Sie sind auf Seite 1von 10

PHYSICS OF FLUIDS 22, 111902 共2010兲

Effect of wing inertia on hovering performance of flexible flapping wings


Bo Yin and Haoxiang Luoa兲
Department of Mechanical Engineering, Vanderbilt University, 2301 Vanderbilt Pl., Nashville,
Tennessee 37235-1592, USA
共Received 31 March 2010; accepted 14 September 2010; published online 10 November 2010兲
Insect wings in flight typically deform under the combined aerodynamic force and wing inertia;
whichever is dominant depends on the mass ratio defined as mⴱ = ␳sh / 共␳ f c兲, where ␳sh is the surface
density of the wing, ␳ f is the density of the air, and c is the characteristic length of the wing. To
study the differences that the wing inertia makes in the aerodynamic performance of the deformable
wing, a two-dimensional numerical study is applied to simulate the flow-structure interaction of a
flapping wing during hovering flight. The wing section is modeled as an elastic plate, which may
experience nonlinear deformations while flapping. The effect of the wing inertia on lift production,
drag resistance, and power consumption is studied for a range of wing rigidity. It is found that both
inertia-induced deformation and flow-induced deformation can enhance lift of the wing. However,
the flow-induced deformation, which corresponds to the low-mass wing, produces less drag and
leads to higher aerodynamic power efficiency. In addition, the wing deformation has a significant
effect on the unsteady vortices around the wing. The implication of the findings on insect flight is
discussed. © 2010 American Institute of Physics. 关doi:10.1063/1.3499739兴

I. INTRODUCTION phase differences between pitching and heaving, which is in


contrast with the relative sensitivity of a rigid wing to this
Aerodynamics of insect flight has drawn considerable
parameter. Full three-dimensional numerical simulations of
attention in recent years due to its promising application in
the desert locust, Schistocerca gregaria, were recently per-
the development of biomimetic microair vehicles.1 During
formed by reconstructing the detailed wing kinematics, in-
flight, insect wings typically experience dynamic
cluding the time-varying camber and spanwise twist of the
deformations,2 i.e., change of the wing shape from its rest
wing surface, from the high-speed digital video of the real
configuration regardless the position or orientation of the
insect flight.6 By comparing the performance of the wing
wing. Even though the deformation magnitude and pattern
vary from species to species, physiological studies have model based on the fully reconstructed kinematics and that
shown that insect wings do not possess an internal actuation of the corresponding wing models without the camber or
mechanism and thus the deformation has to be passive.3 That twist, Young et al.6 found that the wing deformation leads to
is, an insect wing is deformed by either the aerodynamic substantial power economy in lift production. Meanwhile,
force from the surrounding air, or the inertial acceleration, or they noticed that the leading edge vortex remains attached to
a combination of both. the wing during the entire flapping cycle in the full-
The structural deformation during flapping may signifi- kinematics model, which may have contributed to the aero-
cantly change the flow behavior around the wing and conse- dynamic power efficiency of the wing. The aerodynamic ad-
quently have an important effect on its aerodynamic perfor- vantage of the passive wing flexibility was also reported for
mance. A few studies have been devoted to the the biomimetic wings that are designed to produce character-
understanding of the effect of the wing flexibility. Using a istic deformation patterns of insects or birds.7
two-link model representing a chordwise wing section, In insect flight, both the inertial force and the fluid force
Vanella et al.4 performed a two-dimensional numerical simu- can be the primary causes of the wing deformation. Combes
lation of the flow-structure interaction in hovering flight. and Daniel8 compared vibrations of the excised hawkmoth,
They found that the structural flexibility can enhance the Manduca sexta, wing in normal air and in helium 共approxi-
aerodynamic performance of the wing by increasing the lift- mately 15% of the air density兲 and noticed that the difference
to-drag and lift-to-power ratios and that the best performance in the wing deformation pattern between the two cases is
is obtained when the flapping frequency is a fraction of the very small. Their result suggests that the Manduca wing de-
natural frequency of the wing structure. In a separate study formation is mainly due to the wing inertia. In another study,
using also the linkage model, Eldredge et al.5 investigated Chen et al.9 performed a vibration test of the dragonfly wing
the effect of the wing flexibility in a range of kinematic and found that the lowest frequency among the natural vibra-
parameters describing the combined pitching and heaving tion modes is on order of 170 Hz, which is much higher than
motion of the wing. They found that a mildly flexible wing the flapping frequency of the insect. Therefore, they con-
has consistently good performance over a wide range of cluded that the inertial force is small compared to the elastic
force for the dragonfly and the wing deformation is mainly
a兲
Author to whom correspondence should be addressed. Electronic mail: due to the aerodynamic force.
haoxiang.luo@vanderbilt.edu. From an aerodynamic point of view, whether the wing

1070-6631/2010/22共11兲/111902/10/$30.00 22, 111902-1 © 2010 American Institute of Physics

Downloaded 07 Jan 2011 to 129.59.78.81. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions
111902-2 B. Yin and H. Luo Phys. Fluids 22, 111902 共2010兲

(a) (b) A0
x0 (t) x0共t兲 = cos共2␲ ft兲, 共1兲
2
.
α (t) M
y n
dl ␣共t兲 = ␣0 + ␤ sin共2␲ ft + ␾兲, 共2兲
l t M where x0共t兲 is the horizontal position of the leading edge,
x q
n
τ ␣共t兲 is the angle between the leading edge and the horizontal
t
axis 共measured in the counterclockwise direction兲, A0 is the
FIG. 1. 共a兲 A schematic of the wing section during hovering flight. 共b兲 A stroke distance of the leading edge, ␣0 is the initial orienta-
wing segment illustrating the in-plane tension ␶, the transverse stress q, and tion, ␤ is the angle amplitude, f is the flapping frequency,
the bending moment M.
and ␾ is the phase difference between the rotation and trans-
lation. In the present work, we choose ␣0 = −␲ / 2 and ␾ = 0,
which corresponds to the symmetrical rotation.15 The wing is
deformation is caused by the wing inertia or the fluid force assumed to be elastic and nearly inextensible, and its dynam-
will not only affect the deformation pattern but also change ics is governed by the nonlinear equation
the phase of the passive wing deflection with respect to the
d 2x ⳵
wing actuation during a flapping cycle. It is therefore worth- ␳ sh = 共␶t + qn兲 + f, 共3兲
while to investigate the aerodynamic consequences of this dt2 ⳵ l
issue. Since both situations exist in insect flight, as seen in where ␳s and h are density and thickness of the wing, respec-
the hawkmoth and dragonfly, an investigation of this issue tively, t is the unit tangent vector pointing in the direction of
may provide some biomechanical insight into the morpho- increasing arc length, l, from the leading edge, n is the unit
logical differences between these insect wings. To address normal vector, and f is the difference between the distributed
the problem, we have performed a two-dimensional numeri- loads on the two sides of the wing. The in-plane tension, ␶, is
cal simulation of the flow-structure interaction in hovering assumed to be proportional to the tangential strain so that

