You are on page 1of 20

Stability Analysis of Longitudinal Dynamics of

Hovering Flapping MAVs/Insects

Ahmed Elsadek
Zewail City of Science and Technology, Giza, 12588, Egypt

G. El-Bayoumi
Cairo University, Giza, 12613, Egypt
Haithem E. Taha
University of California Irvine, Irvine, CA, 92697, USA

Flapping-Wing Micro-Air-Vehicles (FWMAVs) are complex systems whose concept is


inspired from biological flyers in nature. The aim of this work is to investigate the sta-
bility of the hovering state in FWMAVs and insects. Flight Dynamics of FWMAVs and
insects constitutes a nonlinear time periodic system with unconventional contributors to
the aerodynamics of flight. In this work, an aerodynamic-body-dynamic model that cap-
tures the dominant physical aspects in the flow field is derived. Full simulations are carried
out showing the inherent instability in the system dynamics. The periodic equilibrium
of the nonlinear time periodic system is captured using the optimized shooting method.
Furthermore, Floquet theorem is used to assess the stability of the captured periodic or-
bit corresponding to hovering equilibrium. The results showed that the system is indeed
unstable.

Nomenclature

AR Aspect Ratio
CD Drag Coefficient
CL Lift Coefficient
H0 Horizontal aerodynamic force
Imn Second moment of area
Kmn Second moment of area multiplied by constants
M General moment force
Mh Hinge moment force
MAF Moment force resulting from asymmetric flapping
R Semi-span of the wing
S Surface area of one wing
T Period in seconds
U Velocity of air relative to wing section
V0 Vertical aerodynamic force
X0 Tilted X force after inclusion of i
Z0 Tilted Z force after inclusion of i
c Airfoil chord length
d Drag force per unit span
f Flapping frequency
Research Assistant, Center of Communications and Aerospace.
Professor, Aerospace Engineering Department.
Assistant Professor, Mechanical and Aerospace Engineering.

1 of 20

American Institute of Aeronautics and Astronautics


l Lift force per unit span
m Mass of the aircraft
r Span-wise distance from wing root
wef f Effective vertical velocity component in Zb direction
X Total force in xb direction
u Velocity component of Cg in xb direction
xh Distance between hinge line and center of gravity
Symbols
Angle of attack
i Induced angle of attack
m Mean angle of attack
Geometric angle of attack
c Mean chord
r Fixed wing Yaw rate
x Distance between hinge line and center of pressure normalized to
the chord
Pitching angle
rot 2D rotational circulation of potential flow
r1 , r2 Moments of wing chord distribution
x0 Position of the pitch axis from the leading edge normalized to the
chord length
Fast flapping wing time scale
Flapping amplitude
Air density
Normalized time ( number of cycles ) / Body motion time scale
Viscous rotational damping torque per unit span
Pitching angle
Flapping angle

I. Introduction
Flight of birds and insects has been always an inspiration to mankind to pursue the dream of flight. It
has been a research topic of interest for biologists and engineers who seek understanding of the underlying
physics and the development of bio-inspired aerial vehicles. The aerodynamics of flapping flight is unsteady
with unconventional mechanisms that contribute to lift generation.1 The wings motion gives the system a
periodic nature which means that the aerodynamic loads are changing periodically with time.2 In this work,
we are only interested in the longitudinal dynamics of the hovering state. The insect used for simulation
purposes is the Hawkmoth, a full description of its morphological parameters is illustrated in appendix A.
Wings motion during hover can be considered to be horizontal. The dynamics of such system is highly
nonlinear, which means that it is very sensitive to initial conditions.
Stability analysis of NLTP systems can be approached using two tools, the averaging theory and Floquet
theory. The first approach is based on the assumption that the body will only feel the average of the cyclic
loads because of the high frequency.3 It has been shown that the first order averaging is not sufficient to
capture the dynamics of the system. The second approach is using Floquet theory which entails finding the
periodic orbit that is mainly an optimization problem.

Optimized Shooting Method4 uses Levenberg Marquardt Algorithm (LMA) as its optimization engine.
LMA is a hybrid optimization scheme which uses Gradient descent to ensure that the solution will converge
regardless of the initial position and Gauss Newton method when near the optimum to enhance speed. The
optimized shooting method uses a residue vector to be minimized using LMA. The residue vector elements
are defined according to the concept of periodicity. This means, for example, that every point in the first
cycle has to be equal to the corresponding points in the other cycles to be in a periodic equilibrium. The
residue vector elements are the difference between each point and its counterpart in the other cycle. The
optimization algorithm is to minimize the residue vector and the solution will be the initial conditions of the

2 of 20

American Institute of Aeronautics and Astronautics


system.

In this work, an aerodynamic-dynamic model is derived, the periodic equilibrium of the system at hand
is captured using the optimized shooting method. Finally, the stability of the periodic orbit is assessed using
Flqouet theory.

