You are on page 1of 187

The Pennsylvania State University

The Graduate School

Department of Aerospace Engineering

SIMULATION AND CONTROL OF

A HELICOPTER OPERATING IN A SHIP AIRWAKE

A Thesis in
Aerospace Engineering
by
Dooyong Lee

c 2005 Dooyong Lee


Submitted in Partial Fulfillment


of the Requirements
for the Degree of

Doctor of Philosophy

August 2005
The thesis of Dooyong Lee has been reviewed and approved* by the following:

Joseph F. Horn
Assistant Professor of Aerospace Engineering
Thesis Adviser, Chair of Committee

Lyle N. Long
Professor of Aerospace Engineering

Edward C. Smith
Professor of Aerospace Engineering

Qian Wang
Assistant Professor of Mechanical Engineering

George Lesieutre
Professor of Aerospace Engineering
Head of the Department of Aerospace Engineering

*Signatures are on file in the Graduate School.


Abstract

This thesis describes a study in simulation and control of a helicopter operating in proximity

to a ship. The helicopter/ship combination used in the study is a UH-60A helicopter

operating off an LHA class ship. This represents the same aircraft ship combination used in

the JSHIP program. The flight dynamics model is based on the GENHEL software and this

flight dynamics model has been updated to include high-order dynamic inflow model and

gust penetration effects of the ship airwake. To simulate the pilot control inputs for typical

shipboard operations, an optimal control model of the human pilot is developed. The pilot

model can be tuned to achieve different tracking performances based on a desired crossover

frequency in each control axis and is designed to operate over a range of airspeeds using a

simple gain scheduling algorithm. The pilot model is then used to predict pilot workload

for shipboard operations in two different wind-over-deck conditions.

Validation studies are conducted using both time and frequency domain analyses to

understand the impact of a time-varying ship airwake on the pilot control activity for the

approach and departure operations. The pilot control input autospectra predicted from the

simulation model are compared to those of flight test data from the JSHIP program. It

is found that the control activities are similar in low frequency range but underestimate

iii
in magnitude in the high frequency range (over 1.5 Hz). There is clear evidence that the

human pilot is continually moving cyclic stick in the maneuver. At this stage of the study

no attempt has been made to optimize the parameters of the human pilot model.

The paper also discusses the application of a stochastic airwake model for more efficient

simulation. This new airwake model is derived from the simulation with the full CFD

airwake by extracting an equivalent six-dimensional gust vector. The spectral properties

of the gust components are then analyzed, and shaping filters are designed to simulate the

gusts when driven by white noise. It is proposed that the stochastic gust model can be used

to optimize the automatic flight control system in order to improve disturbance rejection

properties of the aircraft.

A stability augmentation system (SAS) is optimized for a UH-60 helicopter operating

in the turbulent ship airwake. For disturbance rejection, a new performance specification is

designed based on the power spectral density of the transfer function from the gust inputs to

aircraft rate responses. The baseline limited authority SAS is modified and optimized using

CONDUIT (Control Designer’s Unified Interface) in order to improve handling-qualities

and stability, and to minimize a weighted objective of gust responses. In addition, a H∞

controller is designed to provide an alternative SAS configuration. The optimized SAS and

H∞ SAS are then tested using the non-linear simulation model with time-varying airwake.

Time domain and frequency domain analyses of the simulation show that the modified SAS

results in significant reduction of pilot workload.

iv
Contents

List of Figures viii

List of Tables xiv

Acknowledgments xv

1 Introduction 1

1.1 Modeling of Helicopter/Ship Dynamic Interface . . . . . . . . . . . . . . . . 4

1.1.1 Helicopter Flight Dynamic Model . . . . . . . . . . . . . . . . . . . . 4

1.1.2 Modeling of Rotor Aerodynamics . . . . . . . . . . . . . . . . . . . . 7

1.1.3 Ship Airwake Model . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

1.1.4 Flight Control Model . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

1.2 Simulation of Helicopter/Ship Dynamic Interface . . . . . . . . . . . . . . . 16

1.3 Problem Statement and Research Objectives . . . . . . . . . . . . . . . . . 18

2 Helicopter Flight Dynamics Model 20

2.1 Overview of the GENHEL Simulation Model . . . . . . . . . . . . . . . . . 20

2.2 MATLAB Implementation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

v
2.3 Peters-He Inflow Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

2.3.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

2.3.2 Basic Equations of Peters-He Inflow Model . . . . . . . . . . . . . . 25

2.4 Gust Penetration Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

3 Pilot Modeling 34

3.1 Optimal Control Model of the Human Pilot . . . . . . . . . . . . . . . . . . 35

3.2 Nonlinear Elements of the Human Pilot . . . . . . . . . . . . . . . . . . . . 41

3.2.1 Hysteresis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

3.2.2 Deadband . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

4 Numerical Examples 44

4.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

4.1.1 Shipboard Departure Trajectory . . . . . . . . . . . . . . . . . . . . 45

4.1.2 Shipboard Approach Trajectory . . . . . . . . . . . . . . . . . . . . . 46

4.2 Effects of Ship Airwake Model . . . . . . . . . . . . . . . . . . . . . . . . . . 48

4.3 Effects of Different Tracking Performance . . . . . . . . . . . . . . . . . . . 65

4.4 Validation with Flight Test Data . . . . . . . . . . . . . . . . . . . . . . . . 70

4.4.1 Frequency Domain Analysis . . . . . . . . . . . . . . . . . . . . . . . 81

5 Task-Tailored Control Design 85

5.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85

5.2 Stochastic Airwake Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . 88

5.3 Optimization of a Stability Augmentation System . . . . . . . . . . . . . . . 110

vi
5.4 H∞ Control of a Helicopter SAS . . . . . . . . . . . . . . . . . . . . . . . . 128

5.4.1 Review of H∞ Control Design Method . . . . . . . . . . . . . . . . . 129

5.4.2 Design of H∞ Controller for a Helicopter SAS . . . . . . . . . . . . . 135

6 Conclusions and Future Works 154

6.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154

6.2 Recommendations for Future Work . . . . . . . . . . . . . . . . . . . . . . . 157

Bibliography 160

vii
List of Figures

1.1 Typical WOD envelope (Ref. [2]) . . . . . . . . . . . . . . . . . . . . . . . . 2

1.2 The modeling components of a helicopter (Ref. [3]) . . . . . . . . . . . . . . 4

1.3 Momentum theory flow model for axial flight (Ref. [3]) . . . . . . . . . . . . 8

1.4 Block diagram of coupled rotor and induced flow dynamics (Ref. [14]) . . . 9

1.5 Schematic of helicopter control system . . . . . . . . . . . . . . . . . . . . . 14

2.1 Block diagram of GENHEL flight simulation model (Ref. [11]) . . . . . . . 21

2.2 Overall structure of MATLAB based simulation program . . . . . . . . . . . 23

2.3 Comparisons of inflow ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

2.4 Vorticity magnitude iso-surface at t = 40 sec (Ref. [39]) . . . . . . . . . . . 31

2.5 Gust penetration model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

2.6 The approach of the overlapped time history of airwake . . . . . . . . . . . 33

3.1 Optimal control model of the human pilot . . . . . . . . . . . . . . . . . . . 35

3.2 Augmented plant model in longitudinal axis . . . . . . . . . . . . . . . . . . 37

3.3 Effect of a sine wave passing through a hysteresis . . . . . . . . . . . . . . . 42

3.4 Effect of a sine wave passing through a deadband . . . . . . . . . . . . . . . 43

4.1 Top view of an LHA class ship . . . . . . . . . . . . . . . . . . . . . . . . . 45

viii
4.2 Shipboard approach operation procedures . . . . . . . . . . . . . . . . . . . 47

4.3 Helicopter position w.r.t. ship coordinate system - Departure task . . . . . 50

4.4 Helicopter velocity [ft/sec] - Departure task (30 knot, 0 degree WOD condition) 51

4.5 Helicopter attitude angles [deg] in the DI mesh - Departure task (30 knot, 0

degree WOD condition) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

4.6 Pilot inputs [%] in the DI mesh - Departure task (30 knot, 0 degree WOD

condition) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

4.7 Helicopter velocity [ft/sec] - Departure task (30 knot, 30 degree WOD con-

dition) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

4.8 Helicopter attitude angles [deg] in the DI mesh - Departure task (30 knot,

30 degree WOD condition) . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

4.9 Pilot inputs [%] in the DI mesh - Departure task (30 knot, 30 degree WOD

condition) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

4.10 Helicopter position w.r.t. ship coordinate system - Approach task . . . . . . 58

4.11 Helicopter velocity [ft/sec] - Approach task (30 knot, 0 degree WOD condition) 59

4.12 Helicopter attitude angles [deg] in the DI mesh - Approach task (30 knot, 0

degree WOD condition) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

4.13 Pilot inputs [%] in the DI mesh - Approach task (30 knot, 0 degree WOD

condition) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

4.14 Helicopter velocity [ft/sec] - Approach task (30 knot, 30 degree WOD condition) 62

4.15 Helicopter attitude angles [deg] in the DI mesh - Approach task (30 knot, 30

degree WOD condition) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

ix
4.16 Pilot inputs [%] in the DI mesh - Approach task (30 knot, 30 degree WOD

condition) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

4.17 Helicopter position error [ft] - 30 knot, 0 degree WOD condition . . . . . . 66

4.18 Pilot control input [%] - 30 knot, 0 degree WOD condition . . . . . . . . . . 67

4.19 Helicopter position error [ft] - 30 knot, 30 degree WOD condition . . . . . . 68

4.20 Pilot control input [%] - 30 knot, 30 degree WOD condition . . . . . . . . . 69

4.21 Helicopter airspeed [knot] - 30 knot, 0 degree WOD condition . . . . . . . . 73

4.22 Helicopter altitude [ft] - 30knot, 0 degree WOD condition . . . . . . . . . . 74

4.23 Angular rate [deg/sec] - 30knot, 0 degree WOD condition . . . . . . . . . . 75

4.24 Pilot stick inputs [%] - 30 knot, 0 degree WOD condition . . . . . . . . . . 76

4.25 Helicopter airspeed [knot] - 30 knot, 30 degree WOD condition . . . . . . . 77

4.26 Helicopter altitude [ft] - 30 knot, 30 degree WOD condition . . . . . . . . . 78

4.27 Angular rate [deg/sec] - 30knot, 30 degree WOD condition . . . . . . . . . . 79

4.28 Pilot stick inputs [%] - 30 knot, 30 degree WOD condition . . . . . . . . . . 80

4.29 Pilot input autospectrum [dB] - 30 knot, 0 degree WOD condition . . . . . 83

4.30 Pilot input autospectrum [dB] - 30 knot, 30 degree WOD condition . . . . . 84

5.1 Task-tailored control system design scheme . . . . . . . . . . . . . . . . . . 87

5.2 Derivation of stochastic airwake disturbances . . . . . . . . . . . . . . . . . 91

5.3 Comparisons of aircraft angular rates [dB] (time-varying airwake vs. equiv-

alent airwake) - 0 degree WOD condition . . . . . . . . . . . . . . . . . . . 93

5.4 Comparisons of aircraft angular rates [dB] (time-varying airwake vs. equiv-

alent airwake) - 30 degree WOD condition . . . . . . . . . . . . . . . . . . . 94

x
5.5 Comparisons of pilot inputs [dB] (time-varying airwake vs. equivalent air-

wake) - 0 degree WOD condition . . . . . . . . . . . . . . . . . . . . . . . . 95

5.6 Comparisons of pilot inputs [dB] (time-varying airwake vs. equivalent air-

wake) - 30 degree WOD condition . . . . . . . . . . . . . . . . . . . . . . . 96

5.7 Power spectral density for vertical airwake disturbance component . . . . . 98

5.8 Power spectral density of longitudinal airwake disturbance component . . . 99

5.9 Power spectral density of lateral airwake disturbance component . . . . . . 100

5.10 Power spectral density of vertical airwake disturbance component . . . . . . 101

5.11 Power spectral density of roll airwake disturbance component . . . . . . . . 102

5.12 Power spectral density of pitch airwake disturbance component . . . . . . . 103

5.13 Power spectral density of yaw airwake disturbance component . . . . . . . . 104

5.14 Comparisons of aircraft angular rates [dB] (time-varying airwake vs. equiv-

alent airwake vs. stochastic airwake) - 30 knot, 0 degree WOD condition . . 106

5.15 Comparisons of aircraft angular rates [dB] (time-varying airwake vs. equiv-

alent airwake vs. stochastic airwake) - 30 knot, 30 degree WOD condition . 107

5.16 Comparisons of pilot inputs [dB] (time-varying airwake vs. equivalent air-

wake vs. stochastic airwake) - 30 knot, 0 degree WOD condition . . . . . . 108

5.17 Comparisons of pilot inputs [dB] (time-varying airwake vs. equivalent air-

wake vs. stochastic airwake) - 30 knot, 30 degree WOD condition . . . . . . 109

5.18 Augmented plant model for a SAS optimization . . . . . . . . . . . . . . . . 112

5.19 Modified SAS configuration . . . . . . . . . . . . . . . . . . . . . . . . . . . 113

5.20 A new disturbance rejection spec design (ex. pitch axis) . . . . . . . . . . . 115

5.21 HQ windows for the original SAS configuration - 30 knot, 30 degree WOD . 116

xi
5.22 HQ windows for the optimized SAS configuration - 30 knot, 30 degree WOD 117

5.23 Aircraft angular rate responses [deg/sec] - 30 knot, 0 degree WOD . . . . . 120

5.24 Pilot control stick inputs [%] - 30 knot, 0 degree WOD . . . . . . . . . . . . 121

5.25 SAS outputs [%] - 30 knot, 0 degree WOD . . . . . . . . . . . . . . . . . . . 122

5.26 Aircraft angular rate responses [deg/sec] - 30 knot, 30 degree WOD . . . . . 123

5.27 Pilot control stick inputs [%] - 30 knot, 30 degree WOD . . . . . . . . . . . 124

5.28 SAS outputs [%] - 30 knot, 30 degree WOD . . . . . . . . . . . . . . . . . . 125

5.29 Control stick input autospectra [dB] - 30 knot, 0 degree WOD . . . . . . . . 126

5.30 Control stick input autospectra [dB] - 30 knot, 30 degree WOD . . . . . . . 127

5.31 Standard compensator configuration . . . . . . . . . . . . . . . . . . . . . . 131

5.32 Additive perturbation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132

5.33 Disturbance rejection at the plant output . . . . . . . . . . . . . . . . . . . 132

5.34 General block diagram of the mixed sensitivity problem . . . . . . . . . . . 134

5.35 Augmented system for airwake disturbance rejection . . . . . . . . . . . . . 136

5.36 Magnitude of weighting functions We and Wu . . . . . . . . . . . . . . . . . 139

5.37 Singular values of controller K∞ . . . . . . . . . . . . . . . . . . . . . . . . 140

5.38 Aircraft position w.r.t. the spot 8 [ft] - 30 knot, 0 degree WOD . . . . . . . 142

5.39 Aircraft angular rate responses [deg/sec] - 30 knot, 0 degree WOD . . . . . 143

5.40 Aircraft attitude responses [degree] - 30 knot, 0 degree WOD . . . . . . . . 144

5.41 Pilot control stick inputs [%] - 30 knot, 0 degree WOD . . . . . . . . . . . . 145

5.42 SAS outputs [%] - 30 knot, 0 degree WOD . . . . . . . . . . . . . . . . . . . 146

5.43 Aircraft position w.r.t. the spot 8 [ft] - 30 knot, 30 degree WOD . . . . . . 147

5.44 Aircraft angular rate responses [deg/sec] - 30 knot, 30 degree WOD . . . . . 148

xii
5.45 Aircraft attitude responses [degree] - 30 knot, 30 degree WOD . . . . . . . . 149

5.46 Pilot control stick inputs [%] - 30 knot, 30 degree WOD . . . . . . . . . . . 150

5.47 SAS outputs [%] - 30 knot, 30 degree WOD . . . . . . . . . . . . . . . . . . 151

5.48 Control stick input autospectra [dB] - 30 knot, 0 degree WOD . . . . . . . . 152

5.49 Control stick input autospectra [dB] - 30 knot, 30 degree WOD . . . . . . . 153

xiii
List of Tables

4.1 Initial profile parameters for the departure task . . . . . . . . . . . . . . . . 46

4.2 Initial profile parameters for the approach task . . . . . . . . . . . . . . . . 48

4.3 Crossover frequencies for different tracking performance (rad/sec) . . . . . . 65

4.4 Initial profile parameters for the approach tasks from JSHIP program . . . 70

5.1 Gust shaping filters for 0 degree and 30 degree WOD conditions . . . . . . 105

xiv
Acknowledgments

I would first like to express my deepest appreciation to all the committee members for

participating on the examination committee. Many thanks go to Professor Lyle N. Long

and fellow graduate student Nilay Sezer-Uzol for sharing their experiences and support of

ship airwake data. Especially, I gratefully thank Professor Joseph F. Horn, my advisor, for

his time, guidance, and patience, which have been uniquely instrumental in the successful

completion of this thesis.

I would also like to thank CDR Kevin J. Delamer, USN and Colin Wilkinson for their

support with regard to the JHSIP flight test data. Dr. Mark B. Tischler, US Army/NASA

Rotorcraft Division, was deeply appreciated for providing the CIFER and CONDUIT soft-

ware. Fellow graduate students Jun-Sik Kim, Nilesh Sahani, Derek Brigdes, Brian Geiger,

Youngtae Ahn, are thanked for their friendship and technical advice.

I am deeply indebted to my family for all the years of support and encouragement that

led to the completion of this thesis. Endless thanks to my parents, my parents-in-law, my

sister and the family of my brother. Finally, I dedicated this thesis to my wife, Kyoungsun

Moon, whose love and support gave me the strength of mind to complete this work on time.

To my daughters, Michelle and Jennifer, I hope this thesis makes you proud.

xv
Chapter 1

Introduction

Many military and commercial helicopters need to be launched and recovered from a ship

at sea. There are several distinctive problems associated with ship-based helicopters that

can limit their operational capability. For example, the pilot has to takeoff and land the he-

licopter from a moving flight deck within the turbulent airwake of the ship’s superstructure.

Since these environments increase the difficulty of the helicopter shipboard operations, the

shipboard operation is one of the most challenging, training intensive and dangerous of all

helicopter flight operations. To assure the compatibility of any helicopter/ship combination,

a set of limits must be established to define the envelopes which safe launch and recovery

can take place. Currently, those limits are determined by conducting flight tests at sea

using particular helicopter and ship combination.

The so called “Helicopter/Ship Dynamic Interface Testing” is a comprehensive term

referring to investigation of all aspects relating to the effect of ship presence on embarked

helicopter shipboard operations [1]. Typically, the dynamic interface tests include analysis

and quantification of the approach, hover, and departure operations under various shipboard

1
CHAPTER 1. INTRODUCTION 2

flight conditions. For example, during hover and landing, the objectives are to investigate

the effects of turbulence and the ship motion and recovery assist system. The dynamic

interface tests are also used to evaluate the adequacy and safety of shipboard aviation

facilities and procedures.

An important issue of the dynamic interface testing is to establish the wind-over-deck

(WOD) flight envelope. The WOD flight envelope depends highly on extent of the winds

encountered. Low winds result in a small envelope, even if the real envelope may be much

larger. Before a WOD envelope is established, a very restricted ‘general envelope’, which

results in a reduction in helicopter/ship operational capability, is used to determine safe

operating conditions [2]. Figure 1.1 shows the general envelope and a typical envelope

expanded by dynamic interface testing.

Currently the only method for determining the WOD flight envelope for U.S. military is

45 10

330 40
20
35
30
1

320
2
25 3
Expanded envelope
through DI testing
20
310 4 3A

15
5
45
10
6
5 60
General envelope
7

270 90

8 9

Figure 1.1: Typical WOD envelope (Ref. [2])


CHAPTER 1. INTRODUCTION 3

by actual flight test. This method of flight envelope definition requires a significant amount

of time, expense, resources, and is limited by the weather and sea conditions. Thus, the

need has arisen to use better simulation tools for analyzing shipboard operations to reduce

the flight test time and cost to establish safe operating envelopes.

Over the last few years, numerous efforts have been devoted to develop helicopter/ship

dynamic interface simulation tools. Such a simulation tool can be used to estimate the best

approach and departure paths, provide improved real-time training simulator for pilots,

and ultimately be used in the design or acceptance testing of future helicopter. Modeling

and simulation of the helicopter/ship dynamic interface is a challenging technical problem.

Rotorcraft themselves are complex and highly nonlinear dynamic systems, but shipboard

operations present further complexity. There are several considerations when modeling the

shipboard launch and recovery flight tasks:

1. Modeling the helicopter flight dynamics.

2. Defining the trajectory of the helicopter.

3. Modeling the pilot feedback loop required to fly this trajectory.

4. Modeling the influence of ship airwake on the helicopter.

5. Modeling the effect of atmospheric turbulence on the helicopter.

6. Modeling the motion of the ship for the given sea state and its influence on the

helicopter and the pilot.

7. Modeling the visual, aural, and motion cues for DI testing.

In this study, the first four of these considerations will be addressed.


CHAPTER 1. INTRODUCTION 4

1.1 Modeling of Helicopter/Ship Dynamic Interface

1.1.1 Helicopter Flight Dynamic Model

The simulation models that are required to describe the flight behavior of the helicopter

include kinematics, dynamics, and aerodynamics of the helicopter subsystems (main rotor,

fuselage, empennage, tail rotor, power plant, and primary flight control system). Since the

dynamics of a helicopter is highly complex and nonlinear (Figure 1.2), obtaining an accurate

mathematical model is a very challenging task.

In general, the equations governing the dynamics of these components can be developed

from the application of physical laws, e.g., Newton’s laws of motion, conservation of energy,

to the individual components. The basic formulation can be found in various textbooks

[3, 4, 5].

For the special case where only the six rigid body degrees of freedom (DOF) are

considered, the state variables compromise the three translational velocity components

(uf , vf , wf ), the three rotational velocity components (pf , qf , rf ), and the three Eu-

Main rotor
flap, lag, pitch, and wake

Rotor downwash
On Empennage and Tail rotor Rotor Downwash
On Fuselage

Fuselage Wake
On Empennage

Figure 1.2: The modeling components of a helicopter (Ref. [3])


CHAPTER 1. INTRODUCTION 5

ler angles (φ, θ, ψ). The resulting equations of motions are :

X
u̇f = − qf wf + rf vf − g sin θ
Mh
Y
v̇f = − rf uf + pf wf + g cos θ sin φ (1.1)
Mh
Z
ẇf = − pf vf + qf uf + g cos θ cos φ
Mh

where X, Y, Z are the external forces acting on the center of mass, g is the gravitational

acceleration, and Mh is the total mass of the helicopter.

L Iy − Iz Ixz
ṗf = + qf rf + (ṙf + pf qf )
Ix Ix Ix
M Iz − Ix Ixz 2
q̇f = + pf rf + (r − p2f ) (1.2)
Iy Iy Iy f
N Ix − Iy Ixz
ṙf = + pf qf + (ṗf − rf qf )
Iz Iz Iz

where L, M, N are the external moments about the center of mass, Ix , Iy , Iz , Ixz are the

moments and product of inertia respectively.

The Euler attitude angles add to the equations of motion through the kinematic rela-

tionship between the fuselage angular rates and the rates of change of the Euler angles.

Using 3-2-1 sequence (yaw-pitch-roll), the kinematic relations are

pf = φ̇ − ψ̇ sin θ

qf = θ̇ cos φ + ψ̇ sin φ cos θ (1.3)

rf = −θ̇ sin φ + ψ̇ cos φ cos θ


CHAPTER 1. INTRODUCTION 6

It is important to note that this six DOF model, while itself complex and widely used, is

still an approximation to the helicopter behavior. All higher degrees of freedom, associated

with the main rotor, powerplant/transmission, control system and the disturbed airflow,

are all represented in a common quasi-steady manner in the equations, having lost their

own individual dynamics and independence in the model reduction. However, this process

is a common feature of flight dynamics, in the search for simplicity to enhance physical

understanding and ease the computational load.

