Download as pdf or txt
Download as pdf or txt
You are on page 1of 320

The Pennsylvania State University

The Graduate School


College of Engineering
DESIGN AND ANALYSIS OF ROTOR SYSTEMS WITH
MULTIPLE TRAILING EDGE FLAPS AND RESONANT
ACTUATORS
A Thesis in
Aerospace Engineering
by
Jun-Sik Kim
c _ 2005 Jun-Sik Kim
Submitted in Partial Fulllment
of the Requirements
for the Degree of
Doctor of Philosophy
December 2005
The thesis of Jun-Sik Kim was reviewed and approved

by the following:
Edward C. Smith
Professor of Aerospace Engineering
Thesis Co-Advisor
Co-Chair of Committee
Kon-Well Wang
William E. Diefenderfer Chaired Professor in Mechanical Engineering
Thesis Co-Advisor
Co-Chair of Committee
Farhan S. Gandhi
Associate Professor of Aerospace Engineering
Joseph F. Horn
Assistant Professor of Aerospace Engineering
Mary I. Frecker
Associate Professor of Mechanical Engineering
George A. Lesieutre
Professor of Aerospace Engineering
Head of the Department of Aerospace Engineering

Signatures are on le in the Graduate School.


Abstract
The purpose of this thesis is to develop piezoelectric resonant actuation systems
and new active control methods utilizing the multiple trailing-edge aps congu-
ration for rotorcraft vibration suppression and blade loads control.
An aeroelastic model is developed for a composite rotor blade with multiple
trailing-edge aps. The rotor blade airloads are calculated using quasi-steady blade
element aerodynamics with a free wake model for rotor inow. A compressible
unsteady aerodynamics model is employed to accurately predict the incremental
trailing edge ap airloads. Both the nite wing eect and actuator saturation for
trailing-edge aps are also included in an aeroelastic analysis.
For a composite articulated rotor, a new active blade loads control method is
developed and tested numerically. The concept involves straightening the blade
by introducing dual trailing edge aps. The objective function, which includes
vibratory hub loads, bending moment harmonics and active ap control inputs,
is minimized by an integrated optimal control/optimization process. A numerical
simulation is performed for the steady-state forward ight of an advance ratio of
0.35. It is demonstrated that through straightening the rotor blade, which mimics
the behavior of a rigid blade, both the bending moments and vibratory hub loads
can be signicantly reduced by 32% and 57%, respectively.
An active vibration control method is developed and analyzed for a hingeless
rotor. The concept involves deecting each individual trailing-edge ap using a
compact resonant actuation system. Each resonant actuation system could yield
high authority, while operating at a single frequency. Parametric studies are con-
ducted to explore the nite wing eect of trailing-edge aps and actuator satura-
tion. A numerical simulation has been performed for the steady-state forward ight
( = 0.15 0.35). It is demonstrated that multiple trailing-edge ap congura-
tion with the resonant actuation system can reduce the required trailing-edge ap
hinge moments by 37% to 61% in each individual actuator compared to single-ap
iii
conguration for high speed ight conditions.
A novel resonant actuation concept is developed to eciently realize the heli-
copter vibration and blade loads control. The resonant actuation system (RAS) is
achieved through both mechanical and electrical tailoring. With mechanical tun-
ing, the resonant frequencies of the actuation system (includes the piezoelectric
actuator and the related mechanical and electrical elements for actuation) can be
adjusted to the required actuation frequencies. This obviously will increase the
authority of the actuation system. To further enhance controllability and robust-
ness, the actuation resonant peak can be signicantly broadened and attened with
electrical tailoring through the aid of an electric network of inductance, resistance,
and negative capacitance.
A piezoelectric resonant actuation system model is derived for active ap ro-
tors. The optimal values of the electrical components are explicitly determined.
An equivalent electric circuit model emulating the physical actuation system is
derived and experimentally tested to investigate the initial feasibility of the piezo-
electric resonant actuation system. It is demonstrated that the proposed resonant
actuation system can indeed achieve both high active authority and robustness.
It is shown that the actuator authority is signicantly increased from 1.25 to 4.5
degrees as compared to the static value, with wide operating bandwidth of 8 Hz.
In addition to this, the RAS is compared to an equivalent mechanical system to
provide better physical understanding. Design guidelines of the RAS are derived
in dimensionless forms. Feed-forward controllers are developed to realize the elec-
tric network dynamics and to adapt the phase variation. The control strategy is
then implemented via a digital signal processor (DSP) system. Performance of the
resonant actuation system is analyzed and veried experimentally on a full-scale
piezoelectric tube actuator for helicopter rotor control. Promising results are illus-
trated that the actuator stroke is increased 2 to 3.5 times compared to its static
value with bandwidth of 5 to 10 Hz.
iv
Table of Contents
List of Figures ix
List of Tables xv
List of Symbols xvi
Acknowledgments xxvi
Chapter 1
Introduction 1
1.1 Background and Motivation . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Literature Review . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.2.1 Helicopter Vibration Reduction . . . . . . . . . . . . . . . . 7
1.2.2 Active Trailing-edge Flaps . . . . . . . . . . . . . . . . . . . 14
1.2.3 Smart Actuation System Development . . . . . . . . . . . . 20
1.2.4 Piezoelectric Networks . . . . . . . . . . . . . . . . . . . . . 26
1.2.5 Summary of Literature Review . . . . . . . . . . . . . . . . 30
1.3 Problem Statement and Objectives . . . . . . . . . . . . . . . . . . 32
1.4 Overview of Dissertation . . . . . . . . . . . . . . . . . . . . . . . . 36
Chapter 2
Helicopter Model 38
2.1 Preliminary Background . . . . . . . . . . . . . . . . . . . . . . . . 39
2.1.1 Vehicle Kinematics and Coordinate Systems . . . . . . . . . 39
2.1.2 Blade Deformed Kinematics and Coordinate Systems . . . . 41
2.1.3 Nondimensionalization and Ordering Scheme . . . . . . . . . 45
2.1.4 Variational Formulation . . . . . . . . . . . . . . . . . . . . 47
2.2 Structural Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
v
2.2.1 Strain Energy of Rotor Blade . . . . . . . . . . . . . . . . . 49
2.2.2 Kinetic Energy of Rotor Blade . . . . . . . . . . . . . . . . . 57
2.3 Aerodynamic Blade Loads . . . . . . . . . . . . . . . . . . . . . . . 64
2.3.1 Quasi-steady Airloads . . . . . . . . . . . . . . . . . . . . . 68
2.3.2 Noncirculatory Airloads . . . . . . . . . . . . . . . . . . . . 72
2.3.3 Quasi-steady Aerodynamics Implementation . . . . . . . . . 73
2.4 Inow and Free Wake Model . . . . . . . . . . . . . . . . . . . . . . 76
2.4.1 Linear Inow . . . . . . . . . . . . . . . . . . . . . . . . . . 76
2.4.2 Free Wake Model . . . . . . . . . . . . . . . . . . . . . . . . 77
2.5 Aeroelastic Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . 80
2.5.1 Aeroelastic Response . . . . . . . . . . . . . . . . . . . . . . 80
2.5.2 Coupled Propulsive Trim . . . . . . . . . . . . . . . . . . . . 85
Chapter 3
Trailing Edge Flap Formulation 91
3.1 Inertial Contribution . . . . . . . . . . . . . . . . . . . . . . . . . . 92
3.2 Aerodynamic Models . . . . . . . . . . . . . . . . . . . . . . . . . . 97
3.2.1 Incompressible Model . . . . . . . . . . . . . . . . . . . . . . 97
3.2.2 Compressible Model . . . . . . . . . . . . . . . . . . . . . . 98
3.3 Active Trailing Edge Flap Control Algorithm . . . . . . . . . . . . . 102
3.3.1 Feedback Form of Global Controller . . . . . . . . . . . . . . 102
3.3.2 Active-Passive Hybrid Design . . . . . . . . . . . . . . . . . 106
Chapter 4
Active Loads Control Using a Dual Flap Conguration 108
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
4.2 Description of Analytical Models . . . . . . . . . . . . . . . . . . . 110
4.3 Numerical Results and Discussions . . . . . . . . . . . . . . . . . . 112
4.3.1 Baseline Articulated Rotor Analysis . . . . . . . . . . . . . . 113
4.3.2 Rigid Blade vs. Elastic Blade . . . . . . . . . . . . . . . . . 118
4.3.3 A Single Flap for Moment Reduction . . . . . . . . . . . . . 118
4.3.4 Dual Flap Performance . . . . . . . . . . . . . . . . . . . . . 119
4.3.5 Multicyclic Control for Moment and Vibration Reduction . . 124
4.3.6 Active-Passive Hybrid Design . . . . . . . . . . . . . . . . . 126
4.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
Chapter 5
Helicopter Vibration Suppression via Multiple Trailing-Edge
Flaps with Resonant Actuation Concept 132
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
vi
5.2 Description of Analytical Models . . . . . . . . . . . . . . . . . . . 135
5.3 Results and Discussions . . . . . . . . . . . . . . . . . . . . . . . . 137
5.3.1 Baseline Hingeless Rotor Analysis . . . . . . . . . . . . . . . 137
5.3.2 Flap Eect to Free-Wake Geometry . . . . . . . . . . . . . . 143
5.3.3 Determination of Trailing-Edge Flap Locations . . . . . . . . 147
5.3.4 Finite Wing Eects . . . . . . . . . . . . . . . . . . . . . . . 149
5.3.5 Eectiveness of Multiple-Flap Conguration . . . . . . . . . 151
5.3.6 Vibration Reduction with Multicyclic Control . . . . . . . . 156
5.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
Chapter 6
Piezoelectric Actuation System Synthesis 164
6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
6.2 Piezoelectric Actuation System Model . . . . . . . . . . . . . . . . 167
6.2.1 Piezoelectric Tube Actuator . . . . . . . . . . . . . . . . . . 167
6.2.2 Inertial and Aerodynamic Loads . . . . . . . . . . . . . . . . 171
6.2.3 Coupled Actuator-Flap-Circuit System . . . . . . . . . . . . 172
6.3 Mechanical Tuning and Electrical Tailoring . . . . . . . . . . . . . . 175
6.3.1 Mechanical Tuning . . . . . . . . . . . . . . . . . . . . . . . 175
6.3.2 Electrical Tailoring . . . . . . . . . . . . . . . . . . . . . . . 176
6.4 Equivalent Electric Circuit Model . . . . . . . . . . . . . . . . . . . 180
6.4.1 Van Dyke Model . . . . . . . . . . . . . . . . . . . . . . . . 180
6.4.2 Analysis and Experimental Verication . . . . . . . . . . . . 183
6.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
Chapter 7
Design and Test of Resonant Actuation Systems 188
7.1 Design Guidelines of the RAS . . . . . . . . . . . . . . . . . . . . . 189
7.1.1 Resonant Actuators with R-L elements . . . . . . . . . . . . 189
7.1.2 Resonant Actuation Systems with Additional Capacitance . 195
7.1.3 Summary of Design Guidelines for the RAS circuitry . . . . 199
7.2 Dynamic Characteristics of the RAS in Forward Flight . . . . . . . 202
7.2.1 A Perturbation Method . . . . . . . . . . . . . . . . . . . . 202
7.2.2 Analysis of Time Responses . . . . . . . . . . . . . . . . . . 206
7.2.3 Vibration Reduction Within Available Actuation Authority 212
7.3 Experimental Realization of the RAS . . . . . . . . . . . . . . . . . 214
7.3.1 Controller Design . . . . . . . . . . . . . . . . . . . . . . . . 214
7.3.2 Bench Top Testing . . . . . . . . . . . . . . . . . . . . . . . 218
7.4 Power Consumption of Piezoelectric Resonant Actuation Systems . 224
7.4.1 Piezoelectric resonant actuators without circuitry . . . . . . 224
vii
7.4.2 Piezoelectric resonant actuators with circuitry . . . . . . . . 226
7.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229
Chapter 8
Conclusions and Recommendations 231
8.1 Summary of Research Eorts and Achievements . . . . . . . . . . . 232
8.2 Recommendations for Future Work . . . . . . . . . . . . . . . . . . 235
Appendix A
Derivation of Strain Measure 238
A.1 Coordinate Transformation . . . . . . . . . . . . . . . . . . . . . . . 240
A.2 Foreshortening Term . . . . . . . . . . . . . . . . . . . . . . . . . . 242
A.3 Deformed Coordinate System . . . . . . . . . . . . . . . . . . . . . 244
Appendix B
Rotor System Matrices and Force Vectors 246
B.1 Strain Energy of Rotor Blades . . . . . . . . . . . . . . . . . . . . . 247
B.1.1 Stiness Coecients of Composite Beam . . . . . . . . . . . 247
B.1.2 Stiness Matrices and Force Vectors . . . . . . . . . . . . . 248
B.2 Kinetic Energy of Rotor Blades . . . . . . . . . . . . . . . . . . . . 249
B.2.1 Stiness Matrix . . . . . . . . . . . . . . . . . . . . . . . . . 249
B.2.2 Mass Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . 251
B.2.3 Damping Matrix . . . . . . . . . . . . . . . . . . . . . . . . 252
B.2.4 Force Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . 253
Appendix C
Formulations using Mathematica 254
C.1 Rotor Strain Energy . . . . . . . . . . . . . . . . . . . . . . . . . . 255
C.2 Rotor Kinetic Energy . . . . . . . . . . . . . . . . . . . . . . . . . . 259
C.3 Rotor Quasi-steady Aerodynamic Loads . . . . . . . . . . . . . . . 266
C.4 Trailing-Edge Flaps Inertial Loads . . . . . . . . . . . . . . . . . . 273
Bibliography 279
viii
List of Figures
1.1 Various sources of a helicopter vibration . . . . . . . . . . . . . . . 2
1.2 Helicopter vibration variation vs. forward ight speed . . . . . . . . 3
1.3 Blade vortex interaction [4] . . . . . . . . . . . . . . . . . . . . . . 3
1.4 Aerodynamic environment in forward ight . . . . . . . . . . . . . . 4
1.5 Vibratory loads transmitted to fuselage [5] . . . . . . . . . . . . . . 5
1.6 Schematic of a dynamic vibration observer [5] . . . . . . . . . . . . 8
1.7 Composite tailoring of helicopter rotor blades [49] . . . . . . . . . . 8
1.8 Active Control of Structural Response (ACSR) systems: (a) engine
platform (b) cabin [20] . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.9 Schematic of Higher Harmonic Control (HHC) . . . . . . . . . . . . 10
1.10 Schematic of Individual Blade Control (IBC) . . . . . . . . . . . . . 11
1.11 Schematic of Active Trailing-edge Flap (ATF) . . . . . . . . . . . . 12
1.12 Schematic of Active Twist Rotor (ATR) . . . . . . . . . . . . . . . 13
1.13 Schematic of semi-active actuators located at blade root region [4] . 13
1.14 Karman SH-2 Seasprite helicopter with servo-aps . . . . . . . . . . 15
1.15 Rotor with on-blade elevons in the NASA Ames Wind Tunnel [30] . 17
1.16 Elevon motion over one rotor revolution with 4/rev voltage excita-
tion (760 RPM, = 0.2) [30] . . . . . . . . . . . . . . . . . . . . . . 17
1.17 Actuator and ap dynamic system model [83] . . . . . . . . . . . . 18
1.18 Blade-pitch indexing for the swashplateless rotor conguration [91] . 19
1.19 Single and dual ap congurations [94] . . . . . . . . . . . . . . . . 20
1.20 AFC being inserted at active blade assembly [3] . . . . . . . . . . . 22
1.21 Macro-Fiber Composite (MFC) actuator . . . . . . . . . . . . . . . 22
1.22 Piezoelectric bender actuators [107] . . . . . . . . . . . . . . . . . . 23
1.23 MD900 Helicopter . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
1.24 Double X-frame actuators for MD900 Helicopter [112] . . . . . . . . 24
1.25 BK117 Helicopter . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
1.26 Flap unit assembly for Eurocopter BK117/EC145 [114] . . . . . . . 25
ix
1.27 Piezoelectric tube actuator for ATF [117] . . . . . . . . . . . . . . . 26
1.28 Passive piezoelectric vibration absorber [121] . . . . . . . . . . . . . 27
1.29 Active-Passive Piezoelectric Network (APPN) [121] . . . . . . . . . 27
1.30 Experimental setup (a) Beam with APPN and negative capacitance,
(b) Circuit diagram of negative capacitance [125] . . . . . . . . . . 28
1.31 Performance comparison: voltage driving response : shunt cir-
cuit without negative capacitance; : shunt circuit with negative
capacitance [125] . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
1.32 Scaling of aerodynamic hinge moment and actuator block torque . . 33
1.33 Conceptual diagram of a resonant actuation system . . . . . . . . . 34
2.1 Vehicle and rotating coordinate systems . . . . . . . . . . . . . . . 39
2.2 Undeformed blade coordinate systems . . . . . . . . . . . . . . . . . 41
2.3 Deformed blade coordinate systems . . . . . . . . . . . . . . . . . . 42
2.4 Cross-section coordinates before and after deformation . . . . . . . 43
2.5 Deformations in terms of Euler angles . . . . . . . . . . . . . . . . . 44
2.6 Schematic of the wake, discretized in space and time [131] . . . . . 77
2.7 Flow chart of an aeroelastic analysis with a free wake . . . . . . . . 79
2.8 Finite elements for composite rotor blades . . . . . . . . . . . . . . 81
2.9 Finite element discretization in time . . . . . . . . . . . . . . . . . . 84
2.10 Vehicle conguration for propulsive trim . . . . . . . . . . . . . . . 89
3.1 Schematic of blade cross-section incorporating a trailing edge ap . 92
3.2 Nomenclature for a thin airfoil with a ap . . . . . . . . . . . . . . 97
4.1 Dual ap conguration for active loads control . . . . . . . . . . . . 110
4.2 Conceptual sketch of dual ap mechanism for active loads control . 111
4.3 Articulated blade coupled ap mode shapes . . . . . . . . . . . . . 115
4.4 Articulated blade coupled lag mode shapes . . . . . . . . . . . . . . 116
4.5 Articulated blade torsion mode shapes . . . . . . . . . . . . . . . . 116
4.6 Control settings of articulated rotor, = 0.35 . . . . . . . . . . . . 117
4.7 Blade tip response of articulated rotor, = 0.35 . . . . . . . . . . . 117
4.8 Comparison of vibratory hub loads for active loads control . . . . . 118
4.9 Dual ap prole for moment reduction with 1/rev control . . . . . . 120
4.10 Control settings of baseline and actively controlled rotors . . . . . . 121
4.11 Harmonics of apwise bending moment along the radial station . . 122
4.12 Flapwise bending moment distribution before control . . . . . . . . 123
4.13 Flapwise bending moment distribution after control . . . . . . . . . 123
4.14 Dual ap prole with 1 and 2/rev control inputs . . . . . . . . . . . 124
4.15 Dual ap prole with 1, 2 and 3/rev control inputs . . . . . . . . . 125
x
4.16 Comparison of vibration index and maximum apwise bending mo-
ment with dierent control inputs . . . . . . . . . . . . . . . . . . . 126
4.17 Vibratory hub shears comparison for active loads control . . . . . . 127
4.18 Comparison of vibration index and maximum apwise moment of
hybrid designed rotor . . . . . . . . . . . . . . . . . . . . . . . . . . 128
4.19 Dual ap proles for retrot and hybrid designs . . . . . . . . . . . 128
4.20 Control settings of baseline, retrot and hybrid designed rotors . . . 129
4.21 Blade non-structural mass and pitch-ap composite coupling sti-
ness, K
25
, distribution . . . . . . . . . . . . . . . . . . . . . . . . . 130
5.1 Various congurations of the rotor with trailing-edge aps . . . . . 135
5.2 Hingeless blade coupled ap mode shapes . . . . . . . . . . . . . . . 139
5.3 Hingeless blade coupled lag mode shapes . . . . . . . . . . . . . . . 140
5.4 Hingeless blade torsion mode shapes . . . . . . . . . . . . . . . . . 140
5.5 Control settings of hingeless rotor, = 0.15 . . . . . . . . . . . . . 141
5.6 Control settings of hingeless rotor, = 0.35 . . . . . . . . . . . . . 141
5.7 Blade tip response of hingeless rotor, = 0.15 . . . . . . . . . . . . 142
5.8 Blade tip response of hingeless rotor, = 0.35 . . . . . . . . . . . . 142
5.9 Vertical wake geometry with the number of beam elements and the
presence of the ap for = 0.15 . . . . . . . . . . . . . . . . . . . . 143
5.10 Vertical wake geometry with the number of beam elements and the
presence of the ap for = 0.35 . . . . . . . . . . . . . . . . . . . . 144
5.11 Blade tip responses with the number of beam elements and the
presence of the ap for = 0.15 . . . . . . . . . . . . . . . . . . . . 144
5.12 Blade tip responses with the number of beam elements and the
presence of the ap for = 0.35 . . . . . . . . . . . . . . . . . . . . 145
5.13 Control settings with the number of beam elements and the presence
of the ap for = 0.15 . . . . . . . . . . . . . . . . . . . . . . . . . 145
5.14 Control settings with the number of beam elements and the presence
of the ap for = 0.35 . . . . . . . . . . . . . . . . . . . . . . . . . 146
5.15 Vibration reduction vs. radial locations of trailing-edge aps . . . . 148
5.16 Lift curve slope vs. aspect ratio for elliptical lift distribution . . . . 149
5.17 Vibration reduction by multi-ap conguration with lift ap . . . . 150
5.18 Flap deection harmonics of multi-ap conguration with the lift
ap, advance ratio: = 0.15, actuator saturation:
sat
f
= 4
o
. . . . 150
5.19 Polar diagram of ap motion for single-ap conguration, advance
ratio: = 0.15, actuator saturation:
sat
f
= 2
o
. . . . . . . . . . . . 152
5.20 Polar diagram of ap motion for dual-ap conguration, advance
ratio: = 0.15, actuator saturation:
sat
f
= 2
o
. . . . . . . . . . . . 153
xi
5.21 Flap deections of dual-ap conguration with 4/rev control input,
advance ratio: = 0.15, actuator saturation:
sat
f
= 2
o
. . . . . . . 154
5.22 Hinge moments in single- and dual-ap congurations with 4/rev
control input, advance ratio: = 0.15, actuator saturation:
sat
f
=
2
o
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
5.23 Comparison of vibratory hub loads, advance ratio: = 0.15, actu-
ator saturation:
sat
f
= 4
o
. . . . . . . . . . . . . . . . . . . . . . . 156
5.24 Flap deections of single-ap conguration, advance ratio: = 0.15,
actuator saturation:
sat
f
= 4
o
. . . . . . . . . . . . . . . . . . . . . 157
5.25 Flap deections of multiple-ap conguration, advance ratio: =
0.15, actuator saturation:
sat
f
= 4
o
. . . . . . . . . . . . . . . . . . 158
5.26 Comparison of vibratory hub loads, advance ratio: = 0.35, actu-
ator saturation:
sat
f
= 4
o
. . . . . . . . . . . . . . . . . . . . . . . 159
5.27 Flap deections of single- and multiple-ap conguration, advance
ratio: = 0.35, actuator saturation:
sat
f
= 4
o
. . . . . . . . . . . . 160
5.28 Hinge moments in single- and multiple-ap conguration, advance
ratio: = 0.35, actuator saturation:
sat
f
= 4
o
. . . . . . . . . . . . 161
6.1 A piezoelectric tube actuator conguration . . . . . . . . . . . . . . 167
6.2 Forces and moments acting on the trailing-edge ap . . . . . . . . . 171
6.3 Schematic of the PZT tube with R-L circuit and negative capacitance173
6.4 Fulcrum amplication mechanism for the PZT tube actuator . . . . 175
6.5 Equivalent electric circuit model of the resonant actuation system . 181
6.6 Trailing-edge ap deections of the resonant actuation system for
Mach-scaled rotor . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184
6.7 Realization of the equivalent electric circuit for the resonant actua-
tion system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
6.8 Comparison of analytical and experimental results of the RAS for a
Mach-scaled rotor . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186
7.1 A piezoelectric network with a series R-L circuit . . . . . . . . . . . 189
7.2 Actuator strokes with the optimal tuning ratios and various coupling
coecients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
7.3 Actuator stroke and bandwidth variations with coupling coecients 191
7.4 Actuator strokes with the optimal inductance tuning and various
resistance tuning values, = 0.4 . . . . . . . . . . . . . . . . . . . . 192
7.5 Schematic of an equivalent mechanical system to a piezoelectric ac-
tuation system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193
7.6 Electric charges with the optimal tuning ratios for various coupling
coecients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194
xii
7.7 A piezoelectric network with a series R-L circuit and an additional
capacitor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195
7.8 Schematic of an equivalent mechanical system to a piezoelectric ac-
tuation system with an additional capacitor . . . . . . . . . . . . . 197
7.9 Actuator strokes with additional capacitance for = 0.5 . . . . . . 198
7.10 Relative actuator stroke and bandwidth variations with modied
coupling coecients . . . . . . . . . . . . . . . . . . . . . . . . . . . 200
7.11 Peak-to-peak ap deections of a resonant actuator with various
ight speeds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207
7.12 Variations of instantaneous frequencies along the azimuth . . . . . . 208
7.13 Time history of ap motions of the actuation system without cir-
cuitry with 4/rev voltage excitation, = 0.35 . . . . . . . . . . . . 208
7.14 Peak-to-peak ap deections of the RAS with various ight speeds . 209
7.15 Time history of ap motions of the RAS with 4/rev voltage excita-
tion, = 0.35 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 210
7.16 Time history of ap motions of the nominal actuation system with
4/rev voltage excitation, = 0.35 . . . . . . . . . . . . . . . . . . . 211
7.17 Comparison of vibratory hub loads for an advance ratio of 0.15
within the available actuator authority . . . . . . . . . . . . . . . . 213
7.18 Comparison of vibratory hub loads for an advance ratio of 0.35
within the available actuator authority . . . . . . . . . . . . . . . . 213
7.19 Controller diagram of the resonant actuation system . . . . . . . . . 216
7.20 Diagram of an adaptive phase controller based on Matlab/Simulink 217
7.21 Experimental set-up for the resonant actuation system . . . . . . . 218
7.22 Equipments used in the experiment . . . . . . . . . . . . . . . . . . 219
7.23 Frequency responses of the PZT tube actuator before and after me-
chanical tuning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
7.24 Analytical predictions of a resonant actuation system: , tuned
system w/o the voltage signal function; , RAS with

= 0.5;
, RAS with

= 0.6 . . . . . . . . . . . . . . . . . . . . . . . . . 221
7.25 Experimental results of a resonant actuation system: , tuned
system w/o the voltage signal function; , RAS with

= 0.5;
, RAS with

= 0.6 . . . . . . . . . . . . . . . . . . . . . . . . . 221
7.26 Time responses of ap deection signal before/after phase control
with 24Hz input signal . . . . . . . . . . . . . . . . . . . . . . . . . 222
7.27 Time responses of ap deection signal before/after phase control
with 26Hz input signal . . . . . . . . . . . . . . . . . . . . . . . . . 223
7.28 Frequency response of current and phase variation for = 0.5 with
dierent resistances . . . . . . . . . . . . . . . . . . . . . . . . . . . 225
7.29 Frequency response of current and its phase with circuitry for = 0.5226
xiii
7.30 Apparent electric power of a piezoelectric actuator with circuitry for
= 0.5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227
7.31 Actuator strokes under the optimal tuning ratios for = 0.5 . . . . 228
A.1 Thin-walled cross section of rotor blade . . . . . . . . . . . . . . . . 239
xiv
List of Tables
2.1 Nondimensionalized parameters . . . . . . . . . . . . . . . . . . . . 46
2.2 Order of terms used in aeroelastic analysis . . . . . . . . . . . . . . 46
2.3 Vehicle properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
3.1 Order of terms for trailing edge aps . . . . . . . . . . . . . . . . . 94
4.1 Constraints and bounds for design variables . . . . . . . . . . . . . 112
4.2 Baseline articulated rotor properties for active loads control . . . . 113
4.3 Trailing-edge ap properties for active loads control . . . . . . . . . 114
4.4 Natural frequencies of baseline articulated rotor . . . . . . . . . . . 114
4.5 Reduction of maximum bending moments maximum moments . . . 121
4.6 Reduction of vibration and moment reductions with dierent control
inputs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
4.7 Comparisons of maximum moment, vibration index and control eorts127
5.1 Hingeless rotor and trailing-edge ap properties . . . . . . . . . . . 138
5.2 Natural frequencies of baseline hingeless rotor . . . . . . . . . . . . 139
5.3 Control input sequences and ap locations for multiple-ap cong-
urations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
5.4 Peak-to-peak hinge moments in single- and dual-ap congurations
with 4/rev control input, = 0.15 . . . . . . . . . . . . . . . . . . . 152
5.5 Peak-to-peak hinge moments in single- and multiple-ap congura-
tions, = 0.35 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
6.1 Piezoelectric material properties of PZT-5H for a Mach-scaled rotor 183
7.1 Design parameters for the RAS circuitry . . . . . . . . . . . . . . . 201
xv
List of Symbols
a Lift curve slope
A Cross-section area of blade
a Acceleration vector
A
f
Cross-section area of the ap
A
ij
, B
ij
, D
ij
Laminate stiness matrices
A
M
Amplication ratio in the actuation system
A
s
Surface area of the PZT electrode
b Airfoil semi-chord, b = c/2
c Blade chord
c
f
The ap chord
c
o
Constant lift coecient
C
e
p
Total capacitance
C
s
p
Piezoelectric material capacitance
C
p
Structural modal damping of the PZT tube actuator
C
add
Added capacitance in the electric network
c
D
55
Elastic constant (open-circuit)
xvi
c
1
Linear lift coecient
C
l
, C
d
, C
m
Lift, drag, and moment coecients

