You are on page 1of 16

Resonance and propulsion performance of a heaving flexible wing

Sébastien Michelin and Stefan G. Llewellyn Smith

Citation: Phys. Fluids 21, 071902 (2009); doi: 10.1063/1.3177356


View online: http://dx.doi.org/10.1063/1.3177356
View Table of Contents: http://pof.aip.org/resource/1/PHFLE6/v21/i7
Published by the American Institute of Physics.

Additional information on Phys. Fluids


Journal Homepage: http://pof.aip.org/
Journal Information: http://pof.aip.org/about/about_the_journal
Top downloads: http://pof.aip.org/features/most_downloaded
Information for Authors: http://pof.aip.org/authors

Downloaded 22 Apr 2013 to 129.174.21.5. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
PHYSICS OF FLUIDS 21, 071902 共2009兲

Resonance and propulsion performance of a heaving flexible wing


Sébastien Michelin1,2,a兲 and Stefan G. Llewellyn Smith1
1
Department of Mechanical and Aerospace Engineering, Jacobs School of Engineering,
University of California San Diego, 9500 Gilman Drive, La Jolla, California 92093-0411, USA
2
Ecole Nationale Supérieure des Mines de Paris, 60-62 Boulevard Saint Michel,
75272 Paris Cedex 06, France
共Received 7 April 2009; accepted 6 June 2009; published online 20 July 2009兲
The influence of the bending rigidity of a flexible heaving wing on its propulsive performance in a
two-dimensional imposed parallel flow is investigated in the inviscid limit. Potential flow theory is
used to describe the flow over the flapping wing. The vortical wake of the wing is accounted for by
the shedding of point vortices with unsteady intensity from the wing’s trailing edge. The
trailing-edge flapping amplitude is shown to be maximal for a discrete set of values of the rigidity,
at which a resonance occurs between the forcing frequency and a natural frequency of the system.
A quantitative comparison of the position of these resonances with linear stability analysis results is
presented. Such resonances induce maximum values of the mean developed thrust and power input.
The flapping efficiency is also shown to be greatly enhanced by flexibility. © 2009 American
Institute of Physics. 关DOI: 10.1063/1.3177356兴

I. INTRODUCTION trolled by the animal: The wing can experience large passive
deformations during the stroke period19 under the action of
Unlike terrestrial animals that can use solid friction and the outside flow and its internal bending rigidity, whose
a fixed support, insects and fishes must generate from the spanwise and chordwise distributions result from the vena-
surrounding fluid the lift and thrust forces necessary to their
tion pattern of the wing.20,21 The forcing is generally applied
motion in their environment.1–5 Beyond the fundamental in-
by the insect on its wing through a main axis whose rigidity
terest of understanding the mechanism of insect flight and
is significantly higher than the rest of the structure 共e.g., the
fish swimming, recent research on propulsion in fluids has
leading edge of the wing兲. An important challenge now is to
also been motivated by the development of microaviation
understand the impact of such passive deformation on the
vehicles and of more efficient propulsion techniques based
propulsive performance of a flapping structure. In particular,
on biomimetics.
one may be interested in potential reduction in energy usage
Insects use thin flapping wings to generate an unsteady
flow around them to produce these forces. The flow is char- induced by the flexibility, and examine if the values of the
acterized by a relatively large Reynolds number Re= UL / ␯ rigidity for these structures lie within an “optimal” range for
with U as the typical wing velocity, L its characteristic chord, which the flapping efficiency is the highest. The definition of
and ␯ as the kinematic viscosity of the surrounding fluid: optimality in this work in terms only of thrust production and
Typically, Re⬃ 100– 5000;5 in that range, the forces on the propulsive efficiency is purposely restrictive: From a biologi-
wings are dominated by the pressure contribution and vis- cal point of view many other factors must be taken into ac-
cous effects are concentrated in thin boundary layers near the count to define the optimal structure of an insect wing, in-
solid’s boundary. These boundary layers separate during the cluding but not limited to material resistance and
unsteady wing motion and roll up into strong coherent maneuverability.
vortices6 that carry momentum away from the insect, thereby The purpose of the present study is to investigate nu-
generating the propulsive forces. The flow around the insect merically the effect of flexibility on the propulsive character-
is highly unsteady, and an ongoing research challenge resides istics of a flapping appendage. In the following, this structure
in the ability to describe the forces on the flapping structure will be referred to as a wing, understanding that the model
without solving explicitly for the details of the flow.7–11 could also be applied to a fish fin if it is allowed to deform
An important physical insight into the generation of pro- passively under the effect of the flow and of its bending
pulsive forces by flapping or deforming solids has been pro- rigidity. Solving for the coupled motion of a flexible solid
vided by the study of active motion and deformation. In such and a fluid is computationally challenging and expensive,
experimental,12,13 theoretical,1,14 and numerical studies,15–18 primarily due to the coupling occurring on a moving bound-
the position of the solid is prescribed and the influence of the ary whose position is a priori unknown and must be solved
swimming stroke on the propulsive performance is analyzed. for as well. Popular techniques to overcome this difficulty
The recent development of experimental imaging tech- are the use of coupled fluid and solid solvers using fitted
niques has however shown that most insect wings are not grids22 and immersed boundary methods.23,24 The use of
purely rigid and that their deformation is not entirely con- low-order models for the flexible wing also simplifies the
computation while still retaining important physical results
a兲
Electronic mail: smichelin@ucsd.edu. on the reaction of the body to the fluid flow.25,26

1070-6631/2009/21共7兲/071902/15/$25.00 21, 071902-1 © 2009 American Institute of Physics

Downloaded 22 Apr 2013 to 129.174.21.5. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
071902-2 S. Michelin and S. G. Llewellyn Smith Phys. Fluids 21, 071902 共2009兲

A ζe (t) zn (t), Γn ness, actuated by the operator 共e.g., main body of the insect兲.
U∞ ζ(s, t)
ζ0 (t) θ(s, t) The operator applies a purely vertical motion to the sheet’s
leading edge, whose orientation is constrained to be strictly
zN (t), ΓN (t)
horizontal. The vertical position and orientation of the wing
at the leading edge are then
L
FIG. 1. Heaving flexible wing in a steady axial flow. The heaving motion of
h共t兲 = h0共1 − cos ␻t兲 and ␪0共t兲 = 0 共1兲
amplitude A is imposed at the leading edge and vortices are shed from the
so that A = 2h0 and f = ␻ / 2␲ are the amplitude and frequency
trailing edge.
of the flapping motion, respectively 共to avoid confusion, ␻ or
its nondimensional form will be referred to as the angular
The model used here focuses on a simplified two- frequency in the following兲. The flapping wing has a chord-
dimensional propulsion problem using a flexible wing of in- wise flexural rigidity per unit length B and a mass per unit
finite span, actuated at its leading edge in a purely heaving area ␳s. Its thickness is negligible compared to L. The wing
motion and reacting passively to the flow forces and its in- is placed in a uniform horizontal flow U⬁ of density ␳. The
ternal elasticity. The present model does not aim to represent motion of the wing’s leading edge is entirely prescribed by
a particular flying or swimming pattern, but rather considers Eq. 共1兲 but the rest of the wing has a purely passive motion
a one degree of freedom forcing to focus on the influence of in response to its internal elasticity, the leading-edge forcing,
flexibility on the performance of the apparatus. In the limit of and the pressure forces applied by the surrounding flow.
high Re, viscous forces are neglected and the viscosity’s in- In the following, L, U⬁ and ␳ are used as reference quan-
fluence on the flow is retained in this potential flow formu- tities to nondimensionalize the problem. The properties of
lation by the irreversible shedding of vorticity from the trail- the elastic sheet are characterized by the mass ratio ␮ and
ing edge of the flapping structure, in the form of point nondimensional rigidity ␩ defined as
vortices whose unsteady intensity is determined so as to sat- ␳s B
isfy the regularity condition at the solid’s trailing edge.27,28 ␮= , ␩= , 共2兲
␳L ␳U⬁2 L3
A similar approach was recently proposed by Alben29 for
a pitching elastic sheet using a vortex sheet representation of and the leading-edge forcing is characterized by the nondi-
the wake. Optimal values of the solid’s rigidity were dis- mensional forcing amplitude ␧ and frequency f̄,
cussed in the limit of negligible solid inertia 共that is particu-
larly relevant in the case of fish swimming兲 and of infinitesi- h0 A fL
␧= = , f̄ = · 共3兲
mally small displacements of the solid. The present work L 2L U⬁
builds on these results and considers the general case of non-
The forcing Strouhal number is defined in accordance with
linear deformations of the sheet with non-negligible inertia
previous experimental studies12 as
共as is the case for a flapping insect wing兲. The use of the
unsteady point vortex model rather than the full vortex sheet fA
description also allows for a simpler treatment. Similar reso- St = = 2␧f̄ . 共4兲
U⬁
nance patterns are observed and a theoretical argument is
provided for their origin and position. The relation between The motion of the wing is described using an inextensible
thrust or drag production and vortex wake structure is also Euler–Bernoulli beam representation.28 We are interested in
investigated. large displacements of the wing so all nonlinear geometric
In Sec. II, the fluid-solid model is presented and the terms must be included. The linear Euler–Bernoulli assump-
propulsive performance quantities of interest are defined. tion remains valid if the curvature radius of the wing is much
Section III discusses briefly the numerical methods used as larger than its thickness, which is assumed here. The position
well as the existence of a periodic regime. Section IV then of the wing is described using complex notation as ␨共s , t兲
investigates the influence of the solid’s rigidity on the pro- = x共s , t兲 + iy共s , t兲 and its orientation is defined as ␪共s , t兲 with
pulsion forces and efficiency and relates them to the structure 0 ⱕ s ⱕ 1 the curvilinear coordinate along the wing. The clas-
of the solid’s wake. Peaks of thrust are observed for particu- sical notation for the complex flow velocity is also used here:
lar values of the rigidity and in Sec. V, the occurrence of w = u − iv with 共u , v兲 the Cartesian components of the veloc-
such peaks is showed to correspond to a resonance between ity vector. The conservation of momentum for each element
the forcing frequency and the natural frequencies of the of the wing and the inextensibility condition can be written
fluid-solid system. Finally, Sec. VI presents some general as
conclusions and discusses the limitations of the model.
␮␨¨ = 关共T − i␩␪ss兲ei␪兴s − i关p兴⫾ei␪, ␨ s = e i␪ , 共5兲

