You are on page 1of 20

J Intell Robot Syst (2012) 65:417435

DOI 10.1007/s10846-011-9644-7
Design and Analysis of a Gyroscopically Controlled
Micro Air Vehicle
Chris E. Thorne Mark Yim
Received: 18 February 2011 / Accepted: 22 November 2011 / Published online: 3 December 2011
Springer Science+Business Media B.V. 2011
Abstract Micro air vehicles have emerged as a
popular option for diverse robotic and teleoper-
ated applications in both open terrain and urban
environments because of their inherent stealth
and portability. To perform many of the tasks
envisioned for micro air vehicles, agility is essen-
tial. To date, research efforts to improve agility
have focused primarily on constructing complex
controllers to enable existing vertical-take-off-
and-landing vehicles, such as remote-controlled
helicopters and quadrotors, to perform aerobatic
maneuvers autonomously. In this work, we adopt
a system-level perspective and analyze a new
design for a rotary-wing micro air vehicle that
utilizes gyroscopic dynamics for attitude control.
Unlike traditional vehicles where attitude control
moments are generated by aerodynamic control
surfaces, the proposed vehicle will leverage the
existing angular momentum of its counter rotat-
ing components. This paradigm has the poten-
tial to yield significant increases in agility when
compared to state-of-the-art micro vertical take-
off and landing vehicles. The proposed design
C. E. Thorne (B) M. Yim
Mechanical Engineering and Applied Mechanics
Department, University of Pennsylvania,
Philadelphia, USA
e-mail: ecthorne@seas.upenn.edu
M. Yim
e-mail: yim@grasp.upenn.edu
reduces mechanical complexity by precluding the
use of complex mechanisms, such as the swash-
plate. The capacity to rapidly generate large
gyroscopic control moments, coupled with the
precision gained from eliminating the need for
complex and restrictive aerodynamic models, im-
proves both agility and adaptability. We present
the development of a gyroscopically controlled
micro air vehicle including comprehensive mod-
els of the dynamics and the aerodynamics with
an emphasis on the design and analysis of such
systems. A dynamics simulator that incorporates
these models and mechanical hardware solutions
to challenges that arose during prototyping will
also be presented.
Keywords Micro air vehicle Gyroscopic
control Vertical take-off and landing
Control moment gyroscope
1 Introduction
The development of Micro Air Vehicles (MAVs),
an effort introduced by the Defense Advanced
Research Projects Agency (DARPA) just two
decades ago, promises to increase the safety and
reliability of military operations by providing ro-
bust support to personnel in the field. Character-
ized as flying machines with no dimension larger
than 100 cm, a weight of up to 1 kg, and a desired
418 J Intell Robot Syst (2012) 65:417435
flight endurance of over 60 min, MAVs are des-
ignated as Class 1 Unmanned Aircraft Systems
(UASs) by the military. As such, they are required
to be backpackable in standard Modular Light-
weight Load-carrying Equipment (MOLLE) with
a flight endurance of at least 15 min.
Several inherent benefits make MAVs an at-
tractive choice for diverse applications in both
open terrain and complex urban environments
where a direct line-of-sight is not always possi-
ble. A low radar cross section and low acoustic
signature increase the ability to maintain covert
positions; a low weight enables transportation
and deployment by a single person with mini-
mal support, and less material and small-scale
manufacturing lead to lower costs and increase
the feasibility of swarms applications. In addition
to stealth, portability, and disposability, a large
thrust-to-weight ratio creates the potential for ex-
treme agility. These characteristics, coupled with
the designed-in ability to carry small payloads and
current miniaturization of many sensing technolo-
gies (e.g. cameras, wireless network links, etc),
make MAVs a versatile platform for numerous
military and civilian tasks including reconnais-
sance, surveillance, search and rescue, traffic mon-
itoring, inspection, biological and chemical agent
detection, aerial photography, manipulation, and
communication relaying.
In this work, we present the Dysc, a new MAV
with counter-rotating rotors that uses gyroscopic
moments for attitude control (Fig. 1). The two
rotors are fixed pitch, eliminating the need for
complex mechanisms like the swashplate, leading
to lower cost and easier maintenance when com-
Fig. 1 Dysc CAD model
pared to traditional helicopter configurations. In
addition, we showed in [12] that the Dysc has the
potential for higher agility than quadrotors.
2 Related Work
Gress [1] has done significant work on the de-
velopment of attitude control for Vertical Take-
Off and Landing (VTOL) aircraft using the
gyroscopic moments generated by the forced
precession of propellers (Fig. 2). He has designed
a family of VTOL aircraft that features a central
airframe with a gimbaled propeller mounted on
either side. By controlling the lateral and longi-
tudinal precession rates of the two gimbaled pro-
pellers, transient gyroscopic torques are created
that are used in attitude control. This aircraft can
also make use of torque vectoring from the hori-
zontal components of the propeller drag torques.
To date, this method has only been successfully
used for pitch control. Specifically, for forward
flight, the propellers are collectively tilted in the
longitudinal direction and simultaneously forced
to precess in opposing lateral directions (away
from the airframe). The forced precession causes
a net nose-down gyroscopic pitching moment that,
when added to the lateral components of the drag
torques, can be used to counteract the adverse
reactionary moment from the longitudinal accel-
eration of the propellers. The tilted thrust vectors
then produce a pitching moment about the aircraft
center of gravity that combines with the lateral
drag torque components to oppose the aerody-
namic pitch-up moment experienced by propellers
in forward flight. In the current implementation,
the yaw is controlled through differential lon-
gitudinal tilting of the propellers while roll is
controlled through differential propeller speeds,
though the general concept should provide full
attitude control through forced longitudinal and
lateral precession as mentioned above. Kendoul
et al. have built prototypes based on Gress design
and derive the equations of motion along with an
effective backstepping controller [3].
Samuel et al. have developed several proto-
types of the micro coaxial rotorcraft (MICOR)
[9]. This vehicle has two coaxial counter-rotating
rotors and is controlled using an active flexible
J Intell Robot Syst (2012) 65:417435 419
Fig. 2 (left) Micor rotor test stand, (center) Gress Aerospace eVader 750, (right) NASA Flying Test Platform
structure (Fig. 2). The structure consists of four
flexible carbon fiber beams that are bent into
semicircles in the vertical plane and evenly spaced
about the vertical axis forming a somewhat
flattened sphere. The rotors are located at the
north and south poles, protected inside the struc-
ture, and each driven by a dedicated electric mo-
tor. Shape Memory Alloy (SMA) wires are em-
bedded in the carbon fiber beams in order to
control their curvature and thereby control the
total thrust vector direction. The SMA wires are
distributed such that the curvature of the top and
bottom halves of each carbon fiber beam can be
controlled. For instance, by increasing curvature
on the bottom half of one beam and the top half
of the axially opposing beam, the thrust vectors of
both rotors are tilted in that direction. While the
incorporation of a flexible structure can lead to
marginal passive stability and mechanical simplic-
ity, it does so at the cost of significant structural
vibration. While no flight test results have been
published (to the authors knowledge), they have
demonstrated successful thrust vector tilting using
the SMA actuators.
Lim and Moerder [5, 6] have performed sub-
stantial research on the use of Control Moment