C
ij
Reduced transformed stiness for laminated composites
C Damping;
Modal damping
C
T
Thrust coecient
C
N
, C
M
, C
H
Unsteady normal, pitching and hinge moment coecients of trailing-
edge ap
C(k) Theodorsens lift deciency function
d Oset from blade elastic axis to the ap hinge
d
o
Constant drag coecient
d
1
Linear drag coecient
d
2
Quadratic drag coecient
d
15
Charge constant in piezoelectricity
D
i
Electric displacement component
e Oset between mid-chord and the ap hinge
e
d
Oset of blade elastic axis and aerodynamic center
e
g
Oset of blade elastic axis and c.g.
E
i
Electric eld component
f
o
Constant moment coecient
f
1
Linear moment coecient
F
A
Centrifugal force in Appendix B
F, F Force vector
f
e
Excitation force in equivalent mechanical system to the piezoelec-
tric actuation system
xvii
F
x
, F
y
, F
z
Resultant blade shear forces;
N
b
/rev vibratory hub shear forces (longitudinal, lateral and verti-
cal)
g
15
Voltage constant in piezoelectricity
H Electric enthalpy density function
h Rotor height above vehicle c.g.
H
1
, , H
4
Beam element shape functions
H
t1
, , H
t6
Temporal element shape functions
h
15
Piezoelectric constant
I Identity matrix
I
b
Blade ap inertia
I
f
Trailing-edge ap mass moment of inertia
J Objective function
k Reduced frequency;
Spring stiness
K Stiness;
Modal stiness;
Spring stiness
K
ij
Composite stiness coecient
K
22
, K
33
, K
55
Flap bending, lag bending, and torsion stiness
K
25
Pitch-ap composite coupling stiness
K
c
Electro-mechanical coupling stiness
K
n
Blade loads vector containing the apwise curvature harmonics
K
Q
Inverse of the PZT tube actuator capacitance
K
a
Q
Inverse of the added capacitance
L Inductance
xviii
L
u
, L
v
, L
w
Blade sectional shear forces
L
C
, D
C
, M
C
Circulatory lift, drag and moment per unit length
L
NC
, M
NC
Noncirculatory lift and moment per unit length
l
p
Length of the PZT tube actuator
L
el
Length of beam element
L
LV
Vibration index
L
m
, R
m
, C
m
Inductance, resistance and capacitance in Van Dyke circuit
M Mach number;
Modal damping
m Blade mass per unit length
m
f
The ap mass per unit length
M
H
The ap hinge moment
M
M
, C
M
, K
M
Mass, damping and stiness of the primary system in equivalent
mechanical system to the piezoelectric actuation system
m
m
, c
m
, k
m
Mass, damping and stiness of the secondary system in equivalent
mechanical system to the piezoelectric actuation system
mk
2
m1
, mk
2
m2
Flapwise and lagwise mass moment of inertia per unit length
m
o
Blade reference mass per unit length
M
x
, M
y
, M
z
Resultant blade sectional moments;
N
b
/rev vibratory hub moments (rolling, pitching and torque)
M
LV
Blade moment index
N Axial force
N
b
Number of blades
N
el
Number of beam elements
N
s
Number of PZT segments
N
t
Number of temporal elements
xix
n, s Contour coordinates in the cross-section of rotor blades
p
i
Applied traction in the piezoelectric actuator
q Generalized displacement
q
R
Normal mode displacement
q
t
Actuator stroke (tip displacement)
Q Electric charge
Q
s
Applied charge density
r Radial blade station;
Resistance tuning ratio
r Position vector
r
p
Radial coordinate in the polar coordinate system
r
I
Trailing-edge ap chordwise c.g. from hinge
r
II
Flap radius of gyration about the ap hinge
R Rotor radius;
Resistance
R
i
Inner radius of the PZT tube
R
o
Outer radius of the PZT tube
S Nondimensional time in unsteady aerodynamics of the ap
s Natural coordinate in the nite element
S
f
The ap rst sectional moment
S
ij
Mechanical strain component in piezoelectricity
t Time
T
1
, , T
14
Theodorsen coecients
T
HI
Rotation matrix between x
H
i
and x
I
i
T
RH
Rotation matrix between x
R
i
and x
H
i
xx
T
uH
Rotation matrix between x
u
i
and x
H
i
T
uR
Rotation matrix between x
u
i
and x
R
i
T

St. Venant torsion


T

Valsov torsion
T
ij
Mechanical stress component in piezoelectricity
u, v, w Displacements in the undeformed coordinate system
U
x
, U
y
, U
z
Relative wind velocity components in the deformed coordinate sys-
tem

V Total relative wind velocity

V
b
Relative wind velocity due to blade motions

V
w
Relative wind velocity due to vehicle forward speed and rotor inow
V
x
, V
y
, V
z
Relative wind velocity components in the undeformed coordinate
system
V
a
Voltage across the PZT
V
c
Control voltage in the actuation system
W Weighting matrices
x, y, z Positions in the undeformed coordinate system
x
CG
, y
CG
Longitudinal and lateral distances from vehicle c.g. to rotor hub
x
e
, y
e
Degrees of freedom in equivalent mechanical system to the piezo-
electric actuation system
x
f
i
, e
f
i
The ap deformed coordinate system
x
F
i
, e
F
i
Vehicle coordinate system
x
H
i
, e
H
i
Hub-xed coordinate system
x
I
i
, e
I
i
Ground-xed inertial coordinate system
x
R
i
, e
R
i
Hub-rotating coordinate system
xxi
x
u
i
, e
u
i
Undeformed blade coordinate system
x Vector containing passive design variables
Z
n
Hub loads vector containing the N
b
/rev harmonics
Angle of attack

s
Longitudinal shaft tilt angle (positive nose down)

w
,
w
,
w
Scalar weighting parameters used in weighting maxtrices
Prandtl-Glauert correction factor

p
Precone angle of the undeformed blade

S
11
Impermittivity in piezoelectricity
Lock number

ij
Engineering shear strain component
Variational operator;
Inductance tuning ratio

f
Trailing-edge deection angle

n
Control vector of trailing-edge ap harmonics

sat
Actuator saturation angle
U First variation of strain energy
V First variation of kinetic energy
W First variation of virtual work done by external forces
Incremental quantity
Small parameter in perturbation method
Small parameter in ordering scheme

ij
Normal strain component

T
11
Dielectric constant
xxii
Coordinate in the deformed coordinate system;
Nondimensionalized damping coecient
Coordinate in the deformed coordinate system

t
Temporal nodal displacement
Coordinate in the polar coordinate system

o
Collective pitch angle (rigid blade pitch)

tw
Linear pretwist angle of a rotor blade

tr
Tail rotor collective pitch

1
Total blade pitch angle

75
Rigid blade pitch angle at 75% radial location of blade

1c
Lateral cyclic pitch angle

1s
Longitudinal cyclic pitch angle

x
,
y
,
z
Euler angles
Vector of control settings

ij
Curvature component

x
,
y
Drees linear inow parameters
Axial skew angle
Rotor inow

i
Induced rotor inow

T
Warping function of blade cross-section
Advance ratio (non-dimensionalized forward ight speed)
Coordinate in the deformed coordinate system;
Generalized electro-mechanical coupling coecient

Modied electro-mechanical coupling coecient


Air density
xxiii

p
Piezoelectric material density

s
Material density of rotor blade
Rotor solidity

ij
Stress component

s
Lateral shaft tilt angle (positive advancing side down)

Elastic twist angle of blade


Elastic twist angle with respect to undeformed elastic axis
Modal matrix used in normal mode transformation

s
Applied electric potential
Azimuth angle
Mode shape function of the PZT tube actuator
Angular velocity vector
Nondimensionalized frequency, /
E

D
Open-circuit frequency

E
Short-circuit frequency

E
Short-circuit frequency with added capacitance

p
,
s
Parallel and series frequencies in Van Dyke circuit
Rotor rotational speed
Superscripts
()
A
Aerodynamic quantities
()
E
Short-circuit properties in piezoelectricity
()
D
Open-circuit properties in piezoelectricity
()
fI
Trailing-edge ap inertial contribution
()
fA
Trailing-edge ap aerodynamic contribution
xxiv
()
H
Terms associated with hub
()
I
Inertial quantities
()
L
Linear terms in energy
()
NL
Nonlinear terms in energy
()
o
Baseline value;
Degree with numbers
()
S
Properties at constant strain in piezoelectricity
()
T
Transpose of matrix or vector;
Properties at constant stress in piezoelectricity
()
V
Vehicle quantities
()
1
Inverse of matrix
()

Derivative with respect to a coordinate x.

() Derivative with respect to time t;


Derivative with respect to azimuth in Appendix B
()

Derivative with respect to azimuth ;


