high angle-of-attack flight characteristics of a small uav

69
HIGH ANGLE-OF-ATTACK FLIGHT CHARACTERISTICS OF A SMALL UAV WITH A VARIABLE-SIZE VERTICAL TAIL By BARON JON JOHNSON A THESIS PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF MASTER OF SCIENCE UNIVERSITY OF FLORIDA 2009 1

Upload: others

Post on 06-Dec-2021

3 views

Category:

Documents


0 download

TRANSCRIPT

HIGH ANGLE-OF-ATTACK FLIGHT CHARACTERISTICS OF A SMALL UAV WITHA VARIABLE-SIZE VERTICAL TAIL

By

BARON JON JOHNSON

A THESIS PRESENTED TO THE GRADUATE SCHOOLOF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT

OF THE REQUIREMENTS FOR THE DEGREE OFMASTER OF SCIENCE

UNIVERSITY OF FLORIDA

2009

1

c© 2009 Baron Jon Johnson

2

I dedicate this to my loving parents, who have fully supported my every endeavor, and to

Michael Morton, a dear friend, roommate, and colleague who was taken far too early.

You will always be missed Morty.

3

ACKNOWLEDGMENTS

I would like to extend thanks to a number of supportive and inspiring people with

whom I’ve had the honor, privilege, and pleasure to work with during my time at the

University of Florida. First of all, thanks to my advisor, Dr. Rick Lind, for providing me

the opportunity to study under him during graduate school in the Flight Control Lab.

His guidance, suggestions, and critiques have proven invaluable. Thanks to Dr. Peter

Ifju for offering me the most exciting and rewarding job I could have ever imagined as an

undergraduate in the Micro Air Vehicle Lab.

A number of fellow students have also supported me in my time at UF, either directly

or indirectly. Scott Bowman has always provided a great deal of support and made

great contributions to my understanding of electronic systems, particularly in the field of

shooting bottle rockets from RC planes. Assistance, inspiration, and constant amusement

has been provided by the current members of the Flight Control Lab, listed in a very

particular order: Dong Tran, Sanketh Bhat, Daniel ’Tex’ Grant, Robert Love, Brian

Roberts, and Ryan Hurley. Other collaborators which have made strong impressions upon

me over the years include Mujahid Abdulrahim, Carlo Francis, Frank Boria, Dan Claxton,

Kyuho Lee, Adam Watts, Jos Cocquyt, and Patricia Miller.

4

TABLE OF CONTENTS

page

ACKNOWLEDGMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

ABSTRACT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

CHAPTER

1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

1.1 High Angle-of-Attack Flight . . . . . . . . . . . . . . . . . . . . . . . . . . 121.2 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131.3 Previous Research . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131.4 Contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

2 EXPERIMENTAL SETUP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

2.1 Aircraft . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172.2 Vertical Tail . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182.3 Avionics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

3 FLIGHT TESTING . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

4 ANALYSIS TECHNIQUES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

4.1 Fourier Transform . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 284.2 Wavelet Transform . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 284.3 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

4.3.1 Example 1: Stationary Sine Wave Signal . . . . . . . . . . . . . . . 304.3.2 Example 2: Stationary Signal with Multiple Frequency Components 314.3.3 Example 3: Chirp Signal with Increasing Frequency . . . . . . . . . 33

5 SYSTEM IDENTIFICATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

5.1 Procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 365.2 Doublet Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

5.2.1 Longitudinal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 385.2.2 Lateral . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 405.2.3 Directional . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

5.3 Steady-State Flight . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 435.3.1 Longitudinal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 435.3.2 Lateral . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 455.3.3 Directional . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

5

6 VERTICAL TAIL RELATIONSHIP TO WING ROCK . . . . . . . . . . . . . . 50

6.1 Upright Tail . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 506.2 Inverted Tail . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 526.3 Parameterization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

7 SUMMARY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

7.1 Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 627.2 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

BIOGRAPHICAL SKETCH . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

6

LIST OF TABLES

Table page

2-1 Mini ShowTime Specifications . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

2-2 Mini ShowTime Components . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

2-3 Vertical Tail Specifications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

2-4 Size and Mass of Avionics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

2-5 IMU/FDR Wiring Sequence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

2-6 Technical Specifications of IMU Sensors . . . . . . . . . . . . . . . . . . . . . . 23

2-7 IMU Raw Output Multipliers and Resulting Units . . . . . . . . . . . . . . . . . 24

7

LIST OF FIGURES

Figure page

1-1 Graphical Representation of Angle-of-Attack . . . . . . . . . . . . . . . . . . . . 12

1-2 F/A-18 HARV and X-29 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

1-3 RQ-11B Raven Small Unmanned Aircraft System . . . . . . . . . . . . . . . . . 15

2-1 Mini ShowTime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

2-2 Vertical Tails . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

2-3 Interchangeable Vertical Tail Mounted on Fuselage . . . . . . . . . . . . . . . . 21

2-4 Airborne Sensors: IMU, GPS Receiver, and Flight Data Recorder . . . . . . . . 21

2-5 Avionics Mounted Under Canopy . . . . . . . . . . . . . . . . . . . . . . . . . . 22

2-6 Wiring Diagram of Mini ShowTime Aircraft for Data Collection . . . . . . . . . 23

3-1 Mini ShowTime in Representative High Angle-of-Attack Flight . . . . . . . . . . 26

3-2 Mini ShowTime in Inverted High Angle-of-Attack Flight . . . . . . . . . . . . . 27

4-1 Representative Wavelet Window Size Pattern . . . . . . . . . . . . . . . . . . . 29

4-2 Representative Morlet Mother Wavelet . . . . . . . . . . . . . . . . . . . . . . . 30

4-3 Signal for Example 1: 10 rad/s Sine Wave . . . . . . . . . . . . . . . . . . . . . 30

4-4 Frequency Analysis of Example 1 Signal . . . . . . . . . . . . . . . . . . . . . . 31

4-5 3-Dimensional Depiction of Example 1 Wavelet Transform . . . . . . . . . . . . 32

4-6 Signal for Example 2: Summation of 5 and 10 rad/s Sine Waves . . . . . . . . . 32

4-7 Frequency Analysis of Example 2 Signal . . . . . . . . . . . . . . . . . . . . . . 33

4-8 Signal for Example 3: Sinusoidal Chirp with Increasing Frequency . . . . . . . . 34

4-9 Chirp Frequency With Respect To Time . . . . . . . . . . . . . . . . . . . . . . 34

4-10 Frequency Analysis of Example 3 Signal . . . . . . . . . . . . . . . . . . . . . . 35

5-1 Elevator and Pitch Rate During Doublets . . . . . . . . . . . . . . . . . . . . . 38

5-2 Individual Contributions to Response from Longitudinal Model . . . . . . . . . 39

5-3 Aileron and Roll Rate During Doublets . . . . . . . . . . . . . . . . . . . . . . . 40

5-4 Individual Contributions to Response from Lateral Model . . . . . . . . . . . . . 41

8

5-5 Rudder and Yaw Rate During Doublets . . . . . . . . . . . . . . . . . . . . . . . 42

5-6 Individual Contributions to Response from Directional Model . . . . . . . . . . 43

5-7 Elevator and Pitch Rate During Steady Flight . . . . . . . . . . . . . . . . . . . 44

5-8 Uncommanded Pitch Rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

5-9 FFT of Measured Pitch Rate, Uncommanded Pitch Rate, and Elevator Input . . 45

5-10 Aileron and Roll Rate During Steady Flight . . . . . . . . . . . . . . . . . . . . 46

5-11 Uncommanded Roll Rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

5-12 FFT of Measured Roll Rate, Uncommanded Roll Rate, and Aileron Input . . . 47

5-13 Rudder and Yaw Rate During Steady Flight . . . . . . . . . . . . . . . . . . . . 47

5-14 Uncommanded Yaw Rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

5-15 FFT of Measured Yaw Rate, Uncommanded Yaw Rate, and Rudder Input . . . 49

6-1 Time Responses in Upright Configuration . . . . . . . . . . . . . . . . . . . . . 51

6-2 FFT in Upright Configuration of Roll Rate and Aileron Deflection . . . . . . . . 52

6-3 Wavelet Transforms of Roll Rate in Upright Configuration . . . . . . . . . . . . 53

6-4 Wavelet Transforms of Aileron Deflection in Upright Configuration . . . . . . . 54

6-5 Time Responses in Inverted Configuration . . . . . . . . . . . . . . . . . . . . . 55

6-6 FFT in Inverted Configuration of Roll Rate and Aileron Deflection . . . . . . . 56

6-7 Wavelet Transforms of Roll Rate in Inverted Configuration . . . . . . . . . . . . 57

6-8 Wavelet Transforms of Aileron Deflection in Inverted Configuration . . . . . . . 58

6-9 Mean Roll Rate Peak Magnitudes and Peak-to-Peak Frequencies . . . . . . . . . 59

6-10 Upper and Lower Bounds of Wing Rock Frequency from FFTs and Wavelets . . 60

6-11 Mean Wing Rock Bandwidths . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

9

Abstract of Thesis Presented to the Graduate Schoolof the University of Florida in Partial Fulfillment of the

Requirements for the Degree of Master of Science

HIGH ANGLE-OF-ATTACK FLIGHT CHARACTERISTICS OF A SMALL UAV WITHA VARIABLE-SIZE VERTICAL TAIL

By

Baron Jon Johnson

May 2009

Chair: Richard C. Lind, Jr.Major: Aerospace Engineering

The angle-of-attack parameter has a significant influence on the aerodynamics, flight

dynamics, and handling qualities of aircraft. Flight at high angle-of-attack conditions

enables many missions; however, the flight dynamics are challenging to model and

are largely influenced by uncommanded and sometimes unpredictable motions. This

study investigates the flight dynamics of a small UAV that is piloted in open air at high

angle-of-attack conditions well beyond wing stall. Models are estimated from the flight

data to indicate some characteristics of the flight dynamics. The lateral dynamics are

linear and dominated by a traditional mode of roll convergence, while the longitudinal

and directional dynamics exhibit nonlinearities and require high-order terms. The models,

which are based upon responses to doublet perturbations, are used to predict steady-state

high angle-of-attack flight with significantly smaller control inputs. Uncommanded

oscillations are identified as the motions not predicted by the doublet-based models.