Gyroscopes (CMGs) for the stability augmenta-
tion of thrust vectoring VTOL platforms. They
have specifically looked at the use of CMGs on
the NASA Flying Test Platform (NFTP) for dis-
turbance rejection in the presence of heavy turbu-
lence (Fig. 2). The work suggests that even a rel-
atively small CMG (relative to the vehicle mass)
can more than halve the bandwidth required of
the thrust vectoring control system, which in turn
reduces the aerodynamic-based modeling uncer-
tainty. They explored several control laws, using
the CMG to control high frequency dynamics and
the thrust vectoring system to maintain trim and
perform low frequency maneuvers. No experi-
mental validation was conducted.
3 Control Moment Gyroscopes
Momentum Exchange Devices (MEDs) are com-
monly employed to reorient spacecraft such as the
International Space Station (ISS), satellites, and
robotic space systems. Their main component is
a spinning wheel or rotor that is used to exert
internal torques on the vehicle for attitude con-
trol. Unlike thrusters, which require fuel, MEDs
are completely electric and are easily powered by
solar arrays.
The simplest type of MED is the reaction wheel
(RW), which consists of a wheel that spins about
an axis fixed in the spacecrafts body frame. The
speed of the wheel is varied to generate the de-
sired reaction torque on the spacecraft. Three
RWs are typically used, each aligned with one
of the principal axes of the spacecraft. Control
moment gyroscopes (CMGs) are an extension of
reaction wheels where the spinning wheel is gim-
baled to the spacecraft.
A common configuration is the single gimbal
CMG shown in Fig. 3. The wheel is maintained
at a constant speed by a feedback control system
while the gimbal is rotated about an axis fixed
420 J Intell Robot Syst (2012) 65:417435
gimbal
wheel
gimbal axis
spin axis
Fig. 3 Control moment gyroscope diagram
in the spacecraft body to generate large gyro-
scopic torques on the vehicle. Several of these
devices are typically combined into a cluster to
enable robust full attitude control. A common
configuration, used on the ISS, consists of four
CMGs placed in a pyramid configuration, where
each CMG is located at the center of one of the
pyramid faces.
Compared to RWs, CMGs are capable of pro-
ducing larger torques for a given input power. For
this reason, they have become more commonly
used for spacecraft control systems. They do, how-
ever, have several drawbacks. They require more
complex control, and they suffer from singular
configurations at which they are incapable of pro-
ducing the desired torque on the spacecraft. There
has been a significant amount of work devoted to
creating strategies for avoiding these singularities
such as singularity robust steering laws. One such
attempt, [10], uses Variable Speed Control Mo-
ment Gyroscopes (VCMGs): single gimbal CMGs
with wheels that have variable speeds. This extra
degree of freedom enables the cluster of CMGs to
act more like RWs near singular configurations to
maintain control authority.
4 Vehicle Operation
The Dysc is a rotorcraft with two counter-rotating
coaxial rotors. Unlike those in conventional coax-
ial rotorcraft, the Dyscs rotors are coincident at
their centers, and are gimbaled to the main body.
The vehicle is composed of f ive rigid bodies:
the airframe, interior gimbal, interior rotor, exte-
rior gimbal, and exterior rotor. These bodies are
illustrated with the relevant reference frames in
Fig. 4a, b. The reference frames are referred to as
the interior gimbal reference frame (G
i
), the exte-
rior gimbal reference frame (G
e
), and the airframe
reference frame (A). They are body-fixed, located
at the respective centers of mass, and aligned with
the principal directions. The nomenclature used
throughout this section is listed in Table 1.
The orientation of each frame is represented
by the unit vector triads { g
s
i
, g
t
i
, g
g
i
}, { g
s
e
, g
t
e
, g
g
e
},
and { a
1
, a
2
, a
3
} respectively. In the case of the
gimbal triads, the first subscript denotes the lo-
cation of the gimbal, interior or exterior; the sec-
ond subscript denotes the axis direction, spin (s),
transverse (t), and gimbal (g). The interior rotor
is connected to the interior gimbal by a revolute
joint and spins relative to it about the g
s
i
axis
with angular velocity
i
. Likewise, the exterior
rotor is connected to the exterior gimbal through
a revolute joint and spins relative to it about the
g
s
e
axis with angular velocity
e
. Both the inte-
rior and exterior gimbals are in turn connected
to the airframe through revolute joints, referred
to here as the gimbal joints, with angular veloci-
ties
i
and
e
about the g
g
i
and g
g
e
axes, respec-
tively. Note that both g
g
i
and g
g
e
are fixed in the
A frame.
As with typical rotorcraft, the Dyscs rotors
generate not only the thrust necessary to hover,
but also the attitude control moments. Unlike typ-
ical rotorcraft, however, the Dysc generates atti-
tude moments without using aerodynamic forces.
The rotors act as VCMGs, enabling the vehicle
to generate pitch and roll moments through the
proper control of the rotor speeds and gimbal
joint velocities. As discussed in the introduction,
VCMGs are commonly used to control the atti-
tude of satellites and even the international space
station. The derivation presented here is similar to
that found in [10, 11].
Near hover, the yaw and heave are controlled
by the rotor speeds. The vehicles yaw, for exam-
ple, is controlled primarily by differentially vary-
ing the rotor speeds. The resulting difference in
J Intell Robot Syst (2012) 65:417435 421
Fig. 4 Dysc diagram
a exploded view,
b isometric view,
c isometric view with
control inputs
(a) (c)
(b)
rotor drag torques causes a net yaw torque on
the airframe. The rotor speeds also can be var-
ied collectively to change the average thrust that
is generated to control heave motion. When the
gimbal angles are nonzero, the drag torques do
not cancel and, therefore, a small external torque
exists that requires compensation.
The pitch and roll of the vehicle are primar-
ily controlled through small gimbal joint motions
that produce gyroscopic moments on the airframe.
Since the rotors are coupled through the airframe
and because their angular momentum vectors are
in opposite directions, both of their moments act
Table 1 Nomenclature
I
G
j
k
, I
R
j
k
Gimbal and rotor inertia components
j (s) spin axis, (t) transverse axis, and
(g) gimbal axis
k (i) interior, (e) exterior
I
A
Airframe inertia matrix
g
j
k
Gimbal reference frame unit vectors
j (s) spin axis, (t) transverse axis, and
(g) gimbal axis
k (i) interior, (e) exterior
a
i
Airframe reference frame unit vectors
i (1) primary axis, (2) secondary axis, and
(3) tertiary axis
in the same direction. As mentioned above, the
control becomes increasingly complex when the
gimbal joint angles become nonzero and cause a
net drag torque on the airframe. Furthermore, the
gyroscopic moments are created by applying gim-
bal joint velocities, thereby making it impossible
to resist steady external moments such as those
generated by the asymmetry of forward flight
without allowing the rotors to continuously rotate
about their gimbal axes. The strategy for compen-
sating for steady moments will be addressed in the
next section.
If the triads that define the gimbal reference
frames are written with respect to A, then the
rotation matrices from G
i
and G
e
to A can be
expressed as
A
R
G
x
= [ g
s
x
g
t
x
g
g
x
] (1)
where x = {i, e} to indicate the interior or exterior
gimbal. Note that in the following derivation, the
lack of the location subscript indicates that the
parameter or expression applies to both interior
and exterior gimbals and rotors.
The inertia matrix of each gimbal with respect
to its reference frame is a constant diagonal ma-
trix. Furthermore, the inertia matrix of each rotor
with respect to its corresponding gimbal reference
422 J Intell Robot Syst (2012) 65:417435
frame is also a constant diagonal matrix due to its
symmetry about the spin axis.
G
I
G
=

