure 5a.
Wing pitch, depicted in subpanel i of Figure 5a, is defined as a motion in which the wing
undergoes a steady increase in angle of incidence, α(t), from zero to some larger fixed value well
beyond its static stall angle. As in the other cases, most recent studies have used Equation 6 to
smooth the change in angle of incidence and avoid the infinite inertial forces due to impulsive
accelerations at the start and finish of this maneuver. For a wing of chord length c translating at
constant speed U and pitching at nominal rate α̇0 , the motion is characterized by three parameters:
the Reynolds number, generally defined in the same manner as for pure translation motion; the
reduced pitch rate, K = α̇0 c/(2U), which is generally explored in the range between 0.01 and 1;
and the pivot axis position, Xp /c, which typically lies between 0 (pivot about the leading edge) and
1 (pivot about the trailing edge).
The combination of rotation and pitch, illustrated in subpanel ii of Figure 5a, is less commonly
explored in the literature and is only briefly reviewed here. Furthermore, the flows generated by
other commonly studied kinematics, particularly those comprising oscillatory pitch (Lind & Jones

84 Eldredge
Review in Advance first posted on
· Jones

August 15, 2018. (Changes may still


occur before final publication.)
FL51CH04_Eldredge ARI 4 August 2018 12:17

2016, Mulleners & Raffel 2012, Ohmi et al. 1991), surge (Granlund et al. 2014, 2016), or transverse
(Calderon et al. 2013b, Panah et al. 2015, Visbal 2011) motions; or canonical (Baik et al. 2012,
Choi et al. 2015, Fenercioglu & Cetiner 2012, Ol et al. 2009, Rival & Tropea 2010) or bio-inspired
(Bomphrey et al. 2009, Hubel & Tropea 2010, Jardin et al. 2012, Krishna et al. 2018, Ozen &
Rockwell 2010, Phillips et al. 2015) combinations of these often contain important LEV dynamics
but are omitted from this review for brevity.

4. THE PHYSICS OF LEADING-EDGE VORTICES

4.1. Rectilinear Translation


Arguably the simplest of the canonical kinematics is rectilinear translation, and this is the case
that was used to present the basic features of the LEV in Section 2. The flow topology, force
Access provided by University of Central Florida on 08/18/18. For personal use only.

histories, and surface pressure distributions resulting from a 2D version of this motion are given
Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

in Figures 2 and 3, respectively. From the start of the motion, the force produced by a wing
rapidly accelerated in translation grows as the LEV forms and falls off when the LEV sheds. The
mechanism of shedding may vary, but the LEV typically sheds within the first few chord lengths
of travel, and lift augmentation ceases at this point.
The fluid dynamic forces on the wing are intimately linked to the strength and position of the
LEV. Logically, varying the wing velocity will vary the rate at which vorticity is generated, and
thus the rate at which the LEV gathers circulation, and so will affect the LEV strength, shedding,
and force history of the wing. Faster wing motions (i.e., greater accelerations) have been found to
generate more vorticity and a tighter, more coherent LEV; slower accelerations produce weaker
vortices and lower forces (Chen et al. 2010, Mancini et al. 2015). Once the LEV moves off of
the wing, however, it follows a very similar trajectory for both cases (Manar et al. 2016, Mancini
et al. 2015). Despite this, the effect of the startup transient persists in the force history for many
convective times; in experiments, the fully developed wake and constant force is not established
until t ∗ > 14 (Mancini et al. 2015).
In addition to effects from the LEV, force history is a direct function of acceleration due to
added mass, discussed further in Section 5 (Mancini et al. 2015, Pitt Ford & Babinsky 2013, Pullin
& Wang 2004). By removing the added mass component of the measured force, it can be shown
that the remaining circulatory force (i.e., that due to the LEV) is less sensitive to acceleration,
especially for t ∗ ≤ 0.5 (Chen et al. 2010, Mancini et al. 2015). At this early stage, the trailing-
edge vortex is still forming and all of the circulation in the flow is contained in the LEVs and
trailing-edge vortices (Pitt Ford & Babinsky 2013). Even for 1 ≤ t ∗ ≤ 1.5, after the vortices shed,
the circulation within them is nearly equal and opposite. Bound circulation on the wing remains
small; the total lift force can be attributed to a combination of added mass and the LEV rather
than attached flow (Pitt Ford & Babinsky 2014, Pullin & Wang 2004).
On finite wings, tip vortex development affects both the formation and shedding of the LEV.
The severity of 3D effects decreases at higher aspect ratios, but all finite wings are susceptible to
LEV bending (resulting in arch vortices), stall cells, and wake instabilities. Numerical simulations
and flow visualization experiments have demonstrated that the flow topology remains the same at
early times for AR = 1, 2, and 4. From the start of the impulsive wing motion, flow separates at
the leading edge, trailing edge, and wing tips, forming a closed-loop vortex system, as shown in
Figure 6 (Taira & Colonius 2009, Freymuth et al. 1987). Past the initial transient, however, the
wake behind the wing becomes more dependent on aspect ratio as the tip vortices gain strength.
On a wing of aspect ratio 1, the tip vortices grow to affect the entire wing span and the vortex sheet
emanating from the leading edge remains attached to the wing. At higher aspect ratios, however,

www.annualreviews.org • Leading-Edge Vortices 85


Review in Advance first posted on
August 15, 2018. (Changes may still
occur before final publication.)
FL51CH04_Eldredge ARI 4 August 2018 12:17

a b

AR = 1 AR = 2
t * = 0.5, 5 t* = 0.5, 5

Figure 6
Vortex structure on rectangular flat-plate wings with aspect ratios AR = 1 and 2 at angle of incidence α = 30◦ and Reynolds number
Re = 300. Isosurfaces of vorticity are shown in light grey; the vortex cores visualized with the Q criterion are highlighted in green.
Time is measured in convective units, t ∗ = tU/c, where U is the plate speed and c is the chord length. Figure adapted with permission
from Taira & Colonius (2009).
Access provided by University of Central Florida on 08/18/18. For personal use only.

the tip vortices are less dominant. At long times, the leading-edge shear layer separates from an
Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

AR = 2 wing and forms an arch-shaped vortex that is not visible on an AR = 1 wing (Kim &
Gharib 2010, Taira & Colonius 2009).
Despite interactions, the LEVs and tip vortices remain distinct structures on rectangular wing
planforms; the sharp corners of the planform suppress vortex stability by impeding spanwise flow
and vorticity convection between the structures (Kim & Gharib 2010, Taira & Colonius 2009).
LEV shedding can be delayed on translating wings with rounder planforms that allow for vorticity
convection into the tip vortices, but vorticity transport in these 2D kinematics is not sufficient to
allow the LEV to remain attached for long times (Mancini et al. 2015, Taira & Colonius 2009).

4.2. Rotational Motion


The LEV on a rotating wing begins to form in a manner very similar to that of the translating
wing. Here, however, the spanwise velocity gradient that results from rotational motion produces
a spatially varying rate of vorticity generation, and thus the LEV that forms is conical, increasing
in both strength and size with radial distance from the axis of rotation. These radial increases
of the LEV are clearly visible in both the computations of Garmann & Visbal (2014), shown in
Figure 7, and in the plot of experimentally measured LEV circulation shown in Figure 8. As the

L EA DI N G EDGE

Cp
ROOT

ψ = 45°
T IP

0
–1
–2
–3
–4
–5
ψ = 90° –6
–7

AR = 1 AR = 2 AR = 4

Figure 7
Surface pressure distribution (represented by pressure coefficient C p ) on the suction side of a rotating wing at angle of incidence
α = 30◦ and different aspect ratios (AR), with overlaid isosurfaces of relative total pressure highlighting the core of the leading-edge
vortex. Arrows mark the expansion of the vortex core due to burst. ψ denotes the instantaneous angle of rotation. Figure adapted with
permission from Garmann & Visbal (2014).

86 Eldredge
Review in Advance first posted on
· Jones

August 15, 2018. (Changes may still


occur before final publication.)
FL51CH04_Eldredge ARI 4 August 2018 12:17

1.6
2.5
a b
2.1 1.2

1.7 0.8
|Γ|U tipc CL
1.3 0.4
y/b
0.9 0.25
0.50 0
0.60
0.5
-0.4
30 60 90 120 150 180 210 240 270 300 330 360 0 90 180 270 360
ψ (°) ψ (°)
Access provided by University of Central Florida on 08/18/18. For personal use only.