Uncommanded oscillations about all three axes are observed with the most notable being

about the roll axis, commonly called wing rock. This wing rock behavior of a small

UAV with different vertical tail sizes and configurations is studied using time, frequency,

and time-frequency analysis techniques. Wing rock is found to be a fairly narrow-band

phenomenon, but with frequency variations in time. The wing rock behavior is found to

be independent of vertical tail size but largely dependent on the vertical tail configuration;

10

specifically, wing rock is pronounced with any size upright vertical tail but virtually

non-existent with any size inverted vertical tail.

11

CHAPTER 1INTRODUCTION

1.1 High Angle-of-Attack Flight

Angle-of-attack is defined as the angle between the chord line or datum line of an

aircraft and the vector representing direction of motion, as shown in Figure 1-1, where

angle-of-attack is represented by α, pitch angle is represented by θ, and the flight path

inclination is represented by γ.

Figure 1-1. Graphical Representation of Angle-of-Attack

Angle-of-attack can be determined as a function of body-frame velocities, as indicated

in Equation 1–1, where vbx and vb

z are body-frame velocities in the x and z directions

respectively.

α = tan−1

(vb

z

vbx

)(1–1)

Many aircraft experience wing stall at angles-of-attack between 10 and 15◦ without

the use of high-lift devices or flow control mechanisms. The angle-of-attack at which stall

12

occurs, the critical angle-of-attack, provides the maximum lift coefficient available for the

wing. The transition through the critical angle-of-attack into the stalled flight regime

is characterized by substantial flow separation along the lifting surfaces and a loss of

the linear relationship between angle-of-attack and lift coefficient. Such stall conditions

typically limit an aircraft’s flight envelope, as flight beyond stall can be very difficult to

control and can result in dangerous spin autorotations. This study examines flight at

angles-of-attack well beyond wing stall, but with forward airspeed. Hovering, in which

the aircraft is pitched to a vertical attitude and maintains zero airspeed (at which point

angle-of-attack is ill-defined), is not within the realm of this study.

1.2 Motivation

The issue of airspeed resulting from high angle-of-attack conditions is a primary

benefit; specifically, the airspeed can be significantly reduced by operating at high

angle-of-attack conditions. Urban operations are especially impacted by airspeed. Such

low airspeed can enable sensing missions that want exposure to targets which are in close

proximity and thus quickly leave the field of view. The low airspeed coupled with a steep

descent can enable maneuvers such as perching, sensor placement, or landing in confined

areas without sacrificing ingress and egress speed (1).

High angle-of-attack flight could also provide greater obstacle avoidance capabilities

when flying in urban terrain since the turning radius in high angle-of-attack flight is

greatly reduced as compared to that of conventional forward flight. Sensing capabilities

could also be enhanced through the use of sensor pointing without the complexity and

weight of morphing aircraft or gimbaled sensors (2).

1.3 Previous Research

A variety of studies have considered the flight dynamics at high angle-of-attack

conditions. A modified F/A-18, called the High Alpha Research Vehicle (HARV), initiated

the study of thrust vectoring to achieve extreme conditions (3; 4). A novel configuration

for the X-29 used forward-swept wings and a canard to demonstrate flight at these

13

conditions (5). In each case, the flight dynamics were found to be challenging to model

because of the influence of aerodynamics. Additional issues, such as wing rock, were also

noted in varying levels at virtually all angle-of-attack conditions in the post-stall regime.

The F/A-18 HARV and the X-29 can be seen in Figure 1-2 (6; 7).

Figure 1-2. A)F/A-18 HARV [http://www.dfrc.nasa.gov/gallery/Photo/F-18HARV/,reprinted with permission] and B) X-29[http://www.dfrc.nasa.gov/gallery/Photo/X-29/, reprinted with permission]

The flight dynamics associated with small UAVs are receiving significant attention in

the community as a result of their mission potential. Several small, man-portable vehicles

feature a deep-stall short landing mode in which the elevator deflects and the power is

reduced while the aircraft steeply descends in a near-level attitude until it impacts the

ground (8; 9). The RQ-11B Raven, shown in Figure 1-3 (10), is an example of such a UAV

that features a deep-stall landing mode. The flight dynamics of small UAVs have also been

studied in hovering modes to enable autonomous control (11–13).

Small UAVs and remote-controlled (RC) aircraft are rapidly maturing in flight

capability for a variety of missions including urban operations. As such, the ability to

operate at high angle-of-attack conditions is a critical requirement for these platforms.

Such capability is usually achieved thanks to the combination of thrust generation

and control surfaces. These aircraft often utilize front-mounted propellers and tractor

propulsion that produces airflow over the wings and tail. The large control surfaces, which

14

Figure 1-3. RQ-11B Raven Small Unmanned Aircraft System[http://www.avinc.com/media gallery2.asp?id=224, reprinted with permission]

constitute 50% or more of the tail area and may deflect 45◦ or more, are thus able to use

this propwash to maintain control authority.

Flight at high angle-of-attack conditions is often characterized by wing rock. This

phenomenon is described as uncommanded self-induced oscillations primarily about the

roll axis. Some research has indicated that, although dominated by roll motions, the

uncommanded wing rock may actually be a lightly-damped dutch roll motion (14–16).

The source of the uncommanded wing rock is not completely known and seems to

vary by aircraft configuration. Some research has led to the belief that the wing rock

phenomenon is a limit cycle oscillation (LCO) caused either by the loss of dynamic roll

damping at high angles-of-attack (17; 18) or an aerodynamic hysteresis which generates

the spring-like forces required to drive the LCO (19–21).

The presence of sideslip has been shown to have an effect on wing rock at high

angles-of-attack, both as a cause and mitigator (20; 22; 23). Additionally, quite a

few studies have found the uncommanded wing rock phenomenon to be somewhat

unpredictable in nature, both in magnitude and periodicity (22; 24–27).

Some research has determined the phenomenon to be caused by vortices from the

leading edges of the wings (14; 28; 29), while other research has determined it to be

caused largely or entirely by vortices generated from slender forebodies impinging upon

15

downstream airframe components such as the vertical tail (26; 28; 30–34). This has led to

research on the wing rock relationship with and without a vertical tail surface (35).

1.4 Contributions

This paper investigates the flight dynamics of a small UAV during piloted flight at

high angle-of-attack. An avionics suite records 3-axis measurements along with servo

commands during steady-state trim flight and in response to doublet maneuvers for

identification purposes. The resulting data is used to estimate models of the flight

dynamics. These models indicate the lateral dynamics are relatively standard and

dominated by a roll convergence; however, the longitudinal dynamics and directional

dynamics require nonlinear terms and lag delays. Uncommanded behavior in steady

high angle-of-attack flight is extracted from the flight data through the use of the

doublet-generated models.

This study also investigates the effect of tail size and configuration on wing rock at

high angles-of-attack using time, frequency, and time-frequency analysis techniques. This

makes several contributions to the community that supplements previous research. First,

it is determined through time-frequency analysis that wing rock is a narrow frequency

band phenomenon, but with frequency variations in time. Second, wing rock magnitude

and frequency do not depend on the size of the vertical tail. Third, wing rock is essentially

eliminated when configured with a lower vertical tail of any size. This could be caused by

either the presence of the lower vertical tail or the absence of the upper vertical tail.

16

CHAPTER 2EXPERIMENTAL SETUP

2.1 Aircraft

The flight characteristics of the Mini ShowTime are studied in a high angle-of-attack

condition. This electric-powered aircraft, shown in Figure 2-1, is a commercially-available

off-the-shelf platform that is commonly used by the remote-control (RC) community.

Figure 2-1. Mini ShowTime

This aircraft is constructed from a lightweight balsa wood structure that allows the

wingspan to be large in comparison with the vehicle weight. The specific platform has a

weight of approximately 820 g along with the characteristics given in Table 2-1 (36). The

aircraft was outfitted with conventional RC components for control and propulsion, which

are listed in Table 2-2

This aircraft is used in the community for aerobatics because of its excellent agility

and outstanding characteristics at high angle-of-attack conditions. In particular, the

aircraft is highly controllable at high angle-of-attack conditions as a result of low

17

Table 2-1. Mini ShowTime Specifications

Parameter ValueWingspan 1090 mmLength 1065 mm

Wing Area 26.7 dm2

Flying Weight 820-850 gWing Loading 30.7-31.8 g/dm2

Flight Speed 0-20 m/s

Table 2-2. Mini ShowTime Components

Component Manufacturer ModelTransmitter Spektrum DX7

Receiver Spektrum AR6100eServos (4) JR DS285

Battery Eliminator Circuit (BEC) Castle Creations CC BECElectronic Speed Control (ESC) E-flite 40-Amp Brushless (V2)

Motor E-flite Park 480 BL Outrunner, 1020 KvPropeller APC 12x6EBattery Thunder Power 3S, 11.1 V, 2100 mAh

wing-loading and high thrust-to-weight ratio along with large control surfaces. These

characteristics enable a pilot to maintain orientations within reasonable limits for extended

durations.

The Mini ShowTime aircraft has fewer asymmetric and more symmetric features

than many aircraft. The most notable asymmetry is the vertical tail, which is above

the fuselage. Other minor asymmetries include the shape of the canopy and the landing

gear. There are several significant symmetric features of the aircraft: all flight surfaces

have symmetric airfoils, the wing and horizontal tail are located essentially along the

centerline of the aircraft, and all control surfaces have equivalent deflections in both

directions. These characteristics permit the comparison between upright and inverted

flight characteristics with essentially a single asymmetrical influence: the vertical tail.

