Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/268556965

System Identification of an Ornithopter Aerodynamics Model

Conference Paper · August 2010


DOI: 10.2514/6.2010-7632

CITATIONS READS
13 171

5 authors, including:

Jared Grauer Evan Ulrich


NASA Aerospace Corporation
75 PUBLICATIONS   626 CITATIONS    10 PUBLICATIONS   206 CITATIONS   

SEE PROFILE SEE PROFILE

James Edward Hubbard Jr. Darryll J Pines


University of Maryland, College Park University of Maryland, College Park
105 PUBLICATIONS   2,423 CITATIONS    191 PUBLICATIONS   4,125 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

guidance, navigation and control View project

micro air vehicle research View project

All content following this page was uploaded by Jared Grauer on 23 January 2021.

The user has requested enhancement of the downloaded file.


System Identification of an Ornithopter
Aerodynamics Model

Jared Grauer,∗ Evan Ulrich,† James Hubbard Jr.,‡ Darryll Pines,§ and J. Sean Humbert¶
Department of Aerospace Engineering, University of Maryland, College Park, MD, 20742

Flapping-wing ornithopters are emerging as a promising vehicle design for developing


small scale unmanned air vehicles with multi-mission capabilities. One of the primary
technical challenges currently impeding the maturation of flapping-wing vehicle systems
is the difficulty in obtaining simple but accurate models of the complex aerodynamic flow
fields which characterize flapping flight. Typically aerodynamic models are based on first
principles, numerical models, or time-averaged wind tunnel testing. System identification
provides an alternative that does not compromise the aerodynamic environment experi-
enced in free flight. In this paper the authors present model structure determination results
of lift and thrust forces generated by the flapping wings of an ornithopter. A visual track-
ing system was used to obtain state measurements of the ornithopter while in trimmed
straight and level mean flight. A multibody vehicle dynamics model, previous experimen-
tal results, model structure determination, and the equation-error parameter estimation
routine used in both the time and frequency domains were used to determine models for
the aerodynamic forces.

Nomenclature

a(p, v) aerodynamic forces α, β angle of attack and side-slip


B input coupling η, , δ orientation quaternion
C(p, v) coupling and damping matrix Φ kinematic Jacobian
C, K body centers and frames φ, s(φ) parameter estimates and standard errors
CL , CD lift and drag coefficients θ joint angles
G gear ratios ν translational velocity
g(p) gravitational forces τ generalized control forces
I identity matrix ω rotational velocity
Jη attitude Jacobian
K stiffness matrix Superscripts
T
l, r body fixed positions matrix transpose
−1
M(p) mass and inertia matrix matrix inverse
p position states ˙ time derivative
R rotation matrix ˆ estimated value
R2 coefficient of determination I ij
, frame expression
S(.) skew operator jω.̃ Fourier transform
r position
u pilot inputs Subscripts
v velocity states a actuator
z rotation axes b rigid body

∗ Graduate Student, Department of Aerospace Engineering, Member AIAA


† Graduate Student, Department of Aerospace Engineering, Member AIAA
‡ Langley Distinguished Professor, Department of Aerospace Engineering, Associate Fellow AIAA
§ Professor, Department of Aerospace Engineering, Fellow AIAA
¶ Assistant Professor, Department of Aerospace Engineering, Member AIAA

