You are on page 1of 121

DEVELOPMENT OF A REAL-TIME FLIGHT SIMULATOR

FOR AN EXPERIMENTAL MODEL HELICOPTER

Diploma Thesis
Cand. aer. Christian Munzinger

Atlanta, December 1998

Georgia Institute of Technology


School of Aerospace Engineering

advised by:
Dr. Anthony J. Calise
Dr. J. V. R. Prasad

IFR
University of Stuttgart

SUMMARY
This work first describes the development of a real-time flight simulator for an R-50
experimental model helicopter. A mathematical model of the helicopter is developed to
represent the dynamics of the real system. This simulation model is used to investigate
and analyze the helicopter dynamics in a hovering flight condition. The importance of a
control rotor used for stability augmentation of the helicopter is emphasized and
investigated in more detail. Combining the model with further flight software and
hardware, a flight simulator is obtained that is capable of real-time flight simulation. This
simulator will be used in the future for detailed studies on new modern control algorithms
used for helicopter flight control.
Experimental flight tests with the real helicopter are performed and analyzed and
allow the identification of a simplified linear model valid close to the hover flight
condition. Results are shown and compared to the linearized model obtained from the
simulation. The system identification employs a frequency response and a step response
method that result in an approximate model for the helicopter dynamics.
Linear models from simulation and flight tests are then used in a recently developed
Neural Network Adaptive Nonlinear Flight Control System and applied to the real-time
simulator and the real helicopter. The results of both applications are then briefly
presented.

ACKNOWLEDGEMENTS
This work was conducted by the School of Aerospace Engineering at the Georgia
Institute of Technology in Atlanta.
I want to thank Dr. Anthony J. Calise of the School of Aerospace Engineering in Atlanta
and Dr. Klaus H. Well of the Institute of Flight Mechanics and Control in Stuttgart for
their support. They both made it possible for me to study at Georgia Tech and enabled me
to realize this work.
I also want to thank Dr. Anthony J. Calise and Dr. J. V. R. Prasad who guided and
assisted me throughout my work and studies and gave me all the support I needed during
this time.
Special thanks to Dr. Eric J. Corban of Guided Systems Technologies, Inc., who went
with me through numerous hardware and software problems related to the flight test
program.
My respect and thanks to the pilot, Mr. Jeong Hur, and the whole flight team, who
suffered only minor heart-attacks during some critical flight maneuvers that would have
brought my work to a sudden end.
Finally I want to thank all my colleagues and friends who supported me during this one
year at Georgia Tech.

Christian Munzinger

Atlanta, Georgia
December 1998

iii

Contents

CONTENTS
Summary .............................................................................................................................ii
Acknowledgements.............................................................................................................iii
Contents ............................................................................................................................iv
Nomenclature ......................................................................................................................vi
List of Figures......................................................................................................................x
List of Tables......................................................................................................................xii
Chapter 1 Introduction.....................................................................................................13
1.1
1.2

Simulation of Flight.................................................................................................. 13
Experimental Flight .................................................................................................. 15

Chapter 2 Helicopter Flight Dynamics............................................................................17


2.1
2.2

General Equations of Unsteady Motion ...................................................................... 17


The Small-Disturbance Theory .................................................................................. 21

Chapter 3 Helicopter Theory ..........................................................................................23


3.1
3.2
3.3

Main Rotor Reference Frames and Notations.............................................................. 23


Hover and Vertical Flight.......................................................................................... 26
Forward Flight ......................................................................................................... 28

3.3.1 Rotor Theory in Forward Flight.................................................................................................... 28


3.3.2 Influences of Rotor Effects and RotorHelicopter Interference .................................................. 29

Chapter 4 Helicopter Stability and Control ...................................................................35


4.1
4.2

Helicopter Control.................................................................................................... 35
Helicopter Stability................................................................................................... 37

4.2.1 Hover............................................................................................................................................... 38
4.2.2 Forward Flight................................................................................................................................ 42
4.2.3 Stability Augmentation with a Control Rotor ............................................................................... 46

Chapter 5 Mathematical Modeling .................................................................................48


5.1
5.2
5.3
5.4
5.5
5.7

General Helicopter Model......................................................................................... 48


Rigid Body Model.................................................................................................... 49
Main Rotor Model.................................................................................................... 49
Control Rotor Model ................................................................................................ 58
Model of Fuselage, Wing and Tail ............................................................................. 63
Simulation Results for the Linarized Model................................................................ 65

Chapter 6 Real-Time Simulation: Hardware and Software .........................................73


6.1
6.2
6.3

Simulation Elements ................................................................................................. 73


Flight System Elements ............................................................................................ 76
Hardware-In-The-Loop-Simulation ........................................................................... 77

iv

Contents

Chapter 7 Validation of the Helicopter Simulation Model ...........................................79


7.1
7.2

Flight Test Data and Requirements ............................................................................ 79


System Identification Procedures ............................................................................... 81

7.2.1 Static Trim Values........................................................................................................................... 81


7.2.2 Frequency Response Analysis........................................................................................................ 84
7.2.3 Step Response Analysis................................................................................................................... 95

Chapter 8 Modern Adaptive Nonlinear Flight Control in


Simulation and Real Flight ..........................................................................101
8.1
8.2

Flight Control System ............................................................................................. 101


Simulation and Experimental Results ....................................................................... 105

References ........................................................................................................................108
Appendix A R-50 Helicopter Data ..............................................................................110
Appendix B Equations of Unsteady Motion of the Rigid Body................................112
Appendix C System and Control Matrices ................................................................113
Appendix D Results of Linear System Analysis for Hover.......................................115
Appendix E R-50 Helicopter System Components....................................................118
Appendix F Simulated and Experimental Bode Plots...............................................120

Nomenclature

NOMENCLATURE
( L, M, N )
( p, q, r )
( u, v, w )
( u, v, w )E
( X, Y, Z )
( x, y, z )a
( x, y, z )B
( x, y, z )E
( , , )
a
a0
a1s
a p , aq , ar

a j
A
A1, A2
A1,SP
AR
b1
B
B1, B2
B1,SP
b1s
bp, bq, br
c
cD0
cm
dhub
eMR

Components of moment about the CG, in body frame


Angular helicopter body rates, in body frame
Velocity components relative to air expressed in body frame
Velocity components relative to the Earth fixed frame
Components of force acting along the (x, y, z) B axes
Helicopter aerodynamic coordinate frame
Helicopter body coordinate frame
Earth fixed coordinate frame
Euler angles
Two-dimensional constant lift curve slope
Coning angle
First harmonic coefficient of longitudinal blade flapping
with respect to shaft (positive for tilt back)
Uncoupled stability derivatives with respect to
uncoupled body angular rates
Estimated system parameters
Rotor disk area
First and second harmonic coefficient of lateral
blade feathering
Lateral swashplate tilt relative to HP
(positive for tilt right)
Blade aspect ratio
First harmonic coefficients of lateral blade flapping
with respect to feathering plane
Number of blades
First and second harmonic coefficient of longitudinal
blade feathering
Longitudinal swashplate tilt relative to HP
(positive for tilt forward)
First harmonic coefficient of lateral blade flapping
with respect to shaft (positive for tilt right)
Uncoupled control derivatives with respect to inputs
resulting in uncoupled body angular rates
Mean blade chord length
Mean profile drag coefficient
Mean profile moment coefficient of control rotor
Horizontal hub distance from helicopter CG
Flap hinge offset of main rotor blade

ft lb
rad/sec
ft/sec
ft/sec
lb

rad
1/rad
rad
rad
1/sec

ft2
rad
rad

rad

rad
rad
rad

ft

ft
ft

vi

Nomenclature

E
E0
E1, E2
F
F1, F2
fwake
G
hhub
is
Ib
Ixy
Ixz
Iyz
kMR
k
K1
K2
Kc
lb
lCR
M &
Mgust
MT
M
R
T
Td

V
wblade
wr
x

Vector of output errors


Mean harmonic coefficient of blade lag motion
First and second harmonic cos-coefficients of
blade lag motion
Matrix of state derivatives
First and second harmonic sin-coefficients of
blade lag motion
db
da
Wake-function for low/high speed effects on 1s , 1s
dv
du
Matrix of control derivatives
Vertical hub distance from helicopter CG
initial shaft tilt (positive back)
Moment of inertia of blade about flapping hinge
Product of helicopter inertia xy dm

xz dm
Product of helicopter inertia yz dm
Product of helicopter inertia

rad
rad

rad

ft
rad
slug ft2
slug ft2
slug ft2
slug ft2

Coefficient defining main rotor blade pitch due to swashplate tilt


Coefficient defining main rotor blade pitch due control rotor tilt
Cross-coupling coefficient due to delta-three-angle
Cross-coupling coefficient due to hinge offset
Total cross-coupling coefficient
Length of aerodynamic blade section of control rotor
ft
Length of control rotor bar
ft
Non-dimensional aerodynamic moment due to blade flapping velocity
Control rotor moment due to wind velocity
ft lb
Torque
ft lb
Non-dimensional aerodynamic moment derivative with
respect to rotor advance ratio
Blade radius
ft
Thrust
lb
Time delay
sec
Total airspeed
ft/sec
Velocity vector relative to the atmosphere
ft/sec
Average velocity of main rotor blade relative to air
ft/sec
Velocity of rotor disk relative to air
ft/sec
State vector of helicopter rigid body motion

vii

Nomenclature

Important Derivatives

da1s
du
db1s
dv
dL
dA1
dL
db1s
dM
da1s
dM
dB1

Derivative of longitudinal TPP tilt with respect to u-velocity

rad/(ft/sec)

Derivative of lateral TPP tilt with respect to v-velocity

rad/(ft/sec)

Roll moment due to lateral cyclic pitch change

ft lb/rad

Roll moment due to lateral TPP tilt

ft lb/rad

Pitch moment due to longitudinal TPP tilt

ft lb/rad

Pitch moment due to longitudinal cyclic pitch change

ft lb/rad

Greek Symbols

c,CR
s
s,CR
3
coll,MR
coll,TR
lat
long
u

coll
twist
0

Parameter vector in system identification


Blade flapping angle
First harmonic coefficient of longitudinal blade flapping
of control rotor with respect to shaft
Control rotor longitudinal TPP tilt
First harmonic coefficient of lateral blade flapping
of control rotor with respect to shaft
Control rotor lateral TPP tilt
Delta-three-angle
Collective main rotor input
Collective tail rotor input
Lateral cyclic input
Longitudinal cyclic input
Input vector to helicopter rigid body motion
Lock number
Phase shift due to time delay
Directional parameter (-1=clockwise, 1=counterclockwise)
induced velocity
Blade pitch angle
Blade pitch due to pilot collective input
Blade twist
Collective main rotor blade pitch

rad
rad
rad
rad
rad
rad
rad
rad
rad
rad
rad
rad
ft/sec
rad
rad
rad
rad
viii

Nomenclature

0,CR

in
off

Constant initial control rotor blade pitch


Density of air
Time constant
Frequency
Flap rate coefficient for in-axis-motion
Flap rate coefficient for off-axis-motion
Rotor rotational speed
Coefficient defining change in natural main rotor frequency
due to hinge offset
Limited extension parameter of control rotor
Rotor blade azimuth

rad
slug/ft3
rad/sec, Hz

rad/sec

rad

Abbreviations
CG
coll
DOF
HP
lat
long
MR
NN
rpm
TR
TPP

Center of gravity
Collective pitch
Degree of freedom
Hub plane
Lateral
Longitudinal
Main rotor
Neural network
Rotor rotational speed
Tail rotor
Tip path plane

ix

List of Figures

LIST OF F IGURES
Figure 2.1: Body axes of the helicopter and notations............................................................................. 18
Figure 2.2: Wind axes of helicopter in forward flight .............................................................................. 19
Figure 2.3: Block diagram, vehicle with plane of symmetry, body axes, flat-earth approximation,
no wind [4].................................................................................................................................. 20
Figure 3.1: Rotor disk and notations........................................................................................................... 24
Figure 3.2: Hub plane, tip-path plane, body axes and notations ........................................................... 24
Figure 3.3: Fundamental blade motion....................................................................................................... 25
Figure 3.4: Rotor blade velocity in forward flight .................................................................................... 28
Figure 3.5: Offset of rotor blade flap hinge ............................................................................................... 31
Figure 3.6: Cross-coupling due to the delta-three-angle ........................................................................ 31
Figure 3.7: Rotor-Fuselage interference in (a) hover and (b) forward flight ....................................... 33
Figure 3.8: Mechanical linkages of the control rotor for the R-50 experimental helicopter ............. 34
Figure 4.1: Longitudinal hover poles dependent on normalized flap frequency =n / ................ 40
Figure 4.2: Typical hover poles for decoupled longitudinal and lateral motion ................................. 41
Figure 4.3: Typical hover poles for coupled longitudinal and lateral motion ..................................... 41
Figure 4.4: Influence of forward speed and horizontal tail on longitudinal poles.............................. 44
Figure 4.5: Influence of forward speed on lateral poles .......................................................................... 45
Figure 4.6: R-50 control rotor providing rate feedback [29] ................................................................. 47
Figure 5.1: Control Rotor of the R-50 Helicopter (view from top).......................................................... 60
Figure 5.2: Poles of coupled longitudinal and lateral motion, no control rotor.................................. 68
Figure 5.3: Poles of coupled longitudinal and lateral motion, with control rotor............................... 70
Figure 5.4: Hover poles of longitudinal motion, with and without control rotor................................. 71
Figure 5.5: Hover poles of lateral motion, with and without control rotor........................................... 72
Figure 6.1: Elements of Simulation Software............................................................................................. 74
Figure 6.2: Display of the R-50 real-time simulator on PC-screen ........................................................ 75
Figure 6.3: Joint GST/Ga Tech real-time hardware-in-the-loop simulation facility [28] .................. 78
Figure 7.1: Trim table for R-50 (simulation), rearward to forward flight ............................................. 83
Figure 7.2: Block diagram of approximated linear helicopter dynamics for hover............................. 85
Figure 7.3: Frequency response in the pitch channel, = 0.75Hz, u q =0.035 rad........................... 87
Figure 7.4: Experimental Bode plot for the pitch channel ....................................................................... 88
Figure 7.5: Experimental Bode plot for the roll channel ......................................................................... 89
Figure 7.6: Experimental Bode plot for the yaw channel......................................................................... 90
Figure 7.7: Experimental and simulated frequency response, 0.31sec time delay, pitch dynamics .. 94

List of Figures

Figure 7.8: Response of identified model (7.17) compared with measured data.................................. 98
Figure 7.9: Response of identified model (7.18) compared with measured data.................................. 100
Figure 8.1: Neural Network Augmented Model Inversion Architecture................................................. 102
Figure 8.2: Multilayered network with one hidden layer......................................................................... 104
Figure 8.3: Simulated system response of R-50, doublet inputs in long. cyclic.................................... 105
Figure 8.4: Pitch response of R-50 in flight test with NN controller in pitch channel ......................... 106
Figure E.1: R-50 fully equipped during flight test .................................................................................... 118
Figure E.2: R-50 fully equipped on transport cart; GST ground control station in background ...... 118
Figure E.3: R-50 avionics box with on-board PC and sensor packet..................................................... 119
Figure E.4: R-50 horizontal tail for improved handling characteristics in forward flight................. 119
Figure F.1: Experimental and simulated frequency response, 0.31sec time delay, roll dynamics .... 120
Figure F.2: Experimental and simulated frequency response, 0.31sec time delay, yaw dynamics .... 121

xi

List of Tables

LIST OF TABLES
Table 4.1: Single rotor helicopter coupling sources [15]........................................................................ 36
Table 5.1: Trim values in simulated hover for R-50, with and without control rotor .......................... 66
Table 5.2: Analytically obtained system matrix in hover, no control rotor........................................... 67
Table 5.3: Analytically obtained control matrix in hover, no control rotor.......................................... 67
Table 5.4: Eigenvalues, damping and frequencies of hover modes, no control rotor ......................... 68
Table 5.5: Analytically obtained system matrix in hover, with control rotor........................................ 69
Table 5.6: Analytically obtained control matrix in hover, with control rotor....................................... 69
Table 5.7: Eigenvalues, damping and frequencies of hover modes, with control rotor ..................... 70
Table 7.1: Measured and simulated trim values for R-50 in hover and forward flight....................... 82
Table 7.2: Identified parameters and system delay in hover using experimental Bode plots ............ 92
Table 7.3: Identified parameters for decoupled lateral motion, step response .................................... 97
Table 7.4: Identified coupled model parameters for system including only angular rates................. 99

xii

Chapter 1

Introduction

Chapter 1
INTRODUCTION
During the early phase of design and development of an aircraft, it needs to be tested
and performance limits need to be validated. For manned flight vehicles it is very
common to use flight simulators before real flight tests, not only to reduce time and costs
during the testing phase, but also to aid in the avoidance of possible loss of pilot and
aircraft in case of a failure. The use of real-flight simulators is therefore mostly related to
manned flight. In this work a simulation for a remotely controlled aircraft of much
smaller size needs to be developed to test a modern controller. The approach and the
mathematical model that will be used are similar to that of a manned helicopter; in fact
the main part of the simulation is taken from a simulation of a full size helicopter. Several
aspects arise from the very different size and therefore result in different dynamic
behavior which need to be taken into account. A brief description of desired capability
and capacity of real flight simulators in general will follow, and the still necessary results
of experimental flight for validation purpose and to build up confidence in handling the
real flight system will be pointed out.

1.1

Simulation of Flight

Rotorcraft and its unique capabilities of vertical take-off and landing, hover, vertical
and forward flight are playing an important roll in commercial and military applications.
Superior hover and low speed performance and agility are coupled with good flight
characteristics even in fast forward flight. The rotor of a helicopter generates the
predominant aerodynamic forces in all flight conditions and is source of forces and
moments on the aircraft that control position, attitude and velocity. An increase of
complexity with respect to rotor and blade dynamics and the still not sufficiently
explained aerodynamic effects of rotor aerodynamic or rotor-body interference
complicate the development of detailed mathematical models of rotors and helicopters in
general.
The use of mathematical models for simulation is therefore limited to some degree of
accuracy and depends on the final objective of the simulation. In general, examining
structural dynamics of single blades or blade sections needs a more accurate model than a
simulation of flight mechanics and aircraft performance. The effort and expense put into
13

Chapter 1

Introduction

modeling and simulation are justified by the helicopters unique capabilities, and the
objective of simulation influences the level of math model fidelity directly related to the
effort and expense required for a given task.
High performance computers made an increase of accuracy of simulation possible,
but the complexity of simulation still rules the capability of operating in real-time. The
development of high-fidelity real-time simulators for research and design, concept
validation and training is strongly dependent on available time and money. Once
developed and validated, the simulator can be very efficiently used to support flight tests
to evaluate handling qualities during the development and design phase of new or
modified helicopter configurations or helicopter components. More advanced
applications, as in unconventional configurations like tilt wings and tilt rotors, unmanned
flight of full size aircraft, simulating emergency situations, validating and testing new
control systems or simple pilot training, underline the usefulness of such a high-fidelity
real-time simulator. Nevertheless, a highly accurate model of one component of a
complex system does not necessarily mean that the simulation of the whole system
behaves like the real physical system. This is especially true for a helicopter, since
components like rotors, fuselage, wings, horizontal or vertical tail, engine and actuators
interact with each other and influence the system response to external and internal
disturbances.
For application in helicopter controls, where the main objective is to control the
dynamic behavior of the helicopter over some desired flight envelope, it is necessary to
find a representative model that shows the same dynamic characteristics as the real
aircraft. On the one hand a detailed model of the main rotor is desired since the dynamics
are governed mainly by the main rotor, but on the other hand a too detailed description
increases the complexity of the simulation and limits the capability of real-time
simulation. Furthermore, most of the existing simulations make a very detailed
knowledge of the simulated system necessary. This knowledge covers exact physical data
of the aircraft geometry, airfoils of blades and wings and aerodynamic data that is gained
in wind tunnel tests. For competitive reasons this data is handled by companies with care
and is therefore generally not available.
If a basic math model can be developed that depends only on basic data sources, then
the inflexibility of very sophisticated models that are now in common use can be
overcome. Additional individual vehicle components can be added and the existing
model can be easily extended or refined. Such a so-called "Minimum-Complexity
Helicopter Simulation Math Model" has been developed by NASA [1]. A modification of
this simulation model will be used and applied to the remotely controlled experimental
14

Chapter 1

Introduction

model helicopter R-50 (see Appendix E), built by YAMAHA [29]. Aerodynamic data
from wind tunnel testing does not exist and the physical data of system components is
known only approximately. The capability of the minimum-complexity math model to be
flexible and easily extendible to other helicopter configurations is not only desired, it is
of great importance for this application. A small-scale model helicopter shows a different
aerodynamic behavior, since it is designed for a very different flight envelope than fullscale helicopters. Further undesirable effects of excessive model complexity are
computational system delays, a great number of system parameters that need to be
determined for each aircraft, inability to easily observe relationships between modeling
parameters and model response (very important for handling quality simulation) and the
inflexibility in temporarily removing undesired dynamics for debugging. The most
important benefit of a minimum-complexity math model is the potential for a more clear
understanding of the cause and the resulting effect. This aspect can become very
important if the dynamic system itself is additionally complicated by a modern control
system that does not allow a fast and easy engineering understanding of the control and
response features.
In summary, the main attribute of a simulator as an effective tool for controller design
is the ability to produce desired results for a specific application and to operate over the
full flight envelope (forward, rearward and sideward flight, hover, transition from hover
to forward flight, vertical climb) with representative handling qualities. Through a manin-the-loop simulation it also becomes a very powerful tool to identify critical manmachine or controller-machine interface issues and allow pilot training within a
reasonable amount of time, costs and risk until confidence in flying with a new system or
flight controller is gained.
Even if satisfying results can be achieved with a high-fidelity real-time simulator, the
results will not be sufficient unless they are confirmed in real flight. In the following, the
objective and the importance of flight tests to validate results will be briefly described.