II. DESCRIPTION OF THE MODEL where 关p兴 is the pressure difference between the top and
A. Solid model bottom sides of the wing and T is the wing chordwise tension
that must be solved for at each time step to enforce the in-
The following two-dimensional model for the flapping extensibility condition. The clamped-free boundary condi-
structure is considered 共see Fig. 1兲. The wing is represented tions imposed by forcing 关Eq. 共1兲兴 are
by an elastic sheet of chord L and infinite span, clamped at
its leading edge on an attachment pole of negligible thick- ␨共0,t兲 = ␨0共t兲 = i␧关1 − cos共2␲ f̄t兲兴, ␪共0,t兲 = 0, 共6兲

Downloaded 22 Apr 2013 to 129.174.21.5. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
071902-3 Resonance and propulsion performance Phys. Fluids 21, 071902 共2009兲

␪s共1,t兲 = ␪ss共1,t兲 = T共1,t兲 = 0. 共7兲 In the absence of viscosity, the tangential velocity of the
flow can be discontinuous across the wing. The potential
Equations 共5兲–共7兲 can be rewritten as a system for ␪ and T flow around the wing is then computed by representing the
only,28 infinitely thin solid as a bound-vorticity distribution ␬.18,34,35
The complex flow velocity is obtained by superposition of
Tss − ␪s2T = − 关p兴⫾␪s − 2␩␪s␪sss − ␩␪ss
2
− ␮␪˙ 2 , 共8兲 the flow at infinity and the contribution of the bound and
wake vorticity
␮␪¨ = − 关p兴s⫾ − ␩␪ssss + 共T + ␩␪s2兲␪ss + 2Ts␪s ,

␪共0,t兲 = ␪s共1,t兲 = ␪ss共1,t兲 = T共1,t兲 = 0,


共9兲

共10兲
w共z,t兲 = 1 +
1
2␲i
冋冕 0
1
␬ds
+兺
⌫n
N

z − ␨共s兲 n=1 z − zn
. 册 共13兲

The regularity condition at the trailing edge imposes

␮␨¨ 0 + ␮ 冕冕 1

0 0
s
ei␪共i␪¨ − ␪˙ 2兲ds⬘ds
w共␨e , t兲 ⫽ ⬁ or equivalently ␬共1 , t兲 = 0. The total circulation at
infinity is conserved and, assuming the system is started
from rest, must be zero at all times. The bound vorticity ␬ is

= − T共0兲 + i␩␪ss共0兲 − i 冕 0
1
关p兴⫾ei␪ds. 共11兲
the solution of a singular Fredholm equation obtained by
applying the continuity of normal velocity on the wing.18
The corresponding system of equations for ␬ and ⌫N is

B. Representation of the flow


around the flapping wing
1
2␲
冕 冋
1
Re
ei␪共s兲
␨共s0兲 − ␨共s兲
␬共s兲ds册
冋冉 冊册
0
N
The flow around the wing is taken potential. The ex- 1 ⌫n
pected boundary layer separation at the trailing edge and its = Im e i␪
1+ 兺
2␲i n=1 ␨ − zn
˙
− ¯␨ , 共14兲
subsequent roll-up into vortices is represented in this inviscid


formulation by a discrete shedding of point vortices from the 1 N
trailing edge ␨e = ␨共1 , t兲 of the flapping wing 共Fig. 1兲. The ␬共s兲ds + 兺 ⌫n = 0, 共15兲
intensity of the last shed vortex is determined so as to cancel 0 n=1
exactly the square-root singularity arising in the velocity
field due to the presence of the flat sharp corner.27,28,30,31 ␬共1,t兲 = 0. 共16兲
When the intensity of the unsteady vortex reaches a maxi-
mum, a new vortex is created from the generating corner, From ␬, the pressure jump 关p兴⫾ across the wing can be com-
thereby expressing the irreversible nature of vortex roll-up in puted by integration of Bernoulli’s theorem along the wing
the formalism of this inviscid model. From that point on, the
intensity of the previous vortex is frozen. The unsteady point
vortices satisfy the modified equation of motion
关p兴⫾共s0兲 = 冕 0
s0
␬˙ 共s兲ds + ␬共s0兲w p共s0兲, 共17兲

with w p as the principal value of the relative tangential ve-


⌫˙ n locity on the wing28

冋 再
żn + 共zn − ␨e兲 = w̃n , 共12兲
⌫n

where zn and ⌫n refer to the position and intensity of the


w p共s0兲 = Re ei␪共s0兲 1 +
1
2␲i
冉冕 1
␬共s兲ds
␨共s0兲 − ␨共s兲

冎册
0


point vortex, respectively. w̃n is the desingularized complex
N
velocity at the vortex position and the overbar denotes a ⌫j
+兺
˙
complex conjugate. Equation 共12兲 is known as the Brown– − ¯␨共s0兲 . 共18兲
j=1 ␨共s0兲 − z j
Michael equation,32 and enforces the conservation of fluid
momentum around the vortex and associated branch cut in an
integral sense.27 The omission of the corrective term on the C. Energy conservation
left-hand side would lead to an unphysical unbalanced force From Eq. 共5兲, the conservation of energy can be written
on the branch cut linking the vortex to its generating corner for the wing
␨e.33
The shedding of vorticity from the leading edge is ne- d
共Ek + E p兲 = W p + Pin , 共19兲
glected here, as we focus mostly on situations where the dt
angle of attack remains small at the leading edge.29 Alterna-
where
tively, this representation can also be seen as the limit case of
a smoothed leading edge of very small curvature radius 共as
in an airfoil profile for example兲. Only one unsteady vortex
1
Ek = ␮
2
冕 0
1
兩␨˙ 兩2ds, 共20兲
is shed at a time 共from the trailing edge兲. Noting N as the


number of vortices at a particular time, all vortex intensities 1
1
⌫n are therefore independent of time except for the last one Ep = ␩ ␪s2ds, 共21兲
⌫N共t兲. 2 0

Downloaded 22 Apr 2013 to 129.174.21.5. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
071902-4 S. Michelin and S. G. Llewellyn Smith Phys. Fluids 21, 071902 共2009兲

Wp = − 冕 0
1
关p兴⫾Im共␨˙ e−i␪兲ds, 共22兲 具g典 =
1

冕0

g共t兲dt, 共29兲

Pin = − Re兵关␨˙ 共0兲e−i␪共0兲兴关T共0兲 + i␩␪ss共0兲兴其 − ␩␪˙ 共0兲␪s共0兲, 共23兲 with ␶ = 1 / f̄ as the nondimensional period of the flapping
motion and P+ as the positive part of P. In the following, the
are the kinetic and elastic potential energy of the wing, the mean power input is understood as 具P+典, thereby assuming
rate of work of the pressure forces on the wing, and the rate that the animal cannot store and reuse the energy possibly
of work of the force and torque applied by the attachment extracted from the fluid if P ⬍ 0 during a fraction of the
pole on the rest of the wing, respectively. In the particular flapping period. For the range of parameter values used here,
case of a purely heaving motion considered here, ␪共0兲 = 0 and it was observed that P ⬎ 0 for most of the flapping period
␨˙ 共0兲 = iḣ is purely imaginary, so and 具P+典 ⬃ 具P典, except for very rigid wings. The results and
discussions presented here are therefore not affected by this
Pin = ␩␪ss共0兲ḣ. 共24兲 choice.
Finally, the following 共nondimensionalized兲 quantities
D. Propulsive performance are defined for convenience:

We are interested in the thrust generated by the flapping — the trailing-edge peak-to-peak flapping amplitude D;
wing. The forces applied on the leading-edge attachment are — the intensity of the wake ⌫m, defined as the mean value
as follows: of the amplitude of the successive vortices 共positive and
negative兲;
— the elastic forces applied by the sheet on its attachment — the induced velocity of the wake vortices V defined as the
关T共0兲 − i␩␪ss共0兲兴ei␪共0兲; horizontal velocity of the wake vortices relative to the
— the force applied by the operator 共or animal兲 to prescribe imposed unit flow. V is positive if the wake vortices
x y
the leading-edge motion Fop = Fop + iFop; move faster than the background flow, and negative
— the suction force created at the leading edge by the in- otherwise.
verse square-root behavior of the pressure. This suction
force is the limit of the suction force obtained on a
smoothed contour when the curvature radius of the air-
foil’s leading edge tends to zero. This suction force is
III. NUMERICAL SIMULATION OF THE INITIAL VALUE
equal to29,36,37 PROBLEM AND CONVERGENCE TO A PERIODIC
␲ei␪共0兲
冉 冊
STEADY STATE
lim关冑s共1 − s兲␬共s兲兴 .
2
Fs = − 共25兲
4 s→0 Equations 共8兲–共12兲 and 共14兲–共17兲 are solved numerically
expanding ␪共s , t兲 into a finite series of Chebyshev polynomi-
Neglecting the inertia of the attachment and defining the
als of the first kind and using a semi-implicit second-order
thrust 共counted positively to the left兲 as T = Fop
x
, the force
time-stepping scheme. Taking advantage of the linear rela-
balance along the horizontal direction together with Eqs. 共6兲
tion between 关p兴⫾ and ␬˙ , added inertia terms can be isolated
and 共7兲 leads to
from the part of the pressure that can be explicitly computed

冉 冊
at each time step, thereby avoiding the use of an iterative
lim关冑s共1 − s兲␬共s兲兴 − T共0兲.
2
T= 共26兲 solver and greatly enhancing the computational efficiency.28
4 s→0
The system is started from rest. At t = 0, the horizontal
The instantaneous power input P by the operator is flow is ramped up to its long time unit value and the motion
of the leading edge 关Eq. 共1兲兴 is imposed. After a transient
˙
P = Re共Fop¯␨0兲 = Fop
y
ḣ共t兲 = ␩␪ss共0,t兲ḣ共t兲 = Pin , 共27兲 regime of a few heaving periods, a permanent periodic re-
gime is achieved for large enough values of the rigidity ␩
and is equal to the rate of work Pin of the attachment pole on
关Fig. 2共a兲兴.
the wing. Note that this equality would not hold if the motion
However, the harmonic heaving forcing can lead to
of the leading edge was a combination of both heaving and
highly unsteady behaviors if ␩ becomes too small: Below a
pitching, as the suction force Fs would have a nonzero rate of
certain critical value of the solid’s rigidity ␩m共␮兲, the purely
work along the vertical direction.
passive elastic sheet 共as in the flag problem兲 becomes un-
The useful power output is simply the rate of work TU⬁
stable to fluttering modes and flapping can occur even in the
of the thrust force in the horizontal motion. In nondimen-
absence of leading-edge forcing28,35,38 关␩m共␮兲 was found
sional units, the flapping efficiency is then defined as the
equal to ␩m = 2 ⫻ 10−3 for ␮ = 0.2 and ␩m = 4.8⫻ 10−2 for
ratio of the average developed thrust to the average input
␮ = 2, using the same point vortex model28兴. In such cases,
power
the spectrum of the trailing-edge motion can display several
具T典 peaks, corresponding to the forcing frequency and to the fre-
r= , 共28兲 quency of the unstable modes 关Fig. 2共b兲兴. The motion is of
具P+典
large enough amplitude for the regime to be nonlinear and
where 具 · 典 is the averaging operator over a flapping period mode coupling is also expected. As ␩ is reduced further, the

Downloaded 22 Apr 2013 to 129.174.21.5. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
071902-5 Resonance and propulsion performance Phys. Fluids 21, 071902 共2009兲

!)!$

+,-. 012*) 345-*-46 734812 5319*2):

+,-. 012*) 345-*-46 734812 5319*2):


!)!' !)!#&

!)!#
!)!%
!)!"&

!)!"
!)!#
!)!!&

! !
! & "! "& ! & "! "&
 
!)# !)"&

!)"
!)"
+,-. 012*) 345-*-46

+,-. 012*) 345-*-46


!)!&
!
!

!)"
!)!&

!)# !)"
! "! #! $! %! &! '! (! ! "! #! $! %! &! '! (!
* *
(a) µ = 2 and η = 0.2 (b) µ = 2 and η = 0.04

FIG. 2. 共Top兲 Frequency spectrum and 共bottom兲 time evolution of the trailing-edge absolute vertical displacement for ␮ = 2, ␧ = 0.05, and f̄ = 5 / 2␲. 共Left兲
␩ = 0.2⬎ ␩m共␮ = 2兲 lies in the stability region for the purely passive elastic sheet. The power spectrum displays only one peak 共thick arrow兲 at an angular
frequency of ␻ f = 2␲ f̄ = 5. 共Right兲 The rigidity ␩ is below its critical value ␩m共␮ = 2兲 = 0.048 and the elastic sheet is unstable to fluttering. The power spectrum
displays two main peaks: one corresponding to the forcing frequency 共␻ f = 5, thick arrow兲 and one corresponding to the unstable fluttering mode 共␻r ⬃ 2.3, thin
arrow兲 that matches the flapping frequency observed in the purely passive case ␧ = 0. In both cases, small peaks can be seen for ␻ ⬃ 3␻ f = 15, corresponding
to the third harmonic of the forced flapping. The fact that only odd harmonics appear in the tail motion was already previously observed in the case of a
passive flag 共Ref. 28兲.

periodicity is lost and the power spectrum is full; in such a flexibility is introduced in the problem; however, the former
case, determining averaged quantities is not possible increases faster and the resulting efficiency is an increasing
anymore. function of flexibility 共decreasing function of ␩兲 for large
This explains the difficulty to observe a steady perma- values of ␩ 共Fig. 3兲. When the solid’s flexibility is further
nent regime when ␩ is decreased below the critical value ␩m. reduced, 具T典 and 具P+典 display successive peaks occurring for
This difficulty was not present in the linear study by Alben29 the same values of ␩. The propulsive efficiency r displays, in
as the inertia of the solid was neglected 共␮ → 0兲. The inertia general, one large peak 共in general, for a different value of
of the solid is essential to the development of fluttering in- the rigidity ␩ than the thrust peaks兲 before dropping sharply
stability and in the limit ␮ → 0, all the modes are linearly to zero as the mean thrust vanishes 共drag-thrust transition兲.
stable.28,35,38 In the following, unless indicated otherwise, the The increase in the flapping efficiency and developed mean
range of parameter values is chosen such that a periodic state thrust for a flexible wing is significant compared to the case
is achieved.
of a rigid wing: For f̄ = 5 / 2␲, the peak value of the mean
thrust can be greater than twice its value in the rigid case and
IV. WING FLEXIBILITY AND PROPULSIVE the efficiency can increase from 27% to almost 60% 关Fig.
PERFORMANCE
3共b兲兴. Similar behavior is observed for f̄ = 1 / ␲ 关Fig. 3共a兲兴.
In this section, the behavior of the propulsive perfor- Flexibility has therefore a significant impact on the perfor-
mance 共mean thrust, mean power input, and efficiency兲 is mance of the propulsive apparatus considered here.
studied when the rigidity ␩ of the wing is varied. Several Figure 3 shows the evolution of the propulsive properties
values of the forcing frequency f̄, forcing amplitude ␧, and for the lowest possible value of ␩ leading to a permanent
mass ratio ␮ were investigated. periodic regime. In the higher frequency case 共f̄ = 5 / 2␲兲, the
drag-thrust transition was clearly observed while at lower
A. Optimal flexibility for thrust generation frequency 共f̄ = 1 / ␲兲, this transition is less well defined as
and propulsion efficiency
unsteady phenomena start developing around the same value
For given mass ratio, forcing amplitude and frequency, of the rigidity. Several factors lead to unsteadiness of the
the mean thrust, mean power input, and propulsive efficiency problem, including transitions in the wake behind the flap-
were computed for each value of the rigidity ␩. As a general ping wing and development of the fluttering instability for
result, starting from the case of a rigid wing 共␩ → ⬁兲, the low values of the rigidity 共see Sec. III兲.
mean thrust 具T典 and power input 具P+典 both increase when The behavior of 具T典, 具P+典, and r collapse rather well for

Downloaded 22 Apr 2013 to 129.174.21.5. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
071902-6 S. Michelin and S. G. Llewellyn Smith Phys. Fluids 21, 071902 共2009兲

(a) f¯ = 1/π (b) f¯ = 5/2π


 


T /T ∞






 


P + /P + ∞




 

 
Efficiency r

 

 

 

     
     
         
 

FIG. 3. Evolution of the mean thrust 具T典 共top兲, power input 具P+典 共center兲, and propulsive efficiency 共bottom兲 with the wing’s rigidity ␩, for a flapping wing
of mass ratio ␮ = 0.2, flapping frequency 共a兲 f̄ = 1 / ␲ and 共b兲 f̄ = 5 / 2␲, forcing amplitude ␧ = 0.1 共solid circle兲, ␧ = 0.2 共dash-star兲, and ␧ = 0.5 共dot-square兲. For
comparison purposes, the mean power input and thrust have been normalized using their rigid-case value as ␩ → ⬁.