II. Aerodynamic-Dynamic Model


A. Dynamic Model Setup
The interest here is to study the longitudinal flight dynamics of hovering Flapping Micro Aerical Vehicles
(FWMAVs) and insects. Rigid body dynamics is considered in this framework while neglecting the inertial
effects of the wing motion. The assumption of neglecting the wing inertial effects is very well known in the
literature, see [514, ]. Usually, researchers use this assumption because the wings mass is very small com-
pared to the bodys and neglecting the multi-body effects results in equations of motion similar to those of
conventional aircraft. Yet, there still a huge difference between flapping and conventional aircraft dynamics,
the first is time varying as a consequence of the periodic motion of the wings. Two frames of reference are
used, the body axes xb , yb , zb and the wing axes xw , yw , zw . The wing fixed frame is placed so that it coincide
with the body axes. Further more, during hover, insects almost flap their wings in a horizontal plane. Which
means that we can assume that the wing motion is restricted to flapping with an angle . On the other
hand, because the dynamics is restricted to the longitudinal motion, only one body rotation is considered
around the pitch axis with an angle .

For these previous settings, the equations of motion of fixed wing aircraft can be considered.15 Recall
the X, Z, M forces and equations

X mg sin = m(u + qw rv)


Z + mg cos cos = m(w + pv qu)
(1)
M = Iy q + rq(Ix Iz ) + Ixz (p2 r2 )
= q cos r sin

where, X, Z are the forces on the center of gravity in xb and zb directions respectively. M is the force
moment around the pitch axis yb . m is the mass of the insect.

Forces, rates and angles that are out of the longitudinal plane can be equal zero. Rewriting the equations
for u,
w, Now, we have the nonlinear equations of the longitudinal motion written in vector form as
q and .
follows
1X


u q w g sin() m
1
w q u + g cos() m Z

= (2)
q 0 1 M
Iy
q
0
This equation can be written in the vector form = f () + ga (, t). Where is the state vector.

3 of 20

American Institute of Aeronautics and Astronautics


Figure 1: Model of flapping insect during hover

B. Aerodynamic Model
The aerodynamics of flapping flight is characterized by non-conventional mechanisms that augment lift force.
The model adopted here captures two important phenomena, Leading Edge Vortex (generated by motion
of the wing at high angle of attack) and Rotational Lift. As stated by Berman and Wang,16 the static
lift coefficient induced by LEV can be modeled by the following formula. A is constant the is determined
experimentally. Taha and Hajj17 stated that the value of A can be fit by equation(4)
CL = Asin(2) (3)

AR
A= q (4)
2
2[1 + ( AR
a0 + 1)]
According to Sane and Dickinson,18 the 2D rotational circulation potential flow can be used to model
the rotational lift as follows
 
2 3
rot = c x0 (5)
4
where is the pitching angular velocity, c is the airfoil chord length, and x
0 is the position of the pitch
axis from the leading edge normalized to the chord length. LEV is a very prominent in delta wings.2
Consequently, its drag coefficient equation can be a good description for the drag coefficient in the current
model

CD = CL tan() = 2A sin2 () (6)


Now, we can formulate the lift and drag forces at a general distance r per unit span using the preceding
representations
 
1 2 3
l(r,t) = Ac(r)U (r, t)sin2(r, t) + x0 U (r, t)c2 (r)(t)
(7)
2 4

d(r,t) = Ac(r)U 2 (r, t) sin2 (r, t) (8)


where, is the air density and U is the velocity of air relative to the wing section.

4 of 20

American Institute of Aeronautics and Astronautics


C. Effect of Body Motion
The latter representation of the forces on the wing is only generated by the motion of the wing itself. In
real hover flight, the body wobbles as the insect flaps its wings. Because our interest here is to study the
stability of the system, the inclusion of the body motion is a must. This movement will affect the angle of
the attack and hence the forces.

The velocity of a wing section is written as r.


As the body moves forward, the wing moves forward.
This motion of the wing induced by the body motion adds another component of the velocity on the wing
section which is u cos . This means that the airfoil velocity is
p
U = ( + u cos )2 + w (9)
The body also rotates with a rate q around its pitch axis which makes becomes + q cos . As the body
wobbles, it changes the position of the wing. Namely, it produces an induced angle of attack as shown in
figure 2. Now, we have

= + i (10)
Where is the angle of attack of the wing that changes instantaneously with the forward and backward
strokes as follows
(
U > 0 forward stroke
= (11)
U < 0 backward stroke

Figure 2: Induced angle of attack as a result of body motion

i is the induced angle of attack.


 w  For any airfoil the angle of attack can be written as a function of the

free stream flow as = tan 1 . Here we have another component that contributes to the value of the
U
vertical velocity over a wing section.

Consider a wing section that is located at a position r on the wing, the distance from this position to
the hinge line is r sin as in figure 3 . Adding to that the distance between the hinge line and the cg
which equals to xh . Therefore the distance between the cg and the wing section is r sin() + xh . If the cg
is pitching with a rate q around the yb axis, then the wing section is moving upwards with a velocity of
q(r sin() + xh ). Because this velocity is upwards it will have a negative sign and will be subtracted from
the velocity component w. This makes the total vertical velocity of the wing section is w r sin() + xh .

