Sie sind auf Seite 1von 12

IEEE/ASME TRANSACTIONS ON MECHATRONICS, VOL. 18, NO.

2, APRIL 2013 787


A Simulation and Design Tool for a Passive Rotation
Flapping Wing Mechanism
Veaceslav Arabagi, Lindsey Hines, and Metin Sitti, Senior Member, IEEE
AbstractThis paper develops a numerical simulation tool for
designing passive rotation apping wing mechanisms. The simu-
lation tool includes a quasi-static model of the piezoelectric ac-
tuator bending, transmission kinematics, and the small Reynolds
number aerodynamic forces governing wing dynamics. To validate
the developed tool, two single-wing systems with distinct resonant
frequencies are manufactured and characterized. Comparison to
experimental results reveals that, although discrepancies exist, the
simulation is able to predict general trends of wing kinematics and
lift behavior as functions of frequency, thus, being useful as a design
tool. Finally, the complex models, ranging fromactuator deection
to wing aerodynamics presented in this paper, allow analysis of
the complete system revealing insight into several wing trajectory
control methodologies, and potentially serving as a design and op-
timization tool for future apping wing robots.
Index TermsFlapping wing, passive wing pitch reversal, sim-
ulation and design tool, wing lift force modulation.
I. INTRODUCTION
A
S in natural yers, miniature apping wing vehicles are
expected to present a signicant advancement in agility
from their xed and rotary wing counterparts. Inspiration and
basic concepts of apping ight are taken from studying the
kinematics of dragonies [1], butteries [2], hawkmoths [3],
bats [4], and ies [5]. The low Reynolds number operation
regime of small-scale iers is dominated by nonlinear aerody-
namics and temporal wingwake interactions [6], [7] that are
the underlying cause for their, unaccounted by conventional
aerodynamics, large lift force production capacity [8], respon-
sible for the acrobatic maneuverability of these insects. One
major milestone to overcome in creating a robotic apping
wing platform capable of controlled hover is the generation
of proper and controlled wing motion. Over time, two differ-
ent approaches to this problem emerged: 1) active control of
both the apping and rotation angles, and 2) active control of
Manuscript received February 8, 2011; revised September 2, 2011 and
October 27, 2011; accepted December 21, 2011. Date of publication February
20, 2012; date of current version January 10, 2013. Recommended by Technical
Editor S. Fatikow. The work of L. Hines was supported by the National Defense
Science and Engineering Graduate Fellowship.
V. Arabagi was with the Department of Mechanical Engineering, Carnegie
Mellon University, Pittsburgh, PA 15213 USA. He is now with the Department
of Cardiovascular Surgery, Childrens Hospital Boston, Boston, MA02115 USA
(e-mail: veaceslav.arabagi@childrens.harvard.edu).
M. Sitti is with the Department of Mechanical Engineering, Carnegie Mellon
University, Pittsburgh, PA 15213 USA (e-mail: msitti@andrew.cmu.edu).
L. Hines is with the Robotics Institute, Carnegie Mellon University,
Pittsburgh, PA 15213 USA (e-mail: lhines@cmu.edu).
Color versions of one or more of the gures in this paper are available online
at http://ieeexplore.ieee.org.
Digital Object Identier 10.1109/TMECH.2012.2185707
wing apping only, while allowing for passive rotation of the
wing.
Wood et al. proposed a apping wing platform capable
of fully controlled wing trajectory, and thus body forces and
torques [9]. However, that robot design relies on a difcult to
manufacture differential mechanism, driving both the wings
leading and trailing edges, and a complex high-bandwidth
controller [10]. However, recent theoretical and experimental
results of Bergou et al. suggest that at least some of the wing
rotation in insects is passively assisted by inertial and aero-
dynamic forces [1]. Furthermore, the apping wing regime is
known to possess several self-stabilizing damping properties for
translation and rotation maneuvers due to the reciprocal nature
of the wing beat [11], [12], and upright maintenance with the aid
of sail type devices [13]; thus, suggesting the possibility of an
underactuated wing drive mechanism. Based on this idea, a pas-
sive wing rotation approach has been undertaken by Wood, with
a single actuator driving the apping motion of the wings while
allowing for passive rotation, with the amplitude of rotation
limited by mechanical stoppers [14]. This mechanism results
in large impact forces on the wing imparted at the end of each
stroke, which could decrease its lifetime and excite undesirable
structural vibration modes.
This paper is based upon an aerial platform design featur-
ing completely passive wing pitch reversal. Without mechanical
stops and suspended from an axis close to its leading edge by a
spring/damper system, the motion of the wing is governed only
by the dynamics of the wing, aerodynamic forces, and the elas-
ticity of the system. Furthermore, this design not only simplies
mechanical complexity, but also presents opportunity for the
eventual introduction of wing angle of attack control via a vari-
able spring stiffness at the wing rotation joint, potentially using
smart, morphing materials. A dynamic simulation that includes
a compilation of small Reynolds number aerodynamic forces,
coupled with dynamics of the driving piezoelectric bending ac-
tuator and kinematics of the four-bar transmission, serves as the
basis for analysis of the systems performance. This model is
employed to simulate the behavior of two single-apping-wing
systems, one resonating at 30 Hz and the other at 75 Hz, and the
results are compared to experimental measurements. Although
the theoretical model is not able to predict the exact lift force
magnitudes, wing rotation dynamics are predicted accurately
with experimental apping angles as input. In addition, the sim-
ulation is able to capture behavioral trends very well, proving
useful as a simulation and design tool. Finally, three methods
to control the lift force of a passively rotating apping wing
mechanism are proposed: modulating the apping frequency,
the amplitude of the actuator oscillation, and the torsional stiff-
ness of the wing rotation joint.
1083-4435/$31.00 2012 IEEE
788 IEEE/ASME TRANSACTIONS ON MECHATRONICS, VOL. 18, NO. 2, APRIL 2013
Fig. 1. Schematic of the complete single-wing system, including the actuator,
slider crank, four-bar, and wing.
II. SYSTEM DYNAMICS MODELING
A numerical simulation of the actuator-transmission-wing
dynamics is developed in MATLAB to serve as a design and
optimization tool. The simulation includes three main compo-
nents: the actuator, a transmission mechanism, and the wing.
A. Actuator
Piezoelectric actuators are selected to power the robotic ap-
per due to their high efciencies, power densities of up to
400 W/kg [15], extensive theoretical modeling [16], [17], easy
incorporation into the composite manufacturing process, and
their use with some success by other groups, such as the Berke-
ley MFI [18], [19] and the Harvard MicroFly [20]. Although
piezoelectric materials also have several limitations, such as a
need for a high-voltage power supply, small strain (less than
0.1% for PZT-5H), and high brittleness, work has been done
to make their application feasible in miniature aerial vehicles
as presented in [21]. Reciprocating wing drive is achieved with
a bending unimorph piezoelectric actuator, whose schematic is
portrayed in Fig. 1. The modeled actuator consists of an ac-
tive bending portion and a rigid extension that increases its
stroke. Classical lamination theory was used to model the ac-
tuator quasi-statically, alike presented in [17]. The structural
elasticity of the unimorph actuator is introduced into the model
as a spring of stiffness k
act
attached to the end of a rigid ac-
tuator that bends under an applied electric eld, as shown in
Fig. 1. The inertial properties of the actuator have been simpli-
ed to the dynamics of a point mass positioned on the tip of
the actuator. The effective mass of the actuator, m
e
, calculated
based on an energy approach presented in [22], was determined
to be 15 mg for the 30-Hz resonant system. Its kinetic energy,
used for the Lagrangian expression, is T
a
=
1
2
m
e