Optimal tuning ratios in the electric network

() Terms and responses associated with added capacitance


Subscripts
()
r
Main rotor contributions
()
f
Trailing-edge ap contributions
()
ST
Static response
xxv
Acknowledgments
It is a pleasure to thank the many people who made this thesis possible. It is
dicult to overstate my gratitude to my Ph.D. supervisors, Dr. Edward C. Smith
and Dr. Kon-Well Wang. With their enthusiasm, their inspiration, and their
great eorts to explain things clearly and simply, they helped to make research
fun for me. They provided encouragement, sound advice, good teaching, and lots
of good ideas. I would have been lost without them. I would also like to express
my appreciation to Dr. Farhan S. Gandhi, Dr. Joseph F. Horn and Dr. Mary I.
Frecker for their helpful comments and being advisory committee members.
I would like to thank the many people who have taught me aerospace engineer-
ing: my undergraduate teachers at Inha University (especially Dr. Si-Yoong Ryu,
Dr. Ki-Ook Kim and Dr. Jin-ho Kim) and my graduate teachers (especially Dr.
Maenghyo Cho).
I am indebted to my many student colleagues for providing a stimulating and
fun environment in which to learn and grow. I am especially grateful to Dr.
Jianhua Zhang, Dr. Joseph Sze, Dr. Phuriwat Anusonti-Inthra and Jose Palacios
at Rotorcraft Center of Exellence, and to Dr. Ronald Morgan, Dr. David Heverly,
David Belasco and Dr. Michael Philen at Structural Dynamics and Control Lab.
David and Michael were particularly helpful to conduct experiments, patiently
teaching me how to use an apparatus.
I wish to thank my korean friends Dr. Dooyong Lee, Youngtae Ahn (especially
for providing a ride) and Seongkyu Lee for their friendship. I musk thank all of
the support sta, specically Debbie Mottin and Robin Grandy in the aerospace
department, and Karen Thal in the mechanical department.
I would like to express my deepest appreciation to my family for their support
and encouragement. I wish to thank my parents, my parents-in-law and my sisters.
Lastly, and most importantly, I would like to thank my wonderful wife Eun-Ae for
her love, patience, and the many sacrices she has made.
xxvi
Dedication
This thesis is dedicated to my lovely wife Eun-Ae Ayu.
xxvii
Chapter 1
Introduction
Many methodologies have been explored to reduce helicopter vibration. Trailing
edge aps for such a purpose have been studied for the past twenty years. A brief
overview of the introduction of active vibration controls using trailing edge aps
and smart actuators is presented in the rst section. In the second section, the
previous works in rotorcraft vibration controls and smart actuator development
are described. In the third section, the problem statement and research objectives
of the thesis are presented. The overview of this thesis is presented in the last
section.
2
1.1 Background and Motivation
A rotorcraft has been a very important mode of aerial transportation due to its
capability of vertical take-o and landing, enabling many unique missions such as
rescue operation at sea. It has, however, also been under several serious constraints
such as poor ride quality due to high levels of vibration [1] and noise, restricted
ight envelope, low fatigue life of the structural components, and high operating
cost.
Unlike the wings of conventional xed wing aircraft, helicopter rotor blades
experience periodic motions that result in 1/rev variation of aerodynamic and in-
ertial loads along the azimuth. They also undergo a signicant load uctuation
along the rotor spanwise direction. These operations will cause two major issues.
One is the vibratory load of the rotor hub, coming from higher harmonic compo-
nents of aerodynamic and inertial loads of rotor blades (Figure 1.1). This is the
principal source of helicopter fuselage vibration and has an impact on helicopter
performance, fatigue life of onboard equipment, and passenger comfort. The other
issue is the bending moment of rotor blades, which is a primary source of blade
fatigue stresses.
Figure 1.1. Various sources of a helicopter vibration
3
Vibratory loads in helicopters arise from a variety of sources such as the main
rotor system, the aerodynamic interaction between the rotor and the fuselage,
the tail rotor, the engine and transmission, and atmospheric turbulence, which
generally has low frequencies. However, the most signicant source of vibration in
a helicopter is the main rotor because of the unsteady aerodynamic environment
acting on highly exible rotating blades. The vibration level is generally low in
hover and increases with higher forward speed. There are two regimes: low speed
ight (transition) and high-speed ight, where the vibration levels are critical
(Figure 1.2). At low forward speed ( = 0.1), the blade tip vortices in the wake
stay close to the rotor disk, causing a severe blade vortex interaction, which results
in a substantially higher harmonic loading (Figure 1.3). At high forward speed,
the rotor disk tilts forward and the wake is swept away from the disk plane; the
wake-induced vibrations become small at high advance ratio.
Figure 1.2. Helicopter vibration variation vs. forward ight speed
Figure 1.3. Blade vortex interaction [4]
4
A typical aerodynamic environment of the helicopter main rotor during for-
ward ight is depicted in Figure 1.4, where helicopter ight velocity adds to the
blade element rotating velocities on the advancing side (0
o
< < 180
o
), and
subtracts from it on the retreating side (180
o
< < 360
o
) [2, 3]. The resulting
aerodynamic environment may be characterized as follows: high tip Mach number
on the advancing side, and blade stall eects on the retreating side. A reverse
ow region is also generated at the inboard on the retreating side. Such a com-
plicated environment results in an instantaneous asymmetry of the aerodynamic
loads acting among the blades at dierent azimuthal locations. This results in a
vibratory response of a exible blade structure, adding more complexity to the air
loads asymmetry.
Figure 1.4. Aerodynamic environment in forward ight
The rotor hub acts as a lter, transmitting to the pylon and then to the cabin
only harmonics of the rotor forces at multiples of N
b
/rev, where N
b
is the number
5
of blades. It has been shown that only the N
b
/rev and N
b
1/rev rotating frame
loads contribute to the N
b
/rev xed frame loads (see Figure 1.5). More specically,
the N
b
/rev thrust and torque are caused by corresponding N
b
/rev rotating frame
loads, whereas the N
b
/rev rotor drag- and side forces, and pitching- and rolling
moments are caused by the corresponding N
b
1/rev rotating frame loads. This
result is based on the assumption that all the blades are identical and have the
same periodic motion. This is strictly true only if the rotor blades are tracked
and the helicopter is in trimmed ight. If the blades are not perfectly tracked,
the blade to blade dissimilarity will result in a 1/rev vibration transmitted to the
fuselage [6].
Figure 1.5. Vibratory loads transmitted to fuselage [5]
Various methodologies for vibration reduction have been proposed in litera-
ture. There are several types of active strategies, such as HHC (Higher Harmonic
Control) [711], IBC (Individual Blade Control) [1215], ACSR (Active Control of
Structural Response) [1621], ATR (Active Twist Rotor) [2225], and ATF (Active
Trailing-edge Flap) [2631]. It has been shown that improvements in helicopter
vibration reduction can be achieved through the implementation of active control
technology by smart material. One of the most promising methods is the IBC
using active trailing-edge aps. Therefore, active trailing-edge aps to reduce the
rotorcraft vibration have been given considerable attention in the research com-
6
munity [32, 33].
It has been shown that improvements in helicopter vibration reduction can
be achieved by smart materials, such as piezoelectric materials. Piezoelectric ac-
tuation systems are expected to be compact, light weight, low actuation power,
and high bandwidth devices that can be used for multi-functional roles such as to
suppress vibration and noise, and increase aeromechanical stability. While piezo-
electric materials-based actuators have shown good potential in actuating trailing
edge aps, they can only provide a limited stroke. This limitation can be critical in
cases where large trailing-edge ap deections are required or with large size rotor
blades. The eorts to improve the piezoelectric actuator performance have been
made by researchers in developing amplication mechanisms of various types [34].
Recently, a piezoelectric stack actuator with a double X-frame amplication de-
vice to deect a full-scale ap on a MD Explorer helicopter has been developed and
tested in whirl tower facility [112]. Flap deection angles of 3.5
o
(at 450 V) were
achieved during whirl tower test. A multiple piezoelectric actuator conguration
has been considered and tested in Eurocopter to adjust the required control power
and surface [114]. In this work, a single ap is segmented into three parts, and all
actuator is controlled by the same command.
On the other hand, multiple trailing-edge ap congurations have been studied
to reduce the vibration of helicopter rotor system, in which each actuator operates
independently. Several research works have shown that dual ap conguration is
superior over a single ap in vibration reduction [78, 94].
In general, a single trailing-edge ap works well for the purpose of vibration
reduction of helicopter rotor. With typical control inputs 3,4,5/rev, it has been
demonstrated via numerical simulations that vibration level can be reduced by
about 80%. As mentioned earlier, however, piezoelectric actuators provide a lim-
ited stroke. For the same trailing edge ap deections, the actuator design speci-
cation in multiple-ap conguration is more relaxed than that in single-ap cong-
uration, because the required hinge moment is less due to the small control surface
area.
7
1.2 Literature Review
This section describes previous studies relevant to the present subject. For conve-
nience, the section is divided into ve parts. The section starts with the overview
of helicopter vibration reduction, where various methodologies were examined in
attempt to reduce the vibration. In Section 1.2.2, previous studies using active
trailing-edge aps are highlightened. The review of smart actuation system devel-
opment is given in Section 1.2.3, followed by piezoelectric networks that can help
to enhance the actuator performance. The last subsection is a summary of the
literature review.
1.2.1 Helicopter Vibration Reduction
There have been considerable eorts to reduce the vibration in helicopter [35, 36],
and vibration alleviation methodologies may be categorized into the four groups:
1. Passive means such as vibration absorbers, isolation devices and blade tai-
loring
2. Active vibration absorbing devices in fuselage
3. Direct modication of the excitation frequencies on the rotor blades
4. Active-passive and semi-active vibration reduction technologies
In the past, numerous vibration control devices have been proposed and de-
veloped. The most common passive devices are the dynamic vibration absorber
and the isolation mount. A typical dynamic absorber is a single degree of free-
dom system with a relatively small mass on a spring (Figure 1.6). If tuned to the
excitation frequency, it can generate opposing oscillating force in resonance with
the excitation to enforce a node on the support structure. For example, the dy-
namic hub absorber such as a simple [37] or bilar pendulum has been successfully
applied to S-76 helicopter [38]. The vibration isolation mount is another typical
passive control device, such as pads of rubber or springs are placed between the
vibrating system, to reduce the transmitted force from the vibrating system to
the support structure. In helicopters, the conventional transmission mounting to
8
the airframe is replaced by elastomeric supports [3941]. The blade structural
and aerodynamic properties are tailored using automated optimization techniques
to reduce the vibration [4246]. Composite tailoring (Figure 1.7) has also been
studied extensively since composite material provides excellent opportunities for
developing light weight/high stiness structures, as well as providing elastic cou-
plings for potential optimal designs [4751]. A passive approach has merit in that it
does not need additional power. Passive methods, however, usually exhibit limited
performance and cannot adapt to system and operating condition changes.
Figure 1.6. Schematic of a dynamic vibration observer [5]
Figure 1.7. Composite tailoring of helicopter rotor blades [49]
In active vibration absorbing devices, active control actions are directly applied
on the airframe. Two successfully ight tested airframe-based active controls are
9
active vibration suppression [16, 17] and Active Control of Structural Responses
(ACSR) tested on a modied S-76B helicopter [1820]. In ACSR system, vibration
sensors are placed at key locations in the fuselage, where minimal vibration is
desired (Figure 1.8). Depending on the vibration levels from sensors, a controller
calculates proper actions for actuators, such as electro-hydraulic, piezoelectric and
inertial force actuators, to reduce the vibration. The ACSR system has successfully
made into production on the helicopter such as the Westland EH101 and the
Sikorsky S-92 Helibus [20].
Figure 1.8. Active Control of Structural Response (ACSR) systems: (a) engine platform
(b) cabin [20]
Passive and active absorbing devices are still used in most of the rotorcraft
ying today although they also bring unavoidable penalties in terms of weight and
tend to aect vibrations only at discrete points. Therefore, eorts to modify di-
rectly the excitation forces have been sought by modifying unsteady aerodynamic
forces acting on the rotor blades. Among them, Higher Harmonic Control (HHC)
systems have received the most attention [711]. In this approach, servo-actuators
are used to excite the conventional swashplate in the collective, longitudinal cyclic
and lateral cyclic modes at the frequency of N
b
, resulting in blade pitching oscilla-
tions at three frequencies of (N
b
1) , N
b
and (N
b
+1) of HHC in the rotating frame
(Figure 1.9). These higher harmonic blade pitch motions can generate additional
10
unsteady aerodynamic and oscillatory inertial loads with the right amplitudes and
phases to alleviate hub vibration. Therefore, the vibration can be suppressed at
the source before it is propagated into the fuselage. Even though HHC is shown
to be highly eective, several drawbacks have impeded the implementation of the
HHC concept on production helicopters. One of drawbacks is that HHC system
uses the primary control system (swashplate and pitch links) to transfer higher
harmonic pitch inputs to the rotor blades. Thus, considerable power is required to
operate the actuators.
Figure 1.9. Schematic of Higher Harmonic Control (HHC)
An alternative to HHC is the Individual Blade Control (IBC) [1215], in which
each blade is individually controlled in the rotating frame over a wide range of
desired frequencies. This control concept is a more general approach that removes
some of the limitations of HHC such as the xed excitation frequency. Usually,
hydraulic actuators are mounted in conjunction with the blade pitch links in the
rotating frame. The control inputs to the actuators are based on the feedback
signals from the sensors mounted on the blades. A hydraulic slip ring unit is
required to transmit the hydraulic power to the actuators in the rotating frame
(Figure 1.10).
Numerous analytical studies, wind tunnel tests and ight tests of HHC or IBC
have demonstrated their potential for the substantial vibration reduction of up
11
Figure 1.10. Schematic of Individual Blade Control (IBC)
to 90 percent. However, both approaches have limitations on their practical im-
plementation. Apart from the considerable weight penalties and high cost, large
actuation power is needed to pitch the entire blades. The complexity of their
actuation systems as well as their adverse impacts on vehicle reliability and main-
tainability has hampered their availability. Furthermore, since both HHC and IBC
introduce control through the conventional swashplate, which is the primary ight
control system of the helicopter, it will inuence the airworthiness of the helicopter.
Another IBC concept using a trailing-edge ap has been explored extensively
for vibration reduction. This concept uses a small ap on each blade to generate
the desired unsteady aerodynamic loads (Figure 1.11). This concept can be equally
eective, but uses less power than the conventional IBC system. In this approach,
a partial span trailing edge ap is located in the outboard region of the blade.
The active ap control inputs aect the blade inertial loads and rotor dynamics
as well as the unsteady aerodynamic loads. Since only a very small portion of the
blade, about 4 5 percent [52], is actuated, this approach needs much less control
power compared to HHC or IBC. Furthermore, this active ap control system is
totally separated from the conventional swashplate; thus, it has little inuence
on airworthiness. A number of analytical simulations [2629, 31], some wind tun-
nel tests [30, 67] and full-scale whirl tower test [112, 114] of the active ap have
demonstrated that it has the potential to signicantly reduce the vibratory loads,
alleviate noise, enhance the rotor performance and handling qualities. Additional
12
information on vibration reduction using active trailing-edge aps can be found in
two recent survey papers [32, 33]. The detailed literature review will be described
on the active trailing-edge aps in Section 1.2.2.
Figure 1.11. Schematic of Active Trailing-edge Flap (ATF)
With the emergence of smart materials, such as Active Fiber Composites (AFC)
[53], Macro-Fiber Composite (MFC) [54] and piezoelectric materials, the Active
Twist Rotor (ATR) concept [2225] has been proposed (Figure 1.12). One of the
advantages is simplicity of its actuation mechanism compared to the active trailing-
edge ap actuation. The ATR concept has also a merit in that it does not increase
the prole drag of the blade just as discrete ap does. While the ATR technology
can produce a signicant vibration reduction, power requirements are expected to
be much higher than those for active trailing-edge aps.
Other vibration reduction technologies are active-passive or semi-active con-
cepts developed to combine the advantages of both purely active and passive con-
cepts. Although active means can more eectively reduce helicopter vibration and
can be adaptive to system operating condition changes, their performances are
often limited by the authority of the actuators. Active-passive approach has been
proposed and investigated to make up for the weak point (actuator authority) in
active ap system. A hybrid design approach can reduce the required ap deec-
tions via active-passive optimization while retaining the same vibration level as
that of the conventional active ap control [5557]. A semi-active approach using
13
Figure 1.12. Schematic of Active Twist Rotor (ATR)
cyclic variation of the eective ap, lag, and torsion stiness or damping varia-
tions has been proposed for helicopter vibration reduction [58, 59]. This approach
involves evaluating sensitivity of hub variation to cyclic changes in stiness and
damping of the blade root region (Figure 1.13).
Figure 1.13. Schematic of semi-active actuators located at blade root region [4]
Active vibration reduction systems are comprised of similar basic components;
sensors, actuators, and a controller. In most active vibration control methodolo-
gies, the major issue is the actuator itself. Thus numerous smart actuators are
under development. These concept designs of smart actuators have been either
wind tunnel tested or bench tested for helicopter vibration reduction. The key to
on-blade vibration control has been the advent of smart structures, in particular
those incorporating electrically driven piezoceramic materials, exhibiting high en-
14
ergy density and high bandwidth. The on-blade smart systems open a new domain
for vibration control, aeromechanical stability augmentation, handling qualities en-
hancement and noise reduction. The detailed literature review of smart actuation
system developments will be described in Section 1.2.3.
1.2.2 Active Trailing-edge Flaps
The trailing edge aps were used to control the 1/rev rotor primary controls (collec-
tive and cyclic) for earlier helicopters,such as Pescara, dAscanio and Kaman. The
extension of this concept for providing control at higher harmonics was identied
in a early work on multicyclic control [60].
In 1970s, the rst investigation of the trailing edge aps (servo-aps, see Fig-
ure 1.14) for multicyclic vibration control was conducted,called as Kaman Multi-
cyclic Controllable Twist Rotor (MCTR). The MCTR utilized the rst harmonic
servo ap inputs incorporated with conventional pitch control, and the multicyclic
control was generated by the electrohydraulic actuators mounted in the rotating
frame. The wind tunnel testing with the maximum ap inputs up to 6 degrees
at frequencies up to 4/rev were conducted, and showed signicant reduction in
vibratory hub loads with appropriate 2/rev inputs [61]. McCloud III has studied
the feasibility of reducing both vibration and blade loads using a single servo-ap,
where he applied the scheduled multi-cyclic control inputs at 1/rev, 2/rev, 3/rev,
and 4/rev to the MCTR [26]. His results have shown that multi-cyclic control can
achieve both vibration and bending moment reductions with a large 1/rev control
input.
With the advances in the development of smart material-based actuators, active
vibration controls using trailing-edge aps are revisited for helicopter vibration
reduction in 1990s. An analytical study on vibration reduction in a four-bladed
hingeless rotor using an actively controlled servo ap was conducted by Millott and
Friedmann [27, 6265]. A time-domain simulation to reduce 4/rev hub loads was
implemented, and the reduction in vibratory hub loads levels around 90 % were
reported. It was noted that spanwise ap position and blade torsional stiness
were key factors governing the performance in vibration reduction. The vibration
15
Figure 1.14. Karman SH-2 Seasprite helicopter with servo-aps
level for rotor with a single trailing-edge ap using both the fully elastic and rigid
blade models was investigated. It was reported that the predicted vibration level
using the rigid blade model is much less than that of using the elastic blade model.
An extensive proof-of-concept investigation of plain trailing edge ap was con-
ducted by McDonnell Douglas Helicopter Systems. A 12-foot diameter Active Flap
Rotor (AFR) model was tested in NASA Langley 14 22 wind tunnel. The trail-
ing edge ap was actuated by a cam-follower and cable arrangement that could
provide various ap inputs by interchanging and rotating the programming cams.
The reductions in hub loads up to 80 % were demonstrated during testing [6668].
Milgram and Chopra [28, 6971] carried out an analytical study on the eec-
tiveness of plain trailing edge aps for vibration suppression. This analysis incor-
porated an unsteady aerodynamic model based on indicial response functions of a
apped airfoil [72, 73] as well as a free wake model [74, 75] that are implemented
into the UMARC [76]. The reductions in vertical hub shear loads up to 98% was
reported using an open-loop controller with actuation frequencies of 3/rev and
4/rev. The validation of the analysis with test results from McDonnell Douglas
AFR test was carried out,and correlation between predicted and measured results
16
was generally fair. It was reported that varying the phase angle of the ap motion
had a signicant eect on the blade 4/rev atwise and inplane bending loads.
Straub and Hassan [29] conducted a conceptual sizing and design study for a full
scale demonstration trailing edge ap system. Structural parameters were investi-
gated to determine a feasible ap/actuator combination with the consideration of
the blade-ap-actuator dynamics.
Based on the previous researches [27, 6265], Mytle and Friedmann developed
a new two-dimensional compressible unsteady aerodynamic model using a rational
function approximation approach for the blade-ap combination [77, 78]. It was
found that the control power and active ap deections are signicantly higher and
larger as compared to those predicted with quasi-steady aerodynamics. They also
examined and compared various ap congurations, such as servo-ap, plain-ap
and dual-ap.
Fulton and Ormiston conducted a hover test on a two-blade, 7.5 ft diameter
dynamic rotor model with 10% on-blade elevons driven by piezoceramic bimorph
actuators in the U.S. Army Aeroightdynamics Directorate hover test chamber
[79]. The test was successful to provide an encouraging basis for wind tunnel
testing. Subsequently, a wind tunnel test [30] was conducted at NASA Ames
Research Center (Figure 1.15). The test was performed at advance ratios from
0.1 to 0.3 at two dierent tip speeds (450 RPM and 760 RPM). Two important
test, the elevon phase sweeps and frequency sweeps, were carried out to provide
a measurement of elevon eectiveness and the rotor /elevon dynamic response
characteristics. It was observed that the azimuthal time history of the elevon
motion with 4/rev voltage excitation includes not only 4/rev elevon motion but
also moderate 1/rev content, due in part to the azimuthal variation of elevon
aerodynamic stiness opposing the PZT actuator stiness (Figure 1.16).
de Terlizzi and Friedmann investigated Blade Vortex Interaction (BVI) eects
on advanced geometry rotors [80]. The rotor wake model used in this study was
extracted from the comprehensive rotor analysis code CAMRAD/JA [81]. The
results indicated that the active trailing edge ap input angles of 15
o
are required
17
Figure 1.15. Rotor with on-blade elevons in the NASA Ames Wind Tunnel [30]
Figure 1.16. Elevon motion over one rotor revolution with 4/rev voltage excitation
(760 RPM, = 0.2) [30]
to alleviate the BVI induced vibratory hub loads at the advance ratio of 0.15. A
phase sweep simulation of an active ap was carried out and correlated with the
wind tunnel test results at NASA Ames by Fulton and Ormiston [30], and the
correlation was found to be good in most cases.
Straub and Charles [31] have simulated vibration reduction by active ap
18
trailing-edge aps using the comprehensive analysis code CAMRAD II [82]. This
study was in support of the development of a full scale rotor test with smart ac-
tuators. The composite, bearingless rotor of MD900 with trailing-edge aps was
studied for active control of vibration, noise and aerodynamic performance. The
coupled blade/ap/actuator dynamics and their eect on ap motions and sys-
tem stability were investigated. Recently, Shen and Chopra [83, 84] developed a
comprehensive aeroelastic analysis of a fully coupled blade-ap-actuator system
based on UMARC (Figure 1.17). Parametric study was conducted to examine the
coupling eect on the vibration reductions. It was shown that actuator dynamics
can not be neglected, especially for a torsionally soft smart actuation system. This
work was extended to consider the stability of active ap rotors [85].
Figure 1.17. Actuator and ap dynamic system model [83]
Other, recent, studies have addressed the issue of individual blade control of a
helicopter with dissimilar rotor blades. The blade control is implemented using a
conventional HHC algorithm coupled with a rened Kalman lter approach [86,87].
The controller was shown to reduce successfully the vibratory hub loads due to
blade dissimilarities.
In addition to oering improved vibration, acoustic noise, and rotor aerody-
namic performance, the on-blade control concept may also perform the primary
ight control function of the rotor and thereby replace the swashplate, pitch-
links, and hydraulic actuators of the traditional helicopter control system [88].
A few studies of advanced rotor control systems for both primary ight control
19
and rotor vibration control using active trailing-edge aps have been performed.
Army-sponsored studies by Bell and McDonell-Douglas Helicopter Company to
explore the potential benets of active control while eliminating the need for por-
tions of the conventional helicopter ight control system [89]. Recently, Shen and
Chopra [9092] developed a comprehensive rotor code to analyze the swashplateless
rotor conguration. A multicyclic controller was implemented, and the feasibility
of trailing-edge ap performing both primary control and active vibration con-
trol was examined. With a large blade pitch index angle of 16 deg (Figure 1.18),
the required half peak-to-peak values of trailing-edge ap deections were below
6 deg. Among other technical issues, either conventional electromechanical and
hydraulic actuators or a large blade pitch index angle introduced a number of
practical drawbacks.
Figure 1.18. Blade-pitch indexing for the swashplateless rotor conguration [91]
On the other hand, multiple trailing-edge ap congurations have been studied
to reduce the vibration of helicopter rotor system, in which each actuator operates
independently. Myrtle and Friedmann [78] have shown that the dual ap cong-
uration (Figure 1.19) is almost completely unaected by the change of torsional
stiness of rotor blade. Recently, Cribbs and Friedmann [93] have developed the
ap deection saturation model through an automated approach to reduce the
required maximum ap deection. They have shown that the imposition of satu-
ration of ap deection could result in the dierent prole and reduced magnitude
of the active ap while maintaining almost the same vibration level as models
20
without actuator saturation. The actuator saturation model has been extended
to reduce the vibration due to dynamic stall using a dual ap conguration [94].
In this work, the eect of dynamic stall was incorporated by using the ONERA
dynamic stall model, and the drag due to the ap deection was also considered.
They showed that dual ap is superior over a single ap in vibration reduction.
Recently, Liu et. al conducted a study of the combined helicopter noise and vibra-
tion problem using a dual ap conguration [95, 96]. An acoustic prediction based
on WOPWOP [97] was combined with the aeroelastic analysis based on a exible
blade. It was observed that a dual ap conguration is more eective than a single
ap conguration when the actuator saturation is considered. It was also reported
that the noise penalty is mainly due to the large ap deections obtained when
saturation limits are not imposed.
Figure 1.19. Single and dual ap congurations [94]
1.2.3 Smart Actuation System Development
There appeared an opportunity of having multiple light weight sensors/actuators
embedded or surface-mounted at several locations in rotor blades and optimally
distribute actuation with the aid of modern control algorithm [34, 98, 99]. By
employing active materials for such sensors/actuators in order to implement indi-
vidual blade control, one can potentially obtain advantages in terms of weight and
power consumption when compared to traditional hydraulic systems. These new
21
actuators only requires electrical power to operate. Two main concepts have been
under development for the active material application: rotor blade ap actuation
and integral blade twist actuation [100].
For the integral blade twist actuation concept, the actuators may be embed-
ded throughout the structure, which provides redundancy in operation. A major
challenge with integral blades is to develop a design that presents sucient twist
authority while providing the torsional stiness required for the aeroelastic per-
formance of the blade. Chen and Chopra [22], based on the piezoelectric actuator
presented in Barrett [101], built and tested a 6-ft diameter two-bladed Froude-
scaled rotor model with banks of piezoceramic crystal elements in 45
o
embedded
in the upper and lower surfaces of the test blade.
Rodgers and Hagood [102] manufactured and hover tested a single one-sixth
Mach scaled model blade of CH-47D where the integral twist actuation was ob-
tained through the use of Active Fiber Composite (AFC) [53]. An intentional
reduction by 50% on the baseline torsional stiness was imposed and regarded
to improve twist actuation. Hover testing on the MIT Hover Test Stand Facility
demonstrated tip twist performance between 1
o
and 1.5
o
in the rotating environ-
ment. Another example of an integral blade twist concept has been studied as part
of a NASA/Army/MIT Active Twist Rotor cooperative agreement program [103].
The structural design of the ATR prototype blade employing embedded AFC actu-
ators (Figure 1.20) was conducted based on a newly developed analysis for active
composite blade with integral anisotropic piezoelectric actuators [104].
The Macro-Fiber Composite (MFC) has been recently developed at NASA
Langley based on the same idea as the AFC in using the piezoelectric bers under
interdigitated electrodes [54]. In this actuator, shown in Figure 1.21, the piezo-
electric bers are manufactured by dicing from low-cost monolithic piezoceramic
wafers. Thus, it retains most advantageous features of the AFC with a poten-
tially lower fabrication cost. This actuator is currently being tested for its basic
characteristics, and it has been considered for use in many aerospace applications.
The controllable twist rotor approach makes it easy to embed smart materials
22
Figure 1.20. AFC being inserted at active blade assembly [3]
Figure 1.21. Macro-Fiber Composite (MFC) actuator
into a rotor blade and results in an aerodynamically clean blade. However, since
the entire blade must be twisted, it is very dicult to achieve the targeted control
authority of 2 degrees of blade tip twist with the current state of the art smart
materials [34].
Another approach is the discrete active trailing edge ap driven by smart mate-
rial actuators. The primary concerns in this approach are to obtain high actuation
force and stroke with minimal weight penalty, and that the actuation mechanism
must be designed to t into the geometric connes of the blade structure. Over
the past decade, several actuators have been developed to deect a rotor blade
trailing edge ap. The ideal trailing edge ap actuator would generate enough
torque and stroke to overcome the aerodynamic, inertial loads, friction, and all
23
other restricting moments. Piezoelectric actuators must be used in conjunction
with an amplication mechanism to be eective in rotorcraft trailing edge ap
applications because of the small-induced strain capability of the smart material
actuators.
Flap actuator designs that incorporate piezoelectric bender actuators have a
low force output and in general, are restricted to small scale wind tunnel models
[105108]. One of the piezoelectric bender actuators developed by Koratkar and
Chopra [107] is shown in Figure 1.22.
Figure 1.22. Piezoelectric bender actuators [107]
Piezoelectric stack actuators, typically used in larger scale applications, have a
larger force output than benders but produce a relatively smaller stroke. In order
to achieve the required ap deections, stack ap actuator designs require more
complex amplication mechanisms than benders. Trading the actuation force with
actuation displacement using mechanical amplication makes it possible to use the
piezostack to actuating the trailing edge ap of a full-scale blade. Lee and Chopra
designed an actuator to drive a full-scale trailing edge ap using two piezostacks
with double-lever [109] and bi-directional double-lever [110] amplication mech-
anism. In a parallel study, Prechtl and Hall designed a X-Frame trailing edge
servo ap stack actuator [111]. The Mach-scaled rotor model with the trailing edge
ap servo-ap actuation was hover tested in wind tunnel. Straub et. al [112, 113]
used a piezostack with a Double X-frame amplication device (Figure 1.24) to
24
deect a full scale ap on a MD Explorer helicopter (Figure 1.23). Flap deection
angles of 3.5 (at 450 V) were achieved during whirl tower test [112].
Figure 1.23. MD900 Helicopter
Figure 1.24. Double X-frame actuators for MD900 Helicopter [112]
Eurocopter [114] has designed a full-scale actuator trailing edge ap for the
BK117/ EC145 helicopter (Figure 1.25). A multiple piezoelectric actuator cong-
uration has been considered and tested to adjust the required control power and
surface. In this work, a single ap is segmented into three parts, and all actuator
is controlled by the same command. A 16% span ap was deected 5 in whirl
tower testing [114]. In this test, the units of the trailing edge ap of the blade was
25
cut out and the foam used as support between the upper and lower blade skin was
substituted by a at box made from carbon ber (Figure 1.26).
Figure 1.25. BK117 Helicopter
Figure 1.26. Flap unit assembly for Eurocopter BK117/EC145 [114]
There are many alternative concepts proposed for actuating trailing edge aps.
Clement et al designed a piezoceramic C-Block actuators for active ap system
and a bench-top test was conducted [115]. Recently, a new piezoelectric induced
shear tube actuator (Figure 1.27) was developed and bench tested by Centolanza
and Smith [116]. By exploiting the high energy density of the d
15
mode of the
piezoceramic material, along with the well-matched shape factor of the piezoelectric
tube, this new actuation system can generate higher force and stroke than the
26
current piezo bender actuators or piezo stack actuators. Moreover, it has a clearer
and simpler amplication mechanism.
Figure 1.27. Piezoelectric tube actuator for ATF [117]
1.2.4 Piezoelectric Networks
Because of their electro-mechanical coupling characteristics, piezoelectric materi-
als have been explored extensively for both active and passive vibration control
applications. In a passive situation (Figure 1.28), the piezoelectric materials are
usually integrated with an external shunt circuit [118,119]. Electrical eld/current
will then be generated in the shunt circuit because of the electro-mechanical cou-
pling feature. It has been shown that with proper design of shunt components
(inductor, resistor, or capacitor), on can achieve the electrical damper or electrical
absorber eects.
The integration of an inductively shunted piezoelectric element with an active
source was rst studied by Agnes [120]. It was observed that the active-passive
hybrid conguration retains the passive damping ability while allowing additional
performance using active control. Because of the inductive shunt, this cong-
uration is best suited for narrow-band applications. It is important to note that
shunting the piezoelectric in a passive manner does not preclude the use of shunted
27
Figure 1.28. Passive piezoelectric vibration absorber [121]
piezoelectric materials as active actuators. Indeed, the integration of the passive
and active approach, such as the one illustrated in Figure 1.29 often referred to
as an active-passive hybrid piezoelectric network (APPN), has shown promising
results [121]. This APPN conguration not only preserves the passive damping
ability of the shunt circuits, but also, has been found to amplify the active control
authority around the tuned circuit frequency [122125]. A few studies have been
performed on vibration control using APPN concept, which includes simultaneous
optimal control/optimization for determining passive parameters and active gains
in APPN, multi-input multi-output applications [126], and integration of APPN
with traditional viscoelastic treatments [127].
Figure 1.29. Active-Passive Piezoelectric Network (APPN) [121]
28
Figure 1.30. Experimental setup (a) Beam with APPN and negative capacitance, (b)
Circuit diagram of negative capacitance [125]
On the other hand, the performance of piezoelectric material-based actuators
can be augmented with electric networks that include inductor, resistor and neg-
ative capacitor [125, 128130]. Tang and Wang performed analysis and experi-
ment for active-passive hybrid piezoelectric networks [125]. It was shown that the
electro-mechanical coupling of the integrated system is increased by introducing
negative capacitance that is realized by a negative impedance convert circuit with
an operational amplier (Figure 1.30). The overall control authority was signi-
cantly improved since the structure can be driven to a higher amplitude given the
same level of voltage input. The bandwidth of the amplication eect was greatly
increased due to the negative capacitance (Figure 1.31).
29
Figure 1.31. Performance comparison: voltage driving response : shunt circuit
without negative capacitance; : shunt circuit with negative capacitance [125]
30
1.2.5 Summary of Literature Review
Literature review shows that a great deal of research has been conducted in de-
signing passive and active devices for helicopter vibration reduction. Several top-
ics were addressed including helicopter vibration reduction methodologies, active
trailing-edge aps, smart actuation system development, and piezoelectric net-
works to enhance the actuator performance.
The traditional passive approaches to vibration reduction, such as absorbers
and isolation mountings have generally not proven to be eective and/or ecient
enough to realize the desired comfort level of jet-smooth rotorcraft ight. Vi-
bration reduction using active control has been extensively explored. Despite the
promising performance of the HHC and IBC concepts for vibration reduction,
many practical concerns have to be addressed for these systems can be used in a
production helicopter. They use the primary control system (swashplate and pitch
links) to transfer pitch inputs to the rotor blades. In the event of failure of con-
trol systems, the pilot may not have full control of the helicopter (airworthiness).
Although the ACSR approach oers signicant vibration reduction performance
for modern helicopters, the vibration reductions obtained may be localized at the
sensor locations, regardless of vibration levels elsewhere in the airframe, and the
vibration levels in the rotor system are left unaltered. The active twist rotor ap-
proach (ATR) makes it easy to embed smart materials into a rotor blade and
results in an aerodynamically clean blade. However, since the entire blade must
be twisted, it is very dicult to achieve the targeted control authority
Active trailing-edge aps have been explored extensively for vibration reduc-
tion, since this concept uses a small ap on each blade to generate the desired
unsteady aerodynamic loads. This can be equally eective, but uses less power
than the conventional IBC system. A number of analytical simulations and wind
tunnel tests of active trailing-edge aps have demonstrated that it has the potential
to signicantly reduce the vibratory hub loads. In addition to oering improved
vibration and acoustic noise, active trailing-edge aps could be used as the pri-
mary control ight control system (i.e., swashplateless helicopter) and active loads
control system with the low frequency inputs (e.g., 1/rev and 2/rev).
31
With the emergence of smart materials, such as AFC, MFC and piezoelectric
materials, the ATR convept was proposed. This has a merit in that it does not
increase the prole drag of the blade. However, it is dicult to achieve the tar-
geted control authority with the current state of the art smart materials. On the
other hand, many actuators based on piezoelectric materials have been developed
for active trailing-edge aps. Among them, recently, two piezoelectric actuators
were tested in whirl tower facilities. One is a Double X-frame actuator for MD
Explorer helicopter, and the other is a DWARF actuator for Eurocopter BK117
helicopter. Flap deection angles of 3.5
o
5
o
were achieved during whirl tower
tests.
On the other hand, the performance of piezoelectric material-based actuators
can be augmented with electric network. Piezoelectric materials with electrical
networks have been utilized to create the shunt damping for structural vibration
suppression. It was also recognized that such networks not only can be used for pas-
sive damping, they can also be designed to amplify the actuator authority around
the tuned circuit frequency. It was demonstrated that the electro-mechanical cou-
pling of the integrated system can be increased by introducing electric components
such as inductor, resistor and negative capacitor.
Based on the literature review, in the problem statement and objective section
that follows, the problem associated with current state-of-the-art actuators and
active trailing-edge ap controls are described, then the objectives of the current
study are specically listed.
32
1.3 Problem Statement and Objectives
From the review of smart actuators for trailing edge aps, it is observed that active
material actuated rotor blade controls are emerging as a viable solution. Active
trailing edge ap experiments with piezoelectric material based actuators have
shown more remarkable results than actively twisted blades with embedded piezo
elements. However, the major drawback of smart material based actuators is their
low stroke. Although many actuation schemes have been examined to amplify
the strokes by trading actuation force with displacement, so far only 3 5
degrees peak to peak active ap deections are achievable in whirl tower test (hover
condition). The achieved active ap deections are far away from the required
control authority for the large size full-scale rotor blades, since the aerodynamic
hinge moment will obviously increase as the size of rotor blades and forward ight
speed increase.
The performance of the smart actuators with varying scale of rotor blade is de-
picted in Figure 1.32 showing the aerodynamic hinge moment and actuator block
torque as a function of blade chord. It can be seen that the aerodynamic mo-
ment increases with blade chord at a larger rate than the block torque available.
The state-of-the-art actuators for trailing edge ap could be able to generate a
useful combination of ap deection and block torque for a light class helicopter
application, such as MD900 and BK117, (blade chord of 10 inches), however, the
performance of the actuators will be drastically reduced if the blade chord length
is increased further. Therefore, the actuator stroke remains a critical design limit
in practical implementation of active aps for large size rotors. In other words,
actuator authority for heavy lift class helicopters presents a major technical barrier.
The goal of this research is to address the current issues and advance the state
of the art in helicopter rotor vibration control by exploring new actuation ideas
and using multiple trailing-edge aps. To achieve this goal, general tasks involve
the development of rotor systems with multiple trailing edge aps and resonant
actuation systems. A resonant actuation system is introduced to enhance the
eectiveness of piezoelectric actuators. Through mechanical tailoring, the reso-
nant frequencies of the actuation system (including the piezoelectric actuator and
33
Figure 1.32. Scaling of aerodynamic hinge moment and actuator block torque
the related mechanical and electrical elements for actuation) can be tuned to the
required actuation frequencies that are 3, 4, and 5/rev for typical four-bladed he-
licopters. This will increase the authority of the actuation system; however, it
could be hard to control because of the narrow operating bandwidth and dramatic
change in phase. These issues can be resolved through electric circuit tailoring.
Then the actuation resonant peak can be signicantly broadened and attened,
and the changes in phase can be much gradual. In this case, one can achieve a
high authority actuation system without the negative eects of resonant problems
(Figure 1.33).
A resonant actuation system, however, may not cover the entire operating
frequencies (3,4,5/rev). This can be resolved using the rotor blades with multiple
trailing-edge aps, in which each ap is designed to operate at a single frequency
that is one of the operating frequencies. In this way, one can achieve the ultimate
solution to overcome the barrier of actuator authority for large size full-scale rotor
blades, and a lighter actuation system could be achieved for MD900 or BK117 class
34
Figure 1.33. Conceptual diagram of a resonant actuation system
helicopters that are currently considered for ight test. The specic research goals
for this new concept, as well as some of the proposed approaches, are described
below:
1. Investigate the feasibility of using multiple trailing edge aps for vibration
reduction and blade loads reduction.
In pursuing this objective, the initial goal is the development of an com-
prehensive rotor aeroelastic analysis incorporating multiple trailing edge
aps. The unsteady compressible model developed by Hariharan and
Leishman [73] is used for trailing-edge aps, which is able to predict re-
liable ap hinge moments. The free-wake model developed by Tauzsig
and Gandhi [131, 132] is incorporated to predict realistic vibratory hub
loads.
Following this, a multicyclic control algorithm for multiple trailing edge
aps with the resonant actuation concept will be developed to reduce
either the blade loads or the vibration. An actuator saturation will be
also considered to reect the available actuator authority and to avoid
the additional aerodynamic interference to the vehicle trim, which is
based on the bisection method.
Active loads control for an articulated composite rotor. An active loads
control method will be developed to reduce the blade loads as well
as the vibration. This can be achieved by dynamically straightening
35
the blade, which mimics the behavior of a rigid blade, via dual ap
conguration. An aeroelastic analysis will be performed for the high
speed ight condition.
A parametric study will be performed to determine the ap location along
the rotor spanwise direction for the vibration reduction. The nite wing
eect of trailing-edge aps will be also examined since a ap span is
smaller than one in a single ap conguration. Finally, vibratory hub
loads will be compared among various trailing-edge ap congurations.
2. Develop a resonant actuation system (RAS) for active trailing-edge aps that
will be used in multiple trailing-edge ap congurations
Develop an analytical model of the resonant actuation system. Up to
date, no systematic methods have been derived for tailoring the electri-
cal parameters, such that a desired actuator authority can be achieved.
An equivalent mechanical system will be examined to provide better
physical understanding.
Investigate the time-varying characteristics of resonant actuation sys-
tem in forward ight. The time-varying characteristics of a RAS will be
analyzed via a perturbation method and frequency response functions
between peak-to-peak ap deections and control voltage input.
Realize the resonant actuation system, especially under high voltage op-
erations. A method of implementing the electric networks will be real-
ized via a digital signal processor (DSP). A controller will be designed to
adapt the change in phase at the vicinity of resonant frequency (one of
operating frequencies). Through this eort, the performance of a RAS
will be validated experimentally via a bench-top test emulating hover
condition.
36
1.4 Overview of Dissertation
This thesis consists of eight chapters, which are organized as follows:
1. The rst chapter introduces the background and motivation for the current
research. A comprehensive review of literature related to the present research
topics is included, and the research objectives are stated.
2. The second chapter describes the physical and mathematical models of the
helicopter rotor system used in the current research. The coordinate sys-
tems of blades and vehicle, ordering schemes, and variational formulation
is described. The composite blade and quasi-steady aerodynamic models
are derived, and the free-wake model is also presented. Finally, an aeroe-
lastic analysis that includes nite element formulations and vehicle trim is
discussed.
3. The trailing-edge ap formulation and optimal controller for both vibration
and blade loads are described in Chapter 3. The inertial loads of trailing-edge
aps are derived based on the previous work [69]. Followed by the discussion
of available aerodynamic models to predict the aerodynamic loads generated
by ap motions. Finally, the optimal controller is described based on the
minimization of an objective function.
4. Chapter 4 investigates the feasibility of multiple trailing-edge aps for si-
multaneous reductions of vibration and blade loads. The concept involves
straightening the blade by introducing dual trailing edge aps in a conven-
tional articulated rotor blade. Numerical simulation is performed for the
steady-state forward ight of an advance ratio of 0.35.
5. In Chapter 5, vibration reduction using the multiple trailing-edge ap con-
guration is investigated. The concept involves deecting each individual
trailing-edge ap using a compact resonant actuation system that is described
in Chapter 6. Each resonant actuation system could yield high authority,
while operating at a single frequency. Numerical simulation is performed for
the steady-state forward ight ( = 0.15 0.35) condition.
37
6. Chapter 6 describes the development of the resonant actuation system, which
is utilized for the blade loads and vibration controls presented in Chapter 4
and 5. A piezoelectric actuation system model is derived for active ap rotors,
and then followed by mechanical tuning and electrical tailoring, where the
optimal tuning parameters for electric networks are explicitly determined.
An equivalent electric circuit model emulating the physical actuation system
is derived and experimentally tested.
7. Chapter 7 extends the new resonant actuation system concept to provide de-
sign guidelines and better physical understandings. Dynamic characteristics
of the RAS is examined for the case of forward ights. Vibration reduction
performance of various ap congurations is evaluated within the available
actuation authority. Bench top tests are conducted by utilizing the designed
controller with electric network dynamics and phase compensation. Finally,
the power requirement of the actuation system is characterized.
8. Finally, the research eorts and achievements in this thesis are summarized
in the last chapter. Recommendations for future work are also discussed.
Chapter 2
Helicopter Model
One of the objectives of current study is to investigate the feasibility of multiple
trailing-edge aps for vibration reduction. It is required to have a fully coupled
ap-lag-torsion model that can be used for an analysis of both hingeless and fully
articulated hubs. The formulation presented in this chapter is adapted from several
sources [76,133,134] and is included here for completeness. Much of the theory has
been re-derived for the current work, and hence some of the equations are slightly
dierent from the cited sources.
This chapter describes the physical and mathematical models used in the
present investigation. In the rst section, the coordinate systems of blade and
vehicle, ordering schemes, and variational formulation will be described as the
preliminary background. In following two sections, the composite blade and quasi-
steady aerodynamic models will be derived. Additional aerodynamic loads gener-
ated by trailing-edge aps will be described in the following chapter. In the fourth
section, main rotor inow and free wake models will be presented. Finally, an
aeroelastic analysis that includes nite element formulations and vehicle trim will
be discussed.
39
2.1 Preliminary Background
In order to understand the following derivations, the preliminary backgrounds, such
as the coordinate systems of vehicle and blade, ordering schemes and variational
formulation, will be described. This will be carried throughout the rest of analysis.
2.1.1 Vehicle Kinematics and Coordinate Systems
Figure 2.1 shows the coordinate systems for describing the motion of the helicopter
system. Rigid body motion of the vehicle is dened relative to a ground-xed iner-
tial coordinate system, superscript I. The vehicle coordinate system, superscript
F, are xed to the vehicle center of gravity. The hub-xed coordinate system,
superscript H, is dened parallel to the vehicle coordinate system.
F
x
F
y
F
z
H
z
H
x
H
y
I
x
I
y
I
z
s
a
s
f
H
x
H
y
R
x
R
y
y

V
C.G.
Figure 2.1. Vehicle and rotating coordinate systems
40
Position vectors based on each coordinate systems are dened as:
r = x
I

i
I
+y
I

j
I
+z
I

k
I
= x
I
1
e
I
1
+x
I
2
e
I
2
+x
I
3
e
I
3
= x
I
i
e
I
i
= x
H
i
e
H
i
= x
F
i
e
F
i
(2.1)
where the boldface denotes the vector. For convenience, the tensorial formulation
is adopted.
The transformation between the inertial system and the hub-xed nonrotating
systems is dened as
_

_
e
H
1
e
H
2
e
H
3
_

_
=
_

_
1 0
s
0 1
s

s

s
1
_

_
_

_
e
I
1
e
I
2
e
I
3
_

_
, e
H
i
= T
HI
ij
e
I
j
, (2.2)
in which
s
is the longitudinal shaft tilt angle (positive nose down), and
s
is the
lateral shaft tilt angle (positive advancing side down). Their amounts are assumed
to be small.
The hub-rotating coordinate system, superscript R, is rotating at constant
angular velocity, = e
H
3
, with respect to the hub-xed nonrotating frame (see
Figure 2.1). The transformation between hub-xed and rotating systems can be
obtained as
e
R
i
= T
RH
ij
e
H
j
, T
RH
=
_

_
cos sin 0
sin cos 0
0 0 1
_

_
, (2.3)
where is the azimuth angle and equals t.
The undeformed blade coordinate system, superscript u, are attached to the
undeformed blade. The undeformed blade is at a precone angle of
p
as shown
in Figure 2.2. The coordinate transformation between the rotating coordinate
system, superscript R, and the undeformed blade coordinates system, superscript
u, is given by
e
u
i
= T
uR
ij
e
R
j
, T
uR
=
_

_
cos
p
0 sin
p
0 1 0
sin
p
0 cos
p
_

_
. (2.4)
41
H
x
H
y
H
z
R
x
R
y
x
z
p
b
y
y
p
b
Figure 2.2. Undeformed blade coordinate systems
Letting T
uH
ij
= T
uR
ik
T
RH
kj
, the transformation between the hub-xed nonrotating
system and the undeformed blade system is written as
e
u
i
= T
uH
ij
e
H
j
, T
uH
=
_