leading-edge flapping amplitudes up to 50% of the wing’s B. Wake structure and thrust production
length and lower frequencies 关Fig. 3共a兲兴. However, for higher 1. Evolution of the wake structure
frequencies, the influence of the forcing amplitude can be with increasing flexibility
seen for values of ␧ as low as 0.1 关Fig. 3共b兲兴. As a general
For given mass ratio ␮, forcing frequency f̄, and forcing
result, the propulsive efficiency and the normalized mean
amplitude ␧, the variation of the flexibility of the wing in-
thrust and power input are observed to decrease with ␧ when duces important changes in the developed thrust and energy
all other parameters are held fixed. The optimal values of ␩ use. As ␩ is decreased from the rigid case 共␩ → ⬁兲, the wake
for maximal thrust or maximal efficiency are also observed behind the flapping wing also undergoes important modifica-
to increase with ␧. tions. Figure 4 shows the evolution of the flow pattern
It must be emphasized here that the decrease in the around the heaving wing for varying ␩. In the high-rigidity
achievable thrust and power input with the forcing amplitude case 关Fig. 4共a兲兴, the trailing-edge deflection is small. The
␧ is only relative to the rigid case: For higher forcing ampli- wake vortices are arranged in a reverse Von Kármán vortex
tudes, the absolute mean thrust and power input are larger in street, in which vortices with positive 共negative兲 intensity are
magnitude and, in the rigid limit, are observed to scale like positioned above 共below兲 the horizontal axis. This arrange-
␧5/2. ment induces an acceleration of the fluid on the horizontal

Downloaded 22 Apr 2013 to 129.174.21.5. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
071902-7 Resonance and propulsion performance Phys. Fluids 21, 071902 共2009兲

(a) Rigid case η = 890 As a comparison, Fig. 4共c兲 shows the flow pattern for the
 value of ␩ corresponding to the thrust-drag transition. One
observes immediately that the intensity of the wake vortices

has decreased significantly and the width of the vortex street
 has almost vanished. In intermediate-Re experiments, the

formation of a classical Von Kármán street is associated with
 drag on the generating solid body. It appears here that even at

the transition between thrust and drag production, the reverse
     
Von Kármán pattern persists, although largely weakened.
 This result is consistent with the observation in recent
experiments13 that the thrust-drag transition does not neces-
(b) Maximum Thrust η = 3.2 sarily occur at the same time as the transition in the wake
 structure. A theoretical argument for this difference is also
presented in the next subsection. Furthermore, in purely in-

viscid simulations, the Von Kármán street in the drag-
 producing case is generally much weaker and instead the

vortices tend to align along the axis.28,35



2. Classical and reverse Von Kármán streets
      and thrust/drag production

Here, theoretical results on vortex streets are used to
(c) Thrust-Drag transition η = 8 10−3 understand the relationship between drag/thrust production
 and vortex wake structures. We focus on highly periodic
cases where the structure of the reverse Von Kármán or clas-

sical Von Kármán streets is easily identified.

 For a given vortex wake intensity and vortex arrange-
ment, the advection velocity of the vortices is the superposi-

tion of the incoming flow velocity 共equal to 1 in nondimen-
 sional units兲 and of the induced velocity V of the vortex
      street, which is itself a direct function of the width of the

vortex street 共vertical distance between the two rows of op-
FIG. 4. Streamlines of the flow over the flapping wing for ␮ = 0.2, ␧ = 0.1, posite sign vortices兲, its intensity, and its wavelength 共hori-
and f̄ = 5 / 2␲ and decreasing rigidity ␩. The positive 共negative兲 vortices are zontal distance between two successive vortices of identical
represented with upward- 共downward-兲 pointing triangles with sizes scaled sign兲. If the vortex street was infinite, its induced velocity V
to the intensity of the vortices. The streamlines are plotted for t = 20 when would be39

冉 冊
the leading edge crosses the horizontal axis, and the leading-edge forcing for
0 ⱕ t ⱕ 20 was the same in all three cases. ⌫w ␲b
V= tanh , 共30兲
2a a
where ⌫w is the magnitude of the vortices, b is the width of
axis to the right and the vortices have a higher velocity than the vortex street, and a is its horizontal wavelength. In Eq.
the imposed background flow 共V ⬎ 0兲. This jet carries mo- 共30兲, the following convention is chosen: b ⬎ 0 共b ⬍ 0兲 for a
mentum to the right which is consistent with the formation of reverse 共classical兲 Von Kármán street in which positive vor-
a mean thrust on the wing. The formation of a reverse Von tices are located above 共below兲 the horizontal axis. If the
Kármán vortex street in thrust-producing propulsion schemes shedding frequency 共equal to the flapping frequency f̄兲 is
is well known and has been observed in several experimental
fixed, then a = 共1 + V兲 / f̄. Measuring the width of the vortex
studies.12,13
street b and its intensity ⌫w, Eq. 共31兲 leads to a nonlinear
When ␩ is decreased, the solid is more flexible and the
equation for VL, the predicted value of the induced velocity
trailing-edge flapping amplitude D increases. As a result, the
that can be solved numerically for given b and ⌫w,

冉 冊
intensity of the wake vortices increases with decreasing ri-
gidity and so does the thrust generated by the flapping mo- ⌫w f̄ ␲bf̄
tion. Figure 4共b兲 shows the streamlines and arrangement of VL = tanh . 共31兲
2共1 + VL兲 共1 + VL兲
the wake vortices for ␩ = 3.2 which corresponds to the first
peak in thrust production in Fig. 3共b兲. The phase between the Figure 5 shows a very good agreement between the mea-
leading and trailing-edge displacements has also been modi- sured value V and the expected value VL for varying ␩ 共the
fied compared to the rigid case. The increase in the vortex other parameters taking the same values as in Fig. 4兲, par-
advection velocity can be seen as the distance between two ticularly above the thrust-drag transition at ␩ = 8 ⫻ 10−3 共Fig.
successive vortices is slightly increased 共the vortex shedding 5兲. This agreement shows that the induced velocity is mostly
frequency is unchanged and equal to the forcing frequency兲. determined by the neighboring vortices, and the semi-infinite

Downloaded 22 Apr 2013 to 129.174.21.5. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
071902-8 S. Michelin and S. G. Llewellyn Smith Phys. Fluids 21, 071902 共2009兲

 4

3.5
Induced velocity V

3

2.5

 2

1.5

1
      
      0.5

0
FIG. 5. 共Square兲 Induced wake vortices velocity V behind the flapping wing −0.5
for varying ␩ and ␮ = 0.2, ␧ = 0.1, and f̄ = 5 / 2␲. The predicted value obtained −1
for an infinite reverse Von Kármán vortex street 共Ref. 39兲 is plotted for 10
−2
10
−1
10
0
10
1
10
2
10
3

comparison 共solid line兲. η

FIG. 7. Evolution of the mean thrust 具T典 共solid line兲, trailing-edge flapping
amplitude D 共dash-circle兲, vortex wake intensity ⌫m 共dot-square兲, and in-
or infinite nature of the vortex street does not influence the
duced vortex velocity V 共dot-triangle兲 for ␮ = 0.2, ␧ = 0.1, and f̄ = 5 / 2␲. All
induced velocity significantly. quantities have been normalized by their rigid-case value 共␩ → ⬁兲.
The drag on a solid body placed in a uniform parallel
flow and shedding vortices in the pattern of a staggered vor-
tex street 共regular or reverse Von Kármán street兲 was com-
puted by Von Kármán using the conservation of momentum In Eq. 共32兲, the thrust consists of two terms. The second
around the solid body and part of its wake.36,40 The only term is positive in the case of a reverse Von Kármán street
important elements in the derivation are again the intensity 共b ⬎ 0兲 and negative in the classical Von Kármán street 共with
⌫w, wavelength a, and width b of the vortex street. In par- b ⬍ 0 and V ⬍ 0, as long as V ⬎ −1 / 2, which always occur in
ticular, the motion 共other than the main translation兲 and de- the weak Von Kármán streets observed here兲. The presence
formation of the body are irrelevant in Von Kármán’s deri- of the first term 共which is always negative and therefore al-
vation. These results can easily be generalized to the case of ways leads to drag production兲 is responsible for the experi-
a reverse Von Kármán street and in our notation, the pre- mentally and numerically observed differences between the
dicted mean thrust TS is obtained as thrust/drag transition and the transition in the wake structure.
As in the passive flapping flag case,28,35 it is possible to have
⌫w2 ⌫ wb net drag produced by a weak reverse Von Kármán street as
TS = − + 共1 + 2V兲, with a = 共1 + V兲/f̄ . 共32兲
2␲a a the first term in Eq. 共32兲 dominates the second one.
This theoretical prediction compares very well to the results C. Resonance and optimal flexibility
obtained for the mean thrust in our simulations 共Fig. 6兲. In for thrust production
particular, the transition between thrust and drag production
is well reproduced. The agreement is lost at low values of ␩: In Fig. 3共b兲, successive peaks in the mean thrust created
For such low values of the rigidity, the highly regular struc- by the heaving wing can be observed, but also a peak in the
ture of the wake is also lost as some natural modes of the mean drag 共or negative thrust兲 below the drag-thrust transi-
passive elastic sheet become unstable to fluttering and the tion that occurs around ␩ ⬃ 8 ⫻ 10−3. Figure 7 shows the evo-
motion of the wing loses its strong periodicity, making the lution of the mean thrust, flapping amplitude D, wake inten-
definition of a, ⌫w, and b difficult. sity ⌫w, and wake induced velocity V for the same values of
the parameters ␮, ␧, and f̄ as in Fig. 3.
A very clear correlation is observed between the occur-

rence of the maxima 共in magnitude兲 for the mean thrust 共or
 drag兲 and for the other quantities, which confirms the argu-
Mean thrust T 

 ment presented in Sec. IV B: The mean thrust and drag peaks
are created by an increase in the flapping motion at the trail-

ing edge, where the vortex wake is formed. An increase in D

共with a constant flapping frequency兲 induces higher relative
 velocity at the trailing edge, and therefore stronger shed vor-

tices. While this increase in ⌫m with D is physical, it is



 