I
G
s
0 0
0 I
G
t
0
0 0 I
G
g

,
G
I
R
=

I
R
s
0 0
0 I
R
t
0
0 0 I
R
t

(2)
The inertia matrices of the gimbal and rotor
can be expressed in the A frame using the
transformations
A
I
G
=
A
R
G
G
I
G
A
R
T
G
= I
G
s
g
s
g
T
s
+ I
G
t
g
t
g
T
t
+ I
G
g
g
g
g
T
g
(3a)
A
I
R
=
A
R
G
G
I
R
A
R
T
G
= I
R
s
g
s
g
T
s
+ I
R
t
g
t
g
T
t
+ I
R
t
g
g
g
T
g
(3b)
5 Attitude Dynamics
We will first consider one gimbal and rotor con-
nected to the airframe and then extend the results
to the two gimbal and rotor case.
The angular velocity of the gimbal with respect
to the airframe is given by

G/A
= g
g
(4)
The angular velocity of the rotor with respect to
the gimbal is given by

R/G
= g
s
(5)
Using the definition in Eq. 1, we can express the
airframe angular velocity in the gimbal frame as
follows
G

A
=

=
A
R
T
G
A

A
=

g
T
s

A/N
g
T
t

A/N
g
T
g

A/N

(6)
where N is the inertial frame and the absence of
an anterior superscript indicates that the quan-
tity is written in frame A. The total angular
momentum of a vehicle with one gimbal and one
rotor has contributions from all three bodies
H = H
A
+ H
G
+ H
R
(7)
where the gimbal angular momentum is given by
H
G
= I
G

G/N
(8)
and
G/N
=
A/N
+
G/A
. Substituting the ex-
pression for I
G
in Eq. 3a and the result in Eq. 6
into Eq. 8, we get
H
G
=
_
I
G
s
g
s
g
T
s
+I
G
t
g
t
g
T
t
+I
G
g
g
g
g
T
g
_

A/N
+I
G
g
g
g
(9)
This can be simplified to
H
G
= I
G
s

s
g
s
+ I
G
t

t
g
t
+ I
G
g
(
g
+ ) g
g
(10)
In order to obtain the rate of change of the
gimbal angular momentum, we must first compute
the rate of change of the gimbal reference frame
unit vectors with respect to the inertial frame and
the rates of change of the airframe angular veloc-
ity projections along the gimbal directions. These
rates can be expressed in terms of the gimbal rate,
, and the aircraft angular rate,
A/N
. The rate
of change of the gimbal angular momentum is
therefore

H
G
=
_
I
G
s
g
T
s
+ (I
G
s
I
G
t
+ I
G
g
)
t
+ (I
G
g
I
G
t
)
t

g
_
g
s
+
_
I
G
t
g
T
t
+ (I
G
s
I
G
t
I
G
g
)
s
+ (I
G
s
I
G
g
)
s

g
_
g
t
+
_
I
G
g
g
T
g
+ I
G
g
+ (I
G
t
I
G
s
)
s

t
_
g
g
(11)
Similarly, the angular momentum of the rotor is
H
R
= I
R
s
(
s
+ ) g
s
+ I
R
t

t
g
t
+ I
R
t
(
g
+ ) g
g
(12)
J Intell Robot Syst (2012) 65:417435 423
The rate of change of the rotor angular momen-
tum is

H
R
=
_
I
R
s
_

+ g
T
s
+
t
_
_
g
s
+
_
I
R
s

s
+ I
R
t
g
T
s
+ (I
R
s
I
R
t
)
s

g
+ I
R
s
( +
g
) 2I
R
t

s

_
g
t
+
_
I
R
t
_
g
T
g
+
_
+ (I
R
t
I
R
s
)
s

t
I
R
s

t
_
g
g
(13)
The rate of change of the airframe angular mo-
mentum is

H
A
= I
A
+I
A
(14)
where is an element of so(3), defined as
=

0
3

2

3
0
1

2

1
0

(15)
We define the rotor assembly (gimbal and rotor)
inertia matrix as
J = I
G
+ I
R
(16)
We further define the total vehicle inertia
matrix as
I = I
A
+ J (17)
The rate of change of the total angular momentum
for a vehicle composed of an airframe and one
gimbaled rotor is therefore
I = I
_
I
R
s

+ J
s

t
(J
t
J
g
)
t

_
g
s

_
(J
s

s
+ I
R
s
) (J
t
+ J
g
)
s

+ I
R
s

g
_
g
t

_
J
g
I
R
s

t
_
g
g
+ Q (18)
where Q is the external torque on the vehicle.
We can extend these results to the full five body
system by writing the equations of motion as
I = I G
s
(
s
Q) G
t

t
G
g

g
(19)
where Q now encapsulates the rotor drag torques
and the total vehicle inertia matrix, I, gimbal
frame triad matrices, G
k
, and effective torques,

k
, are
I = I
A
+

k=i,e
J
s
k
g
s
k
g
T
s
k
+ J
t
k
g
t
k
g
T
t
k
+ J
g
k
g
g
k
g
T
g
k
(20)
G
s
=
_
g
s
i
g
s
e
_
, G
t
=
_
g
t
i
g
t
e
_
, G
g
=
_
g
g
i
g
g
e
_
(21)

s
=
_
I
R
s
i

i
+ J
s
i

i

t
i
(J
t
i
J
g
i
)
t
i

i
I
R
se

e
+ J
s
e

e

t
e
(J
t
e
J
g
e
)
t
e

e
_
(22a)

t
=

(J
s
i

s
i
+I
R
s
i

i
)
i
(J
t
i
+ J
g
i
)
s
i

i
+I
R
s
i

g
i
(J
s
e

s
e
+I
R
se

e
)
e
(J
t
e
+ J
g
e
)
s
e

e
+I
R
se

g
e

(22b)

g
=
_
J
g
i

i
I
R
s
i

t
i
J
g
e

e
I
R
se

t
e
_
(22c)
We also can calculate the control torques by
revisiting the rate of the change of the angular
momentum of the rotors about the spin axes and
the rotor assemblies about the gimbal axes (

H
R
and

H
R
+

H
G
). The rotor spin control torques
are
q
s
=
_
I
R
s
i
_

i
+ g
T
s
i
+
i

t
i
_
I
R
se
_

e
+ g
T
s
e
+
e

t
e
_
_
(23)
and the gimbal joint torques are
q
g
=
_
J
g
i
_
g
T
g
i
+
i
_
J
i

s
i

t
i
I
R
s
i

t
i
J
g
e
_
g
T
g
e
+
e
_
J
e

s
e

t
e
I
R
se

t
e
_
(24)
5.1 Aerodynamic Model and Verification
Blade Element Momentum Theory (BEMT) is a
method commonly employed in helicopter design
and analysis to estimate the aerodynamic forces
on a rotor and is an integral part of the aero-
dynamic model and dynamics simulation in this
work. The simulation uses BEMT-based models
for the combined rotor system and is valid for
hover and lowspeed forward flight. Typical graph-
ical output of the simulation is shown in Fig. 5.
424 J Intell Robot Syst (2012) 65:417435
0.4
0.3
0.2
0.1
0
0.1
0.2
0.3
0.25
0.2
0.15
0.1
0.05
0
0.05
0.1
0.15
0.2
0.25
4
2
0
2
4
6
8
10
12
x 10
3
x (m)
Thrust Distribution
y (m)
T
h
r
u
s
t