2
, where
is the displacement of the slider crank end. The free, no load,
displacement of the actuator tip is modeled as
= g(E, actGeom) (1)
where E is the electric eld in the PZT layer and actGeom is
the geometry of the actuator.
B. Transmission
Afour-bar-type transmission is employed to amplify the small
tip displacements of the piezoelectric actuator. Due to actuator
Fig. 2. Displacement (left axis) and torque transmission curves (right axis)
for 75-Hz resonance single-apping-wing system.
out
is the torque output of
the four-bar. The lengths of these four-bars links are {L0, L1, L2, L3} = {13.5,
12.1, 10, 5} mm.
bending, a slider crank mechanism is implemented to transfer
the actuator output forces into the transmission. The four-bar
joints are modeled as ideal pin joints, with the exception of the
joint at the output link of the four-bar, designated by in Fig. 1,
which has a stiffness value assigned to it, as it undergoes the
largest range of motion and hence can signicantly affect system
dynamics. Based on the four-bar transmission kinematics mod-
eled in [23], the output angle and torque transmission function
of the four-bar is
= h(links
i
, config, ) (2)

M
drive
= f(links
i
, config,

F
in
) (3)
where

M
drive
is the moment that is applied to the wing about
its apping axis (E
3
axis from Fig. 3), links
i
are the lengths
of the 4 four-bar links, config is the conguration of the four-
bar, i.e., its position through the stroke, and

F
in
is the force
input from the actuator, as portrayed in Fig. 1. These general
equations are sufcient to describe any transmission system
transforming linear input displacement into angular output mo-
tion. In the case of a piezoelectric actuator, F
in
= k
act
( )
and represents how the actuator stiffness enters into the theo-
retical model. Numerical optimization is employed to achieve a
particular transmission ratio and symmetry over the actuator dc-
bias voltage, yielding transmission curves portrayed in Fig. 2.
Generally, higher torque transmission ratios imply a higher res-
onance frequency if everything else is kept constant. Intuitively,
this arises from the fact that higher transmission ratios impart
smaller wing inertial forces on the actuator and hence the latter
senses a smaller effective wing inertia.
C. Wing Aerodynamics
Passive rotation of the wing is achieved by attaching the wing
to a driver spar with an elastic joint, thus, allowing the elas-
tic/damping properties of the exure material to inuence wing
motion. In fact, the interaction of these elastic and damping
torques with the wings inertia and aerodynamic forces is essen-
tial to the generation of proper apping wing trajectory. Fig. 3
portrays the geometric setup of the simulation model and the
sets of Euler angles for coordinate transformations.
Throughout its motion, the wing is affected by aerodynamic
forces stemming from various small Reynolds number effects.
ARABAGI et al.: SIMULATION AND DESIGN TOOL FOR A PASSIVE ROTATION FLAPPING WING MECHANISM 789
Fig. 3. Schematics of the passive apping wing setup. Coordinate sets repre-
sent transformations established by (apping angle) and (rotation angle).
The coordinate systems are shifted for clarity, while in simulation they are all
centered at the point labeled coordinate sets origin. The

E

coordinate system
is attached to the wing.
Fig. 4. (a) Wing SolidWorks model and (b) wing prole illustration of aero-
dynamic force due to rotation. The wing center of mass and rotation axis are
illustrated.
Overall, the total force on the wing is
F
tot
= F
t
+ F
rot
+ F
air
+ F
wc
(4)
where F
t
is the translational force, F
rot
is the force due to wing
rotation, F
air
is the force due to added air mass, and F
wc
is the
force created by wake capture. Translational force estimates are
quasi-steady approximations adapted from thin airfoil theory.
F
rot
is the force generated by air resistance to wing rotation as
portrayed in Fig. 4(b). The added air mass forces are imparted
on the wing by the surrounding air and are generated due to
unsteady wing motion [24]. This is also the only force that aids
the rotation of the wing during stroke reversal, since the inertia
of the air trapped in the vicinity of the wing tends to continue
its linear motion, thereby imparting a torsional moment on the
wing [25]. The wake capture effect is a nonsteady phenomenon
occurring when the wing traverses vortices and air circulation
generated by its motion prior to the direction change. Since a
closed form expression for this force cannot be determined, it is
not included in the dynamical model. Thus, the expressions for
aerodynamic forces acting on chordwise wing strips, dr, take
the form
dF
t
=
1
2
U
2
c(r)[C
2
l
() + C
2
d
()]
1/2
dr (5)
dF
rot
=
1
2
C
rot
| |

c(r)
z

|z

|dz

dr (6)
dF
air
=

4
(

cos

sin)rc(r)
2
dr

z
RA

c(r)
2

c(r)
2
dr (7)
where is the air density, r is the radial position of a wing strip
from the wings apping axis, c(r) is the chord length of the
particular strip, and are the apping and rotation angles,
respectively (portrayed in Fig. 3), is the angle of attack of
the wing, or (/2 ), and z
RA
is the location of the rotation
axis measured from the wings leading edge. The translational
force component is obtained by vector addition of mutually
orthogonal lift and drag forces. U is taken to be the velocity in
the

E

2
direction of each wing strips midchord point
U(r) = v
mc


E

2
= r

+
c(r)
2
cos (8)
where v
mc
is the midchord velocity of each wing division. The
use of midchord velocity in the translational lift force expres-
sion is justied by the fact that the incoming air stream velocity
is linearly distributed along each chord of the wing; hence, the
best approximation to a constant velocity eld, which was used
to obtain the translational lift coefcients in [8], is the mean ve-
locity along each wing chord, observed at the midchord point.
C
l
, C
d
, and C
rot
are the translational lift, drag, and rotational
force coefcients, respectively, where the translational coef-
cient expressions were tabulated experimentally by Sane and
Dickinson for a Reynolds number (Re) of 136 [25]
C
l
() = 0.225 + 1.58 sin(2.13 7.2) (9)
C
d
() = 1.92 1.55 cos(2.04 9.82) (10)
where is in degrees. The position of the rotation axis has been
set at 1/4 chord length from the leading edge of the wing, at
its aerodynamic center (according to thin-plate airfoil theory).
The Reynolds number for our system is Re = 2800, calculated
by averaging the Re numbers obtained for each wing strip, is
similar, in aerodynamic sense, to RoboFlys one, Re = 136,
justifying the use of Sane and Dickinsons ndings as means of
approximating the aerodynamic forces on the wing. The value
for the rotational drag coefcient C
rot
was taken to be 2, as
this is the theoretical result of rotational drag on a at plate
subjected to normal ow [26], [27]. To account for the radial
variability of wing velocity, the aforementioned equations need
to be integrated numerically to obtain the aerodynamic forces
on the entire wing.
D. Equations of Motion of the Passive Rotation Wing System
The wing has a carbon ber leading edge spar with an ex-
truding vein for chordwise stiffness and a Kapton membrane
as the main airfoil (see Fig. 4). The inertial properties of the
wing required for the dynamic equations are obtained directly
from the SolidWorks model. In addition, the three carbon ber
links of the four-bar and the slider crank mechanism were mod-
eled as inertial elements. The dynamic equations are formulated
in Lagrangian form, with generalized coordinates of and
being the apping and rotation angles, respectively. Thus, we
790 IEEE/ASME TRANSACTIONS ON MECHATRONICS, VOL. 18, NO. 2, APRIL 2013
formulate the Lagrangian as follows:
L = T V =
1
2
mv v +
1
2
J
+