Accordingly, the induced angle of attack will be in the form


 
w q(rsin + xh )
i = tan1 (12)
U
Let wef f = w r sin() + xh . The induced angle of attack i is assumed to be small. This means that
wef f
higher orders of i will be equal zero and i u .
|U |
Expanding the expression of the angle of attack

5 of 20

American Institute of Aeronautics and Astronautics


Figure 3: Vertical velocity of the wing section during hover

sin 2 = sin 2(i + ) = sin 2 cos 2i + cos 2 sin 2i


because i is very small

sin 2 = sin 2 + 2i cos 2


The lift and drag forces can be written again as a function of the new angle of attack.
 
1 2 wef f (, t; r)
l(, r; t) = Ac(r)U (r, t) sin 2 (t) + 2 cos 2 (t)
2 |U (, t; r)|
  (13)
3
+ 0 U (r, t)c2 (r)[(t)
x + q cos (t)]
4

d(, r; t) = Ac(r)U 2 (, t; r) sin2 ( + i )

sin2 ( + i ) = (sin cos i + cos sin i )2


= (sin + i cos )2
= sin2 + 2i sin cos + i2 cos2
= sin2 + i sin 2
wef f
= sin2 + sin 2
|U |
 
wef f (, t; r)
d(, t; r) = Ac(r)U 2 (, t; r) sin2 + sin(2 ) (14)
|U (, t; r)|
The induced angle of attack i will undoubtedly affect the direction of lift and drag forces as shown in
figure 4. The X and Z forces can now be written as

X 0 = sgn(U )H 0
(15)
Z 0 = V 0

where, H 0 is the horizontal aerodynamic force per unit span opposing instantaneous velocity and V 0 is the vertical aerodyna

The values of V 0 and H 0 can be expressed in terms of lift and drag

H 0 = d cos i l sin i u d l i
(16)
V 0 = l cos i + d sin i u l + d i

6 of 20

American Institute of Aeronautics and Astronautics


Figure 4: Resolution of forces on the wing section as a result of the induced angle of attack

Now the equation of X 0 becomes

X 0 = cos (d li )
     
wef f wef f 1 wef f
= cos AcU 2 sin2 + sin2 AcU 2 sin2 + 2 cos 2
|U | |U | 2 |U |
  
3
+ x 0 c2 U ( + q cos )
4 (17)
" !#
2
1 2 1 wef f
= cos AcU (2U sin () 2sin(2)wef f ) AcU sin(2)wef f + 2 cos 2
2 2 |U |
 
3
wef f x0 c2 ( + q cos )
4
   
0 1 3
X = cos AcU (2U sin2 + sin(2)wef f ) wef f c2 x
0 ( + q cos ) (18)
2 4
As for the Z 0

Z 0 = V 0
= (l + di )
    
1 2 wef f 3 2
= AcU sin 2 + 2 cos 2 + x
0 c U ( + q cos )
2 |U | 4
 
wef f wef f
+ AcU 2 sin2 + sin 2
|U | |U |
   
1 2 2 3
= AcU (U sin 2 + 2wef f (cos 2 + sin )) + c U x 0 ( + q cos )
2 4

cos 2 + sin2 = cos2 , the Z 0 equation becomes


   
0 1 2 2 3
Z = AcU (U sin 2 + 2 cos wef f ) + c U x
0 ( + q cos ) (19)
2 4
The moment equation will consist of four effects, the moment acting on the hinge that is generated by
the integration of the vertical force on the wing, force on the hinge with an arm of x + h, the effect of wing
rotation and the asymmetric flapping effect. The moment equation can be written as

M (, t; r) = Mh0 cos Z 0 xh + MAF (20)

Mh0 = (V 0 cos + sgn(U )H 0 sin )


x c(r) (21)

0
MAF = Z 0 r sin (22)

7 of 20

American Institute of Aeronautics and Astronautics


where, Mh0 is the hinge moment, MAF is the asymmetric flapping, is the viscous rotational damping
torque per unit span, and x is the distance between the hinge line and the center of pressure normalized
to the chord
c
The center of pressure can be considered to be located at . On the other hand, can be written using
4

the equation stated by Berman and Wang16 = ( 16 )c4 (1 f + 2 ||)
1 and 2 depend on the fluid and
is recommended by Berman and Wang to take a value of 0.2.
The total pitching moment per unit span will be,

M 0 (, t; r) = Z 0 (, t; r)[
xc(r) cos (t)(t) + xh + r sin (t)]
(23)
+ X 0 (, t; r) sin (t)
xc(r) + (, t; r) cos

D. Aerodynamic Dynamic Coupling


Finally, for the forces and moment equations to be fully described in terms of the system parameters and
the system states, and because only first order terms of the state vector are considered, the velocity of
wing section can be approximated as U u r + u cos


0 1
X = cos Ac(r + u cos )[2(r + u cos ) sin2
2
  (24)
3 2
+ (w q(r sin + xh )) sin 2] x
0 c [w q(r sin + xh )]
4

Neglecting higher order terms

1
X 0 = Ac[2r2 2 cos sin2 + (4r cos2 sin2 )u
2
+ (r sin 2 cos )w r cos sin 2(r sin + xh )q]
   
2 3 2 3
+ c cos x0 w c cos x0 q
4 4

X 0 = Acr2 2 sin2 (2Acr cos2 sin2 )u


  
1 3
Acr sin 2 cos c2 x 0 w
2 4 (25)
   