(
N
)
0.25
0.2
0.15
0.1
0.05
0
0.05
0.1
0.15
0.2
0.25
0.25
0.2
0.15
0.1
0.05
0
0.05
0.1
0.15
0.2
0.25
8
6
4
2
0
x 10
3
x (m)
Exterior Rotor Disc Velocity
y (m)
z

(
m
)
Fig. 5 BEMT typical results: (left) interior and exterior rotor thrust distributions, (right) exterior velocity vector field
It includes aerodynamic coefficient lookup tables
containing experimental low Reynolds number
coefficient data published by Okamoto et al. [7],
that was augmented with data generated by the
software package XFOIL to provide a more accu-
rate model that includes both the angle of attack
and Re effects.
It is well known in the literature [2, 4, 8], that a
rotors inflow in hover can be modeled as

i
=
_
_
R

C
l

16R
2


c
2
_
2
+
R

C
l

r
8R
2

_
R

C
l

16R
2


c
2
_
(25)
where
i
is the inflow ratio, is the rotor solidity,
is the blade pitch as a function of radial position,
r is the radial distance to the blade element,
c
is the climb ratio, and C
L

is the linearized lift


coefficient. This result is obtained by equating
the blade element and momentum theories for
a differential annular element of the rotor disk.
In this derivation, R

and R
2
are Dysc specific
correction terms to account for the annular nature
of the exterior rotor. Equation 25 allows us to
solve for the axisymmetric inflow distribution in
hover as a function of blade twist ((r)), blade
chord ((r)), and blade profile geometry, which
are all functions of the radial position.
For first order estimation purposes, we can ne-
glect the dependence of C
L
and C
D
on Re and
Ma and assume C
L
varies linearly with up to
stall, and that C
D
can be modeled as a 2nd order
polynomial in . The blade element thrust and
drag coefficients can then be approximated as
_
dC
T
dC
Q
_
=
1
2
r
_
r
2

2
i
_
1

i
r

i
r
_ _
C
L
()
C
D
()
_
dr (26)
This form is convenient for initial estimates of
rotor hover performance, however, the complete
form of the BEMT equations are used in the
numerical simulation.
A parametric CAD model of a rotor blade
was created using the twist, chord, and profile
distributions found by an optimizer written in
matlab that directly used the performance results
predicted by the BEMT simulation. The optimizer
was configured to maximize the rotor Figure of
Merit (FM).
In addition to driving the rotor blade design
with the simulator aerodynamic model, several
low Re design guidelines from the literature were
also used. Specifically, the selected blade section is
a thin circular arc with sharp leading and trailing
edges. Many researches, including Okamoto et al.,
have noted that thin simple airfoils with sharp
leading and trailing edges exhibit a significant
improvement in performance when compared to
traditional airfoil sections.
J Intell Robot Syst (2012) 65:417435 425
Fig. 6 Test propeller: (top) top view, (bottom) front view
Since the accurate and repeatable manufacture
of the rotor blades is a critical part of Dysc proto-
type development, the use of a rapid prototyping
process such as stereolithography (SLA) was ex-
plored as a method for producing functional rotor
blades. Additive manufacturing processes are ad-
vantageous for components like rotor blades be-
cause of the relative ease with which it can create
complex geometry. There are several challenges
to using these processes for functional parts, the
most important of which are the minimum feature
size and material stiffness. The minimum feature
size limits not only how thin the airfoil can be but
also the leading and trailing edge sharpness. Rapid
prototyping materials are generally less stiff than
their bulk machinable counterparts, and in this
case, a reduction in stiffness can result in excessive
tip deflection and dramatic loss of lift.
A pair of test propeller blades was printed
using SLA out of one of the stiffest materials
availableAccura Bluestone (see Fig. 6). Prior
to printing, finite element analysis (FEA) was
used to evaluate the tip deflection and root stress
on the blade using the combined centrifugal and
aerodynamic loads. The latter being predicted
by the BEMT analysis. We not only tested the
performance of the propeller against Consumer
Off-The-Shelf (COTS) propellers of the same
diameter, but also compared the thrust and drag
to that predicted by the BEMT simulation.
A test stand equipped with two force transduc-
ers and a closed-loop speed controlled BLDC mo-
tor was used to collect thrust and drag data from
the propellers. Both the thrust and drag agree very
well with the BEMT model as shown in Fig. 7. The
reduction of drag at high speeds in the drag plot is
an artifact of drift in a faulty force transducer.
Figure 8 shows that the SLA propeller (labeled
CT-alpha and CT-alpha2) out-performs all of the
COTS propellers tested in terms of FM. Despite
the increase in noise near the maximum FM, the
SLA propeller clearly performs as good or better
than the best COTS propeller. It should be noted
0 500 1000 1500 2000 2500 3000 3500
0
0.5
1
1.5
2
2.5
3
3.5
4
4.5
SLA Propeller Thrust Validation Experiments
Speed (RPM)
T
h
r
u
s
t

(
N
)


Experimental, col = 7
o
BEMT Model
0 500 1000 1500 2000 2500 3000 3500
0
10
20
30
40
50
60
70
80
SLA Propeller Drag Torque Validation Experiments
Speed (RPM)
D
r
a
g

(
N
m
m
)


Experimental, col = 7
o
BEMT Model
Fig. 7 Test propeller comparison with BEMT: (left) thrust, (right) drag
426 J Intell Robot Syst (2012) 65:417435
0 5 10 15 20 25 30 35 40 45 50
0.05
0.1
0.15
0.2
0.25
0.3
0.35
0.4
0.45
0.5
Disk Loading (N/m
2
)
F
M


Fig. 8 Figure of merit versus disk loading for SLA test pro-
peller and several COTS propellers of the same diameter
that the FM of two propellers must be compared
at the same disk loading. The performance gain
is most likely due to compromises made in the
design to reduce manufacturing cost, such as mold
cost for injection molded propellers
6 Control and Simulation
Two controllers were constructed for evaluation
in the simulator; a controller based on the lin-
earized attitude dynamics to stabilize the vehicle
attitude around hoverthe linearized system al-
lows us to examine properties such as controlla-
bility (in a linear sense) and allow us to intuit the
contributions of various design parameters to the
attitude response of the vehicle, and a nonlinear
controller that employs feedback linearization.
We should note here that the final control
scheme used on the vehicle should include a
method for dealing with steady external moments.
For example, this behavior can be mitigated by
the addition of a mechanism capable of generating
steady moments. One simple option is to add
small flaps in the downwash of the rotors to coun-
teract the effects of steady moments such as those
due to an offset mass in the vehicle (perhaps due
to a payload), and asymmetric lift distributions
from wind gusts or forward flight.
6.1 Linear Controller
The steps to obtain the linearized attitude dy-
namics are omitted from this treatment for the
sake of length. Assume now that we have the
linearized system in standard form, Ax + Bu.
The yawheave dynamics are decoupled from the
rollpitch dynamics in the linearized system and
can be considered as separate subsystems. Con-
sidering the gimbal joint velocities and the rotor
velocities as the control inputs, we can reformu-
late the subsystems as shown in [12].
The linearized yaw and heave dynamics are
z =C
9

i
+ C
10

e
q
3
=C
7

i
+ C
8

e
+ B
3

i
+ B
4

e
(27)
where
i
=
i

i
and
e
=
e

e
. The
expressions for the coefficients can be found in
Table 2. It is convenient to refer to the control
variables as effective collective and yaw controls,

d
c
and
d
y
. We can then solve for the desired rotor
speeds with the following mapping
_

d
i

d
e
_
=
1
C
9
C
8
C
10
C
7
_
C
8
C
10
C
7
C
9
_ _

d
c

d
y
_
(28)
Table 2 Linearization coefficients
Coeff. Expression Coeff. Expression
C
1