1
2
m
Li
v
Li
v
Li
+

1
2
J
Li

Li

Li
+
1
2
m
e

1
2
k
rot

2
(11)
where m is the wing mass augmented with the added air mass
[as described in (19)], v is the velocity at the wings center
of gravity (CG), J is the wings inertia matrix taken about the
wings center of mass, k
rot
is the torsional spring stiffness at the
rotation axis, m
Li
, J
Li
, v
Li
, and
Li
are the masses, rotational
inertias, linear and angular velocities of the four-bar links 13,
respectively, m
e
is the effective mass of the actuator,

is the
horizontal velocity of the tip of the actuator (scalar), and is
the angular velocity of the wing dened as
=

3
+

1
=

1
+

sin()

2
+

cos()

3
(12)
expressed in terms of basis vectors of the coordinate frame at-
tached to the wing. Furthermore, the expressions for angular
and linear velocity of the four-bar links are complex in closed
form and, hence, have been substituted by numerically com-
puted trajectories parametrized by the output apping angle
(i.e., v
2
(t) = f((t)) and hence
dv
2
(t)
dt
=

(t)
f
(t)
). Den-
ing the conventions f


f
(t)
and

f
df
dt
, and substituting
all knowns into the Lagrangian L, we obtain the following
equation:
L = 1/2(k
rot

2
+ sin()
2
J
yy

2
+ (cos()J
xz

+J
xx
)+cos()

(cos()J
zz

+ J
xz
)
+ 1/2m((2R
2
CG
+
2
cos(2))

2
+ 4R
CG
cos()


+ 2
2

2
) +
3

i=1
m
Li

2
x

Li

2
+
2

i=1
J
Li

2
Li
+ J
L3

2
+ m
e

2
) (13)
where R
CG
and
CG
are the radial and vertical positions, re-
spectively, of the center of mass fromthe origin of the coordinate
systems, as portrayed in Fig. 4, J
ii
are the components of the
wings inertia matrix, x
Li
is the CG position of link i of the
four-bar, and
Li
is the orientation of the links (note that the
four-bar motion is planar and equations are simplied to reect
2-D motion). Note that for a thin wing, the inertial components
J
yz
and J
xy
are extremely small and have been omitted from
the equations. Thus, the Lagrange equations describing the wing
apping and rotation dynamics are as follows:
d
dt

=

M
aero


E

1
d (14)
d
dt

= (

M
drive
+

M
aero
)

E
3
D

(15)
where d is the rotational damping coefcient, D is the ap-
ping damping coefcient arising from the transmission exure
damping,

M
drive
is the driving apping torque, and

M
aero
is the
moment due to the aerodynamic forces, explicitly dened as

M
aero
=

F
i

i
() (16)
where

F
i
s are the translational and rotational aerodynamic
forces and

i
s are the respective positions of their centers of
pressure from the wing rotation axis, explicitly dened as
|
t
()| =

0.82

|| + 0.05

c(r) (17)
|
rot
| =

span
z

dF
rot

span
dF
rot
(18)
where
t
denes the position of the translational lift center of
pressure as a function of angle of attack for each wing chord
strip c(r) and is obtained from experimental results of Dickson
et al. [28].
rot
is the effective moment arm of the rotational
force distribution of Fig. 4(b).
Given that there is no analytical solution for the added air
mass force on a wing planform moving in 3-D fashion, dF
air
of (7) does not accurately describe the moments exerted by
the added air mass on the apping and rotation angles. As an
approximation, the effect of added air mass forces was imple-
mented in the form of virtual mass that augmented the physical
mass of the wing, similar to the inertial implementation in [26]
and [29]. Stemming from the concept that the added mass effect
can be estimated as dm =

4
c(r)
2
dr [30], the expression for
m emerges as follows:
m = m
w
+

span

4
c(r)
2
dr (19)
where m
w
is the physical mass of the wing. This approximation
allows a simplied treatment of the added air mass effects that
is sufcient for the development of a simulation tool able to
predict general dynamics and lift force trends. Although this
mass augmentation approach is used in the differential equations
dening the systemdynamics, the expression for added air mass
force of (7) is used to estimate the aerodynamic lift generated
by this effect in all lift plots presented in this paper.
The previously presented Lagrange equations (14) and (15)
are fully general, governing the dynamics of a passive rotation
apping wing driven by a torque

M
drive
around the apping
axis. In our case of the driving mechanismconsisting of a piezo-
electric actuator coupled to a four-bar transmission,

M
drive
is
given by (3). Thus, substituting all the known quantities into (14)
and (15) and simplifying, we obtain the nonlinear differential
equations governing the apping and rotation angles that can be
found in the Appendix, due to their length and complexity.
III. MANUFACTURING AND EXPERIMENTAL
PARAMETER DETERMINATION
In order to best quantify the theoretical simulations perfor-
mance, two experimental systems consisting of an actuator, a
four-bar and wing were manufactured, having different reso-
nant frequencies of 30 and 75 Hz. Both systems feature the
ARABAGI et al.: SIMULATION AND DESIGN TOOL FOR A PASSIVE ROTATION FLAPPING WING MECHANISM 791
Fig. 5. Overall dimensions of the manufactured wing and actuator for the
(a) 30-Hz resonant system and (b) 75-Hz system. The actuator mount is 3-D
printed from VisiJet HR 200 material, and assembled with Cyanoacrylate glue.
same construction components, differing only in their dimen-
sions and lengths.
A. Component Manufacturing
All the parts comprising the apping wing system were man-
ufactured using the Smart Composite Microstructures (SCM)
methodology and manually assembled afterward [9]. This last
manual step generally introduces some misalignment and uncer-
tainty in the nal assembly, which tends to adversely affect the
dynamics of the system, generally in the formof unwanted com-
pliance. The transmission members are formed of two preim-
pregnated, unidirectional M60J ultrahigh modulus carbon ber
layers of 60-m thickness (Toray) sandwiching a 12.7-m layer
of Kapton (DuPont Kapton 50HN) and cured together. The wing,
whose Kapton layer serves as a membrane, is manufactured from
thinner Kapton of merely 6.3 m (DuPont Kapton 25HN), as
the minimization of the wing weight is of major importance.
The actuators active layer consists of two layers of carbon
ber bonded to a 125-m-thick layer of PZT-5H (Piezo Sys-
tems), while the passive extension has an additional layer of
s-glass, increasing stiffness. Some dimensions of the actuator
and the wing used in the experimental apping wing systems
are portrayed in Fig. 5.
B. Actuator Experimental Characterization
Since the piezoelectric actuator is of key importance in den-
ing systemdynamics, experiments were performed to obtain the
actual actuator blocking force, tip stiffness, and material prop-
erties. Using laminar plate theory, the actuator bending stiffness
with a force applied at its tip, labeled k
act
in Fig. 1, replicating
the operational scenario, was calculated at 160 N/m. Through
indentation experiments, the actual actuator stiffness was found
to be 169 N/m, measured by the ratio of force over given tip
displacements.
C. Wing Rotational Flexure Characterization
The dynamic parameters of the rotational exure were exper-
imentally measured after manufacturing, to account for any im-
perfections. The stiffness and damping properties of each exure
were determined by applying an impulse impact to it, after an
additional known mass was added. The resulting displacements
were recorded with a laser scan micrometer (Keyence LS-3100,
Woodcliff Lake, NJ), and, being second-order responses, al-
lowed the determination of stiffness and damping parameters
fromthe dominating frequency and envelope of the signal. These
test were performed under vacuum such as to eliminate aerody-
namic damping, which would constitute the main component of
damping otherwise.
D. Transmission Compliance Characterization
Given that the slider crank and four-bar exures are sources
of unwanted compliance, the stiffness and damping properties
of the transmission are measured experimentally. Compliance
of the slider crank is measured via an indentation test at the tip
of the rst transmission link, with the actuator rigidly xed. The
recorded tip stiffness is transformed via the transmission ratio
to stiffness of the slider crank and modeled in simulation in
series with the actuator tip stiffness k
act
. Throughout our exper-
iments, slider crank stiffness ranges as low as k
sc
= 300 N/m
and as high as k
sc
= 1650 N/m. The large span of stiffness val-
ues is attributed to the easy misalignment of the slider crank
links, resulting in exure buckling. For all simulations in this
paper, the 30-Hz resonance apping system was simulated with
a nominal slider crank stiffness of k
sc
= 1400 N/m, while the
75-Hz system had a value of k
sc
= 300 N/m, per experimental
measurements.
The damping coefcient of the four-bar was measured by im-
parting an impulse on wing-driving mechanism, with the wing
rotated 90