Figure 8
Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

Leading-edge vortex (LEV) circulation at spanwise locations y/b = 0.25, 0.50, and 0.60, where b is the span (a), and integrated lift
coefficient (b), versus angle of rotation ψ, for a rotating aspect ratio AR = 2 flat-plate wing at Reynolds number Re = 2,500 and angle
of incidence α = 45◦ . As the wing accelerates through ψ = 30◦ (shaded region), both LEV circulation and lift increase rapidly. Figure
adapted with permission from Medina & Jones (2016).

wing accelerates through ψ = 30◦ (Figure 8b), both LEV circulation and lift increase rapidly.
Past this point, the wing continues to rotate at a constant velocity and the LEV continues to grow,
albeit more slowly. In the example of Figure 8, lift reaches a global maximum near ψ = 120◦ ,
which in this case corresponds to the point at which the outboard portion of the LEV has grown
to cover the entire wing chord (see Figure 9b). The LEV remains attached to the wing and
of a similar size and strength as the wing completes a full rotation, resulting in near-constant
lift.

k– 1 k k+ 1
dz 25
a Ω
c
12.5
Vorticity
(ωc/U tip)
0

–12.5
1
b
0.5
Spanwise
velocity
(vy /U tip)
0

–0.5

Figure 9
Discretization of the flow field on an aspect ratio AR = 2 rotating wing at midspan. (a) Schematic of the wing and planar views. (b) Dye
flow visualization of the leading-edge vortex at rotation angle ψ = 120◦ . (c) Chordwise planes of spanwise vorticity (top) and velocity
(bottom) at ψ = 115◦ , centered at k = b/2 and separated by a spanwise increment, dz = 0.012b. Figure adapted with permission from
Medina & Jones (2016).

www.annualreviews.org • Leading-Edge Vortices 87


Review in Advance first posted on
August 15, 2018. (Changes may still
occur before final publication.)
FL51CH04_Eldredge ARI 4 August 2018 12:17

Due to the fundamental difference in LEV structure, the sustained lift values achieved on a
rotating wing are significantly greater than those on a translating wing (Lentink & Dickinson
2009, Manar et al. 2016, Usherwood 2008, Usherwood & Ellington 2002). The strength of the
LEV burst: occurs
when spanwise flow LEV is highly (nearly linearly) dependent on angle of attack (Ozen & Rockwell 2012, Wojcik &
stagnates or reverses; Buchholz 2014a), but neither velocity profile nor Reynolds number has a significant effect on the
opposite-signed nondimensional LEV strength (DeVoria & Ringuette 2012, Ozen & Rockwell 2012, Wojcik &
vorticity is entrained Buchholz 2014a) or the forces on the wing (Usherwood 2008). Measurements of the circulation of
in the LEV, and the
the LEV on the rotating wing at a given spanwise location exceed the bound circulation predicted
size of the vortex
expands rapidly by thin airfoil theory at that local velocity (Wojcik & Buchholz 2014a) but agree well with the
measured forces ( Jones & Babinsky 2010), supporting claims that lift on the wing is primarily
a result of the LEV and not bound circulation (Hemati et al. 2014; Pitt Ford & Babinsky 2013,
2014). Further evidence for this is given in Figure 7, where pressure contours are plotted under the
LEV, revealing a clear low-pressure region under this structure with higher pressure elsewhere.
Access provided by University of Central Florida on 08/18/18. For personal use only.

The attached LEV is unique to the rotating wing case. Early hypotheses as to the source of
Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

LEV stability focused on the convection of vorticity via spanwise flow (Ellington et al. 1996) but
were confounded by reports that blocking this flow did not result in vortex shedding (Birch &
Dickinson 2001). Later work demonstrated that, while spanwise flow absent rotational motion can
allow the LEV to remain attached for short times, it is insufficient for the attached LEV observed
on rotating wings at long times (Beem et al. 2011, Jardin & David 2014). On a rotating wing,
spanwise flow from root to tip, v y > 0, is generated by virtue of the linear variation of incident
flow along the span. It is typically of the same order of magnitude as the wing tip speed (Carr et al.
2013, Garmann & Visbal 2014, Harbig et al. 2013, Medina & Jones 2016, Wolfinger & Rockwell
2014) and is enhanced on the underside of the vortex core by the Coriolis term, which induces
an outboard-oriented “force” on any fluid element moving relatively slower than the incident
flow (Garmann & Visbal 2014, Jardin 2017). Although some have described this spanwise flow
as a mechanism for vorticity transport, there is no explicit representation of it in Equation 5; its
effect is canceled by spanwise vortex stretching. The third term of that equation does, however,
represent tilting of the nonspanwise components of vorticity toward the spanwise direction. This
tilting includes the influence usually described as the Coriolis effect ( Jardin 2017, Lentink &
Dickinson 2009). In canonical rotating motion of a wing at angular velocity  and angle of attack
α, for which z =  cos α > 0, vorticity is generated in the fluid adjacent to the wing of strength
ω̂z ≈ 2z > 0 (or ωz ≈ 0 in the corotating frame). On the underside of the control volume, where
the spanwise flow is strongest and where n· ω̂ ≈ −ω̂z < 0, this flow tilts this component of vorticity
into a negative spanwise component, opposite that of the LEV. This process, which might be more
appropriately called Coriolis tilting, is illustrated schematically in Figure 4c. In addition to Coriolis
tilting, vorticity annihilation can be a significant contributor to the outflux of circulation (Medina &
Jones 2016, Wojcik & Buchholz 2014b). By careful accounting of the vorticity within the flow, as in
the control volume analysis of Section 2.1, researchers have shown that LEV stability—equivalent
to relatively little growth in LEV circulation and persistent attachment to the wing—must be
achieved through a balance of vorticity transport mechanisms including convection, stretching,
Coriolis tilting, and annihilation ( Jardin 2017, Wojcik & Buchholz 2014b). These mechanisms,
however, vary along the span in different manners, so vorticity equilibrium is only achievable along
some fraction of the wing. In fact, Jardin (2017) found that by artificially modifying the magnitude
of the Coriolis term and thus the Rossby number of the flow, the spanwise extent of this stability
can be altered.
At the point along the wingspan where the balance of vorticity becomes untenable, the LEV
bursts. Burst typically occurs near or beyond the midspan but moves inboard with increases in
Rossby number (Carr et al. 2013, Garmann & Visbal 2014, Harbig et al. 2013, Kolluru Venkata

88 Eldredge
Review in Advance first posted on
· Jones

August 15, 2018. (Changes may still


occur before final publication.)
FL51CH04_Eldredge ARI 4 August 2018 12:17

& Jones 2013). It is visible near the midspan in the bottom row of Figure 7 and in Figure 9b. The
bottom row of Figure 9c provides measurements of the spanwise flow near the burst point of the
example from Figure 8. Spanwise flow at the k − 1 plane is from the wing root to the wing tip but
reverses direction at k, and a region of tip-to-root flow is visible at k + 1. The top row in Figure 9c
gives the vorticity field at each of the planes, where the process of negative vorticity entrainment
and LEV expansion is visible. During this process, the vortex line representing the core of the
LEV tilts aftward, the outboard portion of the vortex moves off of the leading edge, and the LEV
lifts off of the wing in a manner not dissimilar from that of the translating wing; measurements
beyond midspan at relatively high Re do not detect an attached LEV ( Jones & Babinsky 2010).

4.3. Translational Pitching


The LEV development on a wing that pitches while in translation is distinct from the previous
Access provided by University of Central Florida on 08/18/18. For personal use only.

two motions and is governed primarily by the reduced pitch rate, K, the location of the axis of
Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

rotation, Xp /c, and the wing’s aspect ratio, AR. Before discussing the physics of the flow’s response
to the pitching maneuver, it is useful to take note of the parameters that, to leading order, are
essentially unimportant in the flow response. As observed with other maneuvers, the Reynolds
number does not have a significant role; studies conducted over the range from O(102 ) to O(105 )
share a common behavior (Ol et al. 2010). Furthermore, the response is not particularly affected by
the wing’s cross-sectional geometry. Most of the studies of the late 1980s were carried out with a
NACA 0015 section, generally pitched about its quarter chord, whereas many of the investigations
of the past decade focused on a flat plate of 2% thickness. Aside from differences in the angles at
which the milestone events occur, the studies in these respective cohorts reveal the same essential
physical processes and the same trends in the parameters.
LEV formation during the pitch-up maneuver, illuminated by experimental flow visualiza-
tion (Granlund et al. 2013, Shih et al. 1992, Stevens & Babinsky 2016, Walker & Chou 1987,
Yilmaz & Rockwell 2012), numerical simulation (Eldredge et al. 2009, Jantzen et al. 2014, Visbal
2017, Visbal & Shang 1989), and measurements of surface pressure (Acharya & Metwally 1992,
Strickland & Graham 1987, Visbal & Shang 1989) and shear stress (Schreck et al. 2002) follows a
characteristic sequence of events. At low angles, a separation point appears near the trailing edge,
and a recirculation region develops downstream of this point; as angle increases, the separation
point moves toward the leading edge. After this point reaches the leading edge, the flow departs
substantially from the quasi-steady behavior observed if the wing were held at any of these an-
gles. A closed separation bubble emerges near the leading edge, inside of which the thickening
boundary layer rolls up into the LEV. Subsequent behavior is similar to that observed in the other
maneuvers.
As pitch rate K increases, all of the same components of this process are observable, but vary in
two critical respects. First, and perhaps most obvious, is that the LEV is increasingly stronger due
to the faster flow speeds and larger pressure gradients brought about by more rapid rotation. The
second aspect, less obvious, is that many of the milestones—the formation of the recirculation
zone, the initiation of the LEV, and its eventual shedding—are increasingly delayed to larger
angles of incidence (Buchner et al. 2012, Granlund et al. 2013, Jumper et al. 1987, Schreck et al.
2002, Strickland & Graham 1987, Visbal & Shang 1989, Walker & Chou 1987).
These delays are also evident as the pitch axis is moved aftward (Visbal & Shang 1989, Yu &
Bernal 2016). It is not surprising that increases in K and Xp /c exhibit similar influences on the
delay. The wall-normal component of velocity at the leading edge, relative to that of the incident
 
flow, is U sin α − 2K Xp /c ; hence, increases in either parameter tend to reduce this component.
However, it is impossible to conflate their influences based solely on this argument, as increases

www.annualreviews.org • Leading-Edge Vortices 89


Review in Advance first posted on
August 15, 2018. (Changes may still
occur before final publication.)
FL51CH04_Eldredge ARI 4 August 2018 12:17

K= 1
16 a 15
b
CL
K= 1 10
14

5
12 2πα
0
10 K = 0.8
–5
CL K = 0.6
8 K = 0.2

K = 0.5 4

6 K = 0.4
K = 0.3 2 CL
α

Access provided by University of Central Florida on 08/18/18. For personal use only.