2.2 Vertical Tail

The Mini ShowTime aircraft is modified to accomodate interchangeable vertical

tails. Vertical tails of varying height are constructed, including the stock tail size and

two taller and two shorter vertical tails. The weight difference between vertical tails is

18

negligible in terms of overall weight and CG location. Vertical tail volume coefficient is a

nondimensional measure of the directional stability of a particular airframe configuration

and CG location. A vertical tail with larger area or located farther behind the CG will

provide a larger restoring moment due to a sideslip and thus provide a larger vertical tail

volume coefficient for a given wing geometry. The vertical tail volume coefficient, Vv, is

calculated with Equation 2–1, where Sv represents vertical tail area, Lv represents the

distance from the CG of the aircraft to the aerodynamic center of the vertical tail, Sw

represents wing area, and b represents wing span (37).

Vv =

(Sv

Sw

) (Lv

b

)(2–1)

The height, area, and vertical tail volume coefficient of each tail are presented in

Table 2-3. The range of tail sizes provides areas and vertical tail volume coefficients

ranging from approximately 50% to 150% of the stock tail size. The vertical tails are

shown in Figure 2-2, and a tail mounted on the modified fuselage is shown in Figure 2-3.

Table 2-3. Vertical Tail Specifications

Tail Height (mm) Area (cm2) Vv

Smaller 64.86 90.67 0.01846Small 90.90 125.22 0.02549

Normal 115.29 162.10 0.03303Big 140.51 201.70 0.04107

Bigger 165.91 241.73 0.04920

2.3 Avionics

An avionics suite is integrated into the aircraft. This suite augments the baseline

configuration with sensors and actuators that are particularly appropriate for high

angle-of-attack testing.

Data describing the flight is obtained using the set of instruments shown in

Figure 2-4. A pair of sensor packages result from an inertial measurement unit (IMU)

and global positioning system (GPS). The IMU, which is a MEMSense nIMU, is a

MEMS-based unit with temperature compensation and digital I2C output of 3-axis

19

Figure 2-2. Vertical Tails: Drawing (top) and Actual (bottom)

accelerations, angular rates, and magnetic flux. The GPS, which is an Eagle Tree

Expander Module, notes location, groundspeed, course, and UTC timestamp at a rate

of 5 Hz. An additional flight data recorder (FDR), which is the Eagle Tree Systems FDR

Pro, logs the sensor outputs along with barometric altitude and servo commands. This

system is able to obtain more than 15 minutes of data at a rate of 25 Hz.

These sensors are relatively small, as noted in Table 2-4 (38; 39), and in each case the

weight is nearly negligible on the flight dynamics.

Table 2-4. Size and Mass of Avionics

Unit Size (mm) Mass (g)IMU 46.5 x 22.9 x 13.9 20GPS 36.0 x 43.0 x 13.0 23FDR 50.0 x 35.0 x 17.0 22

20

Figure 2-3. Interchangeable Vertical Tail Mounted on Fuselage

Figure 2-4. Airborne Sensors: IMU (left), GPS Receiver (middle), and Flight DataRecorder (right)

21

The IMU is mounted on a specially-installed shelf within the fuselage to lie very

close to the center of gravity along all three axes. The GPS and FDR, along with the

conventional RC receiver and battery eliminator circuit (BEC), are installed under the

canopy. The mounted location of each can be seen in Figure 2-5.

Figure 2-5. Avionics Mounted Under Canopy: A) FDR, B) GPS, C) RC Receiver (UnderGPS), D) IMU, E) BEC

An adapter was assembled to connect the IMU, which uses a Hirose HR30 6-pin

connector, to the FDR, which uses a 4-wire plug. The connection sequence is shown

in Table 2-5 (38). The complete wiring diagram of the experimental aircraft, including

conventional RC components, is shown in Figure 2-6. The BEC must be programmed to

output a voltage in the range of 5.4 - 7.0 V to properly power the IMU, FDR, and RC

systems simultaneously (38; 39).

Table 2-5. IMU/FDR Wiring Sequence

Hirose Port No. I2C Function FDR Wire1 SDA Yellow2 VDD Red3 Not Used NA4 Not Used NA5 GND White6 SCL Brown

22

Figure 2-6. Wiring Diagram of Mini ShowTime Aircraft for Data Collection

These avionics have been demonstrated as highly accurate when compared to

high-quality avionics. When compared to a high-quality IMU with laser-ring gyros, the

nIMU provided measurements that yielded velocity estimates within standard deviations

of approximately 0.2 m/s on all axes and attitude estimates within standard deviations

of approximately 0.2◦ in roll and pitch and 0.35◦ in heading (40). The manufacturer

specifications are given in Table 2-6 for each of the sensors (38).

Table 2-6. Technical Specifications of IMU Sensors

Sensor Dynamic Range Digital Sensitivity Offset/Drift NoiseGyro ±600 o/s 0.01831 o/s ±1 o/s 0.56 o/s

Accelerometer ±5 g 1.5259e− 4 g ±30 mg 4.87 mgMagnetometer ±1.9 Gauss 5.79e− 5 Gauss 2700 ppm/oC 5.6e− 4 Gauss

23

The data output from the nIMU and recorded by the FDR is in the form of signed

16-bit integers. The gyro, accelerometer, and magnetometer are converted to appropriate

units through the use of Equation 2–2 (38).

TrueV alue = RawV alue

(DynamicRange

21.5

32768

)(2–2)

Based on the dynamic range of each sensor as presented in Table 2-6, Equation 2–2

results in multipliers of the raw sensor output and resulting units as presented in

Table 2-7.

Table 2-7. IMU Raw Output Multipliers and Resulting Units

Sensor Multiplier Resulting UnitsGyro 2.747e-2 deg/s

Accelerometer 2.289e-4 GMagnetometer 8.698e-5 Gauss

24

CHAPTER 3FLIGHT TESTING

Open-loop flight testing was performed to collect data during high angle-of-attack

flight, both in a steady, straight and level trim condition and while executing doublet

maneuvers. Only straight passes were used for data collection, as turning flight often

required throttle and angle-of-attack changes. The passes for data collection ranged over

approximately 100 m in length. This length was constrained due to the visual acuity of

the pilot and the ability to perform small corrections to maintain the high angle-of-attack

flight condition. All flights were performed in a clear field with no obstructions to flight

and no turbulence-creating terrain. All flights were performed on days with little to no

wind at surface level. Any flight maneuvers that were visibly affected by an external

disturbance or pilot error were repeated. The altitude for the data collection passes was

approximately 30 m. Each flight was limited to no more than 10 minutes in length due

to both the onboard battery supply and the available memory of the FDR. The pilot

attempted to perform all maneuvers at an angle-of-attack of approximately 45◦ while

maintaining level horizontal flight with as few control inputs as possible to maintain such

a flight condition. Straight and level passes provided the convenience that angle-of-attack

was approximately equivalent to pitch angle. Variations in both vertical velocity and pitch

angle caused angle-of-attack to vary an estimated ±10◦. The Mini ShowTime in high

angle-of-attack flight is shown in Figure 3-1.

A series of flights was performed to collect data during doublet maneuvers and steady

high angle-of-attack flight with the stock (normal) vertical tail. Each flight began by

taking off and establishing the aircraft in high angle-of-attack flight. A complete pass

at high angle-of-attack was first performed to establish the trim position for all control

surfaces. Each subsequent pass then included three distinct actions: establishing the high

angle-of-attack trim condition at the desired heading, performing a doublet maneuver

with a control surface, and reestablishing the trim condition for the remainder of the

25

Figure 3-1. Mini ShowTime in Representative High Angle-of-Attack Flight

pass length. Each flight involved doublets of varying size by a single control surface. The

largest doublet for each control was constrained either by maximum control deflections or

by the pilot’s ability to quickly reestablish straight and level trim flight. A doublet was

performed with maximum deflection and several were performed with progressively smaller

deflections. A minimum of three passes with doublet maneuvers were performed for each

flight.

A series of flights was then performed to collect data during steady high angle-of-attack

flight with ten different vertical tail configurations: upright and inverted with each of the

five vertical tails. Inverted high angle-of-attack was performed in a similar fashion

to upright but with down (positive) elevator deflection to maintain the pitch angle.

Additionally, slight differences in rudder trim existed between upright and inverted due to

the large yaw moment from p-factor in high angle-of-attack flight. The Mini ShowTime in

inverted high angle-of-attack flight is shown in Figure 3-2.

Each flight consisted of establishing the aircraft in either upright or inverted high

angle-of-attack flight to determine approximate trim conditions. A minimum of five

straight, horizontal passes in high angle-of-attack flight were performed with each tail

configuration.

26

Figure 3-2. Mini ShowTime in Inverted High Angle-of-Attack Flight

An assistant with a stopwatch recorded the times at which each high angle-of-attack

pass began and ended, as well as when doublets were performed. After each flight, the

data from the FDR was downloaded to a computer with the USB cable and the FDR’s

buffer was cleared. The trim conditions for the aircraft remained consistent from flight to

flight. The flight testing procedure involved 14 flights which spanned four days.

27

CHAPTER 4ANALYSIS TECHNIQUES

4.1 Fourier Transform

A common method of signal analysis is the Fourier transform, which transforms a

signal from the time-domain to the frequency-domain. Fourier transforms accomplish this

by decomposing the signal into multiple infinite-length sine waves of various frequencies.

The correlation between the signal and a sine wave of a particular frequency represents

the energy of the signal at that particular frequency across the entire time span of the

signal. Fourier transforms are often represented graphically by plotting the relative

magnitude of the energy at each frequency across an appropriate frequency range. Strong

correlations will be represented by peaks in the magnitude at a particular frequency.

Fourier transforms are particularly useful for analyzing a stationary signal, or one whose

periodic behavior varies little over time, because the signal is compared against stationary

signals (41).

The Fourier transform of a discrete signal, such as that sampled from a continuous

signal, is called a discrete Fourier transform (DFT). The most common method with

which a DFT is applied is through a fast Fourier transform (FFT), which is a class of

efficient algorithms that compute the DFT (41).

4.2 Wavelet Transform

The wavelet transform is another signal analysis method which can provide frequency

information along with time-domain information, which makes it ideally suited to analyze

signals which are not stationary and may contain segments of periodicity, abrupt changes,

trends, and other momentary or local characteristics (41).

The wavelet transform is a windowing technique which utilizes long time intervals

to extract low frequency information and short time intervals to extract high frequency

information. The time-frequency plane becomes partitioned into windows of constant area,

as shown in Figure 4-1 (42), which results in a multiresolution analysis.