I
R
s
i

i
+ I
R
se

e
I
A
11
+ J
g
e
+ J
t
i
C
2
D
i
(

i
)
I
A
11
+ J
g
e
+ J
t
i
C
3

I
R
s
i

i
I
A
11
+ J
g
e
+ J
t
i
C
4
I
R
s
i

i
+ I
R
se

e
I
A
22
+ J
g
i
+ J
t
e
C
5

D
e
(

e
)
I
A
22
+ J
g
i
+ J
t
e
C
6
I
R
se

e
I
A
22
+ J
g
i
+ J
t
e
C
7
dD
i
(
i
)
d
i
|

i
2(I
A
33
+ J
s
i
+ J
s
e
)
C
8
dD
e
(
e
)
d
e
|

e
2(I
A
33
+ J
s
i
+ J
s
e
)
C
9
dL
i
(
i
)
d
i
|

i
m
C
10
dL
e
(
e
)
d
e
|

e
m
B
1

J
g
e
I
A
11
+ J
g
e
+ J
t
i
B
2

J
g
i
I
A
22
+ J
g
i
+ J
t
e
B
3

I
R
s
i
2(I
A
33
+ J
s
i
+ J
s
e
)
B
4

I
R
se
2(I
A
33
+ J
s
i
+ J
s
e
)
J Intell Robot Syst (2012) 65:417435 427
We can then setup a hierarchical control scheme
with a simple feedback control for heave and yaw
_

y
_
=
_
K
cp
_
z
d
z
_
+ K
cd
_
z
d
z
_
K
yp
_
q
d
3
q
3
_
+ K
yd
_
q
d
3
q
3
_
_
(29)
Assuming we have direct control over the rotor
spin angular accelerations (a good assumption
near hover), we can use a proportional controller
to set desired rotor speeds using the result of
Eq. 28. If we examine this further, we see that the
influence of the collective control input on the yaw
through the angular acceleration terms in Eq. 27
can be eliminated if the following relationship
between the angular acceleration gains holds:
K
t
i
=
B
4
C
7
B
3
C
8
K
t
e
(30)
where K
t
i
and K
t
e
are the angular acceleration
gains. This relationship involves ratios of the in-
terior and exterior spin inertias and drags. This is
an intuitive result, stating that if you accelerate the
rotors according to Eq. 30 the inertial torques will
always cancel during a heave command and have
no net impact on the yaw motion.
The linearized roll and pitch dynamics are

1
= C
1

2
+ C
2

i
+ C
3

i
+ B
1

e

2
= C
4

1
+ C
5

e
+ C
6

e
+ B
2

i
(31)
This can be reformulated into a compact state
space representation

x
1
x
2
x
3
x
4
x
5
x
6

0 0 1/2 0 0 0
0 0 0 1/2 0 0
0 0 0 0 1 0
0 0 0 0 0 1
0 0 0 0 0 C
1
0 0 0 0 C
4
0

x
1
x
2
x
3
x
4
x
5
x
6

0 B
1
/2
B
2
/2 0
C
3
+C
1
B
2
0
0 C
6
+C
4
B
1
C
2
C
1
C
6
+B
1
C
1
C
4
C
4
C
3
+C
4
C
1
B
2
C
5

_

i

e
_
(32)
using the following definitions
x
1
= q
1
x
2
= q
2
x
3
=
1
B
1

e
x
4
=
2
B
2

i
x
5
=
1
B
1

e
(C
3
+ C
1
B
2
)
i
x
6
=
2
B
2

i
(C
6
+ C
4
B
1
)
e
(33)
where q
1
and q
2
are the roll and pitch components
of the airframe quaternion,
1
and
2
are the
roll and pitch components of the vehicle angular
velocity and
i
and
e
are the gimbal joint angular
velocities that are being controlled.
The gain matrix, K
rp
, must now be tuned such
that the actual nonlinear system has the desired
response to perturbations around hover. Typical
simulation results are shown below (Fig. 9).
_

i

e
_
= K
rp
_
x
1
x
2
x
3
x
4
x
5
x
6
_
T
(34)
6.2 Feedback Linearization
For linearization by state feedback, we take the
rotor actuator torques and the gimbal angular
rates as our control inputs
u =
_
q
s

_
(35)
As is typical with helicopters, we take the out-
put vector to be
y =
_
q
z
_
(36)
where q is the quaternion describing the attitude
of the airframe and z is the heave velocity of
the vehicle. To linearize by state feedback, we
differentiate the output twice to obtain an expres-
sion that contains the inputs
y =
_
1
/2(

q + q)

F
z
/m
_
=
_
1
/2(

q +
1
/2
2
q)

F
z
/m
_
=

1
/4
_
2Q
_
0

_
+
2
q
_

F
z
/m

(37)
428 J Intell Robot Syst (2012) 65:417435
Fig. 9 Linear hover
controller performance:
(left) airframe Euler
angles, (right) gimbal
joint velocities
0 0.5 1 1.5 2
0
5
10
15
Time (seconds)
E
u
l
e
r

A
n
g
l
e
s

(
d
e
g
r
e
e
s
)


roll
pitch
yaw
0 0.5 1 1.5 2
0
20
-20
-40
-60
-80 -5
40
60
80
100
Time (seconds)
A
n
g
u
l
a
r

V
e
l
o
c
i
t
y

(
d
e
g
/
s
e
c
)


Interior
Exterior
where is given by Eq. 19. is a skew-symmetric
matrix that is used to calculate the rate of change
of the quaternion and Q is the matrix form of the
quaternion product
=

0
1

2

3

1
0
3

2

2

3
0
1

3

2

1
0

,
Q =

q
0
q
1
q
2
q
3
q
1
q
0
q
3
q
2
q
2
q
3
q
0
q
1
q
3
q
2
q
1
q
0

(38)
Assuming the vertical drag force is negligible, F
z
is the vertical component of the combined thrust
force from the two rotors and can be expressed as
F
z
=
_
0
z
_
q
_
0
T
total
_
q

=
_
0
z
_
q
__
0
G
s
_ _
T
i
(
i
)
T
e
(
e
)
__
q

(39)
where T
i
(
i
) and T
e
(
e
) are the interior and ex-
terior rotor thrusts respectively and z is the unit
vector in the world z direction. q
_
0 T
T
total
_
T
q

is the composite quaternion product that rotates


the total thrust vector into the world frame. After
differentiating Eq. 39 with respect to time and
performing some simplification, we can write