1.2

Experimental Flight

To increase the fidelity of the mathematical model and to decrease effort required to
create and validate the model, a systematic system identification approach is used to
identify parameters and data that tunes the simulator to fit the highly complex and
nonlinear characteristics of helicopter flight. This approach will be described in Chapters
5 and 7. The data used for the identification process is produced during flight tests. The
aircraft motion is measured, and a model that reflects the physical behavior of the system

15

Chapter 1

Introduction

needs to be found and investigated. The complexity of helicopters, and for the application
described in this work, the reduced size and low payload capability further reduce the
possibility of adding numerous sensors and hardware to measure velocities, rates or
angles in the rotating as well as in the non-rotating frame.
The on-board system of the experimental helicopter consists of a 200 MHz Pentium
based flight control processor and an integrated avionics system. Available on-board
sensors include a 3-axis gyro and accelerometer package, differential GPS with 2 cm
accuracy, 3-axis magnetometer and an 8 channel ultrasonic ranging system. A wireless
digital data link provides a communication link with a mission control ground station.
Representative flight tests are mainly basic maneuvers about hovering, level forward
flight and step inputs to the control stick from carefully trimmed flight conditions. Once
this basic validation is done, more complex maneuvers could be flown, measured and
compared with the simulation results. It is also important to understand operating rules
and human interaction with the flight system. The pilot information given in reports about
handling qualities are not measurable with sensors and are also dependent on personal
experience and pilot skills. However, to evaluate the overall accuracy of the simulation
and its real-time capability, this aspect needs to be investigated. It is obvious that the
visual channel strongly influences the pilots response to a directly observed change in
aircraft motion. This is only one weak point of some real-time simulations, since the hard
and software might not provide the pilot with the 3-D picture he is used to during flight.
Flight tests will give us further information how these facts have to be evaluated and how
they influence the simulation results. Engineering and piloted validation is therefore
necessary, and the quantitative and qualitative handling quality flight tests are described
in more detail in Chapter 7.
In the following work flight tests will mainly be used to create the basis on that a
fine-tuned model can be built on. This is a first important step for the procedure that will
follow to evaluate and investigate advanced control algorithms on model helicopters. The
Uninhabited Flight Research Facility (UFRF) located in the School of Aerospace
Engineering at Georgia Tech was initiated in June 1997 and is dedicated to flight testing
for this purpose. It presently contains two Yamaha model helicopters of the Type R-50.
Testing capability and performance of the simulator is the main objective of this work.
After validation, this simulator will be used to test a neural network (NN) based, adaptive
flight controller, recently developed at Georgia Tech. More details on theory and design
of the developed control algorithm are given in [25, 26, 27]; Chapter 8 only gives a brief
overview of important aspects and should provide the reader with the basic understanding
of Feedback Linearization and Adaptive Neural Networks in flight controls.
16

Chapter2

Helicopter Flight Dynamics

Chapter 2
HELICOPTER FLIGHT DYNAMICS
Like most other flight vehicles, the helicopter body is connected to several elastic
bodies such as rotor, engine and control surfaces. The physical nature of this system is
very complex in shape and motion, and simple mathematical modeling seems not to be
very precise. Nonlinear aerodynamic forces and gravity act on the vehicle, and flexible
structures increase complexity and make a realistic analysis difficult. Several
assumptions can be made to reduce this complexity to formulate and solve relevant
problems. This chapter describes assumptions necessary for a satisfactory modeling of
the helicopter motion and introduces the fundamental motion of the flight vehicle in
general. Some features for the helicopter case are emphasized and explained with respect
to stability analysis and system identification as needed.

2.1

General Equations of Unsteady Motion

Derived from first principles, equations can be found that describe the aircraft as a
rigid body with six degrees of freedom (DOFs), free to move in the atmosphere.
Aerodynamic forces and moments and gravity are incorporated directly in those
equations. In [5] the derivation of the general motion of a mass particle is given, then the
dynamic and kinematic equations for an arbitrary deformable vehicle in flight are
derived. Treating the earth as flat and stationary in inertial space simplifies the model
significantly. For most problems of airplane flight this is acceptable, and for an
experimental helicopter flying at low speed at very low altitude this approximation is
valid. The derived equations contain only a few further assumptions:

The aircraft can be treated as a rigid body with any number of rigid spinning rotors.
There is a plane of symmetry, so that Ixy = Iyz = 0.
The axes of spinning rotors are fixed in the direction relative to the body axes and
have constant angular speed relative to the body axes.

The third assumption seems to be rather crude for the helicopter case since helicopter
motion is mainly controlled by tilting the main rotor relative to the body axes and
therefore creates additional moments and forces. This assumption is justified by assuming
17

Chapter2

Helicopter Flight Dynamics

only changes in rotor tilt of a few degrees relative to the body axes. This is in general true
and no modification needs to be made at this point.
For deriving the equations of motion, it is convenient to choose body axes since all
of the inertias remain constant in the body frame. The x-axis is fixed to a longitudinal
reference line in the aircraft; the y-axis is oriented to the right and the z-axis downward.
A common assumption is the exact symmetry of the aircraft with respect to the xz-plane.
The used body reference frame and its notations for forces, moments and angular rates
about its axis are shown in Figure 2.1.

yB ,Y, M, q

x B, X, L, p

zB, Z, N, r

Figure 2.1: Body axes of the helicopter and notations

Another reference frame that is very useful in formulating the equations of motion is
the earth-fixed frame. With the assumption of a flat and stationary earth, this frame
becomes an inertial system in which Newton's laws are valid. The origin is arbitrary, the
x-axis is horizontally pointing in any convenient direction, the z-axis is pointing
vertically downward. The y-axis is perpendicular to both. For all frames of reference a
right-handed coordinate frame is assumed. The center of gravity (CG) of the aircraft is
equal to the mass center, and its location is given by its Cartesian coordinates relative to
the earth-fixed frame.
To describe the aerodynamic forces the wind frame becomes important, since all of
these forces depend on the velocity relative to the surrounding air mass. This wind frame
is as well fixed to the aircraft, but the x-axis is now oriented along the velocity vector V
of the vehicle relative to the atmosphere. The z-axis lies again in the plane of symmetry;

18

Chapter2

Helicopter Flight Dynamics

the y-axis is perpendicular to both. The origin is again located at the CG of the aircraft.
The wind frame and notations for a helicopter in forward flight are shown in Figure 2.2.

yB ,ya

xa
xB

za

zB

Figure 2.2: Wind axes of helicopter in forward flight

In the case of no velocity relative to the atmosphere, which occurs for a helicopter in
hover with no additional wind, the wind frame is not defined. That makes it necessary to
change back to the body axis if the helicopter motion from or to hover condition needs to
be computed. This again increases the complexity of the computations, but high
performance computers used in flight simulation allow an easy transformation to any
desired frame without remarkable effort. Important transformation matrices can be found
in [4, 5]. The resulting general equations of unsteady motion are also taken from [4] and
listed in Appendix B. This set of equations includes kinematic and dynamic equations
and is presented with respect to body axes. For the case of no wind the equations can be
viewed in a block diagram shown in Figure 2.3. The mathematical system consists of 12
independent equations and the same number of dependent variables.
These variables are:
Position of the Center of Gracity (CG):
Attitude angles (Euler angles):
Velocities relative to the earth fixed frame:
Angular velocities:

xE, yE, zE
, ,
uE, vE, w E
p, q, r

19

Chapter2

,
u,v,w
p,q,r
Control forces

u
Force Equations

p
Moment Equations

Control Moments
,
p,q,r
u,v,w
, ,

v
w

u,v,w
p,q,r

Helicopter Flight Dynamics

q
r

Kinematics 1

Kinematics 2

xE
yE
zE

Figure 2.3: Block diagram, vehicle with plane of symmetry, body axes, flat-earth approximation,
no wind [4]

Each block represents a set of equations with inputs and outputs. The generated
outputs on the right-hand side are the inputs to the left-hand side. Control forces and
moments are dependent on the control inputs. For a helicopter, these inputs are collective
and two cyclic (lateral and longitudinal) stick inputs to the main rotor, pedal inputs
controlling the tail rotor and the throttle controlling the power. In Chapter 5 the effects on
the main rotor tilt and body reactions due to control inputs are examined in more detail.
For stability and control analysis these equations of motion are frequently linearized
about a specific flight condition. From some steady flight condition it is assumed that the
aircraft motion consists of only small deviations from this reference condition. Since the
system of linear equations can also be used for identification purposes of the developed
simulation math model for several flight conditions, relevant aspects of the smalldisturbance theory will be formulated in the following. With respect to the control aspects
it should be pointed out that various flight control structures require at least an
approximate linear model for a vehicle, valid for one or even various flight conditions.
This allows vehicle dynamics to be easily inverted and used, for example, in methods
based on an inverting control scheme [20].

20

Chapter2

2.2

Helicopter Flight Dynamics

The Small-Disturbance Theory

The nonlinear equations of motion listed in Appendix B are still accurate and
therefore very useful for engineering purposes when linearized about some unaccelerated
steady-state flight condition. Aerodynamic effects can be assumed to be linear functions
of disturbances, and the values of linear and angular velocity perturbations are usually
small for many cases [4]. Limitations for this method are disturbances or flight
maneuvers that result in large changes of angles and rates and cause large nonlinearities,
as for example in flight at high angle-of-attack. The detailed derivation of the basic
equations is given in [4].
The linearization process assumes small disturbances, so only first-order terms are
kept, and squares and products are assumed to be negligible. For a steady-state flight
condition all disturbances are set equal to zero. Linear relations to eliminate reference
forces and moments acting on the vehicle in this trimmed flight condition are obtained.
Then the classic assumption of linear aerodynamic theory allows us to express
aerodynamic forces in terms of stability or, more generally, in aerodynamic derivatives.
In determining these derivatives more effort is necessary, and engineers use analytical
and experimental means to find reasonable and accurate results. At this point the
equations of motion for airplanes and helicopters need to be treated separately. An
assumption of decoupled equations of longitudinal and lateral airplane motion is in
general not valid for the helicopter. Derivatives of lateral forces and moments (Y, L, N)
with respect to longitudinal motion variables (u, w, q) are no longer zero, and some
derivatives with respect to rate changes of variables, often negligible in the airplane case,
need to be considered for a helicopter.
The system and control matrices F and G for hover listed in Appendix A show all of
the important gravitational terms that can be obtained analytically and partial derivatives
arising from aerodynamic forces and moments necessary to describe the linear set of
equations for a helicopter. The linear, first-order set of differential equations is then of the
form

x& = F x + G u ,

(2.1)

where x represents the perturbation of state variables uB, wB, B, qB, vB, pB, B and rB
from a steady-state reference flight condition, the trim state. The control vector u
contains deviations from the trim control inputs long, coll,MR, lat and coll,TR. The trim
states as part of the elements inside the matrix are noted with the subscript 0. It is

21

Chapter2

Helicopter Flight Dynamics

convenient to use only the variable names in this matrix form instead of adding another
subscript to denote perturbations. This linear representation is only valid for the initial
angular velocities (pB, qB and rB) equal to zero. The system matrix F includes derivatives
due to small perturbations of system states; the control matrix G represents the
derivatives due to small perturbations of control inputs. It can be seen that the throttle is
not considered to be a control input. For a wide range of flight conditions, the rotational
speed of the rotor does not change, and a variation of throttle is made only to adjust
power keeping some desired rotational rotor speed constant.
One way to obtain the force and moment derivatives is to sequentially perturb the
states and control inputs, positively and negatively from trim values by some small
amount . Then the forces and moments due to both perturbed conditions are computed,
and the derivatives can be obtained by the following equation.

Xu =

X X( u 0 + u) X( u 0 u)

u
2 u

(2.2)

The force X and the state u in this equation represent all the forces, moments, states and
control inputs in the equations of motion.
This approach is used in the later described simulation routine to compute the linear
system matrices for any desired trimmed flight condition. Linear system analysis is very
useful and convenient to examine eigenvalues or eigenvectors, system responses to step
inputs, frequency response and other stability characteristics of a dynamic system. The
system matrices for hover are analyzed in more detail in Chapter 5.
As this chapter has summarized the dynamics of the helicopter treated as a rigid body,
the following chapter introduces the main features of helicopter theory. The main rotor is
essentially responsible for thrust, control forces and moments and is therefore the main
subject of the following investigation. Rotor dynamics and aerodynamics influence the
previously mentioned rigid body dynamics, result in cross-coupling of longitudinal and
lateral motion, and affect the stability of the dynamic system. Chapters 3 and 4
summarize the most important aspects of helicopter theory with respect to the main rotor
and its dynamics.

22

Chapter 3

Helicopter Theory

Chapter 3
HELICOPTER THEORY
As previously mentioned, helicopter dynamics and aerodynamics are mainly affected
by the main rotor. Therefore this chapter will first introduce new main rotor reference
frames and notations that are useful to describe the main rotor in detail. Then it will cover
important aspects of hover and vertical flight and discuss forward flight. A section on
helicopter stability and control follows. Finally the influence of an additional control or
servo rotor to improve stability and handling qualities is described, and the importance of
this dynamic subsystem is pointed out. In general simplifying assumptions are introduced
to allow a faster analysis of the complex main rotor system, and some of the introduced
aspects will be neglected in some applications.

3.1

Main Rotor Reference Frames and Notations

Basically there are three different rotor reference frames used in this main rotor
analysis. The first reference frame is the rotor disk as shown in Figure 3.1. Basic
variables necessary to define and derive the basic equations of main rotor and blade
motion are described. The orientation of rotor rotation in the equations derived is
counterclockwise. Since the main rotor for the experimental helicopter, as for most
remotely controlled helicopters, rotates clockwise, this major difference can be
compensated with a simple sign change in some equations. This additional parameter
describing the direction of rotation will be pointed out when necessary. Further sign
changes in force and moment components due to torque and tail rotor will also be
necessary, as explained in the mathematical modeling of main and tail rotor in Chapter 5.
It is important to mention that the azimuth angle of the blade is defined as zero in the
downstream direction. Azimuth and rotational speed are for now defined positive
counterclockwise.
The two additional rotor reference frames are shown in Figure 3.2. The hub plane
(HP) axes are defined with respect to the main rotor hub that remains fixed relative to the
rigid body of the helicopter. The tip-path plane (TPP) axes are defined with respect to the
motion described by the main rotor blade tips. A first simplifying assumption is that the
thrust vector is always perpendicular to the TPP, which is true in hover and vertical flight
and still very accurate in forward flight. The tilt of the TPP with respect to the HP can be
23

Chapter 3

Helicopter Theory

defined by the two angles, a1s and b1s , referring to longitudinal and lateral tilt of the TPP.
For some helicopters the shaft and therefore the HP is designed with a forward tilt by a
small angle, is , with respect to the helicopter body axes frame. This shaft tilt is not shown
in Figure 3.2, but it will be included in the final force and moment equations for the
helicopter math model.

= 90
Forward
Velocity

Blade

= 180

= 0

Rotor
Disk

= 270
Figure 3.1: Rotor disk and notations

body axes

TPP tip path plane

T
zHP

zTPP

zTPP

HP hub plane

T
zHP

yHP, yTPP

yHP
b1s

a1s

xHP

yTPP
xHP, xTPP

xTPP

yB

TTR

xB
zB
Figure 3.2: Hub plane, tip-path plane, body axes and notations

24

Chapter 3

Helicopter Theory

The motion of the single main rotor blades governs the final tilt of reference planes.
The basic blade motion as treated in this analysis is essentially rigid body rotation about
the root attached to the hub. The degrees of freedom are the angles , , shown in
Figure 3.3. The angle of rotation about an axis in the disk plane, perpendicular to the
blade axis is called flap angle. The lag angle is defined by the rotation about an axis
normal to the disk plane, parallel to the rotor shaft, and the pitch angle is the angle of
rotation about an axis in the disk plane parallel to the blade spar. For a detailed main rotor
analysis more complex motion than this fundamental blade motion needs to be
considered. In this work it is assumed that only the basic flap and pitch motion contribute
to major force and moment calculations and describe the most important rotor
characteristic influencing stability and control of the rigid helicopter.

Blade

Rotor
Shaft
Figure 3.3: Fundamental blade motion

The steady-state blade motion is periodic around the azimuth. Using a Fourier series
expansion, the flap, lag and pitch motion can be written as
= a 0 a1 cos b1 sin a 2 cos 2 b2 sin 2 ...
= E 0 + E1 cos + F1 sin + E 2 cos 2 + F2 sin 2 + ...

(3.1)

= 0 A1 cos B1 sin A2 cos 2 B2 sin 2 + ...


Since the mean and first harmonics (subscript 0 and 1) are most important to rotor
performance and control, all higher order terms will be neglected. The accuracy of the
25

Chapter 3

Helicopter Theory

model remains high whereas the analysis is simplified. The most important angles used in
Equations 3.1 are the coning angle a0, pitch and roll angles of the TPP, a1 and b1,
collective pitch, 0, and cyclic pitch, A1 and B1, commanded by the pilot. The notation is
mainly taken from [1, 6].
It is convenient to compute main rotor forces and moments in the TPP and then
transform them into the HP or directly into the body axes frame. The components can
then be easily added to aerodynamic and gravitational forces and moments acting on the
rigid body. The nature of main rotor forces and moments is briefly described in the
following for different flight conditions, hover, vertical flight and forward flight.

3.2

Hover and Vertical Flight

Hover and vertical flight implies axial symmetry of the rotor and can therefore be
treated as a special case. Analysis is greatly simplified compared to forward flight
described later, and the necessary equations can be written in a nearly closed form.
Momentum theory and lift line theory will be used to determine inflow velocities, thrust
and power for main and tail rotor.
Momentum theory treats the rotor as an actuator disk with zero thickness and circular
surface, able to support a pressure difference and thus accelerate air through the disk.
This resulting airflow is called induced inflow. An approximation of uniform inflow over
the rotor disk is valid in hover and vertical flight and also represents a rough estimation
for forward flight. A more accurate modeling, as described by vortex wake theory and
dynamic wake theory [6, 8], will increase computational cost and is therefore not always
useful for most real-time simulations. For modeling forward flight, a triangular induced
velocity field can be used to increase accuracy. To compute thrust and induced velocity
in general, momentum theory is applied to a specific inflow model. In any case the
iteration of thrust and inflow velocity converges quickly and gives a good estimate of
induced velocity for most flight conditions. More detailed information on rotor wake
theory can be found in [6, 8, 13, 14].
Uniform induced velocity yields minimum induced power loss of an ideal rotor for
given thrust. Other power losses due to non-uniform inflow, non-optimal rotor design,
blade drag and swirl in the wake need to be considered if accuracy is to be increased. For
a finite number of blades additional tip losses reduce effective thrust, affect inflow
characteristics and increase power losses of the rotor. Engine and transmission losses also
affect power calculations, but for most applications a detailed engine model is not
necessary. The constant rotor speed assumption through the entire flight envelope

26

Chapter 3

Helicopter Theory

simplifies this influence on power calculations. Especially in hover, interference between


main rotor, fuselage and tail rotor cause additional power losses and need to be
considered. For forward flight this effect becomes less important due to increasing other
aerodynamic forces. In the case of vertical flight an additional climb or descent velocity
of the helicopter needs to be added, and total power must be further changed due to
necessary climb power.
To compute forces and moments acting on the single blades due to velocity relative to
the surrounding air, blade element theory is used. Linear lifting-line theory is applied to a
rotating wing, the rotor blade, and it is assumed that the aerodynamic forces are produced
by the two-dimensional airfoil of each blade section. The induced angle of attack at a
blade section influences again induced velocity, and therefore again wake-body
interference and further related aerodynamic aspects. Blade element theory is capable of
dealing with the detailed flow and blade loading and allows a very accurate description of
rotor aerodynamics, rotor performance and flight characteristics. To compute the blade
angle of attack a first estimation of the induced velocity at the rotor disk is needed and
can be provided by the momentum theory. The mathematical modeling of the previously
mentioned aspects of rotor aerodynamics is described in Chapter 5. Furthermore,
parameters might be used to adjust constraints or aerodynamic limits like stall speed or
maximum side velocity. If the mathematical modeling of some phenomena can not be
easily done, the existing model needs to be fine-tuned for this very specific experimental
helicopter.
Helicopter stability and control is based on the equations of motion for force and
moment equilibrium on the entire aircraft. In hover, all of the forces and moments are due
to gravity, main rotor, tail rotor or rotor-body interference. Thus the main rotor primarily
governs the stability characteristics of the helicopter for this flight condition. More
detailed remarks about stability and control for hover and forward flight are made in
Chapter 4 . The following section introduces some very important aerodynamic features
of the main rotor in forward flight. The resulting forces and moments strongly influence
flight characteristics of the helicopter and are therefore very important for a desired
accurate modeling and the overall simulation of flight.

27

Chapter 3

3.3

Helicopter Theory

Forward Flight

Rotational motion of the rotor and translational motion of the helicopter are combined
and become a source of additional complexity of rotor theory in forward flight.
Axisymmetry as assumed in hover or vertical flight is no longer valid, because the
aerodynamic environment varies periodically with rotation of rotor blades. The velocity
of the blade relative to the air now governs blade motion. The resulting blade motion and
its influence on forces and moments are the subject of this chapter.
3.3.1

Rotor Theory in Forward Flight

Necessary background of rotor theory in forward flight is provided in this section and
important aspects are explained. For more detailed derivations see [6, 8, 9].
In Figure 3.4 the rotor blade velocity is shown dependent on the azimuth. The
velocity of the advancing blade relative to the air is higher than that of the retreating
blade due to forward velocity and rotation of the blade. Because of this periodic motion,
with the fundamental frequency equal to the rotor speed , the blade aerodynamics as
well as the blade dynamics are first considered.

r+V

Reverse
Flow

Retreating
Side

Advancing
Side
x

Figure 3.4: Rotor blade velocity in forward flight

28

Chapter 3

Helicopter Theory

Momentum theory can again be used to obtain the power due to induced velocity. At
high forward speed this power loss is small compared to other components due to high
velocity, and the assumption of uniform inflow over the entire disk is again valid. But for
transition from hover to fast forward flight the inflow model must be treated more
carefully. This approach is given in [8]. In [1] a mathematical formulation is presented
for the entire range from hover to fast forward flight assuming uniform inflow over the
rotor disk and a triangular velocity field for the rotor wake acting on the fuselage (see
also Chapter 5).
Lifting-line theory for single blade sections integrated over the blade length is used to
derive the force and moment equations for forward flight. Several assumptions can be
made to simplify the analysis and to obtain the equations of blade motion. The following
chapter introduces and explains the influences of blade motion on rigid body dynamics
qualitatively. This provides also the background for the equations used to describe those
influences if desired and necessary for the analysis in Chapter 5.
3.3.2

Influences of Rotor Effects and RotorHelicopter Interference

Influences on power losses, thrust and blade motion depend on blade and rotor
geometry, aerodynamic phenomena and dynamic characteristics of the rotating system
and the rigid body motion. The most important effects are briefly mentioned in the
following sections.
Nonuniform Inflow
Using a linear variation over the disk instead of a uniform inflow can extend the
computation of induced velocity. Additional coefficients can be found dependent on the
forward speed of the helicopter that define a linear distribution of inflow at the rotor.
Typical coefficients result in an inflow model with a small induced velocity at the leading
edge of the rotor disk and about twice the mean value at the trailing edge. The mean
value can still be obtained with the assumption of uniform inflow. This kind of analysis
can be implemented in existing models, and improvements are expected in computation
of mean and first harmonic quantities influencing rotor performance and blade flapping.
If higher harmonics are subject of the analysis, the much more complicated and nonlinear
inflow models become very important. Inflow variation mainly affects the rotor cyclic
flapping and cyclic pitch trim; longitudinal flapping due to inflow variation is small.

29

Chapter 3

Helicopter Theory

Tip Loss and Root Cutout


Finite number of blades instead of a solid rotor disk result in an additional
performance loss. At blade tips the lift decreases to zero, thrust will therefore be reduced
and induced power loss will increase. This tip loss can be accounted for by introducing a
tip loss factor such that there is drag but no lift starting at the radial position defined by
this factor. Major influences can be found on the thrust magnitude. Near the blade root
the airfoil of the blade is different compared to the rest of the blade. Due to reverse flow
at the blade root in forward flight (Figure 3.4), the blade profile is cut out near the blade
root in order to decrease drag. In general, the influences on thrust and flapping moment
are small and can be neglected as long as performance and control aspects are concerned.
Effects of Natural Frequency on Flap Motion
Assuming that the flap hinge of a blade is located at the center of rotation without a
spring producing any additional moment, the natural frequency of the flapping motion is
equal to the frequency of the rotational speed. The resulting moment at the hinge is zero
since the blade is free to move about this hinge. As soon as a spring, some hinge offset or
both (Figure 3.5) is introduced, the resulting moment on the rotor hub is no longer equal
to zero. The additional force of the spring and the offset of the hinge must be considered
in the derivation of the flapping equation of motion. The offset results in a moment arm
for centrifugal, inertial and aerodynamic forces and the additional hub moment must be
added to the moments acting on the rigid body. The natural frequency of the flap motion
for a blade with hinge offset and spring becomes larger than the rotational natural
frequency. The primary effect of the hinge offset and the spring on the flap response is a
coupling of longitudinal and lateral control due to this change of natural frequency. Some
rotor systems instead are designed without a flap hinge and are called hingeless rotors.
They can be treated similarly as described in this section, and [6, 8] deal with further
details of these kind of rotors. At this point it should be pointed out that an additional
control or servo rotor described later, strongly influences the helicopter stability and
performance. This kind of rotor system can approximately be treated like a teetering
rotor. In most applications, for a teetering rotor the coning angle, a0, of the blades in
Equation 3.1 can be disregarded.