冉冏 冏 冊 冉冏 冏 冊
flight and have systematically studied the effect of the mass
ratio of the wing, defined as mⴱ = ␳sh / 共␳ f c兲, where ␳sh is the ⳵x ⳵l
␶=E −1 =E −1 , 共4兲
surface density of the wing, ␳ f is the density of the air, and c ⳵ l0 ⳵ l0
is the characteristic length of the wing. The deformable wing
where E is the stretching coefficient of the plate and l0 is the
section is represented by an elastic plate, which may undergo
arc length in the unstretched state. The transverse stress, q, is
large displacements. Therefore, the wing model in the
linearly related to the bending moment, M, by
present study has infinite degrees of freedom and can have a
smooth camber, as opposed to the previous linkage ⳵ M ⳵ 共EB␬兲
models.4,5 q= = , 共5兲
⳵l ⳵l
Note that experimental measurements of the aerody-
namic force in insect flight are typically carried out in water where EB is the bending modulus and ␬ is the curvature.17
or oil in order to scale up the size of the wing model while The boundary conditions at l = 0 include the specified posi-
keeping the Reynolds number in its physiological tion and orientation, i.e.,
regime.10–12 In those cases, the hydrodynamic pressure is ⳵x
much higher than the inertial force of the wing. Therefore, x = x0共t兲, = 共cos ␣,sin ␣兲. 共6兲
the effect of the wing inertia could not be addressed in those ⳵l
studies. We also point out that the effect of the wing inertia At the trailing edge, l = c, both the bending moment and the
was a topic in a few previous theoretical studies.13,14 How- transverse stress vanish, which requires ␬ = 0 and ⳵␬ / ⳵l = 0.
ever, these works mainly focused on the thrust generation, Therefore, we have
not lift production, of the flexible wings. Furthermore, in
Zhu,13 a potential flow was assumed and the viscous effects ⳵ 2x ⳵ 3x
= 0, = 0. 共7兲
including flow separation were omitted; in Michelin and ⳵ l2 ⳵ l3
Smith,14 point vortices were introduced into their inviscid
The flow is governed by the viscous incompressible Navier–
flow model to account for the vortex shedding, but the stroke
Stokes equation and the continuity equation,
distance of the wing was very small compared to the chord
length and there was no flow separation at the leading edge. ⳵ vi ⳵ v jvi 1 ⳵p ⳵ 2v i
+ =− + ␯f 2 ,
⳵t ⳵xj ␳ f ⳵ xi ⳵xj

II. METHODS ⳵ vi
= 0, 共8兲
A. Problem description and numerical approach ⳵ xi
We consider a two-dimensional hovering wing section where vi is the velocity, ␳ f and ␯ f are the fluid density and
with the chord length c, as shown in Fig. 1. The wing under- viscosity, and p is the pressure. No-slip and no-penetration
goes a combined translational and rotational motion specified conditions are specified at the flow-solid boundary. To pa-
at the leading edge,15,16 rametrize the system, we define the nondimensional groups

Downloaded 07 Jan 2011 to 129.59.78.81. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions
111902-3 Effect of wing inertia on hovering performance Phys. Fluids 22, 111902 共2010兲

(a) (b) (c)


1
3 0.6

0.8

2 0.4

CL /CD
0.6

CD
CL

0.4
1 0.2
0.2

0 0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
ωf /ωn ωf /ωn ωf /ωn

(d) (e) (f)


12
4 0.6
10
3
8
0.4

CL /CP
CP
CP

2 6

4 0.2
1
2

0 0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
ωf /ωn ωf /ωn ωf /ωn

FIG. 2. 共a兲 Lift, 共b兲 drag, 共c兲 lift-to-drag, 共d兲 net power, 共e兲 modified power, and 共f兲 lift-to-modified-power coefficients of the flexible wing without active
rotation and mⴱ = 1 共solid line兲, 5 共dashed line兲, and 25 共dash-dotted line兲.

including the normalized wing stroke, Reynolds number, structure interaction. The flow-structure interaction is solved
mass ratio, and frequency ratio, which are given by at each time step by iterating the flow and the structural
dynamics until convergence is reached.21