, i.e., = 0

, so as to eliminate aerodynamic drag


from the measurements. The simulated impulse response was
then matched to experimentally recorded vibrations via altering
the apping damping parameter D from (15), converging to a
nal value of D = 100 Nmms, employed in simulation.
IV. EXPERIMENTAL SETUP AND WING KINEMATICS
Experimental lift was measured by mounting the single-wing
assembly to a load cell through a mechanical amplication link-
age to increase the sensed lift forces, and the low-friction, single-
degree-of-freedom (DOF) pivot isolates the vertical forces gen-
erated by the wing [see Fig. 6(b)]. A 25-g load cell (Transducer
Techniques) for the 75-Hz resonant prototype (30-g load cell
for the 30-Hz system) was attached to the main shaft via a
contact connection, such that no torques in the shaft would be
transmitted into the load cell. The actuator, wing, and four-bar
mechanism were mounted directly to a 3-D printed (Invision
HR 3D Printer) plastic mount that was bolted onto an aluminum
shaft, portrayed in Fig. 6(a). Given the large inertia of the bal-
ance system, its resonant frequency was determined at 13 Hz.
This is sufciently distant from the operating frequencies of
30 and 75 Hz to ensure lift measurements were unexaggerated;
however, functioning as a low-pass lter, measurements are re-
stricted to mean lift force only.
During each test, the input voltage to the actuator, a sinusoidal
wave specied by the peak-to-peak voltage V
pp
and a constant
dc offset V
dc
. Wing kinematics were captured using a strobe light
and a high-denition camcorder (Sony HDR-SR11). Video was
taken from the top view of the wing, with the camera placed
792 IEEE/ASME TRANSACTIONS ON MECHATRONICS, VOL. 18, NO. 2, APRIL 2013
Fig. 6. Photographs of the experimental lift acquisition setup with main com-
ponents labeled. (a) Actuator-four-bar-wing experimental prototype mounted
onto a 3-D printed plastic part. (b) Lift balance mechanical amplication
device.
in an offset plane parallel to the stroke plane. Three points on
the wing were tracked, based on edge contrast, from which the
wing trajectory was calculated.
A. Wing Kinematics Comparison
Given the previously described setup to obtain kinematics,
a quick comparison is performed to ensure that the numerical
simulation behaves similar to the system it models: experimen-
tal kinematic wing data are compared to predicted results in
Fig. 7(a) for the 30-Hz resonant system. The kinematics feature
a discrepancy of 10

peak to peak in apping angles, propa-


gating in a discrepancy of 20

peak to peak in rotation, likely


due to the asymmetry of the torque transmission curve of the
physical four-bar transmission. The simulated lift force compo-
nents shown in Fig. 7(b) portray that translational lift constitutes
most of the total lift force with a small contribution from added
air mass lift. The rotational lift component is very low, instead
the effect of this force is manifested mostly through the torque
it imparts on the wing, thereby having a great impact on shaping
the kinematics of wing rotation and inuencing total lift force
in that manner.
V. RESULTS AND DISCUSSION
Characterization of the developed theoretical model of a
single-wing-apping system is attempted through comparison
to experimentally obtained wing kinematics and mean lift force
measurements. The full dynamic model can be analyzed by sep-
arating it into two subsystems: 1) the actuator/transmission drive
and wing-apping dynamics; 2) the aerodynamic force models
and wing rotation dynamics. The performance of the second
subsystem can be quantied by using the experimentally mea-
sured wing apping trajectory as a displacement source in the
simulation of wing rotation only. Running the full simulation,
Fig. 7. (a) Overlayed experimental and simulated wing kinematics, and
(b) trajectories of simulated lift forces of the 30-Hz resonance single-wing
system. System parameters: V
pp
= 180 V, V
dc
= 30 V, frequency = 30 Hz,
k
sc
= 1400 N/m, k
rot
= 20 mNmm, and d = 10 Nmms.
with the input of a sinusoidal signal driving the actuator, enables
analysis of the wing driving mechanism. This characterization
of the simulation performance is done for both the 30- and
75-Hz resonant systems, in order to give an idea about the
models shortcomings in simulating systems of different reso-
nances.
A. Analysis of the 30-Hz Resonant System
Since a frequency response characterizes the behavior of a
dynamic system well, this methodology is applied to analyze
the performance of the numerical model. For the 30-Hz reso-
nant experimental system, the frequency response was obtained
by recording the wing kinematics and load cell output while
sweeping the frequency of the input voltage signal at discrete
steps of 3 Hz. This relatively large value of the frequency step-
size is acceptable due to the low-quality factor (Q 2) of a
apping wing system, ensuring smoothness of the frequency re-
sponse curve and, thus, no important missed dynamics. In order
to validate the quasi-steady aerodynamic models and underlying
rotation dynamics, the leading edge of the simulation wing is
driven based on experimental kinematics. The simulation takes
as input for the apping angle a sinusoid signal with the ampli-
tude obtained from measured wing kinematics. The kinematics
plot portrays the two fundamental resonant frequencies of the
system manifested by peaks in the apping amplitude, occur-
ring around 30 and 36 Hz, as shown in Fig. 8. The two resonant
points of the system correspond to an optimal balance between
the internal stresses of the actuator, inertial forces, and nonlin-
ear aerodynamic damping effects acting on the wing. Simulated
wing rotation closely resembles the experimental observations
ARABAGI et al.: SIMULATION AND DESIGN TOOL FOR A PASSIVE ROTATION FLAPPING WING MECHANISM 793
Fig. 8. Frequency sweep of the 30-Hz resonance system simulating only wing
rotation, with apping angle set by a sinusoid wave with experimentally obtained
amplitudes. (a) Experimental and simulated wing kinematics, and (b) lift forces.
System parameters: V
pp
= 180 V, V
dc
= 30 V, k
rot
= 20 mNmm, and d =
10 Nmms.
except for high frequencies; this is thought to occur due to the
excitement of second resonance mode and out-of-phase rota-
tion, addressed in detail in Section V-C. The lift force plots of
Fig. 8(b) portray very similar magnitudes as the experimental
mean lift curve. Translational lift is the dominant aerodynamic
force throughout this frequency range, with the added air mass
lift comprising half of this amount. Similar to the kinematics, the
simulated lift force curve diverges fromexperiments at high fre-
quencies, suggesting the dominance of unmodeled aerodynamic
effects present in that operating regime.
Simulating the complete apping wing system along with
actuator and transmission when driven by a sinusoid signal,
we obtain the plots presented in Fig. 9. Both the simulated
kinematics and total lift force are comparable in magnitude to
their experimental measurements for frequencies below the sec-
ond resonance peak. The model predicts well the occurrence
of the rst resonant peak while overpredicting the occurrence
of the second peak at 50 Hz, although being highly damped.
The simulated frequency sweep is performed for a larger range
of frequency values, while experimental measurements were
stopped earlier due to extreme wing rotations after the second
resonance, destructive to the prototype. Furthermore, the curves
labeled with a slider crank stiffness value, k
sc
, represent total
lift force curves obtained from simulation runs of single-wing
systems featuring bounding values for the corresponding values
of slider crank stiffness, namely k
sc
= 300 and 1650 N/m. The
change in the magnitude of the total lift curve as well as the
observed shift in system resonance illustrates the systems sen-
sitivity to sources of compliance in the wing drive mechanism.
Fig. 9. Frequency sweep obtained with the complete system simulation, in-
cluding the actuator drive and wing dynamics. Systemparameters: V
pp
= 180 V,
V
dc
= 30 V, k
sc
= 1400 N/m, k
rot
= 20 mNmm, and d = 10 Nmms. Curves
labeled with a slider crank stiffness are obtained with simulated systems having
critical values k
sc
= 300 and 1650 N/m.
Fig. 10. Frequency sweep of the 75-Hz resonant system simulating only wing
rotation, with apping angle set by a sinusoid wave with an experimentally
obtained amplitude. (a) Experimental and simulated wing kinematics, and
(b) lift forces. System parameters: V
pp
= 150 V, V
dc
= 50 V, k
rot
=
37.5 mNmm, and d = 13 Nmms.
B. Analysis of the 75-Hz Resonant System
To obtain the 75-Hz resonant system frequency response, lift
was measured with a balance system (see Fig. 6) when the ac-
tuator was driven via a continuous frequency sweep. Fig. 10(a)
portrays good matching between the experimental wing rota-
tion angles and the simulation modeling only wing rotation.
794 IEEE/ASME TRANSACTIONS ON MECHATRONICS, VOL. 18, NO. 2, APRIL 2013
Fig. 11. Frequency sweep obtained with the complete simulation. System
parameters: V
pp
= 150 V, V
dc
= 50 V, k
sc
= 300 N/m, k
rot
= 37.5 mNmm,
and d = 13 Nmms. The lift bounding curve corresponding to a slider crank
stiffness of k
sc
= 300 N/m coincides with the nominal frequency response,
while the other bounding curve corresponds to k
sc
= 1650 N/m.
The experimental apping kinematics portray two resonance
peaks spaced close to each other, similar to the low resonant
frequency system. The simulated lift forces in Fig. 10(b) are
slight underestimates from experimental results. The largest
lift contribution comes from added air mass lift, with transla-
tional lift being half of its value. This result arises from the fact
that the high-frequency low-amplitude operation regime of the
75-Hz systemmaximizes the accelerations of the wing, and thus
the air in its vicinity, i.e., added air mass, while large apping
amplitudes and translational wing motion dominate the behavior
of the 30-Hz resonant system. The large increase in simulated
lift at 95 Hz seen in Fig. 10(b), not present in the experimental
response, arises from the increase in apping amplitude at that
frequency. This occurs due to the excitement of the second res-
onance mode in the experimental system, predicted poorly by
the simulated system, due to unmodeled aerodynamic effects,
discussed in detail in Section V-C. Furthermore, after examin-
ing the phase shift between the apping and rotation sinusoids,
a shift of 81