Typically, the reduced dynamic model is widely used in the field of controller design. For

example, a six degree of freedom linear model of the Bell 205 helicopter was used to describe

a typical steady hover condition [6]. Postlethwaite et al validated linear model of the Bell

205 helicopter against flight test with integration of uncertainty in the specific frequency

range [7]. Frost et al [8] used 6 DOF and 10 DOF linear models of the unaugmented UH-60

at a variety of flight conditions. These models had been previously identified from flight test

data using the Comprehensive Identification from Frequency Response (CIFER) software.

A more complex linear model was applied by Takahashi [9]. The total 23 states linear

model represented the helicopter as a six DOF rigid fuselage with rigid rotor blades each

with a flap and lag DOF. Lead-lag damper models were also included. Rotor RPM DOF

and engine-governor dynamics were not included in this model. Linear two dimensional

quasi-steady theory was used to model the rotor blade aerodynamic forces and a three state

Pitt-Peters dynamic inflow model was used to describe the unsteady wake effects.

A 40 states linear model was used as the basic helicopter dynamic model by Ingle

et al [10]. In this analysis, basic model consists of fuselage rigid body modes, flap and lag

dynamics, a simplified representation of blade torsion, first harmonic dynamic inflow for the
CHAPTER 1. INTRODUCTION 7

main rotor, tail rotor dynamic inflow, and an approximation for the delay of the downwash

effects on the tail rotor and empennage. The rotor speed was assumed constant and the

quasi-steady blade element aerodynamics included the effects of compressibility and stall.

Finally, a 32 state model was augmented with an 8-state model of the actuator dynamics.

Several nonlinear helicopter dynamic models have been developed and can be found in

public domain. Perhaps one that is most widely used in dynamic interface simulation is the

GENHEL (General Helicopter) flight dynamics simulation code. The GENHEL provides

operational and verified engineering simulation of the UH-60 Black Hawk helicopter [11].

This work was originally developed by Sikorsky Aircraft and documented under contract

from NASA. The solution in terms of helicopter motion is obtained iteratively by summing

the forces and moments acting at the helicopter center of gravity and subsequently obtaining

the accelerations. The helicopter model is divided into components for the purpose of

modeling the aerodynamics(the main rotor, fuselage, empennage, tail rotor). The detailed

definitions of each component are given in Reference [11].

1.1.2 Modeling of Rotor Aerodynamics

The unique problem for helicopter flight simulation is the main rotor unsteady aerodynam-

ics. The presence of compressibility effects and an unsteady rotor flow field, even when

the helicopter is moving with uniform speed, makes the analysis of helicopter motion very

complex. Typical issues on rotor aerodynamics include the rotor wake and the ground effect.

The simplest representation of the rotor wake is based on the momentum theory, utilizing

the conservation laws of mass, momentum and energy. The rotor inflow is assumed to be

steady, inviscid and incompressible with a well defined slipstream between the flow field
CHAPTER 1. INTRODUCTION 8

generated by the actuator disc and the external flow (Figure 1.3). Further assumptions are

discussed in detail by Johnson [4] and Bramwell [5].

The early representations of unsteady rotor aerodynamics reduced the main rotor’s flow

field to two dimensions. The classic 2D unsteady approximations are the Theodorsen and

Sears functions [4]. The Theodorsen function describes an airfoil oscillation in pitch and

plunge. The Sears function describes an airfoil with a transverse harmonic gust. These func-

tions were derived from a non-rotating reference frame. However, they provide a convenient

reference for rotating flows.

While changing to rotational coordinates creates some confusion as to how boundary

conditions are referenced and measured, Johnson [12] showed the proper derivation of the

boundary conditions. Loewy developed a 2D representation of a rotating rotor’s unsteady

flow field with harmonically occurring blade passages [13]. The lift deficiency function

resembles the Theodorsen function. For high inflow rates, the Loewy function is approaching

the Theodorsen results. The most important result from Loewy is that the wake geometry

and phasing is the primary cause of unsteady rotor loading.

v=0 v = Vc v = Vd - 2vi

v = Vd - vi
v = vi v = Vc + vi

v = 2vi v = Vc + 2vi v = Vd

(a) Hover (b) Climb (c) Descent

Figure 1.3: Momentum theory flow model for axial flight (Ref. [3])
CHAPTER 1. INTRODUCTION 9

Miller proposed a three dimensional rotating rotor wake theory [4]. The theory rep-

resents the wake as a cylindrical shell of shed vorticity. Johnson discusses some of the

implications of Miller’s results [4]. While the Miller’s theory represents a three dimensional

rotor wake, it can not predict unsteady rotor performance with forward motion.

Numerous people suggested improvements to the Miller’s method based on harmonic

theory. Pitt and Peters showed a model based on coupled inflow and harmonic blade theories

[14, 15]. The Pitt-Peters model stands out as a premier dynamic inflow model within the

conceptual framework of a global approximation (Figure 1.4). It has been verified on the

basis of flap response data. It can be easily adapted in nonlinear version for use in time

history solutions.

In the most recent work, Peters and He have extended the modeling to an unsteady three-

dimensional finite-state wake [16, 17, 18], that holds the traditional theories of Theodorsen

and Loewy. The Peters-He finite-state dynamic inflow theory, also called the generalized

dynamic wake theory, is characterized as representation of induced inflow as dynamic degrees

of freedom in a system of first order differential equations in the time domain. Due to

INDUCED
Inflow
FLOW
THEORY

+
+ Angle of Attack LIFTING Circulation and Loads
THEORY
+

Blade Motions BLADE


DYNAMICS

BODY
DYNAMICS

Figure 1.4: Block diagram of coupled rotor and induced flow dynamics (Ref. [14])
CHAPTER 1. INTRODUCTION 10

its dynamic nature and computational efficiency, the Peters-He model is finding a wide

application in flight dynamics and aeroelasticity analyses of helicopter. Especially, it has

been implemented in major flight simulation programs currently used in helicopter industry,

such as GENHEL, FLIGHTLAB, etc. In addition, this finite-state inflow model has been

updated using parameter estimation technique by Krämer and Gimonet [19]. The finite-

state wake model was validated using measured flight test data for BO105 helicopter by

Hamers and Basset [20].

Operating helicopters close to a ship deck introduces a range of special characteristics

in the flight dynamics behavior since the downwash field is strongly altered in order to meet

the non-penetration boundary condition at the solid surface. The most significant is the

effect on the induced velocity at the rotor and hence, the rotor thrust, hub moments and

power required. In general, a helicopter rotor within proximity to ship deck can be subject

to various kinds of ground effect, such as inclined ground effect, partial ground effect, and

dynamic ground effect. These ground effects have been examined using both empirical

methods and analytical method of mirror-image rotors [4].

Recently, Zhang, Prasad and Peters used an image method to develop a finite-state

dynamic inflow model for the in-ground-effect of inclined surface [21, 22]. A simple ground

effect model for implementation in real time simulations as was extracted from the analysis

of the aerodynamic interaction results obtained from a computationally intensive method

using spline curve fitting method [23]. Xin and Prasad developed the dynamic ground effect

model [24]. The partial ground effect was been examined using finite-state dynamic inflow

model by Xin [25]. Xin, He and Lee introduced the panel ship deck model for partial ground

effect [26].
CHAPTER 1. INTRODUCTION 11

1.1.3 Ship Airwake Model

The simulation of the helicopter in itself is a challenge. The response of helicopter to tur-

bulence in particular is more complex. Helicopters always operate in the lowest part of the

atmosphere, where the turbulence length scale is relatively small, and due to the fact that

the lifting surface moves through the local atmosphere, the effects of the disturbances are

critical. Because of this, the modeling of the ship airwake and its effect on helicopter be-

havior is considered one of the most significant technical challenges. General characteristics

of ship airwake flow field are unsteadiness, vorticity, large regions of separated flow, and

low Mach number.

For many years, numerous studies have been focused on investigation of the ship airwake.

In general, there are three ways to investigate the ship airwake, numerical simulation,

model-scale testing, and full-scale testing. Since all the relevant aerodynamic qualities of

the atmosphere and geometric qualities of the ship are measured, full-scale testing is most

accurate approach. However, this approach is limited by cost and environmental testing

condition. Model-scale testing is more affordable and offers a controlled testing environment,

but both accurate simulation of the atmosphere and geometric characteristics of the ship

are difficult due to environmental and scale errors. A brief history of those experimental

testing is presented in References [27, 28, 29].

In past years, many researchers have studied numerous methods for the ship airwake

CFD modeling with different classes of ships. Tai and Carico calculated the ship airwake

about a simplified DD class ship using a thin-layer Navier-Stokes method [30]. The airwake

around an LHD class ship was also calculated with the same Navier-Stokes method by Tai
CHAPTER 1. INTRODUCTION 12

[31]. Modi et al calculated the airwake around the general ship shape using the parallel

unstructured flow solver PUMA [32]. To describe the unsteady flow field around the ship,

turbulent velocity model has been developed and combined with steady CFD solutions

[27, 33, 34, 35]. Modeling of the turbulent airwake velocity component is achieved by

passing independent white noise processes through spectral filters whose transfer functions

yield the desired forms of power spectral density of experimental data. While this approach

provides good approximation for the characteristics of unsteady airwake, it is still involving

the experimental test.

Latest advances in the field of CFD make it possible to simulate the full scale flow

field with unsteady turbulence around the ship. Recently, time-accurate computational

CFD were performed at the Naval Air Warfare Center/Aircraft Division to characterize the

unsteady nature of the airwake produced by a LHA class ship [36, 37]. In these works,

the parallel unstructured flow solver COBALT was used to calculate the full-scale flow

field with different numerical methods. It was shown that the full-scale solutions could

predict the dominant frequencies in the unsteady flow field. The airwake around the same

LHA class ship was calculated using the PUMA2 [38, 39]. It was integrated with a finite

volume formulation of the Euler/Navier-Stokes equations for 3D, internal and external,

non-reacting, compressible, steady/unsteady solutions for complex geometries.

However, these CFD simulations produced a large amount of data. For example, with

three velocity components, 40 seconds of data sampled every 0.1 second, for 55890 points

in the region around the ship, the total number of values stored for a single relative wind

condition was 67,068,000 [39]. Thus, data storage requirements are extensive, making real-

time implementation somewhat difficult. A number of recent studies have shown that
CHAPTER 1. INTRODUCTION 13

an equivalent disturbance model (e.g. stochastic gust model) can be used to investigate

the helicopter gust response [33, 34, 40]. These studies paid attention to understanding

the aircraft responses with turbulence models more representative of helicopter operating

environments (e.g. nap-of-the-earth and helicopter/ship DI testing). In fact, pilot comments

indicated that there was no significant difference between the complex CFD models and

the simple stochastic models, thus the stochastic gust model is good enough to provide

a reasonably accurate aircraft gust response [34]. The implication of this observation is

that simple airwake representation can be used for real-time application. Furthermore, a

simplified model that incorporates the statistical characteristics of the turbulent airwake

may provide some insight to the effects of the airwake and assist in the design of future

flight control systems.

1.1.4 Flight Control Model

A brief history of helicopter control systems is presented in Reference[41]. The control

problem of helicopters is a challenging task because the helicopter dynamics are highly

nonlinear, inherently unstable, fully coupled, and subject to parametric uncertainties. The

control of a helicopter is a truly multivariable problem in which one usually considers four

inputs and four outputs with significant interaxis coupling. True multivariable analysis

and design for helicopter control system can possibly provide improved performance in the

on-axis loops while offering good decoupling behavior. Moreover, the use of modern multi-

variable control analysis tools allows a more rigorous analysis and, consequently, improves

stability robustness of the closed-loop system.

In past decade, numerous efforts have been focused on multivariable design of helicopter
CHAPTER 1. INTRODUCTION 14

flight control systems for robustness. Manness et al used the eigenstructure assignment

method to meet response-type and dynamic response requirement described by the handling

quality criteria [42].

Rotor-state feedback technique was investigated by Takahashi [9]. It was focused on

high-gain feedback in roll and pitch axes of helicopter in hovering condition. It was shown

that roll and pitch dynamics have second-order behavior. Hess applied quantitative feedback

theory (QFT) to the design of the longitudinal flight control system for a linear model of

the BO-105C helicopter [43]. This work points out the fact that the inclusion of rotor and

actuator dynamics, while obviously essential to a practical design, does not alter the basic

design procedure of QFT.

Bogdanov et al introduced the model predictive neural control design [44]. In this work,

model predictive control was integrated with neural network feedback controller in combi-

nation of linear quadratic controller. Recently, autonomous adaptive flight control systems

have been developed for an unmanned helicopter by Hovakimyan et al[45], Corban[46],

Johnson[47], and Krupadanam[48]. These approaches show that adaptive controller can

provide a stable, robust, and significant improvement in performance over other control

designs.

Helicopter
Reference
Controller Dynamic
Model

Estimator

Figure 1.5: Schematic of helicopter control system


CHAPTER 1. INTRODUCTION 15

The inverse simulation problem has been discussed for a possible solution to determine

the control inputs which enable to complete some specified maneuver [49, 50, 51]. The

integration based inverse simulation method in Reference [50] involves a numerical differ-

entiation of the output vector with respect to the input vector for calculating the Jacobian

matrix. Since this approach requires Newton iterations at each time step, the computational

expense increases exponentially when the helicopter model becomes more complex.

Avanzini et al proposed two-timescale inverse simulation technique [51]. In this work,

the timescale was divided into slow timescale (collective) and fast timescale (cyclic and

pedal). While this approach significantly reduces computational time, it is not appropriate

for long time duration due to accumulation of error. Alternatively, Xin et al [49] developed

the combined technique for inverse simulation by modifying and combining the integration

and differentiation based inverse simulation methods in such a way that the numerical dif-

ferentiation and Newton iterations were only performed at selected points instead of at each

time step. In addition, inverse simulation method was integrated with on-line compensator

to eliminate steady-state error. But inverse simulation can be still time consuming and

difficult to implement computationally.

Another issue on helicopter flight controller design is modeling of the human pilot. The

development of a pilot model as a dynamic control element that can replace the pilot-in-the-

loop simulations for workload investigations and handling qualities offers various important

advantages, such as cost and test time. Moreover, any analysis can be done without the

ambiguities and variations in piloted simulation or flight tests.

Early research on the human pilot model was devoted to understanding the characteris-

tics of the human as a controller of single input, single output linear time-invariant systems.
CHAPTER 1. INTRODUCTION 16

McRuer et al used a set of quasi-linear models that are adept at predicting human behavior.

The quasi-linear model, so-called crossover model, is very useful for analyzing closed-loop

compensatory tracking or state regulation tasks in which human operator attempts to min-

imize some displayed system error [52]. Alternatively, Bradley and Turner have developed

a general pilot model called SyCos (Synthesis through Constrained Simulation) for Lynx

MK3 helicopter [53, 54]. SyCos model includes the liner time-invariant inverse model and

crossover model. The inverse model represents the pilot’s adaptation to the helicopter’s

dynamics. In the crossover model, the pilot adjusts his behavior to compensate for the

perceived dynamics of the system being controlled. Heffley et al [55] and Lee et al [56]

have developed a pilot model using classical control techniques. These studies shows that

the pilot control inputs can be determined using forward simulation in conjunction with a

feedback controller for a given desired trajectories.

1.2 Simulation of Helicopter/Ship Dynamic Interface

Computer simulation of the dynamic interface provides an alternative to the current heli-

copter/ship system development and testing scenario, predictions of system performance.

Flight envelopes can be also estimated through simulation. All environmental conditions

can be specified, including sea state, and winds. In urgent operational scenarios, an enve-

lope estimated through simulation provides good approximation to support operating ships

and aircraft in previously untested conditions or combinations.

Over the past years, there is a wide range of activities associated with simulation of

helicopter/ship dynamic interface problem. Mello et al provided a brief insight into the
CHAPTER 1. INTRODUCTION 17

relevant helicopter/ship aerodynamic interaction phenomena such as ship ground effect,

airwake effect on trimmed control positions [57]. Bradely and Turner have investigated the

pilot control activity with SyCos pilot model [53], [54]. ART (Advanced Rotorcraft Tech-

nology Inc.) has developed a simulation tool for helicopter/ship dynamic interface testing

[28], [49]. The trajectory generation tool was developed for ship deck landing operations of

a VTOL by Avanzini et al [51]. Benefits of a pilot assisted landing system was demonstrated

by Perrins and Howitt [58]. Colwell provided the effects of flight deck motion in high seas on

the hovering helicopter with Fourier and correlation analysis [59]. Lee et al have developed

a simulation program for helicopter/ship dynamic interface testing [38], [39].

The Joint Shipboard Helicopter Integration Process (JSHIP) has been applied to in-

crease the interoperability of joint shipboard helicopter operations for helicopters that are

not specifically designed to go aboard Navy ships [2], [60], [61]. As a part of JSHIP, the

Dynamic Interface Modeling and Simulation System (DIMSS) was established to define and

evaluate a process for developing WOD flight envelopes. Using the DIMSS, the fidelity stan-

dards for the shipboard launch and recovery tasks have been discussed for combination of an

LHA class ship and the UH-60. The goal of this simulation tool is to prove the process for

determining wind-over-deck launch and recovery envelopes using piloted flight simulation.

In order to validate the fidelity of this simulated dynamic interface, experimental trials have

been conducted.

Although numerous studies have focused on simulation of dynamic interface testing,

much additional work is required to provide a high fidelity simulation program.


CHAPTER 1. INTRODUCTION 18

1.3 Problem Statement and Research Objectives

Considerable work has done on the helicopter/ship dynamic interface simulation tool devel-

opment during past decades. However, there is relatively little work to provide an accurate

pilot workload analysis tool without pilot-in-the-loop testing for the shipboard operations.

Moreover, little work has been focused on how to optimize the pilot model for the helicopter

shipboard tasks. For example, the inverse simulation method is widely used to estimate

the pilot control activity. Although a number of schemes have been examined to improve

the inverse simulation method by combining with crossover model or on-line compensation,

these pilot control inputs achieved in inverse simulation methods are far from the required

control authority predicted for an accurate pilot workload analysis. Thus, the human pi-

lot model remains a critical limit in practical implementation of helicopter/ship dynamic

interface simulation tool for a pilot workload analysis.

Therefore, the primary objective of the present research is to develop an advanced

helicopter/ship dynamic interface simulation tool that can help both to understand the

potential of helicopter simulation, and to investigate the pilot workload issues in the dynamic

interface. In addition, an automatic control system is developed to reduce the high pilot

workloads that typically result in the flight safety limitations associated with shipboard

operations.

In this new study, through utilizing the optimal control model as a human pilot model,

a tunable human pilot model is designed to observe the relative behavior for different levels

of tracking precision and is designed to operate over a range of airspeeds using a simple gain

scheduling algorithm. This pilot model is then used to calculate the required control inputs
CHAPTER 1. INTRODUCTION 19

for the specified trajectories for shipboard operations in two different WOD conditions.

Then the simulation model is validated with JSHIP flight test data.

Once the feasibility and potential of this helicopter/ship dynamic interface simulation

model is established, a stability augmentation system is optimized for a UH-60 helicopter

operating in a turbulent ship airwake. To do this, a new airwake model is derived from

the simulation with the full CFD airwake by extracting an equivalent six-dimensional gust

vector. The spectral properties of the gust components are then analyzed, and shaping

filters are designed to simulate the gusts when driven by white noise.

For disturbance rejection, a new performance specification is designed based on the

power spectrum density of the transfer function between the gust inputs and aircraft rate

responses. The baseline limited authority SAS is modified and optimized using CONDUIT

in order to improve handling-qualities and stability, and to minimize a weighted objective

of gust responses. In addition, a H∞ controller is designed to provide an alternative SAS

configuration. The optimized SAS and H∞ SAS are then tested using the non-linear simu-

lation model with time-varying airwake. A series of comparisons are also made to provide

a sound validation of the new SAS models.


Chapter 2

Helicopter Flight Dynamics Model

A helicopter/ship dynamic interface simulation model of a UH-60A operating off an LHA

class ship is developed. The helicopter flight dynamics model is based on the GENHEL

simulation model. The simulation model is then converted to MATLAB/SIMULINK based

simulation model to facilitate model improvement and controller design. A higher order

Peters-He inflow model is employed for the main rotor inflow. The gust penetration model

is used to model the effects of a ship airwake on the helicopter flight dynamics.

2.1 Overview of the GENHEL Simulation Model

The GENHEL was originally developed by Sikorsky Aircraft and documented under con-

tract from NASA. This model is a total systems definition of the Black Hawk helicopter

and provides handling qualities analysis tool for the Black Hawk helicopter which can even-

tually be extended to a real time pilot-in-the-loop simulation. The mathematical model is

generalized analytical representation of a total helicopter system. It normally operates in

20
CHAPTER 2. HELICOPTER FLIGHT DYNAMICS MODEL 21

the time domain and allows the simulation of any steady or maneuvering flight condition

which can be experienced by a pilot [11]. The overall block diagram of the GENHEL is

presented on Figure 2.1.

The basic model is a total force, nonlinear, large angle representation in six rigid body

degrees of freedom. In addition, main rotor rigid blade flapping, lagging and hub rotational

DOF are represented. The hub rotational degree of freedom is coupled to the engine and

fuel control. Motion in the lag degree of freedom is resisted by a nonlinear lag damper

model.

The main rotor model is developed using a blade element approach with five equal-

annuli segments on each of the four blades. In the air mass degree of freedom, a uniform

Main
MainRotor
Rotor
Positions,
Velocities,
Attitudes, … Engine
Engine

Fuselage
Fuselage
Equations
Equationsofof
Servo
ServoModel
Model ++ +
Motion
Horizontal Motion
HorizontalTail
Tail

Control
ControlSystem Vertical
System VerticalTail
Tail

Tail
TailRotor
Rotor
Pilot
Pilot
Sensor
SensorModel
Model
AFCS
AFCSModel
Model
Display
Display System
SystemModel
Model

Figure 2.1: Block diagram of GENHEL flight simulation model (Ref. [11])
CHAPTER 2. HELICOPTER FLIGHT DYNAMICS MODEL 22

downwash is derived from momentum considerations, passed through a first order lag, and

then distributed first harmonically as a function of rotor wake skew angle and the aerody-

namic hub moment. The lift and drag forces of each blade are loaded from wind tunnel

data which is a function of Mach number and angle of attack.

The fuselage is represented by six components of aerodynamic characteristic which are

defined from wind tunnel data. The angle of attack at the fuselage is calculated using the

free stream and interference effects of the main rotor. These interference are based on rotor

loading and rotor wake skew angle.

The aerodynamics of the empennage are treated separately from the forward airframe.

This separate formulation allows good definition of non-linear tail characteristics. The

angles of attack at the empennage are developed from the free stream velocity, plus rotor

wash. Additional dynamic pressure effects from fuselage is accounted for by factoring the

free stream velocity component.

The tail rotor is represented by the linearized closed form Bailey theory solutions. Terms

in tip speed ratio greater than square of advance ratio have been eliminated. An empirical

blockage factor, due to the proximity of the vertical tail, is applied to the thrust output.

The flight control system consists of the primary mechanical flight control system and

the Automatic Flight Control System (AFCS) which includes the Stability Augmentation

System (SAS), the Pitch Bias Actuator (PBA), the Flight Path Stabilization (FPS) and

the Stabilator mechanization. The flight control model represents the control system in a

complete manner except for the FPS. In this case, only the attitude hold and turn features

have been defined. The detailed definitions of each component are given in Ref. [11].
CHAPTER 2. HELICOPTER FLIGHT DYNAMICS MODEL 23

2.2 MATLAB Implementation

MATLAB based simulation model is implemented to facilitate model improvement and

control law development (Figure 2.2). The new simulation model includes main rotor, fuse-

lage, empennage, tail rotor and other subsystems (primary flight control system, stability

augmentation system, sensors, etc.). The simulation model provides the simulation of any

steady or maneuvering flight condition. In addition, numerically linearized dynamic model

can be obtained for controller design.