30

Chapter 3

Helicopter Theory

Flap Hinge Offset


e

Blade

Flap Hinge

Rotor
Shaft

Figure 3.5: Offset of rotor blade flap hinge

A further coupling of blade pitch and flap motion arises from the geometry used to
control blade pitch. For a pitch bearing outboard of the flap hinge, the blade experiences
a pitch change due to flapping displacement of the blade if the pitch link is not in line
with the flapping hinge (see Figure 3.6).

Blade

Pitch Horn

3
Flap Hinge

Hub

Figure 3.6: Cross-coupling due to the delta-three-angle

The 3-angle between the virtual hinge axis and the real flap hinge axis defines the
nature and magnitude of this coupling effect. It also introduces an aerodynamic spring,
and the effective natural frequency of the flap motion is again increased and influences
the flapping response.
31

Chapter 3

Helicopter Theory

Flap Motion due to Pitch and Roll Velocities


This effect becomes important to flying qualities since additional damping is added
due to the effect of roll or pitch velocities on the rotor tilt. Gusts change the attitude of
the helicopter and therefore the tilt of the rotor shaft. The change in main rotor tilt lags
behind by an amount proportional to pitch or roll rate and rotor moment of inertia due to
this additional damping term. The TPP wants to follow the shaft tilt due to a pitch or roll
rate of the body as long as there is a hinge that provides a moment from the shaft to the
rotor blades. Asymmetric flapping velocities over the azimuth with respect to the shaft
results in different aerodynamic forces and moments over the azimuth, and the TPP
follows the shaft. Furthermore, looking at pitch or roll rates it can be seen that the blade
angle-of-attack also changes. To maintain equilibrium, the blade compensates for this
with off-axis flapping. A pure pitch rate for example results in a change of lateral
flapping. It is obvious that this aspect introduces an additional cross-coupling term and
becomes important for evaluating flying and handling qualities.
Dihedral Effect
As in the airplane case, this effect is desirable since it helps the pilot to fly the
aircraft. For wind producing sideslip angles, the aircraft tends to roll away from the
approaching wind. For helicopters, the source for this positive dihedral effect is the blade
flapping. Since for zero side slip angle the blades over tail and nose experience no
velocity due to pure forward flight, a non-zero side slip angle results in additional
velocities for those blade positions over the azimuth. As a result of wind coming directly
from the right, for example, the blade over the nose will become the retreating blade, and
the blade over the tail the advancing blade (for a rotor spinning counterclockwise). The
rotor plane will flap down on the left. Therefore thrust is tilted to the left, and the
helicopter rolls away from the wind. Even so the advancing and retreating blades are
different for a rotor spinning clockwise, the dihedral effect results in a tilt of the rotor
plane to the same side. This dihedral effect is mainly responsible for the phugoid-like
response of the helicopter in forward flight.
Rotor-Body Interference
Due to induced velocity of the main rotor, the additional airflow over the surface of
the helicopter body causes drag counteracting the thrust created by the main rotor. For
some helicopter configurations with additional wings or other aerodynamic surfaces, this
effect becomes even more important. As Figure 3.7 illustrates, this aerodynamic aspect is

32

Chapter 3

Helicopter Theory

different for hover and forward flight [7]. Assuming that the shape of the rotor wake can
be described in such a simple manner, the influence of this wake in hover contributes
mainly to the horizontal force component. In forward flight this wake is diverted due to
the airspeed and also acts on the tail rotor, horizontal tail and vertical tail. The transition
from hover to forward flight is an intermediate condition partially affecting all
components and causing great difficulty. The exact shape of this rotor wake is
furthermore not known and its influence on body, wings or horizontal tail can only be
approximated if a simple model is desired.

(a)

(b)

Figure 3.7: Rotor-Fuselage interference in (a) hover and (b) forward flight

Main Rotor Control


Among different types of rotors, different types of rotor control schemes were
developed. The conventional way is through cyclic pitch changes of the individual
blades. A pilot stick input is transformed by means of links and actuators into a tilt of the
swashplate. The swashplate tilt occurs in the non-rotating and in the rotating frame and is
then transferred to the individual blades through further mechanical links. This can be
achieved directly over linkages and joints, or, as in the case of a so-called servo or control
rotor, via an additional smaller rotor mounted on top of the main rotor. In some literature
this smaller rotor is referred to as a fly bar, servo or Hiller rotor. The mechanical linkages
of the R-50 hub from swashplate to control and main rotor are shown in Figure 3.8. The
TPP of the control rotor then governs the lateral and longitudinal inputs of the main rotor.

33

Chapter 3

Helicopter Theory

Figure 3.8: Mechanical linkages of the control rotor for the R-50 experimental helicopter

One advantage is, that rotor systems of this kind prevent a feedback of rotor forces into
the control system of the non-rotating frame. Forces and therefore the loads on
swashplate actuators are minimized. Another even more important effect is the additional
stability produced by the control rotor if arranged properly. Additional damping is added
to the rotating system, which, as seen in Chapter 4, provides pitch and roll rate feedback
as well as translational velocity feedback to the main rotor. The dynamic effects of the
control rotor on stability are described in general. In Section 5.7 the influences of the
control rotor on the rigid body dynamics of the helicopter are evaluated and investigated
for the linearized equations of motion for the R-50 helicopter.
This chapter referred to individual and coupled effects of aerodynamic, dynamic and
kinematic features and their influence on blade motion or body dynamics. The overall
stability of a helicopter is discussed in the next chapter. Since, for simulation, the
response of the physical system is much more important than individual blade motion, the
background for stability and control analysis important for helicopter dynamics are
provided.

34

Chapter 4

Helicopter Stability and Control

Chapter 4
HELICOPTER STABILITY AND CONTROL
This work refers to a single rotor two bladed helicopter with a tail rotor. Therefore
any further explanations and analysis are related to only this kind of helicopter. To keep
the model more general, non-rotating surfaces generating additional lift through wings
and a horizontal tail are not excluded from the analysis. For the investigation of the
simulated model helicopter additional wings are not present. To provide better flight
characteristics in forward flight a simple horizontal tail was added. In the mathematical
model the general case of a single main rotor and tail rotor helicopter is treated.
This chapter will first describe how control is accomplished for this kind of helicopter
and summarizes how the control inputs can influence off-axis motion. Stability aspects of
the decoupled and coupled rigid body dynamics are then presented for hover and forward
flight. A final section on stability augmentation with a control rotor follows.

4.1

Helicopter Control

Direct control of the helicopter is obtained mainly by controlling moments. Except


the vertical force component of the main rotor thrust, which can be controlled directly, all
main rotor control inputs result in a tilt of the main rotor and produce a moment about the
aircraft CG. Commanded are therefore changes in pitch and roll angles, resulting in
lateral and longitudinal forces and finally in the desired translational helicopter motion.
Controlling particular moments also makes some compensating control inputs in other
axes necessary, since most inputs are coupled with off-axis motion as mentioned in the
previous chapter.
The pilot's controls consist of a collective stick to control the vertical force, a cyclic
stick to control longitudinal and lateral moments, pedals to control the yaw moment via
the tail rotor thrust and the throttle to adjust rotor speed. Collective stick is used to trim
the thrust of the main rotor for some desired forward flight condition and for height
control in hover. This input changes the collective pitch of all blades equally, so that only
the magnitude of thrust, not the orientation of the thrust vector, is influenced; similarly
for the pedal input, which provides torque balance and directional control from the tail
rotor. Only the collective blade pitch of the tail rotor is changed to control tail rotor thrust
and the resulting yaw moment due to the moment arm relative to the center of gravity. To
35

Chapter 4

Helicopter Stability and Control

vary thrust and forward speed, and to maintain the desired constant rotor speed, the
required rotor power changes and throttle must be adjusted. Since a speed governor on
the engine can manage this, the assumption of constant rotor speed over the entire flight
envelope in regards to helicopter performance, is valid. The engine dynamics are
assumed to be fast compared to the rigid-body dynamics and are therefore neglected. An
increase in throttle and engine power results in higher available power and does not
directly influence the helicopter stability. Cyclic pilot stick displacements are connected
to the blade pitch, such that the rotor tilts to the desired direction. In small manned
helicopters this can be done directly by mechanical linkages, for bigger helicopters
electro-hydraulic actuators can be used to convert rotor control inputs. In the case of the
remotely controlled model helicopter, purely electric actuators are mounted on the
aircraft. For full-sized helicopters it is important for the pilot to have a proper feedback of
control forces due to pitch moments of the blades to the pilot's stick to improve handling
qualities. A mechanical linkage automatically provides this feedback. If actuators are
used, an artificial-feel-unit can simulate these forces in the pilot's stick.
Response
Input
Longitudinal
Stick

Pitch
Pure (Prime)

Roll
1. Lateral flapping
due to longitudinal
stick
2. Lateral flapping
due to load factor
Pure (Prime)

Lateral Stick

1. Longitudinal flapping
due to lateral stick
2. Longitudinal flapping
due to roll rate

Pedals
(Rudder)

Negligible

1. Roll due to tail


rotor thrust
2. Roll due to side slip

Collective

1. Transient longitudinal
flapping with load
factor
2. Steady longitudinal
flapping due to climb
and descent in forward
flight caused by rotor
flapping
3. Pitch due to change in
horizontal tail lift

1. Transient lateral
flapping with load
factor
2. Steady lateral
flapping due to
climb and descent
3. Side slip induced
by power change
causes roll due to
dihedral effect

Yaw
Negligible

1. Undesired in
hover, caused by
directional
stability
2. Desired for turn
coordination and
heading control
in forward flight
Pure (Prime)

Power change
varies requirement
for tail rotor thrust

Climb or
Descent
Desired for
vertical flight
path control
in forward
flight
Descent with
bank angle at
fixed power

Undesired
due to power
changes in
hover
Pure (Prime)

Table 4.1: Single rotor helicopter coupling sources [15]

36

Chapter 4

Helicopter Stability and Control

To find the controls required in a trimmed flight condition, all forces and moments on
the helicopter must be zero. An iterative routine is necessary that varies the pilot inputs
until the six force and moment components are simultaneously zero. Predicting the
control inputs for a trim condition is difficult due to the complexity of the rotor dynamics.
For validation purposes it is therefore necessary to compare simulated results with flight
experiments. Since for some identification studies linear models about steady-state flight
conditions are computed, it is very important to verify that the system response about
these steady-state trim conditions correspond with the real system. The basic behavior of
rotor control from hover to forward flight is briefly described in the following.
Summarizing sources of inertial and aerodynamic coupling of longitudinal and lateral
helicopter motion, Table 4.1 describes various sources of cross-coupling.
In forward flight, a longitudinal cyclic input creates a lateral moment on the rotor disk
necessary to cancel the changes of blade-angle-of-attack due to flapping and to
compensate for the higher velocity of the advancing blade. Due to hinge offset and hinge
spring there is also a cross-coupling effect and the phase shift of input and TPP tilt is less
than 90. A lateral tilt of the TPP and a resulting lateral moment for longitudinal cyclic
input proportional to the natural frequency of the rotating system develop and partially
compensate for the effect of the higher velocity on the advancing blade. To maintain
forward flight a longitudinal cyclic input (cyclic forward stick) is required to change the
TPP tilt as speed increases. A lateral cyclic input (cyclic left/right stick, dependent on
direction of rotor rotation) is required to compensate for the lateral TPP tilt due to lateral
flapping.

4.2

Helicopter Stability

In terms of dynamic stability and response to control inputs, the rigid body degrees of
freedom are mainly involved in the flight dynamic analysis. Separate longitudinal and
lateral motion can usually be assumed to simplify analysis and to observe the most
important stability characteristics. A further simplification is to use only low frequency
dynamics of the main rotor. No additional degrees of freedom are therefore added to the
system. In fact, the low frequency model for the main rotor response is a very good
approximation even for a more complex analysis. Later it will be shown how the dynamic
characteristics of the rigid body are influenced by an additional control rotor. In contrast
to the main rotor dynamics, these rotor dynamics are then coupled with the rigid body
dynamics. Also the coupling of longitudinal and lateral rigid body motion can be
considerable and important for handling qualities. The use of a fully coupled simulation

37

Chapter 4

Helicopter Stability and Control

model for the experimental helicopter characteristics will allow a more realistic
representation of the dynamics.
In general, a helicopter shows different characteristics in hover and forward flight.
The most important stability characteristics will now be investigated for decoupled
longitudinal and lateral dynamics to show the basic features of helicopter motion. If
principal aircraft axes are assumed, inertial cross-coupling of yaw and roll can be
neglected. Furthermore yaw motion is assumed to be fully decoupled from the other
degrees of freedom.
The following analysis is a summary of the basic helicopter motion for the flight
conditions hover and forward flight and explains the typical helicopter behavior. More
details can be found in [6, 8].
4.2.1

Hover

Vertical force equilibrium is given by the equation of motion for the helicopter
vertical velocity z& E . Collective pitch control is directly related to main rotor thrust, and
for now it is assumed that there is no pitch-flap-coupling. The resulting first-order
differential equation describing vertical dynamics has only a single pole. The time
constant increases with rotor speed, blade loading and gross weight. This root is in
general small, justifying the low frequency rotor response with respect to vertical motion.
Rotor speed will always be assumed to remain constant to simplify analysis. A variable
rotor speed would add another degree of freedom and modeling height control would
become more difficult.
For the yaw motion, only moments due to main rotor torque and tail rotor thrust will
be considered. Perturbations due to side velocity will be included. The low frequency
response of the tail rotor thrust leads to a first-order differential equation for the yaw rate.
The time constant is approximately the same as the time constant of vertical motion.
Since sideward velocity changes with a change in tail rotor thrust, the lateral translation
and yaw motion in general are coupled. This coupling is small compared to other
coupling effects. A change in lateral cyclic causes a sideward velocity and requires small
pedal input to maintain heading. For constant rotor speed, a change in thrust will also
vary the main rotor torque, and therefore couple vertical and yaw control. To maintain the
heading while thrust is changed, a coordinated pedal input is necessary. The tail rotor is
operating in adverse aerodynamic environment due to the main rotor wake, fuselage and
vertical tail. Modeling all these aspects is very complex and can often be simplified by an
approximation only. However, these effects will become important in forward flight since

38

Chapter 4

Helicopter Stability and Control

yaw damping and directional control are greatly influenced. Because of only low yaw
damping in hover, a helicopter is very sensitive to tail rotor thrust changes. Most real
helicopters therefore require at least yaw rate feedback to show handling qualities that
allow a reasonable control over the helicopter.
Longitudinal dynamics consist of the pitch motion, longitudinal velocity and vertical
velocity. Corresponding longitudinal inputs are longitudinal cyclic stick, longitudinal
gust velocity and collective pitch. The characteristic equation has three solutions
representing the open loop poles of the longitudinal dynamics. One is a stable root on the
real axis, the other two are a mildly unstable complex pair. This instability is a result of
the coupling between pitch moment due to longitudinal velocity and the longitudinal
component of the gravitational force due to pitch. The stable real root is mainly due to
main rotor pitch damping. It can be found that the frequency of the roots is small
compared to the rotor speed, justifying the assumption of only important low frequency
response. The real root is about as fast as the vertical root mentioned previously. Since
period and time to double amplitude of longitudinal dynamics are large, the motion in
hover is still controllable by the pilot. However, it requires the pilot to know about the
influence of cross-coupling and compensation in order to stabilize the unstable mode.
As described in detail in [8] several feedback loops can be introduced to improve
stability characteristics in hover. One possibility is a lagged feedback of the pitch rate,
introducing damping to the system. This kind of feedback can be provided by a
mechanical feedback system. Since this method is commonly used, it will be described
and examined in detail in Section 5.7. The additional lagged pole of this feedback
subsystem should be to the left of the pitch root to satisfy the intention of damping and
improving stability. Considering short and long time response to cyclic control and
longitudinal wind velocity, it can be found that the primary response is pitch acceleration.
Hence the longitudinal response to a gust is small. Because of low damping, the response
to cyclic control is usually large for roll and pitch motion, depending on inertia and
position of CG relative to the rotor hub.

39

Chapter 4

Helicopter Stability and Control

Im s
1.15
=1.0

1.15

1.1

1.05

=1.0

Re s

Figure 4.1: Longitudinal hover poles dependent on normalized flap frequency =n /

The natural frequency of the main rotor also influences the rigid body dynamics, such
that additional cross-coupling is introduced for any natural frequency different from the
rotor rotational speed. An additional spring or damper in the flap hinge and hinge offset
from the axis of rotor rotation changes this natural frequency and results in off-axis
motion due to purely longitudinal inputs. The importance of the natural frequency is
shown in Figure 4.1, where the pole position of the longitudinal motion is shown as a
function of the flap frequency normalized with respect to the main rotor rotational speed.
The lateral dynamics contain lateral velocity and roll angle, lateral cyclic control and
lateral wind velocity. The basic physical system of lateral and longitudinal motion is
similar, except that the roll moment of inertia is much smaller than the pitch moment of
inertia. This increases the magnitude of roll stability derivatives relative to pitch
derivatives. Poles can again be found as one real stable pole due to roll damping and an
unstable complex conjugate pole pair due to the rotor dihedral effect and speed stability.
Damping moments in roll and pitch are similar in hover. However, the roll inertia is
smaller than the pitch inertia, and the lateral mode has therefore a higher frequency than
the longitudinal mode. The smaller roll inertia results in a shorter period and less
damping of lateral modes which makes it more difficult for the pilot to control the lateral
motion in hover. As for the longitudinal dynamics, rate and attitude feedback would be
required to stabilize the system. A control rotor can again provide the mechanical rate
feedback of the roll rate and improve stability characteristics. Figure 4.2 shows a typical
plot of decoupled hover poles for longitudinal and lateral motion of a single main rotor
helicopter.

40

Chapter 4

Helicopter Stability and Control

Im s
0.04

long. poles
lat. poles

Re s
-0.1

-0.05

0.025

-0.04

Figure 4.2: Typical hover poles for decoupled longitudinal and lateral motion

Even for a natural frequency equal to the rotor rotational speed, there is a crosscoupling of lateral and longitudinal dynamics due to the moments of inertia. For small
enough roll inertia, one of the oscillatory modes might be stabilized, and the other might
be destabilized as shown in Figure 4.3.

Im s
0.04

long. poles
lat. poles

Re s
-0.1

-0.05

0.025

-0.04

Figure 4.3: Typical hover poles for coupled longitudinal and lateral motion

Even if the roots are not significantly influenced, roll and pitch motions are coupled, and
the dynamic behavior compared to the uncoupled dynamics changes. Hence the coupled
system should be considered if handling qualities and dynamic response are the subject of
the investigation. Later in Section 5.7 the poles are examined for the R-50 experimental
helicopter. Most helicopters typically show similar characteristics to those just described.

41

Chapter 4

Helicopter Stability and Control

However, the very different mass inertia of the model helicopter will result in differences
compared to a full-size helicopter.
With an increase of forward speed the dynamics of the helicopter change. Additional
aerodynamic forces and moments are acting on the helicopter due to the higher body
velocity. The most important aspects are discussed in the following section.
4.2.2

Forward Flight

Centrifugal forces due to angular body velocities, aerodynamic forces on fuselage and
tail, and rotor forces proportional to forward speed arise in forward flight and
significantly influence handling qualities. Rotor forces and body accelerations are now an
additional source for the coupling of longitudinal and lateral dynamics. In this section, it
will be assumed that lateral velocity, yaw rate and roll attitude can be analyzed separately
from longitudinal, vertical velocity and pitch rate. For the final analysis of the simulation
in Chapter 7 all six coupled DOFs are examined. To investigate the dynamics of the
helicopter in forward flight the rigid body equations are linearized about some trimmed
flight condition. Zero angular rates and constant forward speed are assumed. The
differential equations of motion perturbed from the given trim condition allow the
computation of stability derivatives, which are the elements of the system and control
matrices F and G in Appendix C.
The forces acting on the helicopter are aerodynamic forces on the fuselage, horizontal
and vertical tail, horizontal stabilizer and wings as present in some helicopter
configurations. To demonstrate basic characteristics of helicopter handling qualities, it is
sufficient to assume that the tail rotor thrust produces a yaw moment only. It can be
found that due to a pitch rate a vertical acceleration evolves, and that a lateral
acceleration is generated due to the yaw rate. These forces introduce coupling of
longitudinal, lateral and vertical dynamics. However, inertial cross-coupling of roll and
yaw will be neglected (Ixz = 0). From the low frequency response of the system, rotor
forces and moments can be found. An important stability is the pitching moment due to
angle-of-attack changes of the helicopter. A downward vertical velocity of the helicopter
results in an increase of the rotor blade angle-of-attack and finally in a nose up pitching
moment. Increasing angle-of-attack produces a pitch up moment, further increasing
angle-of-attack due to rotor dynamics and aerodynamics. The main rotor is therefore a
source of an angle-of-attack instability for the helicopter in forward flight.
Aerodynamic effects of non-rotating aerodynamic surfaces like wings, vertical and
horizontal tail and fuselage are similar to fixed wing aircraft. Therefore those stability

42

Chapter 4

Helicopter Stability and Control

and control derivatives are similar. The fuselage drag produces a damping force, and all
influences of the airframe are proportional to the square of forward speed. Damping
derivatives with respect to translational velocities are thus increase with an increase in
flight speed. Another contribution to instability arises from the helicopter fuselage angleof-attack changes, resulting in an unstable pitch and yaw moment. The horizontal tail
adds a pitch moment due to angle-of-attack changes and counteracts the destabilizing
effect of the main rotor. It also adds pitch damping due to the helicopter pitch rate. Lift of
the vertical tail resulting in a side force counteracts the destabilizing yaw moment of the
fuselage due to side velocity. The major effect of the vertical tail is yaw damping and
directional stability as for a fixed wing aircraft. Additional drag from a horizontal and
vertical tail further increases speed stability.
The complex interactions of main rotor, tail rotor, fuselage, vertical and horizontal
tail often prevent a simple analysis. In forward flight, the aerodynamic characteristics of
the tail and the tail rotor are strongly influenced by the shape of the fuselage and the rotor
wake produced by the main rotor. Experimental data from wind tunnel testing or flight
tests can be used to model these aerodynamic effects, and good results can be achieved
with this approach.
The uncoupled longitudinal dynamics consisting of longitudinal velocity, pitch
attitude and vertical velocity are controlled by collective and longitudinal cyclic stick
inputs. Vertical and longitudinal gusts are included as well. The characteristic equation
shows two negative real poles for vertical and pitch motion and a slightly unstable long
period oscillation, similar to those obtained for hover. With increasing speed the
longitudinal modes vary as illustrated in the following Figure 4.4. Shown are the roots
with and without horizontal tail to point out its importance for the angle-of-attack
stability. Damping becomes less without horizontal tail for increasing speed. Including
the horizontal tail usually moves the long period hover mode into the left half-plane, and
pitch and vertical real roots are transformed into a short period highly damped oscillatory
mode. Since the tail effectiveness is reduced at lower speed due to dominating rotor and
fuselage wakes, this result may not always be true. But the basic tendency can always be
found for most helicopter analysis, and handling qualities are improved significantly. The
primary response of the helicopter to control inputs and gusts is again vertical and pitch
accelerations with minimal longitudinal acceleration. Controlling the longitudinal motion
is therefore, as in hover, indirect and can be observed only in the long-term response. For
further simplified analysis, short and long term response could be treated separately as
done in [8].