28

Figure 4-1. Representative Wavelet Window Size Pattern

Within each window the signal is compared to a mother, or original, wavelet of a

given scale and shift to determine a level of correlation. A wavelet is a limited duration

waveform which begins at, ends at, and has an average value of zero. The scale alters

the frequency of the wavelet by stretching or compressing the mother wavelet along the

time axis. The shift alters the location of the wavelet along the time axis. There are

many different types of mother wavelets which can be applied, but the Morlet wavelet is

the most common. The Morlet wavelet is a sine wave which is localized by a Gaussian

envelope and can be seen in Figure 4-2 (41).

The wavelet transform determines the correlation of a signal to the mother wavelet of

various scales and displays the correlation on a time-scale plot. The dominant frequency

of a wavelet can be approximated from the scale. A wavelet with a large scale is stretched

in time, which allows it to correlate with low frequency signals, and a wavelet with a small

scale is compressed in time, which allows it to correlate with high frequency signals (41).

The resulting wavelet transformation can be represented either as a 3-dimensional

plot, with axes representing time, frequency, and correlation magnitude, or as a 2-dimensional

plot with time and frequency axes and intensity representing correlation magnitude. High

correlation at a particular frequency appears on the wavelet transformation plot as

29

−4 −2 0 2 4−1

−0.5

0

0.5

1

Time (s)

Mag

nitu

de

Figure 4-2. Representative Morlet Mother Wavelet

alternating positive and negative correlation magnitudes due to phase shifts between the

mother wavelet and periodic signal.

4.3 Examples

4.3.1 Example 1: Stationary Sine Wave Signal

The first example demonstrates the Fourier and wavelet transforms on a stationary

sinusoidal wave with a frequency of 10 rad/s. The signal is shown in Figure 4-3.

0 2 4 6 8 10−1.5

−1

−0.5

0

0.5

1

1.5

Time (s)

Mag

nitu

de

Figure 4-3. Signal for Example 1: 10 rad/s Sine Wave

30

The Fourier transform is applied to the signal with an FFT algorithm. The

resulting frequency-domain representation displays a peak in magnitude at 10 rad/s.

The wavelet transform is also applied to the signal using a Morlet wavelet, and the

resulting representation displays a band of strong correlation centered at 10 rad/s along

the entire time period. The FFT and wavelet transform representations of the example

1 signal are shown in Figure 4-4. In this case the FFT and wavelet representations both

provide similar information about the signal, as it is stationary.

100

101

102

103

10−15

10−10

10−5

100

105

Frequency (rad/s)

Mag

nitu

de

Figure 4-4. Frequency Analysis of Example 1 Signal: FFT (left) and Wavelet (right)

The wavelet transform in Figure 4-4 can be viewed as a 3-dimensional plot in

Figure 4-5. The correlation magnitude is depicted along the z-direction, along with the

same time and frequency axes. It can be seen that along 10 rad/s, the wavelet transform

indicates alternating strong positive and negative peaks. This is due to the phase shift; at

times the scaled mother wavelet matches the signal well resulting in positive correlation,

and at other times it matches it well in an inverse relationship resulting in negative

correlation.

4.3.2 Example 2: Stationary Signal with Multiple Frequency Components

The second example demonstrates the Fourier and wavelet transforms on a signal

which is composed of multiple stationary signals of different frequencies. The signal is the

31

Figure 4-5. 3-Dimensional Depiction of Example 1 Wavelet Transform

summation of the 10 rad/s sine wave used in example 1 along with a second stationary

sine wave with a frequency of 5 rad/s. The signal for example 2 is shown in Figure 4-6.

0 2 4 6 8 10−1.5

−1

−0.5

0

0.5

1

1.5

Time (s)

Mag

nitu

de

Figure 4-6. Signal for Example 2: Summation of 5 and 10 rad/s Sine Waves

The Fourier transform is applied to the signal with an FFT algorithm, and the

resulting frequency-domain representation displays peaks at both 5 and 10 rad/s. The

wavelet transform is also applied to the signal using a Morlet wavelet, and the resulting

32

representation displays bands of strong correlation centered on both 5 and 10 rad/s along

the entire time period. The FFT and wavelet transform representations of the example

2 signal are shown in Figure 4-7. In this case, the FFT and wavelet representations both

provide similar frequency information about the signal, as it is again stationary. Both

techniques identified both dominant frequencies contained in the signal.

100

101

102

103

10−15

10−10

10−5

100

105

Frequency (rad/s)

Mag

nitu

de

Figure 4-7. Frequency Analysis of Example 2 Signal: FFT (left) and Wavelet (right)

4.3.3 Example 3: Chirp Signal with Increasing Frequency

The third example demonstrates the Fourier and wavelet transforms on signal

without consistent frequency components. The signal is a sinusoidal chirp with increasing

frequency, as shown in Figure 4-8.

The chirp signal shown in Figure 4-8 clearly increases frequency with time. The

frequency migration with respect to time is shown in Figure 4-9.

The Fourier transform is applied to the signal with an FFT algorithm, and the

resulting frequency-domain representation displays similar energy across all frequencies

below approximately 10 rad/s. The wavelet transform is also applied to the signal using

a Morlet wavelet. The resulting representation displays bands of strong correlation which

initially are at low frequency and increase with time until reaching approximately 10 rad/s

at the end of the sample. The FFT and wavelet transform representations of the example

3 signal are shown in Figure 4-10.

33

0 2 4 6 8 10−1.5

−1

−0.5

0

0.5

1

1.5

Time (s)

Mag

nitu

de

Figure 4-8. Signal for Example 3: Sinusoidal Chirp with Increasing Frequency

0 2 4 6 8 100

5

10

15

20

Freq

uenc

y (r

ad/s

)

Time (s)

Figure 4-9. Chirp Frequency With Respect To Time

In this case, the difference between the information that Fourier and wavelet

transforms provide is easily identified. The FFT identifies that there is strong correlation

across a range of frequencies, but gives no indication as to whether the correlations exist

weakly across all times or strongly over short time segments. The wavelet transform, on

the other hand, maintains the time-domain information. It is clearly seen that at any

particular time instance there is a strong correlation at a particular frequency and not

a wide range of frequencies. The overall range of frequencies identified by the wavelet

transform corresponds to the range of frequencies identified by the FFT, but the wavelet

34

100

101

102

103

10−20

10−10

100

1010

Frequency (rad/s)

Mag

nitu

de

Figure 4-10. Frequency Analysis of Example 3 Signal: FFT (left) and Wavelet (right)

transform also provides frequency information localized in time. It can be seen that the

centers of the correlation peaks of the wavelet in Figure 4-10 follow the trend of the chirp

frequency as shown in Figure 4-9.

35

CHAPTER 5SYSTEM IDENTIFICATION

5.1 Procedure

Least-squares regression is a common and effective method of fitting data. It is

particularly useful for modeling an overdetermined system with measurements that contain

random errors. Residuals are the differences between the measured and estimated values

at each data point, and the least-squares method determines the model which minimizes

the sum of the squared residuals. The least-squares equations can be represented with the

single matrix equation in Equation 5–1, where A is an n x m matrix of regressors, X is an

m x 1 vector of unknown coefficients, L is an n x 1 vector of measurements, and V is an

n x 1 vector of residuals, where n is the number of samples and m is the number of states

in the least-squares model (43).

AX = L + V (5–1)

In the specific case where n = m and A is a square matrix, the simple inverse of A

could be used to solve for the unknown coefficients. However, to guarantee an invertible

matrix for the general solution, the pseudoinverse of A is applied. The general solution of

the unknown coefficients is then calculated with Equation 5–2 (43).

X = (AT A)−1AT L (5–2)

Least-squares models are created for roll, pitch, and yaw rates, in which the regressors

are based on direct measurements of angular rates and control input commands. The

complete models also include nonlinear quadratic, cubic, and quartic terms in both

angular rates and control inputs. The models are based solely on direct measurements

from the onboard avionics, as a post-processing INS solution was not available at the time

of this research. Knowledge of aircraft attitude and angles-of-attack and sideslip could be

incorporated into system identification models when they are available.

36

Each regressor in the least-squares models is an average of three data points in order

to provide low pass filtering within the model. A lag of approximately 160 ms is added

into the control input regressor terms to compensate for the speed of the servos. This

lag is introduced by making the control input regressor terms functions of control inputs

from several prior sampling periods and not the most recent control inputs. A constant

regressor is included in each model to account for bias in the data. Each control surface

position is measured during the trimmed high angle-of-attack flight before and after

each doublet maneuver. These trim positions are not necessarily the trim positions in

the normal flight regime, particularly for elevator. The control input data is shifted such

that the trim positions for each control surface correspond to zero in order to remove any

unintended effect due to the large deflections at trim. All control inputs are in units of

percent of maximum deflection and angular rates are in units of deg/s. Other units can

be used with this system identification method, but would result in different regressor

coefficients.

High angle-of-attack flight is especially vulnerable to uncommanded motions, either

momentary disturbances or underdamped oscillatory motion. The models are fit to data

obtained during doublet maneuvers, and only short periods of time before and after each

doublet are included in order to minimize the influence of uncommanded motions. The

data from several doublets on each axis was concatenated such that each model is based

on several doublet maneuvers to reduce the impact of remaining uncommanded motions

and create a more robust model.

A reduced model can be created by removing one or more regressors from the

complete model. Removing regressors whose associated numerical values are small or have

large errors will result in higher accuracy for the estimates of the retained regressors (43).

This is accomplished through backward elimination, in which erroneous or insignificantly

small individual components in each complete model are removed iteratively. At each

iteration the model is used to simulate the output with the same inputs as the model is

37

based upon. The contribution of each regressor term is plotted to identify insignificant

or erroneous contributions which can then be removed. If the removal of a particular

regressor term reduces the accuracy of the model in predicting the doublet responses, the

term is retained. The final reduced models are presented in the following sections.