_
cos
p
cos cos
p
sin sin
p
sin cos 0
sin
p
cos sin b
p
sin cos
p
_

_
. (2.5)
2.1.2 Blade Deformed Kinematics and Coordinate Systems
Accurate modeling of elastic rotor blades requires that deections are considered
to be moderately large. Although deections are assumed to be moderately large,
strains are still assumed to be small. These assumption are so called deections
with small strains and moderately large rotation. In view of solid mechanics, it is
similar to von Karman partial nonlinearity.
The strain-displacement relationships are developed based on the deformed
conguration (Eulerian formulation) so that the same reference is used for both
the stress and strain tensors. To second order which was dened by Reference [134],
however, the two results, Lagrangian based on the undeformed conguration and
Eulerian formulation, are equivalent. Details of this procedure and the strain
42
x
e
r
V
e
r
h
e
r
o
p
p
u
v
w
z
y
x
Figure 2.3. Deformed blade coordinate systems
measure are given in Appendix A.
The deformed blade is characterized by the deformed blade coordinate system,
superscript d. As shown in Figure 2.3, e
d
2
and e
d
3
are aligned with the principal
cross-section axis. A point P
o
on the undeformed elastic axis undergoes deections
u, v and w in the e
u
1
, e
u
2
and e
u
3
directions and moves to a point P as shown in
Figure 2.3. Then the blade cross section containing P undergoes a rotation
1
about the deformed elastic axis. The total blade pitch,
1
, is dened as

1
=
o
+

, (2.6)
where
o
is the rigid pitch angle due to control pitch and pretwist. In general,
the pretwist can be an arbitrary function of radial location, i.e.,
o
=
o
(r). For a
blade with linear a pretwist, the rigid pitch angle can be expressed by

o
=
75
+ ( x .75)
tw
+
1c
cos +
1s
sin . (2.7)
Total blade pitch,
1
, also includes elastic twist

. The elastic twist is dened
as

=
_
x
0
w
x

2
v
x
2
dx. (2.8)
The is the elastic twist about the undeformed elastic axis, while the

can be
viewed as the elastic twist about the deformed elastic axis. This is a nonlinear
43
u
e z
3
,
r
v
u
e y
2
,
r
w
d
e
2
,
r
h
d
e
3
,
r
z
f q q + =
o 1
^
o
q
u
e y
2
,
r
u
e z
3
,
r
h
V
Undeformed Cross Secti on
Deformed Cross Secti on
V
h A.C. C.G. E.A.
d
e
g
e
Figure 2.4. Cross-section coordinates before and after deformation
kinematic eect arising from moderately large rotation [134], see Figure 2.4.
The coordinate transformation between the undeformed and deformed blade
coordinate systems is given by
e
d
i
= T
du
ij
e
u
j
, (2.9)
where
T
du
=
_

_
cos
x
cos
y
cos
y
sin
z
sin
y
cos
z
sin
x
sin
y
cos
x
sin
z
cos
x
cos
y
sin
x
sin
y
sin
z
cos
y
sin
x
cos
x
cos
z
sin
y
+sin
x
sin
z
cos
z
sin
x
cos
x
sin
y
sin
z
cos
x
cos
y
_

_
. (2.10)
As shown Figure 2.5, the Euler angles,
x
,
y
, and
z
are dened in terms of blade
44
x
y
z
ds ds dx
dx
dw
dx
dv
- - =
2 2
) ( ) ( 1
ds
ds w'
ds v'
dx
dv
dx
dw
x
e
h
e
V
e
1
q
Figure 2.5. Deformations in terms of Euler angles
deformations as

x
=
1
,
y

dw
dx
= w

,
z

dv
dx
= v

, (2.11)
and the sequence of body xed rotation (w.r.t. the undeformed axis) is

z

y

x
. (2.12)
Substituting the above relations (Equation 2.11) and simplifying to second
order terms yield the transformation between deformed and undeformed blade
positions
T
du
=
_

_
1
v
2
2

w
2
2
v

cos
1
w

sin
1
(1
v
2
2
) cos
1
v

sin
1
(1
w
2
2
) sin
1
v

sin
1
w

cos
1
(1
v
2
2
) sin
1
v

cos
1
(1
w
2
2
) cos
1
_

_
. (2.13)
For small elastic twist angle, the trigonometric terms, cos
1
and sin
1
, can be
45
approximated by
cos
1
cos
o


sin
o
, sin
1
sin
o
+

cos
o
, (2.14)
By substituting the above equation into T
du
, we can decompose of it as following
constant, linear, and nonlinear matrices
T
du
(T
du
)
o
+ (T
du
)
L
+ (T
du
)
NL
: O(
2
), (2.15)
where
(T
du
)
o
=
_

_
1 0 0
0 cos
o
sin
o
0 sin
o
cos
o
_

_
, (2.16)
(T
du
)
L
=
_

_
0 v

cos
o
w

sin
o

sin
o

cos
o
v

sin
o
w

cos
o

cos
o

sin
o
_

_
, (2.17)
(T
du
)
NL
=
_

1
2
(v
2
+w
2
) 0 0

(v

sin
o
w

cos
o
)

1
2
v
2
cos
o
v

sin
o

1
2
w
2
sin
o

(v

cos
o
+w

sin
o
)
1
2
v
2
sin
o
v

cos
o

1
2
w
2
cos
o
_

_
. (2.18)
2.1.3 Nondimensionalization and Ordering Scheme
Many of the derivations contain complex expressions that are multiplied together.
In order to reduce the total number of terms in the formulation, and ordering
scheme is applied. This provides a method for systematically neglecting terms
based on their relative magnitude. The larger terms are kept, while the smaller
terms are neglected, or ordered out.
In developing the analysis, the physical quantities are nondimensionalized by
the reference parameters given in Table 2.1. Nondimensional quantities are only
used in the subsequent analysis.
46
Table 2.1. Nondimensionalized parameters
Physical Quantity Reference Parameter
Length R
Time 1/
Mass/Length m
o
Velocity R
Acceleration
2
R
Force m
o

2
R
2
Moment m
o

2
R
3
Energy or Work m
o

2
R
3
In formulating Hamiltons principle, it is important to neglect higher order
terms to simplify the analysis. Terms up to second order are retained in the analysis
by introducing the nondimensional quantity , such that 1. Some third order
terms related to elastic torsion are also retained in the energy expressions. The
order of various terms is given in Table 2.2. The Mathematica programs, which
are provided in Appendix C for reference, are used to derive the equations that
will be presented in Chapter 2 and to apply the ordering scheme.
Table 2.2. Order of terms used in aeroelastic analysis
Term List Order

C
ij
O(
2
)
x, h, x
cg
, y
cg
, m,

,

x
O(1)
, cos , sin ,
o
,
tw
,
75
,
1c
,
1s
, c
1
, d
2
O(1)
v, w,

,
p
, , , n,
s
,
s
O()
, c
o
, d
1
, f
o
O()
u, d
o
, f
1
,
T
O(
2
)
e
d
, e
g
O(
3/2
)
Special notice is taken to the torsion terms k
2
m
1
and k
2
m
1
and the wall thickness
of blade n. These terms normally would be order
2
, but are treated as being
in this analysis. The reason is that in the torsion equations, most of the terms
are small compared to the ap and lag bending terms. The wall thickness n is
retained to capture the eect of wall thickness. Therefore, k
2
m
1
, k
2
m
2
and n have
47
been treated as O() and O(), respectively.
The denition of the reference mass per unit length, m
o
, requires special con-
sideration. It is dened as the mass per unit length of a uniform blade with the
same ap inertia as the blade being considered. Thus, the reference mass per unit
length of a non-uniform blade is calculated by
m
o
=
3I
b
R
3
=
3
_
R
0
mr
2
dr
R
3
. (2.19)
Also, time derivatives are slightly dierent when non-dimensionalized. They are
related by = t, so derivatives are given by
= t, (2.20)
d()
dt
=
d()
d
d
dt
=
d()
d
= ()

, (2.21)
d
2
()
dt
2
=
d
2
()
d
2
d
2

dt
2
=
2
d
2
()
d
2
= ()

. (2.22)
2.1.4 Variational Formulation
Hamiltons variational principle is used to derive the system equations of motion.
For a conservative system, Hamiltons principle states that the true motion of
a system, between prescribed initial conditions at time t
1
and nal conditions
ate time t
2
, is that particular motion for which the time integral of the dierence
between the potential and kinetic energies is a minimum. For an aeroelastic system,
e.g., the rotor, there are nonconservative forces which are not derivable from a
potential function. The generalized or extended Hamiltons principle, applicable to
nonconservative systems, is expressed as
=
_
t
2
t
1
(U T W) dt = 0, (2.23)
where U is the rst variation of strain energy, T is the rst variation of kinetic
energy, and W is the virtual work done by external forces.
48
The motion of rotor blade results in the elastic deformation and the acceleration
of blade. Thus, the main source of the strain and kinetic energies is a main rotor
blade. External forces are caused by aerodynamic forces and moments. These vari-
ations have contributions from the rotor, the trailing edge ap, and the actuators.
The contributions from each terms can be expressed as the sum of contributions
from each blade. The variations can be written as
U =
N
b

b=1
(U
r
+U
a
) , (2.24)
T =
N
b

b=1
(T
r
+T
f
+T
a
) , (2.25)
W =
N
b

b=1
(W
r
+W
f
+W
a
) , (2.26)
where the subscript r denotes contribution from the rotor, the subscript f de-
notes contribution from the trailing edge ap (TEF), and the subscript a denotes
contribution from the actuator for TEF.
The main rotors structural and aerodynamic contributions are described in
following two sections, Section 2.2 and Section 2.3, respectively. The trailing-edge
aps inertial and aerodynamic contributions will be described in Chapter 3. In
this study, the contribution from actuator is neglected in the analysis of helicopter
dynamics. Coupled actuator and trailing-edge ap dynamics will be discussed in
Chapter 6.
49
2.2 Structural Model
The strain and kinetic energies of rotor blades are formulated by deriving the
integral equations in the blade deformed coordinate and undeformed coordinates
y and z. They are discretized for use with the nite element method. The strain
measure used in the strain energy section and framework in the kinetic energy
section are given in Appendix A.
2.2.1 Strain Energy of Rotor Blade
Assuming that each rotor blade is a long slender anisotropic laminated composite
beam, the constitutive relationships between stress and strain in an orthotropic
ply are given by
_

11

22

33

23

13

12
_

_
(k)
=
_

_
Q
11
Q
12
Q
13
0 0 0
Q
12
Q
22
Q
23
0 0 0
Q
13
Q
23
Q
33
0 0 0
0 0 0 Q
44
0 0
0 0 0 0 Q
55
0
0 0 0 0 0 Q
66
_

_
(k)
_

11

22

33

23

13

12
_

_
(k)
, (2.27)
where 1, 2, 3 denotes the material coordinate system. The subscript (k) denotes
the k th lamina.
Let us assume a cylindrical geometry that has a curvilinear coordinate system
(, s, n). The constitutive equation (Equation 2.27) can be transformed into a
physical coordinate system as follows:
_

ss

nn

sn

s
_

_
(k)
=
_

Q
11

Q
12

Q
13
0 0

Q
16

Q
12

Q
22

Q
23
0 0

Q
26

Q
13

Q
23

Q
33
0 0

Q
36
0 0 0

Q
44

Q
45
0
0 0 0

Q
45

Q
55
0

Q
16

Q
26

Q
36
0 0

Q
66
_

_
(k)
_

ss

nn

sn

s
_

_
(k)
, (2.28)
where

Q
ij
denotes the transformed stiness. These transformation relations can
50
be found in Reference [140].
For a long slender geometry, the uniaxial stress assumption (
nn
=
ss
=
sn
=
0) is valid. In addition, it can be assumed that the transverse shear strain
n
is
small, because the warping of thin wall due to exure can be neglected. These
assumption allow us to use the classical lamination theory [140].
By applying these assumption to Equation 2.28, the corresponding relationship
between stresses and strains can be obtained by
_

s
_
(k)
=
_

C
11

C
16

C
16

C
66
_
(k)
_

s
_
(k)
, (2.29)
where

C
ij
are dened as

C
11
=

Q
11
+

Q
2
13

Q
22
2

Q
12

Q
13

Q
23
+

Q
2
12

Q
33

Q
2
23


Q
22

Q
33
, (2.30)

C
16
=

Q
16
+

Q
13

Q
23

Q
26
+

Q
12

Q
26

Q
33
+

Q
13

Q
22

Q
36


Q
12

Q
23

Q
36

Q
2
23


Q
22

Q
33
, (2.31)

C
66
=

Q
66
+

Q
2
26

Q
33
2

Q
23

Q
26

Q
36
+

Q
22

Q
2
36

Q
2
23


Q
22

Q
33
. (2.32)
Substituting Equations A.33 and A.34 into Equation 2.29 and integrating with
respect to n, which is a thickness direction of the thin wall, give the following
relationship between resultant forces and strains of thin wall
_

_
N

N
s
M

M
s
_

_
=
_

_
A
11
A
16
B
11
B
16
A
16
A
66
B
16
B
66
B
11
B
16
D
11
D
16
B
16
B
66
D
16
D
66
_

_
_

o
s
k

k
s
_

_
, (2.33)
where
(A
ij
, B
ij
, D
ij
) =
_
n/2
n/2

C
(k)
ij
(1, n, n
2
) dn. (2.34)
51
Then the rst variation of strain energy of the blade can be expressed by
U
r
=
_
R
0
_
s
_
N

+N
s

o
s
+M

+M
s
k
s
_
ds dx, (2.35)
where stress resultant N
ij
, moment resultant M
ij
, axial strain
o

, and in-plane
shear strain
o
s
can be decomposed into linear and nonlinear terms
N

= N
L

+N
NL

, M

= M
L

+M
NL

, (2.36)
N
s
= N
L
s
+N
NL
s
, M
s
= M
L
s
+M
NL
s
. (2.37)
The strain energy can be also decomposed as
U
r
= U
L
r
+U
NL
r
, (2.38)
where
U
L
r
=
_
R
0
_
s
_
N
L

+N
L
s

L
s
+M
L

+M
L
s
k
s
_
ds dx, (2.39)
U
NL
r
=
_
R
0
_
s
_
N
NL

+N
NL
s

L
s
+M
NL

+M
NL
s
k
s
_
ds dx
+
_
R
0
_
s
_
N

NL

+N
s

NL
s
_
ds dx. (2.40)
The rst variation of linear strains and curvatures (Equations A.33 and A.34)
is given by

= u

e
y
o
v

z
o
w

, (2.41)

L
s
=
t
s

, (2.42)
k

= cos

w

sin

v

+q

, (2.43)
k
s
= 2

. (2.44)
Substituting the above equations into Equation 2.39 gives another form of linear
strain energy. The variation of linear strain energy of composite beam can be
52
rewritten by
U
L
r
=
_
R
0
_
N
x
u

e
M
y
w

+M
z
v

+T

+T

_
dx, (2.45)
where
N
x
=
_
N
L

ds, (2.46)
M
y
=
_
_
z
o
N
L

+M
L

cos

_
ds, (2.47)
M
z
=
_
_
y
o
N
L

M
L

sin

_
ds, (2.48)
T

=
_
_
q M
L

T
N
L

_
ds, (2.49)
T

=
_
_
2 M
L
s
+N
L
s

t
s
_
ds, (2.50)
in which the direction of moments is based on the undeformed coordinate system
e
u
i
and the direction of torque and force is based on the deformed coordinate system
e
d
i
. N
x
denotes the axial force resultant, M
y
is the apwise bending moment and
M
z
represents the lagwise bending moment. T

is the torque due to warping, which


is so called Valsov torsion, and T

is the St. Venant torque.


These force, moment and torque can be expressed in terms of degrees of free-
dom, such as u
e
, v, w and

. The constitutive equation of composite beam for
linear strain can be written as follows:
_

_
N
x
M
y
M
z
T

_
=
_

K
11

K
12

K
13

K
14

K
15

K
12

K
22

K
23

K
24

K
25

K
13

K
23

K
33

K
34

K
35

K
14

K
24

K
34

K
44

K
45

K
15

K
25

K
35

K
45

K
55
_

_
_

_
u

e
w

_
, (2.51)
where

K
ij
are dened as

K
12
= K
12
cos
o
+K
13
sin
o
, (2.52)

K
13
= K
13
cos
o
K
12
sin
o
, (2.53)
53

K
22
= K
22
cos
2

o
+K
33
sin
2

o
+K
23
sin 2
o
, (2.54)

K
23
= (K
33
K
22
) sin
o
cos
o
+K
23
cos 2
o
, (2.55)

K
24
= K
24
cos
o
+K
34
sin
o
, (2.56)

K
25
= K
25
cos
o
+K
35
cos
o
, (2.57)

K
33
= K
33
cos
2

o
+K
22
sin
2

o
K
23
sin 2
o
, (2.58)

K
34
= K
34
cos
o
K
24
sin
o
, (2.59)

K
35
= K
35
cos
o
K
25
sin
o
. (2.60)
The other

K
ij
terms are dened to equal to K
ij
which denotes

K
ij
of which the
rigid pitch angle
o
is set to zero, i.e.,
K
ij
=

K
ij
[
o0
. (2.61)
The K
ij
are dened in Appendix B.1.1. The stiness coecients K
11
, K
22
, K
33
,
K
44
, and K
55
are related to in-plane stiness EA, ap bending rigidity EI
y
, lag
bending rigidity EI
z
, warping rigidity EI

, and torsional rigidity GJ in the


isotropic beam, respectively. The coupling terms K
25
and K
35
denote the ap-
torsion coupling and the lag-torsion coupling stiness, respectively. The coupling
term K
13
is also related to EA
e
A
of the isotropic beam. Denite of EA, GJ, and
EA
e
A
etc. can be found in Reference [76].
It is assumed that the other coupling terms except K
13
, K
15
K
25
, and K
35
can be vanished or neglected, because the purpose of present study is to reduce
the helicopter rotor vibration and blade loads using the tension-torsion, the ap-
torsion, or the lag-torsion coupling. The terms related to warping,
_
A
11
ds and
_
A
11
ds, are therefore discarded. For such a conguration, the stiness matrices
can be rewritten, in terms of shape function, as follows:
[K
uu
] =
_
K
11
H
T
u

H
u

dx, (2.62)
[K
uw
] =
_
K
13
sin
o
H
T
u

H
w

dx, (2.63)
[K
uv
] =
_
K
13
cos
o
H
T
u

H
v

dx, (2.64)
54
[K
up
] =
_
K
15
H
T
u

H
p

dx, (2.65)
[K
ww
] =
_
(K
22
cos
2

o
+K
33
sin
2

o
) H
T
w

H
w

dx, (2.66)
[K
wv
] =
_
(K
33
K
22
) cos
o
sin
o
H
T
w

H
v

dx, (2.67)
[K
wp
] =
_
(K
25
cos
o
+K
35
sin
o
) H
T
w

H
p

dx, (2.68)
[K
vv
] =
_
(K
33
cos
2

o
+K
22
sin
2

o
) H
T
v

H
v

dx, (2.69)
[K
vp
] =
_
(K
35
cos
o
K
25
sin
o
) H
T
v

H
p

dx, (2.70)
[K
pp
] =
_
(K
44
H
T
p

H
p

+K
55
H
T
p

H
p

) dx. (2.71)
The rst variations of nonlinear strains (Equations A.33 and A.34) are given
by

NL

= (
2
+
2
)

o
_

+w

+v

_
+ (
2
+
2
)
_

+w

+w

+v

_
+ (z
o
v

y
o
w

(z
o
v

y
o
w

), (2.72)

NL
s
=
t
s
v

+
t
s
w

. (2.73)
Substituting Equations 2.72 and 2.73 into Equation 2.40 yields another form of a
strain energy. Then the variation of nonlinear strain energy of composite beam is
rewritten by
U
NL
r
=
_
R
0
_
N
NL
x
u

e
M
NL
y
w

+M
NL
z
v

+T
NL

+T
NL

+ N
NL
w
w

+N
NL

_
dx, (2.74)
where
N
NL
x
=
_
N
NL

ds, (2.75)
55
M
NL
y
=
_
_
z
o
N
NL

+M
NL

cos

y
o
N

_
ds, (2.76)
M
NL
z
=
_
_
y
o
N
NL

M
NL

sin

+ N

_
z
o

+ (
2
+
2
)(w

o
+w

+w
2
v

)
_
+N
s

t
s
w

_
ds, (2.77)
T
NL

=
_
_
q M
NL


T
N
NL

_
ds, (2.78)
T
NL

=
_
_
2 M
NL
s
+N
NL
s

t
s
+N

(
2
+
2
)(

o
+w

)
_
ds, (2.79)
N
NL
w
=
_
_
N

(
2
+
2
)(

o
v

+w

v
2
) + N
s

t
s
v

_
ds, (2.80)
N
NL

=
_
N

(y
o
w

+z
o
v

) ds. (2.81)
Note that some linear terms are appeared in Equations 2.75 2.81, because they
are treated as the nonlinear terms (see Appendix A.1).
Nonlinear strain energy terms are complicate, so the ordering scheme up to
order O(
2
) based on Table 2.2 is applied to reduce the complexity. The laminate
stiness B
ij
and D
ij
have been discarded, because the warping through the wall
thickness is very small, when compared to the beam cross-section warping. Then
the following reduced nonlinear terms are obtained.
N
NL
x
= Ak
2
11
(
1
2

2
+

o
v

) + Ak
2
11

o
+

K
13
(w

cos
o
v

sin
o
) +

K
15
v

, (2.82)
M
NL
y
= (

K
33


K
22
)(v

cos 2
o
+w

sin 2
o
) +

K
13
cos
o
u

+

K
25
(v

cos
o

sin
o
) +

K
35
(v

sin
o
+

cos
o
)
+ (AB
1
sin
o
+AB
2
cos
o
)

o
, (2.83)
M
NL
z
= (

K
33


K
22
)(w

cos 2
o
v

sin 2
o
)

K
13
sin
o
u

+

K
55
w

+Ak
2
11

o
u

e
w

+

K
15
u

e
w

+

K
25
_
(w

) cos
o
2w

sin
o
_
+

K
35
_
(w

) sin
o
+ 2w

cos
o
_
+ (AB
1
cos
o
AB
2
sin
o
)

o
, (2.84)
T
NL

=

K
44
w

, (2.85)
56
T
NL

=

K
35
(w

cos
o
v

sin
o
)

K
25
(w

sin
o
+v

cos
o
)
+

K
55
w

+Ak
2
11
u

+Ak
2
11
u

o
, (2.86)
N
NL
w
=

K
25
(v

cos
o
v
2
sin
o
) +

K
35
(v

sin
o
+v
2
cos
o
)
+

K
55
v

+Ak
2
11

o
u

e
v

+

K
15
u

e
v

, (2.87)
N
NL

= (

K
33


K
22
)
_
v

cos 2
o
+ (w
2
v
2
) sin
o
cos
o
_
+

K
13
(w

e
cos
o
v

e
sin
o
)
+

K
35
(w

cos
o
v

sin
o
)

K
25
(w

sin
o
+v

cos
o
),(2.88)
where
Ak
2
11
=
_
A
11
(
2
+
2
) ds, (2.89)
AB
1
=
_
A
11
(
2
+
2
) ds, (2.90)
AB
2
=
_
A
11
(
2
+
2
) ds, (2.91)
the terms

K
ij
are dened as K
ij
of which the stretching bending coupling B
ij
and
bending rigidity D
ij
of thin wall are set to zero, i.e.,

K
ij
= K
ij
[
B
ij
,D
ij
0
, (2.92)
and the underline terms are linear terms.
The nonlinear terms except the underline terms are treated as the nonlinear
force vector. The stiness matrices due to the linear terms can be expressed as
[K
up
] =
_
Ak
2
11
H
T
u

p
dx, (2.93)
[K
wp
] =
_
(AB
1
sin
o
+AB
2
cos
o
)

o
H
T
w

p
dx, (2.94)
[K
vp
] =
_
(AB
1
cos
o
AB
2
sin
o
)

o
H
T
v

p
dx, (2.95)
and the nonlinear forces in terms of shape functions are given by
F
u
=
_
N
NL
x
H
T
u

dx, (2.96)
57
F
w
=
_
_
M
NL
y
H
T
w

+N
NL
w
H
T
w

_
dx, (2.97)
F
v
=
_
M
NL
z
H
T
v

dx, (2.98)
F
p
=
_
_
T
NL

H
T
p

+T
NL

H
T
p

+N
N

H
T
p
_
dx, (2.99)
where the linear terms are discarded. Total stiness matrices of strain energy and
explicit force vectors are given in Appendix B.1.2.
2.2.2 Kinetic Energy of Rotor Blade
The kinetic energy of the rotor blade, T
r
, depends on the blade velocity. This
velocity is generally due to: (1) blade motion relative to the hub, as well as (2) the
motion of the hub itself. This relationship is expressed mathematically as

V =

V
b
+

V
f
, (2.100)
where

V
b
is the velocity of the blade relative to the hub and

V
f
is the velocity (at
the blade) induced by the motion of the fuselage. In the present analysis, the hub
is assumed to be rigidly attached to the fuselage. The velocity due to the fuselage
motion has been neglected in this study.
The position vector, r, of an arbitrary point on the blade after deformation is
given, in the mixed coordinate system, as
r = (s +u
e

)e
d
1
+e
d
2
+e
d
3
,
= (x +u
T

)e
u
1
+ve
u
2
+we
u
3
+e
d
2
+e
d
3
, (2.101)
where the rst equations are based on the deformed coordinate system and the
second equations are based on the undeformed coordinate system. Note that the
transformation relationships between the undeformed and deformed coordinate
system is valid in sense of the innitesimal quantities.
The innitesimal length along the deformed axis, ds, may be equal to the
58
innitesimal length along the undeformed axis, dx. The dierence between two
incremental length can be expressed by
(dx +du) =
_
1 (v
2
+w
2
)(ds +du
e
),

_
1
1
2
(v
2
+w
2
)
_
(ds +du
e
), (2.102)
and
du du
e

1
2
(v
2
+w
2
)dx, (2.103)
u = u
e

1
2
_
x
0
(v
2
+w
2
)dx, (2.104)
where u
e
is very small, so we can usually neglect this term. This relationships
can be explained in that the innitesimal displacement, du, shall be shrunk as the
beam rotates rigidly.
Assuming that the beam is inextensible, the length of beam along the deformed
axis, s, in terms of x is given by
se
d
1
=
_
x
1
2
_
x
0
(v
2
+w
2
)dx
_
e
u
1
+ve
u
2
+we
u
3
. (2.105)
Referring to Equation 2.14, the relationship between the deformed and undeformed
coordinate system is given by
e
d
2
(v

cos
1
+w

sin
1
) e
u
1
+ cos
1
e
u
2
+ sin
1
e
u
3
, (2.106)
e
d
3
(v

sin
1
w

cos
1
) e
u
1
sin
1
e
u
2
+ cos
1
e
u
3
. (2.107)
By substituting Equations 2.106 and 2.107 into Equation 2.101, the deformed
position vector can be rewritten, in the undeformed coordinate system, as follows:
r = x
1
e
u
1
+y
1
e
u
2
+z
2
e
u
3
, (2.108)
where
x
1
= x +u
e

1
2
_
x
0
(v
2
+w
2
)dx
T

y w

z, (2.109)
y
1
= v + y, z
1
= w + z, (2.110)
59
y = cos
1
sin
1
, z = sin
1
+ cos
1
. (2.111)
Then the velocity of an arbitrary point on the blade can be calculated as

V
b
=
r
t
+ r , = e
H
3
(
p
e
u
1
+e
u
3
), (2.112)
and the velocity components are given by

V
b
= V
bx
e
u
1
+V
by
e
u
2
+V
bz
e
u
3
, (2.113)
where
V
bx
= x
1
y
1
, (2.114)
V
by
= y
1
+ x
1
z
1

p
, (2.115)
V
bz
= z
1
+ y
1

p
. (2.116)
The derivatives of the arbitrary point on the blade with respect to time are
given as
x
1
= u
e

_
x
0
(v

+w

)dx
T

( v

+w

1
) y ( w

1
) z, (2.117)
y
1
= v

1
z, (2.118)
z
1
= w +

1
y, (2.119)
and the kinetic energy variation can be expressed by
T
r
=
_
R
0
_
A

V dAdx. (2.120)
Substituting Equations 2.114 to 2.116 into Equation 2.120 and integrating by parts
yield


T
r
=
T
r
m
o

2
R
3
=
_
1
0
_
A

s
(T
x1
x
1
+T
y1
y
1
+T
z1
z
1
) dAdx, (2.121)
60
where
s
is the material density of structure and
T
x1
= x

1
+ 2y

1
+x
1
z
1

p
, (2.122)
T
y1
= y

1
2x

1
+y
1
+ 2z

p
, (2.123)
T
z1
= z

1
2y

p
x
1

p
+z
1

2
p
, (2.124)
in which
x

1
= u

e

_
x
0
_
v

+v

+w

+w

_
dx
T

_
v

+w

1
+ 2

1
w

_
y
_
w

1
2

1
v

_
z, (2.125)
y

1
= v

1
z

1
2
y, (2.126)
z

1
= w

1
y

1
2
z. (2.127)
The variations of the deformed position vector components, x
1
, y
1
and z
1
, are
given by
x
1
= u
e

yv

y zw

_
x
0
(v

+w

)dx, (2.128)
y
1
= v + y, (2.129)
z
1
= w + z, (2.130)
where
y = z

, z = y

. (2.131)
Now the variation of kinetic energy (Equation 2.121) can be rewritten by