4
Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

K = 0.2
K = 0.1 0
2
Circulatory component
Added mass component
0 –2
0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 0 10 20 30 40
t* α (°)

Figure 10
(a) Lift coefficient, C L , from pitch-up maneuvers about the leading edge of a flat plate from 0◦ to 45◦ at various pitch rates and
Reynolds number 1,000. Vorticity contours (in all cases, between ±30U/c) are depicted on the left for reduced pitch rate K = 1, 0.6,
and 0.2 (from top to bottom) when the plate is instantaneously at 30◦ incidence, corresponding to the circles on the respective lift
curves. (b) Lift coefficient versus instantaneous angle of incidence for K = 1 (top) and K = 0.2 (bottom), decomposed into the added
mass (dotted dashed line) and circulatory (dashed line) components; the quasi-steady result, 2π α (dashed gray line), is shown for reference.
Figure adapted with permission from Eldredge & Wang (2010).

in Xp /c do not intensify the LEV strength as K does. Visbal & Shang (1989) found that the
peak suction exerted by the LEV, a reasonable measure of its strength, increases steadily with K
but decreases with increasing Xp /c. The initial formation of the LEV cannot be predicted by a
kinematic parameter such as K or Xp /c. However, the leading-edge suction of the wing, which
measures the instantaneous state of the flow at the wing’s nose, appears to reach a critical value
across a variety of motions (Ramesh et al. 2014). This will be discussed further in Section 5.3.
The force generated by a pitching wing has several notable features that distinguish this ma-
neuver from those at fixed angle. Because the wing is simultaneously pitching and translating
throughout the maneuver, the force exerted on the wing is a mix of the influence from the vortices
developing at the edges and from the inertial reaction of the fluid (i.e., added mass). Examples of
this force are shown in Figure 10a, which depicts the histories of the lift generated at various pitch
rates, obtained from high-fidelity simulation of a flat plate pitched upward about its leading edge
to 45◦ at Reynolds number 1,000 (Eldredge & Wang 2010). Both the lift and time are displayed in
their conventional scalings with the translational velocity. Figure 10b shows the lift at two pitch
rates, K = 0.2 and 1, now plotted versus the instantaneous angle of incidence of the plate; in these
panels, the lift is decomposed into the added mass and circulatory contributions.
Each pitch-up begins with a smooth acceleration from α = 0◦ over a fixed time interval that
starts just before t ∗ = 1 and finishes soon afterward; the short-lived peak centered at this instant
is due to the added mass from acceleration, as Figure 10b confirms. Since the final angle is
achieved earlier at higher pitch rate, the duration of the history is compressed as K increases.

90 Eldredge
Review in Advance first posted on
· Jones

August 15, 2018. (Changes may still


occur before final publication.)
FL51CH04_Eldredge ARI 4 August 2018 12:17

This also explains why the added mass peak appears relatively wider at larger K in the plots versus
angle of incidence. The most significant feature of the lift response to this increase in K is its
dramatic increase in peak magnitude, from approximately 3 at K = 0.1 to nearly 16 at K = 1.
Though some of this increase in magnitude is attributable to added mass (proportional to K , as
Equation 11 confirms below), this part is relatively weak, as Figure 10b shows. Thus, most of the
lift enhancement is due to the circulatory part of the force, overcoming the negative lift due to the
decelerating rotation at the end of the maneuver. At angles below static stall, this circulatory lift
exceeds its quasi-steady value at all pitch rates larger than approximately 0.02 (Yu & Bernal 2016).
The excess lift is due to the rates of circulation accumulation in the trailing-edge vortex and the
LEV, which both increase with K ; the increased strength of the LEV is apparent in the vorticity
depicted in Figure 10. Several researchers (Eldredge et al. 2009, Granlund et al. 2013, Strickland
& Graham 1987) have attempted to develop mathematical expressions for the lift coefficient that
account for pitch rate and pivot axis location, but these have been only moderately successful due
Access provided by University of Central Florida on 08/18/18. For personal use only.

to the flow memory encoded in the vortex dynamics. Recent models for the force, discussed in
Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

Section 5, have had more success by explicitly including these dynamics.

4.4. Comparisons and Combinations of the Canonical Motions


There are many similarities in LEV physics between the three canonical motions described above,
but there are also some key differences. Both of the inherently 2D kinematics (translation and pitch)
shed their LEV within a few chord-lengths of travel. They produce high lift at early times before
the LEV sheds, but lift drops off rapidly as the vortex moves off of the wing. The pitching wing
in particular produces extremely high lift due to a combination of the inertial reaction of the fluid
and the presence of the LEV. On a rotating wing, however, the spanwise flow that develops is
strong enough to contribute to LEV stability and so the fully developed flow is fundamentally
different—the LEV remains of a near-constant strength and in a near-constant position on the
wing.
Figure 11 gives a comparison of results from a number of research groups that have per-
formed experiments on rectangular flat-plate wings undergoing the four canonical motions given
in Figure 5. Note that in Figure 11, the x-axis is the convective time based on the constant
final velocity of the wing, U, so while acceleration (in translation, rotation, or pitch) is performed
over one chord-length of travel for all motions, the end of the unsteady portion of the maneuver
appears on the axis at t ∗ = 1 for the pitching cases, where U is constant, and at t ∗ = 2 for the
translating and rotating cases, where U varies with t. Note also that the lift coefficient given here
is the measured lift coefficient, which includes not only the direct effect of the LEV but also
any inertial effects that arise due to unsteady wing motion. Direct comparison of these results
highlights the similarities between the force curves of the translation and rotation cases and the
marked difference that is effected by adding the pitch-up motion in either case. Dynamic pitch
dramatically increases the buildup and magnitude of the lift transient during the unsteady part of
the maneuver; fixed-incidence translation and rotation result in much slower lift growth with less
overshoot. The primary difference between the lift histories of a pitching wing in rotational or
rectilinear motion is that the forces relax to a steady state much more quickly in rotation than in
translation.
In the results shown in Figure 11, the lift coefficient at long times approaches similar values
for both translational and rotational wing motions (except for the AR = 2 translating wing,
where the difference can be attributed to aspect ratio effects). However, several other studies
have reported that, due to the attached LEV, a rotating wing in fully developed flow produces
greater lift than does a translating one (Manar et al. 2016, Percin & van Oudheusden 2015). This

www.annualreviews.org • Leading-Edge Vortices 91


Review in Advance first posted on
August 15, 2018. (Changes may still
occur before final publication.)
FL51CH04_Eldredge ARI 4 August 2018 12:17

5.0

4.5 Translation, AR = 4 (UMD/AFRL)


Translation, AR = 2 (UMD)
4.0 Rotation, AR = 2 (TUD)
Rotation, AR = 2 (UMD)
3.5
Rotation + pitch, AR = 2 (TUD)
3.0
Rotation + pitch, AR = 2 (UMD)
Pitch, AR = 4 (UMD/AFRL)
C L 2.5

2.0

1.5
Access provided by University of Central Florida on 08/18/18. For personal use only.

1.0
Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

0.5

0
0 1 2 3 4 5 6 7 8 9 10
t*

Figure 11
Lift histories from the four canonical wing motions given in Figure 5: translation, rotation, pure pitch, and pitch and rotation. Time is
measured in convective units, t ∗ = tU/c, where U is the plate speed and c is the chord length. Figure adapted with permission from
Jones et al. (2016). Abbreviations: AFRL, Air Force Research Laboratory; AR, aspect ratio; TUD, Delft University of Technology;
UMD, University of Maryland.

apparent contradiction is attributable to the choice of lift normalization. Recall from Section 3
that the computation of C L on a rotating wing requires an arbitrary choice of reference velocity,
affecting the magnitude of the nondimensional lift force. The necessity of this arbitrary choice
highlights the need for flow models that capture the underlying physics of the LEV rather than
blindly producing force predictions.