A0 ␲A0 fc ␳ sh ␻ f 2␲ f B. Simulation setups


, Re = , mⴱ = , ␻ⴱ = = , 共9兲
c ␯f ␳fc ␻n ␻n In the present simulations, we choose the stroke distance
A0 / c = 2.5, Reynolds number Re= 150, and rotational angle
␤ = 0 or ␤ = ␲ / 8. The parameters describing the wing kine-
respectively, where ␻ f = 2␲ f is the flapping frequency and
␻n = 共k2n / c2兲冑EB / ␳sh with kn = 1.8751 is the frequency of the
matics are selected based on previous work on insect
flight.4,15,16 Three mass ratios are considered, mⴱ = 1, 5, and
first natural vibration mode of the wing.18 Physically, mⴱ
represents the ratio between the inertial force of the wing and 25, which represent a light, a medium, and a heavy wing,
the aerodynamic pressure, and ␻ⴱ represents the wing rigid- respectively. For each of these mass ratios, the frequency
ity. ratio ␻ⴱ is chosen to vary among 1/1.25, 1/1.5, 1/2, 1/2.5,
The equations governing the system, Eqs. 共3兲 and 共8兲, 1/3, 1/4, 1/6, and 0, where ␻ⴱ = 1 / 1.25 means that the wing is
are solved numerically in an implicitly coupled manner using most flexible and flaps near the resonant frequency, and ␻ⴱ
an in-house solver. Specifically, the incompressible flow is = 0 means that the wing is rigid.
solved using a sharp-interface immersed-boundary The computational domain has a size of 20c ⫻ 35c. We
method19,20 with a special treatment to suppress the pressure have done extensive tests to make sure that the domain is
oscillations associated with the moving boundaries.21 In this large enough to achieve satisfactory accuracy of the results.
method, a single-block Cartesian grid is used to discretize The entire domain consists of a nonuniform Cartesian grid of
the Navier–Stokes equation on a rectangular domain, and the 320⫻ 448 points. The grid contains a horizontal band and a
ghost nodes and hybrid nodes are defined near the fluid-solid vertical band of width 3c in which the grid points are uni-
interface to facilitate the boundary treatment at the interface.
formly and densely distributed such that the grid spacings
The infinitely thin membranous wing is augmented with an
⌬x = ⌬y = 0.02c. A total number of 100 nodes are used to
artificial thickness that is about three times of spacing of the
Cartesian grid and is automatically reduced as the grid is discretize the wing and its governing equation. The time step
refined. The wing section is discretized by a set of Lagrang- size is ⌬t = 0.0025T, where T = 1 / f is the period of a flapping
ian points initially distributed uniformly along the wing. A cycle. The flow solver and the structure solver have been
standard central finite difference scheme is used to discretize validated separately as shown in Appendixes A and B. In
Eqs. 共3兲–共7兲, and Eq. 共3兲 is solved iteratively as an inner loop addition, grid refinement has been performed to make sure
embedded within the implicit algorithm for the flow- that the simulation results are grid-independent.

Downloaded 07 Jan 2011 to 129.59.78.81. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions
111902-4 B. Yin and H. Luo Phys. Fluids 22, 111902 共2010兲

III. RESULTS AND DISCUSSION (a)

Displacement
0.5
A. Flexible wings without rotation 0
-0.5
We first consider the flexible wing driven by the pure -1
translation, i.e., ␤ = 0, at the leading edge where the wing is 12 12.5 13 13.5
t/T
14 14.5 15

clamped. Therefore, the wing has to rely on its deflection to (b)


4
generate a nonzero lift. Figure 2共a兲 shows the lift coefficient,
CL = 2FL / 共␳ f U2c兲, averaged from 15 flapping cycles, where 2

CL
FL is the total lift and U = ␲A0 f is the maximum translational 0

velocity of the leading edge. The result is shown for the mass 12 12.5 13 13.5 14 14.5 15
ratio mⴱ = 1, 5, and 25 and frequency ratio ␻ⴱ from 0 to 0.8.
t/T
(c)
Two interesting phenomena can be observed from this figure. 8
First, for each mass ratio, there is a particular frequency ratio

CD
4
at which the lift force peaks. The peak lift coefficients for the 0
three mass ratios are very close to each other, and all are 12 12.5 13 13.5 14 14.5 15
t/T
around 0.9. Second, this particular frequency ratio depends
(d)
on the mass ratio of the wing. For mⴱ = 25, CL peaks at ␻ⴱ 30
= 0.8. For mⴱ = 5 and 1, the peaks of CL are shifted to the left 15

CP
and take place at ␻ⴱ = 0.5 and 0.4, respectively. That is, as the 0
-15
mass ratio is reduced, the frequency ratio for the peak lift -30
12 12.5 13 13.5 14 14.5 15
also decreases. Physically, this result means that for the t/T

heavy wing 共large mⴱ兲, flapping near the resonant frequency FIG. 3. Histories of 共a兲 the tail displacement in the x-direction, 共b兲 CL, 共c兲
produces higher lift, and for the light wing 共small mⴱ兲, CD, and 共d兲 C P for ␻ⴱ = 0.5 and mⴱ = 1 共solid line兲, 5 共dashed line兲, and 25
flapping at a frequency much lower than the resonant fre- 共dash-dotted line兲.
quency would produce higher lift. To understand this result,
we point out that when the mass ratio is large, the fluid force
becomes insignificant compared to the wing’s inertial force, The dependence of the lift-to-drag ratio on mⴱ and ␻ⴱ is
and the flapping actuation has to be close to the natural vi- shown in Fig. 2共c兲. It can be seen that the low-mass wings
bration mode in order to produce significant wing deforma- clearly outperform the high-mass wings when the wing rigid-
tions for lift production. On the other hand, when the mass ity is the same. For mⴱ = 25, the lift-to-drag ratio increases
ratio is low, the wing is deformed by the fluid force, and a nearly monotonically as ␻ⴱ is raised and the wing becomes
lower flapping frequency would suffice for the necessary de- more flexible; for mⴱ = 5, this ratio increases first and then
formation. If the flapping frequency is too high in the case of reaches a plateau of 0.31, and for mⴱ = 1, the lift-to-drag ratio
low mⴱ, the wing deformation may become exceedingly first increases, then reaches a peak of 0.56 around ␻ⴱ = 0.6,
large, and the lift would drop, as shown in Fig. 2共a兲 for large and finally drops as ␻ⴱ is further raised.
␻ ⴱ. To analyze the power consumption, we first define the
Figure 2共b兲 shows the drag coefficient defined as CD net power P共t兲 as
= 2FD / 共␳ f U2c兲, averaged over the 15 flapping cycles. Here, P共t兲 = Fxẋ0 + M z␣˙ , 共10兲
FD is the total horizontal fluid force on the wing defined to
be positive when it is against the translational motion of the where Fx and M z are the total force and total torque, respec-
leading edge. This force can be temporarily negative in a tively, applied at the leading edge to actuate the wing. Note
cycle due to the wing deformation, as seen later. For mⴱ that the second term is zero when the wing has no active
= 25 and 5, the average CD changes approximately within rotation. Next, we adopt a conservative assumption that the
20% as ␻ⴱ is varied. For mⴱ = 1, the average CD decreases negative power in the either one of the two terms in Eq. 共10兲
drastically from 2.76 to 0.69 as ␻ⴱ is increased from 0 to 0.8, is not reusable. Therefore, we define an alternative power
and the drag is generally much lower than the other two measurement, P̂, which represents only the positive contri-
wings with the same rigidity. Such a drag reduction is be- butions from the two terms in Eq. 共10兲. The averaged net and
cause the light wing is deflected by the fluid force and its modified power coefficients, C P and C Pˆ, defined as the
shape is adapted to the drag by reducing the frontal area. power normalized by 共1 / 2兲␳ f U3c, are plotted in Figs. 2共d兲
When ␻ⴱ is large and the wing is more flexible, this self- and 2共e兲, respectively. For all three mass ratios, the two
adaptation effect becomes more pronounced. On the other power measurements overall exhibit a similar trend. For mⴱ
hand, when ␻ⴱ goes to zero, that is, the wing behaves essen- = 25 and 5, the net power coefficient increases up to 4.2 and
tially as a rigid plate, the lift coefficient vanishes and the 3.0, respectively, as ␻ⴱ is raised, which indicates a significant
drag coefficient approaches to the same constant for all the energy loss due to the large wing deformation. The instanta-
mass ratios as expected. Note that in the cases of mⴱ = 25 and neous power consumption, as plotted in Fig. 3共d兲 for ␻ⴱ
5, the average drag of a flexible wing 共␻ⴱ ⬎ 0兲 can be higher = 0.5, shows that at these two mass ratios, the net power
than that of a corresponding rigid wing 共␻ⴱ = 0兲, which indi- fluctuates at a large amplitude in a flapping cycle because the
cates that the inertia-dominated deformation may substan- energy changes its form between the kinetic energy of the
tially augment the drag force. wing and the elastic potential stored in the wing. Therefore,