1 of 13

American Institute of Aeronautics and Astronautics


I. Introduction
nmanned aerial systems are progressively changing the world in which we live. Currently these vehicles
U are controlled by a remote pilot and can be used for tasks such as wildlife population monitoring, airport
wildlife control, remote sensing applications, and hobby flying. It is hoped that in the future these vehicles
will achieve full mission autonomy and will be able to independently complete tasks involving reconnaissance,
surveillance, crop surveying, weather observation, and chemical substance detection. Concept analysis and
demonstrations have shown that as the vehicle size decreases, for missions such as flying through indoor
or cluttered environments, flapping-wing aircraft enjoy advantages in agility and efficiency over similarly
sized fixed-wing and rotary-wing aircraft.1 It is envisioned that an ornithoptic flight vehicle will integrate
the agility and maneuverability needed for perching and indoor flight, as well as the flight efficiency and
robustness needed for outdoor flight, into a single and effective multi-mission capable flight platform.
A prerequisite for implementing autonomous flapping-wing vehicles is a thorough understanding of the
aerodynamics and flight dynamics, thereby facilitating the construction of efficient aircraft designs and the
synthesis of agile flight controllers. However for several reasons, these two topics remain areas of active
research with flapping-wing aircraft. The flapping motions create unsteady, complex, and three dimensional
aerodynamic flow fields2 that are at low Reynolds numbers and are difficult to model.3 The small vehicle
mass and inertia also introduces structural deflections and aeroelastic effects into the flight data, with which
some researchers have contended by integrating finite element codes into their flight models.4–6 Above a
certain vehicle size, the variable mass distribution of the vehicle and the fast motions arising from low inertia
dictate that accurate modeling of the flight dynamics requires multibody models, which add considerable
complexity to the dynamics model.7–9 Additionally the small payload capacity of these vehicles requires
sensors based on MEMS technology, which suffer from temperature dependencies, calibration drifts, and low
signal to noise ratios.
Despite these challenges, there have been several efforts to develop analytical, numerical, and experimental
models of flapping-wing aerodynamics. Analytical models typically involve blade element discretizations of
the wing and employ thin airfoil theory and quasi-steady aerodynamic forces.10, 11 These models often lack
experimental validation or represent drastic simplifications of the physical phenomenon experienced in free
flight. Computational fluid dynamics has also played an important role in illuminating the basic flight
mechanisms of flapping-wing flight.2, 12 Some experimental models determined from data exist for insect-
based flyers;13 however, most of the empirical results for avian-sized vehicles comes from time-averaged data
measured from a constrained force balance in a wind tunnel. This method constrains the natural motions
of the test vehicle and removes significant frequency content from the model.
This paper aims to provide a low order aerodynamic model for a specific ornithopter test vehicle, obtained
in free flight and useful for stability analysis and control synthesis. A flight test was performed using
a visual tracking system to obtain state measurements. Using a multibody flight dynamics model, the
aerodynamic forces were extracted and system identification techniques, implemented in a MATLAB toolbox
called SIDPAC,14 were used to determine their functional dependence on the state variables and to estimate
the model parameters. The paper begins with a description of the test vehicle, followed by the development
of the multiple rigid body dynamics model. Experimental results are presented for actuator dynamics, tail
aerodynamics, and then wing aerodynamics.

II. Test Vehicle Description


The modeling work presented herein was centered about a modified version of the “Slow Hawk” Kinkade
ornithopter,15 shown in Figure 1. This ornithopter was chosen for experimental work due to its robust
construction and relative ease of piloting. The wing sail is an arrangement of ripstop polyester segments,
fastened together with Dacron tape. Although a flexible membrane, the wing shape is maintained with a
primary carbon fiber spar spanning the leading edge, a secondary carbon fiber spar running from the aft
section of the ornithopter to the wing tip, and a set of small carbon fiber spars spread along the trailing
edge. The fuselage is cut from a 3.2-mm-thick carbon fiber sheet. The aircraft nose has been reinforced with
additional plies of carbon fiber to keep the vehicle robust to rough landings
The ornithopter is flown in the same manner as other remotely controlled vehicles: a pilot launches the
ornithopter and uses a radio frequency transmitter to relay joystick commands to an onboard receiver, which
then parses the commands to the individual actuators. The throttle input adjusts the speed of a DC motor

2 of 13

American Institute of Aeronautics and Astronautics


Figure 1. Ornithopter research platform

which is connected through a gear box and a four-bar linkage to the wings, thereby governing the flapping
frequency and generation of lift and thrust forces. The longitudinal and lateral inputs control the armature
position of servomotors, which pitch and roll the tail relative to the fuselage in a serial manner, creating
pitching and yawing aerodynamic control torques on the aircraft.
Experience conducting flight tests and laboratory experiments with this ornithopter has illuminated
several characteristics of its flight. Computer aided drawing software was used to model the aircraft geometry,
which resulted in a mass distribution where approximately 76% of the vehicle mass was affixed to the fuselage,
19% comprised wings and 5% comprised the tail. Additionally the center of mass migrates approximately
0.1 m over each wing stroke, and the mass distribution experiences dramatic changes in the primary pitching
and yawing moments of inertia. Constrained flapping experiments performed in a laboratory with a visual
tracking system have shown that for this particular ornithopter the bending moments of the leading edge
spar are relatively mild, whereas smaller ornithopters can have a pronounced bending of the wing.16 These
aircraft have a flapping bandwidth up to about 7 Hz, which results in forcing frequencies that are significantly
closer to the expected rigid body vehicle modes than are seen in conventional aircraft. Flight test experiments
using a custom avionics package have shown that this aircraft experiences heave accelerations of 46.12 m/s2
and pitch rates of 5.62 rad/s.17