5. LOW-ORDER MATHEMATICAL MODELING


There has been significant effort over the past few decades to distill the transient physics of leading-
edge flow separation into simple dynamical models. The objectives of this modeling effort are
multifold. First and foremost, one generally seeks a fundamental understanding of the process
of LEV evolution. Furthermore, by capturing the essential physics of the flow in a mathematical
model and, particularly, the response of the flow to various forms of external input (e.g., actuation,
gust encounter, or motion or deformation of the wing or control surface), we can better conceive
of ways to manage the flow to some desired end.
Modeling strategies for leading-edge separation can be categorized in different ways. One
Force model: only
captures the force (and categorization is based on whether the strategy leads to a force model or a flow model. A flow model
moment) response to should have far fewer degrees of freedom than, say, a fully resolved Navier–Stokes simulation,
motion since it is meant to be a distillation of the physics into some simplified form. Members of either
Flow model: captures category can be roughly grouped into two approaches: models that build from first principles,
the flow field response generally using vortex models from inviscid flow theory, and models that are reductions of the
as well as the force full equations of motion. An example of this latter form is the work of Brunton et al. (2013),
in which they linearized the Navier–Stokes equations about a stable equilibrium at low angle of

92 Eldredge
Review in Advance first posted on
· Jones

August 15, 2018. (Changes may still


occur before final publication.)
FL51CH04_Eldredge ARI 4 August 2018 12:17

incidence and then used the eigensystem realization algorithm to obtain a reduced-order model.
Vortex element
The model was then used to successfully predict the lift response to a modest 10◦ pitch-up/pitch- model: based on
down maneuver. Although such model-reduction approaches are a rich area of research, they are inviscid flow theory,
generally limited in their ability to capture the full range of behaviors of the LEV because they are predicts flow evolution
based on linearization about some state or on a decomposition into modes that are only optimal through advecting
point vortices or
for a particular motion. Vortex element models, in contrast, are naturally adaptive to the changing
vortex sheets
flow conditions and embody the essential nonlinearities of the flow through mutual interactions
Added mass:
between elements. The current review primarily focuses on these models and particularly their
coefficients that assess
most recent incarnations, in which they are informed by empirical data. the extra fluid inertia
that must be overcome
whenever the wing is
5.1. Computing the Force accelerated
A key to the success of unsteady aerodynamics models is to isolate the part of the force due to
Access provided by University of Central Florida on 08/18/18. For personal use only.

inertial reaction of the fluid, i.e., the added mass terms, so that the model is dedicated to computing
Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

the remaining circulatory force. Ultimately, a fluid exerts its influence on the wing through surface
traction: pressure and viscous stress. However, its influence can also be accounted for by the rate
of change of momentum—or, more specifically, the impulse—in the fluid. The great advantage
of the impulse perspective is that, whereas the surface pressure depends nonlinearly on the fluid
velocity field, the impulse depends on it linearly and can thus be decomposed into parts from
fluid vorticity, wing motion, and uniform flow. The vorticity, or circulatory, portion includes the
direct influence of the fluid vorticity, as well as an indirect influence from the wing due to the
no-penetration condition, that is, the impulse of the bound vorticity, which in a viscous setting is
the vorticity in the boundary layer. The parts of impulse resulting from the wing motion relative
to the fluid—the noncirculatory components—comprise the inertial reaction of the fluid and
give rise to the added mass. For example, for a 2D flat plate of length c and infinitesimal thickness,
the added mass component that gives rise to linear momentum due to translation normal to the
plate is ρπ c2 /4.
The focus in this section is on the application of an inviscid vortex model to flow about a flat
plate in which fluid vorticity is represented by discrete vortex elements, as depicted in Figure 12.
Suppose that the plate, instantaneously at angle α(t), is rotating at angular velocity α̇ about a
pivot point while translating at velocity (−U, ḣ) relative to the ground frame. The pivot point is a
distance a = c/2 − Xp forward of the plate centroid. The plate is surrounded by a set of Nv vortex
elements with positions xJ and strengths  J = J ez , for J = 1, . . . , Nv , each of which advects with
velocity uJ . Any point x in ground coordinates can also be expressed in plate coordinates, x̃ = (x̃, ỹ),
via x = XC + x̃τ + ỹn, where n = (sin α, cos α) is the plate’s normal vector, τ = (cos α, − sin α) is
its tangent vector, and XC is the position of the plate’s centroid. The velocity of this centroid, U,
has tangent and normal components

Uτ = −U cos α − ḣ sin α, 9.
Un = −U sin α + ḣ cos α − a α̇. 10.

It can be shown, e.g., via conformal mapping from a unit circle (Milne-Thomson 1996), that the
rate of change of impulse leads to an expression for force on the plate–vortex system given by

d 
Nv
π 2 π
f = −ρ c U̇n n − ρ c2 α̇Un τ + ρ  J × (xJ − XC ) + fv ; 11.
4 4 dt
J=1

the moment can be shown to have an analogous expression. The leading two terms comprise the
inertial reaction (i.e., the effect of added mass) and depend only on the plate kinematics, including

www.annualreviews.org • Leading-Edge Vortices 93


Review in Advance first posted on
August 15, 2018. (Changes may still
occur before final publication.)
FL51CH04_Eldredge ARI 4 August 2018 12:17

(−U, h)

~ ~
y, n
α̇ xJ
α
x
~ XC
a x, τ
Access provided by University of Central Florida on 08/18/18. For personal use only.

Figure 12
Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

Schematic of a flat plate in general rigid body motion about a pivot point. The ground coordinate system is
denoted by (x, y) and the plate coordinates by (x̃, ỹ). The plate translates with velocity (−U, ḣ) (relative to
the ground) and rotates with angular velocity α about a pivot point at a distance a forward of the centroid.
The plate’s centroid and angle of incidence are instantaneously XC and α, respectively. The J-th vortex
element in the surrounding fluid is located at x̃J relative to the plate centroid.

U̇n , the rate of change of the plate centroid’s normal velocity:

U̇n = −U̇ sin α + ḧ cos α − a α̈ − U α̇ cos α − ḣ α̇ sin α. 12.

All of the remaining terms are circulatory forces. The third term in Equation 11 constitutes the
direct effect of all fluid vortex elements on force via their motions and changes in circulation. In
steady flow at low angle of incidence, this term accounts for the classical Kutta–Joukowsky lift
through the advection of the starting vortex: The vortex moves away from the plate at the velocity
of the incident flow on the wing, −U, and its constant strength,  1 , is equal and opposite to the
bound circulation on the plate,  0 , so we have f = −ρU ×  0 .
The remaining circulatory contributions to force in Equation 11 are contained in the final term,
fv . It was noted above that the impulse of a vortex contains both its direct contribution as well
as the indirect contribution from its associated bound vorticity (most easily imagined as the effect
of the vortex’s potential flow image inside the plate); fv contains this indirect influence from
each vortex. This force is significant only when the centers of vorticity are near the edges of the
plate, where the bound vorticity has its greatest influence; it vanishes as these centers move away
from the edges. Rather than provide a complete derivation of this force here, the reader is referred
to Li & Wu (2016). For the purposes of developing intuition, it is reasonable to assume that
fv is zero, so that the circulatory force is well approximated by the third term in Equation 11.
It remains to describe the means of entry and subsequent evolution of vortex elements in the
fluid.

5.2. Vortex Modeling of the Separated Flow


For more than a century, inviscid flow theory has been our most effective language for describing
the aerodynamics of the attached flow past a wing, furnishing us with results that remain unsur-
passed in their brevity and reliability. Much of this success is attributable to the minimal incursion
of viscosity on the problem. In attached flows, the boundary layer remains thin and on the surface

94 Eldredge
Review in Advance first posted on
· Jones

August 15, 2018. (Changes may still


occur before final publication.)
FL51CH04_Eldredge ARI 4 August 2018 12:17

of the wing, and thus it can be suitably ignored in favor of a bound vortex sheet. In those parts
of the problem where viscosity’s role is physically essential—in setting the flux of vorticity into
the wake—this role can be robustly replaced by a constraint, the Kutta condition, that prohibits
infinitely large velocities at the trailing edge of the wing.
A flow that separates before reaching the trailing edge poses obvious challenges to this strictly
inviscid description. The separated flow, instead of passing immediately into the wake, now has
many opportunities to interact again with the boundary layer of the wing, and even the point
(or line) of separation may be in question and determined by subtle viscous mechanisms. But for
flows that separate near the leading edge, as are the focus of this review, the mechanics remain
mostly inviscid, and we propose that these flows can also be described with models that are drawn
from the potential flow toolbox: advecting vortex elements that induce velocity upon each other
and are subject to the no-penetration condition of the wing. However, it should be recognized
that, in contrast to the trailing edge, where the vorticity flux is set by the Kutta condition that the
Access provided by University of Central Florida on 08/18/18. For personal use only.