Downloaded 07 Jan 2011 to 129.59.78.81. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions
111902-5 Effect of wing inertia on hovering performance Phys. Fluids 22, 111902 共2010兲

if the negative power is not reusable as assumed for the movement. In either case, the drag is reduced considerably.
modified power, a large portion of the input energy would be The corresponding power coefficients for the three mass ra-
lost. This effect is reflected in Fig. 2共e兲, where the mⴱ = 25 tios are plotted in Fig. 3共d兲. Note that since the rotational
wing requires a much higher modified power than the other term in Eq. 共10兲 vanishes for ␤ = 0, C Pˆ is exactly equal to the
two wings. Especially when the wing rigidity is low, the positive portion of C P in this case. Comparing the three mass
modified power coefficient reaches an amount of 12 for mⴱ ratios, we find that the power coefficient has least fluctua-
= 25. On the other hand, for mⴱ = 1, the modified power co- tions for mⴱ = 1. In addition, the power peaks have more time
efficient is close to the net power coefficient, and both in delay when mⴱ is lower. For mⴱ = 5 and 25, the power reaches
general decrease as the wing becomes more flexible. As seen its maximum magnitude near the stroke reversals due to the
in Fig. 3共d兲 for ␻ⴱ = 0.5, the instantaneous power coefficient inertial acceleration or deceleration of the wing.
at this mass ratio is almost always positive, and its magni- The instantaneous vorticity field in an entire flapping
tude is much smaller than that for mⴱ = 5 and 25. Figures 2共d兲 cycle and the corresponding wing configuration are shown in
and 2共e兲 show that the average power consumption for mⴱ Fig. 4 for mⴱ = 5 and 1 at ␻ⴱ = 0.5. There are several similari-
= 1 is lowest among the three mass ratios. This result is un- ties in the vortex behavior between the two cases. For ex-
derstandable since the wing with mⴱ = 1 experiences a ample, a leading edge vortex 共LEV兲 is generated during each
smaller drag resistance compared to the other two wings. The half-stroke and is then recaptured by the wing during its
power economy of the low-mass wing is further seen in Fig. return trip after the stroke reversal 共e.g., the positive vortex
2共f兲, where the ratio between the average lift coefficient and blob at t / T = 13.0 and the negative blob at t / T = 13.5兲. The
the average modified power coefficient is shown. For the LEV moves downward along the cambered wing and may
same amount of power input, the wing with mⴱ = 1 may pro- merge with the trailing edge vortex 共TEV兲22 of the same sign
duce more than twice amount of lift than the other two that is being formed 共e.g., the positive blob at t / T = 13.1 and
wings. The peak performance for mⴱ = 1 is around 0.58, the negative blob at t / T = 13.6兲. The merged vortex is
which takes place at ␻ⴱ = 0.5. strengthened and meanwhile stretched by the trailing edge as
We select a specific frequency ratio, ␻ⴱ = 0.5, to analyze shown by the positive vortex band at t / T = 13.3 and also by
the details of the force characteristics and flow field. From the negative band at t / T = 13.8. The gradually thinned trailing
Fig. 2, the wing flexibility at this frequency ratio has nearly edge vortex eventually pinches off in the middle 共e.g., the
the best lift-to-modified-power performance for all three filament between two positive blobs at t / T = 13.5兲. The por-
mass ratios. Figure 3共a兲 shows the x-component of the dis- tion that attaches to the trailing edge, now termed the end-
placement of the wing tail with respect to its undeformed of-stroke vortex 共ESV兲,4 has been evolving during the stroke
configuration for mⴱ = 25, 5, and 1. At this frequency ratio, reversal while the wing is restoring its shape and then de-
the maximum displacement is more than 50% of the chord forming in the other way 共e.g., t / T = 13.3 to 13.6兲. After the
length in all three cases, while the wing with mⴱ = 1 has larg- TEV breaks off from the ESV, it travels downward in the
est amplitude. The positions of the positive and negative wake, while the ESV later also detaches from the trailing
peaks in the displacement indicate that there is a significant edge but may temporarily move upward before disappearing
phase difference among the three cases. For mⴱ = 25, the or merging into the downwash. The wake below the hovering
maximum displacement almost always takes place at the wing is marked by a pair of TEVs with opposite signs that
stroke reversals, e.g., t / T = 12.0, 12.5, 13.0, and so on. For are generated by the two half-strokes in a complete cycle.
mⴱ = 1, there is a phase delay of approximately ␲ / 4 in the The differences in the flow field between mⴱ = 5 and 1
maximum displacement. In addition, the wing displacement are also evident. First, the size of the LEV for mⴱ = 5 is gen-
in this case becomes highly asymmetric between the forward erally larger than that for mⴱ = 1. Second, the stretched trail-
and backward strokes, as indicated by the appearance of two ing edge vortex for mⴱ = 5 is aligned more in the horizontal
peaks in the first half-stroke. This interesting deformation direction, while it is more in the vertical direction for mⴱ
pattern will be discussed later together with the flow field. = 1, as shown in the frames from t / T = 13.4– 13.7. These flow
Figures 3共b兲 and 3共c兲 show the corresponding lift and features are consistent to the higher drag formation for mⴱ
drag coefficients to the wing displacement shown in Fig. = 5. Furthermore, a substantial portion of the flow in the case
3共a兲. It can be seen that the two heavy wings, especially of mⴱ = 5 travels in the horizontal direction or even in an
mⴱ = 25, produce a large amount of negative lift every time upward direction, while in the case of mⴱ = 1, the flow mainly
after the wing passes the middle point of the stroke, at which travels downward, leading to superior energy efficiency of
the wing has recovered to from its deformation and is over- this wing. For mⴱ = 1, the unsteady vortices may cause ape-
shooting and bending forward. In comparison, the light wing riodic vibration of the wing. This phenomenon is illustrated
with mⴱ = 1 has typically a non-negative lift coefficient in the frames t / T = 13.2– 13.4, where a blob of negative vor-
throughout the multiple flapping cycles. In the drag history, tex passes underneath the wing, causing the wing to deflect
the two heavy wings cause much higher drag than the light for the second time in the same half-stroke.
wing, and the peak drag takes place when the wing has the The distinct vortices have been indicated in Fig. 4. To
maximum translation and is nearly in a vertical position. For quantify the strength of these vortices, we first visualize the
mⴱ = 1, the peak drag happens when the wing has the maxi- vorticity field using contour lines. After each vortex is manu-
mum translational velocity and the least frontal area or when ally identified, a closed contour line is generated around this
the wing is restoring its shape prior to the stroke reversal, at vortex with the specified level, and then the circulation ⌫ is
which point the wing has slowed down its translational computed along this line. Though the magnitude of the cir-