43

Chapter 4

forward flight roots


with horizontal tail

Helicopter Stability and Control

Im s
hover roots

Re s

forward flight roots


without horizontal tail

Figure 4.4: Influence of forward speed and horizontal tail on longitudinal poles

In general, pitch control sensitivity is high, and a longitudinal command results in a


reasonable normal acceleration. Nevertheless, there is a delay after applying cyclic
control input until normal acceleration develops. It can cause difficulties in helicopter
control if this lag is too large. To evaluate handling qualities with respect to this
characteristic, the time until maximum acceleration is achieved can be used. Common
specifications define the desired response due to step inputs and are given in [8]. The goal
is to ensure satisfying acceleration and handling qualities during flight maneuvers.
Important influence of pitch and roll damping can again be provided by rate feedback
with a stabilizer bar or a control rotor as already mentioned earlier. Longitudinal flying
qualities are much improved, since steady-state angular rates develop faster, and a faster
response of normal acceleration due to control input is achieved.
Lateral dynamics of forward flight consist of side velocity, yaw rate and roll attitude,
including side gust velocity. Yaw velocity of the body axes results in a lateral
acceleration. Main rotor side forces and moments due to a yaw rate are small. Likewise
small are torque changes due to power changes compared to the yaw moment produced
by the tail rotor thrust. Therefore, the coupling of yaw and lateral dynamics is mainly due
to the yaw moment produced by lateral velocity and the lateral acceleration produced by
the yaw rate. Directional stability is governed by the influence of side velocity on the tail

44

Chapter 4

Helicopter Stability and Control

rotor thrust. As in hover, the lateral dynamics result in a real stable yaw root, a real roll
root, and an often slightly unstable lateral oscillation. For increasing forward speed,
directional stability is increased, and the inertial coupling transforms the oscillatory mode
in hover into a stable, short period oscillation. A typical root locus plot for the lateral
dynamics is shown in Figure 4.5 for increasing forward speed.

Im s
uncoupled roll roots
forward flight roots
hover roots

Re s

Figure 4.5: Influence of forward speed on lateral poles

Both real roots move to the right; the yaw mode might even approach the origin for a
further increase in forward speed. In contrast to hover the lateral velocity produces a
force opposing the motion. Lateral cyclic input commands mainly a roll rate with a small
lag due to inertial forces; lateral velocity builds up again slowly. As for the longitudinal
motion the lateral velocity can only be influenced indirectly through a commanded roll
rate.
To include all aerodynamic effects due to fuselage, rotor and tail interference, the
fully coupled dynamics should be considered. All the influences should be modeled and
included again in the force and moment calculations. Depending on the application, some
of these effects may be neglected.
45

Chapter 4

4.2.3

Helicopter Stability and Control

Stability Augmentation with a Control Rotor

As mentioned in the previous chapters, handling qualities of a helicopter can be much


improved by using an automatic control system like a control rotor. An additional gyro
sensing the inertial forces is introduced which acts on the main rotor. In the case of a
control or servo rotor this control system is fully mechanical. The control rotor is
attached to the rotor shaft on top of the main rotor, with its blades parallel to the main
rotor plane with a 90 phase shift. It rotates therefore with the same angular velocity
about the shaft axis as the main rotor.
As for the main rotor, the motion of the smaller control rotor can be described by a
lateral and longitudinal tilt of the rotor plane relative to the non-rotating hub frame. When
roll and pitch inertia damping is added, this system acts as a mechanical damper as long
as no airfoil or an additional blade section is attached to the control rotor. In Section 5.4
the equations for the control rotor with aerodynamic blade sections are given, and the
similarity with the main rotor is used to derive the equations of motion. A direct
mechanical link connects control rotor and main rotor blades. The control rotor plane tilt
is transferred to the main rotor blades as cyclic blade pitch. Blade-feathering moments are
also transferred back from the main rotor blades to the control rotor. If cyclic control
inputs are fed through the control rotor to the main rotor blades, the control system
provides lagged rate feedback of helicopter pitch and roll motion. This greatly improves
handling qualities in hover and forward flight. The damping for roll and pitch axes is
equal for this mechanical system.
The control rotor with two rods and small airfoils on each end as designed for the R50 Helicopter is shown in Figure 4.6. A rotor of this kind is often called a Hiller control
rotor. More details on the mechanical linkage can be found in Figure 3.8 and [29]. The
pilot controls cyclic pitch of the control rotor airfoils, producing an aerodynamic moment
on the gyro resulting in a tilt of the control rotor plane. Some control rotor designs might
also provide a direct input from the swashplate to the main rotor blades. Both inputs of
control rotor tilt and direct pilot input are then mechanically mixed inputs to the main
rotor. An additional effect of the Hiller control rotor important for full size helicopters is
that control forces are less sensitive to the main rotor motion itself. Feedback of control
forces into the pilot sticks or into the actuators can be reduced. This might not be always
desirable since the pilot likes to feel the necessary control forces to keep the aircraft in a
specific flight condition. This force can of course be provided by an artificial-feel-unit,
resulting in a higher complexity of the control system. The control rotor shown in Figure
4.7 is a typical control rotor as used for two-bladed, single-rotor helicopters. The

46

Chapter 4

Helicopter Stability and Control

mechanical linkage can be fairly complex depending on the desired portion of input
going through the control rotor or directly to the main rotor. The system used for the R50 experimental helicopter is described in Chapter 5.4 based on measured geometry of
the linkage and rotor data available.

Figure 4.6: R-50 control rotor providing rate feedback [29]

Some assumptions are made to simplify modeling and assure real-time capability.
The mathematical models for the helicopter simulation will be described in the following
chapter for the single components, and the basic dynamic features will then be examined
for a linearized helicopter system.

47

Chapter 5

Mathematical Modeling

Chapter 5
MATHEMATICAL MODELING
This chapter describes the mathematical helicopter model that is used in the current
version of the real-time flight simulator. It is based on a Minimum-Complexity
Helicopter Simulation Math Model developed by NASA [1]. Additional components like
control rotor, actuator models and sensor models are added to increase the fidelity of the
simulation.

5.1

General Helicopter Model

When the Minimum-Complexity Helicopter Simulation Math Model was developed


by NASA in 1988, aspects like computational delays, costs and inflexibility of commonly
used complex simulation models made a real-time simulation not very efficient with
respect to better engineering understanding of specific handling quality features. In
addition, a detailed knowledge of a specific helicopter was and still is not always given.
This fact further complicates the implementation of any helicopter of interest into a realtime simulation process. The goal was to develop a build-up of individual vehicle
components described by equations addressing the features associated with those
components. Only basic data sources such as flight manuals or system component
specifications should be necessary to achieve satisfactory results. The model that was
used throughout the entire flight envelope was then further investigated and validated for
some specific helicopter systems. Refinements and extended modeling was necessary to
obtain satisfactory results for specific helicopters. Important features of the desired math
model are:

First-order flapping dynamics for main rotor (coupled or uncoupled)


Main rotor induced velocity computation
Rigid body degrees of freedom
Realistic power requirements over the desired flight envelope
Rearward and sideward flight without computational singularities
Hover dynamic modes (longitudinal and lateral hover cubic, rotor-body coupling
with flapping)
Dihedral effect
48

Chapter 5

Mathematical Modeling

Correct transition from hover to forward flight


Potential for rotor RPM variation

More detailed aspects on some of the components used and modified are given in the
following chapter. A model for engine RPM variation is about to be developed at Georgia
Tech but not yet included in the simulation software.

5.2

Rigid Body Model

The six rigid body DOF model was already introduced in Chapter2.1. For manned or
unmanned simulation of flight, this model is accurate in describing system response for
flight mechanical and controlling purposes. Therefore the basis for the helicopter model
are the equations of unsteady motion of the rigid body as already mentioned and is
derived in [4]. The equations were rederived in the body axes system to avoid
singularities due to undefined values for angle-of-attack and side-slip-angle in hover
condition. These equations are listed in Appendix B. The equations are in general valid
over the entire flight envelope for a typical aircraft. For rotorcraft gyroscopic effects due
to main and tail rotor become important, and the force and moment components for a
two-bladed main rotor, a tail rotor and a control rotor are added. The assumption of the
xB-zB- plane as a plane of symmetry is also made.
The equations will be given for a counterclockwise spinning main and tail rotor
(viewed from above and left respectively). A change of directional rotation will affect
some of the derived expressions. Introducing a directional variable, , the direction of
rotation can be chosen arbitrarily, dependant on the system that needs to be simulated.
The use of this parameter is based on a tail rotor in pull configuration. This implies that
the vertical tail is always mounted in the downstream direction of the tail rotor. Since
rotating systems add dynamics to the rigid body equations, the two most important
subsystems, main and control rotor, are described in the next section in more detail.

5.3

Main Rotor Model

Responsible for the unique and complex characteristics of a helicopter, the main rotor
model described in [1] is modified in order to fit the very different dynamics of the model
helicopter it is applied to and still guarantee real-time capability. For the final
implementation in the simulator the nonlinear dynamic equations for the main rotor are
substituted by its steady-state representation. This is justified for several reasons, as
described later in this chapter.

49

Chapter 5

Mathematical Modeling

The basis for the model is taken from [13], extended in [14] and simplified in [2, 3] to
represent the desired first-order flapping model of the main rotor dynamics. Higher-order
flapping dynamics were not considered. This first-order model results in a TPP lag
affecting the main rotor thrust vector orientation and therefore the immediate response to
control inputs. Uniform inflow distribution is first assumed to compute thrust and
induced velocity. Moment equations are rederived from [15, 16] in the body-axis system.
Based on classical momentum theory the equations to compute thrust and induced
velocity can be found to be
T = (wblade v i )

R 2 aBc
4

(5.1)

and
v 2
v =
2
2
i

T
v 2

+
,
2
2 A

(5.2)

where

wblade = wr +

and

2
3

R coll + twist ,
3
4

(5.3)

w r = w + (a1s + is ) u b1s v ,

(5.4)

v 2 = u 2 + v 2 + w 2 (wr 2 v i ) ,

(5.5)

A = R2 .

(5.6)

Since all angles are small, it is assumed that sin() and cos()1. In Equation 5.3 the
collective input coll to the main rotor blades occurs, influencing directly main rotor
thrust. Most blade designs are based on a fixed change of pitch angle over the blade
radius, the so-called blade twist twist. Assuming a linear twist from the root to the blade
tip is mostly sufficient for this kind of analysis. The total velocity of the main rotor blade
relative to the air mass is defined by w blade, w r is velocity of the main rotor disk relative to
air due to body velocities ub, vb and wb. The velocities need to be transformed with the
TPP tilt angles a1s , b1s and the shaft tilt angle is from the body frame to the TPP. Induced
velocity and thrust can be computed for a given thrust with only a few iterations using
Equations 5.1-5.5 until convergence is reached.

50

Chapter 5

Mathematical Modeling

The dihedral effect of the main rotor describing the change of TPP tilt due to lateral
and longitudinal velocity is dependent on thrust and induced velocity respectively.
The equation for the lateral tilt derivative can be found in [6] to be

db1s
CT
2 8 CT
,
=

+
dv
R a
2

(5.7)

where = Bc/R is the rotor solidity. Assuming equal effects on lateral and longitudinal
motion, the longitudinal derivative is
da1
db
= 1 .
du b
dvb

(5.8)

The assumption of constant derivatives in hover is true as long as uniform inflow is


assumed. For fast forward flight non-uniform inflow becomes necessary and this
assumption is only an approximation. From experimental results on full size helicopters a
more accurate estimation is given by introducing an empirical variable, fwake , considering
low or high-speed effects on those derivatives, as described in [1]. This variable will be
used in the dynamic rotor equations.
With respect to cross-coupling of longitudinal and lateral TPP tilt two already
mentioned components need to be considered. Mechanical linkages from swashplate and
control rotor to the main rotor cause an additional flap angle due to a commanded blade
pitch change. The geometry of this linkage and the rotor hub design result in an off-axis
motion. Defined by the angle between flap hinge axis and an imaginary line drawn from
the hinge to the pitch horn of the main rotor blade, as shown in Figure 3.6, the so called
delta-3-angle, a cross-coupling coefficient is obtained in the form
K 1 = tan 3 .

(5.9)

This coefficient also describes the mechanical feedback from flap angle to blade pitch,
automatically decreasing pitch for an increasing flap angle.
The second effect of cross-coupling can be derived from the fact that a flapping hinge
with an offset from the axis of rotation also changes the natural frequency of the dynamic
system as mentioned in Chapter 3.3.2. Since the R-50 helicopter is designed with such a

51

Chapter 5

Mathematical Modeling

hinge offset, this additional effect might become important and therefore needs to be
considered in the flapping equations.
With the ratio computed by the hinge offset eMR relative to the main rotor radius RMR, this
coefficient can be written as
K2 =

3 e MR MR

,
4 R MR f

(5.10)

where MR is the rotor rotational speed. The coefficient


f =

MR
16

8 e
1 + MR
3 R MR

(5.11)

describes the change of natural frequency due to the hinge offset. This coefficient is also
a function of the so-called Lock number, expressing the ratio of aerodynamic and blade
inertial forces. With the assumptions of a blade section two-dimensional lift curve slope
(a6.0), approximately constant over the entire blade length, and with a constant chord
length c, this non-dimensional Lock number for the main rotor can be written as
a c R 4
.
MR =
Ib

MR

(5.12)

The blade chord length, usually also a function of the radius, was found to be constant
except at the blade root, where a profile modification avoids an excessive aerodynamic
loss due to reverse flow. For a given blade inertia Ib relative to the flap hinge, the total
cross-coupling coefficient Kc can then simply be written as the sum of the previously
computed coefficients
K c = K1 + K 2 .

(5.13)

In the final analysis of the helicopter dynamic response, this coefficient might be
adjusted to fit experimental data, to correct for unmodeled spring-damper dynamics in the
hinge or to compensate for further unmodeled cross-coupling effects not considered in
the given approximation.

52

Chapter 5

Mathematical Modeling

Due to the interaction of the rotor wake with the fuselage especially at low speed, the
dihedral effect will become more effective. This will directly influence the cyclic
flapping due to changes in forward or sideward velocity. The empirical value for fwake
dependants on flight speed and can be chosen based on experience or experimental data,
if available and necessary. For the R-50 flying at low speed, this variable will always be
equal to 1.
The final equations for the TPP tilt angles relative to the HP can then be written in the
form
da
a1s = a1 B1, MR + K c b1s + 1s u (1 + 2 f wake )
(5.14)
du
db
and
b1s = b1 + A1, MR K c a1s + 1s v (1 + f wake ) ,
(5.15)
dv
with b1s and a1s as the lateral and longitudinal TPP tilt relative to the HP, the TPP tilt
angles with respect to the swashplate are defined by b1 (lateral) and a1 (longitudinal). The
influence of the cross-coupling is dependent on the direction of rotor rotation. Therefore
the directional parameter needs to be considered in those terms. The inputs to the blade
pitch angle are given by A1,MR and B1,MR. These are the resulting blade pitch changes
commanded by the pilot stick inputs, transmitted by the swashplate tilt and the
mechanical linkage from swashplate to the pitch horn. The cross-coupling in Equations
5.14 and 5.15 influences the TPP tilt and depends on the direction of rotor rotation. The
lateral and longitudinal swashplate tilt A1,SP and B1,SP commanded by the pilot is
distributed by mechanical linkages to the control rotor as well as directly to the main
rotor. The largest portion is fed through the control rotor. The resulting tilt of the control
rotor plane due to control rotor blade pitch changes, commands the main rotor blade pitch
through mechanical linkages. The resulting main rotor input can then be defined as

and

A1, MR = k MR A1, SP + k s ,CR

(5.16)

B1, MR = k MR B1, SP k c ,CR ,

(5.17)

independent on the direction of rotor rotation. The coefficients kSP and k prescribe the
amount of commanded swashplate tilt and resulting control rotor plane tilt finally acting
on the main rotor blade pitch. The additional states s,CR and c,CR describe the lateral
longitudinal TPP tilt of the control rotor with respect to the HP as shown in Chapter 5.4.
The equations of motion for the control rotor are given in the next chapter, the

53

Chapter 5

Mathematical Modeling

coefficients kMR and k are found by measuring linkage lengths and angles of pitch
changes due to a swashplate tilt.
The collective input commanded by the pilot is fed directly to the main rotor and
therefore does not appear explicitly in the main rotor flapping equations. It is assumed
that the collective pitch directly influences the induced velocity through the rotor disk
and therefore thrust. For an equally over the rotor azimuth distributed change of blade
pitch due to collective stick input and resulting swashplate displacement, it can also be
assumed that effects on lateral and longitudinal TPP tilt are small and will be neglected.
The induced velocity and thrust iteration uses the collective input while computing
induced velocity due to collective pitch and linear blade twist in Equation 5.3.
The final equations of main rotor flapping motion in the rotating rotor frame relative
to the swashplate can be written in the simplified form

and

a&1 = in a1 off b1 q

(5.18)

b&1 = in b1 + off a1 p .

(5.19)

The simplified flap rate coefficients in and off for in- and off-axis changes of the flap
angle can be computed with no flap-cross-coupling,

off = 0

and

in = f ,

(5.20)

or including the flap-cross-coupling,


off =

MR

1 + MR
f

and

in =

MR
off .
f

(5.21)

Equations 5.21 give an approximation for the effect of cross-coupling in the flap rates
including cross-coupling in the flap angles. In [1] it is also mentioned, that a better fit to
experimental data might be obtained for a neglected rate-cross-coupling in the simulation.
With respect to the hinge offset one more aspect should be considered. A pilots cyclic
input causes a change in flapping and finally TPP tilt angles, also causing an
aerodynamic blade moment acting on the helicopter only for a nonzero hinge offset. For
no hinge offset the blades are free to flap and no moments are transferred to the rotor hub.

54

Chapter 5

Mathematical Modeling

The only rotor moments acting on a helicopter with no hinge offset (eMR/RMR=0) are
generated by rotor forces and the moment arm given by the hub distance to the helicopter
CG. An approximation for the in-axis blade moment due to cyclic inputs with a hinge
offset is given in [6] by Equation 5.22. Equation 5.23 represents the aerodynamic and
inertial flap-cross-coupling moment including a hinge offset. The computed moments can
be written in the body-fixed frame as
dM
dL

1 e
2
2
=
= B a c R MR
VTip
MR ,
dB1 dA1
2
6 RMR

and

2
( RMR ) 2 a eMR
3 B c RMR
dM
dL
=
=

4
RMR
da1s db1s

(5.22)

. (5.23)

In the previous chapter it has already been mentioned, that the main rotor dynamics
are usually very fast compared to the rigid body dynamics of the helicopter. Therefore a
so-called quasi-steady-state model of the main rotor can be used to describe the dynamics
of the body-rotor dynamics. This approach is also very useful since additional dynamics
result in additional states that need to be integrated. This should be avoided to maintain
real-time capability of the code. The DOFs to be added are longitudinal and lateral tilt
angles of the main rotor TPP and of the control rotor TPP. This would result in a model
extension from 6 to 10 DOFs. Investigating the influence of control rotor in Chapter 5.7,
it is found that the control rotor dynamic modes are about as fast as the helicopter rigid
body modes. The control rotor dynamic modes become coupled with rigid body modes,
and a steady-state model for the control rotor is not sufficient to represent the real system
dynamics.
To avoid the 10 DOF model the steady-state model for the main rotor will be used to
represent the rotating subsystem dynamics. Characteristics with respect to flight
dynamics and control aspects remain accurate for this approximation. Nevertheless the
two control rotor states are added to the equations of rigid body motion. This model is
described in Chapter 5.4 and in [10] in more detail. The control rotor is treated like a
smaller rotor with only a limited blade profile. The following section gives the quasisteady-state equations for the main rotor derived from the previous dynamic equations.

55

Chapter 5

Mathematical Modeling

Quasi-Steady-State Equations for Main Rotor Dynamics


For the valid steady-state blade motion it is assumed that the desired orientation of the
TPP is reached instantaneously. In fact, the transient flap motion is in general a highly
damped oscillation with an approximate time constant 16/ (see [6]). The resulting
time-to-half amplitude, t1/2=0.693, typically corresponds to 90 of the rotor azimuth,
and the transient response dies out after less than one revolution of the main rotor.
Steady-state equations are obtained by the necessary conditions for steady-state

a&1 = 0, b&1 = 0 .

(5.24)

Solving Equations 5.14 and 5.15 for a1 and b1 and substituting in Equations 5.18 and
5.19, the steady-state equations for the main rotor TPP angles relative to the HP are

a1s =

and

1
( T1 ( q + in T3 off T4 ) + T2 ( p in T4 off T3 ))
T + T22
2
1

b1s =

1
( p in T4 off T3 a1s T2 ) .
T1

(5.25)

(5.26)

The terms T1, T2, T3 and T4 represent abbreviations given in Equations 5.27.

T1 = ( in off K c )
T2 = ( in K c + off

(5.27)

da

T3 = B1,MR 1s u (1 + 2 f wake )
du

db

T4 = A1, MR + 1s v (1 + f wake )
dv

The final Equations 5.25 and 5.26 are then used to compute the TPP tilt angles a1s and b1s
relative to the main rotor HP.
The assumption of a thrust vector perpendicular to the TPP is used to transform the
thrust components into the body-fixed frame by

56

Chapter 5

and

Mathematical Modeling

X MR ,b = TMR sin (a1s + is ) ,

(5.28)

YMR ,b = TMR sin (b1s )

(5.29)

Z MR,b = TMR cos (a1s + is ) cos (b1s ) .

(5.30)

For a given flight condition, the main rotor power calculations need to include
induced power Pi, profile power Ppr due to friction between air and blade surface, climb
power Pc and parasite power Ppa to overcome fuselage drag. The final equations are given
in Equations 5.31.
Ppr =

c D 0 b c R MR
2

R MR (R MR ) + 4.6 u 2 + v 2
2
4

))

(5.31)

Pi = TMR vi

(5.32)

Pc = m g z& E

(5.33)

Ppa = ( X Fus u ) (YFus v ) (Z Fus (w v i ))

(5.34)

Equation 5.31 represents a commonly used approximation for profile power, where the
effective frontal area of the main rotor producing the aerodynamic drag is given in [6] by
cD0bcRMR. The representative profile drag coefficient of the blade is defined by cD0,
taken from airfoil data for a particular rotor blade profile. Force components XFus, YFus
and ZFus are computed in the body-fixed frame.
The main rotor torque can then be computed by
PMR
,
MR

(5.35)

PMR = Ppr + Pi + Ppa + Pc .

(5.36)

M T ,MR =

where

Main rotor moments have to be transformed into the body-fixed frame, and the final
equations are
57

Chapter 5

LMR ,b = YMR hhub +

dL
dL
b1s +
(a1s + A1, MR k1 b1s ) ,
db1
dA1

M MR ,b = Z MR d hub X MR hhub +

and

Mathematical Modeling

N MR ,b = M T , MR .

dM
dM
a1s +
( b1s + B1 k1 a1s )
da1
dB1,MR

(5.37)

(5.38)

(5.39)

The vertical and horizontal distance from hub to helicopter CG represent the moment
arms hhub and dhub, measured in the body-axis frame along the zB- and xB-axis. It is
assumed that there is a negligible offset of the hub in direction of the yB-axis. Since for
dynamic characteristics the most important components are the main and control rotor,
only those two subsystems are described in more detail. Further components like
fuselage, tail rotor, horizontal tail and wings will be briefly mentioned, but are described
in more detail in [1, 8].
Steady-state equations for the main rotor were assumed to be sufficient in the
previous chapter. In the following chapter the dynamic control rotor equations of motion
are given, which will strongly affect the helicopter dynamics. Some results comparing the
dynamic characteristics of the helicopter with and without control rotor will be shown in
Chapter 5.7.