is observed in simulation and one of 138

in
experiment, suggesting that only the experimental system has
begun the transition to out-of-phase rotation, thus, reducing the
produced lift.
The full simulation of the 75-Hz resonance system, portrayed
in Fig. 11, overall portrays much worse agreement of simula-
tion to experiment. The twofold discrepancy in total lift force
produced is attributed primarily to the greatly underpredicted
apping amplitude in simulation, 20

peak to peak. The larger


amplitudes also cause an eightfold increase in translational lift
from the results of Fig. 10(b), rendering it the dominant lift
component, alike in the 30-Hz resonance system. Similar to
previous results, the simulation predicts well the rst resonance
mode at 75 Hz, while overestimating the second one at 115 Hz,
while experimentally it is observed at 95 Hz. The curve of
k
sc
= 1650 N/m, of Fig. 11, represents simulated total lift of a
Fig. 12. Frequency spectrum of the system impulse response. The motion of
the rst four-bar link was recorded after an impulse has been applied. The rst
and second system resonant peaks are indicated by arrows. Note that the 50-
and 100-Hz peaks of the experimental system are due to electrical noise in the
environment and do not reect physical resonant frequencies.
system featuring a stiff slider crank. The other limiting value of
stiffness k
sc
= 300 N/mis the nominal magnitude for this exper-
imental prototype. Again, a change in this parameter propagates
into a resonant frequency increase by more than 10 Hz, having
great impact on the behavior of the system.
C. Discussion
The presented comparison between experimental measure-
ments and both wing rotation only and the complete system
simulation effectively illustrates the capabilities and limitations
of the theoretical model of the single-apping-wing system.
The overall good agreement to experimental results of both
kinematics and lift forces produced by the simulation of wing
rotation only suggests that wing inertial properties, exure stiff-
ness, and damping properties, as well as the rough, quasi-steady
aerodynamic models are largely correct. The complete simu-
lation, encompassing the actuator drive, yields less acceptable
results, overpredicting lift force, especially for the 75-Hz reso-
nant frequency system. Given that most of this discrepancy is
attributed to the underprediction of apping amplitude and that
wing rotation dynamics are correct, as established earlier, the
underlying cause of the problem is traced to the apping mo-
tion of the wing. The decrease in wing apping displacement
could arise from actuator depolarization, inconsistencies in the
manual alignment and assembly of the four-bar transmission, as
well as from spanwise aerodynamic moments not encompassed
by quasi-static 2-D approximations.
One of the limitations of the developed theoretical simulation
is its inability to model accurately the behavior of the systemfor
frequencies higher than the rst resonance. These discrepancies
are thought to be products of nonlinear aerodynamic effects
caused by large stroke and rotation amplitudes. As a test to
this hypothesis, the 75-Hz resonant systems impulse response
is recorded, featuring small apping and rotation stroke am-
plitudes (<5

) and thus rendering the aforementioned dynamic


effects not important. Comparing the frequency components
of the experimental and simulated impulse responses obtained
with the fast Fourier transform algorithm, shown in Fig. 12, we
observe overlapping peaks at 78 and 129 Hz, suggesting that
the inertial, force transmission, and stiffness properties of sys-
tem components were modeled adequately in the numerical
ARABAGI et al.: SIMULATION AND DESIGN TOOL FOR A PASSIVE ROTATION FLAPPING WING MECHANISM 795
simulation. Furthermore, the second resonance peak is de-
creased from 129 to 95 Hz in actuated apping experiments,
suggesting a large effect of added air mass and potentially other
unmodeled forces on wing dynamics. The two resonance peaks
should not be viewed as the apping and rotation resonances
of the wing, but rather as different modes of vibration of the
system. Given that these resonances are reected as peaks in
apping amplitude, they portray an optimally balanced state
between internal actuator stresses and aerodynamic forces im-
parted on the wing. The rst resonance achieves this balance
via approximately symmetric wing rotation, with wing pitch
reversal occurring roughly at the time of stroke reversal (the
exact timing of rotation is a complex function of many differ-
ent parameters), thus resulting in a considerable amount of lift
force. The second resonant peak maximizes apping amplitude
because the drag on the wing signicantly decreases, thus, al-
lowing larger deections of the actuator. The reduction of drag
occurs because the wing begins to rotate out of phase with ap-
ping, such that the apping and rotation curves of Fig. 7 are 180