So far, the simulation model presented in Figure 2.2 has been updated to include Peters-

He inflow model, gust penetration, and pilot model. Details are presented in Section 2.3,

Section 2.4, Chapter 3.

Figure 2.2: Overall structure of MATLAB based simulation program


CHAPTER 2. HELICOPTER FLIGHT DYNAMICS MODEL 24

2.3 Peters-He Inflow Model

2.3.1 Background

Dynamic inflow modeling in helicopter flight dynamics is a means of accounting for the low

frequency wake effects under unsteady or transient conditions. It has been known for years

that the induced flow field associated with a lifting rotor responds in a dynamic fashion to

changes in either blade pitch or rotor flapping angles. In recent years, it has been found

that dynamic inflow for steady response in hover can be treated by an equivalent Lock

number [15]. For more general conditions, such as transient conditions or a rotor in forward

flight conditions, it has been determined that the induced flow can be treated by additional

degrees of freedom of the system.

The most popular model of dynamic inflow is that of Pitt and Peters [14]. The theory

of Pitt-Peters dynamic inflow model relates the airloads of a rotor (CT , CL , and CM ) to

the induced flow distributions (λ0 , λs , and λc ) where CT , CL , and CM are the aerodynamic

perturbation in thrust, roll moment, and pitch moment; and λ0 , λs , and λc are the magni-

tude of uniform, lateral and longitudinal variations in induced flow. The induced inflow is

assumed to have the following variations in the wind-axis coordinates.

r r
λ(r, ψ) = λ0 + λs sin ψ + λc cos ψ (2.1)
R R

where r is blade radial coordinate and R is rotor radius. The time histories of λ0 , λs , and
CHAPTER 2. HELICOPTER FLIGHT DYNAMICS MODEL 25

λc are governed by the following first-order differential equation.

     
     



 λ̇0 






 λ0 






 CT 




 
 
 
 
 

     
−1
[M ] λ̇s  + [L] λs  =  CL (2.2)

  
   


 
 
 
 
 


 
 
 
 
 

 λ̇c 
   λc  
   CM 

aero

where [M ] is the matrix of the apparent mass terms (a time delay effect due to the unsteady

wake), [L] is the nonlinear version of the inflow gains matrix. It should be noted that the

subscript “aero” implies that only aerodynamic contributions are considered in CT , CL , and

CM (i.e. inertial terms are omitted).

Peters-He inflow model, also called the generalized dynamic wake model, is characterized

as representation of induced inflow as dynamic degrees of freedom in a system of first order

differential equations in the time domain. The “Pitt-Peters” inflow model can be thought

of as a special case of this theory but with only 3 inflow expansion terms (uniform, lateral,

and longitudinal). Based on an unsteady wake model, Peters-He inflow model represents

essentially the unsteady wake-induced flow through the rotor disk excited by aerodynamic

loads in a global fashion.

2.3.2 Basic Equations of Peters-He Inflow Model

In the generalized dynamic wake model [18], the induced flow distribution can be represented

in terms of a harmonic variations in azimuth and arbitrary radial distribution functions.

∞ ∞ h i
φrj (r̄) αjr (t̄) cos(rψ) + βjr (t̄) sin(rψ)
X X
w (r̄, ψ, t̄) = (2.3)
r=0 j=r+1,r+3,···
CHAPTER 2. HELICOPTER FLIGHT DYNAMICS MODEL 26

where r̄ and t̄ are nondimensional blade radial coordinate (r̄ = r/R) and nondimensional

time (t̄ = Ωt), respectively. The radial expansion function φrj (r̄) has the following form,

j−1
q (−1)(q−r)/2 (j + q)!!
φrj (r̄) r̄q
X
= (2j + 1)Hjr (2.4)
q=r,r+2,···
(q − r)!!(q + r)!!(j − q − 1)!!

where

(j + r − 1)!!(j − r − 1)!!
Hjr = , (n)!! = (n)(n − 2) · · · (2) or (1) (2.5)
(j + r)!!(j − r)!!

Equation (2.3) gives a harmonic content of induced inflow that can be truncated at any

harmonic of interest; at the same time, it provides a complete description of the radial

variation of induced inflow at rotor disk. With pressure and induced velocity represented

by the above expansion, pressure coefficients τ ’s and the inflow coefficients αjr and βjr can

be related in the following matrix form

n o∗ n o 1 mc
[Mnm ] αjr + V [Lc ]−1 αjr = {τ } (2.6)
2 n
n o∗ n o 1 ms
[Mnm ] βjr + V [Ls ]−1 βjr = {τ } (2.7)
2 n

where [Mnm ] is called the apparent mass matrix and is given by

2 m
Mnm = H (2.8)
π n
CHAPTER 2. HELICOPTER FLIGHT DYNAMICS MODEL 27

and where [Lc ] and [Ls ] are the induced inflow influence coefficient matrices and depend on

the wake skew angle χ (χ = 0◦ in axial flow through χ = 90◦ in pure-edgewise flow).

h ic h i
L0m
jn = X m Γ0m
jn (2.9)
h ic  h i
Lrm
jn = X |m−r| + (−1)l X |m+r| Γrm
jn (2.10)
h is  h i
Lrm
jn = X |m−r| − (−1)l X |m+r| Γrm
jn (2.11)

where l = min(r, m), and X = tan |χ/2|. Note that 0 ≤ X ≤ 1. All sine and cosine elements

on the same coefficients Γrm


jn that can be found in closed-form as follows.

(−1)(n+j−2r)/2
p
2 (2n + 1)(2j + 1)
Γrm
jn = for r + m even (2.12)
(j + n)(j + n + 2) [(j − n)2 − 1]
q
Hnm Hjr
π sgn(r − m)
Γrm
jn = q p for r + m odd, j = n ± 1 (2.13)
2 Hnm Hjr (2n + 1)(2j + 1)

Γrm
jn = 0 for r + m odd, j 6= n ± 1 (2.14)

[L] matrix is partitioned such that the superscripts are row-column indices of the r,

m partition, and the subscripts j, n are the row-column indices of the elements within

each partition. It should be noted that these indices do not take the traditional matrix

values of 1, 2, 3, · · ·. Instead, for the cosine equation, m = 0, 1, 2, 3, · · · ; for sine equation,

m = 1, 2, 3, · · · ; and for either set, n = m + 1, m + 3, m + 5, · · · (r and j follow the same

convention).

In Equations (2,6) and (2.7), V is the mass flow parameter to account for energy added
CHAPTER 2. HELICOPTER FLIGHT DYNAMICS MODEL 28

to the flow from the rotor

µ2 + (λ + λm )λ
V = p (2.15)
µ2 + λ2
λ = λm + λf (2.16)

where V comes from momentum considerations, µ, λf are the nondimensional inplane and

normal components of V∞ , and λm is the momentum theory value of steady induced flow

for a trimmed rotor. According to the approach followed for the low frequency dynamic

inflow model in Reference [17], a completely nonlinear version of Equations (2.6) and (2.7)

can be obtained by

µ2 + λ2m corresponding to the first column (r = 0) of [Lc ]−1 , but


p
1. taking V as VT =

as V for r 6= 0.

2. treating all quantities as total induced flow rather than perturbation.


3. replacing the static λm by the unsteady value, 3α10 .

In order for the model to be coupled with blade lift theory, the τnmc and τnms need to be

appropriately related to the blade lift. The pressure harmonic coefficients τnmc and τnms can

then be given by

Q Z 1
1 X Lq

τn0c = 0
φ (r̄)dr̄ (2.17)
2π q=1 0 ρΩ2 R3 n
Q Z 1
1X Lq

τnmc = φm
(r̄)dr̄ cos (mψq ) (2.18)
π q=1 0 ρΩ2 R3 n
Q Z 1
1X Lq

τnms = m
φ (r̄)dr̄ sin (mψq ) (2.19)
π q=1 0 ρΩ2 R3 n
CHAPTER 2. HELICOPTER FLIGHT DYNAMICS MODEL 29

where Lq is blade sectional lift that can be evaluated from a lift theory and Q is number of

blades.

Thus, Equations (2.3) to (2.19) comprise a complete, three-dimensional unsteady wake

model written in terms of a finite number of states αjr and βjr . The fundamental idea of

this theory and some initial validations have been included in References [16] ∼ [18]. In

this study, a 15 state Peters-He is used. Figure 2.3 shows the difference between Pitt-Peters

inflow model and Peters-He inflow model. It can be observed that Pitt-Peters inflow model

only estimates the average inflow distribution over the rotor disk.
Induced inflow ratio

Induced inflow ratio

Advancing Advancing
side Trailing side Trailing
Y edge Y edge
X X
Retreating Leading Retreating Leading
side edge side edge

(a) Pitt-Peters 3 state inflow model (b) Peters-He 15 state inflow model

Figure 2.3: Comparisons of inflow ratio

2.4 Gust Penetration Model

The gust penetration model is used to model the effects of a three-dimensional ship airwake

on the helicopter flight dynamics. There is a fundamental assumption that the velocity field

of the airwake affects the aerodynamic forces on the helicopter, but the helicopter does not

affect the ship airwake.


CHAPTER 2. HELICOPTER FLIGHT DYNAMICS MODEL 30

The ship airwake velocity field can be found from output of CFD programs. For instance,

the CFD solver called PUMA2 (Parallel Unstructured Maritime Aerodynamics)provides

time-varying ship airwake solutions around an LHA class ship [39]. It uses a finite volume

formulation of the Euler/Navier-Stokes equations for 3-dimensional, internal and external,

non-reacting, compressible, steady/unsteady solutions for complex geometries (Figure 2.4).

PUMA2 can be run so as to preserve time accuracy or as a pseudo-unsteady formulation

to enhance convergence to steady-state. It is written in ANSI C using the MPI library for

message passing so it can be run on parallel computers and clusters. It is also compatible

with C++ compilers and coupled with the computational steering system POSSE. It uses

dynamic memory allocation, thus the problem size is limited only by the amount of memory

available on the machine. Large eddy simulations can also be performed with PUMA2 [39].

The flow case represents both 0 and 30 degrees yaw angles and 30 knot of relative wind

speed. A 4-stage Runge-Kutta explicit time integration algorithm with Roe’s flux difference

scheme is used with CFL numbers of 2.5 and 0.8 for the steady and unsteady computations,

respectively. A zero-normal-velocity boundary condition is applied on the ship surface and

water surface (bottom surface of the domain) and a Riemann boundary condition is applied

at all other faces of the domain. The pseudo steady-state computations are performed using

local time-stepping and initialized with freestream values. The time-accurate computations

are started from the pseudo steady-state solution, and the simulation time step (480 sec) is

determined by the smallest cell size in the volume grid. The computations are performed

on a parallel PC cluster Lion-XL consisting of 256 2.4 Ghz P4 processors with 4 GB ECC

RAM and Quadrics high-speed interconnect [39]. Iso-surfaces of vorticity magnitude of 0.8

sec−1 for both 0 and 30 degree WOD cases are shown in Figure 2.4.
CHAPTER 2. HELICOPTER FLIGHT DYNAMICS MODEL 31

The 40 seconds of the time history data is stored for every 0.1 seconds to be used for the

DI simulations. Each flow solution file is 41 Mbytes in size, whereas a DI velocity data file

is only 5.2 Mbytes. In this study, the discussion will be limited to how to use the outputs

of the CFD program.

a) 0 degree WOD case b) 30 degree WOD case

Figure 2.4: Vorticity magnitude iso-surface at t = 40 sec (Ref. [39])

The ship airwake velocity field from Reference [39] has (81 × 30 × 23) rectangular grid

points with 5ft equal intervals for the part of the ship where the helicopter is expected to fly.

In this study, a 3-dimensional look-up algorithm (including linear interpolation algorithm)

is implemented to calculate the local velocity disturbances at each aerodynamic center of

rotor blade elements, fuselage, empennage and tail rotor. The overall gust penetration

model is presented on Figure 2.5.

The ship airwake velocity field provided by the CFD database is defined with respect to

a ship-fixed coordinate system. Thus, the velocity field must be transformed to inertial axes,

and then to the specific axis system used for each of the helicopter component models. For

the fuselage, empennage, and tail rotor, the following coordinate transformation is required.
CHAPTER 2. HELICOPTER FLIGHT DYNAMICS MODEL 32

Time-Accurate Ship Airwake Account for Local Velocities


at Blade Elements, Fuselage,
Velocities from CFD
Empennage, Tail Rotor

Linear look-up
algorithm

Coordinate transformation
r wake target r wake
Vtarget = [T]ship Vship
3-D uniform grid

Figure 2.5: Gust penetration model

~body = [T ]ib [T ]si V


V ~ship (2.20)

~ship is the ship wake velocity vector in ship coordinate system, V


where V ~body is the ship

wake velocity vector in body coordinate system, and [T ]si , [T ]ib are the coordinate transfor-

mation matrices from ship coordinate system to inertial coordinate system and from inertial

coordinate system to helicopter body coordinate system, respectively.

For each main rotor blade element, the airwake velocity in blade coordinate system is

obtained by

~rotor = [T ]hb [T ]bh [T ]ib [T ]si V


V ~ship (2.21)

~rotor is the ship airwake velocity vector in each blade coordinate system, [T ]b , [T ]h
where V h b

are the coordinate transformation matrices from helicopter body to hub coordinate system
CHAPTER 2. HELICOPTER FLIGHT DYNAMICS MODEL 33

and from hub to each blade coordinate system, respectively.

Since only finite time-varying solutions of ship airwake can be processed and stored in

computers, a simulation of shipboard operation may exceed the range of the data duration.

To overcome this limitation, the solutions are overlapped with sinusoidal filter for first and

last 5 seconds once the simulation time exceeds the airwake data duration (Figure 2.6). The

approach of the overlapped time history variation prevents any sudden jump in the airwake

velocities at the time when the simulation lasts longer than data duration. If time step of

airwake solution is different from that of simulation program, a simple linear interpolation

can be applied to calculate the airwake velocity field at every simulation time step.

Figure 2.6: The approach of the overlapped time history of airwake


Chapter 3

Pilot Modeling

Since the introduction of modern manual control research of dynamic systems during the

1940’s, the control theory which has evolved in the intervening years has been useful in

quantifying control-related human behavior [62]. The so-called “crossover model” employs

classical control methods to model human feedback control of single input, single output

(SISO) systems [52]. The method is based on the expected crossover frequency of the open

loop transfer function of the human and controlled process. In fact, Bradley and Turner

applied the crossover model, coupled with inversion control methods, specifically to model

pilot workload for helicopter shipboard operations [53]. The main drawback of the crossover

model is that the helicopter piloting task is inherently multi input, multi output (MIMO)

system.

To analyze more complex manual control systems, most efforts have been concentrated

on the problem of developing linear models for the human controller model in MIMO sit-

uations. As regards this problem, two basic approaches have been emerging. The first ap-

proach is to extend the classical crossover model developed for SISO system to the MIMO

34
CHAPTER 3. PILOT MODELING 35

system [63]. Their approach is based on classical multi-loop control theory and depends

highly on judgments concerning the closed-loop system structure.

The second approach is based on modern MIMO control theory. It is capable of treating

multivariable systems within a single conceptual framework using state-space forms which

are more naturally suited to the analysis of complex man-machine systems, particularly since

the design algorithms are readily automated using modern software such as MATLAB.

3.1 Optimal Control Model of the Human Pilot

The optimal control model (OCM) of the human pilot is applied for MIMO systems by

solving the Linear Quadratic Gaussian (LQG) problem. The pilot control inputs are based

on a compensatory tracking model of the human pilot. An optimal control model is used to

allow realistic computer simulations of the shipboard operations. References [62] and [63]

provide information on the model and introduce its application in linear or nonlinear flight

dynamic models. Figure 3.1 shows a schematic of the optimal control model used in this

study.

Optimal Control Model


Optimal
Optimal
Desired
Desired yd Deadband Kalman
Kalman
Deadband Time
Trajectory
+-
(collective only) TimeDelay
Delay Estimator
Feedback
Feedback
Trajectory y
(collective only) Estimator Gains
Gains

UH-60
UH-60Flight
Flight Neuromotor Dynamics
+ u Hysteresis
Hysteresis
Dynamic
Dynamic 1
+ (collective only)
(collective only)
Model
Model t ns +1

Disturbance

Figure 3.1: Optimal control model of the human pilot


CHAPTER 3. PILOT MODELING 36

In this study, the human pilot’s basic task is to control the helicopter to follow a specified

shipboard trajectory. To design the human control model, the helicopter is represented with

linearized equations of motion in state-space form:

ẋ(t) = Ax(t) + Bu(t) + w(t) (3.1)

y(t) = Cx(t) + v(t) (3.2)

where x(t) is the state vector, u(t) is the pilot’s control input vector, w(t) is a vector of

external disturbances, y(t) is the output vector (parameters perceived by the pilot), and

v(t) represents observation noise.

The linearized system model used in this study is obtained through numerical lineariza-

tion of the simulation model. The resultant linear model is a 24 state model, which includes

9 rigid body states and 15 states associated with rotor dynamics and inflow. For this study,

the model is linearized at every 10 knots (e.g. hover, 10 knot, 20 knot, ... , 140 knot).

Assuming quasi-static rotor and inflow dynamics, the linear model is reduced to a 9 state

/ 6 DOF model of the rigid body motion. The linear model is decoupled into a 3 state

longitudinal model, a 5 state lateral-directional model, and a 1 state vertical motion model.

Finally, the linear models must be augmented to include shaping filters for the gust dis-

turbances and a dynamic model of the SAS for each axis. Integrators are added so that

position and integrated position can be included in the performance index. A schematic of

the augmented flight dynamics model used for longitudinal control is shown in Figure 3.2.

A similar model is used for lateral-directional control, which includes pilot inputs in both

the lateral and directional axes.


CHAPTER 3. PILOT MODELING 37

òx
Gust disturbance ò
Linearized model
(24 state) K turb x
Longitudinal dynamics ò
s + w turb
éu& ù é X u X q - g cosq 0 ù éu ù
Reduced model ê q& ú = ê M ú êq ú
u
ê ú ê u Mq 0 úê ú
(9 state) + q
+ êëq&úû êë 0 1 0 úû êëq úû
Pilot - é X d long ù
Decoupled q
+ êê M d long úúd long
input
model êë 0 úû

Include
Pitch SAS
shaping filter for the gust, 6.2 s ( s + 1)
SAS dynamics ( s + 0.5)( s + 0.143)

Figure 3.2: Augmented plant model in longitudinal axis

The OCM of the human pilot in Figure 3.1 is represented as an optimal linear regulator

in combination with an optimal state estimator (Kalman estimator). Both the estimator

and feedback gain matrix are determined by solving the LQG control problem. Assuming

the linear dynamics of Equations (3.1), (3,2) and the disturbance are white noise signals,

the objective is to find a dynamic compensator that minimizes the quadratic performance

index given by

( )
1
Z T h i
J = E lim xT (t)Qx(t) + u̇(t)Ru̇(t) dt (3.3)
T →∞ T 0

where Q and R are the state and control weighting matrices. The estimator and feedback

gains are readily solved from a pair of matrix Riccati equations [64]. Note that when

modeling human operators it is customary to use control rates instead of the control position

in the performance index. A simple augmentation of the plant dynamics model is used to

achieve this [62]. Details of the complete optimal control model of the human operator are

discussed in References [62] and [63].


CHAPTER 3. PILOT MODELING 38

The optimal control model is essentially specified by (1) the weighting matrices Q and

R in Equation (3.3) , (2) the covariances of the observation and external noises, and (3) the

magnitude of the operation time delay. These parameters can be selected to yield optimal

control model transfer functions. The effect of those parameter variations on the resulting

model is not as clear as in the case of the crossover model of the human operator. This

is due to the fact that the optimal control model parameters are essentially inputs to an

optimization scheme that involves the solution of sets of nonlinear algebraic equations [63].

However, Reference [63] outlines an approximate method for selecting these parameters to

achieve a desired crossover frequency for each control axis.

Typically, both the Q and R matrices are assumed to be diagonal. The weighting

parameters in the Q matrix are selected such that each state variable is scaled by its

maximum expected deviation [63]. This leaves the task of selecting the control weighting

parameters in R. In this study, the objective is to develop a pilot model that can be

easily tuned to adjust the tracking tolerance in each control axis, where a high crossover

frequency corresponds to “tight” tracking and a low crossover frequency corresponds to

“relaxed” tracking.

Consider the longitudinal axis where the transfer function from longitudinal control

input to pitch attitude can be expressed as

K sm + am−1 sm−1 + am−2 sm−2 + · · · + a1 s + a0



θ
(s) = (3.4)
δlong (sn + bn−1 sn−1 + bn−2 sn−2 + · · · + b1 s + b0 )

An approximate but very useful relationship exists between the weighting coefficients, the
CHAPTER 3. PILOT MODELING 39

controlled-element dynamics, and the closed-loop system bandwidth [62], [63].

h i1/(n−m+1)
ωBW ≈ K(qθ /rδlong ) (3.5)

where qθ is the weighting parameter for pitch attitude, rδlong is the longitudinal control

weighting, and ωBW is the closed-loop bandwidth. The precise definition of ωBW is the

frequency where the amplitude of the closed-loop transfer function is 6dB below its zero-

frequency value. Equation (3.6) shows an approximate relation between the open-loop

corssover frequency and the closed-loop bandwidth.

ωc ≈ 0.56ωBW (3.6)

Thus, given a desired crossover frequency in each control axis, Equations (3.4), (3.5) and

(3.6) provide a method of determining appropriate control weighting parameters in R.

From the previous study, it was shown that the actual crossover frequencies were slightly

different from the desired crossover frequencies in the OCM design, due to the approximate

nature of Equations (3.4) - (3.6) [63]. In this study, an iterative method is used to obtain the

exact desired crossover frequencies. The design process begins with initial guess of control

weighting parameters based on Equations (3.4) - (3.6). Then the actual crossover frequency

is calculated for the full order model. The weighting parameters are adjusted proportional

to this discrepancy between the actual and expected values of the crossover frequency, and

the process is then repeated. This iteration process is automated in MATLAB, and in all

cases the iteration is repeated until the differences between the actual and desired crossover
CHAPTER 3. PILOT MODELING 40

frequencies are negligible.

The blocks labeled “neuromotor dynamics” and “time delay” (Figure 3.1) represent

psycho-physical limitations inherent in the human pilot. It should be noted, for example,

that rapid control movements are rarely produced by trained pilots. Alternatively, these

terms can be used to account indirectly for the physiological limitation on the ability of

human pilots to make corrective actions. The neuromotor dynamics is often approximated

linearly by an adjustable first-order lag filter [62]. The time delay represents an actual

delay. In this study, τn = 0.1 is used and time delay is set to nominal value of 0.1 sec.

The procedure above provides a method for designing OCM gains for a single operating

point. In order to operate over a range of airspeeds, the procedure is repeated in 10 knot

increments from hover out to the maximum speed of the aircraft. A simple gain scheduling

approach is used to adjust the OCM gains as the airspeed changes in flight. For example,

the system matrices of controller for 30 knot and 40 knot are represented as

for 30 knot : Ac30knot , Bc30knot , Cc30knot , Dc30knot


(3.7)
for 40 knot : Ac40knot , Bc40knot , Cc40knot , Dc40knot

And if the helicopter airspeed is 35 knot, then the final control system matrices are obtained

using linear interpolation of the matrices.

Ac35knot = 0.5Ac30knot + 0.5Ac40knot

Bc35knot = 0.5Bc30knot + 0.5Bc40knot


(3.8)
Cc35knot = 0.5Cc30knot + 0.5Cc40knot

Dc35knot = 0.5Dc30knot + 0.5Dc40knot


CHAPTER 3. PILOT MODELING 41

3.2 Nonlinear Elements of the Human Pilot

Control records from human helicopter pilots have shown that there is a stepped appearance

in the collective input [53]. Pilots tend to make discrete rather than continuous adjustments

to the collective lever. This effect can be modeled using nonlinear elements in the pilot

model. A hysteresis is attached to the control leading to the helicopter and deadband

is placed across the error prior to its processing by control model. Descriptions of these

elements are presented in the following Sections 3.2.1 and 3.2.2

3.2.1 Hysteresis

The hysteresis represents a system in which a change in input causes an equal change in

output. However, when the input changes direction, an initial change in input has no effect

on the output. The amount of side-to-side play in the system is referred to as the deadzone.