T
r
=
_
1
0
_
V
I
x
u +V
I
y
v +V
I
z
w +M
I
z
v

M
I
y
w

+M
I
x

_
dx, (2.132)
where
V
I
x
= m
_
x +u
e

1
2
_
x
0
(v
2
+w
2
)dx u

e
61
+
_
x
0
_
v

+v

+w

+w

_
dx
+ 2v

w
p
e
g
cos
o
_
v

+ 2

o
2
v

o
w

o
w

_
e
g
sin
o
_
2

1
+
p
+w

o
v

o
2
w

+ 2

o
v

_
, (2.133)
V
I
y
= m
_
v v

+ 2w

p
2u

e
+ 2
_
x
0
(v

+w

)dx
+ e
g
cos
o
_
1 + 2v

+ 2

o
+

o
2
+

+ 2

p
+ 2

o
w

_
+ e
g
sin
o
_

1


+ 2w

o
2

o
v

__
, (2.134)
V
I
z
= m
_
w

(x + 2v

)
p
e
g
cos
o
(

o
2

)
+ e
g
sin
o
_
2

o
2
+

+ 2

p
__
, (2.135)
M
I
z
= m
_
e
g
_
(x 2v

) cos
o
+x

sin
o
_
+ (k
2
m1
sin
2

o
+k
2
m2
cos
2

o
)
_
2

+v

o
2
v

o
w

o
w

_
(k
2
m2
sin
2

o
+k
2
m1
cos
2

o
)2

(2.136)
+ cos
o
sin
o
(k
2
m2
k
2
m1
)
_
w

+ 2

1
+
p
+

o
v

o
2
w

+ 2

o
v

_
,
M
I
y
= m
_
e
g
_
(x 2v

) sin
o
x

cos
o
_
+ (cos
2

o
k
2
m1
+ sin
2

o
k
2
m2
)
p
+ sin
o
cos
o
(k
2
m2
k
2
m1
)(v

+ 4

o
2
v

o
w

o
w

)
+ (k
2
m1
cos
2

o
+k
2
m2
sin
2

o
) (2.137)
_
w

+ 2

1
w

o
v

o
2
w

+ 2

o
v

_
,
M
I
x
= m
_
e
g
sin
o
(v +v

+xv

) e
g
cos
o
_
w

+x
p
+xw

+2v

p
_
+ sin
o
cos
o
(k
2
m1
k
2
m2
)(1 + 2v

+2

1
w

) (2.138)
k
2
m

1
+

(k
2
m1
k
2
m2
) cos 2
o
2w

(k
2
m1
cos
2

o
+k
2
m2
sin
2

o
)
_
,
where mk
2
m1
and mk
2
m2
represent the apwise and lagwise mass moments of inertia
per unit length, respectively, and are given by
m(e
g
, k
2
m1
, k
2
m2
) =
_
A

s
(,
2
,
2
)dA, (2.139)
62
and
mk
2
m
= mk
2
m1
+mk
2
m2
. (2.140)
By substituting Equation 2.104 into Equation 2.132, the variation of kinetic
energy can be also expressed by


T
r
=
_
1
0
_

V
I
x
u
e
+

V
I
y
v +

V
I
z
w +M
I
z
v

M
I
y
w

+M
I
x

_
dx, (2.141)
where

V
I
x
= V
I
x
+m
1
2
_
x
0
(v
2
+w
2
)d V
I
x
, (2.142)

V
I
y
= V
I
y
m(x + 2v

)v

+v

_
1
x
m( + 2v

)d (2.143)
V
I
y
mv 2
_
x
0
m v

d, (2.144)

V
I
z
= V
I
z
m(x + 2v

)w

+w

_
1
x
m( + 2v

)d (2.145)
V
I
z
mw 2
_
x
0
m w

d, (2.146)
in which the underline terms denote the curvature eect due to the geometric
nonlinearity. This eect may be neglected under assumption of a moderately large
rotation. A helicopter rotor system is under a strong centrifugal force, so that the
deformed geometric shape is almost a straight line except near root of the blade.
Here we developed two types of the variation of a kinetic energy. One is the
forces and moments based on the undeformed coordinate system, and the other
is based on the mixed coordinate system. In previous section, the variation of a
strain energy was derived in the deformed coordinate that corresponds to u
e
.
The variation of a kinetic energy given in Equation 2.141, therefore, should used
to be consistent with the strain energy formulation. Equation 2.132, however, may
63
be used to get the blade root loads with the force summation method, since the
hub loads are dened in the hub-xed coordinate system.
The matrix form of kinetic energy is implemented using the weak form of strain
energy. For example, the weak form of a double integral term can be expressed as
follows:
m(x + 2v

)
_
x
0
(v

+w

)d =
__
1
x
m( + 2v

)d
_
(v

+w

). (2.147)
The system matrices, such as stiness, damping and mass matrices, and force
vectors due to the kinetic energy of rotor blades are given in Appendix B.2.
64
2.3 Aerodynamic Blade Loads
The linear aerodynamic model is based on a quasi-steady strip theory. The blade
velocity is derived in the blade undeformed frame and then translated to the de-
formed frame to calculate the airloads. In the present analysis, motion of the
fuselage is not considered. The aerodynamics only take into account the wind
velocity, including the helicopter forward and climb speed and the rotor rotation,
and the motion of the blades relative to the hub.

V =

V
w
+

V
b
, (2.148)
where

V
w
is the wind velocity with contributions from the vehicle forward speed
and the rotor inow and

V
b
is the blade velocity relative to the hub xed frame
resulting from blade rotation and blade motions.
The blade velocity due to the helicopter forward speed and the rotor rotation is
expressed as

V
w
= R e
H
1
R e
H
3
, (2.149)
where is the rotor advance ratio, is the rotor non-dimensional inow, and is
the rotor angular velocity. The rotor inow ratio, , for small longitudinal shaft
tilt angles, consists of two components and is expressed by
= tan
s
+
i

s
+
i
. (2.150)
These velocities,

V
w
and

V
b
, must be transformed through the blade precone,

p
, and around the rotor azimuth, . This can be done through direction cosine
matrices about body xed axes, rst about the rotor azimuth, then through pre-
cone. From Equation 2.5, the transformation matrix between the hub coordinates,
e
H
i
, and the blade coordinates that are the undeformed coordinates, e
u
i
, is given
by
e
u
i
= T
uH
ij
e
H
j
. (2.151)
The wind velocity on the undeformed coordinate systems with a small angle
65
assumption for precone,
p
, is expressed as

V
w
= V
w
i
e
u
i
= (Rcos R
p
) e
u
1
Rsin e
u
2
(Rcos
p
+R) e
u
3
. (2.152)
The blade velocity,

V
b
, with respect to the rotating undeformed frame can be
written as

V
b
=
r
t
+ r = V
bi
e
u
i
. (2.153)
From Equations 2.112-2.119, the blade velocity components are given by

V
b1
= u
T

( v

+w

1
) y ( w

1
) z (v + y), (2.154)

V
b2
= v

1
z + (x +u
T

y w

z) (w + z)
p
, (2.155)

V
b3
= w +

1
y + (v + y)
p
, (2.156)
where
u = u
e

1
2
_
x
0
(v

2
+w

2
)dx , u = u
e

_
x
0
(v

+w

)dx, (2.157)

+w

,
1
=
o
+

, (2.158)
y = cos
1
sin
1
, z = sin
1
+ cos
1
. (2.159)
For quasi-steady aerodynamics, the rotor blade aerodynamic loads are calcu-
lated using a blade section strip analysis based on the angle of attack at the three-
quarter chord location. In the cross section, this point is at =
r
and = 0,
simplifying the blade velocity expressions. Then the total velocity,

V =

V
b


V
w
,
can be written as

V = V
i
e
u
i
, (2.160)
66
in which
V
x
= u
T

(v +
r
cos
1
) + R
p
Rcos

r
(

1
w

+ v

) cos
1

r
( w

1
v

) sin
1
, (2.161)
V
y
= v + (u +x) + sin
r

1
sin
1

p
(w +
r
sin
1
)

r
(v

cos
1
+w

sin
1
), (2.162)
V
z
= w +R +R
p
cos + (v +
r
cos
1
)
p
+
r

1
cos
1
. (2.163)
The blade section loads are calculated using the resultant velocity and aerody-
namic angle of attack in the deformed blade. Therefore, these velocities should be
transformed to the blade deformed frame. The transformation matrix is given in
Equation 2.13.
e
d
i
= T
du
ij
e
u
j
,

V = V
i
e
u
i
= V
i
T
du
ij
1
e
d
j
= U
j
e
d
j
. (2.164)
Thus the velocity components, U
j
, using the transformation matrix property,
T
du
ij
1
= T
du
ji
, are given by
V
i
T
du
ji
= T
du
ji
V
i
= U
j
. (2.165)
The velocities in the deformed frame are used to nd the blade angle of attack
and to determine the radial ow along the blade. While performing the transfor-
mation,
1
is replaced with
o
+

to modify cos
1
and sin
1
, and assuming

is a
small angle.
sin
1
= sin(
o
+

) sin
o
+

cos
o
, (2.166)
cos
1
= cos(
o
+

) cos
o


sin
o
. (2.167)
After transformation to the deformed frame using Equation 2.165, the resultant
blade velocity in the deformed rotating frame neglecting the warping term and the
higher order terms referring to Table 2.2 is expressed as
67
U
x
R
= u

v +(
p
+w

) +v

sin + +xv

+v

+w

+ cos
_
1 +
p
w

+
1
2
(v

2
+w

2
)
_

r
(1 +v

) cos
o
+
r
(

) sin
o
, (2.168)
U
y
R
= cos
o
_
u +x +v

+ ( +w

w
p
+vv

p
v

1
2
xv

2
+ sin (1
v

2
2
) + cos (
p

+w

+v

)
_
+ sin
o
_
(1
p
w

1
2
w

2
) + w

x(

+v

) v

+v
p
+vw

+ cos (
p

+w

) sin (

+v

)
_
, (2.169)
U
z
R
=
r
(

o
+
p
+w

)
+ sin
o
_
u x v

+(
p
v


) w

+w
p
vv

+
1
2
xv

2
+ sin (1 +
v

2
2
) cos (
p

+w

+v

)
_
+ cos
o
_
+w

+v
p
+vw

(
p
w

+
w

2
2
) xv

+ cos (
p
v

+w

) sin (

)
_
, (2.170)
where U
x
, U
y
, and U
z
denote the radial, tangential, and perpendicular velocities,
respectively. These velocities are also referred to as U
R
, U
T
, and U
P
. The displace-
ments, u, v, w, and the radial position, x, are the normalized quantities with the
blade length, R. The superscript, ()

, represents the derivative with respect to the


azimuth .
68
2.3.1 Quasi-steady Airloads
In this section, the blade loads are modeled as quasi-steady. This means that
although the loads change with time, at a particular instant, they are assumed
to be only a function of conditions at that instant. In other words, for a certain
section of the blade at a certain azimuth, the circulatory lift is only dependent
upon the angle of attack of that section at that time.
The blade circulatory airloads per unit length in the rotating deformed frame,
e
d
i
, can be written as
L
C
=
1
2
U
2
cC
l
, (2.171)
D
C
=
1
2
U
2
cC
d
, (2.172)
M
C
=
1
2
U
2
c
2
C
m
, (2.173)
where U is the incident velocity and C
l
, C
d
, C
m
are the section lift, drag, and
moment coecients, respectively. c is the chord of blade section. The aerodynamic
coecients are expressed, in terms of the angle of attack, as
C
l
= c
o
+c
l
, (2.174)
C
d
= d
o
+d
1
[[ +d
2

2
, (2.175)
C
m
= f
o
+f
1
= c
mac
+f
1
, (2.176)
in which c
o
is the lift coecient when the airfoil is at a zero angle of attack, c
1
is the
lift curve slope, often given by C
l
or C
n
, and d
o
is the prole drag of the blades
due to viscous eects and the nonlinear term, d
2
, causes a drag rise with large
angles of attack. f
o
or c
mac
is the zero angle pitching moment coecient about
aerodynamic center and f
1
is the slope of the moment curve. These relations are
restricted to the incompressible attached ow conditions.
Compressibility eects are accounted by modifying the lift curve slope, c
l
, as
c
l
=
c
l
M=

, =

1 M
2
, (2.177)
69
where is the Prandtl-Glauert factor. This correction factor works quite well
at low Mach numbers, M 0.85, characteristic of a rotor blade. For the most
part of rotor disk, the rotor blade sees normal ow conditions. In other areas,
where the angle of attack is extremely high or the ow is reversed, the dynamic
pressure is low, so there is not much eect on the overall response. Near this
region, particularly near the boundary forward ow and reverse ow, where the
tangential velocity U
T
is near zero, the small angle approximation breaks down.
The detail description of reverse ow model can be found in References [6, 76, 133].
The axial force, chord force, normal force, and pitching moment about the
elastic axis are given by
L
C
u
= D
C
sin , (2.178)
L
C
v
= L
C
sin D
C
cos , (2.179)
L
C
w
= L
C
cos +D
C
sin , (2.180)
M
C

= M
C
e
d
L
C
w
, (2.181)
where L
C
i
are the external loads along the deformed frame, e
d
i
, M
C

is the pitching
moment about the deformed elastic axis, is the axial skew angle due to the radial
component of incident velocity, U
R
, acting on the blade, and e
d
is the chordwise
oset of the aerodynamic center behind the elastic axis.
The equations are non-dimensionalized by the Lock number, , the blade iner-
tia, I
b
, and the standard non-dimensionalization factor and R. The Lock number
and uniform blade inertia are given by
=
acR
4
I
b
, I
b
=
m
o
R
3
3
, (2.182)
where a is the lift curve slope, C
l

. All forces, moments, and velocities are non-


dimensionalized using the scheme described in Table 2.1. The section lift and drag
forces, L
C
and D
C
, are non-dimensionalized as follows:
1
2
cU
2
=
1
2
_
R
2
m
o
_
_
c
R
_
_
U
2

2
R
2
_
=

6a
_
U
R
_
2
=

6a
U
2
. (2.183)
70
The airloads, Equations 2.178-2.181, are nondimensionalized by Equation 2.183.
Then,
L
C
u
=
U
2
6a
(C
d
sin ) , (2.184)
L
C
v
=
U
2
6a
(C
l
sin C
d
cos ) , (2.185)
L
C
w
=
U
2
6a
(C
l
cos +C
d
sin ) , (2.186)
M
C

=
U
2
6a
_
c
R
C
m
e
d
L
C
w
_
. (2.187)
Substituting the expression for C
l
,C
d
, and C
m
from Equations 2.1742.176, and
using the following approximations
sin , cos 1 , U U
T
,
U
P
U
T
, sin
U
R
U
T
, (2.188)
yield
L
C
u
=

6a
_
d
o
U
R
U
T
_
, (2.189)
L
C
v
=

6a
_
d
o
U
2
T
(c
o
U
P
d
1
[U
P
[)U
T
+ (c
1
d
2
)U
2
P
_
, (2.190)
L
C
w
=

6a
_
c
o
U
2
T
(c
1
+d
o
)U
T
U
P
+d
1
[U
P
[U
P
_
, (2.191)
M
C

=

6a
_
c
R
_
f
o
(U
2
T
+U
2
P
) f
1
U
P
U
T
_
e
d
L
C
w
_
. (2.192)
The nondimensionalized aerodynamic forces and pitching moment acting on a
blade section can be expressed by the following vector form.
L
C
= L
C
i
e
d
i
, M
C
= M
C

e
d
1
. (2.193)
It is consistent with the velocity in the deformed frame,

V = U
i
e
d
i
. The aerody-
namic forces in the undeformed frame, e
u
i
, are obtained by using the orthogonal
coordinate transformation. The transformation matrix between the deformed and
undeformed coordinate systems, e
d
i
= T
du
ij
e
u
j
, has already developed in Equation
71
2.13 on Page 44. The aerodynamics forces in the undeformed frame are given by
L
C
= L
C
i
e
d
i
= L
C
i
T
du
ij
e
u
j
= L
C
i
e
u
i
, M
C
= M
C

e
d
1
M
C

e
u
1
, (2.194)
L
C
i
= L
C
j
T
du
ji
= T
du
ij
L
C
j
, M
C

M
C

, (2.195)
where the pitching moment, M
C

, is assumed small, so that higher order terms can


be neglected as follows:
M
C

e
d
1
= M
C

(1
v

2
2

w

2
2
)e
u
1
+M
C

e
u
2
+M
C

e
u
3
,
M
C

e
u
1
. (2.196)
The virtual work done by aerodynamic forces can be written by
W
C
aero
=
_
1
0
_
L
C
u
u +L
C
v
v +L
C
w
w +M
C

_
dx. (2.197)
This variational form should be expressed in terms of the elastic axis displacement,
u
e
, to be consistent with the previous formulations for strain and kinetic energy of
rotor blades. This can be achieved by using the relationship between the elastic
displacement and the axial displacement in the mixed coordinate system, Equation
A.23 on page 242.
W
C
aero
=
_
1
0
_

L
C
u
u
e
+

L
C
v
v +

L
C
w
w +M
C

_
dx, (2.198)
where

L
C
u
L
C
u
, (2.199)

L
C
v
= L
C
v
L
C
u
v

+v

_
1
x
L
C
u
d, (2.200)

L
C
w
= L
C
w
L
C
u
w

+w

_
1
x
L
C
u
d, (2.201)
and another form is available for nite element formulation, which is called the
72
weak form. The weak form of Equation 2.198 is given by
W
C
aero
=
_
1
0
_
L
C
u
u
e
+L
C
v
v +L
C
w
w +M
C

+ (v

+w

)
_
1
x
L
C
u
d
_
dx, (2.202)
where the underline term could be neglected, because it is the higher order term,
L
C
u
the foreshortening term. Thus Equation 2.202 can be simplied as:
W
C
aero

_
1
0
_
L
C
u
u
e
+L
C
v
v +L
C
w
w +M
C

_
dx. (2.203)
2.3.2 Noncirculatory Airloads
The airloads acting on the rotor blade can be classied into two categories that are
circulatory and noncirculatory. The derivation of the circulatory loads has been
performed in the previous section. The noncirculatory loads, which are so called
apparent or virtual forces, will be derived in this section.
When the airfoil has a general motion, the lift and moment of the noncirculatory
origin (the apparent mass forces) must be added. For a airfoil section undergoing
plunge motion, h, and pitch motion, , the noncirculatory lift and pitching moment
[139] are given by,

L
NC
w
= L
2
+L
3
= b
2
(

h a
h
b ) + b
2
U ), (2.204)

M
NC

= a
h
bL
2
(
1
2
a
h
)bL
3
+M
a
, M
a
=
b
4
8
, (2.205)
in which L
2
is a lift force with center of pressure at the mid-chord, of amount
equal to the apparent mass, b
2
, times the vertical acceleration at the mid-chord
point. L
3
is a lift force with center of pressure at the three-quarter chord point,
of the nature of a centrifugal force, of amount equal to the apparent mass, b
2
,
times U . M
a
is a nose-down moment equal to the apparent moment of inertia,
73
b
2
(b
2
/8), times the angular acceleration, .
U = R(x +sin ) , a
h
b = (e
d
+
c
4
) ,

h = w (2.206)
=
1
=
o
+

, b =
c
2
(2.207)
where U denotes the free stream tangential velocity, a
h
b is the distance from mid-
chord to the elastic axis (positive aft),

h is the plunge acceleration (positive down),
is the pitch angle (positive nose up) and b is the airfoil semi-chord. The noncir-
culatory airloads are assumed to act directly on the blade undeformed section.
The virtual work done by the noncirculatory airloads that are normalized is
given by
W
NC
aero
=
_
1
0
(L
NC
w
w +M
NC

)dx, (2.208)
where
L
NC
w
=
c
12a
_
w

+ (c/4 +e
d
)(

o
+

) + (x +sin )(

o
+

)
_
, (2.209)
M
NC

=
c
12a
_
(c/4 +e
d
)w

(c/4 + e
d
)
2
(

o
+

)
(c/2 +e
d
)(x +sin )(

o
+

)
c
2
32
(

o
+

)
_
. (2.210)
2.3.3 Quasi-steady Aerodynamics Implementation
The virtual work by the total airloads, which includes both the circulatory and
noncirculatory airloads contributions, is given by
W
aero
= W
C
aero
+W
NC
aero
,
=
_
1
0
_
L
C
u
u
e
+L
C
v
v + (L
C
w
+L
NC
w
)w
+ (M
C

+M
NC

_
dx. (2.211)
74
This equation is discretized using the nite element method. Because many terms
in Equation 2.211 contain displacements, the discretization can be carried out to
produce a complex forcing vector, treating the displacements as known.
In order to improve the eciency of the numerical implementation, the terms of
velocities (Equations 2.1682.170) are divided into constant, linear, and non-linear
groups as follows:

U
R
=

U
RC
+

U
RL
+

U
RNL
, (2.212)

U
T
=

U
TC
+

U
TL
+

U
TNL
, (2.213)

U
P
=

U
PC
+

U
PL
+

U
PNL
, (2.214)
where

U
RC
=
p
cos
r
cos
o
, (2.215)

U
RL
= u

v +xv

+w

+(v

sin +
p
w

cos )
v

r
cos
o
+
r
_

_
sin
o
, (2.216)

U
RNL
= v

+w

+
1
2

_
v

2
+w

2
_
cos , (2.217)

U
TC
= (x +sin ) cos
o
+ ( +
p
cos ) sin
o
, (2.218)

U
TL
=
_
u +v

w
p
+
p

cos +v

cos
p
v

_
cos
o
+
_
w

sin +v
p
+w

cos
p
w

_
sin
o
, (2.219)

U
TNL
=
_
w

+vv

1
2
v

2
(x +sin ) + w

cos
_
cos
o
+
_
v

+vw

xv

cos v

sin
1
2
w
2
_
sin
o
, (2.220)

U
PC
= (

o
+
p
)
r
+ ( +
p
cos ) cos
o
(x +sin ) sin
o
, (2.221)

U
PL
=
_

+w

_
u +v

+(


p
v

)
p
w +cos (
p

+v

)
_
sin
o
+
_
w

+v
p

p
w

+w

cos

sin
_
cos
o
, (2.222)

U
PNL
=
_
w

vv

+
1
2
xv

2
+
1
2
v

2
sin w

cos
_
sin
o
+
_
v

+vw

1
2
w

2
xv

cos +v

sin
_
cos
o
. (2.223)
75
By substituting Equations 2.212 2.214 into Equations 2.189 2.192 and using
Equation 2.15 on Page 45 that was also broken into constant, linear, and nonlinear
terms, the stiness, damping, and mass matrices are made. The detailed derivation
of these matrices are omitted here for the sake of brevity.
76
2.4 Inow and Free Wake Model
There are several types of inow models, such as uniform or linear inow , pre-
scribed wake, free wake, and nite state wake models. Accurate modeling of the
induced inow is important for analysis of ying qualities, detailed power calcu-
lations, noise. For the rotor response, loads and vibrations, a simple model is
normally sucient. At low speed and decent ight conditions, however, the wake
stays close to the rotor disk and has a dominating inuence on the rotor inow. In
this study, two models are used to calculate the rotor inow. First, a simple linear
ow model will be described, and followed by a sophisticated wake model.
2.4.1 Linear Inow
The induced velocity in the rotor plane is the most non-uniform, it being strongly
aected by the presence of discrete tip vortices that sweep downstream near the
rotor plane. In forward ight, the time-averaged longitudinal and lateral inow
can be approximately represented by the variation
= tan +
i
(1 + k
x
x cos +k
y
x sin ) , (2.224)
where
i
is the induced inow ratio, r represents the radial location along the rotor
blade, and

i
=
C
T
2
_

2
+
2
, (2.225)
k
x
=
4
3
_
_
(1 1.8
2
)

1
_

_
2

_
_
, (2.226)
k
y
= 2, (2.227)
which is the Drees model.
77
2.4.2 Free Wake Model
The free wake model (Figure 2.6) used in this study was developed by Tauszig and
Gandhi [131,132] following the free wake methodology initially developed by Bagai
and Leishman [74, 75]. This wake model was implemented at Penn State and has
been used to study new rotor design to alleviate Blade Vortex Interaction. In the
present study, two modules were extracted from Tauszig and Gandhis model. One
module determines the free wake geometry; the other module performs the induced
inow distribution calculation. Once the vorticity strength and wake geometry are
known, the induced velocity (rotor disk inow) is evaluated using the Biot-Savart
Law.
Figure 2.6. Schematic of the wake, discretized in space and time [131]
A general equation describing the positions of the vortex laments in the rotor
wake can be derived from Helmholtzs law (the vorticity transport theorem) by as-
suming that for each point in the ow the velocity is convected at the local velocity,

V
loc
. Considering a single element of a trailed vortex lament, the fundamental
equation describing the transport of the lament is
dr
dt
=

V
loc
(r, t), (2.228)
78
where r = r(
w
,
b
) is the position vector of a point on the lament at a time or
wake age
w
that was trailed from the blade when it was at an azimuth angle
b
.
By assuming that every vortex lament is convected through the ow eld at
the local velocity, a governing equation the geometry of a single element of the
vortex lament can be written as:
r

w
+
r

b
=

+
1

V
ind
[r(
b
,
w
)], (2.229)
where the summation is carried out over the total number of trailed vortex l-
aments, which is the number of azimuthal discretization, that contribute to the
induced velocity eld at any given point. The induced velocity

V
ind
, can be deter-
mined using the Biot-Savart law.
The numerical solution of Equation 2.228 can be accomplished by discretiz-
ing the wake and using nite dierence to approximate the derivatives. Land-
grebe [145] developed the explicit time nite dierence schemes to solve the Equa-
tion. 2.228. The convergence was achieved by obtaining the periodic distorted
wake geometry. Time stepping schemes, however, have been found to be rather
susceptible to numerical instabilities, particularly at low advance ratios, often ne-
cessitating the use of articial numerical damping. Bagai and Leishman [75] devel-
oped the more stable method by enforcing the wake periodicity as the boundary
condition. This method is called the Pseudo-Implicit Predictor-Correct (PIPC)
method by its developers. In this method, the computational domain was dened
as a discretized grid in time
b
and space
w
. An initial wake geometry with a
linear inow is used to start the free-wake calculations. The convergence of this
method was achieved when the RMS change in wake geometry (L
2
norm), between
two successive iterations is within some prescribed tolerance. Some modications
have been made by Tauszig and Gandhi [132] to the wake calculation methodology.
For instance, instead of using one straight segment with a constant vortex core to
represent the trailing vortices, multiple segments that match an arc in a piecewise
linear manner with increasing a vortex core were used to accurately determine the
release tip vortex strength and near wake.
79
Figure 2.7. Flow chart of an aeroelastic analysis with a free wake
The ow chart of an aeroelastic analysis with a free wake is presented in Figure
2.7. The procedure is started by specifying a desired advance ratio and thrust
level. The iterative procedure is carried out to achieve the convergence. Two
steps are used to get the nal results that are the wake geometry and vehicle trim.
The rst step is to provide the initial condition to the free wake analysis using
an aeroelastic analysis with a linear ow and a rigid wake geometry. Then the
wake geometry calculation is carried out until the RMS change in wake geometry
becomes suciently small. Based on new wake geometry, the induced velocity over
the rotor disk is obtained . With this, an aeroelastic analysis is again carried out
to provide the bound circulation to the wake geometry routine. This process is
repeated until both the wake geometry and vehicle trim meet their convergence
criterion.
80
2.5 Aeroelastic Analysis
In this section, a nonlinear periodic system equation is described based on previous
formulations presented in Section 2.2, Section 2.3 and 2.4. The solution of an
aeroelastic response problem is carried out in two steps. First, spatial discretization
based on the nite element method is used to eliminate the spatial dependence, and
subsequently the combined structural and aerodynamic equations are discretized
to nd the solution via the nite element method in time.
To reduce the computational eciency, the normal mode transformation is
applied to the blade nite element equations. This reduces the number of spatial
degrees of freedom from that of blade degrees of freedom to a smaller number of
modes. The force summation method is applied in the calculation of the blade
loads and hub loads. A coupled trim procedure is used to obtain the vehicle trim
condition. The inertial and aerodynamic loads due to the presence of the trailing-
edge aps are treated as an additional load vector for the blade response. This
will be discussed in Chapter 3 on Page 91.
2.5.1 Aeroelastic Response
The blade nite element discretization is shown in Figure 2.8. The blade is dis-
cretized into 5 to 10 beam elements, each consisting of 12 degrees of freedom. The
shape functions based on Hermitian polynomials are used within the elements,
which are the slope continuous functions. For example, the apping displacement
w
i
for the i-th beam element is interpolated by
w
i
= H
1
H
2
H
3
H
4
|
_