F
z
= f
z
(x) + g
z
(x)u (40)
We can also write the airframe angular accelera-
tion in a similar fashion
= f
r
(x) + g
r
(x)u (41)
where x is the vehicle state vector. We can now
write Eq. 37 as
y = A + Bu (42)
with
A =

1
/4
2
q +
1
/2Q
_
0
f
r
(x)
_
(
1
/m) f
z
(x)

,
B =

1
/2Q
_
0
g
r
(x)
_
(
1
/m)g
z
(x)

(43)
We choose new control inputs, v, that linearize the
system. The original inputs become
u = B
1
v B
1
A (44)
It should be noted that there are four outputs and
four inputs, the first row of Eq. 42 involves the
scalar component of the quaternion and is kept
in the derivation for convenience. With v as our
control inputs, we define a new state vector as
=

q
1
q
2
q
3
z
z
q
1
q
2
q
3
z

(45)
J Intell Robot Syst (2012) 65:417435 429
with the simple linear controller

q
1
q
2
q
3
z
z
q
1
q
2
q
3
...
z

0 0 0 0 0 1 0 0 0
0 0 0 0 0 0 1 0 0
0 0 0 0 0 0 0 1 0
0 0 0 0 1 0 0 0 0
0 0 0 0 0 0 0 0 1
k
p
1
0 0 0 0 k
d
1
0 0 0
0 k
p
2
0 0 0 0 k
d
2
0 0
0 0 k
p
3
0 0 0 0 k
d
3
0
0 0 0 k
p
z
k
d
z
0 0 0 k
dd
z

q
1
q
2
q
3
z
z
q
1
q
2
q
3
z

(46)
The response to an initial condition is shown
in Fig. 10. The time axis is enlarged to show the
transient detail at the expense of showing the
Euler angles reaching steady state hover.
6.3 Simulation
A simulator was constructed in matlab using
the Recursive Newton Euler framework, to de-
velop and evaluate the various control schemes.
The simulator enables us to test the efficacy of
different mechanical designs and tune the perfor-
mance to meet specific agility goals.
6.4 Five Body Recursive Newton Euler
For the purposes of simulation, it is convenient
to construct a compact form of the multibody
dynamics using linear algebra. This maximizes
flexibility when implementing various control
schemes and aerodynamic models. It also provides
an easy method for calculating constraint forces
and torques. Key results of the Recursive Newton
Euler (RNE) equations are highlighted here.
The kinematics of the Dysc can be written in
recursive form
p
1
= p
1
p
2
D
2
p
1
= H
2

1
+
2
p
3
D
3
p
2
= H
3

2
+
3
p
4
D
4
p
3
= H
4

3
+
4
p
5
D
5
p
4
= H
5

4
+
5
(47)
where p
i
is the 6 element vector composed of the
translational acceleration of the center of mass of
body i and the angular acceleration of body i. The

i
s are the four joint accelerations: interior rotor,
interior gimbal, exterior gimbal, exterior rotor,
respectively. The D
i
, H
i
, and
i
are derived from
the joint kinematics
D
i
=
_
I v
i
i1
+v
i1
i
0 I
_
, H
i
=
_
v
i1
i
u
i,i1
u
i,i1
_

i
=
_

i1

i1
v
i
i1

i

i
v
i1
i
+v
i1
i

i
u
i,i1

i1

i
u
i,i1

i1
_
(48)
Fig. 10 Feedback
linearization controller
performance: (left)
airframe Euler angles,
(right) gimbal joint
velocities
0 0.5 1 1.5 2
0
1
2
3
4
5
6
Time (seconds)
E
u
l
e
r

A
n
g
l
e
s

(
d
e
g
r
e
e
s
)


roll
pitch
yaw
0 0.5 1 1.5 2
0
10
20
30
Time (seconds)
A
n
g
u
l
a
r

V
e
l
o
c
i
t
y

(
d
e
g
/
s
e
c
)


Interior Actual
Exterior Actual
Interior Cmd
Exterior Cmd
430 J Intell Robot Syst (2012) 65:417435
where v
j
i
are vectors between consecutive body
centers of mass, u
i, j
are unit vectors in the joint
directions,
i
are the body angular velocities, and

[ ] indicates the skew-symmetric matrix form of


the cross product. It is more convenient to treat
the airframe, or body 3, as the base body. In that
case, we can reorder and rewrite the equations as
p
3
= p
3
p
2
D
1
3
p
3
= D
1
3
H
3

2
D
1
3

3
p
1
D
1
2
p
2
= D
1
2
H
2

1
D
1
2

2
p
4
D
4
p
3
= H
4

3
+
4
p
5
D
5
p
4
= H
5

4
+
5
(49)
The system kinematics can then be encapsulated
as
D p = H p
J
+ (50)
or
p = B
J
p
J
+
J
(51)
where
B
J
= D
1
H,
J
= D
1
,
p =
_
p
3
p
2
p
1
p
4
p
5
_
T
, p
J
=
_
p
3

4
_
T
(52)
The dynamics for the system can be expressed as
M
d
p = F
e
+ F
v
+ F
c
(53)
where M
d
is the system mass matrix, which in-
cludes the body masses and inertia tensors, F
e
is
the vector of external forces and torques, F
v
is the
vector of quadratic velocity terms, and F
c
is the
vector of constraint forces and torques. Using the
kinematic relations, we can write the equations
of motion with the joint accelerations and the
airframe body accelerations

M
d
p
J
=

F (54)
where

M
d
= B
T
J
M
d
B
J

F = B
T
J
(F
e
+ F
v
M
d

J
)
(55)