5.4

Control Rotor Model

The importance of the control rotor due to damping added to the system and its
function as a rate feedback system were already discussed. In order to model the
helicopter more accurate it is necessary to investigate the control rotor in more detail.
Therefore this rotating system is treated as a smaller rotor with similar DOFs. The kind
of control rotor used at the R-50 (Figure 4.7) is called a teetering rotor. More details on
those rotor systems can be found in [6, 8, 15]. Perhinschi investigated in [10] the
influence of a control rotor on the linear system dynamics of a model helicopter as part of
a controller design. The basic equations of motion are taken from main rotor flapping
equations with the similar assumptions made previously. Additionally, since a teetering
rotor is modeled, the coning angle is negligible. With only small blades at the rod ends,
the aerodynamic forces are small compared to the inertial forces. This will result in a very
small Lock number for the control rotor.

58

Chapter 5

Mathematical Modeling

Writing the equation of basic blade motion for a clockwise rotating system,

CR = c ,CR cos b, CR s ,CR sin b, CR ,

(5.40)

and computing moment equilibrium for a rotating and flapping blade, we obtain the
moment equation in the form

M
&&CR + 2CR CR = a, CR .
I b ,CR

(5.41)

Equations 5.40 and 5.41 include only variables referring to the control rotor. Notice that
the control rotor is mounted on top of the main rotor with a 90 phase shift. The moment
due to aerodynamic forces on the control rotor blade is represented by Ma,CR. An
additional moment due to gyroscopic effects of the rotating control rotor in the rotating
rigid body system needs to be considered.

M gyro = 2 CR q sin b,CR 2 CR p cos b ,CR

(5.42)

The angular velocities of the rigid body motion p and q as well as the rotational speed
CR of the control rotor are included. Since both rotating systems, control and main rotor,
are linked on top of each other, their rotational speed is the same.
The aerodynamic moment due to blade flapping and feathering, rigid body roll and
pitch rates and additional moments due to changes in wind velocities (Mgust) can be found
to be


M a = I b ,CR 2 CR & + CR ( A1,CR cos b ,CR B1,CR sin b ,CR )
8
8
(5.43)
CR
CR

p sin b ,CR +
q cos b,CR + M gust ( b,CR )
8
8

In Equation 5.43 the additional gust moment is written in a general form as a function of
the control rotor blade azimuth b,MR. The equation of feathering blade motion is of the
form
CR = A1,CR cos

b,CR

B1, CR sin

b,CR

(5.44)

59

Chapter 5

Mathematical Modeling

since there is no collective input into the control rotor. Collective input from the pilot is
directly fed to the main rotor blades. The so-called limited extension, , of the control
rotor blade is introduced to describe the limited aerodynamic force due to the only small
profile at the end of the bar. It is assumed that the aerodynamic moment of the blade
section of the control rotor (Ma,CR) is equal to the difference of the aerodynamic moment
(Ma,ent) of a blade with a profile over the entire control rotor radius RCR, and the moment
(Ma,lim) of a blade whose profile would extend from the hub to the point where the true
control rotor blade starts Rlim (see Figure 5.1).

lb

R lim

R CR

CR

Figure 5.1: Control Rotor of the R-50 Helicopter (view from top)

The aerodynamic moment equation therefore becomes

M a ,CR = M a ,ent M a ,lim .

(5.45)

With the linear blade theory it follows that


4
4
M a ,CR = aCR c CR R CR
2 c m a CR c CR R lim
2 cm .

(5.46)

The chord length cCR is measured, and the moment coefficient cm is estimated from
available airfoil data and assumed to be constant. The limited radius can be expressed as
Rlim = RCR lb and substituting in 5.46 it produces

M a ,CR = aCR c CR c m R
2

4
CR


l
1 1 b
R CR

(5.47)

60

Chapter 5

Mathematical Modeling

The last term in parenthesis is defined as the limited extension parameter



l
= 1 1 b
RCR

(5.48)

A linear lift curve slope aCR can be computed as a function of the blade aspect ratio AR

a CR =

2
,
2
1+
AR

with AR =

2
lCR
.
lb c CR

(5.49)

The gust terms in 5.43 due to change of translational velocity are estimated according
to [8 and 10]. The flapping components can be found to be

and

M
s,CR =
M &


v
RCR
CR

M
c ,CR =
M &

u
,

R
CR
CR

(5.50)

(5.51)

where
2 CT 1 C T

M =
+

a 4
2
CR

(5.52)

and

l
2
l c R a 0
2

CT =
2
R 2 ( R )

CR

(5.53)

The aerodynamic derivative due to the flapping velocity is approximately equal to

M & =


.
8

(5.54)

61

Chapter 5

Mathematical Modeling

This represents a very good approximation for hover and slow forward flight
considering the limited extension coefficient and the small Lock number for the control
rotor. For fast forward flight this approximation is only a rough estimate. Additionally,
other aerodynamic effects due to increasing velocity become more important, and the
constant moment derivative with respect to flapping velocity can still be used for most
applications.
Substituting these equations in Equations 5.50 and 5.51, the additional flapping angles
due to wind velocity are obtained as
s ,CR =

and

c ,CR =

0,CR

(5.55)

u .

(5.56)

R
0,CR

The constant control rotor blade pitch 0,CR can be measured.


Computing the derivatives with respect to time in Equation 5.40 and substituting in
5.41, the harmonic balancing method can be applied. Two equations are obtained
including in- and off-axis flap velocities. Neglecting the very small amount of crosscoupling for the control rotor, the equations can be written with respect to the hub plane
in the form

CR
CR
& s,CR =
s,CR
16
16

A1,CR + + s ,CR + p

(5.57)

CR
CR
& c, CR =
c ,CR +
16
16

B1,CR + c ,CR q .

(5.58)

These equations are independent of rotor rotational direction since cross-coupling terms
are totally neglected. With Equations 5.57 and 5.58 there are two additional DOFs added
to the six rigid body DOFs used for the helicopter simulation model. The 8 DOF model
will be used for further analysis. The control rotor inputs are given by the geometry of the
mechanical linkage of control rotor and swashplate, represented by the coefficient kCR in
Equations (5.59).

62

Chapter 5

Mathematical Modeling

For the R50 it is measured that kCR = 1.95, and it follows that

A1,CR = k CR A1, SP
(5.59)

B1,CR = k CR B1, SP
The tilt of the control rotor plane represented by the two additional DOFs also
represents the output of the control rotor and via further mechanical links finally the input
to the main rotor as already described in Equations 5.16 and 5.17. The helicopter
components not yet described in detail will be summarized in the following section.

5.5

Model of Fuselage, Wing and Tail

To complete the modeling, equations are needed expressing the aerodynamic effects
arising from fixed aerodynamic surfaces like wings, fuselage, horizontal and vertical tail,
but also the important influence of the tail rotor mainly responsible for yaw stability and
control. This chapter intends to give a brief overview of the used models. For details
References [1, 6, 8] are very helpful.
Tail Rotor
The tail rotor of the R-50 in a pull-configuration mounted on the left (viewed from
aft) is modeled similar to the main rotor. Flapping degrees of freedom and any crosscoupling are neglected. Thrust, induced velocity and power calculations including drag
and induced power are used similarly as already explained in the main rotor model. The
only control input to the tail rotor is collective pitch, influencing directly induced velocity
and therefore thrust. The rpm of the tail rotor is assumed to be constant since the main
and tail rotor are directly coupled over a belt drive. For power calculations the additional
power to run the tail rotor also needs to be considered.
Fuselage
The fuselage is modeled as a virtual flat plate creating no lift, but drag only. A
quadratic aerodynamic form expresses drag in forward and backward flight limiting the
maximum speed. Drag is also computed for sideward flight, and rotor related down wash
results in additional drag forces on the fuselage. These effects are directly related to
power losses for that the main rotor has to compensate. The quadratic form is taken from
[1] representing a simplified application of Lambs theory [17]. Only approximated data is

63

Chapter 5

Mathematical Modeling

necessary to compute the fuselage forces and moments. The estimated effective fuselage
areas with respect to all three axes and the estimated position of the center of pressure of
the fuselage are the main data for fuselage force calculations. Note, that the modeled
experimental helicopter has a position of CG different from the original helicopter. Not
yet considered changes of mass and moments of inertia due to system modifications for
the flight tests need to be measured and computed (e.g. data link, sensor packages, power
supply and on-board PC).
Wings, Horizontal and Vertical Tail
For the horizontal and vertical tail it is assumed that mainly lift is produced. The
helicopter as purchase from YAMAHA does not include a horizontal tail. For better
handling quality and for mounting GPS- and data link antenna a horizontal tail was
designed (see Appendix E). Since drag due to these components is small compared to
their vertical aerodynamic forces, additional drag will not be considered in this model.
Nevertheless, the effects of aerodynamic stall are included in the equations. The basic
model is again the quadratic form previously mentioned. For the vertical tail it must be
considered that it experiences airflow due to the tail rotor induced velocity. Dependent on
the flight condition, the horizontal tail in general might be located in the down wash field
of the main rotor. To include this effect in the computations, a down wash field of
triangular shape has been assumed in [1].
For additional wings the lift and drag components can also be computed similar to the
horizontal tail approach. Additional drag terms are then becoming important for further
power calculations. As the tail a wing experiences the additional airflow due to the main
rotor downwash field and influences the aerodynamic wing forces and moments. The R50 helicopter does not possess such an additional wing. The calculation is therefore not
necessary but will be included in the simulation model for a use in future applications.
After modeling the important helicopter components and combining them in the
simulation, a linear model can be extracted from the equations of motion as mentioned in
Chapter 2. Results and the analysis for model structures including and neglecting the
effect of the control rotor are shown in the following section.

64

Chapter 5

5.7

Mathematical Modeling

Simulation Results for the Linarized Model

After modeling the single components in the previous chapter, it is of interest to


determine how similar the overall system dynamics are to the real helicopter. Therefore
this section comments on a linearized model obtained by the small-disturbance theory
described in Chapter 2.2, for which the dynamics can be written in the previously
introduced form (Equation 2.1)

x& = F x + G u .
Of further interest is the influence of the control rotor on the pole position of the rigid
body modes. This chapter investigates the dynamic characteristics mainly for hover, since
a comparison with flight test data for forward flight was not possible at this time. Limited
range of the pilots transmitter to guarantee safe operation does not allow a steady-state
forward flight for the time necessary to collect reasonable flight data. The maneuvers that
needed to be flown for the purpose of system identification of the unmanned helicopter
were already challenging enough without the risk of loosing control of the aircraft. More
details on the system identification and the flight tests can be found in Chapter 7. The
actual validation of the simulation model with flight tests was done only for the hover
condition for the reasons given above. Since simulation is not restricted with respect to
time and limited range, almost all maneuvers and flight conditions can be simulated.
Using small perturbations in the states and control inputs from the trimmed flight
condition, the system response of the nonlinear model is computed and force and moment
derivatives can be obtained. These stability and control derivatives represent the elements
of the system and control matrices F and G. The investigated system obtained from the
simulation is the 8 DOF system consisting of the 6 rigid body DOFs and the two
additional control rotor states. This linearized system is given based on the following
definition of state and input vectors
x = [u

w q

s,CR

c ,CR ] T ,

coll , MR Main Rotor Coll. Pitch

B1,SP Longitudinal Swashplate Tilt

.
u =
=
A1,SP Lateral Swashplate Tilt

coll ,TR Tail Rotor Coll. Pitch

(5.60)

(5.61)

65

Chapter 5

Mathematical Modeling

To compare this model including the control rotor with a simulation model neglecting
the control rotor, the two state variables for the control rotor TPP tilt angles s,CR and
c,CR are dropped. The mixing coefficients of the input to main rotor and control rotor are
kMR=1 and k=0. Furthermore it should be mentioned that the inputs to the simulation
without control rotor, A1 and B1, are directly given to the main rotor blades, such that a
swashplate tilt results in a blade pitch change of the same magnitude. This results in a
one-to-one transformation of swashplate tilt to main rotor blade pitch, for both collective
and cyclic inputs. As already mentioned this one-to-one transformation of cyclic inputs is
not true for the system including the control rotor. Therefore differences in the G-matrix
are expected in the elements related to cyclic inputs. Appendix C shows all elements of
the system and controls matrices for the extended model as functions of force and
moment derivatives, trim states, basic helicopter data given by mass and inertia about
principal axis and control rotor data. The assumptions made in Chapter 5.4 to simplify the
control rotor modeling are already included.
The trimmed flight conditions at which the system is linearized are computed by
iteratively varying the helicopter states and inputs until force and moment equations
result in the desired flight condition. For hover, the sum of all forces and moments on the
helicopter CG have to be zero. The values of trim states for hover are dependent on
helicopter forces and moments only. The control rotor does not generate any additional
forces or moments transferred to the helicopter body. Therefore the only difference for
both systems in trim, with or without control rotor, can be found in the required control
inputs.
0

B1,SP (long. cycl.)

A1,SP (lat. cycl.)

collMR

collT R

[rad]

[rad]

[rad]

[rad]

[rad]

[rad]

Hover, no Control rotor

-0.0464

0.0561

-0.0447

-0.0108

0.1085

0.0046

Hover,

-0.0464

0.0561

-0.007

-0.0291

0.1085

0.0046

with Control Rotor


Table 5.1: Trim values in simulated hover for R-50, with and without control rotor

The linearized matrices for this flight condition excluding the control rotor are given in
Tables 5.2 and 5.3, with the longitudinal and lateral derivatives listed in the matrix form
given in Appendix C. Columns and rows are marked with the states and inputs they are
referring to.

66

Chapter 5

Mathematical Modeling

theta

phi

-0.0469

-0.0296

1.4106

-32.1394

-0.0034

0.21

-0.0311

-0.6892

-0.1084

1.4901

-0.0032

-0.0012

-1.8032

0.1559

-0.068

-5.8228

0.0465

0.7644

theta

0.9984

-0.0561

0.0047

-0.0039

0.2102

0.0837

-0.0998

-1.443

32.0888

0.4197

0.2176

-0.0148

-2.3849

-0.4276

-17.881

0.428

phi

-0.0026

-0.0464

-0.0116

-0.2397

0.2881

0.1293

-1.7523

Table 5.2: Analytically obtained system matrix in hover, no control rotor

coll MR

B1

A1

coll TR

-18.1857

32.5193

3.4817

-391.015

-1.4548

-0.3507

-30.9937

-85.857

-50.0334

theta

beta_c_cr

-2.1633

-2.3784

-3.4852

32.5522

-43.3123

-9.0806

-155.97

497.447

-43.3504

phi

-75.505

180.7083

beta_s_cr

2.1633

Table 5.3: Analytically obtained control matrix in hover, no control rotor

To investigate the helicopter dynamics without the control rotor, the eigenvalues for the
system matrix are computed. Damping and frequencies of dynamic modes are listed in
Table 5.4. Figure 5.2 illustrates the pole position of the hover modes in the complex plane
for the coupled longitudinal and lateral motion. The two fastest real modes are the roll
and pitch mode. Both are strongly dependent on the helicopter moment of inertia and the
position of the CG, not exactly known for the investigated model helicopter. As shown in
Figure 4.1 for the longitudinal motion, the pole position is also changing with the varying
natural frequency of the flap motion. Mainly responsible for a change in this natural
frequency is the flapping hinge offset. Decreasing the estimated equivalent hinge offset
for the R-50 rotor system will further reduce the frequency of the roll and pitch mode.
Adjusting the parameters can be used to obtain better agreement with the experimental
results discussed in Chapter 7. Heave and yaw modes have a similar relationship as
discussed in Chapter 4. Both oscillatory modes are slightly unstable at almost the same
frequency.

67

Chapter 5

Mathematical Modeling

Re

Im

Damping

Freq.
(rad/sec)

Mode

0.0976

1.0065

-0.0965

1.0113

Lateral Oscillation

0.0976

-1.0065

-0.0965

1.0113

0.0102

0.8473

-0.012

0.8474

0.0102

-0.8473

-0.012

0.8474

-0.6887

0.6887

Heave

-1.8383

1.8383

Yaw

-2.1633

2.1633

Long. Control Rotor

-2.1633

2.1633

Lat. Control Rotor

-6.1697

6.1697

Pitch mode

-17.811

17.8108

Roll mode

Longitudinal Oscillation

Table 5.4: Eigenvalues, damping and frequencies of hover modes, no control rotor

Poles of Lateral and Longitudinal Motion in Hover, no Control Rotor


Im
1.5

0.5

Re

-0.5

-1

-1.5
-20

-18

-16

-14

-12

-10

-8

-6

-4

-2

Figure 5.2: Poles of coupled longitudinal and lateral motion, no control rotor

68

Chapter 5

Mathematical Modeling

Including the control rotor states in the linearization about hover, the results for the
system and control matrices are obtained and shown in Table 5.5 and 5.6. The same
analysis of eigenvalues leads to results listed in Table 5.7 and Figure 5.3.
u

theta

beta_c_cr

phi

-0.0469

-0.0296

1.4106

-32.1394

-19.8693

-0.0034

0.21

beta_s_cr
2.1273

-0.0311

-0.6892

-0.1084

1.4901

0.8889

-0.0032

-0.0012

-1.8032

-0.2143

0.1559

-0.068

-5.8228

52.4587

0.0465

0.7644

-30.5704

theta

0.9984

-0.0561

beta_c_cr

0.0101

-1

-2.1633

-0.0237

0.0047

-0.0039

0.2102

0.0837

2.1295

-0.0998

-1.443

32.0888

0.4197

19.8894

0.2176

-0.0148

-2.3849

95.3006

-0.4276

-17.881

0.428

303.9402

phi

-0.0026

-0.0464

-0.0116

-0.2397

0.2881

0.1293

-1.7523

beta_s_cr

0.0237

-0.0101

-1

-2.1633

Table 5.5: Analytically obtained system matrix in hover, with control rotor

coll MR

B1

A1

coll TR

-18.1857

11.2387

1.2033

-391.015

-0.5028

-0.1212

-30.9937 -29.6722 -17.2916

theta

0
0

beta_c_cr

-4.2184

-2.3784

-1.2045

11.25

-43.312

-9.0806

-53.9049

171.918

-43.35

phi

-75.505

180.708

beta_s_cr

4.2184

Table 5.6: Analytically obtained control matrix in hover, with control rotor

69

Chapter 5

Mathematical Modeling

Re

Im

Damping

Freq. (rad/sec)

Mode

0.0309

0.7663

-0.0402

0.7669

Lateral Oscillation

0.0309

-0.7663

-0.0402

0.7669

-0.0043

0.6423

0.0067

0.6423

-0.0043

-0.6423

0.0067

0.6423

-0.6838

0.6838

Heave

-1.9241

1.9241

Yaw

-4.0211

7.7222

0.4619

8.7064

Pitch + longitudinal CR

-4.0211

-7.7222

0.4619

8.7064

-10.011

15.3329

0.5467

18.3116

-10.011

-15.333

0.5467

18.3116

Longitudinal Oscillation

Roll + lateral CR

Table 5.7: Eigenvalues, damping and frequencies of hover modes, with control rotor

Poles of Lateral and Longitudinal Motion in Hover, with Control Rotor


Im
20

15

10

0 Re
-5

-10

-15

-20
-12

-10

-8

-6

-4

-2

Figure 5.3: Poles of coupled longitudinal and lateral motion, with control rotor

Two new complex pole pairs are found at frequencies similar to those obtained for the
pitch and roll mode without control rotor. Stability of both oscillatory modes is improved,
and the longitudinal oscillatory mode even becomes stable. Heave and yaw modes show
only small changes. To illustrate the strong influence of the control rotor, Figure 5.4 and
5.5 show the longitudinal and lateral modes for both cases with and without control rotor

70

Chapter 5

Mathematical Modeling

respectively in one plot. Also shown are the control rotor poles for the case of no
coupling between control rotor and rigid body.

Open Loop Poles of Longitudinal Motion, Hover


Im
10
8
6
4
2
0

Re

-2
-4
-6
no Control Rotor
decoupled Control Rotor

-8

with Control Rotor

-10
-7

-6

-5

-4

-3

-2

-1

Figure 5.4: Hover poles of longitudinal motion, with and without control rotor

Figure 5.4 shows that the fast pitch mode and the longitudinal control rotor mode
become one complex pole pair from the previous picture. As discussed in Chapter 4.2.3
the control rotor modes are expected to be located to the left of the fastest rigid body
modes to increase damping. This result can not be achieved with the data set used for the
simulation model. Changing the mass inertia of the helicopter and again the crosscoupling effect due to the hinge offset, we can push the fast pitch mode to the right.
Varying the control rotor data should be avoided. The blade inertia is already calculated
accurately, and other parameters do not have a big influence on the control rotor mode.
Unless the neglected cross-coupling terms for the control rotor states are considered,
there are no changes in the control rotor modes expected. Those coupling terms are
considered to be less critical than the more important unknown helicopter roll and pitch
inertia.

71

Chapter 5

Mathematical Modeling

Open Loop Poles of Lateral Motion, Hover


Im
20

15

10

Re

-5

-10
no Control Rotor

-15

decoupled Control Rotor


with Control Rotor

-20
-20

-18

-16

-14

-12

-10

-8

-6

-4

-2

Figure 5.5: Hover poles of lateral motion, with and without control rotor

The picture for the lateral motion shows similar changes. The new complex pair
evolved from the fast roll mode that coupled with the lateral control rotor mode. The roll
motion is much faster since roll inertia is much smaller than pitch inertia. Effects of
inaccurate helicopter data become therefore even more important in the lateral motion
than in the longitudinal dynamics. In both axis only small changes can be observed in the
remaining modes. Nevertheless, it becomes clear that the control rotor adds damping to
the system as expected from the previous analysis in Chapter 4.

72

Chapter 6

Real-Time Simulation: Hardware and Software

Chapter 6
REAL-TIME SIMULATION: HARDWARE AND SOFTWARE
For the purpose of testing modern control schemes for flight control in a simulator, a realtime simulation is necessary that provides the same immediate response of a helicopter model as
the real physical system. The mathematical models from the previous chapters combined with the
equations of unsteady motion provide this basis for the real-time simulation code. This chapter
describes how the simulation code is implemented together with other necessary components like
flight system hardware and software, pilot-system interface and real-time 3-D visualization of
flight.

6.1

Simulation Elements

Various components of the simulation software are presented and summarized in Figure 6.1,
showing also the order in which computations are done. In Chapter 6.3 these components are
shown in interaction with other components of the real-time-hardware-in-the-loop simulation
facility. This section gives a brief overview of how the software components work together to
provide a real-time system response similar to the real aircraft. Some components listed in Figure
6.1 are optional, and their influence on the system or finally the controller performance can be
studied separately.
For a given pilot or controller input the simulated actuator response is transmitted to the
control rotor, main rotor and tail rotor, resulting in the desired calculated blade pitch. The amount
of input fed through the control rotor and directly to the main rotor depends on the linkage
geometry. Additional wind components due to wind gusts can be added if desired. The terrain
model is important for simulation of ground effect during take-off, low altitude flight and landing
is about to be developed, and therefore not yet included in the simulation. The force and moment
calculations include the interaction of several helicopter components like main rotor-fuselage,
main rotor-horizontal tail, main rotor-tail rotor and tail rotor-vertical tail. Engine modeling is
neglected in the current version. Therefore it was assumed that the engine is capable of providing
enough torque to maintain a constant rpm for main and tail rotor. The non-linear equations for
forces and moments are then used to compute changes in the aircraft motion.