The deadzone is centered about the output. A system can be in one of three modes:

1. Disengage - in this mode, the input does not drive the output and the output remains

constant.

2. Engaged in a positive direction - in this mode, the input is increasing (has a positive

slope) and the output is equal to the input minus half the deadzone width.

3. Engaged in a negative direction - in this mode, the input is decreasing (has a negative

slope) and the output is equal to the input plus half the deadzone width.

For example, Figure 3.3 shows the effect of a sine wave passing through a hysteresis

element with deadzone width of 1.


CHAPTER 3. PILOT MODELING 42

0.8

0.6
(c)
0.4

0.2
(b)

0
(a)
−0.2

(d)
−0.4

−0.6

−0.8

−1
0 1 2 3 4 5 6 7 8 9 10
Time (sec)

Figure 3.3: Effect of a sine wave passing through a hysteresis

(a) Input engages in positive direction. Change in input causes equal change in output.

(b) Input disengages. Change in input does not affect output.

(c) Input engages in negative direction. Change in input causes equal change in output.

(d) Input disengages. Change in input does not affect output.

3.2.2 Deadband

The deadband represents a threshold of the perception of departure from the reference

values. The deadband generates zero output within a specified region. The lower and

upper limits of the deadband are specified as the start of deadband and end of deadband

parameters. The output depends on the input and deadband:

1. If the input is within the deadband (greater than the lower limit and less than the
CHAPTER 3. PILOT MODELING 43

upper limit), the output is zero.

2. If the input is greater than or equal to the upper limit, the output is the input minus

the upper limit.

3. If the input is less than or equal to the lower limit, the output is the input minus the

lower limit.

Figure 3.4 shows the effect of the deadband element (deadband width = 1) on the sine wave.

While the input (the sine wave) is between −0.5 and 0.5, the output is zero.

0.8

0.6

0.4

0.2

−0.2

−0.4

−0.6

−0.8

−1
0 1 2 3 4 5 6 7 8 9 10
Time (sec)

Figure 3.4: Effect of a sine wave passing through a deadband


Chapter 4

Numerical Examples : Departure

and Approach Operations

4.1 Overview

The dynamic interface flight simulation model has been applied to simulate a UH-60A

operating near an LHA class ship. The simulation has been performed for two different

shipboard operations (departure, approach). Kinematic profiles of these shipboard tasks

are given in Reference [49], and these profiles are modified slightly in this study. As discussed

in Reference [49], the kinematic profile is determined using an Earth fixed coordinate frame

with the origin at the sea surface directly under the initial position of the helicopter. The

X-axis is along with the North direction, Z-axis is downward, and Y-axis is along with the

East direction. The target spot of shipboard operations is spot 8 of the LHA class ship

(Figure 4.1).

The optimal control model is used to determine the required pilot control inputs for

44
CHAPTER 4. NUMERICAL EXAMPLES 45

3A
3

9
1
Spot 8

7
2

8
Figure 4.1: Top view of an LHA class ship

given shipboard approach and departure tasks. The OCM of the human pilot used in the

sections 4.2 and 4.4 is designed for the desired crossover frequency of the open-loop transfer

function in the lateral (2.75 rad/sec), longitudinal (1.8 rad/sec), collective (2.0 rad/sec),

and pedal (2.0 rad/sec). These crossover frequencies are selected to have similar control

activities from JSHIP flight test. For both departure and approach operations, the ship is

assumed to be still with a steady-state wind of 30 knots for two different WOD conditions

(0◦ and 30◦ ). In this study, 30 degree WOD condition is considered as worst case because

the flight test results from Ref. [2] showed substantial increase in pilot workload due to

turbulence and this case was awarded as ‘unacceptable’.

4.1.1 Shipboard Departure Trajectory

Typical shipboard departure procedures include all actions that are required to conduct

an ascending, acceleration departure from a stationkeeping, ending in steady, level forward

flight [49]. Starting from the stationkeeping location, pilots typically initiate the departure

phase by yawing and/or translating the helicopter at a relatively constant altitude to a

position outboard of the recovery spot that is clear of obstructions. The entire shipboard

departure task can be divided into the following three phases:

1. Phase I : from the stationkeeping position, accelerating to a desired climb rate and a
CHAPTER 4. NUMERICAL EXAMPLES 46

desired horizontal acceleration.

2. Phase II : keeping a constant climb rate and a constant horizontal acceleration.

3. Phase III : reducing the climb rate and horizontal acceleration to zero, and ending in

a steady level flight

In this study, in order to clear obstructions, the departure begins with 60 ft translational

maneuver to the left (port side). The helicopter then transitions to a desired climb rate

and horizontal acceleration. Finally, the helicopter achieves desired level flight speed. The

key parameters for defining the departure trajectory profile are the helicopter initial level

flight speed, initial altitude, and desired final altitude for stationkeeping. The initial profile

parameters of the departure operation are given in Table 4.1.

Table 4.1: Initial profile parameters for the departure task

Trajectory parameters Departure task


Initial altitude 80 ft (17 ft above deck)
Final altitude 300 ft
Initial speed 0 knot
Final speed 60 knot

4.1.2 Shipboard Approach Trajectory

Similar to the departure operation, typical shipboard approach procedures include all ac-

tions that bring the rotorcraft from a point far away from the ship down to a point much

closer to the recovery spot [49]. The key parameters for defining the approach profile are

the helicopter initial level flight speed, initial altitude, initial distance from the ship, and

desired final altitude for stationkeeping.


CHAPTER 4. NUMERICAL EXAMPLES 47

The entire shipboard approach task can be divided into three phases:

1. Phase I : From steady level flight, the helicopter transitions to a desired descent rate

and horizontal deceleration.

2. Phase II : The helicopter maintains a constant descent rate and horizontal decelera-

tion.

3. Phase III : The descent rate and horizontal deceleration are reduced to zero, ending

in stationkeeping over a landing spot.

In this study, the helicopter approaches the ship from the port side at a 45 degree angle,

and then performs a 45 degree left turn to align itself with the longitudinal axis of the ship

after it crosses over the deck. This is similar to the trajectory used in the JSHIP study. For

simplicity, the ship is assumed to be stationary in a 30 knots steady wind. Both a head wind

and a wind from 30 degree starboard of the bow are considered. The entire procedures of

shipboard approach are shown in Figure 4.2. The initial profile parameters of the departure

operation are given in Table 4.2.

Phase III Phase II Phase I

0 deg WOD Vi
60 knot
30 deg WOD

H300
i ft
8 9
45
H
80f ft

Top view
Shipboard approach - 45 deg approach

Figure 4.2: Shipboard approach operation procedures


CHAPTER 4. NUMERICAL EXAMPLES 48

Table 4.2: Initial profile parameters for the approach task

Trajectory parameters Approach task


Initial altitude 300 ft
Final altitude 80 ft (17 ft above deck)
Initial air speed 60 knot
Final air speed 0 knot

4.2 Effects of Ship Airwake Model

In this study, the helicopter/ship dynamic interface simulation has been performed for three

different airwake cases (no airwake, steady-state airwake, time-varying airwake) in 0 and 30

degree WOD conditions. The ship’s time-varying airwake and steady-state (time-averaged)

airwake solutions are calculated using PUMA2 by Sezer-Uzol et al [39]. These solutions

present the airwake velocity field over the 3-dimensional full-scale LHA geometry.

Figures 4.3 - 4.9 show the simulation results for departure operation in 0 and 30 degree

WOD cases. The dotted lines represent the no airwake condition, the dashed lines repre-

sent the steady airwake condition, and the solid lines represent the time-varying airwake

condition. The helicopter trajectory with respect to the ship coordinate system is shown in

Figure 4.3 for both 0 degree and 30 degree WOD conditions. Figures 4.4 and 4.7 show the

helicopter velocity in NED (North-East-Down) coordinate system. Figures 4.5 and 4.8 show

the helicopter attitude responses. The pilot stick inputs provided by the optimal control

model are shown in Figures 4.6 and 4.9. The conventions for these control positions are

as follows; full left lateral cyclic, full forward longitudinal cyclic, full down collective pitch,

and full left pedal correspond to 0 %, full right lateral cyclic, full aft longitudinal cyclic,

full up collective pitch, and full right pedal correspond to 100 %. The results show that the
CHAPTER 4. NUMERICAL EXAMPLES 49

helicopter trajectory and velocities are very similar in each case. This is because the opti-

mal control model of the human pilot is regulating these parameters. These variables are

essentially constrained in the simulation; the optimal control model is effectively calculating

the control inputs and aircraft attitude required to track the desired trajectory. However,

when the helicopter is operating within the DI mesh, there is significant difference in the

aircraft attitude response and the pilot control activity. The steady airwake results differ

only slightly from the results with no airwake in that the trimmed controls and attitude

are different. However, the time-varying airwake results in significant oscillations and pilot

activity, particularly when hovering over the ship deck. This difference in the results with

the steady and time-varying airwake was not entirely expected, since a stationary gust field

can appear to be time-varying to the aircraft when it is moving (and especially to the rotor

blades which are constantly moving). As discussed in the previous section, there is strong

unsteadiness in the crossflow (y-component) and vertical (z-component) components of the

velocity due to bow separation, deck-edge vortices and complex island wake for 30 degree

WOD condition. These effects can be clearly observed from results of aircraft attitude

responses and pilot control activities (Figures 4.8 - 4.9). Note that the differences in the

beginning are due to the different trim conditions. From the results, the 30 degree WOD

condition results in significantly larger oscillations and higher pilot control activity, partic-

ularly when hovering over the ship deck. This is consistent with the JHSIP flight test data

[2, 60].
CHAPTER 4. NUMERICAL EXAMPLES 50

400

200
Y, [ft]

0 Ship
Helicopter trajectory

−200 DI mesh

−400
−4000 −3500 −3000 −2500 −2000 −1500 −1000 −500 0

400

300
Helicopter trajectory
Z, [ft]

200 DI mesh

100

Ship
0
−4000 −3500 −3000 −2500 −2000 −1500 −1000 −500 0
X, [ft]

Figure 4.3: Helicopter position w.r.t. ship coordinate system - Departure task
(30 knot, 0 and 30 degree WOD conditions)
CHAPTER 4. NUMERICAL EXAMPLES 51

100
N

50
V

Escape from DI mesh


0
0 10 20 30 40 50 60 70 80 90

0
VE

−5

−10
0 10 20 30 40 50 60 70 80 90

−2
D
V

−4

−6
0 10 20 30 40 50 60 70 80 90
Time, [sec]

Figure 4.4: Helicopter velocity [ft/sec] - Departure task (30 knot, 0 degree WOD condition)
CHAPTER 4. NUMERICAL EXAMPLES 52

No airwake
Steady-state airwake
Time-varying airwake

2
0
−2
PHI

−4
−6 Escape from DI mesh

−8
0 5 10 15 20 25 30 35 40 45

2
THETA

−2
0 5 10 15 20 25 30 35 40 45

0
PSI

−0.5

−1
0 5 10 15 20 25 30 35 40 45
Time, [sec]

Figure 4.5: Helicopter attitude angles [deg] in the DI mesh - Departure task (30 knot, 0
degree WOD condition)
CHAPTER 4. NUMERICAL EXAMPLES 53

No airwake
Steady-state airwake
Time-varying airwake

54 Escape from DI mesh


Lateral

52
50
48

0 5 10 15 20 25 30 35 40 45
Longitudinal

58

56

54

0 5 10 15 20 25 30 35 40 45
63
Collective

62
61
60
59
0 5 10 15 20 25 30 35 40 45
42
40
Pedal

38
36
34
0 5 10 15 20 25 30 35 40 45
Time, [sec]

Figure 4.6: Pilot inputs [%] in the DI mesh - Departure task (30 knot, 0 degree WOD
condition)
CHAPTER 4. NUMERICAL EXAMPLES 54

100
N

50
V

Escape from DI mesh


0
0 10 20 30 40 50 60 70 80 90

0
E
V

−5

−10
0 10 20 30 40 50 60 70 80 90

−2
D
V

−4

−6
0 10 20 30 40 50 60 70 80 90
Time, [sec]

Figure 4.7: Helicopter velocity [ft/sec] - Departure task (30 knot, 30 degree WOD condition)
CHAPTER 4. NUMERICAL EXAMPLES 55

No airwake
Steady-state airwake
Time-varying airwake

0
PHI

−5
Escape from DI mesh

−10
0 5 10 15 20 25 30 35 40 45

4
THETA

0 5 10 15 20 25 30 35 40 45

1
0.5
PSI

0
−0.5
−1

0 5 10 15 20 25 30 35 40 45
Time, [sec]

Figure 4.8: Helicopter attitude angles [deg] in the DI mesh - Departure task (30 knot, 30
degree WOD condition)
CHAPTER 4. NUMERICAL EXAMPLES 56

No airwake
Steady-state airwake
Time-varying airwake

60
55
Lateral

50
45 Escape from DI mesh

0 5 10 15 20 25 30 35 40 45
60
Longitudinal

55
50
45
0 5 10 15 20 25 30 35 40 45

65
Collective

60
55
50
0 5 10 15 20 25 30 35 40 45

38
36
Pedal

34
32
30
28
0 5 10 15 20 25 30 35 40 45
Time, [sec]

Figure 4.9: Pilot inputs [%] in the DI mesh - Departure task (30 knot, 30 degree WOD
condition)
CHAPTER 4. NUMERICAL EXAMPLES 57

Figures 4.10 - 4.16 show similar simulation results for the approach operation. As

expected, the attitude changes and control activity are fairly benign in the early part of the

maneuver, when the helicopter is relatively far from the ship. Near the end of the maneuver,

the helicopter begins to interact significantly with the time-varying airwake, as indicated

by the fluctuations in attitude and increased control activity. It can also be observed that

the oscillations immediately after entering the DI mesh are similar for the steady and time-

varying airwake. At this point, the aircraft is still moving with significant velocity so the

steady gust field has a time-varying appearance to the aircraft. However, once the aircraft

approaches hover, the results indicate the time-varying airwake results in larger oscillations

and higher pilot control activity than the steady airwake. This reflects the so-called cliff

edge effect [2], where strong shear layers from the ship’s superstructure are blown across

the landing spot with winds from 30 degrees.

For both shipboard operations, compared to the case with no airwake, the trim condi-

tions of the helicopter with the ship airwake are different (in terms of pilot control inputs

and helicopter attitude). These differences are clearly induced due to the ship airwake.

From the results, the optimal control model is reasonably effective in tracking the desired

flight path for both approach and departure operations. The results clearly indicate that

the time-varying airwake has a significant impact on aircraft response and pilot control

activity when the aircraft is flown for specified approach and departure trajectories. The

differences are most notable when the helicopter is operating in or near a hover relative to

the ship deck (stationkeeping). In the past, gust models for fixed-wing aircraft simulation

have often used a stationary or frozen field model. This is adequate when the aircraft is

moving at a significant forward speed. However, the model clearly breaks down as airspeed
CHAPTER 4. NUMERICAL EXAMPLES 58

approaches zero. The same appears to be true of helicopters operating in turbulent ship

airwake. The time-varying nature of the ship airwake becomes dominant as the helicopter

approaches hover. And, the 30 degree WOD condition showed a substantial increase in pi-

lot workload. Thus, these ship airwake effects can increase the pilot workload and possibly

degrade handling qualities during shipboard launch and recovery operations.

0
Ship DI mesh
−500

−1000
Y, [ft]

−1500
Helicopter trajectory
−2000

−2500

−3000
−1000 −500 0 500 1000 1500 2000 2500 3000

400

300
Z, [ft]

200 DI mesh Helicopter trajectory

100

Ship
0
−1000 −500 0 500 1000 1500 2000 2500 3000
X, [ft]

Figure 4.10: Helicopter position w.r.t. ship coordinate system - Approach task
(30 knot, 0 and 30 degree WOD conditions )
CHAPTER 4. NUMERICAL EXAMPLES 59

No airwake
Steady-state airwake
Time-varying airwake

100
Entering DI mesh
N

50
V

0
0 10 20 30 40 50 60 70 80 90

2
E
V

−2

0 10 20 30 40 50 60 70 80 90

4
D
V

0 10 20 30 40 50 60 70 80 90
Time, [sec]

Figure 4.11: Helicopter velocity [ft/sec] - Approach task (30 knot, 0 degree WOD condition)
CHAPTER 4. NUMERICAL EXAMPLES 60

No airwake
Steady-state airwake
Time-varying airwake

Entering DI mesh
−2
PHI

−4

−6

45 50 55 60 65 70 75 80 85 90

6
THETA

45 50 55 60 65 70 75 80 85 90

−10
PSI

−20

−30

−40
45 50 55 60 65 70 75 80 85 90
Time, [sec]

Figure 4.12: Helicopter attitude angles [deg] in the DI mesh - Approach task (30 knot, 0
degree WOD condition)
CHAPTER 4. NUMERICAL EXAMPLES 61

No airwake
Steady-state airwake
Time-varying airwake

44
Lateral

42

40 Entering DI mesh

45 50 55 60 65 70 75 80 85 90
68
Longitudinal

64

58
45 50 55 60 65 70 75 80 85 90

65
Collective

60

55
45 50 55 60 65 70 75 80 85 90

55
Pedal

50
45
40
45 50 55 60 65 70 75 80 85 90
Time, [sec]

Figure 4.13: Pilot inputs [%] in the DI mesh - Approach task (30 knot, 0 degree WOD
condition)
CHAPTER 4. NUMERICAL EXAMPLES 62

No airwake
Steady-state airwake
Time-varying airwake

100
Entering DI mesh
N

50
V

0
0 10 20 30 40 50 60 70 80 90

2
E
V

−2

0 10 20 30 40 50 60 70 80 90

4
D
V

0 10 20 30 40 50 60 70 80 90
Time, [sec]

Figure 4.14: Helicopter velocity [ft/sec] - Approach task (30 knot, 30 degree WOD condition)
CHAPTER 4. NUMERICAL EXAMPLES 63

No airwake
Steady-state airwake
Time-varying airwake

Entering DI mesh
−2
PHI

−4

−6

−8
45 50 55 60 65 70 75 80 85 90

8
THETA

2
45 50 55 60 65 70 75 80 85 90

0
−10
PSI

−20
−30
−40

45 50 55 60 65 70 75 80 85 90
Time, [sec]

Figure 4.15: Helicopter attitude angles [deg] in the DI mesh - Approach task (30 knot, 30
degree WOD condition)
CHAPTER 4. NUMERICAL EXAMPLES 64

No airwake
Steady-state airwake
Time-varying airwake

48
Lateral

43

38 Entering DI mesh

45 50 55 60 65 70 75 80 85 90
Longitudinal

65
60
55

45 50 55 60 65 70 75 80 85 90

70
Collective

63

55
45 50 55 60 65 70 75 80 85 90

55
Pedal

45

35
45 50 55 60 65 70 75 80 85 90
Time, [sec]

Figure 4.16: Pilot inputs [%] in the DI mesh - Approach task (30 knot, 30 degree WOD
condition)
CHAPTER 4. NUMERICAL EXAMPLES 65

4.3 Effects of Different Tracking Performance

The OCM of the human pilot is designed for three different levels of tracking performance

by varying the desired crossover frequency of the open-loop transfer function in the pitch,

roll, and yaw axes. The three different cases are termed “normal”, “relaxed”, and “tight”

tracking. Table 4.3 summarizes the desired crossover frequencies for each case and each

control axis. In this study, the desired crossover frequencies are arbitrarily chosen. However,

it is not necessary to specify exact values, only to develop a “tunable” human pilot model

and observe the relative behavior for different levels of tracking precision.

Table 4.3: Crossover frequencies for different tracking performance (rad/sec)

Longitudinal Lateral Collective Yaw


Case 1 (relaxed) 0.5 1.0 0.5 1.0
Case 2 (normal) 1.5 1.75 1.5 1.5
Case 3 (tight) 2.0 2.75 2.0 2.1

Three different pilot models are compared for approach operation for 0 and 30 degree

WOD conditions (time-varying airwake model). Figures 4.17 ∼ 4.18 show the simulation

results in 0 degree WOD condition. Figures 4.19 ∼ 4.20 show the simulation results in 30

degree WOD condition. The results of position errors (Figure 4.17, Figure 4.19) indicate

that when using a lower crossover frequency, as in case 1 (relaxed) pilot model, there are

significantly larger errors in the tracking but less control activity (Figure 4.18, Figure 4.20).

This represents a situation where the pilot is under controlling, and allowing the airwake

turbulence to move the helicopter about with relatively little compensation. On the other

hand, when using a higher crossover frequency, as with the case 3 (tight), there is actually

relatively little improvement in tracking performance, but significantly more control activity.
CHAPTER 4. NUMERICAL EXAMPLES 66

This is example of a pilot over controlling the helicopter, increasing workload with relatively

little payoff in terms of holding the desired trajectory.

Case 1 (relaxed)
Case 2 (normal)
Case 3 (tight)

40

20
∆X

−20

−40
0 10 20 30 40 50 60

60

40
∆Y

20

−20
0 10 20 30 40 50 60

10

5
∆Z

−5

−10
0 10 20 30 40 50 60
Time, [sec]

Figure 4.17: Helicopter position error [ft] - 30 knot, 0 degree WOD condition
CHAPTER 4. NUMERICAL EXAMPLES 67

Case 1 (relaxed)
Case 2 (normal)
Case 3 (tight)

60
Lateral

50

40

30
0 10 20 30 40 50 60
70
Longitudinal

60

50

40
0 10 20 30 40 50 60
80
Collective

60

40

20
0 10 20 30 40 50 60
80

60
Pedal

40

20
0 10 20 30 40 50 60
Time, [sec]

Figure 4.18: Pilot control input [%] - 30 knot, 0 degree WOD condition
CHAPTER 4. NUMERICAL EXAMPLES 68

Case 1 (relaxed)
Case 2 (normal)
Case 3 (tight)

20
∆X

−20
0 10 20 30 40 50 60 70 80 90

20

10
∆Y

−10

0 10 20 30 40 50 60 70 80 90

10

0
∆Z

−10

0 10 20 30 40 50 60 70 80 90
Time, [sec]

Figure 4.19: Helicopter position error [ft] - 30 knot, 30 degree WOD condition
CHAPTER 4. NUMERICAL EXAMPLES 69

Case 1 (relaxed)
Case 2 (normal)
Case 3 (tight)

60
Lateral

50

40
0 10 20 30 40 50 60 70 80 90
Longitudinal

60

50

40
0 10 20 30 40 50 60 70 80 90

60
Collective

50

40
0 10 20 30 40 50 60 70 80 90

60
Pedal

50

40

30
0 10 20 30 40 50 60 70 80 90
Time, [sec]

Figure 4.20: Pilot control input [%] - 30 knot, 30 degree WOD condition
CHAPTER 4. NUMERICAL EXAMPLES 70

4.4 Validation with Flight Test Data

In this section, a typical shipboard approach trajectory was simulated. At this point,

measured performance and flight test data for departure tasks are unavailable for verifying

the simulation. Thus, the verification in this section is made based on the flight test data

from JSHIP program for the approach tasks only. The simulation has been performed

for two different WOD conditions (0 degree and 30 degree) with the time-varying airwake

solutions. The approach trajectory profile is modified slightly to be consistent with approach

maneuvers used in the JSHIP program. Since the exact aircraft positions with respect to

the ship are not available at this point, the initial parameters for each approach operation

are defined using the information from the flight test data, and these profile parameters are

given in Table 4.4.

