Anda di halaman 1dari 21

69

Simulator Development for


Transition Flight Dynamics of a
VTOL MAV
Diyang Chua, Jonathan Sprinklea, Ryan Randallb
and Sergey Shkarayev*,b
aDepartment of Electrical and Computer Engineering, The University of Arizona PO Box
210104, Tucson, AZ, USA, 85721-0104
bDepartment of Aerospace and Mechanical Engineering, The University of Arizona PO Box
210119, Tucson AZ, USA, 85721-0119
ABSTRACT
This paper presents the simulation and validation of models for transition between
forward flight and hover for a micro air vehicle (MAV). A dynamical simulation
environment is developed, based on measured aerodynamic properties of a fixed-wing
tail-sitter MAV with vertical take-off and landing (VTOL) capabilities. The simulator is
validated against recorded flight data of transition from forward flight to hover using the
physical platform. Specifically, the flight transition maneuver in the vertical plane
(vertical-to-horizontal and horizontal-to-vertical) is studied in detail. The development
and evaluation of a VTOL MAV called Mini-Vertigo 2 (MV2) is presented. A non-linear
longitudinal dynamic model of MV2 flight is derived, and semi-empirical aerodynamic
formulas for thrust, lift, drag and pitching moment based on wind tunnel testing are
applied to the dynamic model. Nonlinearities during transition to/from hover are
addressed, and (using the developed dynamic model) a numerical vehicle dynamic
simulator is presented. The MV2 autopilot is discussed and flight test data are presented.
Flight data from pitch-up maneuvers was compared to results of simulations. This
comparison enables the future application of the simulation system throughout the flight
envelope. This work presents a significant extension of [1].

1. INTRODUCTION
Development of automatic controls for micro air vehicles (MAVs) with vertical take-off and landing
(VTOL) capabilities presents a challenging research and engineering problem. The development of
such vehicles has been discussed in [2, 3, 4], but the ability of autopilots to robustly control VTOL
MAVs has not been well-established, and autopilot improvement is vital to the progress of VTOL MAV
technology.
Design methodology and general theoretical control principles for large VTOL aircraft are welldeveloped and have been summarized in [5, 6]. There is a sufficient amount of relevant data and
knowledge to begin exploring autonomous flight concepts with applications to VTOL MAVs.
For tail-sitter VTOL MAVs, transition between hover and forward-flight modes is a necessary
abstraction to permit frequent takeoff and landing, but it is also an important maneuver for free flight.
In this paper one such tail-sitter platform is discussed and details of the complex aerodynamics that the
platform exhibits are presented. The platform shown in this paper is capable of rapid transition with
only a small loss in altitude between these two flight modes when under human control, but the
nonlinearities in the system model require careful design to ensure that a simulation environment
accurately reflects the physical platform. In order to automate the behavior of the vehicle under human
control advanced analysis and simulation are needed. This enables prototyping new controllers in
simulation, rather than directly on the hardware of the vehicle, but is only possible if the simulator
closely approximates the vehicles model. Further, existing vehicle autopilots typically present only

*Corresponding

author

Preprint submitted to the International Journal of Micro Air Vehicles August 11, 2010

Volume 2 Number 2 2010

70

Simulator Development for Transition Flight Dynamics of a VTOL MAV

PID control for online behaviors. The development of a simulator that accurately reflects the vehicle
characteristics will permit advanced study of several control strategies, before the embedded controller
design is undertaken to validate behavior on the platform.
The main contribution of this paper is the simulator that is built-upon the aerodynamic features of
the aircraft and which is shown to reflect the behavior of the hardware platform when compared to flight
data. Numerical and physical experiments are used to validate that our simulator results match the
analytical and measured results. Numerical analysis of the steady-state behaviors of the platform
provides a flight envelope that enables future work on robustness and stability of the platform. Through
the use of this simulator future work can push the technological achievement of the platform faster than
through empirical or hardware-based experimentation of controllers for the platform.
1.1. Autopilots for MAVs
Available off-the-shelf autopilots and micro-avionics components (especially micro-sensors) are expensive, relatively heavy, and of poor quality. Industrial developments, being generally proprietary, do not
disseminate knowledge with regard to autonomous MAV design. Some software and hardware designs
are freely available: the result of non-industrial works (e.g., references [7, 8]). Documentation and
schematics for freely available autopilots lack technical support, making them difficult to work with.
Conventional design of large aircraft is a two-step process. First, airframe design focuses on
satisfying performance requirements. Second, control laws are designed for required stability and
controllability. A strong coupling between airframe and control system design has been found for
MAVs [9, 10, 11, 12].
Taylor et al. [13] developed an attitude stabilization system based on thermal-horizon-detection. The
system is reliable both day and night, consumes little power, and operates quickly from a cold start. The
system does not function properly when weather conditions are foggy or when the temperature
differential between the sky and the ground is slight.
MEMS technology has the potential to provide size and weight savings, along with reduced power
consumption. Such technology could be applied to MAV autopilot hardware. The application of microsensors to micro-flight controls has been reviewed in a previous study [14]. Micro-sensors and
controllers suffer inaccuracies and instabilities despite progress in sensor technology. Hybridization is
a potential solution to these problems that combines different sensors into one system.
In [15] the authors used a Kestrel autopilot for their platform. A similar system utilizing a Kestrel
autopilot was developed in [16]. Lastly, a control system was developed in [17] for MAV autonomous
surveillance. The system was used on an aircraft that had a 40 cm wingspan and a total mass of 250 g.
Autopilot augmentation via GPS receiver and telemetry system was utilized with a waypoint navigation
algorithm. The augmented system originally consisted of three piezoelectric gyros, two pressure
sensors (static and dynamic) and a microcontroller. The microcontroller applies control algorithms and
sends signals to actuating servos. Algorithms that were added to the system provide altitude hold,
velocity hold and azimuth control. Proportional control is used to command heading. The altitude hold
and speed hold functions use proportional-integral (PI) control. Newly developed control laws were
integrated into a ground station which allowed gain factors and waypoints to be modified during flight.
A micro-scale inertial measurement system, the MR01 [18] was developed for the micro-scale
autopilot system, and used in the Konkuk Batwing platform discussed in [19]. The MR01 consists of a
one-axis gyroscope sensor and a two-axis accelerometer. Longitudinal and lateral stabilities improved
when attitude data measured by the MR01 was used as feedback. Successful surveillance missions have
been flown using MAVs with 13-15 cm wingspans in 5 m/s headwinds.
1.2. Simulation and Control in Transition
R. H. Stone et al. [20] proposed a controller design for transition maneuvering of a tail-sitting T-wing
aircraft. Dynamic equations of motion were presented and numerical simulations of the closed-loop
controller were performed. Euler angles were utilized for the horizontal flight mode and a /2 (90 deg)
tilted frame was used for the vertical (hovering) mode. For transitions, a quaternion based frame was
utilized. Linear control laws for vertical, horizontal and transition modes of flight were implemented in
the custom autopilot. The vehicle was also flown in autonomous take-off and transitions. The aircraft had
some difficulties with horizontal-to-vertical transition, when the angle of attack exceeded /3 (60 deg).
Control laws developed by Knobel et al. [9] for tail-sitter UAVs feature quaternion-based attitude
estimation. Simulation, using commanded pitch and thrust, produced satisfactory results (in terms of
International Journal of Micro Air Vehicles