III. Rigid Body Dynamics


The ornithopter body dynamics are modeled using a multiple body, or multibody, model. While this
model is more complex than conventionally used for aircraft flight dynamics, it is necessary to capture the
large amplitude and fast motions observed in flight, as well as the configuration dependent mass distribution
of the ornithopter.
The multibody configuration is shown in Figure 2 and consists of seven rigid bodies approximating the
fuselage, each wing, two components of the tail mechanism, and two components creating a four-bar linkage
between the fuselage and wings. Each rigid body has a coordinate frame Kij placed at its local center of mass
Cij , where subscripts ij are used to denote each rigid body, where i is the linkage number and j is the chain
number. Body fixed vectors lij and rij denote the locations of the in-board and out-board revolution joint
positions relative to the local center of mass, which pass through axis vectors zij and z(i+1)j , respectfully.
Angles θij indicate the rotation of the body ij about the axis zij , relative to its adjacent in-board body.
Additionally an inertial reference frame is placed on the surface of the Earth.
The state variables to describe the ornithopter flight kinematics were chosen to parallel conventions
used in both the aircraft and robotics literature: the ornithopter is viewed as a central fuselage body with

3 of 13

American Institute of Aeronautics and Astronautics


C12
z24 C11
C24 z14
l11 θ12
r24 z12
C14 z11
C0 θ11
l14 r04 r01
θ24 r14
θ14 r0,I

CI

Figure 2. Multibody ornithopter assembly

articulated joints. The generalized position state variables are


 
rI0,I
 I
p =  η0,I (1)


θ
I
where rI0,I is the position of the fuselage center of mass, expressed in the inertial frame, η0,I = [ T δ ]T
T
is a unit quaternion describing the orientation of the fuselage, and θ = [ θ13 θ14 θ24 ] is a vector
concatenating the articulated joint angles of the system. The wing motion is admitted by a four-bar linkage
mechanism, which can be described by a single degree of freedom. The crank angle θ13 was chosen to
parametrize the wing motion because although choosing the wing angle θ11 or θ12 would facilitate simpler
expression for describing the aerodynamic forces on the wings, numerical problems arise when integrating
those equations of motion.
The generalized velocity states of the ornithopter are
 
0
ν0,I
 0 
v =  ω0,I  (2)
θ̇
0 0
where ν0,I and ω0,I are the translational and rotational velocities of the fuselage, expressed in the fuselage
frame, and where θ̇ are the joint rates.
The kinematic equations of motion describe how the position states evolve in time as a function of the
velocity states. As the position states are expressed in different reference frames than the velocity states,
the equations admit a non-flat relationship
    
ṙI0,I RI,0 0 0 0
ν0,I
 I    0 
 η̇0,I  =  0 Jη 0   ω0,I  (3)
θ̇ 0 0 I θ̇
or more succinctly as
ṗ = Φv (4)
where the rotation matrix and attitude Jacobian matrix are defined as
RI,0 = (δ 2 − T )I + 2T − 2δS() (5)

4 of 13

American Institute of Aeronautics and Astronautics


" #
1 δI − S()
Jη = (6)
2 −T
and where S(.) is a matrix skew operator.
The dynamic equations of motion describe the evolution of the velocity states as a function of the vehicle
position, velocity, and control. Using the Boltzmann-Hamel equations, the dynamic equations of motion
may be cast into the form8
Mb (p)v̇ + Cb (p, v)v + g(p) + a(p, v) = τ (7)
which is a canonical form in the nonlinear control of Euler-Lagrange systems such as spacecraft and robotic
manipulators.18, 19 The matrix Mb (p) describes the generalized mass and inertia properties of the multibody
system, and is dependent on the joint angles of the system. The second term describes coupling forces arising
from Coriolis and centripetal accelerations, as well as forming the equations of motion in a moving reference
frame. Vectors g(p), a(p, v), and τ represent the generalized forces due to gravitational accelerations,
aerodynamic forces, and exogenous forcings on the velocity states.