M
d
is the generalized system mass matrix and

F is
the generalized force vector.
7 Prototype Design
The Dysc concept presents several mechanical
design challenges, the most significant of which
is supporting and driving the rotors. The gimbal
actuation and the size and shape of the airframe
also are important design features. They will be
discussed in the subsequent sections.
7.1 The Rotors and Gimbals
Many rotorcraft use cyclic blade pitch to tilt the
thrust vector and generate the necessary attitude
moments. This pitch variation usually is created by
a swashplate mechanism. The pilots inputs to the
control system change the blade pitch in a cyclic
manner by tilting the swashplate, which also can
be translated along the rotor shaft to change the
collective blade pitch and the average thrust. The
swashplates mechanical complexity contributes
to the high maintenance costs associated with
helicopters.
In contrast to the swashplate, the Dysc employs
directly gimbaled rotors, and mechanical simplic-
ity is gained at the cost of supporting and driving
the coaxial rotors at their centers. Each rotor of
the Dysc must be supported by its gimbal in such
a way that the the aerodynamic and gyroscopic
moments are transmitted to the airframe without
relying on structures that may interfere with the
other rotor.
Co-location of the rotor drive motors and the
gimbal actuators at the center of the vehicle with
structural spokes radiating out to the respective
rotors is infeasible, as they would inevitably inter-
fere at nonzero gimbal joint angles. The inherent
flexibility of these spokes also could cause colli-
sion between the rotors.
The method for connecting the rotors to their
respective gimbals, therefore, must act as a revo-
lute joint, constraining relative translation and all
but the spin rotation without having any structural
component pass over the other rotor. The exterior
rotor should be supported along its inner edge,
allowing the airframe and the interior rotor to
occupy the space inside it. Similarly, the interior
rotor should be supported along its outer edge.
At the outset, off-the-shelf thin section bearings
were considered in the early design stages for their
J Intell Robot Syst (2012) 65:417435 431
Fig. 11 (left) Idler wheel,
(center) exterior rotor
ring, (right) exterior rotor
ring cross section
ease of implementation and availability. To resist
the combined radial, axial, and moment loads
generated by the thrust and the aerodynamic and
gyroscopic moments, either a duplexed back-to-
back configuration of angular contact bearings or
a single four point contact (type-X) bearing is
required. While thin section bearings are avail-
able in sizes that are applicable to the Dysc, they
are almost exclusively manufactured from steel,
making them prohibitively heavy ( 250 g each).
In addition, they are rated for speeds roughly an
order of magnitude less ( 500 RPM) than those
necessary to produce sufficient thrust for this size
vehicle. This speed limitation is primarily due to
bearing seizure from the excessive heat buildup.
In the case of a rotorcraft, it may be possible to
prevent seizing and operate a thin section bearing
beyond its rated speed because it will be in the
rotor slipstream. Several bearing manufacturers
were consulted regarding ultra-light custom thin
section bearings. Manufacturing thin section bear-
ings from alternative materials such as aluminum,
titanium, or composite not only drastically in-
creases the cost, but also decreases the maximum
speed due to inferior heat dissipation.
Due to the inherent limitations of thin section
bearings, a new joint design was developed to
simultaneously support and drive the rotors. This
design employs a friction drive that uses small
wheels, shown in Fig. 11, that ride in a v-shaped
groove on the rotor.
Consider first the exterior rotor. An aluminum
ring with a c-shaped cross section (see Fig. 11) acts
as the hub, containing the attachment points for
the rotor blades and providing structural rigidity
for the whole rotor assembly. The inner face of
the ring contains a v-shaped groove, shown on the
right in Fig. 11, that runs its full circumference.
The corresponding exterior rotor gimbal houses
only three small wheels so as not to over-constrain
the ring. The wheels are comprised of an alu-
minum hub that is grooved to accommodate a
buna-N rubber o-ring. These o-rings roll along
the groove on the ring. Two of these wheels are
idler wheels, free to rotate with respect to the
gimbal. The third wheel is driven by a brushless
DC (BLDC) motor, creating a friction drive with
the ring. This configuration is particularly suited
for the high speeds typical of BLDC motors be-
cause of the rather large gear ratio between the
drive wheel and the ring. The large difference in
radii of the wheels and the ring also causes the
wheels to spin at relatively high speeds ( 10
20 times faster than the ring). Significant care
was taken to design the wheel assemblies and to
ensure the wheel bearings are operating within
their speed and dynamic load specifications. The
interior rotor was designed in a similar fashion,
except its ring contains the groove on the outside
and the blade attachment points on the inside (see
Fig. 12). The interior rotor blades also attach to
a small central hub to help maintain structural
rigidity.
Fig. 12 Interior gimbal assembly CAD modelcutaway
exposing drive motor, idler wheel, and ring groove
432 J Intell Robot Syst (2012) 65:417435
The fit between the wheels and the ring is
crucial to the performance of the rotor joints.
As such, significant time was devoted to ensur-
ing that the manufacturing tolerances remained
within specification. The design of the wheel hub
was guided by standard design practices for parts
with o-ring grooves. The wheel hub groove has a
circular cross section to maximize contact with the
rubber o-ring and to minimize slip. The geometry
of the ring groove was dictated by the available
equipment for manufacturing the grooving tool.
The groove therefore was designed to have a v-
shaped cross-section with an included angle of 90
degrees and a small fillet at the tip. This geometry
proves to be sufficient to resist the various ap-
plied loads. Several experiments were performed
with a prototype wheel and a linear groove to
determine the necessary interference between the
o-ring and the ring groove. An interference of
0.002 inches provided adequate transmission of
the motor torque while keeping the normal force
relatively lowso as not to significantly increase the
wear of the components.
7.2 Airframe
The airframe houses all of the Dyscs electronics,
including hardware for onboard high-level con-
trol, motor control, and sensing, as well as the
lithium-polymer batteries. It is connected to the
interior and exterior gimbals through their re-
spective gimbal joints, whose axes are orthogo-
nal. The attitude control moments are generated
by controlling the velocities of these joints. It is
then critical to have high resolution low speed
control over these joints. Several actuators were
considered to control the gimbal joints, including
SMA wires, piezo bending actuators, hobby servo
motors, and BLDC direct drive motors.
SMA wires are very light weight and can pro-
duce relatively large forces, but have a small
stroke ( 4% strain) and low cycle time ( 0.1 s).
Piezo bending actuators are also light weight and
have sufficient stroke, but produce small forces,
and require large amplifiers to drive them. They
would also require the use of a mechanism to con-
vert the bending motion into a revolute motion.
Hobby servos weigh the most of the actuators
considered. They are capable of producing large
torques for their size, but have a limited maximum
speed and suffer from geartrain backlash and
friction.
Direct drive BLDC motors are typically lighter
than hobby servos of the same size, but produce
an order of magnitude less torque. They have
minimal friction, no backlash, and do not have a
prescribed velocity profile. They can also be put
in a free-wheel state where they have a negligible
effect on the joint motion. The nature of VCMGs
lends itself to the use of the BLDC motor for
several reasons. BLDCs allowfor the most control
over the gimbal joint velocities because they do
not have the complications of a built-in maxi-
mum displacement or backlash and friction like
servo motors. Also, the advantage of VCMGs is
their characteristic large gain between the gimbal
torque and the resulting attitude moment. The
BLDC motors should therefore be able to provide
sufficient torque while providing the benefits of
no backlash and no friction. The use of direct
drive BLDC motors also opens up the possibility
of using the same motors to drive the gimbal joints
as those used to drive the rotors which would
simplify the design by helping to create an axisym-
metric airframe.
7.3 Electrical
A BLDC motor controller board was developed
for use on the Dysc (Fig. 13). The board uses
a TI TMS320f2808 microcontroller and a 12-bit
Fig. 13 Brushless motor controller board
J Intell Robot Syst (2012) 65:417435 433
Fig. 14 Dysc prototype on heave-yaw test stand
magnetic encoder to perform Space Vector Mod-
ulation (SVM) control on the three-phase BLDC
motors that constitute the rotor friction drives and
the gimbal joints. The high clock speed of the
microcontroller (60 MHz) and the high resolution
of the encoder facilitate the accurate control of
the brushless motors at both the high speeds of
the rotors and the low speeds of the gimbal joints.
The use of four identical motor controller boards
on the Dysc also make component and board
replacement easier and reduces debugging time as
well as helps to evenly distribute the mass in the
airframe.
An off-board computer runs the high level con-
trol and communicates with the four motor con-
troller boards via a can bus tether.
7.4 Manufacturing
A prototype vehicle was constructed to demon-
strate the performance benefits of the Dysc con-
cept (Fig. 14). All of the non-3D printed com-
ponents were manufactured in-house from alu-
minum 6061-T6. This includes the motor mounts
and joint frames. The primary structural com-
ponents of the airframe and the gimbal frames
were manufactured by CNC machining light-
weight composite sheets from Dragonplate. The
most complex parts are the interior and exterior
rings. The exterior ring, shown in Fig. 11, is 274
mm in diameter and was manufactured using a
lathe and a CNC milling machine from a single
billet of aluminum 6061-T6.
8 Experimental Results
The first set of experiments focused on test-
ing the performance of the ring friction drives.
First, the assembled rotors were balanced using
a gravity-driven balancing frame. Each rotor as-
sembly was then assembled into its corresponding
gimbal frames and spun at low speeds ( 1000
RPM) while the friction and potential slip be-
tween the rings and wheels were evaluated. The
rotors were then spun at high speeds ( 2000
RPM and higher) to check the performance at and
above typical operating speeds.
The same test stand used to validate the SLA
test propeller was used to collect thrust and drag
data for both the interior and exterior rotors.
The exterior rotor data is shown in Fig. 15. The
BEMT simulator over predicts both the thrust
and the drag at high speeds. This is most likely
due to flattening of the blades at higher speeds
resulting in a lower angle of attack. The interior
rotor thrust and drag (not shown) also exhibits the
same trend of underperforming at higher speeds
Fig. 15 Exterior rotor
thrust and drag data
comparison with BEMT:
(left) thrust, (right) drag
0 2500 5000 7500 10000 12500 15000
0
1
2
3
4
5
6
Motor Speed (RPM)
T
h
r
u
s
t