 
 
 however not possible to find a simple scaling of ⌫m with D,
 as other factors that depend on ␩ must be taken into account
共e.g., the orientation of the trailing edge relative to its veloc-
FIG. 6. 共Square兲 Mean thrust 具T典 produced by the flapping wing for varying
ity兲. The induced velocity on the vortex street V is therefore
␩ and ␮ = 0.2, ␧ = 0.1, and f̄ = 5 / 2␲. The predicted value TS 关Eq. 共32兲兴 based
also increased and the wake carries a larger fluid momentum
on impulse conservation 共Refs. 36 and 40兲 is also plotted for comparison
共solid line兲. TS was computed from Eq. 共32兲 using the values of the vortex downstream, thereby creating a greater thrust on the profile.
street intensity, width, and induced velocity obtained with our model. It is also interesting to notice that the maximum drag

Downloaded 22 Apr 2013 to 129.174.21.5. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
071902-9 Resonance and propulsion performance Phys. Fluids 21, 071902 共2009兲

 
Flapping amplitude D

Flapping amplitude D









   
  

 


 









FIG. 8. Evolution of the normalized trailing-edge flapping amplitude D with
冑␩ / ␮ for ␧ = 0.05, f̄ = 5 / 2␲ and for ␮ = 0.2 共solid squares兲 and ␮ = 2 共dash- FIG. 9. Influence of the forcing amplitude ␧ on the position of the reso-
circle兲. The flapping amplitude was normalized using the asymptotic rigid- nances in the trailing-edge flapping amplitude for ␮ = 0.2 and f̄ = 5 / 2␲. The
case limit 共␩ → ⬁兲. results are plotted for ␧ = 0.01 共solid circles兲, ␧ = 0.05 共dashed line兲, ␧ = 0.1
共dot-dashed line兲, ␧ = 0.2 共dotted line兲, and ␧ = 0.5 共solid line兲. The trailing-
edge flapping amplitude was normalized by its rigid-case value 共␩ → ⬁兲.

observed for ␩ ⬃ 5 ⫻ 10−3 is also associated with a maximum


in D and ⌫m. In that case, the large amplitude of motion at
the trailing edge opposes the imposed flow and creates a net 兵wing+ outside uniform flow其 should be considered. A com-
drag on the body. parison with linear analysis predictions is proposed below in
Sec. V A 3.
V. INFLUENCE OF FLEXIBILITY ON THE FLAPPING Before comparing the numerical results to the linear
AMPLITUDE AND MODE SHAPE analysis, we study the influence of the forcing amplitude ␧
on the position of the resonances. It was noticed in Sec. IV A
A. Resonances between the forcing frequency that the normalized thrust and power input follow similar
and the natural frequencies of the system patterns for values of ␧ up to 0.2–0.5 depending on the value
In this section, we are interested in the origin of the of ␮. In an attempt to determine the exact value of ␩ leading
maxima in the trailing-edge flapping amplitude D. The suc- to a resonance, we study how the resonance peaks in D are
cessive peaks in D as ␩ is varied suggest a resonance phe- modified as ␧ is varied between 0.01 and 0.5. Starting from
nomenon. By varying the rigidity of the solid ␩, the natural small ␧, one observes that the increase in ␧ induces a shift of
frequencies of the system are also modified. For an elastic the resonance peaks toward larger values of ␩ and a smooth-
sheet in vacuum, these frequencies scale like 冑B / ␳s. Looking ing of the resonance peaks 共Fig. 9兲. Also, the convergence
for solutions of Eqs. 共8兲–共11兲 in the linear limit with no fluid toward the limit case of small ␧ is faster for the second and
forcing, the fundamental angular frequencies ␻0n = 2␲ f n of a third resonance peaks than for the first one 共labeling the
clamped-free elastic sheet in vacuum are obtained in nondi- peaks from the right as ␩ is decreased from the rigid-case
mensional form as limit兲.

␻0n = ␭2n 冑 ␩

, with 1 + cosh ␭n cos ␭n = 0. 共33兲 2. Absolute and relative trailing-edge
flapping amplitude
The ratio 冑␮ / ␩ can be thought of as the nondimensional
In Secs. II–IV, D was defined as the peak-to-peak am-
time scale associated with the frequency of the sheet’s natu-
plitude of the trailing-edge flapping motion in the laboratory
ral oscillations in vacuum, or alternatively as the outside
frame 共thereafter referred to as absolute flapping amplitude兲.
flow velocity nondimensionalized by the characteristic ve-
In the following, we are also interested in the motion of the
locity associated with the sheet’s properties.
wing in the frame moving with the leading edge. In this
frame, the peak-to-peak trailing-edge flapping amplitude is
1. Influence of ␮ and ε on the position Dⴱ 共thereafter referred to as relative flapping amplitude兲. D
of the resonance peaks in D and Dⴱ are not necessarily equal because of the phase differ-
However, the resonance peaks observed in Fig. 7 do not ence between the motion of the leading and trailing edges
correspond to a resonance with the natural frequency of the 共Fig. 10兲. This delay is, in general, a decreasing function of
elastic sheet in vacuum. If this were the case, then the reso- the solid’s rigidity. In the limit of a rigid wing ␩ → ⬁, the
nance would be achieved for all values of the mass ratio ␮ at trailing edge flaps in phase with the leading edge with the
the same value of 冑␩ / ␮. Such a coincidence does not occur same amplitude 共Dⴱ = 0 and D = 2␧兲. As ␩ is decreased, a
共Fig. 8兲: The position of the successive resonances is actually delay appears between the motion of the leading and trailing
strongly influenced by the fluid-solid inertia ratio ␮. This edges: The elasticity of the wing takes more time to carry
difference makes sense physically, as the eigenfrequencies of along its length the signal imposed at the leading edge. As a
the system are modified by the presence of the forcing hori- result, the relative amplitude Dⴱ increases.
zontal flow, and such effects as added inertia are expected to One of the main consequences of the appearance of such
be important. Instead, the natural frequencies of the system a delay is the noncoincidence of the first resonance peak

Downloaded 22 Apr 2013 to 129.174.21.5. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
071902-10 S. Michelin and S. G. Llewellyn Smith Phys. Fluids 21, 071902 共2009兲

! &

#$%
"$'

()*++,-. *0+),1234 #

'()*+),-.
"$&
!$%

"$%
!

"$# "$%
/01 3"$#
5*6 "
" # " # !"
#
!"
"

!" !" !" 



'
&

&
#$%

%
#

'()*+),-.
()*+,

$ !$%

# !

! "$%
-./ /01 3#
" # " # " ! " ! #
!" !" !" !" !" !" !"
 