Figures 4.21 - 4.24 show the simulation results for the approach in 0 degree WOD con-

dition. The dashed lines represent the flight test data from JSHIP program, the solid lines

are results from simulation model. Figure 4.21 shows helicopter air speed in knots. Figure

4.22 shows the height above ground level (representative of a radar altitude measurement

on the aircraft). The sudden jump in the time history corresponds to the aircraft flying

over the edge of the ship deck. Figure 4.23 shows the aircraft angular rate responses. The

Table 4.4: Initial profile parameters for the approach tasks from JSHIP program

Trajectory parameters 0 degree WOD 30 degree WOD


Initial altitude 280 ft 250 ft
Final altitude 60 ft (10 ft above deck) 63 ft (13 ft above deck)
Initial air speed 83.5 knot 68 knot
Final air speed 38 knot 30 knot
CHAPTER 4. NUMERICAL EXAMPLES 71

pilot stick inputs provided by the optimal control model are shown in Figure 4.24.

Similarly, Figures 4.25 - 4.28 show the simulation results for the approach operation in

30 degree WOD condition. The time domain results in Figures 4.21 - 4.28 are intended to

show that the simulation faithfully recreated the same maneuver conducted in flight test

and to provide a qualitative comparison of the transient aircraft responses and pilot control

activity.

The results show that the helicopter trajectory and speed are very similar in each case.

This is expected since the optimal control model of the pilot is designed to track these

trajectories. The oscillation in the airspeed from flight test is assumed to be sensor noise,

which was not modeled in the simulation. The 30 degree WOD condition results in sig-

nificantly larger oscillations and higher pilot control activity, particularly when helicopter

hovering over the ship deck. This reflects the so-called a cliff edge effect [60, 65], where

strong shear layers from the island are blown across the spot 8 with winds from 30 degrees.

This is backed up by qualitative results of the JSHIP flight test program, which showed

that the 30 degree WOD condition at spot 8 resulted in high pilot workload.

The attitude changes and control activity predicted by the simulation and those mea-

sured in the flight test are somewhat different in the early part of the maneuver, when the

helicopter is relatively far from the ship. This difference is likely due to the presence of the

atmospheric turbulence in the flight tests which is not modeled in this simulation. There

is also a discrepancy in the prediction of trim when the helicopter is far from the ship.

However, the results are qualitatively similar when the helicopter operates near the ship

deck, where the helicopter interacts significantly with the ship airwake.

Figures 4.24 and 4.28 are comparisons of control activity as predicted by simulation and
CHAPTER 4. NUMERICAL EXAMPLES 72

measured in the JSHIP flight tests. The 30 degree WOD condition results in significantly

larger oscillations and higher pilot control activity, particularly when the helicopter is hover-

ing over the ship deck. The control activity predicted by the simulation, and those measured

in the flight test, are somewhat different in the early part of the maneuver, when the heli-

copter is relatively far from the ship. This difference is likely to be due to the presence of

atmospheric turbulence in the flight tests, which is not modeled in this simulation.

There is also a discrepancy in the prediction of trim when the helicopter is far from the

ship, particularly in the pedals. There are two factors that contribute to this discrepancy.

First, the OCM pilot model uses a zero sideslip trim for all low speed flight conditions;

whereas the flight test pilot appears to use more of a zero bank angle trim strategy until

the airspeed is very near zero. There are also some discrepancies in the trim characteristics

between the aircraft and the simulation model. However, the pilot control activity results are

qualitatively similar when the helicopter operates near the ship deck, where the helicopter

interacts significantly with the ship airwake. These last phases of the approach maneuvers

are of the most interest. The magnitude and frequency of the collective and pedal control

activity, when the helicopter is operating near the ship deck, appear to have good qualitative

agreement with flight test.

The simulation seems to predict a somewhat lower level of longitudinal and lateral

activity. The increased level of high frequency control activity observed in flight test might

be due to a number of factors including: details in the ship airwake not captured in the

CFD solutions, the presence of vestibular feedback in the pilot feedback loop not used in the

OCM pilot model, or even biomechanical feedback effects due to vibration on the aircraft.
CHAPTER 4. NUMERICAL EXAMPLES 73

Flight test
Simulation

90

80

70
Vehicle Air Speed

60

50

40

30
0 10 20 30 40 50 60
Time, [sec]

Figure 4.21: Helicopter airspeed [knot] - 30 knot, 0 degree WOD condition


CHAPTER 4. NUMERICAL EXAMPLES 74

Flight test
Simulation

250

200
Helicopter altitude

150

100

50

0
0 10 20 30 40 50 60
Time, [sec]

Figure 4.22: Helicopter altitude [ft] - 30knot, 0 degree WOD condition


CHAPTER 4. NUMERICAL EXAMPLES 75

Flight test
Simulation

10

0
P

−5

−10
0 10 20 30 40 50 60

10

5
Q

−5

−10
0 10 20 30 40 50 60

10

5
R

−5

−10
0 10 20 30 40 50 60
Time, [sec]

Figure 4.23: Angular rate [deg/sec] - 30knot, 0 degree WOD condition


CHAPTER 4. NUMERICAL EXAMPLES 76

Flight test
Simulation

60

50
Lateral

40

30
0 10 20 30 40 50 60
80
Longitudinal

70

60

50
0 10 20 30 40 50 60
80
Collective

60

40

20
0 10 20 30 40 50 60
80
Pedal

60

40

0 10 20 30 40 50 60
Time, [sec]

Figure 4.24: Pilot stick inputs [%] - 30 knot, 0 degree WOD condition
CHAPTER 4. NUMERICAL EXAMPLES 77

Flight test
Simulation

75

70

65

60

55
Vehicle Air Speed

50

45

40

35

30

25

20
0 10 20 30 40 50 60 70 80
Time, [sec]

Figure 4.25: Helicopter airspeed [knot] - 30 knot, 30 degree WOD condition


CHAPTER 4. NUMERICAL EXAMPLES 78

Flight test
Simulation

250

200
Helicopter altitude

150

100

50

0
0 10 20 30 40 50 60 70 80
Time, [sec]

Figure 4.26: Helicopter altitude [ft] - 30 knot, 30 degree WOD condition


CHAPTER 4. NUMERICAL EXAMPLES 79

Flight test
Simulation

10
P

−10

0 10 20 30 40 50 60 70 80 90

0
Q

−5

0 10 20 30 40 50 60 70 80 90

5
0
R

−5
−10
−15
0 10 20 30 40 50 60 70 80 90
Time, [sec]

Figure 4.27: Angular rate [deg/sec] - 30knot, 30 degree WOD condition


CHAPTER 4. NUMERICAL EXAMPLES 80

Flight test
Simulation

60
Lateral

40

20
0 10 20 30 40 50 60 70 80
80
Longitudinal

60

40
0 10 20 30 40 50 60 70 80
80
Collective

60

40
0 10 20 30 40 50 60 70 80
80

60
Pedal

40

20
0 10 20 30 40 50 60 70 80
Time, [sec]

Figure 4.28: Pilot stick inputs [%] - 30 knot, 30 degree WOD condition
CHAPTER 4. NUMERICAL EXAMPLES 81

4.4.1 Frequency Domain Analysis

In general, simulation requires the adoption of a priori engineering assumptions to allow the

formulation of model equations. These simulation models are then used to predict aircraft

or subsystem motion. To achieve sufficient accuracy in simulations of helicopter-ship com-

binations, model verifications are required for simulated testing of the helicopter shipboard

operations. It is well known that a frequency domain analysis provides good correlation

between test and simulation data. The most common method of acquiring frequency in-

formation is to use spectrum averaging, i.e. averaging is performed in frequency domain.

This method has the advantages that it is relatively easy to use and it will work with

any type of signal. In the spectrum averaging method, the Fourier transform is applied to

blocks of time history data, possibly after a weighting function has been used. The Fourier

spectrum is then squared, to yield a real-valued spectrum called the autospectrum (some-

times also referred to as power spectrum or mean square spectrum). In this study, CIFER

(Comprehensive Identification from FrEquency Responses) is used to get frequency domain

comparisons against flight test data from the JSHIP program. CIFER is an interactive fa-

cility for system identification and verification developed by U.S Army/NASA and Sterling

Federal Systems [66].

Figures 4.29 - 4.30 show comparisons of input autospectra from pilot stick inputs with

those from optimal control model and flight test data. The results are averaged over 5

different simulation runs for different airwake starting points (e.g. 0 sec, 8 sec, 16 sec, 24

sec, 32 sec). In this study, 4 different window sizes (3sec, 5sec, 10sec, and 15sec) were used

in the FFT analysis to obtain a composite average that is accurate over a wide range of
CHAPTER 4. NUMERICAL EXAMPLES 82

frequencies. This is a feature available in the CIFER software package. Only the last phase

of the approach maneuver where the aircraft is interacting with the airwake, is considered.

From the figures, it can be observed that there is reasonable agreement in the collective

and pedal input autospectra for the frequency range of 0.2 to 1.8 Hz but the lateral and

longitudinal cyclic autospectra both underestimate the control activities for the frequency

region over 1 Hz. There are some additional discrepancies in the lateral control activity for

the 0 degree WOD case over the entire frequency range. Control activity in the frequency

range of 0.2 to 2 Hz has the most significant impact on pilot workload [2]. Although the

present work has provided a good initial estimate of pilot control activity, some improve-

ment is warranted. In particular, it is critical to improve the accuracy of the lateral and

longitudinal control activity predictions in the 1-2 Hz region. It is expected that this could

be achieved with further tuning of the OCM pilot model.


CHAPTER 4. NUMERICAL EXAMPLES 83

Flight test
Simulation

0
−20
Lateral

−40
−60
−80
−1 0 1
10 10 10
0
Longitudinal

−20
−40
−60
−80
−1 0 1
10 10 10
0
Collective

−40

−80
−1 0 1
10 10 10
0
−20
Pedal

−40
−60
−80
−1 0 1
10 10 10
Frequency, [Hz]

Figure 4.29: Pilot input autospectrum [dB] - 30 knot, 0 degree WOD condition
CHAPTER 4. NUMERICAL EXAMPLES 84

Flight test
Simulation

0
−20
Lateral

−40
−60
−80
−1 0 1
10 10 10
0
Longitudinal

−40

−80
−1 0 1
10 10 10
0
Collective

−20
−40
−60
−80
−1 0 1
10 10 10
0
−20
Pedal

−40
−60
−80
−1 0 1
10 10 10
Frequency, [Hz]

Figure 4.30: Pilot input autospectrum [dB] - 30 knot, 30 degree WOD condition
Chapter 5

Task-Tailored Control Design

5.1 Overview

From the previous section, shipboard helicopters operate in an environment where task

performance can be easily affected by ship airwake, which contains large velocity gradients

and areas of turbulence. In fact, ship airwake or wind-over-deck conditions can be a factor

in limiting these shipboard operations. For this reason, it would be desirable to include

task-tailored modes in the automatic flight control system that are specifically designed to

compensate for airwake disturbances.

The task-tailored control system lets a pilot fly the aircraft throughout its operational

flight envelope with optimized control augmentation supplied by the system at each oper-

ating point in day, night or adverse weather operations. Operating conditions are defined

as points in the operating space of the aircraft with dimensions including, but not lim-

ited to, mission profile, mission task, weather conditions, visual conditions, air speed, alti-

tude, glide slope, side-slip angle, attitude, g-loading, aircraft-failure state, and level-of-flight

85
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 86

control-system augmentation. The rotorcraft control laws are tailored to provide a seamless

transition between control/system modes and maintaining desired handling qualities.

The use of a task-tailored stability augmentation system could potentially improve safety

for shipboard operations and even expand the operational envelope, by allowing the aircraft

to operate in WOD conditions currently deemed unsafe. The SAS should also conform to

the handling qualities requirements currently dictated in military specifications [67]. There

are a number of technical challenges in the design of such a system, as it is possible to

approach limits on stability, robustness, and other constraints of the rotorcraft.

There have been some detailed simulation studies of the potential improvements of dis-

turbance rejection flying qualities using flight test data [34, 68, 69]. These studies mainly

focused on meeting current handling qualities specifications, and did not attempt to extend

the specification beyond the requirements in ADS-33E. Although these studies have per-

formed disturbance rejection handling qualities studies, very few quantitative or qualitative

parametric studies in the ADS-33 handling qualities requirement exist. In fact, there are

little or no supporting data for the disturbance rejection requirements in ADS-33 [67]. Fur-

thermore, design requirements for rotorcraft handling qualities involve a large set of design

specifications, including metrics based on bandwidth, cut-off frequency, actuator saturation,

and disturbance response [70]. Many innovative flight control design methods have been

proposed to improve flight control system performance. However, there is rarely an effort

to optimize the control parameters to achieve the disturbance rejection requirement.

The present work investigates the optimization of a flight control system for the UH-

60A Black Hawk helicopter operating in the turbulent airwake of a LHA class ship. A

stochastic model of the ship airwake is derived from simulations with a full time-accurate
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 87

CFD solution of the airwake. The stochastic model can be used to simplify and facilitate

off-line or real-time simulation models. It can also be readily applied for flight control

system design. The main objective is to optimize the automatic flight control system in

order to improve disturbance rejection properties of the helicopter when operating in the

airwake. This could be achieved by minimizing the magnitude of the transfer function from

the gust shaping filter input to the aircraft response. In this study, the gains of a SAS

are optimized using CONDUIT in order to ensure good handling qualities and stability,

while minimizing a weighted objective of gust response and actuator saturation. Then the

optimized SAS are tested using a full non-linear simulation model. The overall schematic

of the current task-tailored control system design is shown in Figure 5.1.

In addition, a H∞ controller is designed to provide an alternative for a helicopter au-

tomatic flight control system. It is widely recognized that a robust control design can

provide methods for addressing the control problems associated with poorly modeled sys-

Designed to fit the


spectral properties
Stochastic airwake model of the airwake
Linear
White noist shaping filter

+
Pilot stick inputs + Helicopter
Dynamics

Optimized to
reject disturbance
SAS

Figure 5.1: Task-tailored control system design scheme


CHAPTER 5. TASK-TAILORED CONTROL DESIGN 88

tems. These robust design methods use frequency information about the disturbances to

limit the system sensitivity. However, there has not been implicit consideration of the effect

that airwake disturbance would create. By incorporating practical knowledge about the

disturbance characteristics, and how it affects the real helicopter, then improvements to the

overall performance should be made.

5.2 Stochastic Airwake Modeling

From previous section, the time-accurate airwake model provided reasonable predictions of

helicopter/ship dynamic interface testing. However, the use of time-accurate ship airwake

data was found to present some practical implementation difficulties, in that the method

requires that the simulation handle large quantities of data.

For every grid point a set of time history data must be stored for each component of

velocity. Memory storage can become an issue, particularly if the simulations are to be run

in real-time, in which case accessing data from disk storage may not be feasible. It was

helpful to select a subset of the flow field when performing the simulations in which the

landing spot is known. However, for real-time simulations the pilot might want to access

different deck spots during the same simulation run.

The use of stochastic airwake model (based on the time-accurate CFD results) might

be an attractive alternative. The stochastic airwake model can be designed using shaping

filters based on the statistical characteristics of the turbulent airwake. This approach of

finding an approximate airwake model promises to provide a better real-time application

capability, and will ultimately be used to optimize the automatic flight control system in
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 89

order to improve disturbance rejection properties of the aircraft.

Methods of simulating the effects of gust on rotorcraft range from the straight forward

approach of superimposing frozen-field turbulence models at the vehicle center of gravity

(the von Karman and Dryden approximations), as in fixed-wing aircraft, to complex rotating

frame turbulence models (SORBET) [40]. Although including gust velocities at the center

of gravity has obtained favorable pilot comments at high speeds, as the aircraft speed is

decreased, this type of gust model has been criticized for its high frequency content and

lack of variation. Improved pilot comment of simulated hover/low speed turbulence has

been achieved through the implementation of complex rotating frame turbulence models

[33]. However, this type of model is not well suited for use in control system design.

To simulate the effects of an empirical turbulence on a rotorcraft, Labows developed a

simple, empirically based turbulence model for UH-60 helicopter [34]. The turbulence model

used white noise driven filters that were scalable with wind speed and turbulence intensity,

to generate equivalent turbulent control inputs. These control inputs could then be fed

directly into the aircraft to create equivalent actuator data traces, which generated aircraft

roll and pitch rates that had spectral characteristics that were comparable to the spectral

characteristics of measured rotorcraft rates from flight test in two levels of atmospheric

turbulence. A technical approach similar to that used by Labows to extract equivalent

airwake model is employed in this study.

In this study, the features of the airwake that primarily affect the flying qualities of

the helicopter is wanted to characterize, while ignoring higher-order flow features that may

only be responsible for vibration and other high frequency effects. Thus, the airwake is

represented as a disturbance vector of 3 velocity and 3 angular rate components, similar


CHAPTER 5. TASK-TAILORED CONTROL DESIGN 90

to the von Karman turbulence model. The current disturbance modeling process is similar

to the process in Reference [34]. In that study, a disturbance model was developed by

extracting the aircraft remnant rates due to atmospheric turbulence from flight test data.

The remnant rates were then put into an inverse model of the aircraft to create equivalent

control inputs that could then be fed into the aircraft actuators to simulate response to

turbulence.

The present disturbance modeling effort is different in that the equivalent disturbances

are expressed in terms of body velocities and angular rates, and that the effort focuses

specifically on modeling the gust velocities due to the turbulent wake of a ship’s super-

structure. In this study, the nonlinear helicopter/ship DI simulation model discussed in the

previous section is used to extract the ship airwake disturbances. Hover tasks for 30 knot, 0

degree and 30 degree WOD conditions were conducted over landing spot 8 on a LHA class

ship (Figure 4.1). Similar to the method used in Reference [34], the first step in modeling

procedure is to extract the remnant aircraft rates caused by the time-varying ship airwake.

The remnant rates are then filtered to reduce the effects of low frequency drift and high

frequency noise. The overall schematic of current modeling process is shown in Figure 5.2.

A 9 state linearized model is used to create an inverse model in order to extract equivalent

disturbances that recreate similar aircraft responses as the full airwake. The 9 rigid body

state linear model and corresponding linear model without gust effects are given by

ẋ = Ax + Bu + Gw (5.1)

ẋng = Axng + Bu (5.2)


CHAPTER 5. TASK-TAILORED CONTROL DESIGN 91

Recorded UH-60
Helicopter
control position dynamic
remnant rates
model +

Recorded
helicopter rate
responses Inverse model
Step 1 of UH-60

Step 2
Step 3
Check Design Equivalent airwake
pilot inputs spectral filter disturbance that
with CIFER cause remnant motions

Figure 5.2: Derivation of stochastic airwake disturbances

where x = [u, v, w, p, q, r, φ, θ, ψ]T is rigid body states, w = [ug , vg , wg , pg , qg , rg ]T is the

equivalent airwake disturbance vector, and G represents a 9×6 gust matrix. By subtracting

Equation (5.1) from Equation (5.2), the remnant state model can be written as

ṙ = Ar + Gw (5.3)

where r = x − xng . Since G is not square, the equivalent disturbances can be obtained

using a pseudo-inverse method:

w = G+ (ṙ − Ar) (5.4)

where G+ is the left inverse of G. The resulting disturbance vector gives the best least

squares fit of the overall effect of the airwake on the aircraft. The three velocity components
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 92

of the gust vector represent the average gust velocity over the body of the aircraft, while

the three angular velocity components represent a linear variation of the gust field over the

body of the aircraft. More complex non-linear variation of the gust field over the body of

the aircraft may not be captured, but it has been shown that this model would be sufficient

for workload analysis [71].

The equivalent disturbance model is verified using autospectrum of aircraft roll, pitch,

and yaw rate responses. There appears to be good agreement with the responses from the

higher-order simulation model as presented in Figures 5.3, 5.4. In addition, the resultant

pilot control inputs are analyzed using CIFER. The autospectra of pilot inputs caused by

the equivalent airwake model are compared to the original control responses of full-time

varying airwake model (Figures 5.5, 5.6). The results (Figures 5.3 - 5.6) are averaged over

5 different simulation runs for different airwake starting points (e.g. 0 sec, 8 sec, 16 sec,

24 sec, 32 sec). Comparisons indicated that the equivalent airwake produced very similar

frequency content of full-time varying airwake.

The final step of the overall modeling process, is to simulate the airwake disturbances

by passing zero mean white noise with variance one through spectral filters whose transfer

function yield the desired power spectral density (PSD). In this study, the power spectral

density function is based on the von Karman turbulence model and is modified to represent

the ship airwake. The filters are approximations of the von Karman velocity-spectra that

are valid in a range of normalized frequencies of less than 50 radians. These filters can be

found in both the Military Handbook MIL-HDBK-1797 and Reference [72].


CHAPTER 5. TASK-TAILORED CONTROL DESIGN 93

Equivalent airwake
Time-varying airwake

−20
P

−40

−60
0 1
10 10
0

−20
Q

−40

−60
0 1
10 10

−20

−40
R

−60

−80
0 1
10 10
Frequency, [rad/sec]

Figure 5.3: Comparisons of aircraft angular rates [dB] (time-varying airwake vs. equivalent
airwake) - 0 degree WOD condition
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 94

Equivalent airwake
Time-varying airwake

20

0
P

−20

−40
0 1
10 10
20

0
Q

−20

−40
0 1
10 10
20

0
R

−20

−40
0 1
10 10
Frequency, [rad/sec]

Figure 5.4: Comparisons of aircraft angular rates [dB] (time-varying airwake vs. equivalent
airwake) - 30 degree WOD condition
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 95

Equivalent airwake
Time-varying airwake

−20
Lateral

−40

−60

−80
0 1
10 10
−20
Longitudinal

−40

−60

−80
0 1
10 10

−20
Collective

−40

−60

−80
0 1
10 10
−20

−40
Pedal

−60

−80
0 1
10 10
Frequency, [rad/sec]

Figure 5.5: Comparisons of pilot inputs [dB] (time-varying airwake vs. equivalent airwake)
- 0 degree WOD condition
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 96

Equivalent airwake
Time-varying airwake

0
−20
Lateral

−40
−60
−80
0 1
10 10
0
Longitudinal

−20
−40
−60
−80
0 1
10 10
0
Collective

−20
−40
−60
−80
0 1
10 10
0
−20
Pedal

−40
−60
−80
0 1
10 10
Frequency, [rad/sec]

Figure 5.6: Comparisons of pilot inputs [dB] (time-varying airwake vs. equivalent airwake)
- 30 degree WOD condition
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 97

For example, the transfer function for the vertical gust component from von Karman

turbulence model is given by:

q   2 
Lw
4σw V 1 + 2.7478 LVw s + 0.3398 Lw
V s
Hw (s) =  2  3 (5.5)
1 + 2.9958 LVw s + 1.9754 Lw
V s + 0.1539 Lw
V s

where Lw represents scale length, the variable σw represents turbulence intensity, and V

is the reference airspeed. Because the von Karman model is typically used to represent

atmospheric turbulence at higher altitudes and speeds, it is not appropriate to model the

ship airwake. Thus, Equation (5.5) is slightly modified to produce the ship airwake power

spectral density for a white noise input.

q   2 
Lw
4σw V 1 + b1 LVw s + b2 Lw
V s
Hw (s) =  2  3 (5.6)
1+ a1 LVw s + a2 Lw
V s + a3 Lw
V s

where the coefficients an (n = 1, 2, 3), bm (m = 1, 2) are obtained from a best fit to the

vertical airwake gust PSD data using nonlinear least-square fitting algorithm automated in

MATLAB.

Figure 5.7 shows the power spectral density for the vertical component of airwake dis-

turbance model for the 30 knots, 0 degree WOD condition (17 ft above landing spot 8).

The results shown in this paper are for V = 30 knots (50.6343 ft/sec) since the aircraft is

hovering in a 30 knot relative wind. And the coefficients of the “best fit” spectral filter in

Equation (5.6) are Lw = 12.56 ft, σw = 16.03 ft/sec, a1 = 11.01, a2 = 7.66, a3 = 9.78,

b1 = 0.56, and b2 = 0.13. The resulting PSD of the spectral filter overlays well in the

region of frequency range of 0.2 ∼ 20 rad/sec. A similar process is applied to the other
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 98

five airwake components. Figures 5.8 - 5.13 show the resulting PSD of 6 gust components

([ug , vg , wg , pg , qg , rg ]) for 30 knot, 30 degree WOD condition. Table 5.1 shows the final gust

shaping filters for 0 and 30 degree WOD conditions.