(
N
)
0 2500 5000 7500 10000 12500 15000
0
50
100
150
200
250
Motor Speed (RPM)
D
r
a
g

(
N
m
m
)
434 J Intell Robot Syst (2012) 65:417435
0 2 4 6 8 10 12 14 16 18 20
0
5
-5
-10
10
15
20
25
Time (seconds)
P
o
s
i
t
i
o
n

(
m
m
)
0 2 4 6 8 10 12 14 16 18 20
0
20
-20
-40
-60
-80
-100
-120
Time (seconds)
A
n
g
l
e
s

(
d
e
g
r
e
e
s
)
Fig. 16 Closed-loop heave and yaw at 20 mm altitude and 80

yaw: (left) position, (right) Euler angles


but to a much greater extent. At the desired hover
speed, the interior rotor only generates half of the
required thrust. The reasons for this are still under
investigation.
Three sets of experiments were performed on
the full Dysc prototype; 1) heave-yaw open-loop
control on a gravity compensated test stand that
constrained roll and pitch, 2) heave-yaw closed-
loop control through a VICON motion capture
system on a cable-connected gravity compensated
test stand, and 3) roll and pitch open-loop control
on a cable-connected gravity compensated test
stand. For all of these tests, high level control was
calculated off-board on a laptop that would send
joint commands to the Dysc over a CAN tether.
Furthermore, gravity was compensated for with a
cable and a spring to account for the reduction in
thrust from the interior rotor. The heave-yaw test
stand is shown in Fig. 14.
Figures 16 and 17 show the closed-loop control
of the heave and yaw motion using the VICON
motion capture system. Figure 18 shows a roll
response to a maximum step command. As is ev-
ident in the figure, the response is highly coupled
because of the torques imposed by the communi-
cation and power tether wires.
0 2 4 6 8 10 12 14 16 18 20
0
Time (seconds)
S
p
e
e
d

(
R
P
M
)


cmd
measured
0 2 4 6 8 10 12 14 16 18 20
0
Time (seconds)
S
p
e
e
d

(
R
P
M
)


cmd
measured
Fig. 17 Rotor speed control inputs for closed-loop heave and yaw at 20 mm altitude and 80

yaw: (left) exterior rotor


speed, (right) interior rotor speed
J Intell Robot Syst (2012) 65:417435 435
12 12.5 13 13.5 14 14.5 15 15.5 16
0
-20
-40
-60
-80
-100
-120
-140
20
40
Time (seconds)
A
n
g
u
l
a
r

V
e
l
o
c
i
t
y

(
d
e
g
/
s
)


x
y
Fig. 18 Open-loop roll angular velocity response to maxi-
mum step input
Closed-loop gravity compensated hover control
was not demonstrated due to bearing failures in
the interior rotor idler wheels and intermittent
CAN bus failures.
9 Conclusion
This work presents the Dysc, a new type of MAV
that uses the gyroscopic moments generated by
tilting two gimbaled rotors for attitude control.
This type of vehicle holds the promise of simpler
manufacture and maintenance than conventional
helicopters and higher agility than quadrotors.
This design utilizes a novel ring friction drive that
simultaneously supports and drives the rotors, re-
moving the need for heavy and expensive bear-
ings. We presented a simulator that validates the
complete model of the attitude dynamics along
with two controllers; a simple hover controller us-
ing linearized dynamics and a nonlinear controller
using feedback linearization. Both controllers are
shown to stabilize the vehicle around hover in
simulation. The simulator and dynamic model al-
low us to evaluate the design parameters that
have the largest impact on performance. Future
work includes addressing the concern of how to
compensate for external steady moments without
reaching joint or speed limits, and the exploration
of flaps in the downwash of the rotors to compen-
sate for steady external moments.
References
1. Gress, G.R.: Using dual propellers as gyroscopes for
tilt-prop hover control. In: Proc. AIAA Biennial Int.
Powered Lift Conf. Exhibit. Williamsburg, VA (2002)
2. Johnson, W: Helicopter Theory. Dover Publications
(1980)
3. Kendoul, F., Fantoni, I., Lozano, R.: Modeling and
control of a small autonomous aircraft having two tilt-
ing rotors. In: 44th IEEE Conference on Decision and
Control, 2005 and 2005 European Control Conference.
CDC-ECC05, pp. 81448149 (2005)
4. Leishman, J.G.: Principles of Helicopter Aerodynam-
ics. Cambridge University Press (2006)
5. Lim, K.B., Moerder, D.D.: CMG-augmented control of
a hovering VTOL platform. In: AIAA Guidance, Nav-
igation and Control Conference and Exhibit. Hilton
Head, SC (2007)
6. Lim, K.B., Shin, J.Y., Moerder, D.D.: Variable speed
CMG control of a dual-spin stabilized unconventional
VTOLair vehicle. In: AIAA3rd Unmanned Unlimited
Technical Conference, Workshop and Exhibit,
pp. 2023. Citeseer, Chicago, IL (2004)
7. Okamoto, M., Yasuda, K., Azuma, A.: Aerodynamic
characteristics of the wings and body of a dragonfly.
J. Exp. Biol. 199(2), 281 (1996)
8. Prouty, R.W.: Helicopter Performance, Stability, and
Control. RE Krieger Pub. Co. (1990)
9. Samuel, P., Sirohi, J., Bohorquez, F., Couch, R.,
Center, A.G.R.: Design and testing of a rotary wing
MAV with an active structure for stability and control.
In: Annual Forum Proceedings-American Helicopter
Society, vol. 61, pp. 1946. American Helicopter Society
(2005)
10. Schaub, H., Vadali, S.R., Junkins, J.L.: Feedback con-
trol law for variable speed control moment gyros.
J. Astronaut. Sci. 46(3), 307328 (1998)
11. Tewari, A.: Atmospheric and Space Flight Dynam-
ics: Modeling and Simulation with MATLAB and
Simulink. Springer (2007)
12. Thorne, C., Yim, M.: Towards the development of
gyroscopically controlled micro air vehicles. In: In-
ternational Conference on Robotics and Automation
(2011)
Reproducedwith permission of thecopyright owner. Further reproductionprohibited without permission.

You might also like