offset in phase. This onset of second resonance, when the center


of mass of the wing performs very little translational motion, is
the cause for the sharp decrease in experimental lift in the two
systems shortly after the rst resonant frequency. The reason
for the simulations overprediction of the second resonant fre-
quency is attributed to unmodeled transient aerodynamic forces,
such as wake capture [6], [25]. These dynamic effects generated
by the wings past trajectory create impulse-like loading on it
during each stroke, thus, exciting the second resonance mode
much earlier than predicted.
It is generally desirable to engineer the systemsuch as to space
the two resonance modes far apart in the frequency domain.
One may attempt to do so by utilizing a stiff wing rotational
joint, such as not to excite the second mode; however, this is
suboptimal from the performance perspective, since the wing
will rotate insufciently in its rst resonance mode, which is the
main operating point. Furthermore, stiffening either the wing
rotation exure or the actuator has the effect of shifting both
peaks of Fig. 12 to higher frequencies, as well as increasing
their separation. The proximity of the two resonance peaks in
an optimized lift system is inherent in the passive wing rotation
system at hand, where the inertia of the wing is signicant and
plays a major role in pitch reversal. By reducing the overall
inertia/mass of the wing and employing a correspondingly stiff
wing rotation joint, it is possible to increase the frequency of
the second resonance much higher, thus, reducing interference
and broadening the frequency range of maximum lift.
Finally, although the model generally overpredicts total lift
and overestimates the systems second resonance frequency,
its usefulness lies in capturing trends in wing kinematics as a
function of frequency, amplitude, spring stiffness, etc. Thus,
the model can be used as a design tool to obtain dimensions
and conguration of the actuator, four-bar transmission, and
wing that would ensure wing lift is maximized at a particular
frequency. The rst resonance peak is predicted accurately in
simulation, but, since it is sensitive to slider crank stiffness,
it should be designed in accordance with the bounding values
of that parameter. Although the predicted magnitude of max-
imum lift is exaggerated, its occurrence at 1020 Hz higher
than the rst resonance peak is well predicted by the simulation
and enforced by experimental measurements. Hence, although
an optimization of system parameters based on the numerical
model might yield overestimated lift forces, the physical opti-
mal point will lie in the close vicinity of the one predicted, as it
is dened by trends and general behavior, captured sufciently
by the numerical model. Furthermore, given that the theoreti-
cal model captures well the aerodynamic effects present in the
large amplitude apping regime, it should be used as an opti-
mization routine for apping wing systems designed to operate
in those conditions. Moreover, given that large apping ampli-
tudes are featured by most natural apping wing insects, this
feature should be sought in future system designs, rendering the
simulation tool extremely useful. The optimization of such sys-
tems should be done in accordance with both resonance modes
of the system (such as the ones portrayed in Fig. 12), with their
proximity dictated by optimality of lift force production at rst
resonance.
VI. PROPOSED WING CONTROL METHODOLOGY
The passive wing rotation design allows proper wing motion
without the weight of additional actuators but, incorporated into
a vehicle, results in an overall underactuated system. Deliberate
underactuated design is not uncommon (see for example un-
deractuated mechanical grippers [31]), but can complicate the
control of the nal platform. A variety of control techniques
exist for certain classes of nonlinear underactuated systems, in-
cluding combinations of hybrid and switching control, passivity
based methods, coordinate transformations, and basic lineariza-
tion [32]. Adaptive control and model learning techniques have
been employed in restricted DOF scenarios [33]; however, as of
yet, no overarching control theory for these systems has been
developed, making the success of any particular system unsure.
Nevertheless, control over the produced wing lift force is a vital
requirement for the incorporation of a apping wing mechanism
into any robotic platform, since future vehicle design could rely
primarily on this method for the production of body torques. We
propose three methods of modulating wing lift: controlling the
apping frequency, the waveform amplitude driving the actua-
tor, and the torsional stiffness of the passive rotation joint. The
effectiveness of these methods is investigated experimentally,
resulting in data for Figs. 13 and 14. Note that these experi-
ments were performed with a different single-wing prototype
than the rest of the experiments in this paper, having a resonant
frequency of 70 Hz; however, the mechanical construction
was identical to the prototypes discussed earlier.
Figs. 9 and 11 portray the system response to changing ap-
ping frequency. From the control standpoint, overall lift in-
creases linearly until the resonant frequency, after which point
the behavior diverges based on system construction. Further-
more, given that system efciency is maximized at resonance,
only slight deviations from the main operating frequency are
envisioned, rendering frequency a poor control input.
Alternatively, varying the amplitude of the driving waveform
of the actuator modulates the apping stroke of the wing, thus,
796 IEEE/ASME TRANSACTIONS ON MECHATRONICS, VOL. 18, NO. 2, APRIL 2013
Fig. 13. Simulated and experimental system response to a sweep in the input
waveform amplitude. Simulation parameters: V
dc
= 50 V, frequency = 70 Hz,
k
rot
= 50 mNmm, and d = 5 Nmms. Experimental lift data are averaged
over three datapoints.
Fig. 14. Simulated system response to a sweep in the rotational spring stiff-
ness. Simulation parameters: V
pp
= 150 V, V
dc
= 50 Vand frequency = 70 Hz,
and d = 5 Nmms. Experimental lift data are averaged over three datapoints.
affecting the lift force. The simulated systemresponse to chang-
ing peak-to-peak driving amplitude, overlayed with experimen-
tal data, is portrayed in Fig. 13. A good correlation between
experimental and simulated data is observed, although simu-
lated lift is consistently higher than in experiment. The linear
behavior of the overall lift force and large span of produced
forces renders this control scenario very promising. Intuitively,
an increase in the lift force could always be achieved by in-
creasing the apping amplitude; yet, this approach is limited by
electrical depoling and stress induced crack propagation of the
piezoelectric actuator.
The third envisioned method of controlling lift is modulating
the stiffness of the rotational exure that suspends the wing.
Unlike the high-bandwidth control architecture required for the
Micromechanical Flying Insect [34], the inherently stable na-
Fig. 15. Experimental lift surface. Represents variation of produced lift with
input voltage and torsional spring stiffness. Experimental parameters: V
dc
=
50 V, frequency = 70 Hz, and d = 5 Nmms.
ture of the passive apping wing allows slow actuation of the
rotational joint stiffness in ight. Five exure joints of vary-
ing thickness were manufactured and tested in the experimental
system in order to quantify the effect of exure stiffness. Exper-
imental and simulated results are portrayed in Fig. 14 and illus-
trate the expected correlation of decreasing lift with increasing
exure joint stiffness. The increase in apping amplitude with
increasing joint stiffness is attributed to the shifting of both res-
onance modes (as illustrated in Fig. 12 to higher frequencies and
an increase in Q, a fact experimentally conrmed. Control over
the stiffness of the passive rotation joint is envisioned via an
actively stiffness changing or smart material, such as dielectric
elastomers [35] or ionic polymer metal composite (IPMC) actu-
ators [36] actuators, with design, integration being future works.
Generally, it should be noted that, although each presented
control method has limited application and effect, a combination
of these mechanisms could be employed in a dual wing robot.
Indeed, control over rotational exure stiffness and actuator in-
put voltage allows production of any lift force on the surface
of Fig. 15, thus, maximizing the range of achievable lift forces.
The lift force surface resembles a plane, suggesting a simple
linear controller implementation. Asymmetrical lift and, there-
fore, body torques could be achieved by altering the rotational
stiffness on only one passive rotation wing joint of a dual wing
aerial platform via a morphing polymer. Furthermore, due to a
reduced number of actuation inputs into the system, some of the
six DOFwill likely be coupled and uncontrolled, requiring either
additional actuation mechanisms or combining control inputs to
achieve full position/orientation of the robot in free space. It
should also be noted that, although being very advantageous in
some respects, the passive pitch reversal wing, proposed in this
paper, would render the aerial platform less agile and mobile
than its active control counterparts. However, the mechanical
simplicity, reduced complexity of the controller, and naturally
stable behavior of wing rotation, that benet a passive rotation
design, could potentially bring us closer to the goal of liftoff
and stable hovering.
ARABAGI et al.: SIMULATION AND DESIGN TOOL FOR A PASSIVE ROTATION FLAPPING WING MECHANISM 797
VII. CONCLUSION
It is of great importance for any design project to have a simu-
lation tool to predict the performance of the nal system. Given
that our goal is the development of a miniature, controlled, ap-
ping ight robot, a full simulation of its subsystema single-
wing designcould greatly speed up the design process, as
well as enable optimization of its components. The numeri-
cal simulation presented in this paper includes a model of the
bending piezoelectric actuator, transmission kinematics, and the
passively rotating wing aerodynamics. Two experimental pro-
totypes featuring different resonance frequencies were manu-
factured from carbon ber composite material, enabling the
comparison of simulation to experimental results. Overall, there
was broad agreement between the apping and rotation trajec-
tory trends between the numerical simulation and experiment.
The major discrepancy between the measured and predicted lift
forces is attributed to manufacturing/material imperfections of
the actuator and/or four-bar transmission, and unmodeled aero-
dynamic forces. However, the model is able to predict the main
resonance peak accurately along with general trends of wing
kinematics and lift behavior as functions of frequency. Thus,
given its limitations in predicting the magnitude of produced
lift force, the simulation can be used as a design tool for opti-
mization of wing shape, actuator and four-bar geometry, as well
as a future dual-wing apping robotic platform. Although the
presented experimental prototype features a lift-to-weight ratio
of 1/6 (measured on a robot employing the same construc-
tion), scaled down variants of this design have demonstrated
liftoff [14]. Thus, given the direct scalability of the SCM man-
ufacturing technique and similarity of Reynolds number, the
presented here theoretical model and analysis is valid and use-
ful for scaled down robotic platforms capable of liftoff.
Compared to other wing rotation ideologies, a completely
passive rotation design benets from mechanical simplicity,
passive stability of wing trajectory, and low controller band-
width. However, the high dependence of wing dynamics on
aerodynamic forces renders the system as lacking in robustness
and consistent performance, requiring an adequate controller to
compensate for these design issues. Control over the amplitude
of the actuator driving waveform, as well as, the potential use of
smart materials in the wing rotation joint, is envisioned to enable
indirect control over the trajectory of the wing. The proposed
passive design is by no means seen as a replacement for active
wing control mechanisms, whose ability to generate arbitrary
forces and torques are essential for execution of their tasks, but
merely as an alternative apping platformthat may be realizable
on a shorter time scale for specic applications.
APPENDIX
FULL EQUATIONS OF MOTION FOR THE
PASSIVE ROTATION WING
Given the denition for the Lagrangian in (13) and perform-
ing the differentiations in (14) and (15), the full equations of
motion of wing motion are obtained. The equation governing
the rotation angle is the simplest of the two as it does not
involve inertial effects of the driving subsystem