FIG. 10. 共a兲 Comparison between the absolute flapping amplitude D 共solid FIG. 11. Position of the resonances for the trailing-edge flapping amplitude
line兲 and relative flapping amplitude Dⴱ 共dashed line兲 for the trailing edge, in the 共␩ , f̄兲-plane obtained using the present model 共symbols兲 with ␧
for ␮ = 0.2, f̄ = 5 / 2␲, and ␧ = 0.1. 共b兲 Phase difference between the leading- = 0.01 and 共a兲 ␮ = 0.2 and 共b兲 ␮ = 2. The different symbols correspond to the
edge and trailing-edge motions. nature of the mode: mode 1 has no neck in the flapping envelope 共circles兲;
mode 2 共squares兲, mode 3 共upward-pointing triangle兲, and mode 4
共downward-pointing triangles兲 have one, two, and three necks, respectively,
共with largest ␩兲 in the flapping amplitude, whether D or Dⴱ in their motion envelope. The black symbols correspond to resonances in the
absolute flapping amplitude D. Open symbols correspond to resonances in
is considered. The position of the subsequent peaks is not the relative flapping amplitude Dⴱ when they differ from the resonances in
significantly affected 关Fig. 10共a兲兴. D. The position of the resonances is compared to the prediction of the linear
analysis 共dashed line兲 for the natural frequency of the purely passive elastic
sheet 共or flag兲 in axial flow 共see the Appendix兲.
3. Comparison with the natural frequencies
of the system
In this section, the conjecture that the trailing-edge flap- 关greater than 20␧ at the resonance in the case of the heavier
ping amplitude peaks are due to a resonance with the natural wing 共␮ = 2兲, thereby representing more than 20% of the
frequencies of the 兵wing+ imposed flow其 system is tested by wing’s length兴. These results are compared to the natural
comparing the value of the rigidity ␩ leading to a peak value frequencies of the system 兵wing+ parallel flow其 as predicted
for D to the value of ␩ for which the forcing frequency by the linear analysis and very good agreement is found be-
matches a natural frequency of the system. The position of tween the numerical results and the theoretical predictions.
the peaks in D is measured in the limit of small forcing The position of the resonances in Dⴱ is also indicated. Reso-
amplitude 共typically ␧ = 0.01兲. The natural frequencies of a nances in Dⴱ and D coincide except for the first resonance
purely passive wing clamped at its leading edge in a uniform peak in the case of the smaller mass ratio ␮.
flow are determined using the linear stability analysis The linear analysis seems to underpredict slightly the
method developed by Kornecki et al.41 A brief summary of values of ␩ corresponding to the resonance for a given fre-
this method is given in the Appendix. For given ␮ and ␩, the quency ␻. This difference is consistent with the amplitude of
eigenfrequencies of the system are computed. The lowest discrepancy between the point vortex model and the linear
frequencies are also associated with the modes of lowest or- stability analysis observed in the study of a purely passive
der 共those with the longest wavelength兲. Here, the following elastic sheet or flag.28 Two other factors can also explain the
equivalent problem is considered: For a given ␮ and f̄, we small discrepancy in the results.
want to find the values ␩ for which a mode of the passive — As pointed out, the motion of the wing is not infinitesi-
elastic sheet in axial flow has the particular frequency f̄, mally small here, even for ␧ = 0.01; it is therefore pos-
regardless of its growth rate. sible that nonlinear effects modify the exact position of
Figure 11 shows the position in the 共␩ , f̄兲-plane of the the resonances. In Sec. V A, the effect of increasing ␧
first resonances observed in the trailing-edge flapping ampli- was shown to shift the resonance peaks toward larger ␩,
tude D for small heaving amplitude ␧, starting from the rigid particularly for the first peak, and this could account for a
case ␩ → ⬁. Although ␧ is small, D can be significant significant part of the observed discrepancy.

Downloaded 22 Apr 2013 to 129.174.21.5. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
071902-11 Resonance and propulsion performance Phys. Fluids 21, 071902 共2009兲

— The amplitude of flapping in the laboratory frame is con- VI. CONCLUSIONS


sidered here. However, even for small ␧, this amplitude
differs from the relative flapping amplitude defined in the Using a reduced-order model for the flow past a two-
frame moving with the leading edge 共see Fig. 10兲 be- dimensional heaving flexible wing, the influence of the
cause of the existence of a nonzero phase between the wing’s flexibility on its propulsive performance 共mean thrust,
motion of the leading and trailing edges. It was observed flapping efficiency兲 was investigated. Starting from the
in Fig. 10 that the resonance in relative amplitude occurs purely rigid case, we observed that the flexibility of the wing
for smaller values of ␩, particularly for the first reso- allowed for a larger trailing-edge flapping amplitude, thereby
nance 共largest ␩兲. In the linear analysis, both amplitudes generating a stronger wake and an increased mean thrust.
共relative and absolute兲 are identical since the leading The energy usage also increases with the introduction of
flexibility, but more slowly than the mean thrust, resulting in
edge is held fixed. For ␮ = 0.2, the agreement is improved
a net increase in the flapping efficiency with reduced rigidity.
for the position of the first resonance if Dⴱ is considered
This efficiency gain can be significant 共up to twice the effi-
instead of D 共Fig. 11兲.
ciency of the rigid case兲. While the mean thrust and power
Figure 11 also indicates the nature of the observed mode, input display several peaks when the rigidity ␩ is decreased
in particular, the number of necks in the envelope of the from the purely rigid case, the flapping efficiency displays
wing’s motion. One observes that resonances located on a one wide peak before falling sharply as ␩ nears the value
same branch of the linear analysis prediction share the same leading to the thrust-drag transition. Below this threshold,
general structure, and the number of necks is consistent with the wing is too flexible to communicate momentum to the
flow and instead starts creating a net drag on the leading-
that predicted by the linear analysis: For given ␮ and ␩, the
edge attachment.
lowest frequency mode has the longest wavelength and no
The relationship between thrust production and wake
neck in its envelope. The next lowest frequency corresponds
structure was then investigated, taking advantage of the dis-
to a mode with one neck, and so on for the successive fre-
crete representation of the wake. Analytical predictions for
quencies. The evolution of the mode shape for varying ␩ is
the induced vortex street velocity and mean thrust in terms of
studied in more detail in the following subsection. the vortex street strength and spatial arrangement were suc-
cessfully compared to our simulation results.
B. Evolution of the flapping mode shape
The peaks in mean thrust were found to correspond to
with the flexibility of the profile
maximum values in the trailing-edge amplitude, and shown
To confirm that the maxima of flapping amplitude actu- to be the result of the resonance between the forcing fre-
ally correspond to resonances between the forcing frequency quency of the heaving motion and the natural frequencies of
and the natural frequency of the passive elastic sheet in a the system. A quantitative comparison showed very good
parallel flow, the evolution of the flapping mode shape with agreement between the optimal values of the solid’s rigidity
␩ is now considered. For comparison with the case of a and the linear analysis predictions for the resonances posi-
purely passive flexible elastic sheet, the mode shape is de- tion. The existence of these resonance phenomena was fur-
fined as the envelope of the motion of the wing in the frame ther confirmed by comparing the flapping mode shape to the
moving with the leading edge. Figure 12 shows the mode mode shape observed for a freely flapping elastic sheet 共e.g.,
shape in permanent regime at the values of ␩ leading to peak the flag problem28兲.
values of the absolute trailing-edge flapping amplitude and to The natural frequencies of the system are strongly de-
the minima between two successive peaks. Note that in Fig. pendent on both the solid’s flexibility and the ratio of fluid
and solid inertia, and so are the optimal values of the solid’s
12, case B seems to correspond to a wider envelope than case
rigidity. For a slender neutrally buoyant fish fin, the inertia
A, although the absolute flapping amplitude is smaller for B
ratio ␮ can generally be neglected.29 However, in the case of
than for A. This is the result of the change in frame: If one
an insect wing, the small thickness-to-chord ratio is balanced
considers the flapping amplitude in the moving frame, the
by the large difference in density for the fluid and solid, and
position of the maxima differs from peaks in absolute flap-
the present analysis shows that the mass ratio ␮ plays an
ping amplitude 共see Fig. 10兲. Cases A, C, and E correspond important role in defining the optimal value of the flexibility.
to the resonances while B, D, and F correspond to local The model presented here used a potential flow represen-
minima of the absolute flapping amplitude. tation and the shedding of point vortices to describe the
The mode shapes C and E are structurally similar to the highly unsteady flow around the insect wing. Vortices were
envelope of the first two flapping modes observed for the shed only from the trailing edge. This approximation is rea-
passive flexible flag,28 which is consistent with C and E cor- sonable for small angles of attack 共proportional to the Strou-
responding to resonances between the flapping frequency hal number St in the case of a purely heaving motion兲. Then,
and the flapping modes 2 and 3 共one- and two-neck modes, the vorticity shed at the leading edge is negligible or merges
respectively兲. The mode shape A corresponds to mode 1 of with the vorticity shed by separation of the boundary layers
the passively flapping elastic sheet, which was not observed at the trailing edge,12 leading to the shedding of two indi-
in the case of the flapping flag study28 as it is always stable42 vidual vortices every flapping period and so-called 2S
and therefore does not lead to spontaneous large-scale flap- wakes.43 However, when the flapping amplitude and fre-
ping. quency are increased, more complex wakes are expected as

Downloaded 22 Apr 2013 to 129.174.21.5. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
071902-12 S. Michelin and S. G. Llewellyn Smith Phys. Fluids 21, 071902 共2009兲

0.8
Flapping amplitude

0.6

0.4

0.2

(F) (E) (D) (C) (B) (A)


0 −2 −1 0 1 2 3
10 10 10 10 10 10
η

0.5
(A) (B) (C)

−0.5
0.5
(D) (E) (F)

−0.5
0 0.5 1 0 0.5 1 0 0.5 1
FIG. 12. Evolution of the mode shape with ␩ for ␮ = 0.2, ␧ = 0.1, and f̄ = 5 / 2␲. 共Top兲 Trailing-edge flapping amplitude in the stationary frame D 共solid line兲
and in the frame attached to the leading edge Dⴱ 共dotted line兲. Note that the value of ␩ leading to a resonance in D does not necessarily correspond to the value
of ␩ leading to a resonance in Dⴱ. 共Bottom兲 Mode shape plotted for the value of ␩ indicated on the top panel. The position of the wing in the frame moving
with the leading edge is plotted every ⌬t = 0.06.