Diyang Chu, Jonathan Sprinkle, Ryan Randall and Sergey Shkarayev

71

positional error) along both horizontal and vertical axes.


Transition was also studied by Johnson et al. [21] and demonstrated on a much larger vehicle (a 33%
scale Edge 540T, with a wingspan of over 8 ft) than the MAV considered in this paper. In that work, an
adaptive controller is used for the inner loop stability control, and through velocity and pitch commands
in the outer loop (linearly decreasing/increasing over time) the transition from horizontal to hover was
achieved. That vehicle featured direct-sensing for position and gyro acceleration values, as well as
altitude and airspeed sensors.
Work on a smaller platform [22] also studied and implemented transition between horizontal flight
and vertical hover. However, this vehicle was large relative to typical MAVs. The vehicle was capable
of reliable indoor sensing which provided accurate values for various flight parameters (which were
acquired at high frequency). An open-loop control algorithm was used for the inner-loop controller in
transition flight mode.
1.3. Platforms and Experimentation
Previous work by the authors [10, 11] has focused on the development of fully autonomous MAVs since
2003. Several research and development studies conducted on the aerodynamics, airframe, and control
design of MAVs have been completed. Exceptional fully autonomous MAVs have been designed, built
and flown. One such MAV is called Mini-Vertigo 2 (MV2) [2]; it is used in the present study and will
be described in the next section.
For MV2, an error-based feedback design (with static gains) may be insufficient for several reasons.
First, position and orientation are subject to error-accumulation during flight-testing due to the limited
space for sensors (and the additional weight they would produce). The slipstream, which is induced by
the propulsion system, causes a nonlinear scaling of lift, drag, and pitching moment as thrust is
increased. Gains for a closed-loop controller would need to be nonlinearly scaled based on the flight
condition. Changing MV2 from horizontal mode to hover mode under human pilot control can be done
without a significant loss or gain in altitude. The techniques applied to the larger scale vehicle in [21],
on the other hand, results in a climb of approximately 30 m during the transition, which took
approximately 10 sec to complete.
There are several ways to address the issue of transition between horizontal and hover mode for MV2:
(1) PID control, which mirrors available technology for autopilot control;
(2) piecewise hybrid controller which changes the PID gains based on the current angle of attack;
(3) a strictly open-loop controller based on pitch-error and elevon/throttle saturation which
utilizes sinusoids as the error goes to 0; and
(4) a feed-forward controller in which state targets are tracked.
In this paper we validate our simulator with results from controller (1), an approximation of the
existing vehicle control hardware. The remaining controller designs are left for future work, noting that
the validation results presented in this paper permit advanced testing of the remaining controllers in
simulation.
The simulator presented in this paper uses existing wind tunnel data, with a nonlinear vehicle model,
to inform design tradeoffs. In a series of aircraft flight tests, telemetry data on control actuation,
altitude, and attitude of the aircraft were collected and analyzed. These data were used in the validation
of the simulator.

Table 1. Components of MV2

Volume 2 Number 2 2010

72

Simulator Development for Transition Flight Dynamics of a VTOL MAV

2. AUTONOMOUS VTOL MAV DESIGN AND SPECIFICATIONS


Mini-Vertigo 1 (MV1) was the predecessor to MV2, and utilized a flat wing which was later replaced
with a cambered reflexed wing. This newer iteration was dubbed MV2; it is described in [2] and
depicted in Fig. 1. MAV specifications and components are presented in Tables 2 and 1, where Ixx, Iyy,
Izz are moments of inertia about the x,y,z axes respectively. MV2 has dorsal and ventral fins for lateral
stability. The total mass of the airplane is 243 g. The controls for this aircraft include elevons (which
control roll and pitch) and a rudder (which controls yaw). MV2 is a tail-sitter which is capable of
vertical take-off, hover, transition, and level flight. Thanks to MV2s 50 g payload capacity it can be
equipped with a video camera capable of capturing and transmitting video to a ground station. MV2 is
an effective tool for surveillance both indoors and out. An autopilot is currently integrated into the
airplane which is capable of automatic hovering, horizontal flight and transition.
Both MV1 and MV2 are outfitted with a Paparazzi autopilot [7]. The autopilot includes an autopilot board, two double-axis MEMS gyroscopes (IDG300), an infrared attitude sensor (FMA Co-Pilot)
and ultrasonic range sensor (Maxbotix LV-EZ1). The autopilot board reads sensor feedback signals and
generates control signals to the motor speed controller, rudder and elevon servos. The autopilot board
features an ARM7 processor running at 60 MHz which performs control loops, navigation, data
acquisition, and telemetry communications.
The attitude of MV2 is provided by infrared sensors. In flight, the autopilot sends telemetry data
back to the ground station using XbeePro modems. Currently, the telemetry data include: GPS-based
location coordinates, groundspeed, altitude, climb-rate, autopilot status, and control-surface position.
Telemetry data (such as location, attitude and altitude) are displayed (and can be recorded) on a groundstation laptop, which is regularly updated.