5.2 Doublet Modeling

5.2.1 Longitudinal

A model of the longitudinal dynamics is generated to relate the pitch rate to the

elevator commands. Such a model is estimated to relate a set of doublet commands to the

elevator and the resulting pitch rate. The measurements of pitch rate and the simulated

values from the model are shown in Figure 5-1 along with the doublets.

0 2 4 6 8 10 12 14−100

−50

0

50

100

150

Time (s)

Ele

vato

r (%

)

0 2 4 6 8 10 12 14−300

−200

−100

0

100

200

Time (s)

Pitc

h R

ate

(deg

/s)

Figure 5-1. Elevator (left) and Pitch Rate (right) during Doublets: Measured (—) andSimulated (−−−)

The model with simulated response in Figure 5-1 is given in Equation 5–3 as a

discrete-time equation, where q is pitch rate in deg/s and δe is elevator deflection in

%. This model depends upon a linear combination of lagged values for elevator angle

corresponding to 5 and 7 previous time steps. The current value of pitch rate also depends

upon lagged values of pitch rate corresponding to average values from 2, 4 and 6 previous

time steps.

38

q(k) = 1.0099− 2.1338δe(k − 4) + δe(k − 5) + δe(k − 6)

3

−0.2077δe(k − 6) + δe(k − 7) + δe(k − 8)

3+ 1.1516

q(k − 1) + q(k − 2) + q(k − 3)

3

−0.3956q(k − 3) + q(k − 4) + q(k − 5)

3− 0.0215

q(k − 5) + q(k − 6) + q(k − 7)

3

(5–3)

The contributions of each term from Equation 5–3 to the response in Figure 5-1

is shown in Figure 5-2. The largest contribution resulted from a negative pitch rate at

time of k resulting from a positive elevator deflection at time of k − 5; however, some

higher-order dynamics is also present because of the need to also retain a contribution

from the elevator deflection at time of k − 7. The state dynamics are evidenced by the

contributions from several lagged values of pitch rate. A positive pitch rate at time of

k − 2 generates a positive contribution to current pitch rate while a positive pitch rate

at either time of k − 4 or k − 6 actually generates a negative contribution to pitch rate.

Such disparity is partly due to out-of-phase states from a short-period mode although the

dynamics at high angle-of-attack do not necessarily have traditional modes.

0 2 4 6 8 10 12 14−300

−200

−100

0

100

200

Time (s)

Pitc

h R

ate

(deg

/s)

Figure 5-2. Individual Contributions to Response from Longitudinal Model:Elevator(k-5) (—o), Elevator(k-7) (—), Pitch Rate(k-2) (...), PitchRate(k-4) (−−−), Pitch Rate(k-6) (—.—.)

39

The model provides reasonable accuracy in reproducing the pitch rate as shown in

Figure 5-1; however, these longitudinal dynamics are expectedly difficult to model at

high angle-of-attack conditions. Certainly the aerodynamics are not necessarily linear

nor finite-dimensional functions of flight condition. Also, the lack of angle-of-attack

measurements does not necessarily limit the fidelity, given that a transfer function

always exists between an input and an output, but the lack almost certainly limits the

interpretation of the resulting model.

5.2.2 Lateral

A set of data relating to the lateral dynamics is generated from the doublets and

resulting roll rate shown in Figure 5-3. The doublets and resulting roll rates actually

vary by roughly a factor of 2 between the first and third command so a rich set of data is

available for the model.

0 1 2 3 4 5 6 7−100

−50

0

50

100

150

Time (s)

Aile

ron

(%)

0 1 2 3 4 5 6 7−200

−150

−100

−50

0

50

100

150

200

Time (s)

Rol

l Rat

e (d

eg/s

)

Figure 5-3. Aileron (left) and Roll Rate (right) during Doublets: Measured (—) andSimulated (−−−)

The model that simulated the roll rate in response to the doublets in Figure 5-3 is

given in Equation 5–4, where δa represents aileron deflection in %. This model simply

generates the roll rate at time of k from a bias term along with affine and quadratic terms

associated with average aileron angle at times of k − 5.

40

p(k) = −0.0744 + 1.4965δa(k − 4) + δa(k − 5) + δa(k − 6)

3

−8.42e− 5δ2a(k − 4) + δ2

a(k − 5) + δ2a(k − 6)

3(5–4)

This model indicates the lateral dynamics are dominated by a traditional mode of roll

convergence. The response is nearly linear in aileron angle, as shown in Figure 5-4, since

the nonlinear contribution is negligible. Such a result is somewhat logical given that any

effects of high angle-of-attack conditions would influence the aerodynamics of longitudinal

motion much more than the lateral motion.

0 1 2 3 4 5 6 7−150

−100

−50

0

50

100

150

200

Time (s)

Rol

l Rat

e (d

eg/s

)

Figure 5-4. Individual Contributions to Response from Lateral Model: Aileron (—) andAileron2 (−−−)

5.2.3 Directional

A pair of rudder doublets and associated yaw rates are used to obtain an estimate of

the directional dynamics. These doublets along with the yaw rates, both measured and

simulated, are shown in Figure 5-5.

The model that simulated the yaw rate in response to the doublets in Figure 5-5 is

given in Equation 5–5, where δa represents aileron deflection in %, δr represents rudder

deflection in %, and p represents roll rate in deg/s. This model requires more terms to

describe the dynamics than either the longitudinal model or lateral model. Essentially, the

41

0 1 2 3 4 5−40

−30

−20

−10

0

10

20

30

Time (s)

Rud

der

(%)

0 1 2 3 4 5−40

−20

0

20

40

60

Time (s)

Yaw

Rat

e (d

eg/s

)

Figure 5-5. Rudder (left) and Yaw Rate (right) During Doublets: Measured (—) andSimulated (−−−)

directional dynamics are estimated as being a nonlinear function of both the rudder and

the aileron along with being a nonlinear function of the roll rate.

r(k) = 1.3714− 0.9916δr(k − 4) + δr(k − 5) + δr(k − 6)

3

+0.0059δ2r(k − 4) + δ2

r(k − 5) + δ2r(k − 6)

3− 0.5073

δa(k − 4) + δa(k − 5) + δa(k − 6)

3

−0.0313δ2a(k − 4) + δ2

a(k − 5) + δ2a(k − 6)

3+ 0.0259

p(k − 1) + p(k − 2) + p(k − 3)

3

−0.0009p2(k − 1) + p2(k − 2) + p2(k − 3)

3(5–5)

The contributions from each term in Equation 5–5 to the simulated response in

Figure 5-5 are shown in Figure 5-6. The response is dominated by the contributions from

the rudder with the linear term providing the significant portion.

The directional dynamics are challenging to model as evidenced by the inconsistent

quality of the fit in Figure 5-5 despite the nonlinear terms in Equation 5–5. The nature of

the linear contributions is not consistent given that a positive rudder generates negative

yaw rate and a positive aileron generates negative yaw rate as adverse yaw but a positive

roll rate generates a positive yaw rate to imply some proverse yaw (44). Even the

42

0 1 2 3 4 5−30

−20

−10

0

10

20

30

40

Time (s)

Yaw

Rat

e (d

eg/s

)

Figure 5-6. Individual Contributions to Response from Directional Model: Rudder (—o),Rudder2 (—x), Aileron (−−−), Aileron2 (—.—.), Roll Rate (...), RollRate2 (—)

nonlinearities are inconsistent since any rudder generates a small positive contribution

to yaw rate while any aileron or roll rate generates a small negative contribution to yaw

rate.

The issue of flight testing must be considered when trying to evaluate the quality of

the model and any associated inconsistencies. In particular, the influence of gusts can be

extreme on the directional dynamics when flying at high angle-of-attack conditions but of

course any gust excitation is not properly represented in the model.

5.3 Steady-State Flight

5.3.1 Longitudinal

Flight data associated with steady-state high angle-of-attack flight is shown in

Figure 5-7 for the elevator commands and associated values of pitch rate. In this case,

the elevator commands in Figure 5-7 are a factor of 5 less than the elevator commands in

Figure 5-1 to indicate the pilot is merely moving the control surfaces to maintain flight

condition and thus not introducing significant energy.

43

0 10 20 30 40 50−40

−30

−20

−10

0

10

20

30

Time (s)

Ele

vato

r (%

)

0 10 20 30 40 50−100

−50

0

50

100

Time (s)

Pitc

h R

ate

(deg

/s)

Figure 5-7. Elevator (left) and Pitch Rate (right) During Steady Flight: Measured (—)and Simulated (−−−)

The contribution to pitch rate that is not predicted by the model in Equation 5–3 is

shown in Figure 5-8. Clearly the model is not able to reproduce the complete response

indicating the elevator is not able to account for the entirety of the measured pitch rate.

0 10 20 30 40 50−100

−50

0

50

100

Time (s)

Pitc

h R

ate

(deg

/s)

Figure 5-8. Uncommanded Pitch Rate

This uncommanded estimate of pitch rate from Figure 5-8 is represented in the

frequency domain in Figure 5-9 from a Fourier transform. The energy is concentrated

across low frequencies but a pair of modes, around 3 rad/s and 6 rad/s, is clearly

evident. These modes may correlate to a short-period mode; however, they are more

likely associated with some unmodeled dynamic of the high angle-of-attack condition.

44

10−1

100

101

102

10−2

100

102

104

Frequency (rad/s)

Mag

nitu

de

Figure 5-9. FFT of Measured Pitch Rate (...), Uncommanded Pitch Rate (—), andElevator Input (−−−)

Figure 5-9 indicates that between approximately 2 and 4 rad/s the uncommanded

pitch rate is larger than the measured pitch rate. This seems improper, as the uncommanded

pitch rate is extracted from the measured pitch rate and should therefore be a subset of

the measured pitch rate. However, phase shifts between the model-predicted pitch rate

and the measured pitch rate can result in uncommanded pitch rates estimated at higher

values than the measured pitch rate.

5.3.2 Lateral

The roll rate measured during steady-state high angle-of-attack flight is shown

in Figure 5-10 along with associated aileron commands. The deflections of the aileron

are nearly an order of magnitude less than the size of the doublets in Figure 5-3 to

indicate the pilot is providing only minimal excitation. Some unexplained drift to negative

deflection is clearly evident in Figure 5-10; however, this drift is extremely low-frequency

and thus can be directly eliminated in the analysis.