vorticity is shed from both the trailing and leading edges, isolated from other factors such as leading-edge and wing-tip
inducing, for example, the formation of the so-called 2P vortices.
wake, where a vortex pair is shed during each half period. Despite these limitations, the present method offers the
The representation of the leading-edge vortex falls however advantage of a considerable reduction in computational cost
beyond the scope of inviscid methods such as the present for two-dimensional fluid-solid simulations. This is particu-
point vortex model or a vortex sheet approach as the effect of larly attractive for situations where the cost of full numerical
viscosity cannot be neglected:27,44 The leading-edge vortex is simulation is prohibitive and a large number of simulations
expected to remain close to the body for a sufficiently long are required 共typically in the case of optimization problems兲.
time to interact with the boundary layers. The use of point The present work was purposely limited to a one degree
vortices also restricts this method to two-dimensional prob- of freedom flapping pattern 共pure heaving兲 in order to limit
lems and therefore does not allow the study of such effects as the number of free parameters and focus on the fundamental
wing-tip vortices that are expected to influence significantly effect of solid flexibility on the flapping performance. In fu-
the flight performance. The present study however does not ture work, more realistic flapping schemes should be consid-
aim at reproducing the exact flow around an insect wing but ered to follow more closely the flapping pattern of an insect
to present a test case where the effect of flexibility can be wing. In particular, a combination of heaving and pitching

Downloaded 22 Apr 2013 to 129.174.21.5. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
071902-13 Resonance and propulsion performance Phys. Fluids 21, 071902 共2009兲

should be considered to determine the influence of the rela- tion can be found in Refs. 41 and 50. Small vertical displace-
tive phase between heaving and pitching motions on the re- ments of the sheet h共s , t兲 are considered 共0 ⱕ s ⱕ 1兲 with 兩h兩
sults presented in this work. The interaction between two Ⰶ 1 so that the equation of motion of solid 关Eq. 共5兲兴 becomes
flapping flexible sheets should also be investigated to under- a linearized form
stand the lift and thrust generation by insects with multiple
pairs of flexible wings 共e.g., dragonfly45兲 or the efficiency of ␮ḧ = − ␩hssss − ⌬p. 共A1兲
fish schooling, and to complement recent experimental stud-
ies on multiple passive flexible filaments.46–49 The wake is represented by a continuous distribution of vor-
ticity ␥共s , t兲 along the horizontal axis 共s ⱖ 1兲 advected by the
ACKNOWLEDGMENTS flow 共this continuous shedding thereby differs from the dis-
crete approach presented in the present paper, but is better
This work was supported by the Human Frontier Science suited to linear stability analysis兲. The pressure forcing can
Program Research under Grant No. RGY 0073/2005. be decomposed into two parts:41,50,51 a noncirculatory part
due to the flow created by the solid’s motion with no net
APPENDIX: NATURAL FREQUENCIES OF A PASSIVE
circulation around the solid and a circulatory part due to the
FLEXIBLE SHEET IN AXIAL FLOW
flow created by the vortex wake 共the absence of circulation at
We present here a brief summary of the linear stability infinity requiring the existence of a net circulation around the
analysis of a clamped-free flexible sheet or flag using a vor- solid equal to the opposite of the wake vorticity兲. Consider-
tex sheet representation of the wake as developed by ing a decomposition of the solid’s motion onto normal modes
Kornecki.41 More details on the method and the full calcula- of the form h共s , t兲 = Re关y共s兲ei␻t兴, y共s , t兲 satisfies41

冑 冑
冨 冨
s ␰

冕 冕冑
+
2i␻ 1
1−s 1−␰ 2 1
␰共1 − ␰兲

冑 冑
− ␮␻2y + ␩y ssss = − log 共i␻y + y ␰兲d␰ + 共i␻y + y ␰兲d␰
␲ 0 s ␰ ␲冑s共1 − s兲 0 s−␰

1−s 1−␰


2
␲ 冋冕 冑
0
1

1−␰
共i␻y + y ␰兲d␰ 册冋 2s − 1
冑s共1 − s兲 + 冑 1−s
s
C共␻兲 , 册 共A2兲

where C共␻兲 is the Theodorssen function,50 Kij = ␭4j ␦ij, M 共K兲 共i兲 共j兲 共i兲 共j兲
ij = F1共␺ , ␺s 兲 + F2共␺ , ␺s 兲,

2H共2兲
1 冉冊 ␻
2

冉冊 冉冊
C共␻兲 = , 共A3兲 M 共G兲 共i兲 共j兲 共i兲 共j兲 共i兲 共j兲
␻ ␻ ij = F1共␺ , ␺ 兲 + F2共␺ , ␺ 兲 + F3共␺ , ␺s 兲,
H共2兲
1 + iH共2兲
0
2 2

and H␯共2兲共x兲 = J␯共x兲 − iY ␯共x兲 共␯ = 0 , 1兲 are Hankel functions of M 共M兲 共i兲 共j兲 共i兲 共j兲
ij = F3共␺ , ␺ 兲, ij = F4共␺ , ␺s 兲,
M C1
the second kind.52
The eigenvalue problem 共A2兲 for ␻ is solved numeri-
cally using a Galerkin method: y共s兲 is decomposed along the
共i兲 共j兲
first N eigenmodes of the clamped-free beam in vacuum ij = F4共␺ , ␺ 兲,
M C2
y共s兲 = 兺␣n␺共n兲共s兲 with ␺共n兲共s兲 satisfying
共n兲
␺ssss = ␭4n␺共n兲, ␺共n兲共0兲 = ␺s共n兲共0兲 = ␺ss
共n兲 共n兲
共1兲 = ␺sss 共1兲 = 0, with the functionals Fk共1 ⱕ k ⱕ 4兲 defined as

and ␭n are the successive positive roots of


1 + cos ␭n cosh ␭n = 0. Equation 共A2兲 is replaced by the non-
linear eigenvalue problem
共M兲 共G兲 共K兲
F1共f,g兲 = −
2

冕冑
0
1
f共x兲
x共1 − x兲
冋冕 冑
0
1
␰共1 − ␰兲
x−␰
g共␰兲d␰ dx, 册
关− ␻ 共␮I + M
2
兲 + i␻M + 共␩K + M 兲 + C共␻兲M C1

+ i␻C共␻兲M 兴 · ␣ = 0.
C2
共A4兲

The coefficients of the different N ⫻ N matrices are defined


as
F2共f,g兲 =
2
␲ 冋冕 0
1
共2x − 1兲f共x兲
冑x共1 − x兲 dx 册冋 冕 冑
0
1
x
1−x
g共x兲dx , 册
Downloaded 22 Apr 2013 to 129.174.21.5. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
071902-14 S. Michelin and S. G. Llewellyn Smith Phys. Fluids 21, 071902 共2009兲

"
! # !"
"%&
$%&
" !
$
!" %&'()*
'()*+, (.+/

'()*+),-.
"%&

! !%&
#
!"
!%&
! +,-&'()*
$
"%& $
$%& !"
# # $ ! " !
" # $ ! " ! ! " ! #
!" !" !" !" !" !" !" !" !" !" !" !" !" !"
  

FIG. 13. For ␮ = 0.2, evolution of 共left兲 the growth rate and 共center兲 frequency of the first four modes 共of lowest positive frequency兲: mode 1 共solid line兲, mode
2 共dashed line兲, mode 3 共dot-dashed line兲, and mode 4 共dotted line兲. 共Right兲 Critical stability curve as obtained in the 共␮ , ␩兲-plane from the linear stability
analysis.


12
2 1 J. M. Anderson, K. Streitlien, D. S. Barrett, and M. S. Triantafyllou, “Os-
F3共f,g兲 = f共x兲 cillating foils of high propulsive efficiency,” J. Fluid Mech. 360, 41
␲ 0 共1998兲.

冑 冑
13

冤冕 冨 冨冥
R. Godoy-Diana, J. L. Aider, and J. E. Wesfreid, “Transitions in the wake
x ␰ of a flapping foil,” Phys. Rev. E 77, 016308 共2008兲.
+ 14
T. Y. Wu, “Swimming of a waving plate,” J. Fluid Mech. 10, 321 共1961兲.
1
1−x 1−␰

冑 冑
⫻ g共␰兲log d␰ dx, 15
Z. J. Wang, “Two dimensional mechanism for insect hovering,” Phys. Rev.
0 x ␰ Lett. 85, 2216 共2000兲.
− 16
1−x 1−␰ Z. J. Wang, “Vortex shedding and frequency selection in flapping flight,” J.
Fluid Mech. 410, 323 共2000兲.