(m
w
R
CG

CG
cos() + J
xz
cos())

2
cos()sin()(m
2
CG
+ J
yy
J
zz
)
+ (m
2
CG
+ J
xx
) + k
rot
=

M
aero


E

1
d

1
(20)
where all the constant denitions are presented in the main body
of this paper.
The equation governing the motion of is more complex as it
involves numerous inertial terms and parameterizations in terms
of . We need to remind ourselves that according to the notation
discussed in the text, we have the following:
f


f
(t)
, f



2
f
(t)
2
,

f
df
dt
,

f
d
2
f
dt
2
. (21)
Then, the equation of motion for becomes

sin
2
()J
yy
+ cos
2
()J
zz
+

R
2
CG
+

2
CG
2

1
2

2
cos(2)m +
3

i=1
m
Li
(x

2
Li
+ y

2
Li
) + m
e

2
+
2

i=1
J
Li

2
Li
+ J
L3

+ cos()(mR
CG

CG
+ J
xz
)

2
sin()(mR
CG

CG
+ J
xz
) +

sin(2)(m
2
CG
+ J
yy
J
zz
) +
3

i=1
m
Li

2
(x

Li
x

Li
+ y

Li
y

Li
)
+
2

i=1
J
Li

Li

Li

2
+ m
e

2
= (

M
drive
+

M
aero
)