The uncommanded portion of the response, as determined by subtracting the

simulated response to the aileron from the measured response, is shown in Figure 5-11.

This portion is quite large in magnitude and actually appears quite periodic.

45

0 10 20 30 40−50

−40

−30

−20

−10

0

10

Time (s)

Aile

ron

(%)

0 10 20 30 40−200

−150

−100

−50

0

50

100

150

Time (s)

Rol

l Rat

e (d

eg/s

)

Figure 5-10. Aileron (left) and Roll Rate (right) During Steady Flight: Measured (—) andSimulated (−−−)

0 10 20 30 40−150

−100

−50

0

50

100

150

Time (s)

Rol

l Rat

e (d

eg/s

)

Figure 5-11. Uncommanded Roll Rate

The periodicity noted in Figure 5-11 is quantified by a Fourier transform on that

data to obtain the frequency-domain representation in Figure 5-12. The frequency-domain

representation of the measured roll rate and the uncommanded roll rate are shown

along with the aileron command. The roll rate and the aileron show a peak at 2.2 rad/s

indicating the pilot is actively controlling some dynamic that is affecting the maintenance

of steady-state flight; however, only the roll rate shows a peak at 3.65 rad/s. The large

amount of energy at this frequency is indicative of a significant level of wing rock at high

angle-of-attack conditions for this aircraft.

46

10−1

100

101

102

10−2

100

102

104

Frequency (rad/s)

Mag

nitu

de

Figure 5-12. FFT of Measured Roll Rate (...), Uncommanded Roll Rate (—), and AileronInput (−−−)

5.3.3 Directional

The steady-state characteristics of the directional dynamics is investigated using the

data in Figure 5-13. This data consists of small rudder commands to maintain the flight

condition and the associated yaw rates which were measured and simulated from the

model in Equation 5–5.

0 10 20 30 40 50−30

−20

−10

0

10

20

30

40

Time (s)

Rud

der

(%)

0 10 20 30 40 50−60

−40

−20

0

20

40

60

80

Time (s)

Yaw

Rat

e (d

eg/s

)

Figure 5-13. Rudder (left) and Yaw Rate (right) During Steady Flight: Measured (—) andSimulated (−−−)

47

The amount of yaw rate which can not be predicted by the model is given in

Figure 5-14. The yaw rate also has some clear periodicity and a substantial magnitude

indicating some element of steady-state flight is not captured by the model generated from

doublets.

0 10 20 30 40 50−40

−20

0

20

40

60

80

Time (s)

Yaw

Rat

e (d

eg/s

)

Figure 5-14. Uncommanded Yaw Rate

The frequency-domain representation of the yaw rate, both measured and uncommanded,

along with rudder commands indicates the periodicity. This data, as shown in Figure 5-15,

has a noticeable mode near 2.2 rad/s. Both the rudder and the yaw rate contain this

mode, as did the aileron and the roll rate, which may mean it correlates to a roll-yaw

coupled mode that the pilot is attempting to damp out and maintain condition.

Figure 5-15 indicates frequency regions where the uncommanded yaw rate is of

larger correlation than the measured yaw rate, particularly between approximately 2.5

and 5 rad/s and 6 and 10 rad/s. This seems improper, as the uncommanded yaw rate is

extracted from the measured yaw rate and should therefore be a subset of the measured

yaw rate. However, phase shifts between the model-predicted yaw rate and the measured

yaw rate can result in uncommanded yaw rates estimated at higher values than the

measured yaw rate.

48

10−1

100

101

102

10−2

100

102

104

Frequency (rad/s)

Mag

nitu

de

Figure 5-15. FFT of Measured Yaw Rate (...), Uncommanded Yaw Rate (—), and RudderInput (−−−)

49

CHAPTER 6VERTICAL TAIL RELATIONSHIP TO WING ROCK

6.1 Upright Tail

A series of flight tests are performed with the airplane in an upright orientation with

different vertical tails. Roll rates were examined as the primary indicator of wing rock

characteristics.

The roll rate measured during these tests is shown in Figure 6-1 along with the

associated aileron deflections. The aileron deflection is dramatically smaller than the

roll rate and thus difficult to distinguish. Clearly the roll rate shows some amount of

periodicity among all the tails; consequently, wing rock appears for any size of these tails.

Also, the magnitude of the roll rate shows variation during the response; however, this

magnitude is actually somewhat consistent despite variations in the tail size.

A frequency-domain representation of the roll rates and aileron deflections from

Figure 6-1 is computed and shown in Figure 6-2 using a Fourier transform. The roll rate

shows a consistent amount of energy around 4 rad/s which correlates with the consistent

magnitude observed in the time-domain responses. Very little energy is introduced by

the aileron around 4 rad/s, indicating that the energy observed in figure 6-2 represents

uncommanded wing rock. Also, this peak in energy is actually somewhat broad for every

tail and ranges from approximately 2 rad/s to 6 rad/s indicating the wing rock is a

broad-band phenomenon.

A time-frequency representation is computed for the roll rate to investigate the

temporal nature of any instantaneous frequencies in the wing rock. These representations

are shown in Figure 6-3 as computed by wavelet transforms using a Morlet wavelet.

The wing rock is evident by the high correlations shown around 4 rad/s; however,

this representation is notably different than the frequency-domain characterization in

Figure 6-2. The wing rock is shown to actually have a narrow band of energy when

50

020

4060

8010

0−

150

−10

0

−5005010

0

150

Tim

e (s

)

Magnitude

020

4060

8010

0−

150

−10

0

−5005010

0

150

Tim

e (s

)

Magnitude

020

4060

−15

0

−10

0

−5005010

0

150

Tim

e (s

)

Magnitude

020

4060

8010

0−

150

−10

0

−5005010

0

150

Tim

e (s

)

Magnitude

050

100

150

−15

0

−10

0

−5005010

0

150

Tim

e (s

)

Magnitude

Fig

ure

6-1.

Tim

eR

espon

ses

inU

pri

ght

Con

figu

rati

onfo

rA

)Sm

alle

rTai

l,B

)Sm

allTai

l,C

)N

orm

alTai

l,D

)B

igTai

l,E

)B

igge

rTai

l:R

ollR

ate

(—)

inde

g/s

and

Ailer

onD

eflec

tion

(...)

in%

51

10−1

100

101

102

100

102

104

106

Roll Rate FFTs

Frequency (rad/s)

Mag

nitu

de

10−1

100

101

102

100

102

104

106

Roll/Aileron FFt

Frequency (rad/s)

Mag

nitu

de

Figure 6-2. FFT in Upright Configuration of Roll Rate (left) and Aileron Deflection(right): Smaller Tail (– –), Small Tail (–.–.), Normal Tail (—), Big Tail (...),Bigger Tail (–x–)

localized in time using Figure 6-3. The broad-band nature observed in Figure 6-2 results

from the variations observed in Figure 6-3 in the central frequency of that narrow band.

The wavelet transforms of the aileron deflection, as shown in Figure 6-4, do not

show significant correlation with the wing rock. In each case, the aileron deflections are

predominately at lower frequencies.

6.2 Inverted Tail

The airplane is flown in an inverted configuration for each of the different sizes of

vertical tail. The roll rate measured during these tests is shown in Figure 6-5 along with

the associated aileron deflections. The roll rates are approximately an order of magnitude

greater than the aileron deflections although both appear to have periodicity. In each case,

this magnitude and periodicity are relatively consistent despite variations to the tail.

The frequency-domain representations, shown in Figure 6-6, agree with the time-domain

analysis of Figure 6-5. The energy of the roll rate is about an order of magnitude greater

than the aileron for each tail. Also, the roll rate shows a minor broad-band peak near

4 rad/s although the energy is not excessive.

A wavelet transform is applied to the time-domain data to generate the representation,

shown in Figure 6-7, in the time-frequency domain. These plots show some correlation

52

Tim

e (s

)

Frequency (rad/s)

020

4060

80

23410 6

2.5

Tim

e (s

)

Frequency (rad/s)

020

4060

8010

0

2

2.534610

Tim

e (s

)

Frequency (rad/s)

010

2030

4050

6070

2

2.534610

Tim

e (s

)

Frequency (rad/s)

020

4060

8010

0

2

2.534610

Tim

e (s

)Frequency (rad/s)

020

4060

8010

012

014

016

0

2

2.534610

Fig

ure

6-3.

Wav

elet

Tra

nsf

orm

sof

Rol

lR

ate

inU

pri

ght

Con

figu

rati

on:

A)

Sm

alle

rTai

l,B

)Sm

allTai

l,C

)N

orm

alTai

l,D

)B

igTai

l,E

)B

igge

rTai

l

53

Tim

e (s

)

Frequency (rad/s)

020

4060

80

2

2.534610

Tim

e (s

)

Frequency (rad/s)

020

4060

8010

0

2

2.534610

Tim

e (s

)

Frequency (rad/s)

010

2030

4050

6070

2

2.534610

Tim

e (s

)

Frequency (rad/s)

020

4060

8010

0

2

2.534610

Tim

e (s

)Frequency (rad/s)

020

4060

8010

012

014

016

0

2

2.534610

Fig

ure

6-4.

Wav

elet

Tra

nsf

orm

sof

Ailer

onD

eflec

tion

inU

pri

ght

Con

figu

rati

on:

a)Sm

alle

rTai

l,b)

Sm

allTai

l,c)

Nor

mal

Tai

l,d)

Big

Tai

l,e)

Big

ger

Tai

l

54

010

2030

4050

60−

150

−10

0

−5005010

0

150

Tim

e (s

)

Magnitude

010

2030

4050

60−

150

−10

0

−5005010

0

150

Tim

e (s

)

Magnitude

010

2030

4050

60−

150

−10

0

−5005010

0

150

Tim

e (s

)

Magnitude

020

4060

80−

150

−10

0

−5005010

0

150

Tim

e (s

)

Magnitude

020

4060

−15

0

−10

0

−5005010

0

150

Tim

e (s

)

Magnitude

Fig

ure

6-5.