冋冕 冑 册冋 冕 冑 册
17
J.-M. Miao and M.-H. Ho, “Effect of flexure on aerodynamic propulsive
1 1
2 1−x x efficiency of flapping flexible airfoil,” J. Fluids Struct. 22, 401 共2006兲.
F4共f,g兲 = f共x兲 dx g共x兲dx . 18
␲ 0 x 0 1−x R. K. Shukla and J. D. Eldredge, “An inviscid model for vortex shedding
from a deforming body,” Theor. Comput. Fluid Dyn. 21, 343 共2007兲.
19
For a given N, these matrices can be precomputed. Then, for R. J. Wootton, “Functional morphology of insect wings,” Annu. Rev. En-
tomol. 37, 113 共1992兲.
given ␮ and ␩, the eigenvalue problem 共A4兲 is solved itera- 20
S. A. Combes and T. L. Daniel, “Flexural stiffness in insect wings I.
tively using a Newton–Kantorovitch algorithm.53 Figure 13 Scaling and the influence of wing venation,” J. Exp. Biol. 206, 2979
shows the evolution of the growth rate −Im共␻兲 and fre- 共2003兲.
quency ␻ / 2␲ for ␮ = 0.2 and varying rigidity ␩. In the rigid 21
S. A. Combes and T. L. Daniel, “Flexural stiffness in insect wings II.
case 共large ␩兲, all modes are stable. As ␩ is decreased below Spatial distribution and dynamic wing bending,” J. Exp. Biol. 206, 2989
共2003兲.
␩m = 2 ⫻ 10−3, one or more modes become unstable 共mode 3 22
B. S. H. Connell and D. K. P. Yue, “Flapping dynamics of a flag in
then mode 4兲. The critical stability curve separates the region uniform stream,” J. Fluid Mech. 581, 33 共2007兲.
of the 共␮ , ␩兲-plane where the elastic sheet’s state of rest is 23
L. Zhu and C. Peskin, “Simulation of flapping flexible filament in a flow-
stable and the region of the parameter space where at least ing soap film by the immersed boundary method,” J. Comput. Phys. 179,
one mode is unstable 共Fig. 13兲. 452 共2002兲.
24
L. Zhu and C. Peskin, “Interaction of two flapping filaments in a flowing
1 soap film,” Phys. Fluids 15, 1954 共2003兲.
M. J. Lighthill, “Note on the swimming of slender fish,” J. Fluid Mech. 9, 25
305 共1960兲. A. Bergou, S. Xu, and Z. J. Wang, “Passive wing pitch reversal in insect
2
M. J. Lighthill, “Hydromechanics of aquatic animal propulsion,” Annu. flight,” J. Fluid Mech. 591, 321 共2007兲.
26
Rev. Fluid Mech. 1, 413 共1969兲. J. Toomey and J. D. Eldredge, “Numerical and experimental study of the
3
S. Childress, Mechanics of Swimming and Flying 共Cambridge University fluid dynamics of a flapping wing with low-order flexibility,” Phys. Fluids
Press, Cambridge, 1981兲. 20, 073603 共2008兲.
4 27
M. S. Triantafyllou, G. S. Triantafyllou, and D. K. P. Yue, “Hydrodynam- S. Michelin and S. G. Llewellyn Smith, “An unsteady point vortex method
ics of fishlike swimming,” Annu. Rev. Fluid Mech. 32, 33 共2000兲. for coupled fluid-solid problems,” Theor. Comput. Fluid Dyn. 23, 127
5
Z. J. Wang, “Dissecting insect flight,” Annu. Rev. Fluid Mech. 37, 183 共2009兲.
共2005兲. 28
S. Michelin, S. G. Llewellyn Smith, and B. J. Glover, “Vortex shedding
6
A. L. R. Thomas, G. K. Taylor, R. B. Srygley, R. L. Nudds, and R. J. model of a flapping flag,” J. Fluid Mech. 617, 1 共2008兲.
Bomphrey, “Dragonfly flight: Free-flight and tethered flow visualizations 29
S. Alben, “Optimal flexibility of a flapping appendage in an inviscid
reveal a diverse array of unsteady lift-generating mechanisms, controlled fluid,” J. Fluid Mech. 614, 355 共2008兲.
primarily via angle of attack,” J. Exp. Biol. 207, 4299 共2004兲. 30
L. Cortelezzi and A. Leonard, “Point vortex model of the unsteady sepa-
7
C. P. Ellington, “The aerodynamics of hovering insect flight,” Philos. rated flow past a semi-infinite plate with transverse motion,” Fluid Dyn.
Trans. R. Soc. London, Ser. B 305, 1 共1984兲.
8 Res. 11, 263 共1993兲.
M. H. Dickinson, F. O. Lehmann, and S. P. Sane, “Wing rotation and the 31
L. Cortelezzi, “On the unsteady separated flow past a semi-infinite plate.
aerodynamic basis of insect flight,” Science 284, 1954 共1999兲.
9
Z. J. Wang, J. M. Birch, and M. H. Dickinson, “Unsteady forces and flows Exact solution of the Brown and Michael model, scaling and universality,”
in low reynolds number hovering flight: Two-dimensional computations vs Phys. Fluids 7, 526 共1995兲.
32
robotic wing experiments,” J. Exp. Biol. 207, 449 共2004兲. C. E. Brown and W. H. Michael, “Effect of leading edge separation on the
10
U. Pesavento and Z. J. Wang, “Falling paper: Navier–Stokes solutions, lift of a delta wing,” J. Aeronaut. Sci. 21, 690 共1954兲.
33
model of fluid forces, and center of mass elevation,” Phys. Rev. Lett. 93, N. Rott, “Diffraction of a weak shock with vortex generation,” J. Fluid
144501 共2004兲. Mech. 1, 111 共1956兲.
11 34
G. Berman and Z. J. Wang, “Energy-minimizing kinematics in hovering M. A. Jones, “The separated flow of an inviscid fluid around a moving
insect flight,” J. Fluid Mech. 582, 153 共2007兲. plate,” J. Fluid Mech. 496, 405 共2003兲.

Downloaded 22 Apr 2013 to 129.174.21.5. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
071902-15 Resonance and propulsion performance Phys. Fluids 21, 071902 共2009兲

35 45
S. Alben and M. J. Shelley, “Flapping states of a flag in an inviscid fluid: Z. J. Wang and D. Russell, “Effect of forewing and hindwing interactions
Bistability and the transition to chaos,” Phys. Rev. Lett. 100, 074301 on aerodynamic forces and power in hovering dragonfly flight,” Phys.
共2008兲. Rev. Lett. 99, 148101 共2007兲.
36
P. G. Saffman, Vortex Dynamics 共Cambridge University Press, Cambridge, 46
J. Zhang, S. Childress, A. Libchaber, and M. Shelley, “Flexible filaments
1992兲. in a flowing soap film as a model for one-dimensional flags in a two-
37
D. N. Gorelov, “Calculation of pressure on an airfoil contour in an un- dimensional wind,” Nature 共London兲 408, 835 共2000兲.
steady separated flow,” J. Appl. Mech. Tech. Phys. 49, 437 共2008兲. 47
L.-B. Jia, F. Li, X.-Z. Yin, and X.-Y. Yin, “Coupling modes between two
38
S. Alben, “The flapping-flag instability as a non-linear eigenvalue prob- flapping filaments,” J. Fluid Mech. 581, 199 共2007兲.
lem,” Phys. Fluids 20, 104106 共2008兲. 48
L.-B. Jia and X.-Z. Yin, “Passive oscillations of two tandem flexible
39
H. Lamb, Hydrodynamics, 6th ed. 共Dover, New York, 1932兲. filaments in a flowing soap film,” Phys. Rev. Lett. 100, 228104
40
N. E. Kochin, I. A. Kibel, and N. V. Roze, Theoretical Hydromechanics 共2008兲.
共Interscience, New York, 1964兲. 49
L. Ristroph and J. Zhang, “Anomalous hydrodynamic drafting of interact-
41
A. Kornecki, E. H. Dowell, and J. O’Brien, “On the aeroelastic instability ing flapping flags,” Phys. Rev. Lett. 101, 194502 共2008兲.
50
of two-dimensional panels in uniform incompressible flow,” J. Sound Vib. R. L. Bisplinghoff, H. Ashley, and R. L. Halfman, Aeroelasticity
47, 163 共1976兲. 共Addison-Wesley, Cambridge, MA, 1955兲.
42 51
C. Eloy, C. Souilliez, and L. Schouveiler, “Flutter of a rectangular plate,” T. Theodorsen, “General theory of aerodynamic instability and the mecha-
J. Fluids Struct. 23, 904 共2007兲. nism of flutter,” NACA Report No. 496, 1935.
43 52
C. H. K. Williamson and A. Roshko, “Vortex formation in the wake of an M. Abramowitz and I. A. Stegun, Handbook of Mathematical Functions
oscillating cylinder,” J. Fluids Struct. 2, 355 共1988兲. With Formulas, Graphs, and Mathematical Tables 共Dover, New York,
44
S. Michelin and S. G. Llewellyn Smith, “Falling cards and flapping flags: 1964兲.
53
Understanding fluid-solid interactions using an unsteady point vortex J. P. Boyd, Chebyshev and Fourier Spectral Methods, 2nd ed. 共Dover,
model,” Theor. Comput. Fluid Dyn. 共in press兲. New York, 2001兲.

Downloaded 22 Apr 2013 to 129.174.21.5. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions

You might also like