_
w
1
w

1
w
2
w

2
_

_
i
, (2.230)
where
H
1
=
1
4
(s + 2) (s 1)
2
, (2.231)
81
H
2
=
L
(i)
el
8
(1 s
2
) (1 s), (2.232)
H
3
=
1
4
(2 s) (s + 1)
2
, (2.233)
H
4
=
L
(i)
el
8
(s
2
1) (s + 1), (2.234)
in which L
(i)
el
is the length of the i-th beam element.
1 2
Trailing edge flaps
S
-1 +1
Figure 2.8. Finite elements for composite rotor blades
The extended Hamiltons principle presented in Section 2.1.4 can be rewritten
in the discretized form as
_
2
0
_
N
el

i=1
q
T
i
([M]
i
q

i
+ [C]
i
q

i
+ [K]
i
q
i
F(q)
i
)
_
d = 0, (2.235)
in which subscript ()
i
indicates the i-th beam element, N
el
is the number of element,
and the elemental nodal displacement vector, q
i
, is given by
q
T
i
= v
1
v

1
w
1
w

1
v
2
v

2
w
2
w

2
|
i
, (2.236)
where w and v represent ap and lag displacements, respectively.

denotes the
elastic twist of a blade.
Derivatives with respect to the azimuth are placed in the damping and mass
82
matrices. First derivatives are damping terms, and second derivatives are mass
terms. They are not physically stiness, mass and damping, but mathematically
behave similarly to stiness, mass, and damping. Then elemental contributions are
assembled into the global governing dierential equation of motion that is given
by
[M]q

+ [C]q

+ [K]q = F(q), (2.237)


where [M], [C], and [K] represent the global mass, damping, and stiness matrices,
respectively. F denotes the global force vector that includes nonlinear terms and
inertial and aerodynamic contributions from trailing-edge aps. q is the global
displacement vector.
In order to reduce the computational cost, the nite element equations in terms
of physical nodal displacements are transformed into the modal space. The natural
frequencies and modes of the blade are calculated through an eigenvalue problem
solving procedure. The frequencies are calculated from an undamped system, such
that,
[M]q

+ [K]q = 0. (2.238)
This equation is solved by an conventional eigenvalue solver. Then the global
blade displacement vector q can be represented by a linear combination of the
eigenvectors. The blade equation is converted to the normal mode equation via a
modal transformation.
q = [] q
R
, (2.239)
where [] is the matrix of eigenvectors or madal matrix, and q
R
represents the
vector of normal mode coordinates. The size of the system is reduced by only
picking certain modes to represent the blade motion. Typically four ap bending
modes, three lag bending modes and two torsion modes are selected. Then the
modal equation of motion is expressed as follow:
[m]q

R
+ [c]q

R
+ [k]q
R
= f(q
R
), (2.240)
where [m], [c], [k], and f are the modal mass, damping, and stiness matrices
83
and force vector, which are given by
[m] = []
T
[M] [], (2.241)
[c] = []
T
[C] [], (2.242)
[k] = []
T
[K] [], (2.243)
[f] = []
T
F. (2.244)
The nonlinear periodic dierential equation presented in Equation 2.240 can be
numerically solved using the nite element method. This equation can be rewritten
as:
_
2
0
q
R

T
([m]q

R
+ [c]q

R
+ [k]q
R
f(q
R
)) d = 0. (2.245)
This can be discretized to eliminate the temporal dependence. The time period
(one revolution) is divided into a number of equally spaced temporal elements, as
shown in Figure 2.9, where each element has three nodes. For instance, one of
modal amplitudes, q
Ri
, for the i-th temporal element can be interpolated by
q
Ri
= H
t1
H
t2
H
t3
H
t4
H
t5
H
t6
|
_

t1

t1

t2

t2

t3

t3
_

_
i
(2.246)
where
H
t1
= s
2

5
4
s
3

1
2
s
4
+
3
4
s
5
, (2.247)
H
t2
=

i
8
_
s
2
s
3
s
4
+s
5
_
, (2.248)
H
t3
= s
2
+
5
4
s
3

1
2
s
4

3
4
s
5
, (2.249)
H
t4
=

i
8
_
s
2
s
3
+s
4
+s
5
_
, (2.250)
84
H
t5
= 1 2s
2
+s
4
, (2.251)
H
t6
=

i
2
_
s 2s
3
+s
5
_
, (2.252)
which are the velocity continuous shape functions based on Hermitian polynomials,
where
i
represents the time-interval of the i-th temporal element, and
t

i
is
the i-th temporal nodal displacement vector. The number of degrees of freedom
of a temporal element depends on the number of selected modes.
1
2
3
4
5
6
7
8
1 2
S
-1 +1
3
Figure 2.9. Finite element discretization in time
Substituting Equation 2.246 into Equation 2.245 yields
N
t

i=1
_

i+1

T
i
( [K
t
()]
i

i
F
t
(
t
, )
i
) d = 0, (2.253)
85
where
[K
t
()]
i
= H

T
i
[m()]H

i
+ H
t

T
i
[c()]H

i
+H
t

T
i
[k()]H
t

i
, (2.254)
F
t
(
t
, )
i
= H
t

T
i
f(
t
, ), (2.255)
and assembling the temporal elements, Equation 2.253 takes the form
[K
t
]
t
= F
t
(
t
). (2.256)
This can be solved by using a Newton-Raphson method that is a gradient-based
method. In this type of iteration scheme, the solution is found by calculating a
gradient between the current point and the solution point.
Performing a rst-order Taylor series expansion to Equation 2.256 about the
current condition,
t

(i)
, yields

(i)
=
_
[K
t
]
F
t

t
=
(i)
t
_
1
_
[K
t
]
t

(i)
F
t
(
(i)
t
)
_
, (2.257)
where

(i)
=
t

(i+1)

(i)
. (2.258)
The initial guess of the response
t

(0)
is solved by

(0)
= [K
t
]
1
F
t
(0). (2.259)
When the error vector given in Equation 2.258,
t

(i)
, is suciently close to
zero, the response is considered converged.
2.5.2 Coupled Propulsive Trim
Once the aeroelastic response has been obtained, the next step is to compute the
blade loads, from which the hub loads can be obtained. Then the vehicle trim
86
condition can be obtained using the hub steady loads together with the forces and
moments acting on the fuselage.
The force summation method is used to calculate the blade loads, which is based
on Newtons law,

F = m a, the sum of the inertial loads must equal the applied
forces. The inertial loads are calculated from closed-form expressions derived in
Section 2.2.2, Equation 2.141 on Page 62. The applied loads to a section of the
blade are the aerodynamic loads, and the elastic forces within the blades. If the
inertial loads and aerodynamic loads are known, then the elastic forces can be
derived as well. Aerodynamic loads are calculated using Equation 2.211 on Page
73 in Section 2.3.3. Then the resulting blade section loads are expressed by
L
u
= L
A
u
+

V
I
x
+L
f
u
,
L
v
= L
A
v
+

V
I
y
+L
f
v
, (2.260)
L
w
= L
A
w
+

V
I
z
+L
f
w
,
M
u
= M
A

+M
I
x
+M
f
u
,
M
v
= v

M
A

+M
I
y
+M
f
v
, (2.261)
M
w
= w

M
A

+M
I
z
+M
f
w
,
where superscripts ()
f
represent trailing-edge ap contributions that include the
inertial and aerodynamic loads due to trailing-edge aps described on Page 101 in
Chapter 3.
These section loads are in the rotating frame and are also in the undeformed
blade coordinate system. To calculate the rotating frame hub loads, additional
integration is required. Equations 2.260 and 2.261 only represent the loads per unit
length. To obtain the blade shear forces and moments at the hub, the spanwise
integration is carried out from the hub center to the blade tip. Then,
_

_
F
x
F
y
F
z
_

_
=
_
1
0
_

_
L
u
L
v
L
w
_

_
dx, (2.262)
87
_

_
M
x
M
y
M
z
_

_
=
_
1
0
_

_
L
v
w +L
w
v +M
u
L
u
w L
w
x +M
v
L
u
v +L
v
x +M
w
_

_
dx. (2.263)
The hub loads are obtained by summing the blade root loads between the N
b
blades at each azimuth position. The rotating frame hub loads are translated to
the xed frame in the vehicle principal directions. The xed frame hub loads are
expressed by
F
H
X
() =
N
b

i=1
_
F
i
x
cos
i
F
i
y
sin
i

p
F
i
z
cos
i
_
, (2.264)
F
H
Y
() =
N
b

i=1
_
F
i
x
sin
i
+F
i
y
cos
i

p
F
i
z
sin
i
_
, (2.265)
F
H
Z
() =
N
b

i=1
_
F
i
z
+
p
F
i
x
_
, (2.266)
M
H
X
() =
N
b

i=1
_
M
i
x
cos
i
M
i
y
sin
i

p
M
i
z
cos
i
_
, (2.267)
M
H
Y
() =
N
b

i=1
_
M
i
x
sin
i
+M
i
y
cos
i

p
M
i
z
sin
i
_
, (2.268)
M
H
Z
() =
N
b

i=1
_
M
i
z
+
p
M
i
x
_
, (2.269)
where F
i
x
, F
i
y
, and F
i
z
are the blade root shear forces from Equation 2.262 due to
the i-th blade, and M
i
x
, M
i
y
, and M
i
z
are the blade root moments due to the i-th
blade from Equation 2.263.
The time histories of the hub forces and moments presented in Equations 2.264
2.269 are not useful by themselves. These forces and moments are broken up
into harmonics of the rotor frequency , 1/rev. For a balanced rotor, only integer
multiples of N
b
/rev are transmitted to the hub. Otherwise, for a dissimilar rotor,
the large 1/rev forces would remain. The steady components of the hub loads are
the rotor thrust, longitudinal and side forces, rolling and pitching moments, and
the rotor shaft torque. The steady loads are therefore used in the helicopter trim
88
process.
For propulsive trim of a helicopter, there are three forces and moments which
must be zero for the vehicle to be equilibrium. These are controlled by the vehi-
cle controls and orientation. The controls and vehicle orientation states, such as
collective pitch
o
, cyclic pitch
1c
and
1s
, tail rotor collective pitch
tr
, and the
shaft tilt angles
s
and
s
, are referred to as control settings. It is assumed that
the engine can supply all the power needed to maintain the ight condition. The
vehicle equilibrium equations to be solved are given by
F
V
() = 0, (2.270)
where
F
V
() = F
V
x
F
V
y
F
V
z
M
V
x
M
V
y
M
V
z
|
T
, (2.271)
=
75

1c

1s

s

s

tr
|
T
(2.272)
in which F
V
x
, F
V
y
, and F
V
x
are the vehicle force residuals, and M
V
x
, M
V
y
, and M
V
z
are the moments. The detailed formulations can be found in References [76, 133].
The nonlinear vehicle equilibrium equations given in Equation 2.270 can be
solved to nd the control settings via a rst order Taylors series expansion with
respect to the control settings.

i
=
_
F
V

=
i
_
1
F
V
(
i
), (2.273)
where

i
=
i+1

i
, (2.274)
the Jacobian matrix is approximated by a forward nite dierence,
F
V

=
i

F
V
(
i+1
) F
V
(
i
)

i
, (2.275)
and the initial guess
0
is estimated by a rigid blade trim analysis based on blade
ap dynamics.
89
The vehicle properties used in this research (see Figure 2.10), which are based
on Reference [135], are listed in Table 2.3.
Figure 2.10. Vehicle conguration for propulsive trim
90
Table 2.3. Vehicle properties
Total vehicle weight 5800 lbs
Vehicle longitudinal c.g. oset, x
CG
/R 0
Vehicle lateral c.g. oset, y
CG
/R 0
Hub location above vehicle c.g., h/R 0.2
Vehicle at plate area, f/R
2
0.01
Tail rotor radius 3.24 ft
Tail rotor solidity,
tr
0.15
Tail rotor location, x
tr
/R 1.2
Tail rotor location above vehicle c.g., h
tr
/R 0.2
Tail rotor lift coecient, c
0tr
, c
1tr
0, 6
Horizontal tail location, x
ht
/R 0.95
Horizontal tail planform area, S
ht
/R
2
0.011
Horizontal tail lift coecient, c
0ht
, c
1ht
0, 6
Chapter 3
Trailing Edge Flap Formulation
This chapter describes the trailing-edge ap formulation and optimal controller for
both vibration and blade loads. In the rst section, the inertial loads of trailing-
edge aps are derived based on the previous work [69]. Followed by the discussion
of available aerodynamic models to predict the aerodynamic loads generated by ap
motions. Finally, in the third section, the optimal controller is described based on
the minimization of an objective function. This objective function includes three
quadratic functions related to vibratory hub loads and blade loads, and active ap
control inputs.
92
3.1 Inertial Contribution
The trailing-edge ap coordinate system is presented in Figure 3.1 showing that
ap coordinates x
f
i
are attached on the trailing edge ap and move along with the
ap by the deection angle
f
.
f
d
y
z
d
x
f
y
f
z
f
w
v
Figure 3.1. Schematic of blade cross-section incorporating a trailing edge ap
The position vector of a mass element on the trailing-edge ap in the ap
coordinate system e
f
i
is given by
r
f
= 0 x
f
2
x
f
3
| e
f
(3.1)
This position vector can be transformed into the blade deformed coordinate system
e
d
i
presented in Section 2.1.2.
r
f
= x
fd
i
e
d
i
, (3.2)
93
in which
x
fd
1
= 0, x
fd
2
= x
f
2

f
x
f
3
d, x
fd
3
= x
f
3
+
f
x
f
2
, (3.3)
where
f
is the ap deection, and d denotes the oset from the trailing-edge ap
hinge line to the blade elastic axis.
This is further transformed into the blade undeformed coordinate system using
Equation 2.9 on Page 43:
r
f
u
= r
f
+ (x +u) v w |e
u
, (3.4)
where
r
f
= x
fu
i
e
u
i
, (3.5)
and
x
fu
i
= x
fd
j
T
du
ji
. (3.6)
The position vector r
f
u
presented in Equation 3.4 is used in the formulation of
ap inertia forces. The position vector r
f
presented in Equation 3.5 is needed to
calculate the ap sectional moments with respect to the blade deformed elastic
axis.
To calculate the sectional loads, the acceleration of a trailing-edge ap particle
with respect to the blade undeformed coordinate system is derived rst. Based on
motion of a particle in a moving coordinate system [69, 142], the acceleration is
expressed as:
a
f
= R

r
f
u
+() +r
f
u

+ 2r
f
u

, (3.7)
where R represents the hub position vector with respect to an inertial frame, and

p
e
u
1
+e
u
3
. (3.8)
The rst term in the underline terms could be dropped out, since the steady state
ight is assumed in this study, which implies that the hub does not accelerate. The
second terms can be also dropped out, if the blade rotational speed is assumed
94
constant.
The inertial contribution to the blade sectional loads associated with a trailing-
edge ap mass element dm
f
can be written as:
L
fI
=
_
A
f
a
f
dm
f
, (3.9)
where A
f
represents the ap cross section, and the trailing-edge aps inertial
contribution to the sectional moments about the blade deformed elastic axis, in
the undeformed basis e
u
, are calculated by
M
fI
=
_
A
f
r
f
a
f
dm
f
. (3.10)
Table 3.1. Order of terms for trailing edge aps
Term List Order

f
O(
1/2
)

f
,

f
, d, x
f
O()
y
f
O(
3/2
)
z
f
O(
5/2
)
The acceleration terms that are independent of trailing-edge ap motions are
already included in the blade inertial loads. Terms associated with trailing-edge
ap deections, such as
f
,

f
and

f
, therefore needs to be considered as the ap
inertial contributions. By applying the small angle assumptions and the ordering
scheme up to O(
3
) (Tables 2.2 and 3.1), which is performed using the Mathematica
program presented in Appendix C.4, the resulting normalized inertial sectional
forces are given by
L
fI
u
= 0, (3.11)
L
fI
v
= S
f
_

f
sin
o
2

o
cos
o
+
f

2
o
sin
o

f
sin
o

o
cos
o
_
, (3.12)
95
L
fI
w
= S
f
_
2

o
sin
o

2
o
cos t
o
+

f
cos
o

o
sin
o
_
, (3.13)
, and the inertial sectional moments are
M
fI
u
= S
f
_

f
v cos
o
d
f
cos
2

o
x
f

p
sin
o

f
v

cos
o

f
w

sin
o
d

f
_
I
f

f
, (3.14)
M
fI
v
= S
f
x
f
cos
o
, (3.15)
M
fI
w
= S
f
x
f
sin
o
, (3.16)
where S
f
and I
f
represent the ap rst sectional moment and ap hinge moment
of inertia, respectively, and they are dened by
S
f
=
_
A
f
x
f
2
dm
f
, (3.17)
I
f
=
_
A
f
(x
f
2
)
2
dm
f
. (3.18)
The inertial trailing-edge ap hinge moment is calculated by
M
I
H
=
_
A
f
(r
f
h
a
f
) e
h
1
dm
f
, (3.19)
where
r
f
h
= 0 (x
f
2

f
x
f
3
) (x
f
3
+
f
x
f
2
) | e
h
, (3.20)
in which e
h
is the coordinate system that has its origin on the trailing edge ap
hinge line with an oset d to the blade elastic axis. Here the acceleration a
f
is expressed in the deformed coordinate system e
d
by transforming Equation 3.7
using Equation 2.9. Applying the ordering scheme up to O(
3.5
), one can obtain
the inertial contribution to the trailing-edge ap hinge moment.
M
I
H
= S
f
_

1
2
d sin 2
o
+ (v x v

) sin
o
+ (x
p
+xw

+w

) cos
o
d

o
_
96
I
f
_
1
2
sin 2
o
+

f
+

o
_
, (3.21)
where the symmetry of trailing-edge ap about x
f
1
x
f
2
plane is assumed,
_
A
f
x
f
3
dm
f
= 0, (3.22)
_
A
f
x
f
2
x
f
3
dm
f
= 0. (3.23)
Note that the inertial hinge moment, Equation 3.21, includes the blade motion as
well as the trailing edge ap motion.
97
3.2 Aerodynamic Models
One of the present study is to reduce the helicopter vibration or the blade loads
using trailing-edge aps. It is important to accurately model the aerodynamic loads
generated by trailing-edge ap motions. In order to accurately predict the ap
performance and ap hinge moments, it is essential to use an unsteady aerodynamic
model. In this section, the widely used classical unsteady aerodynamic model
derived by Theordorsen [143] and the subsonic compressible ow model developed
by Hariharan and Leishman [72, 73] are described.
3.2.1 Incompressible Model
An unsteady airfoil theory with an oscillating trailing edge ap has been devel-
oped by Theodorsen [143]. The most important assumptions behind this approach
are that thin airfoil theory holds and ow is incompressible. Although the incom-
pressible assumption may not be accurate enough, this theory is a good starting
for understanding the aerodynamics due to trailing-edge aps.
Considering a large aspect ratio wing in incompressible and inviscid ow, a
thin airfoil with a ap is undergoing two degrees of freedom: heave motion h(t)
and pitch motion (t) about the blade elastic axis, as shown in Figure 3.2, where
a and e are the normalized distance from the mid-chord to the blade elastic and
to the ap hinge, respectively. The aerodynamically unbalanced trailing-edge ap
rotates about the ap hinge by the angle
f
(t) relative to the chord line.
a
d
f
ab
eb
c
-b
+b
U
Elastic axis
Flap hinge
h
Figure 3.2. Nomenclature for a thin airfoil with a ap
98
The unsteady lift coecient, C
N
, of a thin airfoil with a ap undergoing os-
cillatory motion resulting from trailing-edge ap deection
f
and rate

f
is given
by
C
N
=
b
U
2
_
UT
4

f
bT
1

f
_
+ 2C(k)
_
T
10


f
+
bT
11
2U

f
_
, (3.24)
where U is the free stream velocity and b is the semi-chord of he airfoil. C(k)
is Theodorsens lift deciency function with reduced frequency, k = b/U, which
represents the integrated eect of shed vortex sheet extended from trailing-edge
ap all the way minus innity. The coecients T
4
, T
1
, T
10
, and T
11
are geometric
terms, which depend on the size of the ap relative to the airfoil chord.
The underline terms in Equation 3.24 represents non-circulatory or apparent
mass terms due to the inertia of the uid. When the incompressible uid assump-
tion is used, these terms are proportional to instantaneous displacements. The
remaining terms are the circulatory components due to creation of circulation and
the eects of the shed wake vorticity. Similar expressions can be derived for the
pitching moment, C
M
, and hinge moment, C
H
, generated by trailing-edge ap
motions. The detailed derivation can be found in References [72, 135, 143].
3.2.2 Compressible Model
In the subsonic compressible ow, both the magnitude and phase of the aerody-
namic loads are aected by compressibility eects, thus the simple correction by
the Prandtl-Glauert factor cannot address the phase issue. Unlike the incompress-
ible ow, both the circulatory and non-circulatory loads are subject to time history
eects. The time dependency, however, originates in dierent phenomena. The
non-circulatory loads display time delays as a result of the nite-speed propagation
of an acoustic wave disturbance created by the initial perturbation, while in case
of the circulatory loads the delays are caused by the nite velocity at which shed
circulation is convected downstream away from the airfoil. The non-circulatory
loads dominate initially and the circulatory loads dominate as time progresses.
The compressible unsteady aerodynamic model used in the present study, which
99
was developed by Hariharan and Leishman [73], captures the unsteady eects in the
time domain via an indicial function representation. All derivations and denitions
in detail, such as the unsteady normal force, pitching moment, ap hinge moment
and drag, can be found in Reference [72]. Here a brief description of the recursive
formulation used in unsteady aerodynamics of a apped airfoil is presented.
It is convenient to express the unsteady lift coecient C
N
in terms of coecient
due to ap deection
f
and coecient due to the ap rate

f
. Then,
C
N
(S) = C
N

(S) + C
N

(S), (3.25)
where S is the non-dimensional time dened as the distance traveled by airfoil in
semi-chord, which is given by
S =
2Ut
c
=
2

U
c
, (3.26)
where

() represents the non-dimensional quantity, and

U x +sin .
Considering an arbitrary ap motion,
f
, the lift coecient due to
f
is given
by
C
N

(S) =
2T
10


eff
f
(S) +
2(1 e)
M
T

_
K
n
N

K
n
N

_
, (3.27)
where is the Prandtl-Glauert compressibility factor, =

1 M
2
, and
eff
f
is
the eective ap deection and is given by

eff
f
(S) =
n
f
X
n
1
Y
n
1
, (3.28)
in which
n
f
is the geometry ap deection at the given instant in time step n.
K
n
N

, K
n
N

, X
n
1
, and Y
n
1
are the deciency functions, which account for the time
history between the forcing and the aerodynamic response, and these functions
can be expressed as one step recursive formulae. T

is the non-circulatory time


constant.
100
For an arbitrary ap rate

f
, the lift coecient is given by
C
N

(S) =
T
11
2
_

f
c
U
_
eff
(S) +
(1 e)
2
2M
T

_
K
n
N

K
n
N

_
, (3.29)
where the rst term represents the circulatory term, and the second term indicates
the non-circulatory term.
_

f
c
U
_
eff
(S) is the eective ap rate, which is given by,
_

f
c
U
_
eff
(S) =
_

f
c
U
_
n
X
n
2
Y
n
2
, (3.30)
in which
_

f
c
U
_
n
is the geometry ap deection at the given instant in time step n,
and K
n
N

, K
n
N

, X
n
2
and Y
n
2
are the deciency functions. T

is the non-circulatory
time constant.
Similar expressions can be derived for the pitching moment, C
M
(S), and hinge
moment, C
H
(S), due to the ap deection
f
and ap rate

f
. The detailed
derivation can be found in References [72].
With the trailing edge ap lift, pitching moment and hinge moment coecients,
additional forces and moment generated by trailing-edge aps based on the blade
deformed coordinate system can be obtained, in the non-dimensionalized form
using Equation 2.183 on Page 69,

L
fA
u
= 0, (3.31)

L
fA
v
=

U
2
6a
C
N
sin , (3.32)

L
fA
w
=

U
2
6a
C
N
cos , (3.33)

M
fA

U
2
6a
c C
M
, (3.34)
and the ap hinge moment in the blade deformed coordinate system is given by
M
A
H
=

U
2
6a
c C
H
. (3.35)
101
These forces and moment should be transformed into the blade undeformed coor-
dinate system to be consistent with the formulation of rotor blade, except the ap
hinge moment. Then, via the transformation matrix T
du
ij
presented in Equation
2.9,
L
fA
i
= T
du
ij
T

L
fA
j
, M
fA



M
fA

. (3.36)
The trailing-edge aps inertial and aerodynamic loads are treated as an ad-
ditional loads for the aeroelastic analysis. The inertial contribution of a ap was
presented in Equations 3.113.16. Now the total force contribution of trailing-edge
aps to the blade sectional loads can be written as:
L
f
u
= L
fA
u
+L
fI
u
,
L
f
v
= L
fA
u
+L
fI
v
, (3.37)
L
f
w
= L
fA
u
+L
fI
w
,
and the moment contribution
M
f
u
= M
fA

+M
fI
u
,
M
f
v
= v

M
fA

+M
fI
v
, (3.38)
M
f
w
= w

M
fA

+M
fI
w
.
These forces and moments are added into the blade sectional forces and moments,
which are presented in Equations 2.260 and 2.261 on Page 86 in Section 2.5.2.
102
3.3 Active Trailing Edge Flap Control Algorithm
In this section, the optimal controller is developed based on the minimization
of an objective function to reduce the rotor induced vibration or rotating frame
blade loads. The objective function includes two quadratic functions related to
vibratory hub loads and blade loads, and active ap control inputs. To limit
the trailing-edge ap deections, the actuator saturation is considered, which is
another optimization loop. An active-passive approach developed by Zhang et.
al [5557] is briey reviewed, which will be used in an active loads control to
reduce the control eorts and blade loads.
3.3.1 Feedback Form of Global Controller
This section describes a frequency domain control algorithm used in this study.
The control problem can be transformed from the time domain to the frequency
domain because of the periodic nature of blade response in forward ights. This
periodic assumption is, however, only valid for steady state conditions. Thus a
frequency domain controller is only applicable to steady state ight conditions.
In the present study, a multicyclic controller developed in Reference [144] is
modied and implemented. This approach is based on minimization of an objective
function, such that
J = Z
T
n
W
Z
Z
n
+ K
T
n
W
K
K
n
+
T
n
W


n
, (3.39)
where Z
n
is a hub loads vector containing the N
b
/rev cos and sin harmonic com-
ponents of the xed hub loads at time step n (three hub shear forces and three hub
moments). K
n
is a blade loads vector containing the apwise curvature harmonics,
K
n
=
(i)
1c

(i)
1s

(i)
2c

(i)
2s
|
T
, i = 1, 2, , n
K
, (3.40)
in which
(i)
represents the apwise curvature 1/rev and 2/rev harmonics at the
i-th radial station. n
K
is the number of curvature sensing locations along the rotor
spanwise direction.
n
represents the harmonics of the control inputs. For a typical
103
four-bladed rotor,

n
=
f1c

f1s

f5c

f5s
|
T
, (3.41)
where subscripts c and s denote the cosine and sine components. With this control
input vector, the trailing-edges ap time history can be expressed by

f
() =
5

i=1
[
fic
cos(i) +
fis
sin(i) ] . (3.42)
The matrices W contain penalty weights for the harmonics of the curvature W
K
,
the vibration W
Z
and the control inputs W

, they are given by


W
K
= (1
w
)
w
I, (3.43)
W
Z
= (1
w
)
_

w
I 0
0 (1
w
)I
_
, (3.44)
W

=
w
I, (3.45)
where I and 0 represent identity and null matrices, respectively.
w
,
w
and
w
are scalar weighting parameters. By changing
w
, the controller is instructed to
weight more or less on vibratory hub shears or moments.
A controller may be based on a local linearization assumption. Then, rst
order Taylor series expansions of Z
n
and K
n
with respect to a current control
input vector
n1
yield
Z
n
= Z
n1
+T
Z
(
n

n1
), (3.46)
K
n
= K
n1
+T
K
(
n

n1
), (3.47)
where the sensitivity matrices T relate the linearized system response to multi-
cyclic control inputs and need not be square (however, it will in general have more
rows than columns). These matrices are numerically calculated by perturbing the
control harmonics individually,
T
Z
=
Z
n1

n1
, T
K
=
K
n1

n1
, (3.48)
104
Substituting Equations 3.46 and 3.47 into Equation 3.39 yields
J = [Z
n1
+T
Z
(
n

n1
)]
T
W
Z
[Z
n1
+T
Z
(
n

n1
)]
+ [K
n1
+T
K
(
n

n1
)]
T
W
K
[K
n1
+T
K
(
n

n1
)]
+
T
n
W


n
. (3.49)
Then minimizing J by solving
J

n
= 0, (3.50)
yields

n
= D
_
T
T
Z
W
Z
T
Z
+T
T
K
W
K
T
K


n1
D
_
T
T
Z
W
Z
Z
n1
+T
T
K
W
K
K
n1

, (3.51)
where
D
_
T
T
Z
W
Z
T
Z
+T
T
K
W
K
T
K
+W

1
. (3.52)
This can be further simplied by dening
C
Z
D T
T
Z
W
Z
, (3.53)
C
K
D T
T
K
W
K
. (3.54)
Then, Equation 3.51 can be rewritten as:

n
= C
Z
Z
n1
+C
K
K
n1
(C
Z
T
Z
+C
K
T
K
)
n1
. (3.55)
The update law of the optimal controller that minimizes the objective function
J is expressed by

n
=
n

n1
,
= C
Z
Z
n1
+C
K
K
n1
(C
Z
T
Z
+C
K
T
K
+I)
n1
. (3.56)
105
From the denitions presented in Equations 3.523.54, one can nd that
DW