velocity remain finite, the condition at the leading edge is ambiguous. Indeed, the model might
Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

be more effective if it is left slightly underdetermined in this respect, so that it can be augmented
by empirical measurements; this is discussed in the next two sections.
In previous work, researchers seeking to model the lift on supersonic delta wing air-
craft used inviscid flow theory to predict the influence of the LEV (Brown & Michael 1954,
Edwards 1954, Mangler & Smith 1959, Polhamus 1966), extending classical thin airfoil theory
with some representation of the separated flow. The models differ only in the manner in which
the LEV is represented: an analytical model of a spiral vortex sheet (Mangler & Smith 1959)
or a single, variable-strength point vortex representing the rolled-up core (Brown & Michael
1954, Edwards 1954). Vortex models of separating flows in agile flight or rotorcraft, notably
those developed by Sarpkaya (1975) and Katz (1981), increasingly relied on emerging compu-
tational resources to provide greater fidelity than earlier analytical models. These 2D models
represent the separating flow and wake with discrete point vortices released in short time in-
crements from the respective edges of the wing. More recent models developed by Ansari et al.
(2006), Xia & Mohseni (2013), and Ramesh et al. (2014) have utilized similar approaches, al-
though with different leading-edge conditions, as discussed below. Vortex sheet models, such
as those by Jones (2003) and Shukla & Eldredge (2007), maintain connectivity between shed
vortex elements, enabling greater authority over the initiation of instabilities and their associ-
ated length scales. Unfortunately, both the vortex sheet model and the discrete vortex element
model become progressively more expensive as time proceeds and the number of elements grows.
Furthermore, as the LEV convects along a wing at modest angle of incidence, these models
exhibit nonphysical behaviors in lieu of the essentially viscous interactions with the boundary
layer.
To avoid these issues, Wang & Eldredge (2013) took a simple approach similar to that of Brown
& Michael (1954) and used a single, variable-strength vortex element at each edge to capture the
formation and convection of the edge’s associated flow structure. Variable-strength vortices re-
quire special care, however, because they necessarily produce a discontinuous pressure along a
curve between the vortex and the wing. Brown & Michael (1954) addressed this by modifying
the advection velocity of the vortex so that its altered force cancels the spurious force generated
by the pressure discontinuity. The velocity modification is proportional to the rate of transfer of
˙ from the wing to the vortex. Unfortunately, the Brown–Michael approach makes
circulation, ,
it impossible to avoid a clumsy jump in force whenever the strength of a vortex is frozen. To
avoid this nonphysical result, Wang & Eldredge (2013) calculated the modified velocity to en-
sure that the force on the wing is independent of ˙ through an approach they called “impulse
matching.”

www.annualreviews.org • Leading-Edge Vortices 95


Review in Advance first posted on
August 15, 2018. (Changes may still
occur before final publication.)
FL51CH04_Eldredge ARI 4 August 2018 12:17

As computational resources have become cheaper and more widely available, it has become
correspondingly faster to achieve high-fidelity simulations of dynamically separated flows. Vortex
models no longer have a clearly advantageous role as predictive tools. However, because of their
unique aptitude for capturing inherently nonlinear flow dynamics (vortex–vortex and vortex–body
interactions) even with a minimal number of elements, vortex models remain powerful for distilling
the full flow physics into a computationally tractable form.

5.3. Leading-Edge Conditions, Vorticity Flux, and the Suction Parameter


Classically, a mathematical model of wing aerodynamics based on 2D inviscid flow theory includes
three basic ingredients: a transport equation for the vortex elements in the fluid, such as those
discussed in the previous section; the Kutta condition, which ensures finite velocity and pressure
at the trailing edge; and Kelvin’s circulation theorem, a constraint that requires that any change
Access provided by University of Central Florida on 08/18/18. For personal use only.

of circulation in the wake is balanced by an equal and opposite change to the bound circulation.
Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

Traditionally, the Kutta condition is enforced—at every instant in an unsteady flow—by choosing
the bound circulation. Kelvin’s theorem, in turn, determines the strength of newly shed wake
vorticity. At low angles of incidence, the circumstances at the leading edge can be ignored, aside
from noting that the flow there will generally have large velocity and (suction) pressure. These
become infinite as the nose radius approaches zero, but the integrated pressure remains finite in
the limit and is called the leading-edge suction.
The three basic ingredients can be generalized to model a flow with expected leading-edge
separation. However, some condition must now be placed on the flow at the leading edge to
determine the strengths of vortex elements that enter the fluid at that point. A reasonable approach
is to express the rate at which circulation enters the LEV, which follows directly from the convective
flux of vorticity via the shear layer, as shown in Equation 5. This can alternatively be written in
terms of the speed of the flow, vLE , just outside the shear layer at the leading edge of the wing:
d 1 2
= − vLE . 13.
dt 2
Some researchers (e.g., Katz 1981, Sarpkaya 1975) have used this equation to determine the
strength δ of new vortex elements entering the flow near the leading edge in each time step.
However, it is considerably challenging to accurately calculate vLE in the vicinity of a discrete
set of singular vortex elements and a singular leading edge, and since the growth of the LEV
circulation depends strongly on this velocity, small errors can lead to significant miscalculations
of the growth.
For a suitable alternative, it is helpful to revisit the concept of edge suction. Enforcement of
the Kutta condition at an edge is equivalent to prescribing its suction force to be zero; thus, this
condition can be expressed as
σ (δ; S(t)) = 0, 14.
in which the strength δ of a nascent vortex element is determined by this condition on the scalar
function σ , which is called the edge’s suction parameter (Ramesh et al. 2014). In addition to δ, this
parameter depends on the current state, S(t), of the flow: the positions and strengths of all existing
vortex elements, as well as the current configuration and motion of the wing. The σ parameter has
two essential properties. First, it has units of velocity, and the edge’s suction force (per unit length)
is equal to ρσ 2 c/2. Second, the sign of σ describes the direction in which the flow is traveling
around the edge. In other words, σ is a measure of the average velocity circumnavigating the edge;
when the Kutta condition is enforced at that edge, this average is zero.

96 Eldredge
Review in Advance first posted on
· Jones

August 15, 2018. (Changes may still


occur before final publication.)
FL51CH04_Eldredge ARI 4 August 2018 12:17

Both the leading and trailing edges of an infinitesimally thin representation of a wing have an
associated edge suction parameter, and each provides an opportunity for generalizing the con-
straint imposed at that edge. It is usually sufficient to leave the suction parameter at the trailing
edge fixed at zero, as in classical aerodynamics. At the leading edge, however, it is reasonable
to relax this condition on its suction parameter, σ L . Ramesh et al. (2014), for example, postu-
lated that the flow around the nose of a wing can tolerate some amount of suction—and the
associated adverse pressure gradient—without separating. Since σ L is defined so that it is gener-
ally positive, they proposed that no vorticity be released as long as σ L is smaller than a critical
value, σmax
L
:
σ L (δ; S(t)) ≤ σmax
L
. 15.
If, instead, the instantaneous σ L exceeds the allowable range, then a new vortex element is released
from the edge, with a strength δ proportional to σ L − σmax L
to ensure that σ L is brought back
within bounds. Clearly, the Kutta condition is equivalent to setting σmax
Access provided by University of Central Florida on 08/18/18. For personal use only.

L
to zero, and Equation 15
Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

can be viewed as a generalization of this condition. The Kutta condition leads to a prediction of
circulation growth that is too strong except at very large angles of incidence (Wang & Eldredge
2013). Larger values of σmax
L
naturally lead to the release of weaker vortex elements from the leading
edge, and there is necessarily a maximum value of σmax L
that σ L never exceeds and that reverts the
model to shedding from the trailing edge only. It is important to stress that, for any nonzero value
of σmax
L
, the leading edge of a thin airfoil will have singular flow velocity and nonzero leading-edge
suction.
To determine the appropriate value of σmax L
, Ramesh et al. (2014) postulated that, for a given
airfoil section and Reynolds number regime, there is a unique value of σmax L
(when nondimen-
sionalized by a characteristic velocity)that holds across a wide variety of airfoil kinematics. They
obtained this value by carrying out both a high-fidelity simulation and a vortex model of some
representative motion, noting the value of σ L in the vortex model at the instant when the high-
fidelity simulation first exhibits the formation of a vortex at the leading edge. Vortex models of
other canonical motions, using this same σmax L
, agreed reasonably well with high-fidelity results.
Although there is evidence from the study of Ramesh et al. (2014) that σmax L
is insensitive
to the motion of the wing, there is also an advantage in assuming that σmax adapts to changing
L

conditions. For example, a time-varying value of σmax L


can serve as a trigger for the transient
aerodynamic response to an incident disturbance or actuation (Darakananda et al. 2018). In fact,
as is discussed further in Section 5.4, σmax
L
provides a natural opportunity for constructing a data-
driven vortex model that is responsive to empirical measurements: Rather than prescribing the
value of σmaxL
from offline calibration, its time-varying value might be obtained through data
assimilation.