Downloaded 07 Jan 2011 to 129.59.78.81. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions
111902-6 B. Yin and H. Luo Phys. Fluids 22, 111902 共2010兲

FIG. 4. 共Color online兲 A series of instantaneous vortex


field in a flapping cycle for ␻ⴱ = 0.5, mⴱ = 5 共first and
third columns兲, and mⴱ = 1 共second and fourth columns兲.
The contour level ranges from −3.2U / c to 3.2U / c.

culation depends on the chosen contour level, the character- two mass ratios. It should be pointed out that the difference
istic behavior of the vortex is not affected by this choice. The between the LEVs and between TEVs shown in Fig. 5 is
computed circulations of the LEV, TEV, and ESV are shown fairly consistent for different wing strokes, but the difference
in Fig. 5 for the vortices indicated in Fig. 4. It can be seen between the ESVs is not. The appearance and strength of the
that the LEV for mⴱ = 5 is much stronger than the correspond- ESV vary from stroke to stroke, and they depend on the
ing vortex for mⴱ = 1 over a significantly long period of time. evolution of the LEV. For example, in Fig. 4 for mⴱ = 5 from
The TEV is initially stronger for mⴱ = 5, but after breaking t / T = 13.0 to 13.1, the LEV moves along the wing to the
up, it has a similar strength for both mass ratios. Further- trailing edge and suppressed formation of the ESV. Conse-
more, the timing of the LEV and the TEV is similar for the quently, the ESV for mⴱ = 1 is stronger during that stroke
reversal 共this ESV is visible in Fig. 4 but is too weak to show
up in Fig. 5兲.
Generally speaking, the strength and timing of the vor-
6 tices have important consequences on the force production of
TEV
4 flapping wings. In the present case, the three types of vorti-
LEV ESV ces have both positive and negative effects on the wing per-
2
Γ/(Uc)

formance. For example, during the translational stage, the


0 LEV and the TEV cause the flow to circulate around the
-2 leading edge and the trailing edge, reducing the pressure dif-
LEV ference between the two sides of the wing and thus lowering
-4 TEV
the lift. However, the shed vortices provide lift augmentation
-6 through the wake-capturing mechanism. This effect can be
12.5 13 13.5 14
t/T seen from Fig. 4 at t / T = 13.6, where the LEV and ESV cre-
ate a flow directed against the wing and thus the lift is en-
FIG. 5. Histories of the circulation for the vortices in the flow at ␤ = 0, ␻ⴱ hanced. The lift enhancement at the moment is seen from
= 0.5, and mⴱ = 1 共solid lines兲 and 5 共dashed lines兲; the vortex type and the
specified contour level are LEV with ⫾3.95U / c 共thick lines兲, ESV with Fig. 3共b兲, where the mⴱ = 5 wing has significantly higher lift
⫾1.9U / c 共squares兲, and TEV with ⫾3.95U / c 共thin lines兲. than the mⴱ = 1 wing at t / T = 13.6.

Downloaded 07 Jan 2011 to 129.59.78.81. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions
111902-7 Effect of wing inertia on hovering performance Phys. Fluids 22, 111902 共2010兲

(a) (b) (c)


4 0.8
1.2

1 3 0.6

0.8

CL /CD
CD
CL

2 0.4
0.6

0.4
1 0.2
0.2

0 0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
ωf /ωn ωf /ωn ωf /ωn

(d) (e) (f)


4 4 0.8

3 3 0.6

CL /CP
2 CP2 0.4
CP

1 1 0.2

0 0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
ωf /ωn ωf /ωn ωf /ωn

FIG. 6. 共a兲 Lift, 共b兲 drag, 共c兲 lift-to-drag, 共d兲 net power, 共e兲 modified power, and 共f兲 lift-to-modified-power coefficients of the flexible wing with active rotation
␤ = ␲ / 8 for mⴱ = 1 共thick solid line兲 and 5 共thick dashed line兲. The corresponding data for the nonrotational wing are shown as thin lines. The dotted lines in
共a兲–共c兲 are for the corresponding rigid wing with a rotation angle of ␤ = ␲ / 4.