1 3
+ Acr cos sin 2(r sin + xh ) c2 cos x
0 (r sin + xh ) q
2 4


0 1
Z = Ac(r + u cos )[(r + u cos ) sin 2 + 2 cos2 (w q(r sin + xh ))]
2
  
2 3
+ c (r + u cos ) x 0 ( + q cos )
4

Neglecting higher order terms of the states,

 
0 1 2 2 2 3
Z = Acr sin 2 + c x 0
2 4
  
2 3
Acr cos sin 2 + c cos x
0 u
4 (26)
Acr cos2 w
 
  
2 3
Acr cos (r sin + xh ) + c r cos x
0 q
4

8 of 20

American Institute of Aeronautics and Astronautics


The moment equation becomes

1
M 0 = (cos cos xc + xh + r sin ) Acr2 2 sin 2
2
 
3
+ c2 r x 0 Ac2 r2 2 x sin3 + cos
4
   
3
+ ( cos cos xc xh r sin ) Acr 2 sin 2 + c2 cos x0
4

+ 2Ac2 r cos2 sin3 x u

1
+ ( cos cos xc xh r sin )(Acr cos2 ) + Ac2 r sin 2 sin x cos (27)
2
 
3
c3 sin x x0 w
4
  
3
+ ( cos cos xc xh r sin )(Acr cos (rsin + xh )) c2 r cos x 0
4
1
Ac2 r cos sin 2 sin x(r sin + xh )
2   
3
c3 sin x cos x 0 (r sin + xh ) q
4

9 of 20

American Institute of Aeronautics and Astronautics


These equations are to be integrated over the semi-span(R) of the wing and multiplied times two for the
two wings. Wing pitch angle can be assumed to vary with a piecewise constant manner, which means that
= 0. Now we can substitute the equations of X 0 , Z 0 and M 0 in the first equation to have a model that
couples the aerodynamics tightly with the dynamics of the system
1 0
mX

u qw g sin 1
0
w qu + g cos m Z

= + (28)
q 0 1 M 0
Iy


q
0

1
u(t)
q(t)w(t) g sin (t) m X0 (t) Xu (t) Xw (t) Xq (t) 0 u(t)
1 Z (t)
q(t)u(t) + g cos (t) m o Zu (t) Zw (t) Zq (t)
w(t) 0 w(t)

= + 1 + (29)
Iy Mo (t) Mu (t) Mw (t) Mq (t)

q(t)
0 0 q(t)

(t) q(t) 0 0 0 0 0 (t)

The stability derivatives are,

X0 = 2K21 | sin2
|
Z0 = K21 | | sin 2
 
M0 = 2|
|
sin K22 x cos + K21 xh cos + K31 sin cos
K11
Xu = 4 cos2 sin2
||
m
K11
Xw = ||
cos sin 2
m
K21
Xq = ||
sin cos sin 2 xh Xw
m
Zu = 2Xw
K11
Zw = 2 cos2
m
K21 Krot12
Zq =2 sin cos2 cos xh Zw
m m
K12 x m
Mu =4 cos2 sin (2Xq xh Zu )
Iy Iy
K12 x K21 mxh
Mw =2 cos cos + 2 sin cos2
|| Zw
Iy Iy Iy
2x
Mq = ||
cos cos (K12 xh + K22 sin )
Iy
1
+ cos (Krot13 x cos cos + Krot22 sin )
Iy
2 K 1 f mxh
|| cos2 sin (K21 xh + K31 sin ) cos2 Zq
Iy Iy Iy
Z R
where, Imn =2 rm cn (r)dr
0
1 1
Kmn = AImn Krotmn = (
x)Imn K = I04
2 2 16

III. Equilibria and Optima


A. Optimized Shooting Method
Optimized shooting method is a numerical method proposed by Botha and Dednam as an extension to
the general shooting method to solve for the periodic solutions of both autonomous and non-autonomous

10 of 20

American Institute of Aeronautics and Astronautics


nonlinear systems. This method is based upon Levenberg Marquart Algorithm LMA which is a non-linear
least squares optimization scheme. The main idea of this method is the minimization of the residue vector,
which is the difference between a point at a specific T + and a point at . Consider the nonlinear
dynamical system

x = f (x, , t) (30)
where x, f Rn . This system is a non-autonomous system because it depends explicitly on time.
For the system to be periodic
x(t) = x(t + T ) t 0 (31)
19
where T is the period (T > 0). LMA is a method for solving nonlinear least squares problems. To know
how it works, suppose that it is desired to fit a function y(t; p) to a set of m data points (ti , yi ). Here, the
independent variable is t while p is a vector of n parameters. For this problem it is necessary to minimize
the sum of the weighted squares of the errors between the measured data and the curve fit function

" m
#2
2
X y(ti ) y(ti ; p)
(p) =
i=1
wi (32)
= y Wy 2y T W y + yT W y
T

where W is the diagonal weighting matrix. Based on the gradient descent method, the perturbation h that
moves the parameters in the direction of steepest descent is,

hgd = J T (y y) (33)

where J is the Jacobian matrix and is the step size. In the same manner it can be shown that the
Gauss-Newton perturbation is given by,