5.4. Data-Driven Models


Because of the ambiguity in succinctly describing the physics of leading-edge flow separation
in a mathematical model, many investigators have proposed semiempirical approaches in which
measurements are used to inform some phenomenological model of the flow. The underlying
model, termed a dynamical template model in this review, captures much of the underlying physics
but leaves one or more parameters unspecified and to be determined from measurements. These
approaches represent a form of system identification, familiar in the field of control theory. Such
a procedure might be carried out as a postprocessing of data from experiments or high-fidelity
computations in order to obtain a simple closed-form description of the flow and force behavior,
or, by restricting the measurements to those obtained from onboard sensors and employing tools

www.annualreviews.org • Leading-Edge Vortices 97


Review in Advance first posted on
August 15, 2018. (Changes may still
occur before final publication.)
FL51CH04_Eldredge ARI 4 August 2018 12:17

from data assimilation, it might serve as a framework for real-time dynamic estimation of the
current flow state.
Inviscid vortex models that are informed from empirical data represent a useful and intuitive
approach to distilling the flow field. For example, Pitt Ford & Babinsky (2013) constructed a
basic potential flow template of a 2D flat plate accelerating from rest but left the positions and
strengths of a general set of point vortices unspecified. Each of these parameters was determined
by applying vortex identification tools to experimental flow data. This allowed them to deduce
aspects of the flow field unavailable from the direct measurements, such as the value of bound
circulation.
The approach taken by Pitt Ford & Babinsky (2013) omits the vortex dynamics (i.e., the motion
of point vortices with the local velocity field) from the template model, leaving the instantaneous
vortex position to be determined empirically along with its strength. Alternatively, the model
could include these dynamics and rely on measurements only to determine the strengths of vortex
Access provided by University of Central Florida on 08/18/18. For personal use only.

elements. The aforementioned study of Ramesh et al. (2014) represents a particularly light touch:
Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

Empirical measurements—from high-fidelity simulation—were used to determine a single static


parameter, σmaxL
, that helped determine the element strengths while the underlying dynamical
model was otherwise unmodified. Other approaches are more heavily reliant on data. For exam-
ple, Hemati et al. (2014) used the variable-strength vortex model of Wang & Eldredge (2013) as
a dynamical template but did not specify constraints at the edges of the wing. Instead, the rates of
change of the two vortex element strengths (i.e., the growth rates of the LEV and trailing-edge vor-
tex circulation) were found by minimizing a cost function—the error between the model-predicted
force and the “true” force obtained from high-fidelity computation over a time window—subject
to the constraint that the vortices obey their usual dynamics about the wing. This approach suc-
cessfully matched empirical force data for multiple cases of wings in pure translation and pitching,
except at low angles of incidence, where the drag was primarily due to skin friction and there-
fore impossible to model with vortex dynamics. The constrained minimization approach taken by
Hemati et al. (2014) requires a full empirical measurement history; an a posteriori phenomenolog-
ical model is iteratively constructed from these historical data. Darakananda et al. (2016) improved
the efficiency of the technique by building the data-optimized model progressively in short time
increments. The vortex growth rates in each increment are first predicted from a regression over
their recent history; these are then refined by minimizing as before, using the new force data in the
increment.
The investigations of Hemati et al. (2014) and Darakananda et al. (2016) demonstrated that
force measurements can provide meaningful information about the unknown rates of vorticity flux
into the LEV and the wake. However, one drawback was that the underlying model of two variable-
strength vortex elements was too simple to capture the mechanics of shedding. Furthermore, the
strategy they used is not of practical value if the objective of the model is to predict force, as is
typically the case. It is helpful to employ tools from the field of data assimilation, such as the
Kalman filter (KF), to develop a truly predictive strategy that is assisted by practical—and likely,
noisy—sensor measurements. The basic premise of the KF is to advance the estimate of a system’s
state in two steps within each time increment. The state of the flow is first predicted with a
dynamical model, and then this prediction is updated by incorporating new measurements. The
ensemble KF (EnKF) is a variant of the KF designed for dealing with high-dimensional nonlinear
systems, such as fluid flows (Evensen 2009). In the framework of an EnKF, Darakananda et al.
(2018) employed a discrete vortex element model with the leading-edge suction criterion for the
predictive step and used surface pressure measurements obtained from a high-fidelity Navier–
Stokes simulation (the “truth” system) for the update. The state vector comprised the positions
and strengths of the vortex elements and, crucially, σmax
L
. This framework was used to estimate the

98 Eldredge
Review in Advance first posted on
· Jones

August 15, 2018. (Changes may still


occur before final publication.)
FL51CH04_Eldredge ARI 4 August 2018 12:17

3
a t* = 3 t* = 5
b
2
Cn
1

0.8

σ Lmax
0.4

0
Access provided by University of Central Florida on 08/18/18. For personal use only.

0 1 2 3 4 5
t*
Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

Figure 13
(a, top row) Snapshots of vorticity fields from the “truth” case (computational fluid dynamics at Re = 500) and (bottom row) snapshots of
vortex elements from the ensemble Kalman filter (EnKF)-based model (circles) at the same times. The strengths of the vortex elements
are proportional to the size of the circles, and the colors indicate sign (red is negative). (b, top panel) Normal force coefficient, Cn , versus
time from the EnKF-based model (blue) and computational fluid dynamics “truth” data (red ). (Bottom panel) EnKF estimate of critical
L . Time is measured in convective units, t ∗ = tU/c, where U is the plate speed and c is the chord
leading-edge suction coefficient, σmax
length. Figure adapted with permission from Darakananda et al. (2018).

flow past a translating flat-plate wing at α = 20◦ and Re = 500. As the results in Figure 13 show,
the estimated vorticity field and normal force agreed well with the truth; the small noise in the
force estimate is due to the fact that the EnKF uses an ensemble of random instances of the model.
As Figure 13b shows, the estimate of σmax L
settled to a reasonably constant value, supporting the
hypothesis of Ramesh et al. (2014).

SUMMARY POINTS
1. The leading-edge vortex (LEV) is an essential flow feature of agile flight in nature and
small-scale vehicles. Its mechanics can be understood by distilling complex wing geome-
tries and kinematics into basic canonical forms.
2. The early development of the LEV and its structure are remarkably common across a
wide variety of wing kinematics and geometries, in both two and three dimensions, and
are relatively independent of Reynolds numbers from O(102 ) to O(104 ).
3. The lift generated by the LEV is significantly larger than achievable in steady, attached
flow, but this enhancement is short lived except on rotating wings.
4. The LEV inevitably sheds from the wing in nearly all kinematics and geometries, ex-
cept over limited spanwise ranges in rotating wings, where Coriolis tilting and vorticity
annihilation collectively balance the growth of LEV circulation.
5. Vortex models can capture the inherent nonlinearities of the LEV flow field with a
modest number of degrees of freedom. Their accuracy can be systematically improved
by data assimilation using sensor measurements.

www.annualreviews.org • Leading-Edge Vortices 99


Review in Advance first posted on
August 15, 2018. (Changes may still
occur before final publication.)
FL51CH04_Eldredge ARI 4 August 2018 12:17

FUTURE ISSUES
1. The canonical motions of rigid wings described in this review set only a foundation of
understanding; ultimately, these motions will be combined and applied to flexible wings to
form agile maneuvers. These higher-order combinations and the resulting fluid–structure
interactions will require systematic study, possibly with help from machine learning to
navigate the large parametric space toward optimal maneuvers.
2. LEV formation or response in incident gusts and other transient disturbances has only re-
cently received systematic attention. We have little understanding of how LEV behavior
will change in an uncertain environment.
3. Data-driven vortex modeling of the LEV has the potential to provide real-time estimation
of the flow state in a closed-loop control setting. Furthermore, 3D vortex models have
Access provided by University of Central Florida on 08/18/18. For personal use only.

been largely unexplored but have the potential to capture the essential influences of the
Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

tip vortices and spanwise flow on low–aspect ratio wings.


4. Closed-loop control of massively separated flows, effected with pulsed actuators or vari-
ations in wing geometry, remains a primary objective for achieving agile flight vehicles
that remain robust to incident gusts. The LEV is likely to play a central role in this
control.

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

ACKNOWLEDGMENTS
The authors gratefully acknowledge support from the Air Force Office of Scientific Research under
grants FA9550-11-1-0098 ( J.D.E.), FA9550-12-1-0251 (A.R.J.), FA9550-14-1-0328 ( J.D.E.), and
FA9550-16-1-0508 (A.R.J.); the US Army Research Laboratory under the Micro Autonomous
Systems and Technology (CTA-MAST) program (A.R.J.); and the National Science Foundation
under grant 1553970 (A.R.J.).