Figure 1. Mini-Vertigo 2

3. DYNAMIC EQUATIONS OF MOTION FOR VTOL MAVs


A control law design process that makes use of the longitudinal equations of motion is followed herein.
Fig. 2 clarifies nomenclature used in subsequent expressions, where L,D, MAW are wing lift, drag and
moment at the wings aerodynamic center, T is thrust force generated by the propulsion system, V is
free-stream velocity, is wing angle of attack, is pitch angle, m is the mass of the aircraft, e is elevon
(elevator) deflection. The x,y,z coordinate system is a body-fixed frame with its origin at the center of
gravity of the aircraft (Fig. 2). The xE ,yE ,zE coordinate system is Earth-fixed with arbitrary origin and
is oriented as shown.

International Journal of Micro Air Vehicles

Diyang Chu, Jonathan Sprinkle, Ryan Randall and Sergey Shkarayev

73

Table 2. MAV specifications and properties

Figure 2. Dynamic model of the vehicle

3.1. Simulation Model Derivation


Kinematic and dynamic expressions were used to describe the motion of the aircraft. Aerodynamic forces
were empirically determined and modeled using wind tunnel test data. The equations of motion are:

(1)

where W is the weight of the aircraft, xAC, zAC are offsets of the aerodynamic center of the wing from
the center of gravity of the aircraft in the body-fixed frame. The expressions in (1) were rotated from a
Volume 2 Number 2 2010

74

Simulator Development for Transition Flight Dynamics of a VTOL MAV

body-fixed coordinate system to an Earth-fixed system using a standard rotation matrix. Trigonometric
product-to-sum formulas were used for simplification. The following expression results:

(2)

The y-axis is initially parallel to the yE axis and the direction of the y-axis does not change as a result
of pitching. The third expression in (1) is not affected by rotation; it remains the same in the Earth-fixed
system.
Expressions for lift, drag and moment in terms of non-dimensional coefficients (and with parameters
of the airframe given in Table 2 held as variables), are:
(3)

(4)

(5)
where CL,CD,Cm are non-dimensional coefficients for wing lift, drag, and moment respectively. S is
wing area (planform area of the wing)), is local air density, c is the mean aerodynamic chord of the
wing.
Free-stream velocity is given by
. The equations of motion can be expressed in
terms of dimensionless coefficients as:

(6)

4. VTOL MAV AERODYNAMICS


Unlike conventional aircraft, the effect of a propeller slipstream on wing aerodynamics may not be neglected for VTOL aircraft. Wind tunnel testing was conducted [2, 3] for a motor, a wing, and an
arrangement of a wing with a motor. The aerodynamic forces on the wing were produced by two mixing
airflows: the free-stream and the propeller-induced pulsating slipstream. Semi-empirical aerodynamic
models presented in this section take into account the effect of the slipstream.

International Journal of Micro Air Vehicles

Diyang Chu, Jonathan Sprinkle, Ryan Randall and Sergey Shkarayev

75

Table 3. Reynolds number (Re) and equivalent free-stream velocity

4.1. Thrust
Previous work [2] noted a quadratic variation of thrust, T , with free stream velocity, V . A series of
wind tunnel tests were conducted on a propulsion system consisting of two coaxial contra-rotating
propellers [3]. The following parameters were varied: throttle setting (t ), free-stream velocity (V ) and
angle of attack ().
The relationship between free-stream velocity and Reynolds number is provided in Table 3. The
paper [3] presented data that was used to formulate a semi-empirical model describing thrust as:

(7)

where t0 is an arbitrary throttle setting about which the model is centered. Note that thrust varies
linearly with throttle-setting and quadratically with free-stream velocity. Throttle-setting was used in
the model rather than a more physically-significant non-dimensional parameter (e.g., propeller advance
ratio). We leave to future work determination of the relationship between propeller speed and throttlesetting. The autopilot directly controls throttle-setting and MV2 does not actively sense propeller
rotation during flight.
4.2. Lift Coefficient
A model of lift was suggested in previous work [3]. The wing and propulsion system were tested under
various conditions and the propulsion system without the wing was tested under the same conditions.
The difference between the two sets yielded aerodynamic forces for the wing alone. An inertial tare was
also utilized for determination of wing-induced moment. The model is broken down into three distinct
domains corresponding to pre-stall, stall, and post-stall. The pre-stall and post-stall regions are described
by linear polynomials; the stall region by a quadratic polynomial. The model is continuous with , but
not smooth. The following expression provides lift-coefficient variation with angle of attack.

(8)

CL is the coefficient lift-curve slope in the pre-stall region


CL(0) is the coefficient of lift at = 0
CL(1) is the coefficient of lift at = 1
1,2,3 are angles of attack where: the lift-curve slope diverges from linearity, the wing stalls, and the
post-stall lift coefficient equals zero, respectively.
The constituent parameters in (8) may vary with V and t (which have restricted allowable values based
on the scope of wind tunnel testing [3]). Fig. 3 shows CL for varying , Re and t .