[J T W J]hgn = J T W (y y) (34)

Since LMA adaptively varies the parameter updates between Gradient Descent and Gauss Newton methods.
It can be seen that the resulting perturbation is given as

[J T W J + I]hlm = J T W (y y) (35)
The optimized shooting method can be applied to any system that can be expressed in the form of (30).
In the original work, Botha normalized the systems used in the paper to the time as the period of the system
was unknown.

x = T f (x, , T ) (36)
The new variable allows the simplification of the boundary conditions in equation equatino (31). = 1
means a full cycle.
In our case, the flapping frequency is known to be 26.5 for the Hawkmoth. Which means that the period
is T = 0.0380 seconds.
The residual can be written as Z 1
R=T f (x, , t) (37)
0
Furthermore, the number of quantities to be optimized are of great effect on the residual which can be
expressed as,

R = (x(1) x(0), x(1 + ) x( ), ..., x(1 + (p 1) ) x((p 1) )) (38)

where is the integration step size. Botha chose the step size to be = 2 10. The same step size is
used in this work. The natural number p in the residual equation is a requirement of the LMA and has to be
chosen so that the number of components of the residual is greater than or equal to the number of quantities
to be optimized. The main goal now is to minimize the residue vector to get the right initial conditions that
will put the system in the periodic equilibrium.

11 of 20

American Institute of Aeronautics and Astronautics


IV. Results
The main problem lies in capturing the periodic orbit that ensures hovering. This means that it is needed
to ensure that the vertical speed of the center of gravity equals to zero. This can be achieved in several ways.
In this work, it was done by introducing a new state h = wsin() ucos(), where h is the vertical speed.
The new virtual state will be treated as if it was a system state, which means that it will be included in the
initial guess vector and in the residual vector. Furthermore, the flapping amplitude will be added to the
initial guess vector as its value to ensure hovering is not known yet.
The system is subjected to a harmonic variation of the flapping angle

(t) = cos(2f t) (39)

with a variation of the angle of attack

mgT 2
 
1
m = sin 1 (40)
2 2 AI21 2

The optimization algorithm will minimize the variation of h along with other elements of the residue
vector. For example, if three elements are chosen for each state, the residue vector will be as follows

u(1) u(0)
u(1 + ) u( )


u(1 + 2 ) u(2 )


w(1) w(0)

w(1 + ) w( )


w(1 + 2 ) w(2 )


q(1) q(0)

R = q(1 + ) q( ) (41)


q(1 + 2 ) q(2 )


(1) (0)

(1 + ) ( )


(1 + 2 ) (2 )


h(1) h(0)

h(1 + ) h( )

h(1 + 2 ) h(2 )

Starting the problem using arbitrary initial guess in table 1

Table 1: Initial guess for the optimization problem

State Initial Guess


u 0.1
w 0.1
q 0.1 * / 180
0.1 * / 180
h 2

The periodic orbit found is represented in the state space (u, w, q) is the periodic solution for the problem
given the previous initial conditions as shown in figure 5 and the states versus number of cycles in figure
6. The required flapping amplitude required to achieve hovering according the optimization results is

12 of 20

American Institute of Aeronautics and Astronautics


= 61.0107 degrees with a variation of zero in h which ensures that it is indeed the hovering periodic orbit.
Initial conditions of this periodic orbit are demonstrated in table 2

Table 2: Results of the optimization problem. Integrating the system using these initial conditions for one
cycle will give a periodic orbit in the state space

State Orbit Initial Condition


u -0.083375
w 1.777
q -1.1677
6.3345

15

10

5
q (m/sec)

10

15
1.82
1.8 0.1
1.78 0.05
0
1.76 0.05
1.74 0.1
w (m/sec) u (m/sec)

Figure 5: Periodic orbit in the three dimensional state space

13 of 20

American Institute of Aeronautics and Astronautics


0.1 1.82

0.05 1.8

w (m/sec)
u (m/sec)
0 1.78

0.05 1.76

0.1 1.74
0 0.5 1 1.5 0 0.5 1 1.5
No. of cycles No. of cycles
20 6.336

10 6.335
q (m/sec)

0 6.334

h m
10 6.333

20 6.332
0 0.5 1 1.5 0 0.5 1 1.5
No. of cycles No. of cycles

Figure 6: System states plotted for one cycle using the results from the optimization problem. It is shown
that all of them are periodic oscillations

Stability Analysis
Stability of Linear Time Periodic (LTP) systems can be assessed using Floquet Theory.20 Looking at
Floquet multipliers of small systems of ODEs. These multipliers are the eigenvalues of the monodromy
matrix which is the solution at t = T for the variational equation

d(t) F
= (t) (42)
dt x x(t)

where
m11 m12 m13 m14
m m22 m23 m24

(t) = 21 (43)
m31 m32 m33 m34
m41 m42 m43 m44
where (t) is the state transition matrix and (0) is the identity matrix. The initial condition is the
result of the optimization problem in the last chapter. Using this initial condition will ensure that the system
is in the periodic orbit and the values of Floquet multipliers will not change. One of the Floquet multipliers
is 1 for autonomous and nonautonomous systems. This multiplier is called the trivial Floquet multiplier
which has eigenvectors tangent to the periodic solution at the initial point x(0). The asymptotic stability
of the solution can be assessed by checking whether the other Floquet multipliers are less than one. The
periodic orbit is said to be asymptotically unstable if at least one of the Floquet multipliers lie outside the
unit circle in the complex plane.
Recall the system in equation, (29), calculating its Jacobian