IV. System Identification Results


IV.A. Actuator Dynamics
The ornithopter has onboard three actuators, consisting of one Feigao brushless motor and two Hitec
servo motors. Dynamic models for these systems were identified for two reasons. First, inclusion of the
actuator dynamics increases the model fidelity and the accuracy to which other forces, such as the wing
aerodynamic forces, are identified. Second, including actuator dynamics facilitates the design of control laws
for the pilot input u, instead of the more abstract joint torque τ , admitting control laws that can be directly
implemented in hardware.
An experiment was performed to determine the model structure and estimate model parameters for these
two types of actuators.20 The actuators were constrained to a torque cell, instrumented with position and rate
sensors, and subjected to a train of pulse commands. It was found that these devices were well-approximated
by linear, second order differential equations of the form

Ma v̇ + Ca v + Ka p + Gτ = Bu (8)

where Ma describes rotary inertia; Ca describes linear damping arising from back electro-mechanical forces,
viscous friction, and electrical resistance; Ka is the synthetic stiffness injected by a proportional control law
in the servo motors; G is a matrix of gear ratios; τ are the torques applied to the joints; and B is an input
coupling matrix.
Model parameters in (8) were estimated from the data using the frequency domain equation-error method
and the time domain output-error method.14 Model fits for the DC motor and servo motor are shown for
both methods in Figure 4. In the frequency domain estimation, a frequency grid using intervals of 0.01 Hz
were used on the range where there was appreciable frequency content. The output-error method converged
after 28 iterations for the DC motor, and after 38 iterations for the servo motor. The DC motor results had
model fits of R2 = 0.89 and R2 = 0.98 for the frequency domain equation-error and time domain output-
error analysis, whereas the servo motor results had model first of R2 = 0.96 and R2 = 0.98, respectively.
Estimated model parameters are listed in Table 1 along with standard errors, corrected for colored residuals.
In general there was good agreement between the two methods and overall, the frequency domain equation-
error method produced smaller error bounds. Based on the standard errors of these two methods, a weighted
average was computed and used for the remainder of the work.
The rigid body equations (7) can be substituted for the joint torque in the actuator dynamics (8) to yield
the equation
M(p)v̇ + C(p, v)v + w(p, v) = Bu (9)
where
M(p) = GMb + Ma
C(p, v) = GCb + Ca (10)
w(p, v) = Gg + Ga + Ka p
which couples the rigid body equations with the actuator equations.

5 of 13

American Institute of Aeronautics and Astronautics


(a) DC motor (b) servo motor

Figure 3. Experimental setup for actuator system identification

Table 1. DC motor and servo motor parameter estimates and standard errors

Actuator Parameter Freq EE Time OE Weighted Average Multiplier


φ φ̂ + s(φ̂) φ̂ + s(φ̂) φ̂ + s(φ̂)
Ma +0.2265 ± 0.0051 +0.2207 ± 0.2149 +0.2264 ± 0.0070 10−3
DC Motor Ca +0.8861 ± 0.0224 +1.0069 ± 0.2477 +0.8961 ± 0.0291 10−3
B −0.5838 ± 0.0130 −0.6316 ± 0.3100 −0.5857 ± 0.0177 10+2
Ma +0.0255 ± 0.0009 +0.0425 ± 0.0006 +0.0279 ± 0.0001 10−3
Servo Motor Ca +0.1554 ± 0.0006 +0.6493 ± 0.0090 +0.1854 ± 0.0008 10−3
B +2.1976 ± 0.0319 +8.8359 ± 2.3396 +2.2869 ± 0.0445 10−3

6 of 13

American Institute of Aeronautics and Astronautics


10000 -2000
data
model

q̇ (rad/s)
-3000
jω q̃

5000

-4000
0
0 2 4 6 0 5 10 15 20

frequency (Hz) time (s)

(a) DC motor

400 1
data
model
q (rad)

200 0

0 -1
0 4 8 12 0 5 10

frequency (Hz) time (s)

(b) servo motor

Figure 4. Actuator model fits using frequency domain equation-error and time domain output-error