Volume 2 Number 2 2010

76

Simulator Development for Transition Flight Dynamics of a VTOL MAV

CL: Nonlinear relationships with Re,

2.5
Re=43000
Re=87000
Re=130000

CL

1.5

0.5

0. 5!
0

10

20

30

40

50
(deg)

60

70

80

90

Figure 3. CL vs. , itemized by (Re, t ) pairs

4.3. Drag Coefficient


The coefficient of drag model was broken down into two regions: pre-stall and post-stall. The variation
of drag with angle of attack was parabolic in the pre-stall region and was linearly approximated in the
post-stall region [3].

(9)

fi are coefficients that vary with V and t


CD(2) is the coefficient of drag at = 2
CD is the drag-curve slope in the post-stall region; it varies with V and t
Fig. 4 shows CD for varying , Re and t.

International Journal of Micro Air Vehicles

Diyang Chu, Jonathan Sprinkle, Ryan Randall and Sergey Shkarayev

77

CD: Nonlinear relationships with Re, t


3.5
Re=43000
Re=87000
Re=130000

2.5

CD

1.5

0.5

10

20

30

40

50
(deg)

60

70

80

90

Figure 4. CD vs. , itemized by (Re, t ) pairs

4.4. Moment Coefficient


The moment coefficient was modeled in three separate regions roughly corresponding to pre-stall, stall,
and post-stall. The coefficient variation with angle of attack was linear in all three regions.
(10)
(11)

4,5 are boundaries between regions; they vary with V


Cm2 ,Cm3 are moment curve slopes in the 2nd and 3rd regions; they vary with V
Cmt ,Cme are changes in moment coefficient with t and e; they vary with V .
Fig. 5 shows Cm for varying , Re and t .

Volume 2 Number 2 2010

78

Simulator Development for Transition Flight Dynamics of a VTOL MAV


Cm: Nonlinear relationships with Re,

0.8

0.6
Re=43000
Re=87000
Re=130000

0.4

Cm

0.2

0. 2

0. 4

0. 6

0. 8 !
0

10

20

30

40

50

60

70

80

90

(deg)

Figure 5. Cm vs. , itemized by (Re, t ) pairs. e = 0.4363 (25 deg) for these curves

Nonlinearities complicate modeling and calculations. Wind tunnel testing [3] showed that lift
coefficient is nonlinearly dependent on angle of attack, free-stream velocity, and throttle-setting (inputs
during testing).
Drag coefficient is nonlinearly dependent on free-stream velocity. To evaluate candidate controllers,
various numerical approximation schemes were developed which directly use the models that
synthesized the curve-fits shown in Figures 3-5. Solutions for relevant ODEs were obtained using the
continuous time integration features of Simulink. Numerical expressions for the coefficients were
obtained through embedded MATLAB functions (to increase simulation speed).
In Figures 3-5 the curves are not smooth, and the piecewise linear approach leaves cusp points.
Although this discontinuity in the derivative may result in some issues for controllers, it was not a
major factor in our design, and we discuss this in the next section.
5. VEHICLE DYNAMICS SIMULATOR
To investigate VTOL MAV transition a vehicle dynamics simulator was developed (Fig. 9). In the
process of developing the simulator, numerous exercises were performed to validate the computational
implementations of the nonlinear parameters. The design strategy began by validating the stability of
the vehicle in steady state flight, and continued from there to the pitch control simulations.
5.1. Calculation of Steady State Flight
MAV dynamics are described by the three equations in expression (6). The semi-empirical model for
aerodynamic system properties ((6)-(11)) is highly nonlinear; it is not feasible to use direct analysis to
determine steady state points so a numerical scheme was utilized.
First, steady state behavior was tested at several points in the flight envelope. By definition, the
steady state flight satisfies the following conditions: xE = zE = E = E = 0. The state space was searched
for values of the 5 state variables, V, , t ,e , that approximately satisfied those conditions. The
International Journal of Micro Air Vehicles

Diyang Chu, Jonathan Sprinkle, Ryan Randall and Sergey Shkarayev

79

algorithm first searches over a reasonable range of pitch-angle, throttle setting, velocity and angle of
attack. Values of these variables are determined which result in x and z accelerations that are smaller
than a user-specified error threshold. Next a practical range of elevon angles, e, are searched until is
smaller than a specified threshold. The algorithm then stores the approximate values of the five state
variables (which is considered to be within an acceptable neighborhood of the steady state point). If the
determined steady state point is considered to be unsatisfactory the search range of the state variables
can be reduced around the roughly determined steady state point. The search step size can then be
reduced for all of the variables.
After sufficient iteration a steady state of acceptable accuracy is obtained. Fig. 6, Fig. 7 and Fig. 8
show the estimated MV2 steady state conditions for various parameter values. The dashed box on each
of the plots encloses the points within the domain of the data used to develop the empirical
aerodynamics model.
5.2. Controller for Validation of Steady State Conditions
The dynamic model (from Fig. 9) was validated using a subset of the offline calculated steady state
values as initial conditions. During the search for steady state points, the search resolution varied
depending on the variable (e.g., a step size of 0.01 was common), so accumulation of error could result
in flight instability after some time due to the intrinsic instabilities of the platform.
This motivated the introduction of a steady state controller that is permitted to make small ([1,1]
deg) changes in e based on the assumption that = 0 in steady state. With this PD controller, the
vehicle can fly arbitrarily long at steady state points, whereas without this controller oscillations
resulted in instability after only a few seconds. Although the details of the controller are interesting,
detailed discussion is left to future papers. It is notable that this controller can be roughly compared to
the infrared sensors onboard the MV2 platform, which improve attitude control.