F f () ga (, t)
= + (44)

where

14 of 20

American Institute of Aeronautics and Astronautics



0 q w g cos()
f () q

0 u g sin()

= (45)


0 0 0 0
0 0 1 0
and

Xu (t) Xw (t) Xq (t) 0
ga Z (t) Zw (t) Zq (t) 0

= u (46)


Mu (t) Mw (t) Mq (t) 0
0 0 0 0
Now equation (42) becomes

0 q w g cos() Xu (t) Xw (t) Xq (t) 0
d(t) q 0

u g sin() Zu (t) Zw (t) Zq (t) 0

= + (t) (47)
dt 0 0 0 0 Mu (t) Mw (t) Mq (t) 0
0 0 1 0 0 0 0 0
Equation (42) is solved for monodromy matrix. The new system states of M is added to the state vector
. The simulation is carried out using the results of the optimization problem and the identity matrix for
Floquet multipliers as follows

Table 3: Initial conditions for periodic orbit

State Orbit Initial Condition


u -0.083375
w 1.777
q -1.1677
6.3345


1 0 0 0
0 1 0 0

i = (48)
0 0 1 0
0 0 0 1
The new linearized states are added to the old states (u, w, q, ). The new system to be solved contains
20 states. The new states added to the system of equations are

m11 m12 m13 m14
d(t) m m22 m23 m24

= 21 (49)
dt

m31 m32 m33 m34
m41 m42 m43 m44
where

2
4K11 m11 ||cos()
sin()2
m11 = m21 q gm41 cos()
m
  (50)
cos()(Krot12 sin()
K21 ||sin(2)sin())

+ m31 w
m

15 of 20

American Institute of Aeronautics and Astronautics


2
4K11 m12 ||cos()
sin()2
m12 = m22 q gm42 cos()
m
  (51)
cos()(Krot12 sin()
K21 ||sin(2)sin())

+ m32 w
m

2
4K11 m13 ||cos()
sin()2
m13 = m23 q gm43 cos()
m
  (52)
cos()(Krot12 sin()
K21 ||sin(2)sin())

+ m33 w
m

2
4K11 m14 ||cos()
sin()2
m14 = m24 q gm44 cos()
m
  (53)
cos()(Krot12 sin()
K21 ||sin(2)sin())

+ m34 w
m
 
Krot10,2 cos()sin(2)

m21 = m11 q +
m
 2
 (54)
Krot12 cos()
+ 2K21 ||cos()
sin()
gm41 sin() + m31 u +
m
 
Krot10,2 cos()sin(2)

m22 = m12 q +
m
 2
 (55)
Krot12 cos()
+ 2K21 ||cos()
sin()
gm42 sin() + m32 u +
m
 
Krot10,2 cos()sin(2)

m23 = m13 q +
m
 2
 (56)
Krot12 cos()
+ 2K21 ||cos()
sin()
gm43 sin() + m33 u +
m
 
Krot10,2 cos()sin(2)

m24 = m14 q +
m
 2
 (57)
Krot12 cos()
+ 2K21 ||cos()
sin()
gm44 sin() + m34 u +
m

1
m31 =
Iy
[m11 (cos()[Krot
10,2 Xh ] + dxKrot10,3 cos()cos() + Krot1,2 sin()

+ 4||cos()sin()(dxK
1,2 cos() + cos()(K1,1 Xh + K2,1 sin())))
(58)
+ m31 ( 2||cos()sin()(dxK
2,2 cos() + cos()(K2,1 Xh + K3,1 sin()))
+ 2cos()(Kv (f 1 + 2 ||))cos()
+ (Krot
1,2 Xh + dxKrot1,3 cos()cos()

+ K2,2 sin())) + (dxKrot


1,3 cos()sin() Kv 2 cos() sgn( + qcos()))]
2

1
m32 =
Iy
[m12 (cos()[Krot
10,2 Xh ] + dxKrot10,3 cos()cos() + Krot1,2 sin()

+ 4||cos()sin()(dxK
1,2 cos() + cos()(K1,1 Xh + K2,1 sin())))
(59)
+ m32 ( 2||cos()sin()(dxK
2,2 cos() + cos()(K2,1 Xh + K3,1 sin()))
+ 2cos()(Kv (f 1 + 2 ||))cos()
+ (Krot
1,2 Xh + dxKrot1,3 cos()cos()

+ K2,2 sin())) + (dxKrot


1,3 cos()sin() Kv 2 cos() sgn( + qcos()))]
2

16 of 20

American Institute of Aeronautics and Astronautics


1
m33 =
Iy
[m13 (cos()[Krot
10,2 Xh ] + dxKrot10,3 cos()cos() + Krot1,2 sin()