73

Chapter 6

Real-Time Simulation: Hardware and Software

Operator Inputs (Pilot)

Command Filter

Controller

Actuator Model
Gust Model
Control Rotor

Tail Rotor

Main Rotor

Aerodynamic Interaction between


Main Rotor, Tail Rotor, Fuselage

Terrain Model

Summation of Forces
and Moments

Rigid Body
Equations of Motion
Linear Perturbation
Model (optional)
System Response

Sensor
Model

3-DVisualization

Figure 6.1: Elements of Simulation Software

74

Chapter 6

Real-Time Simulation: Hardware and Software

Integration over the simulation time results in non-linear motion for the 6 DOF rigid body
helicopter plus the additional control rotor states. The resulting vehicle response drives a
PCbased real-time 3-D visualization, and it is input to a sensor model of the on-board motion
pack, including models for sensor noise, bias and sensor latency. The controller-in-the-loop
receives the simulated sensor data as inputs and can correct for any deviations from the desired
control objective. The pilot obtains his feedback via the visual queue provided by the PC-display
only. The screen displays a helicopter and a simple model of the surrounding terrain, including a
grid on the earth surface for better orientation as shown in Figure 6.2.

Figure 6.2: Display of the R-50 real-time simulator on PC-screen

A cockpit view and different views from the outside of the aircraft are possible to create a
picture as real as possible for the pilot. Improving displays and the visual pilot-system interface
will allow an even better simulation of the real system. Further improvements in this direction
would therefore be desirable.
A further component allows the computation of trim for each desired flight condition in the
entire flight envelope of the aircraft. Using a linear perturbation routine this part also provides the
linear stability and control matrices for a trimmed flight condition. This linearized model is very
useful for system identification and validation as mentioned in Chapter 5. It should be noticed
that the Adaptive Neural Network Controller that needs to be tested with the simulator also needs
an approximation of the linear system matrices for hover condition. In the simulation these
matrices can be accurately calculated by this routine and the controller should show good results
close to hover even without the network. The adaptive controller is based on an inverting control

75

Chapter 6

Real-Time Simulation: Hardware and Software

scheme that is accurate if the inverse of the system and control matrix is known. Chapter 8 and
[25] provide details on the controller. The trim routine can be either used for advance calculations
of trim tables, or it can be activated during the simulation process via the keyboard.
The command filter and controller in Figure 6.1 are components of the flight system including
pilot-controller interface, hardware and software switches and a ground control station providing
system information and communication with the on-board flight control computer. The elements
of the flight system will be described in the next section.

6.2

Flight System Elements

While the elements of the simulation mostly contain mathematical models and software code,
the flight system includes most of the hardware components like actuators and control surfaces,
mechanical linkages and the ground control station. The part of most interest is the flight control
system. Coded in the C programming language, it is able to provide the real time simulation in
double precision with an update rate of 100 Hz on a 200 MHz Pentium-based Single Board
Computer (SBC). A previously developed commercial-grade flight control system, known as the
R1 [28], provides management of sensor data, hardware and software interface to actuators and
pilot, and it manages the telemetry links to the ground control station. The flight control system
can consist of any appropriately coded flight control scheme developed for helicopter control.
The control system of recent interest is the Adaptive Nonlinear Flight Controller described in
detail in [25] and summarized in Chapter 8. With any control system integrated in the unmanned
R-50 helicopter, the result is a very capable testbed for evaluating or testing modern control
schemes and investigating handling quality and system performance improvements for aircraft or
controller.
During the hardware-in-the-loop simulation the real helicopter is used with all of its on-board
system components. All the commands and signals are generated and processed as in real flight.
Only sensor data is generated by the simulation code. This allows a detailed examination of
actuator and system response to changes in controller parameters, variables or controller
configuration. The best controller performance can be obtained by fine tuning the controller in
simulated flight while observing helicopter motion on the display and actuator displacements on
the real system to assure proper behavior and actuator response in each flight phase.
Pictures of the R-50 helicopter manufactured by YAMAHA and ubgraded by Georgia Tech
and GST are given in Appendix E. Further details can be obtained from the flight manual [29].
The upgrade of the Flight Control System as used in flight tests, allows a high speed data
transmission between the four Motorola processors, an eight channel ultrasonic ranging system,
differential GPS of optional accuracy down to 2cm, the Attitude and Heading Reference System

76

Chapter 6

Real-Time Simulation: Hardware and Software

(AHRS) based on a 3-axis magnetometer and a 3-axis state rate sensor. In planned future flight
tests the AHRS will be substituted by a complete GPS-aided inertial navigation platform for
further improvement of data accuracy and evaluation of the controller and system performance.
The performance of navigation and sensor components, as well as PC performance also influence
data transfer speed. It can cause time delays, noise or an undesirable large and varying bias in
sensor data used by the controller, causing undesirable non-linearities, unknown dynamics or
even instability. It was also observed during tests that the transfer rate of sensor data to the
ground station via the data link is very limited, and that saving the data on board at too high
frequency causes a delay in the data processing and controller computations. Therefore the
frequency and size of the data set saved for evaluation purposes must be kept as low as possible.

6.3

Hardware-In-The-Loop-Simulation

The real-time-hardware-in-the-loop test facility is schematically illustrated in Figure 6.2 from


[28]. The left half of the figure shows necessary model components to simulate the whole
helicopter system as described in section 6.1. The elements of the flight system in the right half
represent the equivalent hardware components and necessary software to combine both. An
interesting opportunity of this facility would be the optional replacement of parts of the
simulation as actuators, mechanical linkages and control surfaces by their mechanical
equivalence. The real dynamic response of actuators could be measured or investigated, real time
ability of the on-board flight control computer can be tested, and the data transfer to the ground
control station can be optimized while the helicopter is still on ground. The simulated system
response is visualized by the pilot on the PC-screen (Figure 6.2) while commanding the necessary
inputs via the transmitter also used in real flight. The advantage of this simulation is that
theoretically for a successful controller test in real flight, the mathematical helicopter model
simply needs to be replaced by the real helicopter. The better the mathematical simulation of the
whole system and all important aspects like disturbances, noise and time delays, the easier is the
transition from simulation to real flight, and perhaps more important, the lower the risk of loosing
the aircraft because of system failure.

77

Chapter 6

Real-Time Simulation: Hardware and Software

SIMULATION ELEMENTS

FLIGHT SYSTEM

Copy of Actuator Interface Logic

True System Dynamics

Simulated Actuator Dynamics

True Sensor Data

Simulated Mechanical Linkage


Software Switch
Software Switch
Gust
Model

Fuselage
Aero

Rotor
Models

Controller
Engine
Models

Command Filters
Operator
Inputs

Software Switch
Terrain
Model

Summation of Forces and Moments


Rigid Body Equations of Motion

Actuator
Interface
Logic

Pilot
Interface
Logic

Serial
Tx
Pilot
Inputs

Sensor Models
Artificial Loads
Sensed Linkage Position

Real-Time 3-D Flight Animation Tool

Safety Switch

PWM
Tx

True Actuator Response

Ground Control Station

Figure 6.3: Joint GST/Ga Tech real-time hardware-in-the-loop simulation facility [28]

The true system dynamics in flight are input to sensors and controller, and all simulation
models are taken out of the loop. If further tuning of controller parameters should be necessary,
changes can still be made in flight via the data link from the Ground Control Station. More
advanced applications like high performance maneuvers or simulation of subsystem failures to
test the controller performance can easily be introduced during flight by limiting actuator
displacements, introducing high latency or sensor failure. These advanced procedures might
follow a more general test. The purpose of a basic test program is to assure reliability and the full
functionality of the controller in normal flight conditions. The advanced test procedures should be
subject of intense real-time simulations before being applied to the real system. A very important
option that should always be available during all flight tests, is a switch back to the open loop
system to give the pilot full authority. It can also be switched to the already approved and fully
tested control system, the Yamaha Attitude Control System (YACS) with the pilot in the loop.
Before the simulation can be used for detailed flight tests, model validation is important to
gain confidence and assure the correct response of the math model compared to the real
helicopter. The validation procedures and some results are presented in the following chapter.

78

Chapter 7

Validation of the Helicopter Simulation

Chapter 7
VALIDATION OF THE HELICOPTER SIMULATION MODEL
In the previous chapters background information for understanding helicopter and
rotor dynamics were presented, and a math model for simulating rigid body and rotor
dynamics, including the control rotor, were described. The purpose of this chapter is to
evaluate and validate this model by comparing simulated results with flight test data. The
complexity of the system combined with limitations related to the operation of the
remotely controlled helicopter make the flight test program a difficult task. Improvements
and further investigations related to the system identification and model validation are
recommended, but require changes in the system to improve the data quality and
quantity. Improvements are necessary related to eventual time delay in the data
processing and increasing the data transfer rate to the ground station.

7.1

Flight Test Data and Requirements

The test site located 40 miles south of Atlanta allows flight maneuvers in the entire
range of the radio transmitter. A mobile ground station operated by Guided Systems
Technologies, Inc., (GST) is available and includes devices for data transfer, flight
recording and flight surveillance. Sensor data is processed on board and sent to the
ground station via a data link. Information from the ground station to the on-board system
to switch controller modes, change system configurations in flight, or to activate the onboard data collection is transferred through the same data link.
For the system identification sensor data of high accuracy and low noise level is
desirable. The configuration of the data link was originally optimized for system
surveillance in flight, allowing a data transfer of numerous parameters and variables at
only a relatively low frequency. Maximum performance of the data link without loosing
the real-time capability of on-board computations is found at a transfer rate of 5Hz,
processing 66 variables and parameters including sensor data, controller variables for
later flight tests, GPS signals and pilot commands. This data can be stored on ground for
the entire flight test. This low transfer rate will limit the accuracy of the system
identification especially with respect to the high frequent system response.
A second recording device is available on board the helicopter. This system runs on a
200MHz PC with an available flash disk memory of 20KB. Sensor data and controller
79

Chapter 7

Validation of the Helicopter Simulation

calculations are processed on board. To assure real-time performance the sampling rate is
limited to a minimum of 100Hz. Saving data on board puts an additional load on the
system, and the minimum sampling rate can not be maintained if frequency, number of
saved variables or recording time is too large. For the actual system configuration it is
possible to achieve a recording frequency of 10Hz for the duration of several flight
maneuvers, with the number of recorded variables limited to 8. Thus the system
identification process was hampered by a low data transfer rate. These limitations require
a careful investigation of useful system identification procedures to support a validation
of the simulation. Two appropriate identification procedures are presented in this chapter.
Angular rate sensors measure the body rotational velocities in the body-fixed frame.
Translational velocities are obtained by a GPS receiver of high accuracy, transmitting
data in a vehicle carried north-east-down reference frame. Euler angles are computed by
integrating angular rates and compensated for bias and measurement noise before the data
is saved.
For further data evaluation, the velocity components need to be transformed from
the vehicle carried frame to the body-fixed frame, and the bias needs to be removed. The
transformation matrix Lbv is given by the following equation

cos cos

(sin sin cos


Lbv =
cos sin )

(cos sin cos


+ sin sin )

cos sin

sin

(sin sin sin


+ cos cos )

sin cos

(cos sin sin


sin cos )

cos cos

(7.1)

and the velocity components in the body-fixed frame are computed as


u

v = Lbv
w

x&
v
y& + b .
z&

(7.2)

v
The vector b contains the bias for all three axis measured on ground or in hover, where
all angular rates and translational velocities are known to be zero.

80

Chapter 7

7.2

Validation of the Helicopter Simulation

System Identification Procedures

One approach chosen to validate the simulation model is to experimentally identify


the parameters of an assumed linear system for a known flight condition, and to compare
the results to the linear system obtained from the simulation by small perturbation theory.
The aircraft motion is measured and a model has to be found reflecting the same physical
behavior of the real system. A valuable tool for model structure and fidelity assessment is
frequency response analysis. This approach will be described in Section 7.2.2 to identify
the most important system parameters.
Section 7.2.3 presents a time domain approach by looking at a static trim position like
hover and then applying step inputs to investigate the dynamic system response in each
axis. This will probably result in a less reliable estimate, because the least-square method
for estimating the approximated state-space model is sensitive to noise, low recording
rate of sensor data, measurement errors and initial starting values for the estimated
parameters.
Flight conditions that should be investigated are hover and forward flight, whereas for
a remotely controlled helicopter a steady-state forward flight condition is hard to
maintain due to the limiting transmitter range. Therefore the parameter identification is
presented for the most important parameters in hover only. Further system identification
tests to improve accuracy and usefulness of the real-time simulation is necessary and will
be the subject of a more detailed study for the R-50 experimental helicopter in the near
future.
7.2.1

Static Trim Values

First, the sensor data is recorded in a hover flight condition at about 30 feet altitude to
determine trim values without any disturbances or influences due to ground effect.
Measured trim values for hover are listed in Table 7.1 and can be directly compared to
the trim values obtained from the simulation. Also listed are the measured and simulated
control inputs for hover.
Trim values for the simulated model partially do not agree with test data. However,
the avionics box including navigation system, sensor pack and PC is not included in the
simulation. Furthermore, changes in the CG position and helicopter inertia due to the
additional component added to the helicopter are completely neglected. The simulated
collective main rotor input, mainly dependant on the helicopter gross weight, is smaller
compared to test results. Pitch and roll angles are strongly dependent on the position of
the CG relative to the rotor hub. Adding estimates for the avionics box to the system data,

81

Chapter 7

Validation of the Helicopter Simulation

the trim values change towards the experimentally obtained values as it can be seen in
Table 7.1 A more detailed model of the avionics box has to be computed and included in
the helicopter data set for an accurate simulation. The analysis in this chapter is based on
the helicopter model including the approximated changes due to the avionics box listed
below.

Flight Test, Hover


Simulation, Hover

B1,SP (long. cycl.)

A1,SP (lat. cycl.)

collMR

collT R

[rad]

[rad]

[rad]

[rad]

[rad]

[rad]

0.0525

0.0661

0.0204

-0.0089

0.1567

0.232

-0.0464

0.0561

-0.007

-0.0291

0.1085

0.0046

0.0073

0.0427

0.0064

0.004

0.1291

0.0166

0.0068

0.0334

0.007

0.0688

0.1136

0.0198

(no avionics box)


Simulation, Hover
(incl. avionics box:
25 lbm, xCG=-3in, zCG =7in)
Simulation,
Forward Flight at 20fps
(incl. avionics box)

Table 7.1: Measured and simulated trim values for R-50 in hover and forward flight

For steady-state forward flight at 20fps, the simulated trim values and the control
inputs are also listed. No flight test data is yet available to compare with, but as described
in Chapter 4, the influence of forward speed is observable in increasing cyclic control
inputs to compensate for the cross-coupling and aerodynamic rotor effects due to the
higher velocity at the advancing blade. Less input in collective main rotor pitch is
required in forward flight, reflecting the decrease in desired thrust for an increasing
forward velocity.
In Figure 7.1 a trim table is computed, and the results are shown as a function of
forward speed from 20fps to +20fps. Roll angle, pitch angle and the inputs show
reasonable values during transition from hover to forward or backward flight. The trim
values show consistent results throughout the investigated speed range, and compare
qualitatively with results that can be found in literature [6, 7, 11]. The control rotor
requires a relatively large input in the longitudinal cyclic control to maintain forward
speed. This commanded swashplate tilt results in a blade pitch of the control rotor and
then, through mechanical linkages, in a blade pitch of the main rotor. Responsible for

82

Chapter 7

Validation of the Helicopter Simulation

TPP and thrust tilt is the actual main rotor blade pitch, which is less than the computed
pilot input. Cyclic inputs are strongly dependent on the approximated mechanical mixing
of direct and indirect inputs from swashplate to control and main rotor. Ground tests
visualizing the real resulting control and main rotor TPP tilt due to a swashplate tilt could
provide more information to model this mixing more accurately. The present model is
based on geometry and single blade pitch measurements over the blade azimuth. The
position of the CG and mass inertia are not accurately known and therefore also
contribute to the only slowly varying pitch and roll angle with changes in forward speed.
With increasing velocity the estimated center of aerodynamic pressure of the fuselage
becomes more important. This variable was estimated based on the computed center of
fuselage projections in all three axes. Changing the flap hinge offset, and therefore the
influence of cross-coupling and additional moments on the body, affects the required
lateral cyclic input needed to compensate for this effect in forward flight. The previously
mentioned parameters should be tuned to achieve good agreement between the simulation
and the real helicopter.

R-50 Trim Table (incl. Control Rotor)

[rad]
0.15

0.1

0.05

Coll. Main Rotor Pitch


Long. Cyclic Pitch
Lat. Cyclic Pitch

-0.05

Coll. Tail Rotor Pitch


Body Pitch Angle
Body Roll Angle
-0.1
-20

-15

-10

-5

10

15

20

Forward Speed [fps]

Figure 7.1: Trim table for R-50 (simulation), rearward to forward flight

83

Chapter 7

7.2.2

Validation of the Helicopter Simulation

Frequency Response Analysis

From the determined hover condition, the goal is now to command computer
generated sinusoidal inputs of specific amplitude and frequencies in each axis and record
the system response. The aircraft should therefore always be in a flight condition near
hover. The flight tests are performed mainly open loop, and the pilot is advised to
command only low frequent inputs to keep the helicopter near hover. Frequency sweeps
with varying frequencies in a single flight maneuver could not be flown due to the
difficulty in maintaining full control over the helicopter throughout the entire maneuver.
Those inputs used for flight test with full size helicopters allow in general a relatively
easy and accurate system identification with reasonable results.
In the yaw channel of the R-50 a yaw damper is present providing rate feedback to
the yaw input. The yaw motion is therefore highly damped, and the response expected in
yaw will be different compared to the simulation neglecting this rate feedback. For future
tests this rate feedback should be removed by calculating the amount of feedback and
removing it from the final input to the collective tail rotor pitch. This will be necessary
for future tests where the NN Adaptive Controller is supposed to provide this feedback.
In addition, mechanical coupling of collective tail and main rotor collective pitch should
be modeled and analytically removed. Since the vertical motion due to collective inputs is
not investigated, this mechanical coupling is also neglected in the simulation model. The
collective main rotor input is kept constant for all flown test maneuvers.
Experimental Bode Plots
Frequency response methods as developed by Bode in the 1930s, are very powerful
when modeling transfer functions from physical data. The concept is that in steady-state,
sinusoidal inputs to a linear system result in a sinusoidal response of the same frequency
as the input, but they differ in amplitude and phase angle. Measuring these differences
and concluding how the desired system can be best approximated is the objective of
generating the experimental Bode plots [18].
The ratio of the output sinusoids magnitude to the input sinusoids magnitude is
defined as the magnitude response, the difference in phase between output and input as
the phase response. Both responses are functions of frequency and represent only the
steady-state sinusoidal response of the investigated system.

84

Chapter 7

Validation of the Helicopter Simulation

The following assumptions are made to simplify the analysis:

Only angular rates in all three axes are investigated and their dynamics are decoupled.
The steady-state response of the helicopter can be approximated at low frequencies by
a first-order system.
Delay due to play in mechanical linkages and signal processing in the electrical
systems result in pure phase shift that needs to be included in the analysis.

Thus the resulting linear system for all three axes can be presented in the following
block diagram (Figure 7.2). The inputs to the uncoupled system for roll, pitch and yaw
motion are pure lateral cyclic, longitudinal cyclic and collective tail rotor pitch
respectively. Outputs are roll, pitch and yaw rate. Equations are presented for the pitch
channel only, but the approach is identical for roll and yaw motion.

up

uq

ur

s Td , p

p& = a p p + b p u p

s Td , q

q& = a q q + bq u q

s Td ,r

r& = ar r + br ur

Figure 7.2: Block diagram of approximated linear helicopter dynamics for hover

Writing the transfer function for the pitch channel it follows that

s T

bq e d , q
q
=
,
uq
(s + a q )

(7.3)

85

Chapter 7

Validation of the Helicopter Simulation

with Td,q representing the time delay in the pitch channel and s the Laplace transform.
The parameters aq and bq are to be identified from the frequency response data.
Substituting s = j in the Equation 7.3, we obtain the pitch transfer function

Gq ( j ) =

bq e

j Td , q

( j + a q )

(7.4)

The magnitude of the frequency response is then


bq
G q ( j ) =

q
( )
u
q

aq

a
q

(7.5)

+1

The phase angle due to pure time delay can be written as

d ,q ( ) = Td ,q ,

(7.6)

resulting in a total phase angle expressed as the phase response

q ( ) = tan 1
aq

Td , q .

(7.7)

For << 1 in Equation 7.5, an estimate independent of the low frequency response,
written as
q
u

bq
aq

(7.8)

Figure 7.3 shows a typical test result for the sinusoidal response of the helicopter in the
pitch channel.

86

Chapter 7

Validation of the Helicopter Simulation

Pitch Channel Frequency Response


Sinusoidal Input: Frequency 0.75 Hz, Amplitude 0.035 rad
0.060
0.040
0.020

Amplitude

0.000
-0.020
-0.040
-0.060
-0.080
Long. Cyclic Input [rad]

-0.100

Pitch Rate [rad/sec]

-0.120
0.0

5.0

10.0

15.0

20.0

25.0

Time [sec]

Figure 7.3: Frequency response in the pitch channel, = 0.75Hz, u q =0.035 rad

Measuring time differences and given the frequency of the exciting sinusoidal input,
the total phase shift q() is
q ( ) = t ,

(7.9)

with t is the measured time difference. The phase shift due to system dynamics and pure
delay are included in Equation 7.9. Results for a typical test data are plotted in a Bode
diagram (Figure 7.4). The mean value of at least three data points in the time history plot
for one flight maneuver is used to compensate for measurement errors. The plots show
results for more than one measurement at each frequency. Different sections in the time
history plot are picked for each measurement. The data needs to be corrected for bias and
trim value. In the pitch channel a positive input (forward tilt of swashplate) results in a
negative body pitch rate (forward pitch of helicopter).