20

18

16

Velocity:m/s

14

12

=0deg
=10deg

10
=20deg
8

=30deg
=40deg
=80deg

=50deg
=90deg

=60deg
=70deg
4
0.4

0.5

0.6

0.7

0.8

0.9

Figure 6. Velocity vs. throttle setting. The dashed box in each figure indicates the wind tunnel testing
domain

Volume 2 Number 2 2010

80

Simulator Development for Transition Flight Dynamics of a VTOL MAV


90

80

70

60

=70deg

(deg)

50
=60deg

=90deg
=50deg

40

=40deg

=80deg

=30deg

30

=20deg
=10deg
20

=0deg

10

0
0.4

0.5

0.6

0.7

0.8

0.9

Figure 7. Angle of attack vs. throttle setting. The dashed box in each figure indicates the wind tunnel
testing domain

15
=90deg

10

=70deg
=80deg

5
=60deg
=50deg
0
(deg)

=40deg

=30deg
5

=20deg
=10deg

10

=0deg

15

20
0.4

0.5

0.6

0.7

0.8

0.9

Figure 8. Elevon deflection angle vs. throttle setting. The dashed box in each figure indicates the wind
tunnel testing domain

International Journal of Micro Air Vehicles

Diyang Chu, Jonathan Sprinkle, Ryan Randall and Sergey Shkarayev

81

Figure 9. Vehicle dynamics simulator

5.3. Physical Constraints


The highest simulated pitch angle that MV2 can achieve with the existing hardware platform configuration without descending is 75 deg. To fly without descent above 75 deg requires a throttle setting that
exceeds the scope of the wind tunnel tests, upon which the dynamic model is based. During horizontalto-vertical transition the upward force changes as a result of two phenomena. First, as the pitch angle
increases the angle of attack increases. If the transition maneuver is properly performed then the angle
of attack will never exceed the stall angle of attack during transition (vertical dashed line in Fig. 3 for
V = 15 m/s;t = 0.60). Prior to stall, wing lift increases with angle of attack (Fig. 3), which increases
the wings contribution to vertical force. Second, as MV2 pitches upward the thrust vector tilts upwards
which increases vertical force. These combined contributions to vertical force can exceed the weight of
MV2 and result in an upward acceleration unless thrust is reduced during transition to prevent climb.
To reach higher steady state pitch angles, the throttle setting must be increased as shown in Fig. 8.
The empirical model was developed from wind tunnel experimentation that did not exceed a throttle
setting of t = 0.70. To reach steady-state pitch angles greater than 70 deg the dynamic model would
have to be applied to throttle settings outside of the wind-tunnel testing domain (Fig. 8). Parameter
values outside of the testing domain are estimated based on known values within the domain (but
outside values have not been experimentally verified). The multiple points at each throttle setting
represent estimated steady state flight conditions. The true steady state solution is expected to lie within
the streak of approximate solutions.
6. CONTROL LAW DESIGNS
The dynamic simulator was applied to several different control scenarios to validate steady-state flight,
and then a feedback controller was used to investigate the behaviors of the simulated vehicle in a
similar scenario to the hardware implementation.
6.1. Open-Loop Control Validation
To validate the simulator, we started the vehicle in a steady state condition in forward flight. We then
selected a rate of increase for the throttle, and a rate of decrease for elevon deflection until it reached
saturation (e 30 deg). Several families of rates were tested, to check the stability of the simulator
within the wind tunnel data ranges. In these validations, the vehicle pitched up, but for some throttle or
elevon deflection rates was not stable when the control inputs saturated. This is not surprising, given
the nonlinearities in the vehicle model, and inherent instability when flying at low speeds without
closing the loop around observed error.

Volume 2 Number 2 2010

82

Simulator Development for Transition Flight Dynamics of a VTOL MAV

6.2. Closed-Loop Transition Maneuver


After validating steady state flight and open-loop behaviors, a closed-loop controller was applied to
transition the vehicle from one steady state region to another. This controller used as its two control
inputs thrust, t, and elevon deflection, e. In order to mirror the existing hardware controller, PD and
PID controllers were developed and simulated. We performed the maneuver with two proportional gain
settings, one for low-rate transition, and one for high-rate transition. The high-rate transition, shown in
Fig. 10, converges faster on the commanded values with its higher gains, but with more overshoot, and
large changes in control inputs. The low-rate transition, shown in Fig. 11 (and used to create Fig. 13)
has smaller gains, and is a less-aggressive maneuver. The PD controller for thrust is based on the error
of the altitude of the MAV.
(12)
Where zEerr = zEcommand zE, KP and KD are proportional, derivative gain coefficients. The low-rate
transition used constant values of KP = 0.005 and KD = 0.5. We constrain t between [0,1]. During
the transition, once the altitude of the MAV decreases, the thrust will be increased to get enough lift to
maintain the altitude. The error between the commanded and measured pitch angle is used to calculate
elevon deflection.
(13)
Where err = command , KI is the integral gain coefficient. The low-rate transition used constant values
of KP = 0.8, KI = 0.03, KD = 0.8. We constrain e to be within the physical range of the vehicle.
The simulation result with the above controller settings are shown in Fig. 11 and Fig. 13. In order to
improve the performance of the closed-loop controller, different P values of the two controllers were
tested. For the t controller, KP = 0.250 is used while for the e controller KP = 40 is used, this highrate transition results are shown in Fig. 10.
In the simulation, the t controller was given a constant reference input zEre f , as the altitude change
during the transition process should be as small as possible. The reference input for e is ref as in
Equation (14), which contains two ramps, one is for flight-to-hover transition started at 10 s and the
other is for hover-to-flight transition started at 40 s, here flight is about 13 deg and hover is about 70
deg. Both controllers sample from their environment at 5 Hz. MV2 begins the transition in steady state
flight; from Fig. 6, the initial pitch angle is in the range of 10 to 20 deg, so the lowest initial speed of
MV2 is around 11 m/s.
s
s
s
s