B. Flexible translating wings with rotation value for the wings with active rotation. For both mⴱ = 5 and
Next, we consider the situation where the wings translate 1, this value is around ␻ⴱ = 0.4. As a reference, we provide
and rotate actively around the leading edge with an ampli- the lift coefficient, the drag coefficient, and the lift-to-drag
tude of ␤ = ␲ / 8. Only mⴱ = 5 and 1 are considered here. Fig- ratio of a rigid wing performing the same translation at the
ure 6共a兲 shows the averaged lift coefficient CL of the wing. leading edge but rotates with an amplitude of ␤ = ␲ / 4. The
The corresponding nonrotational cases are also included in Reynolds number is also Re= 150 for this rigid wing. The
the figure for comparison. It can be seen that for both mⴱ results are shown in Figs. 6共a兲–6共c兲 as the dotted lines. Com-
= 5 and 1, adding a moderate amount of active rotation sig- paring this rigid wing with the flexible wings that has less
nificantly increases the lift force for a range of ␻ⴱ. When ␻ⴱ active rotation, we notice that the flexible wings may gener-
is large and the wing is very flexible, the active rotation ate higher lift when they have a proper rigidity, but the ref-
reduces the lift instead due to the upward swinging motion of erence wing has considerable lower drag. In terms of the
the wing tail. The frequency ratio of the rotational wing at lift-to-drag ratio, only the light wing with mⴱ = 1 can outper-
which the lift peaks is lower compared to that of the corre- form the reference wing by a small amount 共around 6%兲.
sponding nonrotational wing. Therefore, when the wing is This result suggests that there may be an optimal combina-
actively rotating, less structural flexibility is needed for lift tion of the active wing rotation and the passive wing defor-
enhancement. Furthermore, like the nonrotational wings, the mation.
lift curves of the rotational wings also have a peak value The active rotation has a similar effect on the power
whose amplitude is insensitive to the wing mass ratio but efficiency of the flexible wings. Figures 6共d兲–6共f兲 show the
whose corresponding frequency ratio exhibits a left-shift net power coefficient, the modified power coefficient, and
trend when the mass ratio is reduced. the lift-to-modified-power coefficient, respectively. Overall,
The average drag coefficient is shown in Fig. 6共b兲. The introducing active rotation reduces the power requirement
active rotation reduces the drag for both mⴱ = 5 and 1 for all and significantly improves the power efficiency. The fre-
the frequency ratios. In addition, the effect of the frequency quency ratio for the optimal power performance is down-
ratio on the drag coefficient for the rotational wings is simi- shifted to 0.35 for the wings with active rotation, regardless
lar to that for the corresponding nonrotational wings. That is, of the mass ratio. Finally, we point out that in both the lift-
for mⴱ = 5, the drag coefficient only changes slightly when to-drag and lift-to-modified-power measurements, the wing
the wing becomes more flexible, while for mⴱ = 1, the drag with mⴱ = 1 significantly outperforms the wing with mⴱ = 5 by
coefficient reduces significantly. The lift-to-drag ratio in Fig. 34% for CL / CD and 71% for CL / C Pˆ.
6共c兲 shows that the combination of the active rotation and Figure 7 shows the x-displacement, lift, drag, and power
passive wing deformation improves the wing performance histories of the rotational wing at ␻ⴱ = 0.4. There is again a
and the improvement is by 66.7% for mⴱ = 5 and 38.7% for phase delay in the displacement of the wing with mⴱ = 1 com-
mⴱ = 1. Compared to the wings without rotation, the optimal pared to that for mⴱ = 5. Although the light wing does not
frequency ratio for the lift-to-drag ratio is shifted to a lower have an obvious double-peak displacement within a single

Downloaded 07 Jan 2011 to 129.59.78.81. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions
111902-8 B. Yin and H. Luo Phys. Fluids 22, 111902 共2010兲

(a) half-stroke as shown earlier for the corresponding nonrota-


Displacement
0.5 tional wing, the deformation has nearly a plateau after the
0
wing passes the midstroke point. The peak lift for mⴱ = 1
-0.5
takes place at the midstroke point where the wing has the
12 12.5 13 13.5 14 14.5 15
t/T maximum translation. For mⴱ = 5, the peak lift point during
(b) second half-stroke also takes place at the midstroke point,
4
but it is brought earlier during the first half-stroke. The drag
2
histories are in phase with each other for the two wings, but
CL

0 the mⴱ = 5 wing clearly produces a higher peak drag when the


12 12.5 13 13.5 14 14.5 15 wing passes the midstroke point. The histories of the net and
t/T

(c) modified power coefficients, plotted in Figs. 7共d兲 and 7共e兲,


8
show that the power input has large fluctuations for mⴱ = 5,
4 especially before and after the stroke reversals when the
CD

0 wing experiences the maximum inertial deceleration or ac-


12 12.5 13 13.5 14 14.5 15 celeration. On the other hand, the power input for mⴱ = 1 is
t/T

(d)
nearly always positive and has a much lower amplitude of
10 fluctuation.
5 The flow field is shown in Fig. 8 for mⴱ = 5 and 1 at
CP


0 ␻ = 0.4. Comparing this figure with Fig. 4, it can be seen
-5
12 12.5 13 13.5 14 14.5 15
that the LEV now becomes more attached to the rear side of
t/T
the wing due to the active rotation and an on-average re-
(e)
10 duced angle of attack at the leading edge. Compared the flow
fields between mⴱ = 5 and 1, we see again that the LEV of the
CP

5
light wing is smaller in size than that of the heavy wing. In
0 addition, the vortices have less upward movement in the case
of mⴱ = 1, and the vortex pairs in the wake are more evenly
12 12.5 13 13.5 14 14.5 15
t/T

spaced compared to mⴱ = 5. Using the same approach de-


FIG. 7. Histories of 共a兲 the tail displacement in the x-direction, 共b兲 CL, 共c兲
CD, 共d兲 C P, and 共e兲 C Pˆ for ␻ⴱ = 0.4, mⴱ = 1 共solid line兲 and 5 共dashed line兲, scribed earlier, we compute the instantaneous circulations for
and the rotational angle ␤ = ␲ / 8.