LITERATURE CITED
Acharya M, Metwally MH. 1992. Unsteady pressure field and vorticity production over a pitching airfoil.
AIAA J. 30:403–11
Akkala JM, Panah AE, Buchholz JHJ. 2015. Vortex dynamics and performance of flexible and rigid plunging
airfoils. J. Fluid Struct. 54:103–21
Ansari SA, Zbikowski R, Knowles K. 2006. Non-linear unsteady aerodynamic model for insect-like flapping
wings in the hover. Part 2: implementation and validation. Proc. Inst. Mech. Eng. G 220:169–86
Baik YS, Bernal LP, Granlund K, Ol MV. 2012. Unsteady force generation and vortex dynamics of pitching
and plunging aerofoils. J. Fluid Mech. 709:37–68
Beckwith RMH, Babinsky H. 2009. Impulsively started flat plate flow. J. Aircr. 46:2186–89
Beem HR, Rival DE, Triantafyllou MS. 2011. On the stabilization of leading-edge vortices with spanwise
flow. Exp. Fluids 52:511–17
Birch JM, Dickinson MH. 2001. Spanwise flow and the attachment of the leading-edge vortex on insect wings.
Nature 412:729–33

100 Eldredge
Review in Advance first posted on
· Jones

August 15, 2018. (Changes may still


occur before final publication.)
FL51CH04_Eldredge ARI 4 August 2018 12:17

Bomphrey R, Taylor G, Thomas A. 2009. Smoke visualization of free-flying bumblebees indicates independent
leading-edge vortices on each wing pair. Exp. Fluids 46:811–21
Brown CE, Michael WH. 1954. Effect of leading edge separation on the lift of a delta wing. J. Aeronaut. Sci.
21:690–94
Brunton SL, Rowley CW, Williams DR. 2013. Reduced-order unsteady aerodynamic models at low Reynolds
numbers. J. Fluid Mech. 724:203–33
Buchner AJ, Buchmann N, Kilany K, Atkinson C, Soria J. 2012. Stereoscopic and tomographic PIV of a
pitching plate. Exp. Fluids 52:299–314
Calderon DE, Wang Z, Gursul I. 2013a. Lift-enhancing vortex flows generated by plunging rectangular wings
with small amplitude. AIAA J. 51:2953–64
Calderon DE, Wang Z, Gursul I, Visbal M. 2013b. Volumetric measurements and simulations of the vortex
structures generated by low aspect ratio plunging wings. Phys. Fluids 25:067102
Carr Z, Chen C, Ringuette M. 2013. Finite-span rotating wings: three-dimensional vortex formation and
variations with aspect ratio. Exp. Fluids 54:1444
Access provided by University of Central Florida on 08/18/18. For personal use only.

Chen KK, Colonius T, Taira K. 2010. The leading-edge vortex and quasisteady vortex shedding on an
accelerating plate. Phys. Fluids 22:033601
Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

Choi J, Colonius T, Williams DR. 2015. Surging and plunging oscillations of an airfoil at low Reynolds
number. J. Fluid Mech. 763:237–53
Darakananda D, Eldredge JD, Colonius T, Williams DR. 2016. A vortex sheet/point vortex dynamical model for
unsteady separated flows. Paper presented at AIAA Aerospace Science Meeting, 54th, San Diego, CA, AIAA
Pap. 2016-2072
Darakananda D, Eldredge JD, da Silva A, Colonius T, Williams D. 2018. EnKF-based dynamic estimation of
separated flows with a low-order vortex model. Paper presented at AIAA Aerospace Science Meeting, 56th,
Kissimmee, FL, AIAA Pap. 2018-0811
DeVoria AC, Ringuette MJ. 2012. Vortex formation and saturation for low-aspect-ratio rotating flat-plate
fins. Exp. Fluids 52:441–62
Dickinson MH, Lehmann FO, Sane SP. 1999. Wing rotation and the aerodynamic basis of insect flight. Science
284:1954–60
Edwards RH. 1954. Leading edge separation from delta wings. J. Aerosp. Sci. 21:134–35
Eldredge JD, Wang C. 2010. High-fidelity simulations and low-order modeling of a rapidly pitching plate. Paper
presented at Fluid Dynamics Conference and Exhibit, 40th, Chicago, IL, AIAA Pap. 2010-4281
Eldredge JD, Wang C, Ol MV. 2009. A computational study of a canonical pitch-up, pitch-down wing maneuver.
Paper presented at AIAA Fluid Dynamics Conference, 39th, San Antonio, TX, AIAA Pap. 2009-3687
Ellington CP. 1984. The aerodynamics of hovering insect flight. VI. Lift and power requirements. Philos.
Trans. R. Soc. B 305:145–81
Ellington CP, van den Berg C, Willmott AP, Thomas ALR. 1996. Leading-edge vortices in insect flight.
Nature 384:626–30
Evensen G. 2009. Data Assimilation: The Ensemble Kalman Filter. Berlin: Springer-Verlag
Fenercioglu I, Cetiner O. 2012. Categorization of flow structures around a pitching and plunging airfoil.
J. Fluid Struct. 31:92–102
Freymuth P, Finaish F, Bank W. 1987. Further visualization of combined wing tip and starting vortex systems.
AIAA J. 25:1153–59
Garmann D, Visbal M. 2014. Dynamics of revolving wings for various aspect ratios. J. Fluid Mech. 748:932–56
Graftieaux L, Michard M, Grosjean N. 2001. Combining PIV, POD and vortex identification algorithms for
the study of unsteady turbulent swirling flows. Meas. Sci. Technol. 12:1422–29
Granlund K, Monnier B, Ol M, Williams DR. 2014. Airfoil longitudinal gust response in separated versus
attached flows. Phys. Fluids 26:027103
Granlund K, Ol MV, Bernal LP. 2013. Unsteady pitching flat plates. J. Fluid Mech. 733:R5
Granlund K, Ol MV, Jones AR. 2016. Streamwise oscillation of airfoils into reverse flow. AIAA J. 54:1628–36
Haller G. 2001. Lagrangian structures and the rate of strain in a partition of two-dimensional turbulence.
Phys. Fluids 13:3365–85
Harbig R, Sheridan J, Thompson M. 2013. Reynolds number and aspect ratio effects on the leading-edge
vortex for rotating insect wing planforms. J. Fluid Mech. 717:166–92

www.annualreviews.org • Leading-Edge Vortices 101


Review in Advance first posted on
August 15, 2018. (Changes may still
occur before final publication.)
FL51CH04_Eldredge ARI 4 August 2018 12:17

Hemati MS, Eldredge JD, Speyer JL. 2014. Improving vortex models via optimal control theory. J. Fluid
Struct. 49:91–111
Huang Y, Green MA. 2015. Detection and tracking of vortex phenomena using Lagrangian coherent struc-
tures. Exp. Fluids 56:147
Hubel TY, Tropea C. 2010. The importance of leading edge vortices under simplified flapping flight conditions
at the size scale of birds. J. Exp. Biol. 213:1930–39
Jantzen RT, Taira K, Granlund K, Ol MV. 2014. Vortex dynamics around pitching plates. Phys. Fluids
26:053606
Jardin T. 2017. Coriolis effect and the attachment of the leading edge vortex. J. Fluid Mech. 820:312–40
Jardin T, David L. 2014. Spanwise gradients in flow speed help stabilize leading-edge vortices on revolving
wings. Phys. Rev. E 90:013011
Jardin T, Farcy A, David L. 2012. Three-dimensional effects in hovering flapping flight. J. Fluid Mech.
702:102–25
Jeong J, Hussain F. 1995. On the identification of a vortex. J. Fluid Mech. 285:69–94
Jones AR, Babinsky H. 2010. Unsteady lift generation on rotating wings at low Reynolds numbers. J. Aircr.
Access provided by University of Central Florida on 08/18/18. For personal use only.

47:1013–21
Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

Jones AR, Manar FH, Phillips N, Nakata T, Bomphrey R, et al. 2016. Leading-edge vortex evolution and lift
production on rotating wings (invited). Paper presented at AIAA Aerospace Sciences Meeting, 54th, San
Diego, CA, AIAA Pap. 2016-0288
Jones MA. 2003. The separated flow of an inviscid fluid around a moving flat plate. J. Fluid Mech. 496:405–41
Jumper E, Schreck S, Dimmick R. 1987. Lift-curve characteristics for an airfoil pitching at constant rate.
J. Aircr. 24:680–87
Kaden H. 1931. Aufwicklung einer unstabilen Unstetigkeitsfläche. Ing.-Arch. 2:140–68
Katz J. 1981. A discrete vortex method for the nonsteady separated flow over an airfoil. J. Fluid Mech. 102:315–
28
Kim D, Gharib M. 2010. Experimental study of three-dimensional vortex structures in translating and rotating
plates. Exp. Fluids 49:329–39
Kolluru Venkata S, Jones AR. 2013. Leading-edge vortex structure over multiple revolutions of a rotating
wing. J. Aircr. 50:1312–16
Krishna S, Green MA, Mulleners K. 2018. Flowfield and force evolution for a symmetric hovering flat-plate
wing. AIAA J. 56(4): 1360–71
Lentink D, Dickinson MH. 2009. Rotational accelerations stabilize leading edge vortices on revolving fly
wings. J. Exp. Biol. 212:2705–19
Li J, Wu ZN. 2016. A vortex force study for a flat plate at high angle of attack. J. Fluid Mech. 801:222–49
Lind AH, Jones AR. 2016. Unsteady aerodynamics of reverse flow dynamic stall on an oscillating blade section.
Phys. Fluids 28:077102
Liska S, Colonius T. 2017. A fast immersed boundary method for external incompressible viscous flows using
lattice Green’s functions. J. Comput. Phys. 331:257–79
Lu Y, Shen GX. 2008. Three-dimensional flow structures and evolution of the leading-edge vortices on a
flapping wing. J. Exp. Biol. 211:1221–30
Lu Y, Shen GX, Lai GJ. 2006. Dual leading-edge vortices on flapping wings. J. Exp. Biol. 209:5005–16
Manar F, Mancini P, Mayo D, Jones AR. 2016. Comparison of rotating and translating wings: force production
and vortex characteristics. AIAA J. 54:519–30
Mancini P, Manar F, Granlund K, Ol MV, Jones AR. 2015. Unsteady aerodynamic characteristics of a trans-
lating rigid wing at low Reynolds number. Phys. Fluids 27:123102
Mangler KW, Smith JHB. 1959. A theory of the flow past a slender delta wing with leading edge separation.
Proc. R. Soc. A 251:200–17
Medina A, Jones AR. 2016. Leading-edge vortex burst on a low-aspect-ratio rotating flat plate. Phys. Rev.
Fluids 1:044501
Milne-Thomson LM. 1996. Theoretical Hydrodynamics. New York: Dover. 5th ed.
Mulleners K, Raffel M. 2012. The onset of dynamic stall revisited. Exp. Fluids 52:779–93
Ohmi K, Coutanceau M, Daube O, Loc TP. 1991. Further experiments on vortex formation around an
oscillating and translating airfoil at large incidences. J. Fluid Mech. 225:607–30