The white noise source utilized herein is a random number generator with a mean of

zero and a variance of 1 for each airwake component. The resulting autospectra of angular

rate responses for 0 degree and 30 degree WOD conditions are plotted in Figures 5.14 -

5.15. The autospectra of pilot inputs caused by the stochastic airwake model are compared

to the original control responses of full-time varying airwake model (Figures 5.16, 5.17).

The resulting frequency domain analyses show good fits over the frequency range of 0.4 ∼

10.0 rad/sec, supporting the current stochastic airwake modeling scheme.


PSD of vertical gust component, [(ft/sec) /(rad/sec)]

Best Fit Spectral Filter


2

Extracted from simulation


with full time-varying airwake

Frequency, [rad/sec]

Figure 5.7: Power spectral density for vertical airwake disturbance component
(30 knot, 0 degree WOD condition)
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 99

PSD of longitudinal gust component, [(ft/sec) /(rad/sec)]

Best Fit Spectral Filter


2

Extracted from simulation


with full time-varying airwake

Frequency, [rad/sec]

Figure 5.8: Power spectral density of longitudinal airwake disturbance component


(30 knot, 30 degree WOD condition)
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 100

PSD of lateral gust component, [(ft/sec) /(rad/sec)]

Best Fit Spectral Filter


2

Extracted from simulation


with full time-varying airwake

Frequency, [rad/sec]

Figure 5.9: Power spectral density of lateral airwake disturbance component


(30 knot, 30 degree WOD condition)
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 101

PSD of vertical gust component, [(ft/sec) /(rad/sec)]

Best Fit Spectral Filter


2

Extracted from simulation


with full time-varying airwake

Frequency, [rad/sec]

Figure 5.10: Power spectral density of vertical airwake disturbance component


(30 knot, 30 degree WOD condition)
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 102

PSD of roll gust component, [(rad/sec) /(rad/sec)]

Best Fit Spectral Filter


2

Extracted from simulation


with full time-varying airwake

Frequency, [rad/sec]

Figure 5.11: Power spectral density of roll airwake disturbance component


(30 knot, 30 degree WOD condition)
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 103

PSD of pitch gust component, [(rad/sec) /(rad/sec)]

Best Fit Spectral Filter


2

Extracted from simulation


with full time-varying airwake

Frequency, [rad/sec]

Figure 5.12: Power spectral density of pitch airwake disturbance component


(30 knot, 30 degree WOD condition)
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 104

PSD of yaw gust component, [(rad/sec) /(rad/sec)]

Best Fit Spectral Filter


2

Extracted from simulation


with full time-varying airwake

Frequency, [rad/sec]

Figure 5.13: Power spectral density of yaw airwake disturbance component


(30 knot, 30 degree WOD condition)
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 105

Table 5.1: Gust shaping filters for 0 degree and 30 degree WOD conditions

Gust component 0 degree WOD 30 degree WOD

0.1042s2 +2.3588s+19.1603 14.2261s2 +306.1786s+198.2738


Hu (s) 0.1075s3 +.1741s2 +2.5696s+1 2.8484s3 +2.3466s2 +11.2279s+1

0.0838s2 +0.5626s+12.9292 0.3103s2 +3.6789s+94.2622


Hv (s) 0.0402s3 +0.2108s2 +1.2048s+1 0.0257s3 +0.4211s2 +1.0154s+1

0.2475s2 +4.4454s+31.9292 11.3834s2 +409.0059s+433.3580


Hw (s) 0.1494s3 +0.4718s2 +2.7311s+1 1.6950s3 +0.9448s2 +12.5721s+1

1.2591s2 +24.3807s+0.3617 0.0628s2 +0.4665s+25.4193


Hp (s) 105.3086s3 +94.7148s2 +738.2469s+1 0.4983s3 +5.5760s2 +20.4112s+1

3.3859s2 +49.8481s+0.5244 0.0475s2 +0.3093s+11.5695


Hq (s) 192.6s3 +279.4s2 +1015.8s+1 0.1955s3 +2.5249s2 +7.0773s+1

0.0763s2 +0.6526s+9.4927 0.0115s2 +0.1520s+4.7283


Hr (s) 0.7871s3 +1.8020s2 +24.0094s+1 0.0069s3 +0.2645s2 +0.3370s+1
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 106

Stochastic airwake
Equivalent airwake
Time-varying airwake

−20
P

−40

−60
0 1
10 10
0

−20
Q

−40

−60
0 1
10 10
0

−20
R

−40

−60
0 1
10 10
Frequency, [rad/sec]

Figure 5.14: Comparisons of aircraft angular rates [dB] (time-varying airwake vs. equivalent
airwake vs. stochastic airwake) - 30 knot, 0 degree WOD condition
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 107

Stochastic airwake
Equivalent airwake
Time-varying airwake

20

−20
P

−40

−60
0 1
10 10
20

−20
Q

−40

−60
0 1
10 10
20

−20
R

−40

−60
0 1
10 10
Frequency, [rad/sec]

Figure 5.15: Comparisons of aircraft angular rates [dB] (time-varying airwake vs. equivalent
airwake vs. stochastic airwake) - 30 knot, 30 degree WOD condition
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 108

Stochastic airwake
Equivalent airwake
Time-varying airwake

0
−20
Lateral

−40
−60
−80
0 1
10 10
0
Longitudinal

−20
−40
−60
−80
0 1
10 10
0
Collective

−20
−40
−60
−80
0 1
10 10
0
−20
Pedal

−40
−60
−80
0 1
10 10
Frequency, [rad/sec]

Figure 5.16: Comparisons of pilot inputs [dB] (time-varying airwake vs. equivalent airwake
vs. stochastic airwake) - 30 knot, 0 degree WOD condition
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 109

Stochastic airwake
Equivalent airwake
Time-varying airwake

0
Lateral

−20
−40
−60
−80
0 1
10 10
0
Longitudinal

−20
−40
−60
−80
0 1
10 10
0
Collective

−20
−40
−60
−80
0 1
10 10
0
−20
Pedal

−40
−60
−80
0 1
10 10
Frequency, [rad/sec]

Figure 5.17: Comparisons of pilot inputs [dB] (time-varying airwake vs. equivalent airwake
vs. stochastic airwake) - 30 knot, 30 degree WOD condition
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 110

5.3 Optimization of a Stability Augmentation System

Previous studies showed that with an increasing magnitude of disturbance responses, an

increasing pilot compensation level was required to achieve desired task performance [34,

68, 69]. If the flight control system has satisfactory handling qualities in a disturbance-

free environment, these results indicate that to meet desired performance in a turbulent

environment an additional design criteria must be required. Currently, the disturbance

rejection requirements in ADS-33E-PRF state that roll, pitch, and yaw responses to control

inputs shall meet the bandwidth threshold limits based on aircraft response to pilot stick

inputs [67]. However, there has not been an effort to optimize the control parameters to

reduce pilot workload.

An effort to optimize the automatic flight control system is attempted in order to improve

disturbance rejection properties of the helicopter when operating in the turbulent ship

airwake. This can be achieved by minimizing the magnitude of the transfer function from

the gust shaping filter input to the aircraft response. In this work, the gains of a basic

stability augmentation system are optimized using CONDUIT in order to ensure good

handling-qualities and stability, while minimizing a weighted objective of gust response

and actuator energy. CONDUIT is a “state-of-the-art” computational tool for integrating

simulation models and control law architectures with design specifications and constraints

for modern fixed-wing and rotary-wing aircraft. In addition, CONDUIT allows for the

optimization of multiple objectives with multiple constraints by tuning a set of selected

design parameters (e.g. controller gains, time constants, etc.). Details of the CONDUIT

design environment can be found in References [73] and [74].


CHAPTER 5. TASK-TAILORED CONTROL DESIGN 111

The case problem is based on the flight dynamics of a UH-60 Black Hawk helicopter

(Figure 5.18). The key elements of the block diagram are:

• 24-state linear model of UH-60

• ship airwake spectral filters

• helicopter SAS

• washed-out pitch and yaw channel feedback

• analog-to-digital filter for roll, pitch, yaw sensors

• longitudinal acceleration feedback control

• pitch attitude feedback control

In this study, the digital SAS of the UH-60A is used as a starting design point and for

comparisons in this study (it is assumed to have 10% authority). In low speed mode, this

SAS features roll, pitch, and yaw rate feedback through separate SAS channels. The roll

SAS also includes limited authority roll attitude feedback. The compensators use a rate plus

lagged rate feedback approach, which is essentially equivalent to a phase-lag compensator.

The pitch and yaw channels also include washout filters to reduce steady-state feedback in

prolonged maneuvers.

In the modified SAS, longitudinal acceleration feedback and pitch attitude feedback are

added as shown in Figure 5.18. The acceleration feedback is expected to improve gust

response, while pitch attitude feedback is added to provide closed-loop stability at low

speed. Figure 5.19 shows a schematic of the modified SAS architecture. The compensators
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 112

Figure 5.18: Augmented plant model for a SAS optimization

are put in the classical phase-lag form. In addition, a phase lead/lag type controller is added

to the roll channel. The SAS gains, the lead/lag time constants, and the pitch attitude and

longitudinal acceleration feedback parameters are all selected as design parameters to be

optimized using CONDUIT. In the diagrams, these design parameters include the prefix

“dpp ”. The analog-to-digital filters and washout filters from the original SAS are retained

and not modified in the optimization process.

In this paper, four design specs are selected from the CONDUIT libraries as constraints.

The relative priority of each spec is designated as indicated by “Hard specifications (H)”

indicated in the upper right corner of the spec. The role of the spec priority in the CON-

DUIT optimization process is described fully in Reference [73]. In summary, the “hard
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 113

Figure 5.19: Modified SAS configuration

specification” selected for this study are:

• Crossover frequencies for the individual broken loops (CrsLnG1) :

This specification is intended as an objective to minimize crossover frequency in phase

3. The boundaries should be set to ensure that the design is in the Level 1 region for

phase 1 and 2.

• All closed-loop eigenvalues (absolute stability) (EigLcG1) :

This criterion is used to ensure that all the real parts of the eigenvalues of the system

are zero or negative, ensuring that all the dynamics are stable or neutrally stable. At

any given iteration, the sum of unstable eigenvlaue real parts or the largest stable

eigenvalue is returned as the spec metric.

• Gain/phase margin requirement (StbMgG1) :

This spec has logic for treating stable, conditionally stable, and unstable systems. It
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 114

also has logic for correctly accounting for right-half plane poles and zeros. A table of

margins is built for all crossings of the 0db and -180 deg lines and displayed in the

supporting plot. The spec returns the minimum gain and phase margin values from

the table. The Level I boundaries are taken from MIL-F-9490D. In this document

that required margins depend on the frequency of the first aeroelastic mode and on

the airspeed. In the CONDUIT gain/phase margin spec, the requirements for rigid

body modes is implemented. Stability margin specs for other frequency ranges are

easily implemented by shifting the splines.

• Bandwidth requirement for roll/pitch axes (BnwAtH1) :

The pitch (roll) response to longitudinal (lateral) cockpit control force or position

inputs shall meet the limits specified. It is desirable to meet this criterion for both

controller force and position inputs. If the bandwidth for force inputs falls outside

the specified limits, flight testing should be conducted to determine that the force feel

system is not excessively sluggish.

The crossover frequency spec (CrsLnG1) is generally intended as an objective to minimize

feedback control activity in the last phase of optimization. However, the current design ob-

jective is not to minimize control activity, but rather to minimize disturbance response while

retaining similar crossover frequencies as the original SAS. Thus, the crossover frequency

spec is enforced as constraints in the optimization.

A new spec (DisRnL1) for disturbance rejection is designed based on the PSD of angular

rate response to corresponding gust input. Here, the design parameters are tuned to attempt

to minimize the magnitude of the transfer function from the gust shaping filter input to the
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 115

Magnitude [dB]
Level III
Transfer function
White Level II
H (s) =
q( s) PSD
noise
qg (s)

Level I
Frequency [rad/sec]

Figure 5.20: A new disturbance rejection spec design (ex. pitch axis)

aircraft response. Figure 5.20 shows a schematic of the proposed disturbance rejection spec.

This new spec is designated as a “summed objective specification (J)”. The use of summed

objective allows the optimization process to improve a set specific performance objectives

of the controllers while maintaining compliance with Level 1 requirements. Currently, the

boundaries for Level 1/Level 2 and Level 2/Level 3 are selected arbitrarily for this spec

since there are no supporting data for the disturbance rejection requirements in ADS-33

at this time. This is sufficient for the current analysis, the level boundaries are simply

used to provide a measure of how well gust disturbances are rejected and are not used as

constraints. Note that the new spec for roll axis has more generous level boundaries as the

aircraft is inherently more sensitive in roll due to lower inertia in that axis.

Figure 5.21 shows the performance of the original SAS for the selected design specs.

The blue region reflects Level 1 handling qualities ratings, the magenta region represents

Level 2, and the red region reflects Level 3 handling qualities. Note that the basic UH-

60A SAS does not fully stabilize the aircraft in hover and low speed flight. There is a

low frequency unstable mode. The mode can be stabilized by the outer loop Flight Path

Stabilization system (FPS) which is not considered in this analysis. Closed loop stability
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 116

is not necessarily required to achieve Level 1 or Level 2 handling qualities, but nonetheless

the eigenvalues and stability margin (pitch axis) specs identify any instability as “Level 3”.

The disturbance rejection requirements defined for this study show Level 3 for pitch and

yaw axes and Level 2 behavior for roll axis, although as noted before these boundaries are

somewhat arbitrary.

For the optimization process, only the 30 degree WOD condition is considered. The

optimized SAS is then tested using non-linear simulation for both 0 and 30 degree WOD

conditions. This is a logical approach since the 30 degree WOD condition resulted in

significant higher pilot workload [60].

After several iterations CONDUIT reaches the final phase, which is a “feasible solution”

where all specs are in the Level 1 region. Figure 5.22 shows the fully converged result. The

CrsLnG1:Crossover Freq. EigLcG1: StbMgG1: Gain/Phase Margins


Roll (1) Roll (1) (linear scale) Eigenvalues (All) (rigid−body freq. range)
H H H
80
Pitch (2) Pitch (2)
Yaw (3) Yaw (3) 60

PM [deg]
CrsLnG1 (1) 40

CrsLnG1 (2)
20
CrsLnG1 (3)
EigLcG1 (1) Ames Research Center Ames Research Center 0 MIL−F−9490D
0 5 10 15 20 −1 0 1 0 10 20
EigLcG1 (2) Crossover Frequency [rad/sec] Real Axis GM [db]

EigLcG1 (3) BnwAtH1:Bandwith (pitch & roll) DisRnL2:Gust Response DisRnL1:Gust Response
Other MTEs;UCE>1; Div Att Roll Pitch/Yaw
0.4 40 40
StbMgG1 (1) H J J
20
StbMgG1 (2) 0.3
20
0
Phase delay [sec]

Magnitude [db]

StbMgG1 (3)
Magnitude [db]

0 −20
BnwAtH1 (1) 0.2
−20 −40
BnwAtH1 (2)
−60
0.1
DisRnL1 (2) −40
−80
DisRnL1 (3)
0 ADS−33D −60 PSU RCOE −100 PSU RCOE
0 1 2 3 4 0 1 0 1
DisRnL2 (1) Bandwidth [rad/sec]
10 10 10 10
Frequency [rad/sec] Frequency [rad/sec]

Figure 5.21: HQ windows for the original SAS configuration - 30 knot, 30 degree WOD
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 117

disturbance rejection properties are improved and moved into Level 1 for the gust rejection

spec. Closed loop stability is also achieved, due to the addition of pitch attitude feedback.

The crossover frequency for each axis is increased somewhat. The optimization effectively

resulted in higher gain in all three axes, so it will be necessary to check that the increased

gain does not result in rate or position saturations. Note that the stability margins in

the roll axis are reduced and are against the 6 dB and 45◦ gain and phase margin Level

1 constraints. This is because the optimization tends to increase the roll axis gain until

it hit stability margin limits. The stability limits on maximum roll gain are due to rotor-

body coupling issues that are typically observed on helicopters with articulated rotors. The

disturbance rejection requirements would probably drive the crossover frequencies and roll

and yaw feedback gains higher if it were not for this stability constraint. It is found that

CrsLnG1:Crossover Freq. EigLcG1: StbMgG1: Gain/Phase Margins


Roll (1) Roll (1) (linear scale) Eigenvalues (All) (rigid−body freq. range)
H H H
Pitch (2) Pitch (2) 80

Yaw (3) Yaw (3) 60

PM [deg]
CrsLnG1 (1) 40
CrsLnG1 (2)
20
CrsLnG1 (3)
EigLcG1 (1) Ames Research Center Ames Research Center 0 MIL−F−9490D
0 5 10 15 20 −1 0 1 0 10 20
EigLcG1 (2) Crossover Frequency [rad/sec] Real Axis GM [db]

EigLcG1 (3) BnwAtH1:Bandwith (pitch & roll) DisRnL2:Gust Response DisRnL1:Gust Response
Other MTEs;UCE>1; Div Att Roll Pitch/Yaw
0.4 40 40
StbMgG1 (1) H J J
20
StbMgG1 (2) 20
0.3
0
Phase delay [sec]

Magnitude [db]

StbMgG1 (3)
Magnitude [db]

0 −20
BnwAtH1 (1) 0.2
−20 −40
BnwAtH1 (2)
−60
DisRnL1 (2) 0.1
−40
−80
DisRnL1 (3)
0 ADS−33D −60 PSU RCOE −100 PSU RCOE
0 1 0 1
DisRnL2 (1) 0 1 2 3 4 10 10 10 10
Bandwidth [rad/sec] Frequency [rad/sec] Frequency [rad/sec]

Figure 5.22: HQ windows for the optimized SAS configuration - 30 knot, 30 degree WOD
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 118

using a lead-lag type compensator resulted in some improvement in the optimization of the

roll axis, as compared to the original lag compensator.

The ability to achieve an optimized solution is quite sensitive to the initial values for the

design parameters. If the design parameters are started too far from a feasible solution, the

optimization tends to wander and often does not reach a satisfactory result because there

are too many design parameters to tune. It is possible that the problem is somewhat over-

parameterized. Future work might focus on reducing the number of controller parameters

to achieve a more robust optimization.

The optimized SAS is converted to discrete form and tested using the non-linear simula-

tion model for the hovering operation on the spot 8 (included time-varying airwake solutions

for 0 and 30 degree WOD conditions). Figure 5.23 and 5.26 show the responses of aircraft

angular rate for 0 degree and 30 degree WOD conditions. The solid lines represent the

results with optimized SAS, the dotted lines are simulation results with the baseline SAS

configuration. The corresponding pilot stick inputs (generated by an optimal control model

of the human pilot) are shown in Figures 5.24 and 5.27. The results with an optimized SAS

show some significant improvement over the original SAS. The aircraft roll rate and lateral

stick input are only slightly improved, due to the limits on the roll axis gain discussed

above and the fact that the aircraft gust response is more sensitive in roll. Nonetheless the

overall aircraft angular rates and pilot control inputs are reduced with the optimized SAS

for both 0 and 30 degree WOD conditions. Note that there is no attempt to optimize the

control system for heave axis. Figures 5.25 and 5.28 show the response of SAS outputs. The

optimized SAS appears to result in slightly higher SAS actuator activity in the 0 degree

WOD condition, while the magnitude and frequency of the SAS actuator activity in the
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 119

30 degree WOD condition is similar to the baseline SAS. In both cases, the SAS actuators

stayed well within the rate and position saturation limits. Note that the difference in the

roll SAS outputs in 0 degree WOD case is mainly due to the small roll attitude feedback

gain from the optimized SAS.

Figures 5.29 and 5.30 show comparisons of input autospectra from pilot stick inputs. In

this study, 4 different window sizes (3, 5, 10, and 15 seconds) are used in the FFT analysis

to obtain a composite average that is accurate over a range of frequencies. From the figures,

it can be observed that the optimized SAS shows some improvement over the original SAS

configuration in the frequency range of 1 to 10 rad/sec. It is generally recognized that control

activity in the frequency range of 0.2 to 2 Hz (about 1.2 ∼ 12 rad/sec) has significant impact

on pilot workload [60].


CHAPTER 5. TASK-TAILORED CONTROL DESIGN 120

Results with the original SAS


Results with the optimized SAS

0.5

0
P

−0.5

−1

0 5 10 15 20 25 30 35 40

0.5
Q

−0.5
0 5 10 15 20 25 30 35 40

0.4

0.2

0
R

−0.2

−0.4
0 5 10 15 20 25 30 35 40
Time, [sec]

Figure 5.23: Aircraft angular rate responses [deg/sec] - 30 knot, 0 degree WOD
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 121

Results with the original SAS


Results with the optimized SAS

0.5
Lateral

−0.5

0 5 10 15 20 25 30 35 40
0.5
Longitudinal

0
−0.5
−1

0 5 10 15 20 25 30 35 40
Collective

−1
0 5 10 15 20 25 30 35 40

0.5
Pedal

−0.5

0 5 10 15 20 25 30 35 40
Time, [sec]

Figure 5.24: Pilot control stick inputs [%] - 30 knot, 0 degree WOD
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 122

Results with the original SAS


Results with the optimized SAS

1.5

1
RSAS

0.5

−0.5
0 5 10 15 20 25 30 35 40

0
PSAS

−1

−2
0 5 10 15 20 25 30 35 40

0.5
YSAS

−0.5

−1
0 5 10 15 20 25 30 35 40
Time, [sec]

Figure 5.25: SAS outputs [%] - 30 knot, 0 degree WOD


CHAPTER 5. TASK-TAILORED CONTROL DESIGN 123

Results with the original SAS


Results with the optimized SAS

4
2
0
P

−2
−4
−6

0 5 10 15 20 25 30 35 40

2
Q

−2

−4
0 5 10 15 20 25 30 35 40

2
R

−2

−4
0 5 10 15 20 25 30 35 40
Time, [sec]

Figure 5.26: Aircraft angular rate responses [deg/sec] - 30 knot, 30 degree WOD
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 124

Results with the original SAS


Results with the optimized SAS

5
Lateral

−5
0 5 10 15 20 25 30 35 40
4
Longitudinal

2
0
−2
−4
−6
0 5 10 15 20 25 30 35 40

5
Collective

0
−5
−10
0 5 10 15 20 25 30 35 40
4
2
0
Pedal

−2
−4
−6
−8
0 5 10 15 20 25 30 35 40
Time, [sec]

Figure 5.27: Pilot control stick inputs [%] - 30 knot, 30 degree WOD
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 125

Results with the original SAS


Results with the optimized SAS

10

5
RSAS

−5
0 5 10 15 20 25 30 35 40

0
PSAS

−5

−10
0 5 10 15 20 25 30 35 40

10

5
YSAS

−5

−10
0 5 10 15 20 25 30 35 40
Time, [sec]

Figure 5.28: SAS outputs [%] - 30 knot, 30 degree WOD


CHAPTER 5. TASK-TAILORED CONTROL DESIGN 126

Results with the original SAS


Results with the optimized SAS

Figure 5.29: Control stick input autospectra [dB] - 30 knot, 0 degree WOD
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 127

Results with the original SAS


Results with the optimized SAS

Figure 5.30: Control stick input autospectra [dB] - 30 knot, 30 degree WOD
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 128

5.4 H∞ Control of a Helicopter SAS

It is widely agreed that high levels of feedback augmentation will be needed, if the enhanced

flying qualities required of the next generation or current rotorcraft are to be realized. The

use of modern multivariable control analysis tools such as robust control method allows

a more rigorous analysis and, consequently, improved stability robustness of the closed-

loop system. The design of robust control laws for a helicopter is, however, difficult due to

several factors: parametric and structural uncertainty in the model caused by the limitation

of current flight-mechanic models to predict helicopter response to control and disturbance

inputs.