FIG. 8. 共Color online兲 A series of instantaneous vortex


field in a flapping cycle for ␤ = ␲ / 8, ␻ⴱ = 0.4, mⴱ = 5
共first and third columns兲, and mⴱ = 1 共second and fourth
columns兲. The contour level ranges from −3.2U / c to
3.2U / c.

Downloaded 07 Jan 2011 to 129.59.78.81. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions
111902-9 Effect of wing inertia on hovering performance Phys. Fluids 22, 111902 共2010兲

6 Their finding is thus consistent to our result that for the low-
4 mass wing to have best performance, the flapping frequency
LEV TEV
should be much lower than the resonant frequency. Our
2 simulation further suggests that the dragonfly could have
Γ/(Uc) ESV
0 taken advantage of the low-mass ratio of its wings for effi-
cient lift production. We should point out that a direct com-
-2
LEV parison of the aerodynamic performance between the drag-
-4 TEV onfly wing and the hawkmoth wing is not possible through
-6 the present work since the three-dimensional effect, which is
12.5 13 13.5 14 an important factor in insect flight, is not considered here.
t/T
Finally, we should note that the mass ratio of an insect
FIG. 9. Histories of the circulation for the vortices in the flow at ␤ = ␲ / 8, wing cannot be reduced arbitrarily, even though a low mass
␻ⴱ = 0.4, and mⴱ = 1 共solid lines兲 and 5 共dashed lines兲; the vortex type and the ratio may improve the wing performance aerodynamically.
specified contour level are LEV with ⫾3.95U / c 共thick lines兲, ESV with This is because the wing needs to maintain at least a minimal
⫾1.9U / c 共squares兲, and TEV with ⫾3.95U / c 共thin lines兲.
thickness to achieve the necessary rigidity and its physiologi-
cal functions. In some insect wings, certain structural fea-
the LEV, TEV, and ESV identified from the vorticity con- tures may help reduce the wing mass while retaining the
tours shown in Fig. 8, and the result is plotted in Fig. 9. Like necessary stiffness, e.g., corrugations of the dragonfly
the nonrotational wing, the LEV and the TEV are again con- wing.23
sistently much stronger for mⴱ = 5 than for mⴱ = 1, while the
difference between the ESVs of the two wings may vary for IV. CONCLUSION
each stroke reversal. Among these three vortices, the LEV In this paper, the fluid-structure interaction of a two-
and the ESV have a significant contribution to the lift during dimensional hovering wing is numerically simulated in order
the wake capture. Therefore, immediately after the right to investigate the effect of the wing inertia on the wing de-
stroke reversals 共e.g., t / T = 13.1兲, the wing with mⴱ = 5 has formation and on the aerodynamic performance. The wing is
both stronger LEV and ESV and thus a much higher lift parametrized by a nondimensional mass ratio and a fre-
coefficient than the wing with mⴱ = 1, while immediately af- quency ratio representing the wing flexibility. The mass ratio
ter the left stroke reversals, the wing with mⴱ = 5 has a stron- is taken from the physiological data of insects. The simula-
ger LEV but a weaker ESV 共e.g., t / T = 13.6兲; thus the wing tion shows that when the amount of deformation is about the
has only a moderately higher lift coefficient than the wing same, the low-mass wing has consistently better performance
with mⴱ = 1, as shown in Fig. 7. than the high-mass wing in term of the lift-to-drag ratio and
power efficiency. Therefore, the present result suggests that
C. Discussion
the fluid force dominated wing flexibility has aerodynamic
According to the insect data in Ref. 8, a hawkmoth advantages over the inertial force dominated flexibility. Fu-
Manduca wing has a chord length of the order of 10 mm and ture work will focus on the three-dimensional effect of the
an approximate thickness of h = 45 ␮m. Using the wing den- wing flexibility.
sity ␳s = 1200 kg/ m3 and the air density ␳ f = 1.2 kg/ m3, we
obtain the mass ratio mⴱ = ␳sh / 共␳ f c兲 = 4.5. Chen et al.9 mea- ACKNOWLEDGMENTS
sured the natural frequency of the dragonfly wing 共Orth-
We gratefully thank Professor J. Eldredge of UCLA for
etrum pruinosum and Orthetrum sabina兲. In their study, the
providing the simulation data that we used for our code vali-
wing has a dimension of 38 mm⫻ 8 mm, and the wing
dation. This work is partially supported by the NSF under
mass is 2.5 mg. Based on these data, the mass per unit wing
Grant No. CBET-0954381.
area is around ␳sh = 8.2 mg/ mm2, and the mass ratio is mⴱ
= 0.85. Therefore, the hawkmoth and dragonfly wings corre-
APPENDIX A: VALIDATION OF FLOW SOLVER
spond roughly to the mⴱ = 5 and 1 cases, respectively, in the
present simulations, and the wing inertia seems to play dif- The simulation of the unsteady flow around a thin and
ferent roles in the structural deformation of these two insects. rigid wing is compared with previous result in Ref. 16. In
Chen et al.9 found that the flapping frequency of the dragon- this test, the wing rotates around its center, whose stroke
fly is only about 16% of the natural frequency of the wing. distance is denoted by A0. The parameters are otherwise de-

2 2

1
1
CD
CL

0
0
-1

-1 -2
0 0.5 1 1.5 2 2.5 3 3.5 4 0 0.5 1 1.5 2 2.5 3 3.5 4
t/T t/T

FIG. 10. Lift and drag coefficients 共defined in the same way as in Sec. III兲 from the present simulation 共solid line兲 and from Ref. 16 共dashed line兲.