Tim

eR

espon

ses

inIn

vert

edC

onfigu

rati

onfo

rA

)Sm

alle

rTai

l,B

)Sm

allTai

l,C

)N

orm

alTai

l,D

)B

igTai

l,E

)B

igge

rTai

l:R

ollR

ate

(—)

and

Ailer

onD

eflec

tion

(...)

55

10−1

100

101

102

100

102

104

106

Roll Rate FFTs

Frequency (rad/s)

Mag

nitu

de

10−1

100

101

102

100

102

104

106

Roll/Aileron FFt

Frequency (rad/s)

Mag

nitu

de

Figure 6-6. FFT in Inverted Configuration of Roll Rate (left) and Aileron Deflection(right): Smaller Tail (– –), Small Tail (–.–.), Normal Tail (—), Big Tail (...),Bigger Tail (–x–)

around 4 rad/s; however, the magnitude of correlation is not excessively high in comparison

to the lower frequencies. This lack of excessive correlation agrees with the Fourier

transforms in Figure 6-6 and is evident for each tail.

A similar time-frequency representation of the aileron data, as computed through

wavelet transform and shown in Figure 6-8, indicates a strong correlation to the roll rates

that are shown in Figure 6-7. In particular, most of the peaks in correlation for roll rate

are matched by a peak in aileron at the same time and frequency. Examples of such

input/output matching can be seen with the smaller tail at approximately 2.5 to 3 rad/s

between 50 and 60 s and with the bigger tail at approximately 3 rad/s between 10 and

20 s and at approximately 3 to 4 rad/s between 45 and 60 s.

6.3 Parameterization

The time-domain plots of roll rate in upright and inverted high angle-of-attack flight,

shown in Figures 6-1 and 6-5 respectively, are analyzed. Peak averages and frequencies are

extracted for each vertical tail size and shown in Figure 6-9.

The time-domain analysis in Figure 6-9 indicates that wing rock frequency and

magnitude does not depend upon the size of the vertical tail. However, the wing rock

magnitude does change based on the vertical tail configuration: the average wing rock

56

Tim

e (s

)

Frequency (rad/s)

010

2030

4050

2

2.534610

Tim

e (s

)

Frequency (rad/s)

010

2030

4050

60

2

2.534610

Tim

e (s

)

Frequency (rad/s)

010

2030

4050

60

2

2.534610

Tim

e (s

)

Frequency (rad/s)

010

2030

4050

6070

2

2.534610

Tim

e (s

)Frequency (rad/s)

010

2030

4050

60

2

2.534610

Fig

ure

6-7.

Wav

elet

Tra

nsf

orm

sof

Rol

lR

ate

inIn

vert

edC

onfigu

rati

on:

A)

Sm

alle

rTai

l,B

)Sm

allTai

l,C

)N

orm

alTai

l,D

)B

igTai

l,E

)B

igge

rTai

l

57

Tim

e (s

)

Frequency (rad/s)

010

2030

4050

2

2.534610

Tim

e (s

)

Frequency (rad/s)

010

2030

4050

60

2

2.534610

Tim

e (s

)

Frequency (rad/s)

010

2030

4050

60

2

2.534610

Tim

e (s

)

Frequency (rad/s)

010

2030

4050

6070

2

2.534610

Tim

e (s

)Frequency (rad/s)

010

2030

4050

60

2

2.534610

Fig

ure

6-8.

Wav

elet

Tra

nsf

orm

sof

Ailer

onD

eflec

tion

inIn

vert

edC

onfigu

rati

on:

A)

Sm

alle

rTai

l,B

)Sm

allTai

l,C

)N

orm

alTai

l,D

)B

igTai

l,E

)B

igge

rTai

l

58

Figure 6-9. Mean Roll Rate Peak Magnitudes (—) and Peak-to-Peak Frequencies (...)

peak magnitude in the inverted flight condition is approximately half of that in the

upright flight condition across all tail sizes. The average wing rock frequency is nearly

constant at approximately 4 rad/s for all tail sizes in both upright and inverted high

angle-of-attack flight, which agrees with the peaks seen in the Fourier transforms in

Figures 6-2 and 6-6.

The difference in magnitude of uncommanded wing rock between upright and inverted

flight is likely more pronounced than Figure 6-9 indicates. Figures 6-3 and 6-4 show little

correlation between aileron input and roll rate frequencies in the upright configuration.

Figures 6-7 and 6-8 indicate, however, that a strong correlation exists between roll

rate and aileron input frequencies at essentially all times when roll rate oscillations are

observed in the inverted configuration. This indicates that uncommanded wing rock is

virtually nonexistent when in the inverted configuration.

The upper and lower bounds of wing rock frequency are extracted from Figures 6-2

and 6-3 and are shown in Figure 6-10. Only the upright configuration is analyzed for such

bounds because only uncommanded wing rock is of interest. The upper and lower bounds

as determined by both the FFTs and wavelets are fairly consistent for all upright tail

59

sizes. These bounds indicate that wing rock exists across a spectrum of frequencies, with

bandwidths of approximately 2 rad/s for all tail sizes.

Figure 6-10. Upper and Lower Bounds of Wing Rock Frequency from FFTs (—) andWavelets (...)

The time-frequency representations shown in Figure 6-3 actually indicate that

the wing rock oscillation is much narrower in bandwidth when localized in time than

Figure 6-10 indicates. The average bandwidth of the wing rock oscillations at all times

for each tail are extracted from the wavelet plots and are shown in Figure 6-11 along with

the broad bands extracted from Figure 6-10. The bandwidth at most times is notably

smaller in Figure 6-11 than in Figure 6-10. This indicates that wing rock is actually a

fairly narrow-band phenomenon, but variations in frequency over time within a broader

band lead to the appearance of a broad-band phenomenon with many classical analysis

techniques such as the FFT.

60

Figure 6-11. Mean Wing Rock Bandwidths: FFT (—), Wavelet Broad Limits (...), andWavelet Narrow Limits (– – –)

61

CHAPTER 7SUMMARY

7.1 Recommendations

High angle-of-attack flight can expand the flight envelope of a small UAV if it has

sufficient thrust-to-weight ratio and control surface size and deflection. The expanded

regions of the flight envelope, including lower airspeed limits and smaller recovery area

requirements, can enhance mission capability if maneuvering control can be achieved. As

such, modeling the flight dynamics is required to analyze the appropriateness of various

control schemes in this unconventional realm of flight.

The models identified from the flight data are somewhat limited in usefulness.

The models may be applicable with slight modifications to other aircraft with similar

configurations and sizing, but their usefulness beyond such a narrow range of aircraft

is unknown. The least-squares models display the general simplicity and complexity of

modeling the various axes of the aircraft dynamics. However, they do not accurately

predict substantial uncommanded motion that appears to be present about all three axes.

Additionally, the lack of angle-of-attack and angle-of-sideslip measurements or

estimates introduces uncertainty in modeling. Work is currently underway to obtain

inertial-based estimates of both such angles from the current avionics suite. The inclusion

of such parameters could greatly improve the accuracy of the models and permit higher

fidelity modeling approaches.

Angle-of-attack and sideslip estimates could also improve the ability to analyze wing

rock behavior in conjunction with time-frequency analysis techniques. Of particular

interest in future studies is the relationship between the wing rock frequency and

the angle-of-attack. A correlation between the variations in wing rock frequency and

fluctuations in angle-of-attack would be an extremely useful result. A lack of any

correlation between such parameters would also be notable finding.

62

The dependency of wing rock on vertical tail configuration is a useful result.

Further study of wing rock characteristics could analyze dependencies on other airframe

geometries, such as V-tail and twin-tail configurations, combined upper and lower vertical

tails, wingtip endplates, and flap/spoiler configurations.

An additional branch of future studies on this topic could include analyzing high

angle-of-sideslip flight. The high angle-of-sideslip flight realm is utilized very little in

practical applications; however, it could potentially increase maneuverability for operations

within urban environments or aid in sensor pointing for various surveillance and sensing

missions.

7.2 Conclusion

The flight dynamics at high angle-of-attack conditions are challenging to model.

An extensive set of data from a remote-controlled Mini Showtime aircraft is analyzed

in response to both commanded doublets and steady-state flight. The responses to

doublets demonstrate that the lateral dynamics are relatively straightforward to model

with a dominant roll convergence mode, while the longitudinal dynamics and directional

dynamics need nonlinearities and numerous states along with lags to properly represent

the flight characteristics. The responses at steady-state flight are particularly interesting

in that the roll rate indicates a significant contribution from wing rock while both the

roll rate and yaw rate have a contribution from a mode that the pilot is attempting to

damp. As such, a partial analysis of the behavior at high angle-of-attack conditions is

generated that demonstrates phenomena and characteristics that are not as evident at low

angle-of-attack conditions.

Uncommanded wing rock at high angle-of-attack flight is a complex phenomenon.

An extensive set of high angle-of-attack flight data from the Mini ShowTime aircraft

with multiple vertical tail sizes and configurations is analyzed. The aircraft has very few

asymmetries, except for the vertical tail, which allows upright and inverted flight to be

compared. Time-frequency analysis indicates that when the vertical tail is in the upright

63

position, pronounced uncommanded wing rock is present at approximately 4 rad/s for all

sizes of vertical tail. Uncommanded wing rock was virtually eliminated when a vertical tail

of any size was in the inverted configuration. These findings indicate that for this aircraft

uncommanded wing rock is greatly influenced by the presence or absence of upright and

inverted vertical tails, but not necessarily the size of such vertical tails.

64

REFERENCES

[1] Wright, K., and Lind, R., “Investigating Sensor Emplacement on Vertical Surfacesfor a Biologically-Inspired Morphing Design from Bats,” AIAA Paper 2007-6488,Aug. 2007.

[2] Grant, D., Abdulrahim, M., and Lind, R., “Flight Dynamics of a Morphing AircraftUtilizing Multiple-Joint Wing Sweep,” AIAA Paper 2006-6505, Aug. 2006.