7 of 13

American Institute of Aeronautics and Astronautics


IV.B. Tail Aerodynamics
To separate the aerodynamic contributions of the wings from those of the tail, a wind tunnel test was
performed.21 The ornithopter fuselage and tail were mounted onto a six degree of freedom load cell and
placed in the test section of a free-jet wind tunnel. The wind tunnel was run at speeds between 0 and 12
m/s while the ornithopter tail was cycled through its full range of motion. Measurements of the lift and
drag forces, as well as the tail deflections angles were made at 500 Hz.
In post processing, lift and drag measurements were reduced to dimensionless coefficients. A multivariate
orthogonal function algorithm was used to choose the best model for the lift and drag coefficients based on
regressors including (potentially nonlinear) combinations of angle of attack and sideslip angle. It was found
that the model equations

CL = CL0 + CLα α + CLαβ αβ + CLα3 α3 + CLα2 β2 α2 β 2 (11)

CD = CD0 + CDα2 α2 (12)


obtained fits to the time domain data that minimized the predicted square error, a metric used to choose
the number of regressors retained in the model.14
Parameter estimates were performed in both the time domain and the frequency domain using the
equation-error method and resulted in two sets of parameter estimates, given in Table 2. The frequency
domain estimates are shown as a response surface in Figure 5. Additionally the two parameter sets were
averaged, based on the inverse of the standard errors, to arrive at a final set of parameters and associated
statistics. Both the frequency domain method and the time domain method results had relatively small
standard errors and predicted parameters fairly consistently. Models had fits of R2 = 0.98 and R2 = 0.99
for the time-domain lift and drag fits, and R2 = 0.92 and R2 = 0.97 for the frequency domain lift and drag
fits, respectively.

-0.2 0.4

-0.4
0.2
CL CD
-0.6
model model
-0.8 data 0 data

-0.2 -1
-0.4 -0.8
0.8 -0.6 α (rad) -0.8 -0.4 -0.6 α (rad)
0.4 β (rad) 0 -0.8 β (rad) 0 -0.4
-0.4 -1 0.4 -0.2
-0.8 0.8

(a) lift coefficient (b) drag coefficient

Figure 5. Tail aerodynamic model fits using frequency domain equation-error

Table 2. Tail aerodynamic parameters using equation-error

Parameter Time EE Freq EE Weighted Average


φ φ̂ ± s(φ̂) φ̂ ± s(φ̂) φ̂ ± s(φ̂)
CL0 +0.1219 ± 0.0227 +0.0403 ± 0.0084 +0.0623 ± 0.0087
CLα +1.1996 ± 0.0566 +0.9147 ± 0.2314 +1.1436 ± 0.0643
CLαβ −0.1874 ± 0.0130 −0.2215 ± 0.0090 −0.2075 ± 0.0075
CLα3 −0.5409 ± 0.0466 −0.2231 ± 0.1279 −0.4560 ± 0.0483
CLα2 β2 +0.2657 ± 0.0341 +0.2261 ± 0.0324 +0.2454 ± 0.0235
CD0 +0.0784 ± 0.0034 +0.1009 ± 0.0053 +0.0872 ± 0.0029
CDα2 +0.3404 ± 0.0080 +0.2863 ± 0.0462 +0.3325 ± 0.0096

8 of 13

American Institute of Aeronautics and Astronautics


IV.C. Wing Aerodynamics
The only term in (9) that does not have a known functional dependence are the aerodynamic contributions
a(p, v). This term can be isolated as

a(p, v) = G−1 [Bu − Mv̇ − Cv − Gg − Ka p] (13)