Robust control theory has evolved since the early 1980s and provides methods for ad-

dressing the control problems associated with poorly modeled systems. Methodologies such

as H∞ optimization now provide systematic procedures for designing robust controllers for

multivariable system. The H∞ design method generates optimal controllers that provide

good performance even when the characteristics of the plant are not modeled exactly or are

varying periodically with time. Frequency-dependent weighting functions are included in

the performance index. It is not necessary to measure or estimate the entire state vector.

The singular value loop shapes are directly prescribed to meet performance and robustness

bounds formulated in the frequency domain.

Recently, the DERA (Defence Evaluation and Research Agency) Bedford have been

collaborating to narrow the gap between the theoretical developments in Robust Control

and the more practical problems associated with designing helicopter flight control systems

meeting exacting flying-qualities requirements in ADS-33 [75, 76, 77]. Earlier work from
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 129

DERA has focused on the Westland Lynx. Yue and Postlethwaite designed H∞ based

controller based on the HELISIM model. That work led to successful piloted simulation,

through which the potential of H∞ based design methods was clearly demonstrated [75].

Building on those results, Walker et al. designed single and two-degree-of-freedom H∞

controllers which were extensively evaluated in the DERA Bedford Large Motion simulator

[76, 77]. Desired handling qualities ratings were consistently obtained during aggressively

performed Mission Task Elements in ground-based simulation on the Lynx model of a

series of two-degree-of-freedom controllers. The limitations of ground-based simulation

were, however, recognized, and it was concluded that appropriately validated mathematical

models containing higher order dynamics would be important if similar results were to be

replicated in flight [77].

In this study, the design of flight controller on the UH-60A is described. It presents a

critical assessment of the use of H∞ control in the design of robust flight controllers for the

UH-60A operating in the turbulent ship airwake conditions. The emphasis in the design

is to meet stringent U.S. Army handling qualities specifications (ADS-33, [67]) against the

constraint of robust stability to plant disturbances. The brief review of H∞ controller

design will be given in the section 5.4.1. In Section 5.4.2, the helicopter control problem is

formulated and evaluated using nonlinear DI simulation program.

5.4.1 Review of H∞ Control Design Method

Every control system is constantly subject to command-disturbance input uncertainties

and plant uncertainties. The issue of robust control is to design a controller such that

the closed-loop system remains stable for all possible plant perturbations, and that the
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 130

response is admissible for every disturbance and command in the prescribed set under all

possible plant perturbations. The prescribed set of disturbances and commands is modeled

by the designer according to the actual environment of the system. For many practical

control problems, the external input signals are not known precisely, but instead belong to

a prescribed norm-bounded (energy-bounded) set. In this case, it is more meaningful for a

design engineer to minimize the maximum error (worst case) that can occur subject to all

possible input signals belonging to this set. This min-max approach (H∞ control law) was

introduced for feedback design from a frequency-domain point of view [78]. Since then, this

research area has attracted many researchers and significant progress has been made. The

basic formulation can be found in various textbooks and articles [7, 78, 79].

Mathematically, H∞ controller design is a frequency-dependent optimization problem.

Therefore all design specifications need to expressed in the frequency domain. This is not

always straightforward, particularly for time-domain specifications, but the approach is

ideally suited for robustness considerations. In the optimization procedure, a controller is

selected which stabilizes a nominal plant model and minimizes the energy gain (H∞ norm)

of a closed-loop transfer function which describes the design objectives.

Figure 5.31 shows the standard compensator configuration. In here, G(s) is a system

with two kinds of inputs and two kinds of outputs. The input ω is an exogenous input

representing the disturbance acting on the system. The output z is an output of the

system, whose dependence on the exogenous input ω to be minimized. The output y is a

measurement, which is used to choose the input u, which in turn is the tool to minimize

the effect of ω on z. A constraint is that this mapping from y to u should be such that the

close-loop system is internally stable. This is quite natural since the states are not wanted
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 131

to become too large while regulating the performance. The effect of ω on z after closing the

loop is measured in terms of the energy and the worst disturbance ω. The measurement,

which will turn out to be equal to the closed-loop H∞ norm, is the supremum over all

disturbances unequal to zero of the quotient of the energy flowing out of the system and

the energy flowing into the system.

w z
G(s)

u K(s)
y

Figure 5.31: Standard compensator configuration

The H∞ norm of a stable transfer function matrix G(s), denoted kG(s)k∞ , is the max-

imum over all frequencies of the largest singular value of the frequency response G(jω). Its

power stems from two important results:

(1) A sufficient condition for closed-loop stability to be robust against a set of plant

perturbation is given by a bound on the H∞ norm of a stable closed-loop transfer

function (model uncertainty problem).

(2) The H∞ norm of a stable transfer function matrix represents a bound on the maximum

energy gain from the input signal to the output (mixed sensitivity problem).

For the model uncertainty problem, consider the feedback system in Figure 5.32, where

the uncertainty in the nominal plant model G(s) is represented by an additive perturbation
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 132

D(s )

+ +
K(s) G(s) +
-

Figure 5.32: Additive perturbation

∆(s). Suppose, for simplicity, that ∆(s) is stable and that k∆(s)W (s)k∞ ≤ 1 where W (s)

is a weight which represents the variation of uncertainty with frequency and also normalizes

the H∞ norm of the uncertainty to a maximum of 1. Then the perturbed feedback system is

stable if the nominal feedback system (∆(s) = 0) is stable and kW −1 K(I + GK)−1 k∞ < 1.

Therefore, minimizing kW −1 K(I + GK)−1 k∞ over the set of all stabilizing controllers for

G(s) maximizes the margin of stability.

The mixed sensitivity problem is a special kind of H∞ control problem. In the mixed

sensitivity problem it is assumed that the system under consideration can be written as

the interconnection where K(s) is the controller which has to satisfy certain prerequisites.

Consider the feedback configuration shown in Figure 5.33, the problem is to regulate the

d(s)
+ u(s) + + y(s)
r(s) K(s) G(s)
-

Figure 5.33: Disturbance rejection at the plant output


CHAPTER 5. TASK-TAILORED CONTROL DESIGN 133

output y(s) of the system G(s) to look like some given reference signal r(s) by designing a

precompensator K(s) which has as its input the error signal, i.e. the input of the controller is

the difference between the output y(s) and the reference signal r(s). To prevent undesirable

surprises internal stability is required. Then the problem can be formulated as “minimizing”

the transfer function from r(s) to r(s) − y(s). As one might expect we shall minimize the

H∞ norm of this transfer function under the constraint of internal stability. The transfer

matrix from r(s) to u(s) should also be under consideration. In practice the process inputs

will often be restricted by physical constraints. This yields a bound on the transfer matrix

from r(s) to u(s). These transfer matrices from r(s) to r(s) − y(s) and from r(s) to u(s)

are given by:

S := (I + GK)−1 (5.7)

T := K(I + GK)−1 (5.8)

Here S is called the sensitivity function and T is called the control sensitivity function. A

small function S expresses good tracking properties while a small function T expresses small

inputs u(s). Note that there is a trade-off: making S smaller will in general make T larger.

Figure 5.34 shows the general form of the mixed sensitivity problem. An external input

w is added to the output y(s) as in Figure 5.34. Then the transfer matrix from ω(s) to

y(s) is equal to the sensitivity matrix S and the transfer matrix from ω(s) to u(s) is equal

to the control sensitivity matrix T . As noted in previous paragraph the H∞ norm can be

viewed as the maximum amount of energy coming out of the system, subject to inputs with

unit energy. However, if the Laplace transform is applied, then we obtain a frequency-
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 134

z 1(s) d(s)

W 1(s) W d(s)

w(s)

r(s) +
K(s)
u(s)
G(s) + + y(s)
W 2(s) z 2(s)
-

Figure 5.34: General block diagram of the mixed sensitivity problem

domain characterization. For a SISO system the H∞ norm is equal to the largest distance

of a point on the Nyquist contour to the origin. Hence the H∞ norm is a uniform bound

over all frequencies on the transfer function. It is assumed that the tracking signal will

have a limited frequency spectrum. It is in general impossible to track signals of very high

frequency reasonably well. On the other hand, since in general the model is only accurate

up to a certain frequency, the system is only required to track signals of frequencies up to a

certain bandwidth. In this situation straightforward application of H∞ control might yield

bad results because it only investigates a uniform bound over all frequencies. Also, certain

frequencies may be more important than others for the error signal and the control input.

Thus in Figure 5.34, the systems W1 (s), W2 (s), Wd (s) are weighting functions which are

chosen in such a way that we put more effort in regulating frequencies of interest than one

uniform bound. For practical purposes the choice of these weights is extremely important.

For SISO systems expressing performance criteria into requirements on the desired shape

of the magnitude Bode diagram is well established. This immediately translates into the

appropriate choice for the weighting functions. On the other hand, for MIMO systems
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 135

it is in general very hard to translate practical performance criteria into an appropriate

choice for the weighting functions. It should be noted that in practical circumstances it is

often better to minimize the integrated tracking error. This can also be incorporated in the

weighting functions and is simply one way to emphasize the interest in tracking signals of

low frequency.

In this way, the interconnection system from Figure 5.34 can be obtained. Note that

the transfer function matrix from the disturbance d(s) to z1 and z2 is:

 
 W1 T W d 
G̃(s) =  (5.9)
 

 
W2 SWd

Note that these weighting functions can be used to stress the relative importance of min-

imizing the sensitivity matrix S with respect to the importance of minimizing the control

sensitivity matrix T by multiplying W1 (s) by a scalar. The solution to this problem is

now well understood and can be computed automatically within a computer-aided design

package such as Matlab/Robust control toolbox [80].

5.4.2 Design of H∞ Controller for a Helicopter SAS

This section presents an application of H∞ optimization to the design of the SAS for a

UH-60A utility helicopter and illustrates how practical problems may be formulated in the

H∞ framework. Particular attention is paid to the presence of the external disturbance and

the motivation behind the selection of the weighting functions. Attention will be restricted

to designing a fixed-gain linear controller which is able to both stabilize the closed-loop

system and maintain some nominal performance in the hovering flight.


CHAPTER 5. TASK-TAILORED CONTROL DESIGN 136

The H∞ optimization framework used here is shown in Figure 5.35. The main steps

in the design process are: (i) augmentation of the plant P (s) at input and output with

weighting functions We , Wu and gust filter Wg ; (ii) synthesis of stabilizing controller K∞

minimizing the H∞ norm of the transfer function from (dg , w) to (u, y, eu , ee ).

The aircraft model used here is a 8-state/6 degree-of-freedom linear model extracted

from the non-linear simulation model. The plant output y is the aircraft angular rate

responses (p, q, r) and H∞ controller produces corresponding SAS outputs (u) for lateral

(rsas) and longitudinal (psas) cyclic and pedal collective (ysas). The inputs for the gust

shaping filter (Wg (s)) are zero-mean white noise with a variance of 1. The same gust

shaping filters from previous section 5.2 are used, in here only 30 degree WOD condition

is considered. In this case, if the pseudo-inverse of the plant distribution matrix (Bp ) is

implemented, then the disturbance could be considered as entering at the plant input. The

disturbance (w) in the model represents as unbounded perturbations to the plant output y.

A state-space realization for the augmented system G(s) can be obtained by directly

realizing the transfer matrix G(s) using any standard multivariable realization techniques

Gust filter
dg
(W g )
w
d
+ + + Aircraft + + y Weighting
r ee
- (P) (W e )

Weighting u H¥ controller
eu
(W u ) (K¥ )

Figure 5.35: Augmented system for airwake disturbance rejection


CHAPTER 5. TASK-TAILORED CONTROL DESIGN 137

(e.g. Gilbert realization). However, the direct realization approach is usually complicated.

Here another way is shown to obtain the realization for G(s) based on the realizations of each

component. To simplify the expressions, it is assumed that r is zero, and P, We , Wu , Wg

have, respectively, the following state-space realization.

       
 Ap Bp   Ae Be   Au Bu   Ag Bg 
P =  , We =   , Wu =   , Wg =   (5.10)
       
       
Cp Dp Ce De Cu Du Cg Dg

That is,

ẋp = Ap xp + Bp (−u + d), yp = Cp xp + Dp (−u + d)

ẋe = Ae xe + Be (yp + w), ee = Ce xe + De (yp + w)

ẋu = Au xu + Bu u, eu = Cu xu + Du u (5.11)

ẋg = Ag xg + Bg dg , d = Cg xg + Dg dg

y = yp + w

Now define a new state vector (x̄), an external disturbance (w̃) and a new output (z) as

" #T
x̄ = xp xe xu xg , w̃ = [dg w]T , z = [ee eu ]T (5.12)

and eliminate the variable yp to get a realization of G as

x̄˙ = Ax̄ + B1 w̃ + B2 u

z = C1 x̄ + D11 w̃ + D12 u (5.13)

y = C2 x̄ + D21 w̃ + D22 u
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 138

Note that the augmented problem must be well posed. In particular, the matrices D12 and

D21 must have full rank. The physical interpretation of these constraints is that, if D12

loses rank, this would lead to unconstrained controller action, and, if D21 loses rank, this

would correspond to a output being uncontrollable.

In general, the selection of weighting functions for a specific design problem often in-

volves ad hoc fixing, many iterations, and fine tuning. It is very hard to give a general

formula for the weighting functions that will work in every case. In choosing the weighting

functions, there are some guidelines by looking at a typical design specification [79]. For

the initial design, a high-gain low-pass filter is used for the output weighting function We ,

since the sensitivity function S must be keep small over a range of frequencies, typically

low frequencies where the disturbances are significant. Correspondingly the control weight-

ing function Wu must be a low-gain high-pass filter to emphasize the control sensitivity

function T at high frequencies so that the robustness is improved, and actuator activity

is reduced. This process gives a general shape for the weighting functions. The desired

frequency responses should be in inverse proportion to the magnitudes of the weights. In

this study, the weighting functions are tuned iteratively based on the disturbance rejection

property described in previous section 5.2. The final design weighting functions We and Wu

are given by

" #
We = diag 1.5 0.025s+10 0.025s+10 0.05s+5
s+10 , 1.5 s+10 , 5 s+5

(5.14)
" #
Wu = diag 0.4 s+0.001 s+0.5 s+0.6
s+15 , 0.05 0.9s+5 , 0.004 0.9s+6
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 139

20

10 We(r)

0 We(p), We(q)

−10

W (rsas)

Magnitude, [dB]
−20 u

−30 Wu(psas)

−40

−50

−60

−70 W (ysas)
u

−80
−2 −1 0 1 2
10 10 10 10 10

Frequency, [rad/sec]

Figure 5.36: Magnitude of weighting functions We and Wu

The magnitude of We and Wu are plotted in Figure 5.36.

The design process leads to a controller with as many as the interconnection structure

of Equation (5,13). In this study, the resulting controller has 14 states. The frequency

response of the controller is shown in Figure 5.37. It can be seen that the controller has

high gain at low frequency, for good tracking, and low gain at high frequency, for robustness.

This is consistent with the specification of the performance weighting function We (s).

Similar to previous section 5.2, the final H∞ controller is then converted to discrete form

and tested using the non-linear simulation model for the hovering operation on the spot

8 (0 and 30 degree WOD conditions, time-varying airwake model). Figures 5.38 and 5.43

illustrate the aircraft relative position with respect to the spot 8. The solid lines represent

the results with H∞ SAS, the dotted lines and dashed lines are simulation results with the
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 140

80

60

40

Magnitude, [dB]
20

−20

−40

−4 −3 −2 −1 0 1 2 3 4
10 10 10 10 10 10 10 10 10
Frequency, [rad/sec]

Figure 5.37: Singular values of controller K∞

baseline SAS and the optimized SAS configurations respectively from previous section for

comparison purpose. Figures 5.39 and 5.44 show the responses of aircraft angular rate for 0

degree and 30 degree WOD conditions. Figures 5.40 and 5.45 show the helicopter attitude

responses. The corresponding pilot stick inputs are shown in Figures 5.41 and 5.46. The

results with the H∞ SAS show some significant improvement over both the original SAS and

the optimized SAS. The aircraft pitch rate and longitudinal stick input are only slightly

improved compared to the optimized SAS case. Nonetheless the overall aircraft angular

rates and pilot control inputs are reduced with the H∞ SAS for both 0 and 30 degree WOD

conditions. Note that there is no attempt to optimize the control system for heave axis in

this design process. Figures 5.42 and 5.47 show the response of SAS outputs. The H∞ SAS

produces higher SAS actuator activity than both original SAS and optimized SAS. In both
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 141

cases, however, the SAS actuators stay well within the rate and position saturation limits.

Figures 5.48 and 5.49 show comparisons of input autospectra from pilot stick inputs.

Same 4 different window sizes (3, 5, 10, and 15 seconds) from previous section 5.2 are

used in the FFT analysis to obtain a composite average that is accurate over a range of

frequencies. From the figures, it can be observed that the optimized SAS shows some

improvement over the original SAS configuration in the frequency range of 1 to 10 rad/sec.

The H∞ controller practically halved the airwake disturbance effect on roll, pitch, and yaw

rate responses against the original SAS configuration.


CHAPTER 5. TASK-TAILORED CONTROL DESIGN 142

Original SAS
Optimized SAS
H¥ SAS

1.2

0.8

0.6
∆X

0.4

0.2

−0.2
−0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6

81

80.5
Altitude

80

79.5

79
−0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6
∆Y

Figure 5.38: Aircraft position w.r.t. the spot 8 [ft] - 30 knot, 0 degree WOD
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 143

Original SAS
Optimized SAS
H¥ SAS

0.5

0
P

−0.5

−1

0 5 10 15 20 25 30 35 40

0.5
Q

−0.5
0 5 10 15 20 25 30 35 40

0.4

0.2

0
R

−0.2

−0.4
0 5 10 15 20 25 30 35 40
Time, [sec]

Figure 5.39: Aircraft angular rate responses [deg/sec] - 30 knot, 0 degree WOD
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 144

Original SAS
Optimized SAS
H¥ SAS

−2.2
−2.4
PHI

−2.6
−2.8
−3
0 5 10 15 20 25 30 35 40

2.4

2.2
THETA

1.8

0 5 10 15 20 25 30 35 40

0.1
PSI

−0.1

0 5 10 15 20 25 30 35 40
Time, [sec]

Figure 5.40: Aircraft attitude responses [degree] - 30 knot, 0 degree WOD


CHAPTER 5. TASK-TAILORED CONTROL DESIGN 145

Original SAS
Optimized SAS
H¥ SAS

0.5
Lateral

0
−0.5

0 5 10 15 20 25 30 35 40
0.5
Longitudinal

0
−0.5
−1

0 5 10 15 20 25 30 35 40
Collective

−1
0 5 10 15 20 25 30 35 40

0.5
Pedal

−0.5

0 5 10 15 20 25 30 35 40
Time, [sec]

Figure 5.41: Pilot control stick inputs [%] - 30 knot, 0 degree WOD
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 146

Original SAS
Optimized SAS
H¥ SAS

1.5

1
RSAS

0.5

−0.5
0 5 10 15 20 25 30 35 40

0
PSAS

−1

−2
0 5 10 15 20 25 30 35 40

0.5
YSAS

−0.5

−1
0 5 10 15 20 25 30 35 40
Time, [sec]

Figure 5.42: SAS outputs [%] - 30 knot, 0 degree WOD


CHAPTER 5. TASK-TAILORED CONTROL DESIGN 147

Original SAS
Optimized SAS
H¥ SAS

2
∆X

−1
−4 −3 −2 −1 0 1 2 3 4

83

82

81
Altitude

80

79

78
−4 −3 −2 −1 0 1 2 3 4
∆Y

Figure 5.43: Aircraft position w.r.t. the spot 8 [ft] - 30 knot, 30 degree WOD
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 148

Original SAS
Optimized SAS
H¥ SAS

4
2
0
P

−2
−4
−6

0 5 10 15 20 25 30 35 40

2
Q

−2

−4
0 5 10 15 20 25 30 35 40

2
R

−2

−4
0 5 10 15 20 25 30 35 40
Time, [sec]

Figure 5.44: Aircraft angular rate responses [deg/sec] - 30 knot, 30 degree WOD
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 149

Original SAS
Optimized SAS
H¥ SAS

0
PHI

−2

−4

0 5 10 15 20 25 30 35 40

4
THETA

2
0 5 10 15 20 25 30 35 40

1
PSI

−1

0 5 10 15 20 25 30 35 40
Time, [sec]

Figure 5.45: Aircraft attitude responses [degree] - 30 knot, 30 degree WOD


CHAPTER 5. TASK-TAILORED CONTROL DESIGN 150

Original SAS
Optimized SAS
H¥ SAS

5
Lateral

−5
0 5 10 15 20 25 30 35 40
4
Longitudinal

2
0
−2
−4
−6
0 5 10 15 20 25 30 35 40

5
Collective

0
−5
−10
0 5 10 15 20 25 30 35 40
4
2
0
Pedal

−2
−4
−6
−8
0 5 10 15 20 25 30 35 40
Time, [sec]

Figure 5.46: Pilot control stick inputs [%] - 30 knot, 30 degree WOD
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 151

Original SAS
Optimized SAS
H¥ SAS

10

5
RSAS

−5
0 5 10 15 20 25 30 35 40

0
PSAS

−5

−10
0 5 10 15 20 25 30 35 40

10
YSAS

−10
0 5 10 15 20 25 30 35 40
Time, [sec]

Figure 5.47: SAS outputs [%] - 30 knot, 30 degree WOD


CHAPTER 5. TASK-TAILORED CONTROL DESIGN 152

Original SAS
Optimized SAS
H¥ SAS

Figure 5.48: Control stick input autospectra [dB] - 30 knot, 0 degree WOD
CHAPTER 5. TASK-TAILORED CONTROL DESIGN 153

Original SAS
Optimized SAS
H¥ SAS

Figure 5.49: Control stick input autospectra [dB] - 30 knot, 30 degree WOD
Chapter 6

Conclusions and Future Works

6.1 Conclusions

A helicopter/ship dynamic interface simulation tool has been developed to model a UH-60A

operating off an LHA class ship. To achieve a high fidelity simulation model, high-order

dynamic inflow model and time-accurate ship airwake solutions of an LHA class ship are

integrated with the flight dynamics simulation model. An optimal control model of the

human pilot has been developed to perform the desired shipboard operations. The optimal

control model of the human pilot proved to be an efficient method for simulating approach

trajectories.

Typical shipboard operations have been simulated from landing spot 8 on the LHA

(CFD results showed significant time-varying flow effects over this spot). Simulation results

are compared to the flight test data for the 0 degree and 30 degrees WOD conditions. Pilot

control activity is then compared to JSHIP flight test data in terms of time domain and

frequency domain analysis.

154
CHAPTER 6. CONCLUSIONS AND FUTURE WORKS 155

In addition, the present work investigates the optimization of a flight control system

for the UH-60A Black Hawk helicopter operating in the turbulent airwake of a LHA class

ship. A stochastic model of the ship airwake is derived from simulations with a full time-

accurate CFD solution of the airwake. In this study, the gains of a SAS are optimized

using CONDUIT, and a H∞ controller is designed to provide an alternative for a helicopter

automatic flight control system. New SAS configurations are then tested using a full non-

linear simulation model.

Overall, the main objectives of this thesis, as stated in section 1.3, have been achieved.

The following general conclusions can be drawn from this work:

1. The simulation results from Section 4.2 clearly indicate that the time-varying airwake

has a significant impact on aircraft response and pilot control activity when the aircraft

is flown for specified approach and departure trajectories. The differences are most

notable when the helicopter is operating in or near a hover relative to the ship deck

(stationkeeping). In the past, gust models for fixed-wing aircraft simulation have

often used a stationary or frozen field model. This is adequate when the aircraft is

moving at a significant forward speed. However, the model clearly breaks down as

airspeed approaches zero. The same appears to be true when helicopters are operating

in turbulent ship airwake. The time-varying nature of the ship airwake becomes

dominant as the helicopter approaches hover on the ship deck. And, the simulation

results of 30 degree WOD condition showed a substantial increase in pilot workload.