+ 4||cos()sin()(dxK
1,2 cos() + cos()(K1,1 Xh + K2,1 sin())))
(60)
+ m33 ( 2||cos()sin()(dxK
2,2 cos() + cos()(K2,1 Xh + K3,1 sin()))
+ 2cos()(Kv (f 1 + 2 ||))cos()
+ (Krot
1,2 Xh + dxKrot1,3 cos()cos()

+ K2,2 sin())) + (dxKrot


1,3 cos()sin() Kv 2 cos() sgn( + qcos()))]
2

1
m34 =
Iy
[m14 (cos()[Krot
10,2 Xh ] + dxKrot10,3 cos()cos() + Krot1,2 sin()

+ 4||cos()sin()(dxK
1,2 cos() + cos()(K1,1 Xh + K2,1 sin())))
(61)
+ m34 ( 2||cos()sin()(dxK
2,2 cos() + cos()(K2,1 Xh + K3,1 sin()))
+ 2cos()(Kv (f 1 + 2 ||))cos()
+ (Krot
1,2 Xh + dxKrot1,3 cos()cos()

+ K2,2 sin())) + (dxKrot


1,3 cos()sin() Kv 2 cos() sgn( + qcos()))]
2

m41 = m31 (62)

m42 = m32 (63)

m43 = m33 (64)

m44 = m34 (65)


Solving the system of equations and calculating the state transition matrix at t = T , the monodromy
matrix becomes

1 0.0007 0.0497 0.3725
0.001 1 0.0027 0.0195

M = (t=T ) = (66)
0.0001 0.0027 0.5933 0.0001
0 0.0195 0.0001 1
The eigenvalues of equation (66) are the Floquet multipliers

0.5933 + 0i
1.0258 + 0i

= (67)
0.9871 + 0.0108i
0.9871 0.0108i
Looking at the Floquet multipliers (eigenvalues) of the system, it is found that the second value is greater
than one which implies that the system is unstable. To make it clearer the Floquet multipliers are plotted
versus the unit circle in the complex plane in figure 7. It is shown that second Floquet multiplier is outside
the unit circle.

17 of 20

American Institute of Aeronautics and Astronautics


2
Floquet Multiplier 1
Floquet Multiplier 2
Floquet Multiplier 3
Floquet Multiplier 4
1.5 Unit Circle

0.5

Imaginary axis
0

0.5

1.5

2
2 1.5 1 0.5 0 0.5 1 1.5 2
Real axis

Figure 7: Floquet multipliers in the complex plane with respect to the unit circle. The system is unstable
as one eigenvalue lies outside the circle

V. Conclusion
An aerodynamic-dynamic model that represents the flight dynamics of FWMAVs was derived. The sta-
bility derivatives are represented as a function of system parameters. Simulations of the dynamics were
carried out and showed inherent instability in the system dynamics. The Optimized Shooting Method was
used to capture the periodic orbits of Lorenz and Rossler systems to show its effectiveness. The periodic
orbit of the FWMAV system was captured using the same method. Finally, the system stability was assessed
using Floquet theory and showed that indeed the system is unstable.

The optimization problem showed high sensitivity to initial conditions which is intuitive because of the
nonlinear nature of the problem. Moreover, the integration time step was very crucial to the stability of
the solution. A very small time step (210 )was used to ensure convergence. Furthermore, the periodic orbit
initial conditions are used to integrate the system over only one cycle. An important observation was that
using more than one cycle may lead to false results as the system states will leave the orbit. This is due to the
unstable nature of the periodic orbits captured. The stability of the periodic equilibrium was assessed using
Floquet theory. The position of floquet multipliers with respect to the unit circle showed that the system
at hand is in fact unstable. A deeper look into the dynamics of FWMAVs is needed. The current model
maybe expanded to include the wings inertial effects. Better analytical aerodynamic models are needed to
simulate and capture the true physics underlying the flow field around flapping flyers.

Appendix
A. Hawkmoth Morphological Parameters
The morphological parameters and the wing planform for the hawkmoth, as given in21 and,22 are

R = 51.9mm, S = 947.8mm2 , c = 18.3mm,


r1 = 0.44, r2 = 0.525, f = 26.3Hz, = 60.5 ,
mb = 1.648gm, and Iyb = 2080mg.cm2 ,

where R is the semi-span of the wing, S is the area of one wing, c is the mean chord, f is the flapping
frequency, is the flapping angle amplitude, mb is the mass of the body, and Iyb is the body moment of

18 of 20

American Institute of Aeronautics and Astronautics


inertia around the body y-axis. The moments of the wing chord distribution r1 and r2 are defined as
Z R
Ik1 = 2 rk c(r) dr = 2SRk rkk .
0

As for the wing planform, the method of moments used by Ellington22 is adopted here to obtain a chord
distribution for the insect that matches the documented first two moments r1 and r2 ; that is,

c  r 1  r 1
c(r) = 1 ,
R R
where h i h i
r1 (1 r1 ) r1 (1 r1 )
= r1 r22
r2
1 , = (1 r1 ) r22
r12
1 ,
R 1 11
and = 0 r (1 r)1 d
r.
The mass of one wing is taken as 5.7% of the body mass according to Wu et al. ? and is assumed uniform
with an areal mass distribution m0 The inertial properties of the wing are then estimated as
RR RR
Ix = 2 0
m0 r2 c(r) dr , Iy = 2 0
m0 d2 c3 (r) dr
R
m0 rc(r) dr
R
2
, Iz = Ix + Iy , and rcg = 0
mw = I2S11
,

where d is the chord-normalized distance from the wing hinge line to the center of gravity line.