by algebraic manipulations. The aerodynamic contributions are assumed generated by two wings and a tail
with no interactions. The left hand side of (13) can be subdivided into contributions from the tail, given
by equations (11) and (12), and by the wings. The only remaining unknown term are the aerodynamic
contributions from the wings. By performing a flight test and obtaining state measurements, substitutions
can be made into the right hand side of (13) to yield time histories of the aerodynamic forces and moments
on the ornithopter. It is then a system identification problem to determine the functional dependence of
how those time histories depend on the states of the ornithopter, and the value of any parameters within
the model.
A flight test was performed where the ornithopter was flown indoors while trimmed for straight and level
mean flight. A Vicon visual tracking system,22 consisting of eight cameras, was used to estimate at 500 Hz
the spatial position of several retro-reflective markers placed on the ornithopter. From the marker positions,
position and velocity state vectors were computed by employing a nonlinear least squares estimator to fit
rigid bodies to the marker locations and extracting out the states. Figure 6(a) displays a photograph taken
of the capture volume and Figure 6(b) shows the rigid body estimation results for a single capture frame.
The data from this experiment was substituted into (13) to produce aerodynamic forces on the wings.
Model structure determination, using a modified step-wise regression algorithm,14 was used to determine the
functional dependence of the lift and drag forces on the ornithopter states. In the past orthogonal regressors23
and a quasi-steady aerodynamics model24 have been used to model the lift coefficient. The method of using
orthogonal regressors resulted in a relatively large model structure, and the quasi-steady model resulted in a
model needing acceleration terms, which in turn requires additional states in the dynamics model.25, 26 In the
current study, only the ornithopter state variables were considered as model regressors. Additionally, step-
wise regression was performed in the frequency domain instead of the time domain to explore the spectral
content of the data.
The model structure for lift and drag were chosen as

CL = CL0 + CLνx νx + CLνz νz + CLωy ωy + CLθ̇ θ̇11 (14)


11

2 2
CD = CD0 + CDθ̇ θ̇11 + CDθ2 θ11 + CDθ̇2 θ̇11 (15)
11 11 11

which can be found from the ornithopter state vector and requires no additional states in the model. Param-
eter estimates were determined in both the time domain and the frequency domain using the equation-error
method, and are shown with residuals in Figure 7. As the frequency domain method subtracts out the bias
and trend of the data before the estimation, a second time-domain estimation was required to find the lift
and drag bias coefficients. The lift estimation had a fit of R2 = 0.84 in the time domain and R2 = 0.61 in
the frequency domain, whereas the drag estimation had a fit of R2 = 0.88 in the time domain and R2 = 0.64
in the frequency domain. These model fits are more accurate than those previously reported for system
identification results with this ornithopter,23, 24 and significantly better than the analytical models27 and
computational fluid dynamics work12 attempted previously. Estimation model parameters with standard
error bounds corrected for colored residuals are displayed in Table 3. With the exception of the coefficient
CLνx , the two methods provide consistent estimates of the model parameters. Generally it is the frequency
domain method that produced the estimates with lower standard errors.

V. Conclusions
Ornithopters are flapping-wing vehicle which have the potential to not only accomplish standard missions
for unmanned aircraft such as way-point navigation, but also multi-mission capabilities such as robust
outdoor flight, agile indoor flight, and precision perching at a variety of speeds and sizes. To realize an
autonomous ornithopter that could fulfill this potential, a simple and accurate model for the aerodynamic
forces must be developed to program into autopilots uniquely tailored for flapping-wing flight dynamics. In

9 of 13

American Institute of Aeronautics and Astronautics


(a) visual tracking system and test vehicle, outfitted with
retro-reflective markers

-1.1

-1
fuselage
-0.9
right wing
-0.8 left wing
tail

3.6

3.8 -0.6
4 -0.3

4.2 0

0.3
4.4
0.6

(b) nonlinear rigid body fitting from retro-reflective marker


data

Figure 6. Experimental setup for wing aerodynamics system identification

10 of 13

American Institute of Aeronautics and Astronautics


0.14
data
1
lift coefficient

model

0.07
0

-1 0

0.07
1
residual

0
0

-1 -0.07
0 0.2 0.4 0.6 0 10 20 30 40 50

time (s) frequency (Hz)

(a) lift coefficient

0.4 0.04
data
drag coefficient

model
0
0.02
-0.4

-0.8 0

0.4 0.02

0
residual

0
-0.4

-0.8 -0.02
0 0.2 0.4 0.6 0 10 20 30 40 50

time (s) frequency (Hz)

(b) drag coefficient

Figure 7. Wing aerodynamic model fits using equation-error in the time and frequency domains