2. An optimal control model of the human pilot is successfully implemented to solve the

“inverse simulation” problem. Given a specified trajectory, the pilot controls can be
CHAPTER 6. CONCLUSIONS AND FUTURE WORKS 156

calculated using forward simulation in conjunction with a feedback controller. This

is found to be highly useful for this research task. Inverse simulations can be time

consuming and difficult to implement computationally. The pilot model is easily tuned

and seemed to produce reasonable predictions of trajectory tracking and pilot control

activity.

3. By comparison with data from flight test data, it is found that the control activities are

similar in low frequency range but underestimate in magnitude in the high frequency

range (over 1.5 Hz). There is clear evidence that the human pilot is continually moving

cyclic stick in the maneuver. At this stage of the study no attempt has been made to

optimize the parameters of the human pilot model.

4. The stochastic disturbance model of ship airwake appears to result in similar overall

behavior of the coupled aircraft/pilot system as the simulation with the full time-

varying airwake. The 40 seconds of the time-varying airwake solutions are stored for

every 0.1 second, and each flow solutions for the target DI mesh is 5.2 Mbytes in

size. Therefore, the proposed stochastic airwake modeling approach avoids the large

computational and storage requirements of the CFD data.

5. The shaping filters used in the stochastic model are readily incorporated in the flight

control optimization using CONDUIT and H∞ control design. The filters provide a

tool for flight control design.

6. The optimized SAS results in higher gain in all three axes compared to the original

SAS, but it does not appear to result in excessive control activity in the SAS. Upper

limits on roll gain are due to stability margin limits from rotor-body coupling. Results
CHAPTER 6. CONCLUSIONS AND FUTURE WORKS 157

with the non-linear simulation model show significant reduction in pilot workload in

pitch and yaw, with slight improvement in roll compared to the original SAS. Although

further improvements could be obtained, the results appear to validate the current

design approach.

7. The potential of H∞ control method for a helicopter flight control system design has

been demonstrated. Successful engagement is achieved on the non-linear simulation.

The final H∞ controller considerably reduces the aircraft angular rate responses and

corresponding pilot control inputs compared to the original SAS. Thus pilot workload

would be significantly reduced, allowing more aggressive maneuvers to be carried out

with a higher degree of precision. Also passenger comfort and safety will be increased.

6.2 Recommendations for Future Work

The following are suggested for future study:

1. Currently, the helicopter/ship aerodynamic interactions are not included. To examine

the effect of complex helicopter/ship aerodynamic interactions, the deck ground effect

should be considered, and the appropriate condition at the ground surface needs to

be determined in order to model unsteady effects.

2. In this study, the stochastic airwake model is designed for one location (17 ft above

the spot 8). In order to increase the computational efficiency, an array of airwake

shaping filters would still need to be designed for different WOD conditions and for

different locations on the ship.


CHAPTER 6. CONCLUSIONS AND FUTURE WORKS 158

3. For further validation, it should be apply an equivalent airwake disturbance method

to validate ship airwake CFD analysis. Airwake disturbance can be extracted from

flight test and compared to simulation with CFD airwake solutions.

4. At this stage of the study no attempt has been made to optimize the parameters of

the human pilot model and this work should be completed in the future work.

5. A new spec for disturbance rejection is designed based on PSD of angular rate response

to corresponding gust input. Currently, the boundaries for Level 1/Level 2 and Level

2/Level 3 are arbitrarily selected, but the optimization with respect to this spec results

in a reduction in the aircraft response and pilot control activity required to stabilize

the aircraft in a turbulent airwake. It would be desirable to perform a more detailed

handling qualities study of the spec to establish appropriate level boundaries.

6. The ability to achieve an optimized solution using CONDUIT is sensitive to the initial

values for the design parameters. The problem may be over-parameterized. Future

work should seek to reduce the number of control parameters used in the synthesis

process in order to get more consistent behavior in the optimization.

7. To design a flight control system to improve handling qualities, a series of design

specifications are required to guarantee the performance of the controller. However,

there is no attempt to establish the performance measurement for flight control system

design with modern control theory. Thus, future work must include an investigation

into different measures of performance to design a robust flight control system.

8. Although performance and robustness are achieved using the current H∞ design
CHAPTER 6. CONCLUSIONS AND FUTURE WORKS 159

method, how adequately this captures the discrepancies between the low-order lin-

ear model used in the design and the true non-linear aircraft dynamics is debatable.

Questions remain concerning the fidelity of the plant model used in this study; large

amounts of flight test data must be gathered to help shed light on that.

9. Future shipbased rotorcraft will require high levels of agility and maneuverability

as well as capabilities for operations in degraded visual environments and adverse

weather conditions. Thus it is desirable to expand flight control design efforts to

establish the autonomous landing flight control system and position hold over ship

deck. This efforts may make these performance requirements achievable by reducing

pilot workload and expanding WOD envelopes.


Bibliography

[1] Langlois, R.G. and Tadros, A.R., “State-of-the-Art On-Deck Dynamic Interface Anal-

ysis,” The American Helicopter Society 55th Annual Forum, Montreal, Canada, May

1999

[2] Wilkinson, C.H., Roscoe, M.F., and VanderVliet, G.M, “Determining Fidelity Stan-

dards for the Shipboard Launch and Recovery Task,” AIAA Modeling and Simulation

Technologies Conference and Exhibit, Montreal, Canada, August 2001

[3] Padfield, G.D., Helicopter Flight Dynamics: The Theory and Application of Flying

Qualities and Simulation Modeling, AIAA Education Series, Virginia, 1996

[4] Johnson, W., Helicopter Theory, Dover, New York, 1994

[5] Bramwell, A.R.S., Helicopter Dynamics, John Wiley & Sons, New York, 1976

[6] Trentini, M. and Pieper, J.K., “Mixed Norm Control of a Helicopter,” Journal of

Guidance Control and Dynamics, Vol. 24, No. 3, pp. 555–565, May 2001

[7] Postlethwaite, I., Smerlas, A., Gubbels, A.W., Baillie, S.W., Strange, M.E., and

Howitt, J., “H∞ Control of the NRC Bell 205 Fly-by-Wire Helicopter,” Journal of

The American Helicopter Society, Vol. 24, No. 3, pp. 276–284, October 1999

160
BIBLIOGRAPHY 161

[8] Frost, C.R., Hindson, W.S., Moralez, E., Tucker, G.E., and Dryfoos, J.B., “Design and

Testing of Flight Control Laws on the RASCAL Research Helicopter,” AIAA Modeling

and Simulation Technologies Conference and Exhibit, Monterey, CA, August 2002

[9] Takahashi, M.D., “Rotor-State Feedback in the Design of Flight Control Laws for a

Hovering Helicopter,” Journal of The American Helicopter Society, Vol. 39, No. 1,

pp. 50–62, January 1994

[10] Ingle, S.J. and Celi, R., “Effects of Higher-Order Dynamics on Helicopter Flight Control

Law Design,” Journal of The American Helicopter Society, Vol. 39, No. 3, pp. 12–23,

July 1994

[11] Howlett, J.J., “UH-60A Black Hawk Engineering Simulation Program: Volume I -

Mathematical Model,” NASA CR-177542 USAAVSCOM TR 89-A-001, September

1989

[12] Johnson, W., “Application of Unsteady Airfoil Theory to Rotary Wings,” Journal of

Aircraft, Vol. 17, pp. 285–286, April 1980

[13] Loewy, R.G., “A Two-Dimensional Approximation to the Unsteady Aerodynamics of

Rotary wings,” Journal of the Aeronautical Sciences, Vol. 24, No. 2, pp. 81–92, Februry

1957

[14] Pitt, D.M. and Peters, D.A., “Theoritical Prediction of Dynamic-Inflow Derivatives,”

Vertica, Vol. 5, No. 1, pp. 21–34, March 1981

[15] Gaonkar, G.H. and Peters, D.A., “Reveiw of Dynamic Inflow Modeling for Rotorcraft

Flight Dynamics,” Vertica, Vol. 12, No. 3, pp. 213–242, 1988


BIBLIOGRAPHY 162

[16] Peters, D.A., Boyd, D.D., and He, C., “Finite-State Induced-Flow Model for Rotors

in Hover and Forward Flight,” Journal of the American Helicopter Society, Vol. 34,

No. 4, pp. 5–17, October 1989

[17] Peters, D.A. and He, C., “Correlation of Measured Induced Velocities with a Finite-

State Wake Model,” Journal of the American Helicopter Society, Vol. 34, No. 4, pp. 59–

70, October 1989

[18] He, C., Development and Application of a Generalized Dynamic Wake Theory for

Lifting Rotors, Ph.D. thesis, Georgia Institute of Technology, July 1989

[19] Krämer, P. and Gimonet, B., “Improvement of Nonlinear Simulation Using Parameter

Estimation Technique,” 26th European Rotorcraft Forum, Hague, The Netherlands,

September 2000

[20] Hamers, M. and Basset, P., “Application of the Finite State Unsteady Wake Model in

Helicopter Flight Dynamic Simulation,” 26th European Rotorcraft Forum, Hague, The

Netherlands, September 2000

[21] Zhang, H, Prasad, J.V.R., and Peters, D.A., “Finite State Inflow Models for Lift-

ing Rotors in Ground Effect,” The American Helicopter Society 52nd Annual Forum,

Washington D.C., June 1996

[22] Prasad, J.V.R., Zhang, H., and Peters, D.A., “Finite State In-Ground Effect Inflow

Models for Lifting Rotors,” The American Helicopter Society 53rd Annual Forum,

Virgina Beach, VA, April 1997


BIBLIOGRAPHY 163

[23] Zhang, H., Mello, O.A.F., Prasad, J.V.R., Sankar, L.N., and Funk, J.D., “A Simulation

Model of Ship Ground Effect for Rotorcraft/Ship Interaction Study,” The American

Helicopter Society 51st Annual Forum, Fort Worth, TX, May 1995

[24] Xin, H., Prasad, J.V.R., Peters, D.A., Nagashima, T., and Iboshi, N., “A Finite State

Inflow Model for Simulation of Helicopter Hovering in Ground Effect,” The American

Helicopter Society 54th Annual Forum, Washington D.C., 1998

[25] Xin, H., Development and Validation of a Generalized Ground Effect Model for Lifting

Rotors, Ph.D. thesis, Georgia Institute of Technology, July 1999

[26] Xin, H., He, C., and Lee, J, “Combined Finite State Rotor Wake and Panel Ship Deck

Models for Simulation of Helicopter Shipboard Operations,” The American Helicopter

Society 57th Annual Forum, Washington D.C, May 2001

[27] Healey, J.V., “The prospects for Simulating the Helicopter/Ship Interface,” Naval En-

gineers Journal, pp. 45–63, March 1987

[28] He, C., Kang, H., Carico, D., and Long, K., “Development of a Modeling and Simu-

lation Tool for Rotorcraft/Ship Dynamic Interface Testing,” The American Helicopter

Society 58th Annual Forum, Montreal, Canada, June 2002

[29] McKillip, R., Boschitsch, A., Quackenbush, T., Keller, J., and Wachspress, D., “Dy-

namic Interface Simulation Using a Coupled Vortex-Based Ship Airwake and Ro-

tor Wake Model,” The American Helicopter Society 58th Annual Forum, Montreal,

Canada, June 2002


BIBLIOGRAPHY 164

[30] Tai, T.C and Carico, D., “Simulation of DD-963 Ship Airwake by Navier-Stokes

Method,” Journal of Aircraft, Vol. 32, No. 6, pp. 1399–1401, November 1995

[31] Tai, T.C., “Simulation and Analysis of LHD Ship Airwake by Navier-Stokes Method,”

The 83rd Fluid Dynamics Panel Symposium on Fluid Dynamics Problems of Vehicles

Operating Near or In the Air Sea Interface, Amsterdam, The Netherlands, October

1998

[32] Modi, A., Long, L.N., and Hansen, R.P., “Unsteady Separated Flow Simulations us-

ing a Cluster of Workstations,” International Conference on Parallel and Distributed

Processing Techniques and Applications, Las Vegas, NV, June 1999

[33] Clement, W.F., Gorder, P.J. and Jewell, W.F., “Development of a Real-Time Simula-

tion of a Ship-Correlated Airwake Model Interfaced with a Rotorcraft Dynamic Model,”

Technical Report TR-91-C-0176, The Naval Air Syatems Command, 1991

[34] Labows, S.J., Blanken, C.L, and Tischler, M.B., “UH-60 Black Hawk Disturbance

Rejection Study for Hover/Low Speed Handling Qualities Criteria and Turbulence

Modeling,” The American Helicopter Society 56th Annual Forum, Virginia Beach, VA,

May 2000

[35] Elliott, A.S. and Chopra, I., “Helicopter Response to Atmospheric Turbulence in For-

ward Flight,” Journal of the American Helicopter Society, Vol. 35, No. 2, pp. 51–59,

May 1990
BIBLIOGRAPHY 165

[36] Polsky, S.A. and Bruner, C.W.S., “Time-Accurate Computational Simulations of an

LHA Ship Airwake,” 18th AIAA Applied Aerodynamics Conference, Denver, CO, Au-

gust 2000

[37] Polsky, S.A., “A Computational Study of Unsteady Ship Airwake,” 40th AIAA

Aerospace Sciences Meeting and Exhibit, Reno, NV, January 2002

[38] Lee, D, Sezer-Uzol, N., Horn, J.F., and Long, L.N., “Simulation of Helicopter Shipboard

Launch and Recovery with Time-Accurate Airwakes,” The American Helicopter Society

59th Annual Forum, Phoenix, AZ, May 2003

[39] Lee, D, Horn, J.F., Sezer-Uzol, N., and Long, L.N., “Simulation of Pilot Control Ac-

tivity During Helicopter Shipboard Operation,” AIAA Atmospheric Flight Mechanics

Conference and Exhibit, Austin, TX, August 2003

[40] McFarland, R. E. and Duisenburg, K., “Simulation of Rotor Blade Element Turbu-

lence,” NASA TM 108862, January 1995

[41] Prouty, R.W. and Curtiss, H.C., “Helicopter Control Systems: A History,” Journal of

Guidance Control and Dynamics, Vol. 26, No. 1, pp. 12–18, January 2003

[42] Mannes, M.A. and Murray-Smith, D.J., “Aspects of Multivariabl Flight Control Law

Design for Helicopters Using Eigenstructure Assignment,” Journal of the American

Helicopter Society, Vol. 18, No. 1, pp. 18–32, July 1992

[43] Hess, R.A., “Rotorcraft Control System Design for Uncertain Vehicle Dynamics Using

Quantitative Feedback Theory,” Journal of the American Helicopter Society, Vol. 39,

No. 2, pp. 47–55, April 1994


BIBLIOGRAPHY 166

[44] Bogdanov, A.A., Wan, E.A., Carlsson, M., Zhang, Y, Kieburtz, R., and Baptista,

A., “Model Predictive Neural Control of a High-Fidelity Helicopter Model,” AIAA

Guidance Navigation and Control Conference and Exhibit, Montreal, Canada, August

2001

[45] Hovakimyan, N., Kim, N., Calise, A.J., Prasad, J.V.R., and Corban, E., “Adaptive

Output Feedback for High-Bandwidth Control of an Unmanned Helicopter,” AIAA

Guidance Navigation and Control Conference and Exhibit, Montreal, Canada, August

2001

[46] Corban, J.E., Calise, A.J., Prasad, J.V.R., Hur, J., and Kim, N, “Flight Evaluation of

Adaptive High-Bandwidth Control Methods for Unmanned Helicopters,” AIAA Guid-

ance Navigation and Control conference and Exhibit, Monterey, CA, August 2002

[47] Johnson, E.N. and Kannan, S.K., “Adaptive Flight Control for an Autonomous Un-

manned Helicopter,” AIAA Guidance Navigation and Control Conference and Exhibit,

Monterey, CA, August 2002

[48] Krupadanam, A.S., Annaswamy, A.M., Mangoubi, R.S., “Multivariabl Adaptive Con-

trol Design with Applications to Autonomous Helicopters,” Journal of Guidance Con-

trol and Dynamics, Vol. 25, No. 5, pp. 843–851, September 2002

[49] Xin, H. and He, C., “A Combined Technique for Inverse Simulation Applied to Rotor-

craft Shipboard Operations,” The American Helicopter Society 58th Annual Forum,

Montreal, Canada, June 2002


BIBLIOGRAPHY 167

[50] Hess, R.A. and Gao, C., “A Generalized Algorithm for Inverse Simulation Applied to

Helicopter Maneuvering Flight,” Journal of the American Helicopter Society, Vol. 38,

No. 4, pp. 3–15, October 1993

[51] Avanzini, G. and Matteis, G., “Two-Timescale Inverse Simulation of a Helicopter

Model,” Journal of Guidance Control and Dynamics, Vol. 24, No. 2, pp. 330–339,

March 2001

[52] McRuer, D.T., Graham, D., Krendel, E.S., and Reisener, Jr., W., “Human-Pilot Dy-

namics in Compensatory Systems,” AFFDL-TR-65-15, July 1965

[53] Bradley, R. and Turner, G., “Simulation of the Human Pilot applied at the Heli-

copter/Ship Dynamic Interface,” The American Helicopter Society 55th Annual Forum,

Montreal, Canada, May 1999

[54] Turner, G., Brindley, G., and Bradley, R., “Simulation of Pilot Control Activity for

the Prediction of Workload Ratings in Helicopter/Ship Operations,” 26th European

Rotorcraft Forum, Hague, The Netherlands, September 2000

[55] Heffley, R.K., Hess, R.A., and Zeyada, Y., “Application of Classical-Control Techniques

for Computer Modeling and Simulation of Helicopter Pilotage Tasks,” The American

Helicopter Society 57th Annual Forum, Washington D.C., May 2001

[56] Lee, D and Horn, J.F., “Simulation and Control of Helicopter Shipboard Launch and

Recovery Operations,” The American Helicopter Society Flight Controls and Crew Sys-

tem Design Specialists Meeting, Philadelphia, PA, October 2002


BIBLIOGRAPHY 168

[57] Mello, O.A.F, Pradad, J.V.R. Sankar, L.N., and Tseng, T., “Analysis of Heli-

copter/Ship Aerodynamic Interactions,” The American Helicopter Society Aerome-

chanics Specialists Conference, San Francisco, CA, January 1994

[58] Perrins, J.A. and Howitt, J., “Development of A Pilot Assisted Landing System for

Helicopter/Ship Recoveries,” The American Helicopter Society 57th Annual Forum,

Washington D.C., May 2001

[59] Colwell, J.L., “Effects of Flight Deck Motion in High Seas on the Hovering Maritime

Helicopter,” The American Helicopter Society 58th Annual Forum, Montreal, Canada,

June 2002

[60] Roscoe, M.F. and Wilkinson, C.H., “DIMSS-JSHIP’s Modeling and Simulation Process

for Ship/Helicopter Testing and Training,” AIAA Modeling and Simulation Technolo-

gies Conference and Exhibit, Monterey, CA, August 2002

[61] Advani, S.K. and Wilkinson, C.H., “Dynamic Interface Modelling and Simulation - A

Unique Challenge,” The Royal Aeronautical Society Conference on Helicopter Flight

Simulation, London, UK, November 2001

[62] Kleinman, D.L., Baron, S., and Levison, W.H., “An Optimal Control Model of Human

Response Part I: Theory and Validation,” Automatica, Vol. 6, No. 3, pp. 357–369, 1970

[63] Hess, R.A., Feedback Control Models - Manual Control and Tracking, Handbook of

Human Factors and Ergonomics, 2nd edition, Wiley, New York, 1997

[64] Lewis, F.L., Applied Optimal Control and Estimation, Prentice-Hall, New Jersey, 1992
BIBLIOGRAPHY 169

[65] Bunnell, J.W., “An Integrated Time-Varying Airwake in a UH-60 Black Hawk Ship-

board Landing Simulation,” AIAA Modeling and Simulation Technologies Conference

and Exhibit, Montreal, Canada, August 2001

[66] Tischler, M.B. and Cauffman, M.G., “Comprehensive identification from frequency

responses: Volume 1 - Class Notes,” NASA CP 10149, USAATCOM TR-94-A-017,

1994

[67] Anon., “Handling Qualities Requirements for Military Rotorcraft,” Aeronautical De-

sign Standard Performance Specification (ADS-33E-PRF), US Army Aviation and Mis-

sile Command, March 2000

[68] Lusardi, J.A., Tischler, M.B., Blanken, C.L., and Labows, S.J., “Empirically Derived

Helicopter Response Model and Control System Requirements for Flight in Turbu-

lence,” Journal of the American Helicopter Society, Vol. 49, No. 3, pp. 340-349, July

2004

[69] Baillie, S.W. and Morgan, J.M., “An In-flight Investigation into the Relationships

Among Control Sensitivity, Control Bandwidth and Disturbance Rejection Bandwidth

Using a Variable Stability Helicopter,” 15th European Rotorcraft Forum, Amsterdam,

The Netherlands, September 1989

[70] Tischler, M.B., Lee, J.A., and Colbourne, J.D., “Optimization and Comparison of

Alternative Flight Control System Design Methods Using a Common Set of Handling-

Qualities Criteria,” AIAA Guidance, Navigation, and Control Conference, Montreal,

Canada, August 2001


BIBLIOGRAPHY 170

[71] Lee, D. and Horn, J.F., “Analysis of Pilot Workload in the Helicopter/Ship Dynamic

Interface Using Time-Accurate and Stochastic Ship Airwake Models,” AIAA Atmo-

spheric Flight Mechanics Conference and Exhibit, Providence, RI, June 2004

[72] Ly, U. and Chan, Y., “Time-Domain Computation of Aircraft Gust Covariance Ma-

trices,” AIAA Atmospheric Flight Mechanics Conference and Exhibit, Danvers, MA,

August 1980

[73] Anon., CONDUIT - The Control Designer’s Unified Interface, Army/NASA Rotorcraft

Division, Raytheon ITSS, User’s Guide, 2003

[74] Tischler, M.B., Colbourne, J.D., Morel, M.R., Biezad, D.J., Levine, W.S., and

Moldoveanu, V., “CONDUIT-A New Multidisciplinary Integration Environment for

Flight Control Development,” AIAA Guidance, Navigation, and Control Conference,

New Orleans, LA, August 1997

[75] Yue, A. and Postlethwaite, I., “Improvement of Helicopter Handling Qualities Using

H-infinity Optimization,” IEE Proceedings Part D 137, May 1990

[76] Walker, D.J., Postlethwaite, I., Howitt, J., and Foster, N., “Rotorcraft Flying Quali-

ties Improvement Using Advanced Control,” Proceedings of American Helicopter Soci-

ety/NASA comference on Piloting Vertical Flight Aircraft: flying qualities and human

factors, 1993

[77] Walker, D.J. and Postlethwaite I., “Advanced Helicopter Flight Control Using Two-

degree-of-freedom H-infinity Optimization,” Journal of Guidance, Control, and Dy-

namics, Vol. 19, No. 2, pp. 461–468, 1996


BIBLIOGRAPHY 171

[78] Zames, G., “Feedback and Optimal Sensitivity: Model Reference Transformations,

Multiplicative Seminorms, and Approximate Inverse,” IEEE Transacitons Automatic

Control, Vol. AC-26, pp. 301–320, 1981

[79] Zhou, K. and Doyle, J.C., Essentials of Robust Control, Prentice-Hall, New Jersey,

1998

[80] Balas, G., Chiang, R., Packard, A., and Safonov, M., Robust Control Toolbox: User’s

Guide, Version 3, The Math Works, Massachusetts, 2004


Vita

Mr. Dooyong Lee was born on December 25, 1971, in Busan, Republic of Korea. He

received his bachelor’s degree in Aerospace Engineering in February 1997 from INHA Uni-

versity, Inchon, Korea. He earned his master’s degree in Aerospace Engineering in February

1999 from the INHA University before entering the Graduate Program in the Aerospace

Engineering at the Pennsylvania State University where he worked as a research assistant.

He was awarded the Vertical Flight Foundation Scholarship in 2002. Mr. Lee is a student

member of the American Helicopter Society and American Institute of Aeronautics and

Astronautics.

You might also like