(14)

6.2.1. Differences in t and e


Simulated data from the high-rate transition maneuver is shown in Fig. 10. Closer examination of Fig.
10a and Fig. 10b (data for t and e) shows that the increase of KP makes the system sensitive to the
error of the system. When the transition begins at 10 s and 40 s, t and e spike for the high-rate
transition. In the low-rate transition, t (Fig. 11a) does not spike as high, and e (Fig. 11b) has fewer
oscillations. For the t value, the big difference between the two cases is around 45 s, where the highrate transition produces a peak value of about 0.72, while the low-rate case does not exceed the thrust
necessary for steady-state flight while hovering.
6.2.2. Differences in
Although the oscillation of e in high-rate transition is relatively large, the result does track the commanded control input quite well as in Fig. 10d. Under the low-rate transition setting, the only
performance difference for the forward-to-hover transition was that it took about 10 s to finish the
transition while the high-rate transition only needs 5 s.
International Journal of Micro Air Vehicles

Diyang Chu, Jonathan Sprinkle, Ryan Randall and Sergey Shkarayev

83

10

0.75

12
0.7

14
16

0.65

(deg)

0.6

20

18

22
0.55

24
26

0.5

28
0.45

20

40

60

80

30

100

20

40

60

80

100

60

80

100

80

100

time(s)

time(s)

45

80

40

70

35
60

(deg)

(deg)

30

25

50

40
20
30
15
20

10

20

40

60

80

10

100

20

40

time(s)

time(s)

16
0.2
0.15

14

0.1

12

0.05
0
Vz(m/s)

V(m/s)

10

0.05
0. 1
0.15

0. 2
0.25

0. 3

20

40

60

80

time(s)

Figure 10. High-rate transition simulation

Volume 2 Number 2 2010

100

20

40

60
time(s)

84

Simulator Development for Transition Flight Dynamics of a VTOL MAV


10

0.75

12
0.7

14
16

0.65

(deg)

0.6

20

18

22
0.55

24
26

0.5

28
0.45

20

40

60

80

30

100

20

40

60

80

100

60

80

100

80

100

time(s)

time(s)

45

80

40

70

35
60

(deg)

(deg)

30

25

50

40
20
30
15
20

10

20

40

60

80

10

100

20

40

time(s)

time(s)

16
0.2
0.15

14

0.1

12

0.05
0
Vz(m/s)

V(m/s)

10

0.05
0. 1
0.15

0. 2
0.25

0. 3

20

40

60

80

time(s)

100

20

40

60
time(s)

Figure 11. Low-rate transition simulation

International Journal of Micro Air Vehicles

Diyang Chu, Jonathan Sprinkle, Ryan Randall and Sergey Shkarayev

85

6.2.3. Differences in
Although the shapes of the curves are similar between Figures 10c and 11c, the values differ
drastically. In Fig. 10c, the high-rate transition produces around 37 degrees when at high-pitch-angle,
and in final forward flight is approximately 7 degrees. The low-rate transition produces a higher
value in these modes (nearly 45 degrees in high-pitch-angle, and 15 degrees when returning to forward
flight). This is explained by the faster transition speed of the high-rate transition, which preserves a large
velocity component in xE -direction, and permits a smaller value for steady-state flight. As that speed
drops, must increase, as demonstrated by the low-rate transition.
6.2.4. Differences in velocity
The velocity of MV2 is the combination of velocities in the xE -direction and the zE -direction, as
the velocity in the zE -direction is small, so velocity in the xE -direction is very close to the total
velocity and can be considered as such.
In high-rate transition, as the first transition is 5 s faster than the low-rate case, the drag force has less time
to decrease the the xE -direction velocity, so the high-rate case has 4 m/s velocity at high-pitch-angle while
the low-rate case only has 3.3 m/s. After the second transition, as the high value of speed in the xE -direction
also makes smaller, from Fig. 3 it can be seen that the velocity should be larger to maintain the lift when
the lift coefficient is smaller. So in the final forward flight, the high-rate case has a larger value than the
low-rate case. From the zE -direction velocity plot, the altitude change can be estimated. The altitude
change is very small considering the value of Vz in Fig. 10f; For low-rate transition (Fig. 11f), the altitude
change will be a bit larger than in the high-rate case. Also, recall that the zE -direction is positive
downwards, so the vehicle increases in height during the transition. As zEref is used as the control reference
for t , the higher P controller value not surprisingly tracks the reference value zEre f better.
Although decoupling is needed for the implementation of controllers for general nonlinear MIMO
systems, it is not necessary to solve the coupling problem in this case because the moment equilibrium
equation and vertical equilibrium equations (6) are not strongly coupled. The velocity in zE -direction
is very small, so it will take some time to have some change in altitude. This means that once there is
a small decrease in altitude, the PD controller will have enough time to increase thrust to provide lift
and maintain altitude. After some tuning of the parameters of the two controllers, the transition is
comparable to that of the flight system, discussed in the next section.
7. VALIDATION OF FLIGHT AND SIMULATION
MV2 featured an autopilot stability augmentation mode [10, 11] which was employed during hovering
and transition flight tests. The response of MV2 to prescribed altitude control commands was
investigated via flight testing in [23]. Telemetry data on altitude response, angular velocities, and servo
signals were acquired during autonomous flights. Control gains of hardware PID controller were
empirically tuned for stability, and effective response to prescribed altitude commands.
7.1. Flight Test Results
The performance of the autopilot in terms of the altitude hold feature is shown in Fig. 12. The aircraft
demonstrates fast response with peak amplitude of about 0.1m. The average amplitudes of oscillation
were found to be less than 0.1 rad/s for pitch, 0.2 rad/s for roll and 0.1 rad/s for yaw. The throttle
commanded by the autopilot to reach the desired altitude is presented in Fig. 14. The changes in throttle
setting are very small, as the full-range of throttle setting is normalized to [0,1].
Transition maneuvers from low-pitch-angle flight to high-pitch-angle flight were investigated in
additional flight tests. The aircraft was flown in stability augmentation mode. The pilot manually
commanded the change of the pitch angle through the standard RC controller box. Altitude and ground
speed were not held during the flight, and were controlled by the test pilot. All angular rates were
controlled by the autopilot using gyroscopes and PID controllers.
The variation of the pitch angle with time is shown in Fig. 15. The aircraft demonstrated a fast
response to the commanded pitch-up maneuvering. The onboard pitch-rate gyroscope showed the
maximum pitch rate of 1.745 /s (100 deg/s) during transition.
7.2. Comparison with Simulator
The simulation result, compared with flight data, is also shown in Fig. 15. As discussed in the previous
section, the simulator is also using PID controllers, which are tuned to approximate the hardware
Volume 2 Number 2 2010