[3] Regenie, V., Gatlin, D., Kempel, R., and Matheny, N., “The F-18 High AlphaResearch Vehicle: A High-Angle-of-Attack Testbed Aircraft,” NASA TM 104253,1992.

[4] Iliff, K., and Wang, K., “Flight Determined Subsonic Longitudinal Stability andControl Derivatives of the F-18 High Angle of Attack Research Vehicle (HARV)With Thrust Vectoring,” NASA TP 97-206539, 1997.

[5] Bauer, J., Clarke, R., and Burken, J., “Flight Test of the X-29A at High Angle ofAttack: Flight Dynamics and Controls,” NASA TP 3537, 1995.

[6] NASA Dryden F-18 High Alpha Research Vehicle (HARV) Photo Collection.July 26, 2001. National Aeronautics and Space Administration. Feb. 23, 2009<http://www.dfrc.nasa.gov/gallery/Photo/F-18HARV/>.

[7] NASA Dryden X-29 Photo Collection. Feb. 3, 2000. National Aeronautics and SpaceAdministration. Feb. 23, 2009 <http://www.dfrc.nasa.gov/gallery/Photo/X-29/>.

[8] AeroVironment Unmanned Aircraft Systems, <http://avinc.com/UAS.asp>.

[9] Quigley, M., Barber, B., Griffiths, S., and Goodrich, M., “Towards Real-WorldSearching with Fixed-Wing Mini-UAVs,” IEEE/RSJ International Conference onIntelligent Robots and Systems, Aug. 2005.

[10] AeroVironment, Inc.: Media Gallery. 2009. AeroVironment, Inc. Feb. 23, 2009<http://www.avinc.com/media gallery2.asp?id=224>.

[11] Frank, A., McGrew, J., Valenti, M., Levine, D., and How, J., “Hover, Transition,and Level Flight Control Design for a Single-Propeller Indoor Airplane,”AIAA Paper 2007-6318, Aug. 2007.

[12] Johnson, E., Turbe, M., Wu, A., Kannan, S., and Neidhoefer, J., “Flight TestResults of Autonomous Fixed-Wing UAV Transitions to and from StationaryHover,” AIAA Paper 2006-6775, Aug. 2006.

[13] Gree, W., and Oh, P., “Autonomous Hovering of a Fixed-Wing Micro Air Vehicle,”IEEE International Conference on Robotics and Automation, May 2006.

[14] Hwang, C., and Pi, W. S., “Some Observations on the Mechanism of Aircraft WingRock,” Journal of Aircraft, Vol. 16, No. 6, pp. 366-373, 1979.

65

[15] Liebst, B., Nolan, R., “Method for the Prediction of the Onset of Wing Rock,”Journal of Aircraft, Vol. 31, No. 6, pp. 1419-1421, 1994.

[16] Regenie, V., Gatlin, D., Kempel, R., and Matheny, N., “The F-18 High AlphaResearch Vehicle: A High-Angle-of-Attack Testbed Aircraft,” NASA TM 104253,1992.

[17] Ramnath, R. V., and Go, T. H., “An Analytical Approach to the Aircraft WingRock Dynamics,” AIAA Paper 2001-4426, 2001.

[18] Go, T. H., and Ramnath, R. V., “An Analysis of the Two Degree-of-Freedom WingRock on Advanced Aircraft,” Journal of Guidance, Control, and Dynamics, Vol. 25,No. 2, pp. 324-333, 2002.

[19] Schmidt, L. V., “Wing Rock Due to Aerodynamic Hysteresis,” Journal of Aircraft,Vol. 16, No. 3, pp. 129-133, 1979.

[20] Lie, F. A. P., and Go, T. H., “Analysis of Single Degree-of-Freedom Wing Rock dueto Aerodynamic Hysteresis,” AIAA Paper 2007-6489, 2007.

[21] Abramov, N., Goman, M., Demenkov, M., and Khabrov, A., “Lateral-DirectionalAircraft Dynamics at High Incidence Flight with Account of Unsteady AerodynamicEffects,” AIAA Paper 2005-6331, 2005.

[22] Bauer, J., Clarke, R., and Burken, J., “Flight Test of the X-29A at HighAngle-of-Attack: Flight Dynamics and Controls,” NASA TP 3537, 1995.

[23] Liebst, B. S., and DeWitt, B. R., “Wing Rock Suppression in the F-15 Aircraft,”AIAA Paper 1997-3719, 1997.

[24] Iliff, K. W., and Wang, K. C., “X-29A Lateral-Directional Stability and ControlDerivatives Extracted from High-Angle-of-Attack Flight Data,” NASA TP 3664,1996.

[25] Davison, M. T., “An Examination of Wing Rock for the F-15,” Master’s Thesis, AirForce Inst. of Technology, AFIT/GAE/ENY/92M-01, Feb. 1992.

[26] Brandon, J. M., and Nguyen, L. T., “Experimental Study of Effects of ForebodyGeometry on High Angle-of-Attack Stability,” Journal of Aircraft, Vol. 25, No. 7,pp. 591-597, 1988.

[27] Meyer, R. R., “Overview of the NASA Dryden Flight Research Facility AeronauticalFlight Projects,” NASA TM 104254, 1992.

[28] Ericsson, L. E., “Various Sources of Wing Rock,” Journal of Aircraft, Vol. 27, No. 6,pp. 488-494, 1990.

66

[29] Walton, J., and Katz, J., “Application of Leading-Edge Vortex Manipulations toReduce Wing Rock Amplitudes,” Journal of Aircraft, Vol. 30, No. 4, pp. 555-557,1993.

[30] Ericcson, L. E., “Sources of High Alpha Vortex Asymmetry at Zero Sideslip,”Journal of Aircraft, Vol. 29, No. 6, pp. 1086-1090, 1992.

[31] Ericsson, L. E., “Wing Rock Generated by Forebody Vortices,” Journal of Aircraft,Vol. 26, No. 2, pp. 110-116, 1989.

[32] Ericsson, L. E., “Further Analysis of Wing Rock Generated by Forebody Vortices,”Journal of Aircraft, Vol. 26, No. 22, pp. 1098-1104, 1989.

[33] L. E. Ericsson, “Effect of Deep-Stall dynamics on Forebody-Induced Wing Rock,”AIAA Paper 1996-3404, 1996.

[34] Alcorn, C. W., Croom, M. A., Francis, M. S., and Ross, H., “The X-31 Aircraft:Advances in Aircraft Agility and Performance,” Progress in Aerospace Sciences,Vol. 32, pp. 377-413, 1996.

[35] Gilbert, W. P., Nguyen, L. T., and Gera, J., “Control Research in the NASAHigh-Alpha Technology Program,” AGARD Fluid Dynamics Panel Symposium onAerodynamics of Combat Aircraft Control and Ground Effects, 1989.

[36] Mini ShowTime 4D ARF by E-flite (EFL2500), Horizon Hobby, Inc.,<http://www.horizonhobby.com>, Champaign, IL, 2006-2009.

[37] Smith, H. C., The Illustrated Guide to Aerodynamics, TAB Books, Blue RidgeSummit, PA, pp. 228-229, 1992.

[38] nIMU Datasheet Version 2.7, MEMSense, LLC, <http://www.memsense.com>,Rapid City, SD, 2007.

[39] Instruction Manual for the USB Flight Data Recorder Version 3.7, Eagle TreeSystems, LLC, <http://www.eagletreesystems.com>, Bellevue, WA, 2003-2005.

[40] Perry, J., Mohamed, A., Johnson, B., and Lind, R., “Estimating Angle of Attack andSideslip Under High Dynamics on Small UAVs,” Proceedings of the 21st Interna-tional Technical Meeting of the Satellite Division of the Institute of Navigation IONGNSS 2008, pp. 1165-1173, 2008.

[41] Misiti, M., Misiti, Y., Oppenheim, G., and Poggi, J., Wavelet Toolbox User’s Guide,Version 2, The MathWorks, Inc., Natick, MA, Chap. 1,6, 2000.

[42] Strang, G., Nguyen, T., Wavelets and Filter Banks, Wellesley-Cambridge Press,Wellesley, MA, Chap. 1, 1996.

[43] Klein, V., and Morelli, E. A., Aircraft System Identification Theory and Practice,AIAA Education Series, AIAA, Reston, VA, Chap. 5, 2006.

67

[44] Yechout, T., Introduction to Aircraft Flight Mechanics, AIAA Education Series,AIAA, Reston, VA, pp. 176-177, 2003.

68

BIOGRAPHICAL SKETCH

Baron Johnson’s aviation pursuits began with his first aerobatic rides a few days

prior to being born in Memphis, TN in 1983. After moving to Ocala, FL, Baron soloed an

L23 Super Blanik sailplane at the age of 14, and then soloed a Piper J-3 Cub on his 16th

birthday. Baron then pursued his private pilot license, which he received at the age of 17,

and his commercial pilot license, which he received at the age of 18, along with instrument

and multi-engine ratings. Baron has accumulated over 600 hours of flight experience,

including very unique opportunities such as flying in part of a 60-ship formation of and

serving as in-flight safety observer in the EAA AirVenture airshow.

After graduating from Belleview High School in 2002, Baron attended the University

of Florida in Gainesville, FL. He graduated cum laude with a Bachelor of Science in

aerospace engineering in 2007, and received a Master of Science in aerospace engineering

in 2009. While studying, Baron has had opportunities to work on many exciting UAV and

MAV programs while employed by the Florida Cooperative Fish & Wildlife Research Unit,

Micro Air Vehicle Lab, and Flight Control Lab. While member and pilot for the UF Micro

Air Vehicle Team, the team captured 4 consecutive victories at the International Micro Air

Vehicle Competition.

Baron has also been an active RC aviator during his years. His primary passions

in RC are IMAC Scale Aerobatics and F3C Helicopter Aerobatics. Baron has had the

opportunity to fly many RC demonstrations and competitions, and has won 4 national

championship titles along with other high finishes at international competitions.

Baron also authored a children’s book titled History Takes A Wild Ride, for which

he was awarded the key to the city of Memphis and appeared on NBC’s The Today Show.

Baron has had some incredible experiences and opportunities in his life for which he is

incredibly grateful.

69