E
3
D . (22)
The coupled equations for (t) and (t) are solved numeri-
cally with MATLABs ode15s solver. The obtained results are
presented in the main body of the paper.
ACKNOWLEDGMENT
The authors would like to thank R. Smith for his work on the
mean lift measuring setup, as well as prototype manufacturing
and testing.
REFERENCES
[1] A. Bergou, S. Xu, and Z. Wang, Passive wing pitch reversal in insect
ight, J. Fluid Mech., vol. 591, pp. 321337, 2007.
[2] R. B. Srygley, A. L. R. Thomas, Unconventional lift-generating mecha-
nism in free-ying butteries, Nature, vol. 420, pp. 660664, 2002.
[3] T. L. Hedrick and T. L. Daniel, Flight control in the hawkmoth manduca
sexta: The inverse problemof hovering, J. Exp. Biol., vol. 209, pp. 3114
3130, 2006.
[4] F. T. Muijres, L. C. Johansson, R. Bareld, M. Wolf, G. R. Spedding, and
A. Hedenstrom, Leading-edge vortex improves lift in slow-ying bats,
Science, vol. 319, pp. 12501253, 2008.
[5] S. Fry, R. Sayaman, and M. H. Dickinson, The aerodynamics of hovering
ight in drosophila, J. Exp. Biol., vol. 208, pp. 23032318, 2005.
798 IEEE/ASME TRANSACTIONS ON MECHATRONICS, VOL. 18, NO. 2, APRIL 2013
[6] J. Birch and M. H. Dickinson, The inuence of wing-wake interactions
on the production of aerodynamic forces in apping ight, J. Exp. Biol.,
vol. 206, pp. 22672272, 2003.
[7] M. Hamamoto, Y. Ohta, K. Hara, and T. Hisada, A fundamental study of
wing actuation for a 6-in-wingspan apping microaerial vehicle, IEEE
Trans. Robot., vol. 26, no. 2, pp. 244255, Apr. 2010.
[8] S. P. Sane and M. H. Dickinson, The aerodynamic effects of wing rotation
and a revised quasi-steady model of apping ight, J. Exp. Biol., vol. 205,
pp. 10871096, 2002.
[9] R. J. Wood, S. Avadhanula, M. Menon, and R. S. Fearing, Microrobotics
using composite materials: The micromechanical ying insect thorax, in
Proc. IEEE Int. Conf. Robot. Autom., 2003, pp. 18421849.
[10] S. Avadhanula and R. S. Fearing, Flexure design rules for carbon ber
microrobotic mechanisms, in Proc. Int. Conf. Robot. Autom., 2005,
pp. 15791584.
[11] B. Cheng and X. Deng, Translational and rotational damping of apping
ight and its dynamics and stability at hovering, IEEE Trans. Robot.,
vol. 27, no. 5, pp. 116, Oct. 2011.
[12] T. L. Hedrick, B. Cheng, and X. Deng, Wingbeat time and the scaling of
passive rotational damping in apping ight, Science, vol. 324, pp. 252
255, 2009.
[13] F. V. Breugel, W. Regan, and H. Lipson, From insects to machines:
Demonstration of a passively stable, untethered apping-hovering micro-
air vehicle, IEEE Robot. Autom. Mag., vol. 15, no. 4, pp. 6874, Dec.
2008.
[14] R. J. Wood, The rst takeoff of a biologically inspired at-scale robotic
insect, IEEE Trans. Robot., vol. 24, no. 2, pp. 341347, Apr. 2008.
[15] E. Steltz and R. Fearing, Dynamometer power output measurements
of miniature piezoelectric actuators, IEEE/ASME Trans. Mechatronics,
vol. 14, no. 1, pp. 110, Feb. 2009.
[16] S. N. Mahmoodi, N. Jalili, and M. Daqaq, Modeling, nonlinear dynamics,
and identication of a piezoelectrically actuated microcantilever sensor,
IEEE/ASME Trans. Mechatronics, vol. 13, no. 1, pp. 5865, Feb. 2008.
[17] R. J. Wood, E. Steltz, and R. S. Fearing, Optimal energy density piezo-
electric bending actuators, Sens. Actuators A, vol. 119, pp. 476488,
2005.
[18] M. Sitti, D. Campolo, J. Yan, R. S. Fearing, T. Su, D. Taylor, and T. Sands,
Development of PZT/PZN-PT unimorph actuators for micromechani-
cal apping mechanisms, in Proc. IEEE Robot. Autom. Conf., 2001,
pp. 38393846.
[19] M. Sitti, Piezoelectrically actuated four-bar mechanism with two exi-
ble links for micromechanical ying insect thorax, IEEE/ASME Trans.
Mechatronics, vol. 8, no. 1, pp. 2636, Mar. 2003.
[20] R. J. Wood, Liftoff of a 60mg apping-wing mav, in Proc. IEEE/RSJ
Int. Conf. Intell. Robots Syst., 2007, pp. 18891894.
[21] M. Karpelson, G. Wei, and R. J. Wood, Milligram-scale high-voltage
power electronics for piezoelectric microrobots, in Proc. Int. Conf. Robot.
Autom., 2009, pp. 22172224.
[22] R. J. Wood, E. Steltz, and R. Fearing, Nonlinear performance limits for
high energy density piezoelectric bending actuators, in Proc. IEEE/RSJ
Int. Conf. Intell. Robots Syst., 2005, pp. 36333640.
[23] S. Avadhanula, The design and fabrication of the MFI thorax based on op-
timal dynamics, Ph.D. dissertation, Dept. Mech. Eng., Univ. California,
Berkeley, CA, 2005.
[24] L. I. Sedov, Two-Dimensional Problems in Hydrodynamics and Aerody-
namics. New York: Interscience, 1965.
[25] M. H. Dickinson, F. Lehman, and S. P. Sane, Wing rotation and the aero-
dynamic basis of insect ight, Science, vol. 284, pp. 19541600, 1999.
[26] A. Andersen, U. Pesavento, and Z. J. Wang, Unsteady aerodynamic of
uttering and tumbling plates, J. Fluid Mech., vol. 541, pp. 6590, 2005.
[27] J. P. Whitney and R. J. Wood, Aeromechanics of passive rotation in
apping ight, J. Fluid Mech., vol. 660, pp. 197220, 2010.
[28] W. B. Dickson, A. D. Straw, C. Poelma, and M. H. Dickinson, An
integrative model of insect ight control, in Proc. AIAA Aerosp. Sci.
Meet. Exhib., 2006, pp. 119.
[29] R. J. Wood, Design, fabrication, and analysis of a 3 DOF, 3 cm apping-
wing mav, in Proc. 2007 IEEE/RSJ Int. Conf. Intell. Robots Syst.,
pp. 15761581.
[30] R. Dudley, The Biomechanics of Insect Flight: Form, Function and Evo-
lution. Princeton, NJ: Princeton Univ. Press, 1999.
[31] S. Montambault and C. M. Gosselin, Analysis of underactuated mechan-
ical grippers, J. Mech. Design, vol. 123, pp. 367375, 2001.
[32] R. Olfati-Saber, Nonlinear control of underactuated mechanical systems
with application to robotics and aerospace vehicles, Ph.D. dissertation,
Dept. Electr. Eng. Comput. Sci., Massachussets Inst. Technol., Cambridge,
MA, 2000.
[33] N. Perez-Arancibia, J. Whitney, and R. Wood, Lift force control of
apping-wing microrobots using adaptive feedforward cancelation sche-
mes, IEEE/ASME Trans. Mechatronics, [Online]. Available:
http://ieeexplore.ieee.org, DOI: 10.1109/TMECH.2011.2163317.
[34] X. Deng, L. Schenato, and S. Sastry, Flapping ight for biomimetic
robotic insects: Part II: Flight control design, IEEETrans. Robot., vol. 22,
no. 4, pp. 789803, Aug. 2006.
[35] P. Lotz, M. Matysek, and H. Schlaak, Fabrication and application of
miniaturized dielectric elastomer stack actuators, IEEE/ASME Trans.
Mechatronics, vol. 16, no. 1, pp. 5866, Feb. 2011.
[36] C. Zheng and T. Xiaobo, Acontrol-oriented and physics-based model for
ionic polymermetal composite actuators, IEEE/ASME Trans. Mecha-
tronics, vol. 13, no. 5, pp. 519529, Oct. 2008.
Veaceslav Arabagi received the B.Sc. degree from
the University of California, Berkeley, in 2006,
and the M.Sc. and Ph.D. degrees from Carnegie
Mellon University, Pittsburgh, PA, in 2008 and 2011,
respectively.
He is currently a Research Fellow at Childrens
Hospital Boston, Boston, MA, where he is involved
in the eld of surgical robotics. His research interests
include apping wing ight, small-scale design and
manufacturing techniques, and bioinspired robotics.
Dr. Arabagi placed rst in the 2011 ASME Grad-
uate Robot Design Competition for his work on the design of a apping wing
aerial robot.
Lindsey Hines received the B.Sc. degree in mechan-
ical engineering and the B.A. degree in mathematics
fromthe University of St. Thomas, Saint Paul, MN, in
2008, and the M.Sc in robotics fromCarnegie Mellon
University, Pittsburgh, PA, in 2011, where she is cur-
rently working toward the Ph.D. degree in robotics.
Her interests include apping ight and robust
control.
Ms. Hines has been awarded both the National Sci-
ence Foundation and National Defense Science and
Engineering Graduate Fellowship, and placed rst in
the 2011 Graduate Robot Design Competition.
Metin Sitti (S94A99M99SM08) received the
B.Sc. and M.Sc. degrees in electrical and electron-
ics engineering from Bogazici University, Istanbul,
Turkey, in 1992 and 1994, respectively, and the Ph.D.
degree in electrical engineering from The University
of Tokyo, Tokyo, Japan, in 1999.
He was a Research Scientist at the University of
California, Berkeley, from 1999 to 2002. He is cur-
rently a Professor in the Department of Mechanical
Engineering and Robotics Institute, Carnegie Mellon
University, Pittsburgh, PA, where he is also the Di-
rector of the NanoRobotics Laboratory and the Center for Bio-Robotics. He
was the Adamson Career Faculty Fellow during 20072010. His research in-
terests include micro/nanorobotics, bioinspired miniature mobile robots, and
micro/nanomanipulation.
Dr. Sitti received the Society of Optics and Photonics Nanoengineering Pi-
oneer Award in 2011. He received a National Science Foundation CAREER
Award in 2005. From 2008 to 2010, he was the Vice President of Technical Ac-
tivities of the IEEE Nanotechnology Council. He was elected as a Distinguished
Lecturer of the IEEE Robotics and Automation Society from 2006 to 2008. He
also received the Best Paper Award at the IEEE/RSJ International Conference on
Intelligent Robots and Systems in 2009 and 1998, the Best Biomimetics Paper
Award at the IEEE Robotics and Biomimetics Conference in 2004, and the Best
Video Award at the IEEE Robotics and Automation Conference in 2002. He is
Co-Editor-in-Chief of the Journal of Micro/Nano-Mechatronics.

Das könnte Ihnen auch gefallen