= C
Z
T
Z
+C
K
T
K
+I, (3.57)
C

. (3.58)
Then, the update law of the controller given in Equation 3.56 can be concisely
expressed as:

n
= C
Z
Z
n1
+C
K
K
n1
C


n1
. (3.59)
Either a local controller or a global controller based on the current derivation
or Reference [144] can be implemented. For a linear system, the global controller
is appropriate. If the system has strong nonlinear properties, the local controller
is needed to avoid error. Since the transfer matrices, T, have to be updated
constantly in a local controller, it will become very time consuming during an
optimization process. In this study, a so-called feedback form of global controller
is implemented, since the system is only moderately nonlinear. In this controller,
the transfer matrices given in Equation 3.48 are assumed to be constant over the
entire range of the control input. This controller, however, is a closed-loop form
when the control input during each step is determined by the feedback of the
measured vibration levels of the previous control step [135].
One important issue associated with the implementation of active trailing-edge
ap systems to the vibration control and blade loads control involves actuator satu-
ration. Saturation can be due to limitations associated with piezoelectric actuation
which can provide ap deections of 4
o
or less. Alternatively, when larger ap de-
ections are possible, for practical reasons, it is desirable to limit ap authority
to 3 4 degrees, so as to avoid interfering with the helicopter handling qualities.
An eective way of limiting saturation without loss of control eectiveness can be
achieved by applying the auto weight approach [93].
In this study, an algorithm based on the bisection method is implemented to
106
nd an optimal weighting value of
w
. This can be expresses as:
minimize
_
|

n
(
w
) |
2

sat
_
2
subject to 0
w
< 1 (3.60)
where

n
represents the optimal control gain vector, and
sat
is the prescribed
limiting value of ap deections or actuator saturation angle.
In this algorithm, if the ap deection is overconstrained or underconstrained,
the weighting matrix W

is appropriately modied to relax or tighten the ap


deection constraint by adjusting
w
. Then the new weighting matrix is input
into the optimal control calculation routine and the process repeats until the ap
is properly constrained.
3.3.2 Active-Passive Hybrid Design
A hybrid active-passive optimization process, which was developed by Zhang et
al. [5557], is implemented by combining an optimal control law with nonlinear
optimization programming for a composite rotor blade model.
The passive optimization is solved with a gradient-based nonlinear constrained
minimization program, the modied feasible direction method [146]. The nal
optimal control/optimization results are obtained when both passive design pa-
rameters and active control actions are optimized.
For an active-passive approach, the constrained optimization problem can be
formulated as:
minimize min
n
[f( x,
n
)]
subject to g
i
( x) i = 1, 2, , n
c
(3.61)
x
L
x x
U
in which x represent the vector containing passive design variables, such as the
blade mass and stiness distributions. n
c
is the number of constraints. f( x,
f
) is
107
the objective function given in Equation 3.39. The design variable vector is given
by
x = m K
22
K
33
K
55
K
25
K
35
|
T
(i)
, i = 1, 2, , N
el
(3.62)
where K
22
, K
33
and K
55
represent the ap bending, lag bending and torsional
bending stiness, respectively. K
25
and K
35
denote the pitch-ap and pitch-lag
composite coupling stiness. The constraints g
i
( x) are given by
g
j
= 1

j

L
j
, j = 1, 2, 3 (3.63)
g
k
=

k

U
k
1, k = 4, 5, 6 (3.64)
where
j
represent the rst natural frequencies of each elastic ap, lag, and torsion
modes. Superscripts L and U denote lower and upper bound, respectively.
The frequency placement constraints will prevent blade resonance as well as
unrealistic blade design. Side constraints are also used to prevent the design vari-
ables from reaching impractical values during the optimization process. For the
optimization tool, a commercial optimization package (DOT [146]) is used.
After the optimization is completed, the trim condition of the vehicle should be
checked, because large pitch-ap coupling stiness K
25
could result in the severe
change of primary control settings. In this case, the vehicle should be re-trimmed
to ensure the proper nal results after solving the optimization problem of Equa-
tion 3.61. The converged trim solutions are usually obtained after three or four
iteration. Detailed description of an active-passive hybrid design method can be
found in References [135].
Chapter 4
Active Loads Control Using a Dual
Flap Conguration
The purpose of this chapter is to investigate the feasibility of multiple trailing-edge
aps for the simultaneous reductions of vibration and blade loads. The concept
involves straightening the blade by introducing dual trailing edge aps in a con-
ventional articulated rotor blade. A classical incompressible theory presented in
Section 3.2.1 is employed to predict the incremental trailing edge ap airloads.
The objective function, which includes vibratory hub loads, bending moment har-
monics and active ap control inputs, is minimized by an integrated optimal con-
trol/optimization process described in Section 3.3. Numerical simulation has been
performed for the steady-state forward ight of an advance ratio of 0.35. It is
demonstrated that through straightening the rotor blade, which mimics the be-
havior of a rigid blade, both the bending moments and vibratory hub loads can be
signicantly reduced.
109
4.1 Introduction
An active control using trailing-edge aps can signicantly reduce the vibration
level of helicopter. It may, however, result in the increase of blade loads. McCloud
III [26] has studied the feasibility of reducing both vibration and blade loads using
a single servo-ap. His results have shown that multi-cyclic control can achieve
both vibration and bending moment reduction with large 1/rev control input.
The vibration level for rotor with a single trailing-edge ap using both the fully
elastic and rigid blade models has been investigated by Millot [63]. His results have
shown that the predicted vibration level using the rigid blade model is much less
than that of using the elastic blade model. This implies that if one can straighten
the rotor blade, the vibratory hub loads can be reduced. The simplest way to
achieve this condition is to make the rotor blade very sti. This will, however,
result in high bending moments although the vibration level will become low. The
other possibility is to straighten the blade through active actions, in which one can
reduce both the bending moment and vibration.
The use of trailing-edge aps for vibration suppression has been studied care-
fully in past decade. However, studies focused on the use of trailing-edge aps
for blade loads control are rare. The objective of this research is to explore the
feasibility of utilizing active control for both rotor blade bending moment and vi-
bration reductions. The dual trailing-edge aps are introduced and applied for
such purposes. One of the active aps is located at the out-board region and the
other one at mid-span. A typical articulated rotor blade is selected as a test bed
for this investigation. A control law for dual ap conguration is developed and
veried through numerical results for a steady-state forward ight condition. An
objective function, which includes the ap-wise curvature harmonics, is dened at
ve dierent radial stations so that minimizing the objective function results in the
straightened blades. Composite tailoring via a pitch-ap coupling is also evaluated
utilizing an active-passive approach.
110
4.2 Description of Analytical Models
The trailing-edge ap model, which is based on the incompressible theory de-
veloped by Theodorsen [143], is formulated and then integrated into the blade
model. Two partial span trailing-edge aps, which are plain type aps, are used
to straighten the blade. One is located at the outboard region of the blade,
0.9R1.0R, and the other is located at around the mid-span, 0.4R0.6R, as shown
in Figure 4.1. Flap span size and location are determined through a parametric
study.
In-board flap
0.4~0.6R
Out-board flap
0.9~1.0R
Figure 4.1. Dual ap conguration for active loads control
In the present analysis, motion of the fuselage is not considered. The aerody-
namics only takes into account the wind velocity, including the helicopter forward
speed and the rotor rotation, and the motion of the blades relative to the hub.
A linear inow distribution, which is the Drees model presented in Section 2.4.1,
is assumed in this study. Sophisticated wake models are important for analysis
of ying qualities, detailed power calculations, and noise. For the rotor response,
loads, and vibration, a simple model is normally sucient, especially for high-speed
forward ight conditions.
The aerodynamic model for trailing-edge ap, which is based on the incom-
pressible theory derived by Theodorsen [143], is used in the present study. This
quasi-steady model may underestimate the hinge moment of trailing-edge ap when
compared to the compressible theory developed by Hariharan and Leishman [72],
especially at high Mach number. However, the hinge moment is not the scope of
this chapter.
A composite rotor blade is discretized into ve nite elements. A coupled
propulsive trim scheme is implemented to simultaneously determine the blade non-
111
linear steady response, vehicle orientation and control setting. The vibratory hub
loads are calculated by integrating the blade and active trailing-edge ap inertial
and aerodynamic loads along the blade using the force summation method, while
bending moments at each radial station are calculated by the curvature method
for the sake of computational eciency.
For the purpose of bending moment reduction, an objective function that re-
tains the curvature information of the blade is introduced, so that minimizing the
objective function results in the straightened blade. A conceptual sketch of dual
ap mechanism is depicted in Figure 4.2. For a four-bladed articulated rotor, the
ap is typically actuated at combinations of 3, 4, and 5/rev to reduce the vibratory
hub loads. The objective of the present study is to minimize bending moments
of the rotor blade. The 1/rev component of apwise bending moment harmon-
ics is dominant, and its maximum occurs at around mid-span. For this reason,
1/rev control input is considered. The control weighting parameters
w
,
w
and

w
are set to 0.2, 0.0 and 1.0, respectively, since the study of this chapter focus on
the bending moment reduction. The optimal control algorithm was described in
Section 3.3 in detail.
Figure 4.2. Conceptual sketch of dual ap mechanism for active loads control
A hybrid active-passive optimization process described in Section 3.3.2 is imple-
mented by combining an optimal control law with nonlinear optimization program-
ming for a composite rotor blade model. The passive optimization is solved with
112
a gradient-based nonlinear constrained minimization program. Both constraints
and bounds for the design variables are presented in Table 4.1. The frequency
placement constraints will prevent blade resonance as well as unrealistic blade
design.
Table 4.1. Constraints and bounds for design variables
Constraints Lower limit Upper limit

1
/
o
1
(ap) 0.95 1.05

2
/
o
2
(lag) 0.9 1.1

3
/
o
3
(torsion) 0.8 1.2
Design variables Lower bound Upper bound
m/m
o
0.90 1.10
K
ii
/ K
o
ii
0.90 1.10
K
25
1 10
4
1 10
4
4.3 Numerical Results and Discussions
For numerical studies, a four-bladed articulated composite rotor with two plain
active aps is investigated. Baseline blade properties and trailing edge ap data
are given in Tables 4.2 and 4.2, respectively. As mentioned in the previous section,
the active aps are set to activate in 1/rev or 1, 2 and 3/rev. The prescribed
ap motion is considered in this study. Results are obtained at a forward ight
speed of an advance ratio of 0.35. All the forces and moments are nondimensional
quantities (see Table 2.1 on Page 46 in Section 2.1.3).
Three cases are computed for comparison among dierent approaches:
1. Baseline: A generic, uniform, articulated rotor blade.
2. Actively controlled blade or Retrot design: Add the active dual aps to the
tuned blade, determines the optimal control inputs, and then trim the blade.
3. Hybrid design: Simultaneously redesign the blade pitch-ap composite cou-
pling stiness K
25
and active control inputs.
113
Table 4.2. Baseline articulated rotor properties for active loads control
Hub type articulated rotor
Number of blades, N
b
4
Rotor radius, R 16.2 ft
Hover tip speed, R 650 ft/sec
Hover tip Mach number, M
tip
0.58
Airfoil NACA 0015
Lift coecients, c
o
, c
1
0, 5.73
Drag coecients, d
o
, d
1
, d
2
0.0095, 0, 0.2
Blade chord, c/R 0.08
Solidity, 0.1
Thrust coecient over solidity, C
T
/ 0.07
Blade linear twist,
tw
8
o
Precone,
p
0
Lock number, 6.34
Flap bending stiness, K
o
22
0.002429
Lag bending stiness, K
o
33
0.022825
Torsional stiness, K
o
55
0.002195
Flapwise mass moment of inertia, mk
2
m1
0.00002114
Lagwise mass moment of inertia, mk
2
m2
0.000369
Blade mass/length, m
o
0.011947 slugs/in
Advance ratio, 0.35
For the purpose of comparison, the vibration and blade moment indices are
dened:
V
LV
=
_
F
2
x
+F
2
y
+F
2
z
+M
2
x
+M
2
z
, (4.1)
M
LV
= max [M
ap
(r, )] , (4.2)
where F
x
, F
y
and F
z
are 4/rev vibratory hub shear forces, and M
x
, M
y
and M
z
are 4/rev vibratory hub moments. M
ap
represents the apwise bending moment
in the rotating frame.
4.3.1 Baseline Articulated Rotor Analysis
The baseline rotor blades are generic, uniform, articulated rotor blades. Blade
natural frequencies and mode shapes are calculated based on the vehicle trim. In
114
Table 4.3. Trailing-edge ap properties for active loads control
In-board ap location 0.4 0.6
Out-board ap location 0.9 1.0
Flap chord ratio, c
f
/c 0.2
Flap mass per unit length, m
f
/m
o
0.0844
Flap chordwise c.g. (after ap hinge), r
I
/c
f
0.149
Flap radius of gyration about ap hinge, r
2
II
/c
2
f
0.109
Oset from blade elastic axis to ap hinge, d/c
f
0.55
this chapter, four ap modes, three lag modes and two torsion modes are used
for the purpose of modal reduction. The baseline blade natural frequencies are
presented in Table 4.4. The blade rst four ap mode shapes, the rst three lag
mode shapes and the rst two mode shapes are shown in Figures 4.3 4.5.
Table 4.4. Natural frequencies of baseline articulated rotor
Mode Frequency, /rev
Flap 1 1.03072
Flap 2 2.68559
Flap 3 5.03369
Flap 4 8.24382
Lag 1 0.24988
Lag 2 3.47887
Lag 3 9.24582
Torsion 1 3.99194
Torsion 2 11.6810
The rotor thrust coecient C
T
is assumed to be 0.007, and quasi-steady blade
element aerodynamics with a linear inow model is used to obtain the blade re-
sponse. The rotor trim control settings and blade tip response are shown in Figures
4.6 and 4.7, respectively, for an advance ratio of 0.35.
115
0 0.2 0.4 0.6 0.8 1
0.05
0
0.05
0.1
0.15
0.2
0.25
0.3
0.35
1st mode
Radial position, x
Flap
Lag
Torsion
0 0.2 0.4 0.6 0.8 1
0.1
0.05
0
0.05
0.1
0.15
2nd mode
Radial position, x
Flap
Lag
Torsion
0 0.2 0.4 0.6 0.8 1
0.06
0.04
0.02
0
0.02
0.04
0.06
0.08
0.1
3rd mode
Radial position, x
Flap
Lag
Torsion
0 0.2 0.4 0.6 0.8 1
0.06
0.04
0.02
0
0.02
0.04
0.06
4th mode
Radial position, x
Flap
Lag
Torsion
Figure 4.3. Articulated blade coupled ap mode shapes
116
0 0.2 0.4 0.6 0.8 1
0.05
0
0.05
0.1
0.15
0.2
0.25
0.3
0.35
1st mode
Radial position, x
Flap
Lag
Torsion
0 0.2 0.4 0.6 0.8 1
0.1
0.05
0
0.05
0.1
0.15
2nd mode
Radial position, x
Flap
Lag
Torsion
0 0.2 0.4 0.6 0.8 1
0.06
0.04
0.02
0
0.02
0.04
0.06
0.08
0.1
3rd mode
Radial position, x
Flap
Lag
Torsion
Figure 4.4. Articulated blade coupled lag mode shapes
0 0.2 0.4 0.6 0.8 1
0
0.05
0.1
0.15
0.2
0.25
0.3
0.35
Radial position, x
1st mode
Flap
Lag
Torsion
0 0.2 0.4 0.6 0.8 1
0.2
0.15
0.1
0.05
0
0.05
0.1
0.15
2nd mode
Radial position, x
Flap
Lag
Torsion
Figure 4.5. Articulated blade torsion mode shapes
117
10
8
6
4
2
0
2
4
6
8
10
C
o
n
t
r
o
l

s
e
t
t
i
n
g
s
,

d
e
g

.75

1c

1s

s

Figure 4.6. Control settings of articulated rotor, = 0.35
0 90 180 270 360
0.1
0.08
0.06
0.04
0.02
0
0.02
0.04
0.06
0.08
0.1
Azimuthal angle ( ), deg
B
l
a
d
e

t
i
p

r
e
s
p
o
n
s
e
Flap
Tosrion
Lag
Figure 4.7. Blade tip response of articulated rotor, = 0.35
118
4.3.2 Rigid Blade vs. Elastic Blade
In order to investigate the vibration level of the rigid blade model, the very large
bending and torsion stinesses are assumed to simulate a rigid blade. The 4/rev
vibratory hub loads are presented in Figure 4.8 and compared to those using the
baseline blade without trailing-edge ap. As expected, predicted vibration levels
of the rigid blade model are much lower than those of the elastic blade model.
This explains snap physics that straightening the blade, which is equivalent to
alleviating the curvature of a blade and leads to attenuating the blade inertial
loads, results in reducing bending moment and vibration.
0
0.001
0.002
0.003
0.004
0.005
Fx Fy Fz
Baseline
Active control (dual flap)
Rigid blade
Figure 4.8. Comparison of vibratory hub loads for active loads control
4.3.3 A Single Flap for Moment Reduction
n general, a single ap located at 0.6 0.8R works well for the purpose of vibration
reduction of helicopter rotor blade. Vibration and apwise bending moment are
examined to investigate a single ap performance. With typical control inputs
3,4,5/rev and the objective function for vibratory hub loads, it is observed that
vibration level is reduced by 88% and apwise moment is increased by 13%. The
119
single ap is also applied to the moment reduction problem. The results show that
vibration and apwise moment are reduced by 54% and 9%, respectively, and the
control settings are severely changed. These results indicate that a single ap is
not only inadequate to reduce the bending moment but also hard to trim a vehicle
when it is used for the purpose of moment reduction.
It is observed that reducing the curvature of blade almost always produces
lower vibration level, while reducing the vibration level does not necessarily yield
lower bending moment.
4.3.4 Dual Flap Performance
Dual ap has been introduced to straighten a rotor blade so that we could re-
duce the bending moment with a reasonable vibration level. Figure 4.2 shows
the dual ap mechanism that generates the additional apwise bending moment
to dynamically straighten the blade. This additional moment is quite eective in
the out-board region. Active control authority is, however, small in the in-board
region due to geometric location of dual ap and low aerodynamic pressure. This
is the reason why we have selected a wider ap at the in-board region.
The vibratory hub shear forces are compared to baseline and rigid blade in
Figure 4.8. Vibratory hub shear forces are reduced to almost the same level as
those of the rigid blade. This is the reason why the dynamically straightened
blade yields lower vibration level than baseline blade without the change of control
settings for trim.
In general, 1/rev control input strongly aects trim since primary control inputs
are 1/rev. Dual ap prole presented in Figure 4.9 explains the reason why control
settings for trim is barely changed. It is observed that in-board ap deection is
larger than out-board ap, approximately 1.4 times, and there is 180 degree out-of-
phase between in-board and out-board aps. This out-of-phase is reasonable since
the bending moment will be reduced by apwise aerodynamic moment due to dual
ap on rotor blade. These out-of-phase and dierence between ap deections
120
make the net lift generated by dual ap to be close to zero. Although net lift
is almost zero, additional apwise moment enforces a vehicle with a lateral tilt.
Indeed, we can see that trimmed control settings have kept after active control, as
shown in Figure 4.10.
0 90 180 270 360
-8
-6
-4
-2
0
2
4
6
8
Out-board flap
In-board flap
F
l
a
p

d
e
f
l
e
c
t
i
o
n
s
,

d
e
g
Azimuth, deg
Dual flap profile
Figure 4.9. Dual ap prole for moment reduction with 1/rev control
Bending moment harmonics along the radial station and maximum bending
moment are presented and compared in Figure 4.11 and Table 4.5, respectively.
Reduction of maximum ap bending moment and lag bending moment are about
32% and 45%, respectively. Note that our objective function currently includes
apwise curvature harmonics only. There, however, is a cross coupling between
ap and lag motion due to blade twist angle. It helps to reduce lagwise curvatures
that directly aect lag bending moment. As it is shown in Figure 4.11, maximum
of 1 and 2/rev harmonics occur in the out-board region. It is shown that the
1/rev harmonic component is greatly reduced at around 0.6R and maximum is
shifted to the in-board region. Flapwise bending moment distributions along the
radial station and azimuth are presented in Figures 4.12 and 4.13, before and after
control, respectively. It is clearly shown that the maximum moment moved from
121
-10
-8
-6
-4
-2
0
2
4
6
8
10
t
h
e
t
a
_
7
5
t
h
e
t
a
_
1
c
t
h
e
t
a
_
1
s
a
l
p
h
a
_
s
p
h
i
_
s
Baseline
Dual flap
Figure 4.10. Control settings of baseline and actively controlled rotors
the out-board region to the in-board region after control. These results indicate
the superiority of the dual ap conguration over the single ap one for bending
moment reduction problem.
Table 4.5. Reduction of maximum bending moments maximum moments
Maximum bending moments Flap Lag
Baseline (10
4
) 2.57 2.50
Dual ap (10
4
) 1.75 1.36
Reduction (%) 31.9 45.4
122
0
0.00005
0.0001
0.00015
0.0002
0.00025
0
.
0
0
.
1
0
.
2
0
.
3
0
.
4
0
.
5
0
.
6
0
.
7
0
.
8
0
.
9
Radial station, r
Baseline 1/rev
Baseline 2/rev
Retrofit 1/rev
Retrofit 2/rev
Figure 4.11. Harmonics of apwise bending moment along the radial station
123
-1 -0.5 0 0.5 1
Y
-1
-0.75
-0.5
-0.25
0
0.25
0.5
0.75
X
Z
0.0121
0.0104
0.0086
0.0069
0.0051
0.0034
0.0017
-0.0001
-0.0018
-0.0036
-0.0053
-0.0071
-0.0088
-0.0106
-0.0123
Advancing side Retreating side
Figure 4.12. Flapwise bending moment distribution before control
-1 -0.5 0 0.5 1
Y
-1
-0.75
-0.5
-0.25
0
0.25
0.5
0.75
X
Z
0.0121
0.0104
0.0086
0.0069
0.0051
0.0034
0.0017
-0.0001
-0.0018
-0.0036
-0.0053
-0.0071
-0.0088
-0.0106
-0.0123
Advancing side Retreating side
Figure 4.13. Flapwise bending moment distribution after control
124
4.3.5 Multicyclic Control for Moment and Vibration Re-
duction
In this case, the objective function includes two quadratic functions that are vi-
bratory hub loads and apwise bending moment harmonics. Multicyclic control
inputs, such as 1, 2 and 3/rev, have been considered to improve both bending mo-
ment and vibration load reductions. 1 and 2/rev inputs contribute to the reduction
of the apwise bending moments, and 3/rev input helps to reduce the vibration
level.
Figures 4.14 and 4.15 show the dual ap proles with multicyclic control inputs.
It is observed that there is still 180 degree out-of-phase between in-board and out-
board aps, irrespective of the number of control inputs. These results indicate
that in-board ap should deect to opposite direction of the out-board ap in order
to reduce the apwise bending moment. Multicyclic control yields better results
than a single control. As mentioned earlier, 3/rev control input is dominant in
reducing vibration level, and tends to increase maximum bending moments.
0 90 180 270 360
-10
-5
0
5
10
15
Out-board TEF
In-board TEF
Dual flap profile
F
l
a
p

d
e
f
l
e
c
t
i
o
n
s
,

d
e
g
Azimuth, deg
Figure 4.14. Dual ap prole with 1 and 2/rev control inputs
The reason of this can be deduced from comparisons of vibration index and
125
0 90 180 270 360
-10
-5
0
5
10
15
Out-board TEF
In-board TEF
Dual flap profile
F
l
a
p

d
e
f
l
e
c
t
i
o
n
s
,

d
e
g
Azimuth, deg
Figure 4.15. Dual ap prole with 1, 2 and 3/rev control inputs
maximum apwise bending moment that are presented in Figure 4.16 and Table
4.6. Control inputs with 1 and 2/rev slightly improve both vibration and moment
reductions. The vibratory hub loads are greatly reduced with slightly increasing
maximum of bending moment when 3/rev input is added to control input vector.
The required ap deections are, however, increased with adding higher harmonic
control inputs. As shown in Figures 4.9, 4.14 and 4.15, the maximum deections
of out-board ap is almost constant with 5 degree, while those of in-board ap
is linearly increased from 7 degrees to 14 degrees. As shown in Figure 4.16, a
conventional single ap performance for the purpose of vibration reduction is quite
good but it normally causes the increase of maximum apwise bending moment as
well as the blade fatigue load.
Table 4.6. Reduction of vibration and moment reductions with dierent control inputs
Control inputs Vibration reduction (%) Max. apwise moment
reduction (%)
1/rev 56.68 31.96
1 and 2/rev 60.09 34.60
1, 2 and 3/rev 96.01 24.42
126
Figure 4.16. Comparison of vibration index and maximum apwise bending moment
with dierent control inputs
Note that dual ap performance with 1/rev control input is not good for vi-
bration reduction since 1/rev does not aect the vibratory hub loads. It, however,
does work for vibration reduction when an objective function includes the bending
moment harmonics. Dual ap, for the purpose of vibration reduction, is superior
over a single ap as reported in Reference [94], but it also yields higher bending
moment than a single ap does.
4.3.6 Active-Passive Hybrid Design
The hybrid design can be applied to reduce the required maximum ap deection.
As shown in Figure 4.11, the 1 and 2/rev harmonics are signicantly reduced at
around 0.7R where the most eective additional aerodynamic moment due to dual
ap occurs. Passive design parameters, such as non-structural mass and composite
pitch-ap coupling stiness, maximum ap deection constraints (4 degrees), and
1/rev control input of active dual ap are considered in the hybrid design, while
ap and lag stiness are not considered.
The current dual ap conguration is quite eective on reducing the objective
function related to bending moment harmonics. There is no room for further re-
127
duction. Thus weighting matrix for bending moment, W
K
, described in Equation
3.39 has been chosen to avoid being overly minimized at around 0.7R, so that pas-
sive parameters help to reduce the required ap deections, vibration level and the
maximum bending moment. In Figure 4.17, vibratory hub loads of hybrid design
are presented and compared to those of baseline and retrot design. Vibration
index and maximum apwise bending moment are compared in Figure 4.18 and
Table 4.7. The improvement is not signicant when compared to the retrot de-
sign, but the required maximum ap deection is reduced by 25% (Figure 4.19).
This dierent ap deection might cause a signicant change of control settings,
but it turns out that there is no signicant change as shown in Figure 4.20.
Table 4.7. Comparisons of maximum moment, vibration index and control eorts
Max. moment Vibration Max. TEF
Design reduction (%) reduction (%) deection, deg
Flap / Lag Outboard / Inboard
Retrot 32 / 45 56.7 4.88 / 6.70
Hybrid 34 / 44 61. 8 5.09 / 4.90
0
0.001
0.002
0.003
0.004
0.005
Fx Fy Fz
Baseline
Retrofit
Hybrid
Figure 4.17. Vibratory hub shears comparison for active loads control
128
0
1
2
3
4
5
6
7
8
Vib. Index (1e3) Max. Flapwise moment (2e3)
baseline
retrofit
hybrid
Figure 4.18. Comparison of vibration index and maximum apwise moment of hybrid
designed rotor
0 90 180 270 360
-8
-6
-4
-2
0
2
4
6
8
Out-board flap
(Retrofit,Hybrid)
In-board flap
(Retrofit)
In-board flap
(Hybrid)
F
l
a
p