Downloaded 07 Jan 2011 to 129.59.78.81. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions
111902-10 B. Yin and H. Luo Phys. Fluids 22, 111902 共2010兲

(a) (b) 4
Ser. B 358, 1577 共2003兲.
0
M. Vanella, T. Fitzgerald, S. Preidikman, E. Balaras, and B. Balachandran,
40 nodes
“Influence of flexibility on the aerodynamic performance of a hovering
0.005
60 nodes wing,” J. Exp. Biol. 212, 95 共2009兲.
−0.005 5
100 nodes J. D. Eldredge, J. Toomey, and A. Medina, “On the roles of chord-wise
Analytical solution
0 flexibility in a flapping wing with hovering kinematics,” J. Fluid Mech.
−0.01 20 nodes
659, 94 共2010兲.
Analytical solution 6
J. Young, S. M. Walker, R. J. Bomphrey, G. K. Taylor, and A. L. R.
−0.005
−0.015
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Thomas, “Details of insect wing design and deformation enhance aerody-
l/c l/c namic function and flight efficiency,” Science 325, 1549 共2009兲.
7
S. K. Chimakurthi, J. Tang, R. Palacios, C. Cesnik, and W. Shyy, “Com-
FIG. 11. Comparison of the simulated vibration mode with the eigenmode. putational aeroelasticity framework for analyzing flapping wing micro air
共a兲 The first mode. 共b兲 The second mode. vehicles,” AIAA J. 47, 1865 共2009兲.
8
S. A. Combes and T. L. Daniel, “Into thin air: contributions of aerody-
namic and inertial-elastic forces to wing bending in the hawkmoth
fined in the same way as in Sec. II. To match the simulation Manduca sexta,” J. Exp. Biol. 206, 2999 共2003兲.
setup, A0 / c = 2.8 and ␤ = ␲ / 4 are chosen for the wing kine-
9
J. S. Chen, J. Y. Chen, and Y. F. Chou, “On the natural frequencies and
matics, and the Reynolds number is Re= 75. The flow is mode shapes of dragonfly wings,” J. Sound Vib. 313, 643 共2008兲.
10
S. Heathcote, D. Martin, and I. Gursul, “Flexible flapping airfoil propul-
initially quiescent. The lift and drag coefficients for first a sion at zero freestream velocity,” AIAA J. 42, 2196 共2004兲.
few cycles are shown in Fig. 10. It can be seen that the 11
S. Heathcote and I. Gursul, “Flexible flapping airfoil propulsion at low
present simulation has very good agreement with that in Ref. Reynolds numbers,” AIAA J. 45, 1066 共2007兲.
12
16 where a vortex particle method was used. S. Heathcote, Z. Wang, and I. Gursul, “Effect of spanwise flexibility on
flapping wing propulsion,” J. Fluids Struct. 24, 183 共2008兲.
13
Q. Zhu, “Numerical simulation of a flapping foil with chordwise or span-
APPENDIX B: VALIDATION OF THE STRUCTURAL wise flexibility,” AIAA J. 45, 2448 共2007兲.
SOLVER 14
S. Michelin and S. G. L. Smith, “Resonance and propulsion performance
of a heaving flexible wing,” Phys. Fluids 21, 071902 共2009兲.
The structural solver is validated by comparing the nu- 15
Z. J. Wang, J. M. Birch, and M. H. Dickinson, “Unsteady forces and flows
merical simulation of the small-amplitude vibration in in low Reynolds number hovering flight: Two-dimensional computations
vacuum with the eigenmodes of the wing structure. To do vs robotic wing experiments,” J. Exp. Biol. 207, 449 共2004兲.
16
this, a sinusoidal translation is specified at the leading edge J. D. Eldredge, “Numerical simulation of the fluid dynamics of 2d rigid
body motion with the vortex particle method,” J. Comput. Phys. 221, 626
and the amplitude of translation is much smaller than c. The 共2007兲.
structural dynamics is simulated in the absence of the fluid 17
C. Pozrikidis, “Buckling and collapse of open and closed cylindrical
so that there is no damping mechanism. The frequency of the shells,” J. Eng. Math. 42, 157 共2002兲.
18
actuation is chosen to be either the first or second eigenfre- E. Volterra and E. C. Zachmanoglou, Dynamics of Vibrations 共CE Merrill,
Columbus, 1965兲.
quency of the corresponding cantilever beam.18 Figure 11 19
R. Mittal, H. Dong, M. Bozkurttas, F. M. Najjar, A. Vargas, and A. von-
shows the simulated vibration modes together with the ana- Loebbeck, “A versatile sharp interface immersed boundary method for
lytical eigenfunctions. For the first mode, 20 nodes on the incompressible flows with complex boundaries,” J. Comput. Phys. 227,
wing are sufficient to capture the deformation pattern accu- 4825 共2008兲.
20
rately. For the second mode, a 100-node mesh leads to a H. Luo, R. Mittal, X. Zheng, S. A. Bielamowicz, R. J. Walsh, and J. K.
Hahn, “An immersed-boundary method for flow–structure interaction in
satisfactory solution. In the flow-structure simulations pre- biological systems with application to phonation,” J. Comput. Phys. 227,
sented here, 100 nodes are used in all cases. 9303 共2008兲.
21
H. Luo, B. Yin, H. Dai., and J. F. Doyle, “A 3D computational study of the
1 flow–structure interaction in flapping flight,” AIAA Paper No. 2010-556,
C. P. Ellington, “The novel aerodynamics of insect flight: Applications to
micro-air vehicles,” J. Exp. Biol. 202, 3439 共1999兲. 2010.
2 22
S. A. Combes and T. L. Daniel, “Flexural stiffness in insect wings II. H. Aono, F. Liang, and H. Liu, “Near-and far-field aerodynamics in insect
Spatial distribution and dynamic wing bending,” J. Exp. Biol. 206, 2989 hovering flight: an integrated computational study,” J. Exp. Biol. 211, 239
共2003兲. 共2008兲.
3 23
R. J. Wootton, R. C. Herbert, P. G. Young, and K. E. Evans, “Approaches S. Sunada, L. Zeng, and K. Kawachi, “The relationship between dragonfly
to the structural modeling of insect wings,” Philos. Trans. R. Soc. London, wing structure and torsional deformation,” J. Theor. Biol. 193, 39 共1998兲.

Downloaded 07 Jan 2011 to 129.59.78.81. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

Das könnte Ihnen auch gefallen