11 of 13

American Institute of Aeronautics and Astronautics


Table 3. Wing aerodynamic parameters using equation-error

Parameter Time EE Freq EE Weighted Average


φ φ̂ ± s(φ̂) φ̂ ± s(φ̂) φ̂ ± s(φ̂)
CL0 −1.8935 ± 1.9391 −2.8556 ± 0.2325 −2.7526 ± 0.2936
CLνx +0.2333 ± 0.2107 +3.1314 ± 0.0669 +2.4330 ± 0.0718
CLνz +0.0346 ± 0.0953 +0.1233 ± 0.0075 +0.1168 ± 0.0098
CLωy +0.2775 ± 0.1051 +0.3389 ± 0.0054 +0.3359 ± 0.0073
CLθ̇ +0.0296 ± 0.0043 +0.0320 ± 0.0004 +0.0318 ± 0.0005
11

CD0 −0.0538 ± 0.0673 −0.0272 ± 0.0099 −0.0306 ± 0.0122


CDθ̇ −0.0072 ± 0.0009 −0.0070 ± 0.0001 −0.0070 ± 0.0001
11

CDθ2 +0.4455 ± 0.1488 +0.4410 ± 0.0146 +0.4414 ± 0.0188


11

CDθ̇2 −0.0013 ± 0.0003 −0.0015 ± 0.0000 −0.0015 ± 0.0000


11

the past this task has been accomplished using first principles modeling, numerical simulations, and time-
averaged wind tunnel data. Each of these methods introduces either drastic simplifications of aerodynamics,
or changes the flow environment experienced by the ornithopter during testing.
In this paper, aerodynamic models are determined using standard system identification techniques applied
to free flight data. Past experience and analysis with the ornithopter has led to the development of a
multiple rigid body model of the flight dynamics. System identification methods were used to identify
models for actuator dynamics in static testing environments, and aerodynamics of the tail in a wind tunnel.
Incorporating these models facilitated the system identification of the wing aerodynamics from a free flight
test. The flight test was conducted indoors using a Vicon visual tracking system to obtain spatial positions
of retro-reflective markers placed on the ornithopter, from which state measurements were extracted. System
identification was then performed to determine the model structure of the lift and drag forces on the wing,
and to estimate the model parameters.
In contrast to previous system identification work with this ornithopter, only the state vector was con-
sidered for model regressors, which led to a simpler model structure and ultimately more accurate results,
particularly for the drag model, than previously reported. Good models fits were obtained and parameter
estimates were generally consistent between the two methods of parameter estimation used. An accurate, low
order model of the wing aerodynamics was presented, suitable for stability analysis and control development
and hardware implementation.

VI. Acknowledgments
This work was funded in part by the Army MAST CTA and the NASA Langley NIA Langley Professor
Program. Flight tests were conducted at the University of Maryland. Conversations with Eugene Morelli
at the NASA Langley Research Center are acknowledged and appreciated. The authors would like to also
acknowledge support from the NASA Langley Research Center, the National Institute of Aerospace, and the
University of Maryland in conducting this research, and would like to thank the members of the Morpheus
Laboratory and the Autonomous Vehicle Laboratory for their continued teamwork and motivation.

References
1 Seshadri, P., Benedict, M., and Chopra, I., “Experimental Investigation of an Insect-based Flapping-Wing Hovering
Micro Air Vehicle,” Aeromechanicanics Specialists’ Conference, American Helicopter Society, 2010.
2 Shyy, W., Lian, Y., Tang, J., Viieru, D., and Liu, H., Aerodynamics of Low Reynolds Number Flyers, Cambridge

Aerospace Series, Cambridge University Press, 2007.


3 Mueller, T. and DeLaurier, J., “An Overview of Micro Air Vehicle Aerodynamics,” Fixed and Flapping Wing Aerody-

namics for Micro Air Vehicle Applications, edited by T. Mueller, Vol. 195 of Progress in Aeronautics and Astronautics, 2001,
pp. 1–9.
4 Larijani, R. and DeLaurier, J., “A Nonlinear Aeroelastic Model for the Study of Flapping Wing Flight,” Fixed and

Flapping Wing Aerodynamics for Micro Air Vehicle Applications, edited by T. Mueller, Vol. 195 of Progress in Aeronautics
and Astronautics, 2001, pp. 399–428.