102 Eldredge
Review in Advance first posted on
· Jones

August 15, 2018. (Changes may still


occur before final publication.)
FL51CH04_Eldredge ARI 4 August 2018 12:17

Ohmi K, Coutanceau M, Loc TP, Dulieu A. 1990. Vortex formation around an oscillating and translating
airfoil at large incidences. J. Fluid Mech. 211:37–60
Ol MV, Altman A, Eldredge JD, Garmann DJ, Lian Y. 2010. Résumé of the AIAA FDTC Low Reynolds Number
Discussion Group’s canonical cases. Paper presented at AIAA Aerospace Sciences Meeting, 48th, Orlando,
FL, AIAA Pap. 2010-1085
Ol MV, Babinsky H. 2016. Unsteady flat plates: a cursory review of AVT-202 research. Paper presented at AIAA
Aerospace Sciences Meeting, 54th, San Diego, CA, AIAA Pap. 2016-0285
Ol MV, Bernal LP, Kang CK, Shyy W. 2009. Shallow and deep dynamic stall for flapping low Reynolds
number airfoils. Exp. Fluids 46:883–901
Ozen CA, Rockwell D. 2010. Vortical structures on a flapping wing. Exp. Fluids 50:23–34
Ozen CA, Rockwell D. 2012. Flow structure on a rotating plate. Exp. Fluids 52:207–23
Panah AE, Akkala JM, Buchholz JHJ. 2015. Vorticity transport and the leading-edge vortex of a plunging
airfoil. Exp. Fluids 56:160
Percin M, van Oudheusden BW. 2015. Three-dimensional flow structures and unsteady forces on pitching
Access provided by University of Central Florida on 08/18/18. For personal use only.

and surging revolving flat plates. Exp. Fluids 56:47


Phillips N, Knowles K, Bomphrey R. 2015. The effect of aspect ratio on the leading-edge vortex over an
Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

insect-like flapping wing. Bioinspir. Biomim. 10:056020


Pitt Ford CW, Babinsky H. 2013. Lift and the leading-edge vortex. J. Fluid Mech. 720:280–313
Pitt Ford CW, Babinsky H. 2014. Impulsively started flat plate circulation. AIAA J. 52:1800–2
Polhamus EC. 1966. A concept of the vortex lift of sharp-edge delta wings based on a leading-edge-suction analogy.
Tech. Note D-3767, NASA, Washington, DC
Pullin DI. 1978. The large-scale structure of unsteady self-similar rolled-up vortex sheets. J. Fluid Mech.
88:401–30
Pullin DI, Wang ZJ. 2004. Unsteady forces on an accelerating plate and application to hovering insect flight.
J. Fluid Mech. 509:1–21
Ramesh K, Gopalarathnam A, Granlund K, Ol MV, Edwards JR. 2014. Discrete-vortex method with novel
shedding criterion for unsteady aerofoil flows with intermittent leading-edge vortex shedding. J. Fluid
Mech. 751:500–48
Rival DE, Kriegseis J, Schaub P, Widmann A, Tropea C. 2014. Characteristic length scales for vortex detach-
ment on plunging profiles with varying leading-edge geometry. Exp. Fluids 55:1660
Rival DE, Tropea C. 2010. Characteristics of pitching and plunging airfoils under dynamic-stall conditions.
J. Aircr. 47:80–86
Sarpkaya T. 1975. An inviscid model of two-dimensional vortex shedding for transient and asymptotically
steady separated flow over an inclined plate. J. Fluid Mech. 68:109–28
Schlueter KL, Jones AR, Granlund K, Ol M. 2014. Effect of root cutout on force coefficients of rotating wings.
AIAA J. 52:1322–25
Schreck S, Faller W, Robinson M. 2002. Unsteady separation processes and leading edge vortex precursors:
pitch rate and Reynolds number influences. J. Aircr. 39:868–75
Shih C, Lourenco L, Van Dommelen L, Krothapalli A. 1992. Unsteady flow past an airfoil pitching at a
constant rate. AIAA J. 30:1153–61
Shukla RK, Eldredge JD. 2007. An inviscid model for vortex shedding from a deforming body. Theor. Comput.
Fluid Dyn. 21:343–68
Shyy W, Liu H. 2007. Flapping wings and aerodynamic lift: the role of leading-edge vortices. AIAA J.
45:2817–19
Spedding GR, Hedenström A. 2008. PIV-based investigations of animal flight. Exp. Fluids 46:749–63
Stevens PRRJ, Babinsky H. 2016. Experiments to investigate lift production mechanisms on pitching flat
plates. Exp. Fluids 58:7
Strickland J, Graham G. 1987. Force coefficients for a NACA-0015 airfoil undergoing constant pitchrate
motions. AIAA J. 25:622–24
Taira K, Colonius T. 2009. Three-dimensional flows around low-aspect-ratio flat-plate wings at low Reynolds
numbers. J. Fluid Mech. 623:187–207
Usherwood JR. 2008. The aerodynamic forces and pressure distribution of a revolving pigeon wing. Exp.
Fluids 46:991–1003

www.annualreviews.org • Leading-Edge Vortices 103


Review in Advance first posted on
August 15, 2018. (Changes may still
occur before final publication.)
FL51CH04_Eldredge ARI 4 August 2018 12:17

Usherwood JR, Ellington CP. 2002. The aerodynamics of revolving wings I. Model hawkmoth wings. J. Exp.
Biol. 205:1547–64
Visbal MR. 2011. Numerical investigation of deep dynamic stall of a plunging airfoil. AIAA J. 49:2152–70
Visbal MR. 2017. Unsteady flow structure and loading of a pitching low-aspect-ratio wing. Phys. Rev. Fluids
2:024703
Visbal MR, Shang JS. 1989. Investigation of the flow structure around a rapidly pitching airfoil. AIAA J.
27:1044–51
Wagner H. 1925. Über die Entstehung des dynamischen Auftriebes von Tragflügeln. Z. Angew. Math. Mech.
5:17–35
Walker JM, Chou DC. 1987. Forced unsteady vortex flows driven by pitching airfoils. Paper presented at AIAA
Fluid Dynamics, Plasma Dynamics and Lasers Conference, 19th, Honolulu, HI, AIAA Pap. 1987-1331
Wang C, Eldredge JD. 2013. Low-order phenomenological modeling of leading-edge vortex formation. Theor.
Comput. Fluid Dyn. 27:577–98
Widmann A, Tropea C. 2015. Parameters influencing vortex growth and detachment on unsteady aerodynamic
Access provided by University of Central Florida on 08/18/18. For personal use only.

profiles. J. Fluid Mech. 773:432–59


Annu. Rev. Fluid Mech. 2019.51. Downloaded from www.annualreviews.org

Willmott A, Ellington CP. 1997. The mechanics of flight in the hawkmoth Manduca sexta. II. Aerodynamic
consequences of kinematic and morphological variation. J. Exp. Biol. 200:2723–45
Willmott AP, Ellington CP, Thomas ALR. 1997. Flow visualization and unsteady aerodynamics in the flight
of the hawkmoth, Manduca sexta. Philos. Trans. R. Soc. B 352:303–16
Wojcik CJ, Buchholz JHJ. 2014a. Parameter variation and the leading-edge vortex of a rotating flat plate.
AIAA J. 52:348–57
Wojcik CJ, Buchholz JHJ. 2014b. Vorticity transport in the leading-edge vortex on a rotating blade. J. Fluid
Mech. 743:249–61
Wolfinger M, Rockwell D. 2014. Flow structure on a rotating wing: effect of radius of gyration. J. Fluid Mech.
755:83–110
Xia X, Mohseni K. 2013. Lift evaluation of a two-dimensional pitching flat plate. Phys. Fluids 25:091901
Xu L, Nitsche M. 2014. Scaling behavior in impulsively started viscous flow past a finite flat plate. J. Fluid
Mech. 756:689–715
Yilmaz T, Rockwell D. 2012. Flow structure on finite-span wings due to pitch-up motion. J. Fluid Mech.
691:518–45
Yu HT, Bernal LP. 2016. Effects of pivot location and reduced pitch rate on pitching rectangular flat plates.
AIAA J. 55:702–18

104 Eldredge
Review in Advance first posted on
· Jones

August 15, 2018. (Changes may still


occur before final publication.)

You might also like