86

Simulator Development for Transition Flight Dynamics of a VTOL MAV

controller behavior. One PID controller is using the elevon deflection to control the pitch angle , and
the other PD controller is using thrust to maintain the altitude of the MAV. From Fig. 15, the simulation
has a similar response and overshoot to the flight test data. This provides a validation of the behavioral
model for this range of parameters. Notably, the simulation does not include disturbances such as
sensor error or environment (e.g., wind), and thus some deviation from flight data is expected.

Figure 12. Measured and desired altitudes of MV2 during flight [23]

vertical(m)

50
0
50

t=0
0

t=10s go to
hover mode
100

t=40s go back to forward flight


200

300

400
500
horizontal(m)

600

700

800

900

Figure 13. Simulation result of low-rate transition of MV2

Figure 14. Throttle setting during flight [23]

International Journal of Micro Air Vehicles

Diyang Chu, Jonathan Sprinkle, Ryan Randall and Sergey Shkarayev

87

80
70

(deg)

60
50
40
30

Flight test data


Simulation result

20
10
1395

1400

1405

1410
1415
time(s)

1420

1425

1430

Figure 15. Comparison of flight test data with simulator results. Note that flight test data includes
noise from wind and measurement, not present in the simulation model

8. CONCLUDING REMARKS
This paper reviewed recent VTOL MAV studies and introduced Mini-Vertigo 2 (MV2). Mini-Vertigo 2
is a tail-sitter VTOL MAV that was developed at the University of Arizona and has proven itself for
horizontal flight, hover, and transition. MV2 is equipped with an autopilot and video camera and is
capable of autonomous, surveillance flight.
The transition phase of flight between horizontal and hover was the focus of this paper. Dynamic equations of motion were derived for a VTOL MAV configuration similar to that of MV2. Empirical models
of wing aerodynamics were developed from wind tunnel tests to accurately quantify the nonlinear effects
of longitudinal flight parameter variation. A vehicle dynamics simulator, based on these aerodynamic
models, was created and validated with calculated steady state flight behaviors, as well as transition
behaviors recorded with flight data. Simulator results may be applied to future autopilot design.
The use of static PID gains for transition between hover and horizontal flight is sufficient for
transitions at low speeds. However, for high-speed transitions, alternate techniques must be explored in
order to prevent stall or significant overshoot of the terminal steady state values. Thus, some kind of
intelligent approach to the transition controller is required.
Among the many reasons that transition (between horizontal and vertical flight modes) is complex
is the nonlinear behavior of many of the system parameters as the angle of attack increases from 0-/2
(0-90 deg). As discussed in Osbornes work [24] there are regions of continuity for CL and CD across
the angle of attack range, but there are also several non-smooth points. Given the eventuality that nondifferentiable points or discontinuities will exist in these parameters, special attention must be given to
how the gains might be modified to account for these shifts in the plant model, and ensure that any
changes to the controller appropriately correspond to changes in the dynamics of the vehicle.
In a series of stability augmentation mode flight tests, the PID controls were adjusted and
telemetry data on control actuation, altitude, and attitude were collected for analysis of control designs.
Stability augmented test flights demonstrated the ability of the aircraft to fly at high pitch angles,
including hover of up to 1.658 (95 deg), as well as at low pitch angles of about 0.08727 (5 deg), and
transition between them in a stable, controlled manner. Future work will validate these results with the
simulator presented in this paper.
With the validated simulation model, controllers that consider the nonlinear nature of the vehicle can
be investigated and prototyped without the need to perform extensive field tests. For example, we can
consider a piecewise hybrid controller, designed to utilize a different control law for each region of
CL,CD,Cm, by changing controller gains while in a certain region of the statespace. Another approach,
as shown in [26], is to explore the use of open-loop controllers, thus removing a dependency on time
for the value that our controller tracks. Finally, the use of feed forward control techniques could
mitigate errors in the model which are treated as a disturbance.
Volume 2 Number 2 2010