References
1 Sane, S. P., The Aerodynamics of Insect Flight, Journal of Experimental Biology, Vol. 206, 2003, pp. 41914208.
2 Dickinson, M. H., Lehmann, F.-O., and Sane, S. P., Wing rotation and the aerodynamic basis of insect flight. Science,
Vol. 284, No. 5422, 1999, pp. 19541960.
3 Taha, H. E., Tahmasian, S., Woolsey, C. A., Nayfeh, A. H., and Hajj, M. R., The need for higher-order averaging in the

stability analysis of hovering, flapping-wing flight, Bioinspiration & biomimetics, Vol. 10, No. 1, 2015, pp. 016002.
4 Dednam, W. and Botha, A., Optimized shooting method for finding periodic orbits of nonlinear dynamical systems,

Engineering with Computers, Vol. 31, No. 4, 2015, pp. 749762.


5 Thomas, A. L. R. and Taylor, G. K., Animal flight dynamics I. Stability in gliding flight. Journal of Theoretical Biology,

Vol. 212, No. 1, 2001, pp. 399424.


6 Khan, Z. A. and Agrawal, S. K., Control of Longitudinal Flight Dynamics of a Flapping Wing Micro Air Vehicle Using

Time Averaged Model and Differential Flatness Based Controller, IEEE American Control Conference, 2007, pp. 52845289.
7 Sun, M. and Xiong, Y., Dynamic flight stability of a hovering bumblebee. Journal of Experimental Biology, Vol. 208,

No. 3, 2005, pp. 447459.


8 Xiong, Y. and Sun, M., Dynamic Flight Stability of a Bumble Bee in Forward Flight, Acta Mechanica Sinica, Vol. 24,

No. 3, 2008, pp. 2536.


9 Dietl, J. M. and Garcia, E., Stability in Ornithopter Longitudinal Flight Dynamics, Journal of Guidance, Control and

Dynamics, Vol. 31, No. 4, 2008, pp. 11571162.


10 Schenato, L., Campolo, D., and Sastry, S. S., Controllability Issues in Flapping Flight for Biomemetic MAVs, Vol. 6,

42nd IEEE conference on Decision and Control, 2003, pp. 64416447.


11 Deng, X., Schenato, L., Wu, W. C., and Sastry, S. S., Flapping Flight for Biomemetic Robotic Insects: Part II Flight

Control Design, IEEE Transactions on Robotics, Vol. 22, No. 4, 2006, pp. 789803.
12 Doman, D. B., Oppenheimer, M. W., and Sigthorsson, D. O., Wingbeat Shape Modulation for Flapping-Wing Micro-

Air-Vehicle Control During Hover, Journal of Guidance, Control and Dynamics, Vol. 33, No. 3, 2010, pp. 724739.
13 Oppenheimer, M. W., Doman, D. B., and Sigthorsson, D. O., Dynamics and Control of a Biomimetic Vehicle Using

Biased Wingbeat Forcing Functions, Journal Guidance, Control and Dynamics, Vol. 34, No. 1, 2011, pp. 204217.
14 Taha, H. E., Hajj, M. R., and Nayfeh, A. H., On the Longitudinal Flight Dynamics of Hovering MAVs/Insects, Journal

of Guidance Control and Dynamics, In Press.


15 Cook, M., V., Flight Dynamics Principles, Elsevier Ltd., 2007.
16 Berman, G. J. and Wang, Z. J., Energy-minimizing kinematics in hovering insect flight, Journal of Fluid Mechanics,

Vol. 582, No. 1, 2007, pp. 153,168.


17 Taha, H. E., Hajj, M. R., and Beran, P. S., Unsteady Nonlinear Aerodynamics of Hovering MAVs/Insects, AIAA-Paper

2013-0504, Jan 2013.


18 Sane, S. P. and Dickinson, M. H., The aerodynamic effects of wing rotation and a revised quasi-steady model of flapping

flight. Journal of Experimental Biology, Vol. 205, 2002, pp. 10871096.


19 Madsen, K., Nielsen, H. B., and Tingleff, O., Methods for non-linear least squares problems, 2004.
20 Lust, K., Improved Numerical Floquet Multipliers, International Journal of Bifurcation and Chaos in Applied Sciences

and Engineering, Vol. 11.

19 of 20

American Institute of Aeronautics and Astronautics


21 Sun, M., Wang, J., and Xiong, Y., Dynamic Flight Stability of Hovering Insects, Acta Mechanica Sinica, Vol. 23, No. 3,

2007, pp. 231246.


22 Ellington, C. P., The aerodynamics of hovering insect flight: II. Morphological parameters, Philosophical Transactions

Royal Society London Series B , Vol. 305, 1984, pp. 1740.

20 of 20

American Institute of Aeronautics and Astronautics

You might also like