87

Chapter 7

Validation of the Helicopter Simulation

Experimental Bode Plot, Long. Cyclic Input to Pitch Rate


25

Gain [dB]

20
15
10
5
0
1.0

10.0

100.0

Frequency [rad/sec]

0
-50

Phase []

-100
-150
-200
-250
-300
-350
1.0

10.0

100.0

Frequency [rad/sec]

Figure 7.4: Experimental Bode plot for the pitch channel

88

Chapter 7

Validation of the Helicopter Simulation

Experimental Bode Plot, Lat. Cyclic Input to Roll Rate


25

Gain [dB]

20
15
10
5
0
1.0

10.0

100.0

Frequency [rad/sec]

Phase []

-50

-100

-150

-200

-250
1.0

10.0

100.0

Frequency [rad/sec]

Figure 7.5: Experimental Bode plot for the roll channel

89

Chapter 7

Validation of the Helicopter Simulation

Experimental Bode Plot, Coll. Tail Rotor Input to Yaw Rate


20
18
16
Gain [dB]

14
12
10
8
6
4
2
0
1.0

10.0

100.0

Frequency [rad/sec]

Phase []

-50

-100

-150

-200

-250

1.0

10.0

100.0

Frequency [rad/sec]

Figure 7.6: Experimental Bode plot for the yaw channel

90

Chapter 7

Validation of the Helicopter Simulation

It can be seen, that the assumption of a first-order system in the pitch, yaw and roll
rate dynamics is only approximately true for low frequencies up to 5Hz. The
following calculations are therefore only valid for the low frequency response. For higher
frequencies the behavior is more like a second-order response with a resonance frequency
for the pitch response at about 9 rad/sec. This mode can be identified as the longitudinal
control rotor mode coupled with the longitudinal fast pitch mode of the body dynamics.
The Bode plot in Figure 7.5 for the roll rate response due to lateral cyclic input shows a
second-order response with a natural frequency at about 13 rad/sec identified as the fast
roll mode. The yaw response due collective tail rotor input is shown in Figure 7.6. If we
want to compare these results with simulation results, a time delay needs to be considered
resulting in a pure phase shift in all three axes. An approximation for this time delay can
be computed using the Bode plots and the previously made assumption of a first-order
response at low frequencies.
The tested frequency range is chosen based on estimated maximum limits causing
uncontrollable instabilities, and lower limits mainly restricted by the helicopter response.
For low frequencies, flight conditions develop which are too dangerous to be maintained
for the time necessary to record the steady-state response. Furthermore, pilot inputs of
about the same magnitude as the sinusoidal input are observed in the test data at very low
frequencies. An accurate analysis for those cases is consequently not possible since the
pure sinusoid can not be extracted from the test data to measure phase shift and
magnitude. Controllable flight maneuvers down to 0.5Hz in the pitch axis, 0.3Hz in the
roll axis and 0.2Hz in the yaw axis are possible. As a maximum limit = 2Hz is chosen
in all three axes.
To obtain approximations for the system parameters of the assumed first-order system
described in Equation 7.3, in each Bode plot at least two data points in the low frequency
range are used. Knowing the phase shift as expressed in Equation 7.7 for two different
frequencies, we have two equations with two unknowns, the system parameter aq and the
time delay Td. Those equations can be solved numerically for aq and Td. Equation 7.5 can
then be solved for the unknown bq. The results for this approximate identification are
shown in Table 7.2 for all three axes.

91

Chapter 7

States

Validation of the Helicopter Simulation

-a

-5.6

-2.772

-1.838

25.6

-14.716

10.34

Td

0.112

0.115

0.099

Parameters

Table 7.2: Identified parameters and system delay in hover using experimental Bode plots

Several system components can be made responsible for the time delay, which could
cause problems for the later implementation of the new control architecture. Processing
sensor data, sampling time synchronization, a Bessel anti-aliasing filter for the rate
sensor, play in mechanical linkages and finally actuators contribute to a time delay that
can add up to a significant value important for stability and control issues.
The next section compares the experimentally obtained results with simulation
results, and the observed time delay not present in the linear simulated model will be
investigated.
Simulated Bode Plots
Generating the corresponding Bode plots for the linearized helicopter model in hover
is much simpler, since all stability derivatives have already been computed. The linear
model is further reduced to a system with decoupled longitudinal and lateral motion. For
the decoupled longitudinal motion the following state-space model taken from the fully
coupled computation of the linear system matrices (see Appendix C) is assumed,
including the longitudinal control rotor tilt and the approximated changes in mass and CG
position for the additional avionics box (Equation 7.10).
Inputs to the collective main rotor pitch are not considered. The longitudinal cyclic
input long remains as the only longitudinal input. The lateral model is similar and can be
extracted from the tables in Appendix D. Results for the longitudinal model are shown in
Figure 7.7 for the transfer function from longitudinal cyclic input to pitch rate.
Experimental data and simulated data are plotted in the same picture. Results are shown
for a pure time delay of Td = 0.31sec acting on the system obtained from the simulation.
Very good agreement of simulated and experimental data can be seen.

92

Chapter 7

u&
w&

q& =
&

&c ,CR

Validation of the Helicopter Simulation

32.1731 19.9033 u
1.413
0.0553 0.0039
0.0027 0.5727 0.0236 0.2358 0.1233 w

0.2373
6.9424
0.002
0
68.2896 q

0
0
0.9991
0
0

0.0101
1
2.1633 c ,CR
0
0
(7.10)

11.2579
0.0698

+ 38.6267 long

4.2184
The simulated time delay of 0.31sec seems to be very large, even if all the mentioned
components are considered. This leads to the assumption that there might be some delay
in the data processing, creating a large phase shift related to not synchronized input and
output data collection or other effects introducing some time delay.
Further tests with the real system revealed that the primary contributor of time delay
for very low frequency is the analog anti-aliasing filter at the input to the rate sensors.
Another source was found in the message passing structure in the software. The
remaining measured time delay from sensor input to actuator output varies from 0.030.1sec. This is due to the fact that the various components in the path have finite duration
processes that are unsynchronized and of varying frequencies. The goal for further flight
tests is to reduce the time delay from rate measurement to actuator response to less than
0.01sec. This amount of time delay would be acceptable to achieve high bandwidth
control.
Comparing the experimentally obtained (Table 7.2) and simulated (Appendix D)
stability and control derivatives for the same case, it is seen that simulated results show a
much higher sensitivity to perturbations of states and control inputs. Differences in the
control matrix expressed by the elements bp, bq, br can be partially related to not exactly
modeled changes of CG position and mass properties due to hardware components added
to the original helicopter. The already mentioned additional yaw damper present in the
real open loop system further contributes to the differences in sensitivity and stability.
The used model for the control rotor also includes an approximation defining the mixing
of direct inputs from the swashplate to the main rotor and inputs via the control rotor tilt
as described in Chapter 5. A significant influence is expected from the coefficients

93

Chapter 7

Validation of the Helicopter Simulation

defining this mixing of inputs to the main rotor cyclic pitch, directly related to the rate
response of the helicopter. The geometric flap hinge offset also contributes to the
sensitivity of system response, since rotor moments are directly related to this input
parameter to the simulation.

Bode Plot, Pitch Dynamics, 0.31sec Time Delay


25
Flight Test

Gain [dB]

20

Simulation

15
10
5
0
1.0

10.0

100.0

Frequency [rad/sec]

0
-50

Phase []

-100
-150
-200
-250
-300
-350
-400
1.0

10.0

100.0

Frequency [rad/sec]

Figure 7.7: Experimental and simulated frequency response, 0.31sec time delay, pitch dynamics

94

Chapter 7

Validation of the Helicopter Simulation

Comparing the natural frequencies of the dynamic modes from flight test data with
those from simulation (Figure 7.8 and Appendix F), we find relatively good agreement in
the roll, pitch and yaw response. In those results the same amount of time delay is added
as that identified in the pitch channel. Due to the additional yaw damper in the real
system, the results for the yaw channel show big differences in the magnitude response,
whereas the frequency response is still similar (Figure F.2).
The following system identification method will focus on the validation of the already
obtained values and on the identification of other important system and control
derivatives
7.2.3

Step Response Analysis

This more deterministic identification technique also assumes a linear relationship


between state derivatives x& , states x and control inputs u. The goal is to predict the
measured states based on a linear combination of measured states and inputs defined by
constant parameters ai. The estimated parameters will be compared to those obtained with
the frequency analysis and those calculated from the simulation model.
The fully coupled system as presented in Appendix C for the general case, and in
Appendix D for results obtained from the simulation, contain too many parameters to be
properly identified. Furthermore some of the parameters are not even practically
identifiable with the input-output flight test data available. The flight tests, consisting of
step and doublet inputs over a short recording time, do not provide optimal excitation of
the state response. Some parameters can be assigned based on the analytically obtained
perturbation matrices as given in Appendix C.

Zero terms in the system and controls matrices F and G are not identified.
Cross-coupling terms will be neglected.
Parameters based on geometric functions of trim values are not identified.
Parameters relating p to , q to and r to are set equal to 1.
Collective input is neglected and not changed during the flight maneuvers.
Immediate response of the helicopter is mainly resulting in a change of angular
rates.

The numerical problem is to minimize the error between experimental and simulated
output data for an estimated set of identifiable parameters. A performance function is
defined as the sum of the squared errors between simulated and experimental output data,

95

Chapter 7

V=

1 t 2
1
e (i ) = E
2 i =1
2

Validation of the Helicopter Simulation

(7.11)

where

E = Y Y

(7.12)

and

Y = f ( u , a j ) .

(7.13)

Estimated parameters a j are the elements of the estimated model that gives the
estimated output Y if the known measured input data set u is input to the estimated
model. Minimizing the performance function to find the best solution for the parameters
in the sense of least-square requires computational steps in the direction of a decreasing
V, and finally finding the minimum value of the performance function for an estimated
combination of parameters.
A powerful method to find steps in this direction is given by a Levenberg-Marquardt
method [19]. It evaluates the gradient of the performance function as the partial
derivatives with respect to the parameters to be identified, known as the Jacobian matrix
J=

V
.
a j

(7.14)

An approximated Hessian matrix representing the second partial derivatives can be


defined for a performance function of the form of a sum of squares as

H = JT J .

(7.15)

The gradient of the error function V is then

g = J T e , where e is the vector of the output

errors, and the update law for the parameters is

a k +1 = a k J T J + I

JT e,

(7.16)

known as the Levenberg-Marquardt-Algorithm. For the scalar = 0 Equation 7.16


represents a pure Newton algorithm using the approximated Hessian matrix H. For large
it becomes the gradient descent method with a small step size. The Newton method is
in general better and converges faster for a minimum search near the error minimum.

96

Chapter 7

Validation of the Helicopter Simulation

Thus is decreased after each successful step towards a minimum and increased when a
step would increase the value of the performance function. As a result, the performance
function will always be reduced for all iterations of the algorithm.
The problem of finding local minima instead of the desired absolute minimum is still
present and makes it necessary to start from a different initial guess for the parameter
T
vector = a1 a 2 ... a j and compare the results of several solutions.

The previously described approach is applied to the decoupled lateral motion of the
helicopter and to a system with model structure similar to that assumed for the frequency
analysis in the previous chapter. Results are listed in the following figures and tables. For
the longitudinal motion a successful identification was not possible. Poor velocity data,
and noise in the angle and angular rate measurements, prevented a successful
identification with this approach. Different and more powerful step response
identification procedures might lead to reasonable results.
For a model of the following form
v&
p&
=
&

r&

a1

a 4
0

a 6

a2

32.070

5 .6

a7

v a8

a 5 p 25.6
,
+
lat
0.0524 0


1.838 r 0
a3

(7.17)

the results obtained from the Bode plot analysis are set as fixed. Furthermore, parameters
that can be computed by geometric functions or that are zeros as given in the system
matrix F in Appendix C are again not identified. Error calculations include velocity, roll
rate and roll angle, the error in the yaw rate is neglected. The identified system with the
parameters in Table 7.3 results in a step response shown in Figure 7.8 together with the
measured output data. Control rotor states are not measured and no information about its
dynamics can be obtained by this very limited analysis method. Values for the control
rotor derivatives obtained from the simulation could be set as known and fixed in an
extended model that includes control rotor states. This approach did not result in a better
solution. For the system shown in Equation 7.17, the parameters listed in Table 7.3 are
identified.
a1

a2

a3

a4

a5

a6

a7

a8

-0.0474

-11.1943

-0.9378

0.0247

-3.3562

0.0026

0.6824

2.499

Table 7.3: Identified parameters for decoupled lateral motion, step response

97

Chapter 7

Validation of the Helicopter Simulation

Step Response, Lateral Cyclic Input


0.2
0.15
Roll Rate [rad/sec]

0.1
0.05
0
-0.05
-0.1
-0.15

Lateral Cycl. Input [rad]

-0.2

Simulation

-0.25

Flight Data

-0.3
0

10

Time [sec]

35
Simulation

Body-v-velocity [ft/sec]

30

Flight Data

25
20
15
10
5
0
-5
0

10

10

Time [sec]

0.4

Roll Angle [rad]

0.3
0.2
0.1
0
-0.1
Simulation

-0.2

Flight Data

-0.3
-0.4
0

5
Time [sec]

Figure 7.8: Response of identified model (7.17) compared with measured data

98

Chapter 7

Validation of the Helicopter Simulation

A model similar to the one identified with the frequency method is now assumed,
except that cross-coupling terms are included.
p&
&
q =
r&

a1

a 4
a 6

a2
a5
0

a 3 p a8

0 q + a9
a7 r 0

lon

(7.18)

As an initial guess only the parameters known from the frequency analysis are given, and
off-diagonal terms are set to zero. The following parameters are identified and the
resulting step response is shown in Figure 7.9.
a1

a2

-4.9153 0.5747

a3

a4

a5

a6

a7

a8

a9

0.5641

-0.8205

-2.7600

-5.9641

-2.9033

2.5303

-11.4687

Table 7.4: Identified coupled model parameters for system including only angular rates.

The diagonal terms are relatively close to the identified parameters from the frequency
response method. Note that this is the response due to longitudinal cyclic input only.
Since most of the cross-coupling terms except rate cross-coupling are neglected, a good
fit in the roll and yaw channels are not expected. Good agreement is also found for the
pitch derivative ( bq ) in the control matrix. An identification of system models including
velocity data is not presented, since no reasonable results were obtained. This might lead
to the conclusion that the velocity data achieved from the GPS system is not very reliable.
More emphasis on the system identification is needed in the future to make the
simulator more useful for flight tests. Comparing the identified parameters with the linear
matrices from the simulator, we can see differences in sign and magnitude. The identified
parameters strongly depend on the helicopter inertia as can be seen in matrix F (Appendix
C). Without accurate helicopter inertia data and a good estimation for the position of the
CG, these significant differences in control and system matrix elements will remain.
Some of these differences can also be a result of the neglected effects from inertial or
aerodynamic cross-coupling. Further identification procedures should therefore
emphasize those cross-coupling terms, also because they influence the natural flapping
frequency and therefore might increase stability as shown in Figure 4.1. A method to
identify the control rotor parameters by looking at the step response is not yet possible,
since the control rotor states are not measurable with the currently used equipment.
Improvements in this direction are planned but not yet fully developed.

99

Chapter 7

Validation of the Helicopter Simulation

Step Response, Longitudinal Cyclic Input


0.4
Long. Cyclic Input [rad]
Simulation
Flight Data

Pitch Rate [rad/sec]

0.3
0.2
0.1
0
-0.1
-0.2
-0.3
0

10

12

10

12

10

12

Time [sec]

0.4
Simulation
Roll Rate [rad/sec]

0.3

Flight Data

0.2
0.1
0
-0.1
-0.2
0

6
Time [sec]

0.4
Simulation

Yaw Rate [rad/sec]

0.3

Flight Data
0.2
0.1
0
-0.1
-0.2
0

6
Time [sec]

Figure 7.9: Response of identified model (7.18) compared with measured data

100

Chapter 8

Modern Adaptive Nonlinear Flight Control in Simulation and Real Flight

Chapter 8
MODERN ADAPTIVE NONLINEAR FLIGHT CONTROL IN
SIMULATION AND REAL FLIGHT
Artificial Neural Networks (NNs) function as highly nonlinear adaptive control
elements and have great advantages over conventional linear parameter adaptive
controllers. To examine performance and limits of such a controller, the School of
Aerospace Engineering at the Georgia Institute of Technology has implemented a
recently developed NN Adaptive Flight Controller in the experimental helicopter R-50
previously described and modeled. The goal is to apply this modern controller and show
its capability in normal and critical flight conditions. In [26] the theoretical development
of a direct adaptive tracking control architecture for flight control using NNs is
presented. This work is extended and applied on helicopter flight control in [27]. In [25]
the same controller is also applied to a tiltrotor aircraft, and some stability issues for the
controller architecture are discussed.
This chapter first describes the controller architecture and summarizes its main
features. Then some results for a simulated flight maneuver are shown and compared to
recently obtained experimental results. In flight tests, the NN Adaptive Controller was
first implemented in the pitch channel. Finally, some comments on future work in this
field concerning simulated and real flight conclude this chapter. Improvements necessary
to facilitate an implementation of modern controllers in the real system are still required
and are related to the simulator and hardware components like data link, on-board
system, sensor performance and accuracy.

8.1

Flight Control System

The changing flight characteristics of a helicopter in different flight conditions


provide a challenging opportunity for the application of adaptive flight control laws. Two
basic types of control augmentation for aircraft are rate command and attitude command
systems, each with advantages and disadvantages dependent on the flight condition and
desired task. The system first implemented in the R-50 controller is a rate command
system.

101

Chapter 8

Modern Adaptive Nonlinear Flight Control in Simulation and Real Flight

The basis for the control scheme is a NN Augmented Model Inversion Architecture
shown in Figure 8.1 for the pitch channel. Roll and yaw channel controllers are
constructed similarly.

Pitch
Rate
Command
q

Pilot

q
c
st

1 Order
Command
Filter

1
qc

~
q

Kp

Uq

Uc q

Approximate
Inverse
Transformation

Kd
Pitch Channel Example

Control
Surface
Deflections
^

RCAH

UAD

(Rate Command,
Attitude Hold)

LON

X
Pitch Channel
Neural Network

Helicopter
Dynamics

Normalization

Bias
Uq

Sensed State, q

Figure 8.1: Neural Network Augmented Model Inversion Architecture

As a feedback linearization method dynamic model inversion [20] is used, where the
model inversion is based on linearized dynamics about the hover condition only. The
states included in the model inversion scheme are the angular rates p, q and r. The
dynamic equation is thus given by
& = A1 x1 + A2 + B ,

(8.1)

where A1, A2 and B represent the aerodynamic stability and control derivatives at the
nominal trim point, respectively.
The rates about the body-axes are contained in the vector = [p q r ]T. The main
rotor collective input is treated in this application as an additional slow translational state,
such that x1 = [ u v w coll,MR]T. The remaining inputs are the controls resulting in a
direct moment about the fixed body axes = [lat lon coll,TR]T.
Slow and fast states are separated and any cross-coupling between them is neglected
for the inverting controller. The NN is used to adapt to errors caused by this linearized
inverted model for hover. A pseudo control U of the form
U = U C + & C U AD

(8.2)

is designed for the rotational degrees of freedom ( p, q, r ), including the output of the
linear controller operating on an error signal q~ (see Figure 8.1), the adaptive control term

102

Chapter 8

Modern Adaptive Nonlinear Flight Control in Simulation and Real Flight

Uad from the NN and the filtered commanded acceleration & C . The term Uc specifies the
tracking error transient for the linear controller
t

UC = KP ~
q + K I q~ d .

(8.3)

t0

With & c = [ p& c q&c r&c ]T as the vector of filtered commanded angular accelerations the lefthand side of Equation 8.1 is set equal to the desired accelerations, which are in this case
identically the pseudo controls for each channel. Substituting & c in Equation 8.1 using
Equation 8.2 and solving for the required control, we obtain
= B 1 {U A1 x1 A2 }.

(8.4)

Since neither A1 and A2 , nor B are exactly known, it follows that

= B 1 U A1 x1 A 2 ,

(8.5)

Furthermore, since the elements in x1 are not presently sensed on board the helicopter, it
is assumed that x1 = 0 in the implementation. The inversion error for each channel can be
written in the form

= & A 2 + B .

(8.6)

This inversion error can be expressed as a function of states and pseudo controls. The
tracking-error dynamics for the pitch channel can finally be written as
t

q~& + K P q~ + K I q~ d = U AD (2 ) q ,

(8.7)

t0

with the left-hand side representing the error dynamics and the right-hand side describing
the network compensation error, which is a forcing function of the tracking error
dynamics.

103

Chapter 8

Modern Adaptive Nonlinear Flight Control in Simulation and Real Flight

The adaptation law assuring boundedness of the tracking error for a linear-in-theparameters network can be written in the form
U AD = W T (X ,U )

(8.8)

where W is the vector of network weights, and is a vector of basis functions. In [25] it
has been proven, that errors and network weights remain bounded, and that the errors
asymptotically approach zero if the network is capable of perfectly mapping the function
given in Equation 8.6. This will depend on the network architecture.
The design in the R-50 flight controller employs a multilayered NN with sigmoidal
activation functions () in the hidden layer, capable of approximately reconstructing the
inversion error. NNs of this structure are also called universal approximators. The inputs
to the multilayered NN with one hidden layer shown in Figure 8.2 consist of the state
variables, the pseudo control and bias terms. Variable network weights of the first-tosecond layer interconnection are noted with V, whereas the second-to-third layer
interconnection weights are represented by W.

()

()

x
()
()

Input Layer

Output Layer
Hidden Layer

Figure 8.2: Multilayered network with one hidden layer

A further problem related to adaptive systems and specifically unmodeled dynamics,


can be solved by using NNs. Nonlinear unmodeled dynamics might cause instabilities. A
Nonlinear Dynamic Damper was recently employed at Georgia Tech, which can work in
conjunction with a multilayered NN. This additional component is also added to the

104

Chapter 8

Modern Adaptive Nonlinear Flight Control in Simulation and Real Flight

tested flight control system as a part of the network learning algorithm, to examine its
performance in the presence of unmodeled dynamics like actuators or unknown delays.
Performance improvements are expected mainly related to command tracking and
robustness during normal operation but also in the case of system failure.

8.2

Simulation and Experimental Results

The results in Figure 8.3 show the simulated system response to a commanded
doublet input in the pitch channel. Also considered in this simulation result are sensor
noise and bias, actuator dynamics, time delay and wind gusts acting on the system. The
adaptive NN controller compensates for nearly all the motion in the other channels due to
cross-coupling effects of rotor or body dynamics. It can be seen in the plot, that the
response follows the filtered command very well.

Pitch Channel Command and Response

0.2
0.15
0.1
0.05
0
-0.05
-0.1
-0.15
Pitch Rate Command [rad/sec]
-0.2

Pitch Rate [rad/sec]


Pitch Attitude [rad]

-0.25
11

13

15

17

19

21

23

25

27

29

Time [sec]

Figure 8.3: Simulated system response of R-50, doublet inputs in long. cyclic

It should be also mentioned, that in this result the used linearized model for hover is also
a model with only non-zero diagonal elements. All off-diagonal elements in the system
matrix obtained from the linearized simulation model are set to zero. A further artificial

105

Chapter 8

Modern Adaptive Nonlinear Flight Control in Simulation and Real Flight

model error is introduced by multiplying all system matrix elements by 5. The controller
performance can be evaluated for different magnitudes of the introduced error.
In flight tests the NN adaptive controller was first tested in the real-time simulator to
assure a reliable switching between an operation with and without the controller, and to
verify the stable dynamics of the NN. The controller was first activated in the pitch
channel only. The remaining channels were flown open-loop by the pilot. Starting with
the inverting controller only, learning rates, robustifying components and parameters of
the nonlinear dynamic damper were raised slowly to see the influence on the system
response and tracking quality. The linearized system matrices are taken from the system
identification results of the frequency response method in Chapter 7. It was assumed that
only diagonal terms are of importance and all off-diagonal terms were set to zero. The
results given in Figure 8.4 show a typical input in the pitch channel and the resulting
systems response, including noise, bias, time delay and actuator dynamics. A similar
behavior than in Figure 8.3 can be seen. Tuning the parameters of the NN controller may
improve the tracking of the commanded input and minimize the oscillations observed in
the pitch rate response.