88

Simulator Development for Transition Flight Dynamics of a VTOL MAV

ACKNOWLEDGMENTS
This project was sponsored by a grant from the RDECOM (Program Manager Eric Stierna), as well as
AFOSR (Program Manager David Luginbuhl). The authors also would like to thank Roman
Krashanitsa, Gavin Kumar and David Addai for their contributions to the aerodynamic design of the
aircraft, and the anonymous reviewer for their insightful comments and suggestions.
REFERENCES
[1] D. Chu, J. Sprinkle, R. Randall, S. Shkarayev, Automatic Control of VTOL Micro Air Vehicle
During Transition Maneuver, in: AIAA Guidance, Navigation and Control Conference, AIAA,
2009, AIAA2009-5875.
[2] S. Shkarayev, J. Moschetta, B. Bataille, Aerodynamic Design of Micro Air Vehicles for Vertical
Flight, Journal of Aircraft 45 (5) (2008) 17151724.
[3] R. Randall, C.-A. Hoffmann, S. Shkarayev, Aerodynamics of VTOL Micro Air Vehicles from 090 Degrees Angle of Attack, in: Journal of Aircraft, 2010 (accepted).
[4] J. Moschetta,S. Shkarayev, On Fixed-Wing Micro-Air Vehicles with Hovering Capabilities, in:
46th Aerospace Sciences Meeting and Exhibit, Reno, NV, 2008, AIAA-2008-0221.
[5] B. W. McCormick, Jr., Aerodynamics of V/STOL flight, Dover Publications, Mineola, NY, 1999,
ISBN: 0486404609.
[6] D. Kohlman, Introduction to V/STOL airplanes, Iowa State Press, 1981.
[7] A. Drouin, P. Brisset, PaparaDzIY: DoIitYyourself UAV, 4th European Micro-UAV Meeting,
Toulouse, France (Sept. 15-17 2004).
[8] R. Dixon, Open Source Software Law, Artech House, 2004.
[9] N. Knoebel, S. Osborne, D. Snyder, T. McLain, R. Beard, A. Eldredge, Preliminary Modeling,
Control, and Trajectory Design for Miniature Autonomous Tailsitters, in: AIAA Guidance,
Navigation, and Control Conference, 2006, AIAA 2006-6713.
[10] R. Krashanitsa, G. Platanitis, D. Silin, S. Shkarayev, Autopilot Integration Into Micro Air
Vehicles, in: Introduction to the Design of Fixed-Wing Micro Air Vehicles Including Three Case
Studies, AIAA Education Series, AIAA, Reston, VA, 2006, Ch. 3, pp. 109149.
[11] R. Krashanitsa, G. Platanitis, D. Silin, S. Shkarayev, Aerodynamics and Controls Design for
Autonomous Micro Air Vehicles, in: AIAA Atmospheric Flight Mechanics Conference and
Exhibit, Keystone, Colorado, 2006.
[12] I. Kajiwara, R. Haftka, Simultaneous Optimum Design of Shape and Control System for Micro
Air Vehicles, in: 30th Plasmadynamics and Lasers Conference, Norfolk, 1999.
[13] B. Taylor, C. Bil, S. Watkins, G. Egan, Horizon Sensing Attitude Stabilisation: A VMC Autopilot,
in: International UAV Systems Conference, Bristol, UK, 2003.
[14] R. Arning, S. Sassen, Flight Control of Micro Aerial Vehicles, in: AIAA Guidance, Navigation,
and Control Conference and Exhibit, 2004, pp. 112.
[15] T. Millet, J. Yates, B. Millar, N. Johnson, Brigham Young University MAV Team, 3nd USEuropean Micro Air Vehicle Competition, Toulouse, France (September 2007).
[16] University of Florida Micro Air Vehicle Laboratory, Development of Flexible Wing Autonomous
MAV, 3rd US-European Micro Air Vehicle Competition, Toulouse, France (September 2007).
[17] H. Quix, W. Alles, Design and Automation of Micro air Vehicles, in: First European Micro Air
Vehicle Conference and Flight Competition, German Institute of Navigation, Germany, 2004.
[18] D. Chung, J. Ryu, I. Nam, K. Jo, K. Yoon, H. Huang, , J. Kim, Development of Fixed Wing MAV
Batwing at Konkuk University, Seoul, South Korea, 9th International Micro Air Vehicle
Competition (2005).
[19] A. Suhariyono, J. H. Kim, N. S. Goo, H. C. Park, K. J. Yoon, Design of Precision Balance and
Aerodynamic Characteristic Measurement System for Micro Aerial Vehicles, Aerospace Science
and Technology 10 (2) (2006) 92 99. doi:DOI:10.1016/j.ast.2005.10.004.
URL http://www.sciencedirect.com/science/article/B6VK2-4HPD2TR-2/2/
dc7ace23792c8d1907ecf344e7a21f6f

International Journal of Micro Air Vehicles

Diyang Chu, Jonathan Sprinkle, Ryan Randall and Sergey Shkarayev

89

[20] R. Stone, P. Anderson, C. Hutchison, A. Tsai, P. Gibbens, K. Wong, Flight Testing of the T-Wing
Tail-Sitter Unmanned Air Vehicle, Journal of Aircraft 45 (2) (2008) 673685.
doi:10.2514/1.32750.
[21] E. Johnson, A. Wu, J. Neidhoefer, S. Kannan, M. Turbe, Flight-Test Results of Autonomous
Airplane Transitions Between Steady-Level and Hovering Flight, Journal of Guidance Control
and Dynamics 31 (2) (2008) 358370.
[22] A. Frank, J. McGrew, M. Valenti, D. Levine, J. How, Hover, Transition, and Level Flight Control
Design for a Single-Propeller Indoor Airplane, in: Proceedings of the AIAA Guidance,
Navigation, and Control Conference and Exhibit, Myrtle Beach, SC, 2007.
[23] D. Poinsot, C. Berard, R. Krashanitsa, S. Shkarayev, Investigation of Flight Dynamics and
Automatic Controls for Hovering Micro Air Vehicles, in: AIAA Guidance, Navigation and
Control Conference, Honolulu, Hawaii, 2008, AIAA 2008-6332.
[24] S. Osborne, Transitions Between Hover and Level Flight for a Tailsitter UAV, Ph.D. thesis,
Brigham Young University (2007).

Volume 2 Number 2 2010

Anda mungkin juga menyukai