12 of 13

American Institute of Aeronautics and Astronautics


5 Sibilski,
K., “Dynamics of Micro-Air-Vehicle with Flapping Wings,” Acta Polytechna, Vol. 44, No. 2, 2004, pp. 15–21.
6 Isogai,
K. and Harino, Y., “Optimum Aeroelastic Design of a Flapping Wing,” Journal of Aircraft, Vol. 44, No. 6, 2007,
pp. 2040–2048.
7 Rashid, T., The Flight Dynamics of a Full-Scale Ornithopter , Master’s thesis, University of Toronto, 1995.
8 Grauer, J. and Hubbard, J., “Multibody Model of an Ornithopter,” Journal of Guidance, Control, and Dynamics,

Vol. 32, No. 5, Sep.-Oct. 2009, pp. 1675–1679.


9 Bolender, M., “Rigid Multi-Body Equations-of-Motion for Flapping Wing MAVs using Kane’s Equations,” No. AIAA-

2009-6158, 2009.
10 Dietl, J. and Garcia, E., “Stability in Ornithopter Longitudinal Flight Dynamics,” Journal of Guidance, Control, and

Dynamics, Vol. 31, No. 4, July-August 2008, pp. 1157–1162.


11 DeLaurier, J., “An Aerodynamic Model for Flapping Wing flight,” The Aeronautical Journal of the Royal Aeronautical

Society, , No. 1853, April 1993, pp. 125–130.


12 Roget, B., Sitaraman, J., Harmon, R., Grauer, J., Hubbard, J., and Humbert, S., “Computational Study of Flexible

Wing Ornithopter Flight,” Journal of Aircraft, Vol. 46, No. 6, Dec 2009, pp. 2016–2031.
13 Sane, S. and Dickinson, M., “The Aerodynamic Effects of Wing Rotation and a Revised Quasi-Steady Model of Flapping

Flight,” Journal of Experimental Biology, Vol. 205, No. 8, April 2002, pp. 1087–1096.
14 Klein, V. and Morelli, E., Aircraft System Idenfication: Theory and Practice, AIAA Education Series, AIAA, 2006.
15 Kinkade, S., “Hobby Technik,” www.flappingflight.com, 2008.
16 Harmon, R., Aerodynamic Modeling of a Flapping Membrane Using Motion Tracking Experiments, Master’s thesis,

University of Maryland, College Park, MD, 2008.


17 Grauer, J. and Hubbard, J., “Inertial Measurements from Flight Data of a Flapping-Wing Ornithopter,” Journal of

Guidance, Control, and Dynamics, Vol. 32, No. 1, Jan.-Feb. 2009, pp. 326–331.
18 Slotine, J. and Li, W., Applied Nonlinear Control, Prentice Hall, 1991.
19 Ortega, R. and Loria, A., Passivity-Based Control of Euler-Lagrange Systems, Springer, 1998.
20 Grauer, J. and Hubbard, J., “Identification and Integration of Ornithopter Actuator Models,” No. AIAA-2009-5937,

2009.
21 Grauer, J. and Hubbard, J., “Experimental Determination of Ornithopter Control Derivatives,” Region 1 Mid-Atlantic

Student Conference, AIAA, Hampton, VA, April 2007.


22 Vicon, “Vicon Motion Tracking Systems,” www.vicon.com, 2009.
23 Grauer, J., Ulrich, E., Hubbard, J., Pines, D., and Humbert, S., “Model Structure Determination of an Ornithopter

Aerodynamics Model from Flight Data,” No. AIAA-2010-41, 2010.


24 Grauer, J. and Hubbard, J., “Modeling of Ornithopter Flight Dynamics for State Estimation and Control,” No. WeA15.2,

American Controls Conference, 2010.


25 Klein, V. and Noderer, K., “Modeling of Aircraft Unsteady Aerodynamic Characteristics,” Tech. Rep. 109120, National

Aeronautics and Space Administration, 1994.


26 Jouannet, C. and Krus, P., “Unsteady Aerodynamic Modelling: A Simple State-Space Approach,” No. AIAA-2005-855,

2005.
27 Harmon, R., Grauer, J., Hubbard, J., and Humbert, S., “Experimental Determination of Ornithopter Membrane Wing

Shapes Used for Simple Aerodynamic Modeling,” No. AIAA Paper 2008-6397, August 2008.

13 of 13

American Institute of Aeronautics and Astronautics

View publication stats

You might also like