Pitch Channel Command, Response and Control

0.2
Pitch Rate Command [rad/sec]
Pitch Rate [rad/sec]

0.15

Pitch Attitude [rad]


0.1

0.05

-0.05

-0.1

-0.15

-0.2
2020

2022

2024

2026

2028

2030

2032

2034

2036

2038

2040

2042

2044

2046

Time Step Number (~0.25 sec each)

Figure 8.4: Pitch response of R-50 in flight test with NN controller in pitch channel

106

Chapter 8

Modern Adaptive Nonlinear Flight Control in Simulation and Real Flight

The shown results give confidence in the use of the simulator to test modern NN
control architectures. Including sensor noise and bias, actuator dynamics, time delays and
wind gusts, the developed real-time simulator provides a very useful tool to find limits of
the controller performance and to identify the influence of various controller parameters
on the system performance. Influences of various components of the simulation and
disturbances can be investigated separately, and the controller can be tested for its ability
to compensate for unmodeled dynamics. Further flight tests are necessary to find the
optimal setting of controller parameters for the experimental helicopter. Also of interest
is the tracking performance of the controller during high performance maneuvers. These
tests will probably be the subject of flight tests in the near future. To improve the
capability of the simulator and to further validate the simulation model, more detailed
studies in the field of system identification are necessary.

107

References

REFERENCES
[1]

R.K. Heffley, M. A. Mnich: Minimum-Complexity Helicopter Simulation Math


Model, NASA CR 177476, USAAVSCOM TR 87-A-7, California, April 1988

[2]

R.K. Heffley, S.M. Bourne, H.C. Curtiss, JR., W.S. Hindson, R.A. Hess: Study of
Helicopter Roll Control Effectiveness Criteria, NASA CR 117404, USAAVSCOM
TR 85-A-5, California, April 1986

[3]

R.K. Heffley, S.M. Bourne, M. A. Mnich: Helicopter Roll Control Effectiveness


Criteria, Manudyne Report 83-2-2 (forthcoming NASA CR/AVSCOM TR), May
1987

[4]

B. Etkin, L.D. Reid: Dynamics of Flight, Stability and Control, 3rd Edition, John
Wiley and Sons, Inc., 1996

[5]

B. Etkin: Dynamics of Atmospheric Flight, John Wiley and Sons, Inc., 1972

[6]

R.W. Prouty: Helicopter Performance, Stability, and Control, Robert E. Krieger


Publishing Company, Florida 1990

[7]

R.W. Prouty: Helicopter Aerodynamics, Rotor and Wing International, Philips


Publishing, Inc., 1985

[8]

W. Johnson: Helicopter Theory, Dover Publications, Inc., New York 1980

[9]

A. Gessow, G.C. Meyers, JR.: Aerodynamics of the Helicopter, Frederick Ungar


Publishing Co., New York 1967

[10] M.G. Perhinschi, J.V.R. Prasad: A Simulation of An Autonomous Helicopter,


Proceedings of the Remotely Piloted Vehicles 13th International Conference,
Bristol, United Kingdom, 30 March-1 April 1998,
[11] G.D. Padfield: Helicopter Flight Dynamics: The Theory and Application of Flying
Qualities and Simulation Modeling, AIAA Education Series, 1996
[12] E.N. Johnson, P.A. DeBitetto: Modeling and Simulation for Small Autonomous
Helicopter Development, AIAA Modeling and Simulation Technologies
Conference, 1997
[13] M.A. Glauert: A General Theory of Autogyro, Royal Aeronautical Establishment R.
and M. No. 1111, November 1926
[14] C.N.H. Lock: Further Development of Autogyro Theory, Royal Aeronautical
Establishment R. and M. No. 1127, March 1927

108

References

[15] P.D. Talbot, B.E. Tinling, W.A. Decker, R.T.N. Chen: A Mathematical Model of a
Single Main Rotor Helicopter for Piloted Simulation, NASA-TM-84281, California,
September 1982
[16] R.T.N. Chen: Effects of Primary Rotor Parameters on Flapping Dynamics, NASA
TP-1431, January 1980
[17] H. Lamb: Hydrodynamics, Dover, New York 1945
[18] N.S. Nise: Control Systems Engineering, The Benjamin/Cummings Publishing
Company, Inc., 1992
[19] W.H. Press, S.A. Teukolsky, W.T Vetterling, B.P. Flannery: Numerical Recipes In
Fortran, The Art of Scientific Computing, 2nd Edition, Cambridge University Press,
1992
[20] L.-J. E. Slotine, W. Li: Applied Nonlinear Control, Prentice Hall, Inc., 1991
[21] D.A. Teare: Modeling and System Identification for Rotorcraft, Thesis, Georgia
Institute of Technology, Atlanta, March 1988
[22] W.D. Lewis: An Aerolastic Model Structure Investigation for a Manned Real-Time
Rotorcraft Simulation, Thesis, Georgia Institute of Technology, Atlanta, August
1992
[23] S.H. Sturisky: A Linear System Identification and Validation of an AH-64 Apache
Aerolastic Simulation Model, Thesis, Georgia Institute of Technology, Atlanta, July
1993
[24] M. Hossein, M.B. Tischler: An Empirical Correction Method for improving OffAxes Response Prediction in Component Type Flight Mechanics Helicopter
Models, AGARD-CP-592, April 1997
[25] A.J. Calise, R.T. Rysdyk: Nonlinear Adaptive Flight Control using Neural
Networks, IEEE Control System Magazine, Vol. 18, No. 6, December 1998
[26] B.S. Kim, A.J. Calise: Nonlinear Flight Control Using Neural Networks, Journal of
Guidance, Control, and Dynamics, Vol. 20, No. 1, January-February 1997
[27] J. Leitner, A.J. Calise, J.V.R. Prasad: Analysis of Adaptive Neural Networks for
Helicopter Flight Control, Journal of Guidance, Control, and Dynamics, Vol. 20,
No. 5, September-October 1997
[28] J.E. Corban, A.J. Calise, J.V.R. Prasad: Implementation of Adaptive Nonlinear
Control for Flight Test on an Unmanned Helicopter, Proceedings of the 37th
Annual IEEE Conference on Decision and Control, December, 1998
[29] YAMAHA: Unmanned Helicopter R-50 Instruction Manual, 1st Edition, YAMAHA
Motor Co., Ltd., Aeronautics Operation, May 1995

109

Appendix A R-50 Helicopter Data

APPENDIX A R-50 HELICOPTER DATA


/*** R50 Parameters ***/
/*** Loading Parameters ***/
FS_CG = 24.99 ;

/* Location of CG in inches from the zero fuselage station

*/

WL_CG = 20.154 ;

/* Vertical CG location in inches above the zero waterline

*/

WT

/* Vehicle Take-Off Gross Weight in lbs. (weight is assumed constant)

*/

= 0.1997*7.345;

/* slug-ft2

*/

IY

= 0.6231*7.345;

/* slug-ft2

*/

IZ

= 0.6*7.345;

/* slug-ft2

*/

IXZ

= 0.0;

/* slug-ft2

*/

IX

= 97.85 ;

/*** Main Rotor Parameters ***/


FS_HUB = 27.50;

/* station line of main rotor hub, measure aft of the zero fuselage station

*/

WL_HUB = 42.25;

/* waterline location of the main rotor hub

*/

IS

/* Main rotor shaft tilt backward of vertical (radians)

*/

/* geometric main rotor flapping hinge offset or effective hinge offset

*/

= 0.0;

E_MR
I_B

= 0*2.25/12.0;

/* flapping inertia of a single blade about the flapping hinge

*/

R_MR

= 0.86754;
= 5.05;

/* main rotor radius in feet

*/

A_MR

= 6.0;

/* effective lift curve slope of main rotor (per radian)

*/

RPM_MR = 870.0;

/* nominal main rotor anglular velocity (rev. per min)

*/

CDO

= 0.010;

/* effective profile drag for main rotor blade cross section

*/

B_MR

= 2.0;

/* number of main rotor blades

*/

C_MR

= 0.354;

/* blade chord in feet

*/

TWST_MR = 0.0;

/* effective blade twist in radians

*/

K1

/* tangent of delta-3, (effective pitch-flap coupling)

*/

/* based on flaping hinge geometry

*/

/* rotor spin direction -1 = clockwise

*/

DIR

= 0.000;
= -1.0;

/*** Fuselage Parameters***/


FS_FUS = 25.25;

/* station line of center of pressure of fuselage (in.), including hub drag

*/

WL_FUS =

/* waterline of center of pressure of fuselage (in.), including hub drag

*/

6.904;

XUU_FUS = -2.322;

/* effective frontal area (ft2) corresponding to profile drag in the x axis

*/

YVV_FUS = -7.849;

/* effective side area (ft2) for sideward flight

*/

ZWW_FUS = -6.960;

/* effective planview area (ft2) for vertical flight and download in hover

*/

/*** Wing Parameters ***/

/*** Note that the R-50 does not have a wing

*/

FS_WN =

0.0;

/* fuselage station of aerodynamic center of wing (in.)

*/

WL_WN =

0.0;

/* waterline location of aerodynamic center of wing (in.)

*/

ZUU_WN =

0.0;

/* (ft2) */

ZUW_WN =

0.0;

/* (ft2) */

ZMAX_WN =

0.0;

/* (ft2) */

B_WN

1.0;

/* Wing span (ft), Arbitrary non-zero number to avoid division by zero

*/

110

Appendix A R-50 Helicopter Data

/*** Horizontal Tail Parameters ***/


FS_HT = 64.50;

/* (in.) fuselage station of aerodynamic center of horizontal tail

WL_HT = 27.30;

/* (in.) waterline location of aerodynamic center of horizontal tail

*/

ZUU_HT =

/* (ft2) effective lift per unit dynamic pressure at zero angle of attack

*/

/* relative to fuselage reference system

*/

0.0;

ZUW_HT = -1.9374;

*/

/* (ft2) effective variation in circulation lift

*/

/* (approx. the neg. product of lift curve slope and surface area)

*/

/* (ft2) max force generated by the horizontal tail when stalled

*/

/* (in.) station location for the effective aerodynamic center of the

*/

/* vertical fin

*/

WL_VT = 29.5;

/* (in.) vertical position of the vertical fin aerodynamic center

*/

YUU_VT =

/* net y-force per unit dynamic press for zero sideslip

*/

/* Assume symmetric airfoil at 0 incidence

*/

ZMAX_HT = -1.2536;
/*** Vertical Tail Parameters ***/
FS_VT = 99.0;

0.0;

YUV_VT = -0.6876;
YMAX_VT = -0.2487;

/* sideforce arising from a side-velocity component

*/

/* (approx. equal to lift curve slope times the net fin area)

*/

/* sets the maximum sideforce generated by the vertical tail at stall

*/

/* (in.) station line of the tail rotor hub

*/

/*** Tail Rotor Parameters ***/


FS_TR = 97.5;
WL_TR = 25.85;

/* (in.) waterline location of the tail rotor hub

*/

R_TR

0.853;

/* radius of the tail rotor in feet

*/

A_TR

3.0;

/* effective lift curve slope of tail rotor

*/

SOL_TR =

0.10807;

RPM_TR = 5400.0;
TWST_TR =

0.0;

/* solidity of the tail rotor (ratio of blade area to disk area)

*/

/* angular velocity of the tail rotor (rev. per minute)

*/

/* effective twist of the tail rotor blade

*/

B_TR

2.0;

/* number of tail rotor blades

*/

C_TR

0.1458;

/* (ft.) tail rotor blade chord

*/

/*** Control Rotor Parameters ***/


R_CR =

1.8537;

/* radius of control rotor in feet

*/

lb_CR =

0.4921;

/* lenght of control rotor blade in feet

*/

A_CR =

2.657;

/* 2*pi/(1+2/AR) and AR = lb_CR*lb_CR/(lb_CR*C_CR)

*/

C_CR =

0.328;

/* chord length in feet

*/

/* control rotor blade inertia in slug*ft^2

*/

/* speed for transition from dihedral wake function in rotor tip

*/

/* path plane dynamics (set empirically)

*/

I_B_CR = 0.04566;
/*** Other Modeling Parameters ***/
VTRANS = 30.0;

111

Appendix B Equations of Unsteady Motion of the Rigid Body

APPENDIX B EQUATIONS OF UNSTEADY M OTION OF THE RIGID BODY


Force and Moment Equations:

X mg sin = m u& E + qw E rv E

(
)
Z + mg cos cos = m (w& + pv qu )
L = I p& I r& + qr (I I ) I pq
M = I q& + rp ( I I ) + I ( p r )
N = I r& I p& + pq (I I ) + I qr

Y + mg cos sin = m v& E + ru E pw E


E

zx

zx

zx

zx

zx

Kinematic Equations:
p = & & sin
q = & cos + & cos sin
r = & cos cos & sin
& = p + (q sin + r cos ) tan
& = q cos r sin
& = (q sin + r cos ) sec

x& E = u E cos cos + v E (sin sin cos cos sin )


+ w E (cos sin cos + sin sin )

y& E = u E cos sin + v E (sin sin sin + cos cos )


+ w E (cos sin sin sin cos )

z& E = u E sin + v E sin cos + w E cos cos

112

Appendix C System and Control Matrices

APPENDIX C SYSTEM AND CONTROL M ATRICES

Xu

Zu

Mu

Iy

Yu

m
F =
I z Lu + I xz Nu

Ic

I L + I N
xz u x u
Ic

0,CR

16 lCR

Xw
m

Xq

Zw
m

Zq

m
m

g cos 0

Xv
m

+ u0

g cos 0 sin 0

Zv
m

Mv
Iy

Mp

Mw
Iy

Mq

cos 0

Yw
m

Yq

g sin 0 sin 0

Yv
m

I z Lw + I xz Nw
Ic

I z Lq + I xz N q

sin 0 tan 0

I xz Lw + I x Nw
Ic

I xz Lq + I x Nq

0
0

Iy

m
Ic

Ic

Xp

w0

Xr
+ v0
m
Z
r
m

Z c
k
m

Mr
Iy

M c
k
Iy

sin 0

g cos 0 cos 0

Yr
u0
m

Yc
k
m

I z Lr + I xz Nr
Ic

I z Lc + I xz Nc
k
Ic

cos 0 tan 0

I xz Lr + I x N r
Ic

I xz Lc + I x Nc
k
Ic


16

m
Zp
m

v0

Iy

Yp
m

+ w0

I z Lv + I xz Nv
Ic

I z Lp + I xz N p

I xz Lv + I x Nv
Ic

I xz Lp + I x N p

Ic

Ic

0,CR


16

16 lCR

0
g sin 0 cos 0

X c
k
m


16
0

Za

k
m

M a

k
Iy

Ya

I z La + I xz Na
k
Ic

I xz La + I x Na
k
Ic

16

X a
k
m

113

Appendix C System and Control Matrices

X e

Ze

M e

Iy

Ye

m
G=
I z Le + I xz N e

Ic

I xz Le + I x N e

Ic

X c
k MR
m
Z c
k MR
m
M c
k MR
Iy

X a
k MR
m
Z a
k MR
m
M a
k MR
Iy

Yc
k MR
m
I z L c + I xz N c
k MR
Ic

Ya
k MR
m
I z La + I xz N a
k MR
Ic

I xz Lc + I x N c
k MR
Ic

k CR
16

I xz L a + I x N a
k MR
Ic

0

k CR
16

Z p

M p

Iy

g sin 0 sin 0

I z Lp + I xz N p

Ic

I z Lp + I xz N p

Ic

X p

x=
p

r

c ,CR
s, CR
e coll , MR

c
long

=
u =
a lat


p coll ,TR

I c = I x I z I xz2

)
long
lat

longitudinal swashplate tilt


lateral swashplate tilt

114

Appendix D Results of Linear System Analysis for Hover

APPENDIX D RESULTS OF LINEAR SYSTEM ANALYSIS FOR HOVER


Control rotor decoupled from rigid body dynamics (rotor rotating clockwise):
F-Matrix:
u

theta

beta_c_cr

phi

beta_s_cr

-0.0469

-0.0296

1.4106

-32.1394

-0.0034

0.21

-0.0311

-0.6892

-0.1084

1.4901

-0.0032

-0.0012

-1.8032

0.1559

-0.068

-5.8228

0.0465

0.7644

theta

0.9984

-0.0561

beta_c_cr

0.0101

-1

-2.1633

-0.0237

0.0047

-0.0039

0.2102

0.0837

-0.0998

-1.443

32.0888

0.4197

0.2176

-0.0148

-2.3849

-0.4276

-17.881

0.428

phi

-0.0026

-0.0464

-0.0116

-0.2397

0.2881

0.1293

-1.7523

beta_s_cr

0.0237

-0.0101

-1

-2.1633

G-Matrix:
trim:
coll MR

B1

A1

coll TR

theta

-18.1857

32.5193

3.4817

phi

-0.0464

-391.015

-1.4548

-0.3507

coll MR

-30.9937

-85.857

-50.0334

B1

-0.0447
-0.0108

0.0561

theta

A1

beta_c_cr

-2.1633

coll TR

0.1085

0.0046

-2.3784

-3.4852

32.5522

-43.3123

a1s

0.0453

-9.0806

-155.97

497.447

-43.3504

b1s

-0.0059

phi

-75.505

180.7083

beta_s_cr

2.1633

Eigenvalues of F:
Re

Im

Damping

Freq. (rad/sec)

Mode

0.0976

1.0065

-0.0965

1.0113

Lateral Oscillation

0.0976

-1.0065

-0.0965

1.0113

0.0102

0.8473

-0.012

0.8474

0.0102

-0.8473

-0.012

0.8474

-0.6887

0.6887

Heave

-1.8383

1.8383

Yaw

-2.1633

2.1633

Long. Control Rotor

-2.1633

2.1633

Lat. Control Rotor

-6.1697

6.1697

Pitch mode

-17.811

17.8108

Roll mode

Longitudinal Oscillation

115

Appendix D Results of Linear System Analysis for Hover

Control rotor coupled with rigid body dynamics (rotor rotating clockwise):
F-Matrix:
u

theta

beta_c_cr

phi

-0.0469

-0.0296

1.4106

-32.1394

-19.8693

-0.0034

0.21

beta_s_cr
2.1273

-0.0311

-0.6892

-0.1084

1.4901

0.8889

-0.0032

-0.0012

-1.8032

-0.2143

0.1559

-0.068

-5.8228

52.4587

0.0465

0.7644

-30.5704

theta

0.9984

-0.0561

beta_c_cr

0.0101

-1

-2.1633

-0.0237

0.0047

-0.0039

0.2102

0.0837

2.1295

-0.0998

-1.443

32.0888

0.4197

19.8894

0.2176

-0.0148

-2.3849

95.3006

-0.4276

-17.881

0.428

303.9402
0

phi

-0.0026

-0.0464

-0.0116

-0.2397

0.2881

0.1293

-1.7523

beta_s_cr

0.0237

-0.0101

-1

-2.1633

G-Matrix:
trim:
coll MR

B1

A1

coll TR

theta

-0.0464

-18.1857

11.2387

1.2033

phi

-391.015

-0.5028

-0.1212

coll MR

0.0561

-30.9937 -29.6722 -17.2916

B1

-0.0447
-0.0108

0.1085

theta

A1

beta_c_cr

-4.2184

coll TR

-2.3784

-1.2045

11.25

-43.312

a1s

0.0453

-9.0806

-53.9049

171.918

-43.35

b1s

-0.0059

phi

-75.505

180.708

beta_s_cr

4.2184

0.0046

Eigenvalues of F:
Re

Im

Damping

Freq. (rad/sec)

Mode

0.0309

0.7663

-0.0402

0.7669

Lateral Oscillation

0.0309

-0.7663

-0.0402

0.7669

-0.0043

0.6423

0.0067

0.6423

-0.0043

-0.6423

0.0067

0.6423

-0.6838

0.6838

Heave

-1.9241

1.9241

Yaw

-4.0211

7.7222

0.4619

8.7064

Pitch + longitudinal CR

-4.0211

-7.7222

0.4619

8.7064

-10.011

15.3329

0.5467

18.3116

-10.011

-15.333

0.5467

18.3116

Longitudinal Oscillation

Roll + lateral CR

116

Appendix D Results of Linear System Analysis for Hover

Control rotor and estimated avionics box included (rotor rotating clockwise):
F-Matrix:
u

theta

beta_c_cr

phi

beta_s_cr

-0.0553

0.0039

1.413

-32.1731

-19.9033

-0.004

0.2104

-0.0027

-0.5727

-0.0236

-0.2358

-0.1233

0.004

-0.0112

-1.3749

0.1946

0.2373

0.002

-6.9424

68.2896

0.0595

0.5964

-32.255
0

theta

2.1309

0.9991

-0.0427

beta_c_cr

0.0101

-1

-2.1633

-0.0237

0.0059

0.0049

0.2104

-0.0101

2.1309

-0.0871

-1.4622

32.1437

0.2913

19.903

0.2791

0.0307

-1.8597

100.6205

-0.6653 -21.5131

0.8169

353.6293
0

phi
r

0.0003

0.0073

0.0022

-0.2894

0.2551

0.2469

-1.4638

0.0237

-0.0101

-1

-2.1633

beta_s_cr

G-Matrix:
trim:
coll MR

B1

A1

coll TR

theta

0.0073

2.3713

11.2579

1.2053

phi

0.0427

-322.9734

0.0698

0.1101

coll MR

0.1291

7.3491 -38.6267

-18.2444

B1

0.004
0.0064

q
theta

A1

beta_c_cr

-4.2184

coll TR

2.9838

-1.2053

11.2578

-31.0981

a1s

-0.0072

18.8329

-56.914 200.0234

-86.3006

b1s

0.009

phi
r

-87.6774

156.1621

4.2184

beta_s_cr

0.0166

Eigenvalues of F:
Re

Im

Damping

Freq. (rad/sec)

Mode

-0.0203

0.702

0.0289

0.7023

Lateral Oscillation

-0.0203

-0.702

0.0289

0.7023

-0.0328

0.5605

0.0584

0.5615 Longitudinal Oscillation

-0.0328

-0.5605

0.0584

0.5615

-0.5749

0.5749

Heave

-1.4704

1.4704

Yaw

-4.4905

8.5089

0.4667

9.6211 Pitch + longitudinal CR

-4.4905

-8.5089

0.4667

9.6211

-11.9143

15.9267

0.599

19.8899

-11.9143

-15.9267

0.599

19.8899

Roll + lateral CR

117

Appendix E R-50 Helicopter System Components

APPENDIX E R-50 HELICOPTER SYSTEM COMPONENTS

Figure E.1: R-50 fully equipped during flight test

Figure E.2: R-50 fully equipped on transport cart; GST ground control station in background

118

Appendix E R-50 Helicopter System Components

Figure E.3: R-50 avionics box with on-board PC and sensor packet

Figure E.4: R-50 horizontal tail for improved handling characteristics in forward flight

119

Appendix F Simulated and Experimental Bode Plots

APPENDIX F SIMULATED AND EXPERIMENTAL BODE P LOTS

Bode Plot, Roll Dynamics, 0.31sec Time Delay


30
Flight Test
Simulation

25

Gain [dB]

20
15
10
5
0
-5
1.0

10.0

100.0

Frequency [rad/sec]

100
0

Phase []

-100
-200
-300
-400
-500
-600
-700
1.0

10.0

100.0

Frequency [rad/sec]

Figure F.1: Experimental and simulated frequency response, 0.31sec time delay, roll dynamics

120

Appendix F Simulated and Experimental Bode Plots


Bode Plot, Yaw Dynamics, 0.31sec Time Delay
( yaw damper incl. in flight test )
50
45

Flight Test

40

Simulation

Gain [dB]

35
30
25
20
15
10
5
0
0.1

1.0

10.0

100.0

Frequency [rad/sec]

0
-50

Phase []

-100
-150
-200
-250
-300
-350
-400
0.1

1.0

10.0

100.0

Frequency [rad/sec]

Figure F.2: Experimental and simulated frequency response, 0.31sec time delay, yaw dynamics

121

You might also like