d
e
f
l
e
c
t
i
o
n
s
,

d
e
g
Azimuth, deg
Figure 4.19. Dual ap proles for retrot and hybrid designs
129
-10
-8
-6
-4
-2
0
2
4
6
8
10
t
h
e
t
a
_
7
5
t
h
e
t
a
_
1
c
t
h
e
t
a
_
1
s
a
l
p
h
a
_
s
p
h
i
_
s
Baseline
Retrofit
Hybrid
Figure 4.20. Control settings of baseline, retrot and hybrid designed rotors
130
Passive design parameters can take care of the lack of aerodynamic lift at
the in-board region. The non-structural mass and composite pitch-ap coupling
stiness, K
25
, distributions are shown in Figure 4.21. The non-structural mass
helps to dynamically straighten the blade by increasing the tip mass that will
increase the centrifugal forces. The coupling stiness could help to redistribute
the blade pitch in order to compensate the lack of in-board ap lift. Through
the ne-tuning of the blade passive parameters, the dual ap works much more
eciently in reducing the bending moment.
k25
-1.0E-04
-5.0E-05
0.0E+00
5.0E-05
1.0E-04
1 2 3 4 5
Mass
-0.1
-0.05
0
0.05
0.1
1 2 3 4 5
Figure 4.21. Blade non-structural mass and pitch-ap composite coupling stiness,
K
25
, distribution
131
4.4 Summary
The purpose of this chapter was to investigate the feasibility of multiple trailing-
edge aps for the simultaneous reductions of vibration and blade loads. The con-
cept involved straightening the blade by introducing dual trailing edge aps in a
conventional articulated rotor blade. An active loads control strategy was simu-
lated for the steady-state forward ight condition, = 0.35.
Based on the present study, the principal conclusions obtained are summarized
as follows:
Straightening the rotor blade using dual ap congurations can reduce both
vibratory hub loads and bending moments without a signicant change of
control settings for trim. Only 1/rev control input is required to reduce
vibratory hub loads with the present method, which is well suited for resonant
actuation systems that will be described in Chapter 6.
The proposed active loads control method can reduce the apwise bending
moment by 32% and the vibratory hub loads by 57%, simultaneously, by
dynamically straightening the blade with the 1/rev control input.
For bending moment reduction, 1/rev control input is dominant, and the
3/rev action is very eective on vibratory hub loads but detrimental on ap-
wise bending moments.
For hybrid design, the maximum apwise bending moment and vibration
are simultaneously reduced by 34% and 62%, respectively. The required
maximum ap deection is reduced by 25% when compared to retrot design.
Chapter 5
Helicopter Vibration Suppression via
Multiple Trailing-Edge Flaps with
Resonant Actuation Concept
The objective of this chapter is to investigate the vibration reduction using the
multiple trailing-edge aps conguration. The concept involves deecting each in-
dividual trailing-edge ap using a compact resonant actuation system (see Chapter
6). Each resonant actuation system could yield high authority, while operating at
a single frequency. The rotor blade airloads are calculated using quasi-steady blade
element aerodynamics with a free wake model. A compressible unsteady aerody-
namics model is employed to accurately predict the incremental trailing edge ap
airloads. Both trailing-edge ap nite wing eects and actuator saturation are in-
cluded in the simulation. A numerical simulation is performed for the steady-state
forward ight ( = 0.15 0.35). It is demonstrated that multiple trailing-edge ap
conguration with the resonant actuation system can reduce the required trailing-
edge ap hinge moments. The analysis and parametric study are conducted to
explore the nite wing eect of trailing-edge aps and actuator saturation.
133
5.1 Introduction
It has been shown that improvements in helicopter vibration reduction can be
achieved by smart materials. Piezoelectric actuation system is expected to be
compact, light weight, low actuation power, and high bandwidth devices that can
be used for multi-functional roles such as to suppress vibration and noise, and
increase aeromechanical stability. While piezoelectric materials-based actuators
have shown good potential in actuating trailing edge aps, they can only provide
a limited stroke. This limitation can be critical in cases where large trailing-edge
ap deections are required. The eorts to improve the piezoelectric actuator per-
formance have been made by researchers in developing amplication mechanisms
of various types (see Section 1.2.3 on Page 20). In general, these devices are still
limited in their performance. To circumvent this limitation, a resonant actuation
system can be used to enhance the eectiveness of piezoelectric actuators, which
will be presented in Chapter 6.
A multiple piezoelectric actuator conguration has been considered and tested
in Eurocopter to adjust the required control power and surface [114]. In this work,
a single ap is segmented into three parts, and all actuator is controlled by the same
command. On the other hand, multiple trailing-edge ap congurations have been
studied to reduce the vibration of helicopter rotor system, in which each actuator
operates independently. Myrtle and Friedmann [78] have shown that the dual ap
conguration is almost completely unaected by the change of torsional stiness of
rotor blade. Recently, Cribbs and Friedmann [93] have developed the ap deection
saturation model through an automated approach to reduce the required maximum
ap deection. They have shown that the imposition of saturation of ap deection
could result in the dierent prole and reduced magnitude of the active ap while
maintaining almost the same vibration level as models without actuator saturation.
The actuator saturation model has been extended to reduce the vibration due to
dynamic stall using a dual ap conguration [94]. They showed that dual ap is
superior over a single ap in vibration reduction. In the previous chapter, Chapter
4, it was demonstrated that the dual ap conguration can be also applied to
reduce both vibration and blade loads by dynamically straightening the blade.
134
In general, a single trailing-edge ap works well for the purpose of vibration
reduction of helicopter rotor. With typical control inputs 3,4 and 5/rev, it has
been demonstrated via numerical simulations that vibration level can be reduced
by about 80%. As mentioned earlier, however, piezoelectric actuators provide a
limited stroke. For the same TEF deections, the actuator design specication
in multiple-ap conguration is more exible than in single-ap conguration,
because the required hinge moment is much less due to small control surface area.
The goal of this chapter is to develop an active vibration control method us-
ing multiple trailing-edge aps. In this study, it is assumed that each individual
trailing-edge ap is operating at a single frequency to utilize a compact resonant
actuation system (see Chapter 6). One of the problems associated with a resonant
actuation system is the operating bandwidth. This can be resolved using the mul-
tiple trailing-edge aps conguration, in which each ap is designed to operate at
a single frequency that is one of the operating frequencies (e.g, 3, 4 and 5/rev for
a four-bladed rotor).
135
5.2 Description of Analytical Models
Aerodynamic loads acting on the blade are calculated using quasi-steady blade
element theory. A free wake model, which is extracted from Tauzsig and Gandhis
code [132] (see Section 2.4.2 on Page 77), is used to determine the non-uniform
inow distribution over the rotor disk. A compressible unsteady trailing-edge ap
aerodynamic model developed by Hariharan and Leishman [73] (Section 3.2.2 on
Page 98), is formulated and then integrated into the blade model. Multiple par-
tial span trailing-edge aps, which are plain types, are used to control the rotor
vibration. Single-, dual- and multiple-TEF congurations (Figure 5.1) are also
considered to compare the performance and required trailing-edge ap deections.
Figure 5.1. Various congurations of the rotor with trailing-edge aps
In multiple trailing-edge aps conguration, a trailing-edge ap is normally
smaller than one in a single ap conguration. Thus the aspect ratio could be
critical in using multiple-TEFs conguration. Based on the approximate lifting
surface theory [147], the lifting curve slope of trailing edge ap can be expressed
136
as:
a
f
a
fo
=
AR
AR + 2(AR + 4)/(AR + 2)
, (5.1)
in which
AR =
L
2
f
L
f
c
f
, (5.2)
where a
fo
is the nominal lifting curve slope of trailing-edge ap, which is normally
2. L
f
and c
f
represent the ap span and chord, respectively.
The unsteady lift of a wing of nite aspect ratio was reported by Jones [148].
Results indicate that the starting lift of nite wing is similar to that of innite wing,
and the change of unsteady lift curve slope with respect to AR is approximately
proportional to quasi-steady lift curve slope correction factor. Thus Equation 5.1
is applied to the normal force coecient C
N
presented in Equation 3.25 on Page 99,
in which C
N
was calculated by a compressible unsteady trailing edge ap model.
A rotor blade is discretized into ve or ten nite elements. In order to reduce
the computational cost, the nite element equations in terms of physical nodal
displacements are transformed into modal space. Four ap, three lag, and two
torsion modes are used in this chapter. Eight temporal elements are used, and
velocity-continuous shape functions, which are fth-order polynomials, are used
within the temporal elements. A coupled propulsive trim scheme is implemented to
simultaneously determine the blade nonlinear steady response, vehicle orientation
and control setting. The vibratory hub loads are calculated by integrating the blade
and active trailing-edge ap inertial and aerodynamic loads along the blade using
the force summation method. The details of aeroelastic analysis were described in
Section 2.5 on Page 80.
For a four-bladed hingeless rotor, the ap is typically actuated at combinations
of 3, 4, and 5/rev to reduce the vibratory hub loads. In this study, the multiple-ap
conguration to utilize the resonant actuation system is introduced, so that each
ap operates at the specic frequency. Then the control input vector presented in
Equation 3.41 is modied as:

n
=
(1)
fc

(1)
fs

(2)
fc

(2)
fs

(3)
fc

(3)
fs
|
T
, (5.3)
137
where superscript ()
(i)
indicates the i-th trailing-edge ap. Flap deections d
f
()
are limited by 2
o
4
o
using the actuator saturation algorithm presented in Equa-
tion 3.60 on Page 106, which automatically determines the control weight param-
eter
w
. For the purpose of the vibration reduction, the other control weighting
parameters,
w
and
w
, are set to 0.2 and 0.0, respectively.
5.3 Results and Discussions
For numerical studies, a four-bladed hingeless rotor with three plain active aps
is investigated. The baseline blade and trailing-edge ap properties are listed in
Table 5.1. Active trailing-edge aps are set to activate at 3, 4 and 5/rev. Each
trailing-edge ap is operated at a single frequency that is one of 3, 4 and 5/rev
in the multiple-trailing edge ap conguration to utilize the resonant actuation.
Results are obtained at forward ight speeds ( = 0.15, 0.35).
Four cases are computed for comparison among the various trailing-edge ap
congurations (see Figure 5.1):
1. Baseline: A generic, uniform, hingeless rotor blade.
2. Single-TEF conguration : Add a single active ap to the generic blade.
3. Dual-TEFs conguration: Add two active aps to the generic blade.
4. Multiple-TEFs conguration: Add three active aps, in which each ap op-
erates at a single frequency, to the generic blade.
For the purpose of comparison, the vibration index dened in Equation 4.1 on
Page 113 is used.
5.3.1 Baseline Hingeless Rotor Analysis
The baseline rotor blades are generic, uniform, hingeless rotor blades. Blade nat-
ural frequencies and mode shapes are calculated based on the vehicle trim. In this
chapter, four ap modes, three lag modes and two torsion modes are used for the
138
Table 5.1. Hingeless rotor and trailing-edge ap properties
Main rotor properties
Hub type hingeless rotor
Number of blades, N
b
4
Rotor radius, R 16.2 ft
Hover tip speed, R 650 ft/sec
Hover tip Mach number, M
tip
0.58
Airfoil NACA 0015
Lift coecients, c
o
, c
1
0, 5.73
Drag coecients, d
o
, d
1
, d
2
0.0095, 0, 0.2
Blade chord, c/R 0.08
Solidity, 0.1
Thrust coecient over solidity, C
T
/ 0.07
Blade linear twist,
tw
8
o
Precone,
p
0
Lock number, 6.34
Flap bending stiness, K
o
22
0.008345
Lag bending stiness, K
o
33
0.023198
Torsional stiness, K
o
55
0.00225
Flapwise mass moment of inertia, mk
2
m1
0.0001
Lagwise mass moment of inertia, mk
2
m2
0.0004
Blade mass/length, m
o
0.011947 slugs/in
Advance ratio, 0.15 0.35
Trailing-edge ap properties
Flap chord ratio, c
f
/c 0.2
Flap mass per unit length, m
f
/m
o
0.0844
Flap chordwise c.g. (after ap hinge), r
I
/c
f
0.149
Flap radius of gyration about ap hinge, r
2
II
/c
2
f
0.109
Oset from blade elastic axis to ap hinge, d/c
f
0.55
purpose of modal reduction. The baseline blade natural frequencies are presented
in Table 5.2. The blade rst four ap mode shapes, the rst three lag mode shapes
and the rst two mode shapes are shown in Figures 5.2 5.4.
139
Table 5.2. Natural frequencies of baseline hingeless rotor
Mode Frequency, /rev
Flap 1 1.1469
Flap 2 3.4041
Flap 3 7.4938
Flap 4 13.4995
Lag 1 0.7499
Lag 2 4.3682
Lag 3 11.0249
Torsion 1 3.5546
Torsion 2 10.4414
0 0.2 0.4 0.6 0.8 1
0.05
0
0.05
0.1
0.15
0.2
0.25
0.3
1st mode
Radial position, x
Flap
Lag
Torsion
0 0.2 0.4 0.6 0.8 1
0.08
0.06
0.04
0.02
0
0.02
0.04
0.06
0.08
0.1
0.12
2nd mode
Radial position, x
Flap
Lag
Torsion
0 0.2 0.4 0.6 0.8 1
0.06
0.04
0.02
0
0.02
0.04
0.06
0.08
3rd mode
Radial position, x
Flap
Lag
Torsion
0 0.2 0.4 0.6 0.8 1
0.04
0.03
0.02
0.01
0
0.01
0.02
0.03
0.04
0.05
4th mode
Radial position, x
Flap
Lag
Torsion
Figure 5.2. Hingeless blade coupled ap mode shapes
140
0 0.2 0.4 0.6 0.8 1
0.05
0
0.05
0.1
0.15
0.2
0.25
0.3
1st mode
Radial position, x
Flap
Lag
Torsion
0 0.2 0.4 0.6 0.8 1
0.08
0.06
0.04
0.02
0
0.02
0.04
0.06
0.08
0.1
0.12
2nd mode
Radial position, x
Flap
Lag
Torsion
0 0.2 0.4 0.6 0.8 1
0.06
0.04
0.02
0
0.02
0.04
0.06
0.08
3rd mode
Radial position, x
Flap
Lag
Torsion
Figure 5.3. Hingeless blade coupled lag mode shapes
0 0.2 0.4 0.6 0.8 1
0
0.05
0.1
0.15
0.2
0.25
Radial position, x
1st mode
Flap
Lag
Torsion
0 0.2 0.4 0.6 0.8 1
0.1
0.08
0.06
0.04
0.02
0
0.02
0.04
0.06
0.08
0.1
2nd mode
Radial position, x
Flap
Lag
Torsion
Figure 5.4. Hingeless blade torsion mode shapes
141
The rotor thrust coecient C
T
is assumed to be 0.007, and quasi-steady blade
element aerodynamics with a free-wake model is used to obtain the blade response.
The rotor trim control settings are presented in Figures 5.5 and 5.6 for = 0.15
and = 0.35, respectively. The blade tip responses are shown in Figures 5.7 and
5.8 for low and high speed ight conditions. The wake eect is signicant at the
low speed ight ( = 0.15), as shown in Figure 5.7 where showing that the blade
tip apping response contains higher harmonic components.
4
2
0
2
4
6
8
10
C
o
n
t
r
o
l

s
e
t
t
i
n
g
s
,

d
e
g

.75

1c

1s

s

Figure 5.5. Control settings of hingeless rotor, = 0.15
10
5
0
5
10
15
C
o
n
t
r
o
l

s
e
t
t
i
n
g
s
,

d
e
g

.75

1c

1s

s

Figure 5.6. Control settings of hingeless rotor, = 0.35
142
0 90 180 270 360
0.03
0.02
0.01
0
0.01
0.02
0.03
0.04
0.05
0.06
Azimuthal angle ( ), deg
B
l
a
d
e

t
i
p

r
e
s
p
o
n
s
e
Flap
Torsion
Lag
Figure 5.7. Blade tip response of hingeless rotor, = 0.15
0 90 180 270 360
0.04
0.02
0
0.02
0.04
0.06
0.08
Azimuthal angle ( ), deg
B
l
a
d
e

t
i
p

r
e
s
p
o
n
s
e
Flap
Torsion
Lag
Figure 5.8. Blade tip response of hingeless rotor, = 0.35
143
5.3.2 Flap Eect to Free-Wake Geometry
Multiple trailing-edge aps could cause discrete lift forces along the rotor spanwise
direction. In subsection, the wake geometry, blade tip response and control settings
are compared to investigate the ap eect to them. Free-wake vertical geometries
are shown in Figures 5.9 and 5.10 for advance ratio of 0.15 and 0.35, respectively.
The changes in wake geometry due to the number of beam elements and the ap are
not signicant, since the tip vortices generated by the ap itself are not considered
in this study. Blade tip responses are presented in Figures 5.11 and 5.12 for
= 0.15 and 0.35, respectively. Torsion tip responses are aected by both the
number of beam elements and the ap, especially for low speed ights (Figure
5.11). Trim control settings for these cases are shown in Figure 5.13 for = 0.15
and Figure 5.14 for = 0.35. It is observed that changes in control settings due
to the number of beam elements and the ap are not signicant. In the study of
this chapter, ve beam elements are used to save the computational time.
1 0 1 2 3 4 5
1.2
1
0.8
0.6
0.4
0.2
0
0.2
Y wake geomtry for =0.15
element 5
element 10
element 10 w/ flap
Figure 5.9. Vertical wake geometry with the number of beam elements and the presence
of the ap for = 0.15
144
2 0 2 4 6 8 10
1.4
1.2
1
0.8
0.6
0.4
0.2
0
Y wake geomtry for =0.35
element 5
element 10
element 10 w/ flap
Figure 5.10. Vertical wake geometry with the number of beam elements and the pres-
ence of the ap for = 0.35
0 90 180 270 360
0.04
0.03
0.02
0.01
0
0.01
0.02
0.03
0.04
0.05
0.06
B
l
a
d
e

t
i
p

r
e
s
p
o
n
s
e
Azimuth, deg
5 beam elements w/ flap
10 beam elements w/ flap
Recalculated freewake w/ flap
Flap
Torsion
Lag
Figure 5.11. Blade tip responses with the number of beam elements and the presence
of the ap for = 0.15
145
0 90 180 270 360
0.04
0.02
0
0.02
0.04
0.06
0.08
B
l
a
d
e

t
i
p

r
e
s
p
o
n
s
e
Azimuth, deg
5 beam elements w/ flap
10 beam elements w/ flap
Recalculated freewake w/ flap
Flap
Torsion
Lag
Figure 5.12. Blade tip responses with the number of beam elements and the presence
of the ap for = 0.35
Figure 5.13. Control settings with the number of beam elements and the presence of
the ap for = 0.15
146
Figure 5.14. Control settings with the number of beam elements and the presence of
the ap for = 0.35
147
5.3.3 Determination of Trailing-Edge Flap Locations
To determine the ap locations for multiple-TEFs congurations, the parametric
study is performed using a small single ap conguration (L
f
= 0.07R) for low-
and high-speed forward ight condition, where the ap motions were limited in
amplitude of two degrees to recognize the ap eectiveness in terms of radial
locations.
Figure 5.15 shows the results showing the eects of ap locations. As expected,
the vibration reductions are directly related to the spanwise location of the ap. In
low-speed ight ( = 0.15), 4/rev control input is very eective since the vertical
hub shear force (F
z
) is dominant. It is noted that the greatest reductions in
vibration observed in Figure 5.15 occurs with ap located away from x 0.8R,
that is, away from the node of second apwise bending mode. Apparently the ap
benets by being positioned where it may produce large generalized forces on the
second ap bending mode [69, 114].
For high-speed ight ( = 0.35), 3/rev control input is relatively eective
compared to the others. The eect of 5/rev control input is not signicant in both
low- and high-speed ight conditions. Thus 5/rev is discarded in multiple-TEFs
conguration to increase the eectiveness of the aps. Resulting control input
sequences for various trailing-edge ap congurations are summarized in Table
5.3.
Table 5.3. Control input sequences and ap locations for multiple-ap congurations
Flap Single Dual Multiple
1st location 0.6 0.8R 0.6 0.74R 0.6 0.67R
control 3,4,5/rev 3,4,5/rev 3/rev
2nd location 0.9 0.97R 0.7 0.77R
control 3,4,5/rev 4/rev
3rd location 0.9 0.97R
control 4/rev
148
Figure 5.15. Vibration reduction vs. radial locations of trailing-edge aps
149
5.3.4 Finite Wing Eects
Figure 5.16 shows the change of lift curve slope with respect to aspect ratio AR.
In single- and multiple-ap congurations, the decreases of lift curve slope due to
the nite wing eect are 15% and 38%, respectively. So the lift by multiple-aps
is reduced by 23% compared to a single-ap conguration.
Figure 5.16. Lift curve slope vs. aspect ratio for elliptical lift distribution
For multiple-ap conguration with lift ap (the pitching moment of ap is
set to zero), vibration reductions are presented in Figure 5.17. It is observed
that performance degradation due to nite wing eect is less than 10% (relative
percentage) in both low- and high-speed ight. Recalling that nite wing correction
is applied to normal force only, it can be deduced that the kinetic energy due to ap
motion contributes to vibration reduction. On the other hand, nite wing eect
can be regarded as the weight to control eorts. This is clearly shown in Figure
5.18 showing that 3rd ap deection increases 1 degree, and 2nd ap phase changes
150
slightly for low-speed ight condition ( = 0.15). Similar trend is also observed
in high-speed ight condition ( = 0.35). This implies that the controller is able
to compensate for changes in ap capability by adjusting the ap deections and
phases. This is the reason why trailing-edge ap performance is not reduced by
38% that is estimated by the quasi-steady correction factor presented in Equation
5.1.
Figure 5.17. Vibration reduction by multi-ap conguration with lift ap
Figure 5.18. Flap deection harmonics of multi-ap conguration with the lift ap,
advance ratio: = 0.15, actuator saturation:
sat
f
= 4
o
151
5.3.5 Eectiveness of Multiple-Flap Conguration
Single-, dual- and multiple-ap congurations are investigated to explore the multiple-
ap motion mechanism. For useful insights, the ap motions are limited by 2
degree, and only 4/rev control input is considered for low-speed ight condition.
Figure 5.19 shows a polar diagram of trailing-edge ap motion for various single-
ap congurations. Each ap span is 20 % of rotor blade length, and location is
varied from 0.6R to 0.8R. As previously noted, vibration reduction is less with
the ap located at near the node of second ap mode. It is observed that phase
dierence between inboard and outboard single ap congurations is 110 degree.
Similar trends are observed in both dual- and multiple-ap congurations, as pre-
sented in Figure 5.20 showing 160 degree phase between inboard and outboard
aps. This clearly indicates that ap motions are nearly out-of-phase to eciently
excite the second apwise bending mode at inboard and outboard regions from the
node (x 0.785R). In all congurations, vibration index is reduced by 56 60%
as shown in Figure 5.20.
Flap prole for dual-ap conguration is presented in Figure 5.21. It is noted
that outboard ap deection (2nd ap) is less than inboard ap one due to the
high dynamic pressure. To investigate the eectiveness of multiple-ap congura-
tion, the hinge moments should be compared since both vibration reductions and
required maximum ap deections by single- and multiple-ap congurations are
similar. Normalized hinge moments for single- and dual-ap congurations are
shown in Figure 5.22. Maximum hinge moment of single-ap is 6.8 10
6
and
occurs at the azimuth angle of 342 degree. Maximum hinge moment of dual-ap
is 4.5 10
6
, and occurs at the azimuth angles of 225 and 342 degrees in 2nd and
1st aps, respectively.
Peak-to-peak values of hinge moments are presented in Table 5.4 showing that
the peak-to-peak hinge moment at 1st ap is a half of that at single-ap. Total
peak-to-peak hinge moment in dual-ap is, however, almost the same as in single-
ap.
152
Figure 5.19. Polar diagram of ap motion for single-ap conguration, advance ratio:
= 0.15, actuator saturation:
sat
f
= 2
o
Table 5.4. Peak-to-peak hinge moments in single- and dual-ap congurations with
4/rev control input, = 0.15
Flap conguration Peak-to-peak hinge moments Reduction, %
Single ap 8.04 10
6
-
Dual 1st ap 4.04 10
6
49.8
Dual 2nd ap 3.96 10
6
50.7
Dual (total) 8.00 10
6
0.47
153
Figure 5.20. Polar diagram of ap motion for dual-ap conguration, advance ratio:
= 0.15, actuator saturation:
sat
f
= 2
o
154
Figure 5.21. Flap deections of dual-ap conguration with 4/rev control input, ad-
vance ratio: = 0.15, actuator saturation:
sat
f
= 2
o
155
Figure 5.22. Hinge moments in single- and dual-ap congurations with 4/rev control
input, advance ratio: = 0.15, actuator saturation:
sat
f
= 2
o
156
5.3.6 Vibration Reduction with Multicyclic Control
The vibratory hub loads for the baseline, single-, dual- and multiple-ap cong-
urations are investigated. Trailing-edge ap deection limit is set to 4 degree for
fair comparison. Control input sequence for multiple-ap conguration is 3/rev,
4/rev and 4/rev from inboard to outboard. Vibratory hub loads for low-speed
ight condition are presented in Figure 5.23 showing that all congurations can
reduce vibration signicantly. Resulting ap proles for single- and multiple-ap
are presented in Figures 5.24 and 5.25. It is shown that outboard ap deections
are nearly 180 degree out-of-phase with inboard aps (1st and 2nd aps).
advance ratio = 0. 15
0.000
0.002
0.004
0.006
0.008
Fx Fy Fz Mx My Mz
Figure 5.23. Comparison of vibratory hub loads, advance ratio: = 0.15, actuator
saturation:
sat
f
= 4
o
For the high-speed ight, corresponding to an advance ratio of 0.35, the vibra-
tory hub loads are presented in Figure 5.26. Vibration reductions by single-, dual-
and multi-ap congurations are 83%, 85% and 79%, respectively. In this case,
in-plane hub shear force F
x
of multi-ap is higher than that of the others, since
5/rev control input is discarded to increase the ap eectiveness in both low- and
high-speed ight conditions. Flap proles of single- and multi-ap are presented in
Figure 5.27 showing that both 2nd and 3rd ap deections are nearly out-of-phase,
and their maximum is 1 degree.
157
0 90 180 270 360
5
4
3
2
1
0
1
2
3
4
5
F
l
a
p

d
e
f
l
e
c
t
i
o
n
s
,

d
e
g
r
e
e
s
Azimuth angle, degrees
Flap profile of singleflap, =0.15
Figure 5.24. Flap deections of single-ap conguration, advance ratio: = 0.15,
actuator saturation:
sat
f
= 4
o
158
0 90 180 270 360
5
4
3
2
1
0
1
2
3
4
5
F
l
a
p

d
e
f
l
e
c
t
i
o
n
s
,

d
e
g
r
e
e
s
Azimuth angle, degrees
Flap profile of single and multipleflap, =0.15
multi 1st TEF
multi 2nd TEF
multi 3rd TEF
Figure 5.25. Flap deections of multiple-ap conguration, advance ratio: = 0.15,
actuator saturation:
sat
f
= 4
o
159
advance ratio = 0. 35
0.000
0.001
0.001
0.002
0.002
0.003
0.003
Fx Fy Fz Mx My Mz
baseline
single (4 deg)
dual (4 deg)
mult i (4 deg)
Figure 5.26. Comparison of vibratory hub loads, advance ratio: = 0.35, actuator
saturation:
sat
f
= 4
o
160
Figure 5.27. Flap deections of single- and multiple-ap conguration, advance ratio:
= 0.35, actuator saturation:
sat
f
= 4
o
161
Once again it is observed that all congurations show similar performance in
terms of vibration index and ap deections. To explore dierences among various
ap congurations, the hinge moments along the azimuth are presented in Figure
5.28. In multi-ap conguration, maximum hinge moment is 5.210
6
and occurs
at the azimuth angle of 22 degrees. Although 3rd ap deections are much less
than that of the 1st ap, maximum hinge moment occurs at 3rd ap (outboard
ap) because of high dynamic pressure due to high-speed ight. In single-ap
conguration, maximum hinge moment is 5.2 10
6
and occurs at the azimuth
angle of 170 degree. Peak-to-peak values of hinge moments are presented in Table
5.5. Peak-to-peak hinge moment is reduced by 37.5% to 61.2% in each individual
actuator, while total peak-to-peak hinge moment increases by 45.3%.
Figure 5.28. Hinge moments in single- and multiple-ap conguration, advance ratio:
= 0.35, actuator saturation:
sat
f
= 4
o
162
Table 5.5. Peak-to-peak hinge moments in single- and multiple-ap congurations,
= 0.35
Flap conguration Peak-to-peak hinge moments Reduction, %
Single ap 1.06 10
5
-
Multiple 1st ap 4.10 10
6
61.2
Multiple 2nd ap 4.65 10
6
56.0
Multiple 3rd ap 6.60 10
6
37.5
Multiple (total) 1.53 10
5
-45.3
One of the merits of multiple-ap conguration is that the resonant actuation
concept can be applied to achieve high action authority, since vibration reduction
via single-ap could not be realized due to actuation limitations. This will be
discussed in Chapter 7.
163
5.4 Summary
In this chapter, an active control method for multiple trailing-edge aps congura-
tion is proposed. The concept involves deecting the trailing-edge aps by intro-
ducing the resonant actuation, in which each ap operates at a single frequency, in
a four-bladed hingeless rotor system. A proposed active vibration control strategy
is simulated for the steady-state forward ight condition. Based on the present
study, the following conclusions can be made:
Blade responses in terms of the number of beam elements are explored. The
eect of trailing-edge aps to the free-wake geometry is also investigated.
The number of beam elements lightly aects the free-wake geometry in the
high speed ight case, while the ap eect is very small because the tip
vortices generated by the ap itself are not considered in the present study.
A nite wing eect is not signicant in the multiple trailing-edge ap con-
guration, since the controller is able to compensate for changes in ap ca-
pability by adjusting the ap deections and phases, and kinetic energy also
contributes to vibration reduction.
For low-speed ight, vibration level is reduced by 56 60 % with a 4/rev
control input. Multiple-ap conguration can reduce the peak-to-peak hinge
moments by 49.8% and 50.7% at in- and out-board aps when compared to
the single-ap approach.
All ap congurations can reduce the vibration level by 80 85% for both
low- and high- speed ight conditions with multi-cyclic control inputs. For
high-speed ight, however, peak-to-peak hinge moment is reduced by 37.5%
to 61.2% in each individual actuator compared to single-ap conguration.
Chapter 6
Piezoelectric Actuation System
Synthesis
The dissertation to this point has discussed the vibration control and blade loads
control of helicopter rotor blades using active trailing-edge aps. A brief discussion
about the resonant actuation concept was given in Chapter 5. With this resonant
actuation concept, the helicopter vibration control and blade loads control could be
realized. The purpose of the research described in this chapter is to develop such
resonant actuation systems and to provide analytical tools. In the rst section, the
background and objective are presented. Next, a piezoelectric actuation system
model is derived for active ap rotors. Utilizing this model, mechanical tuning
and electrical tailoring methods are developed, where the optimal tuning param-
eters for electric networks can be explicitly determined. In the fourth section, an
equivalent electric circuit model emulating the physical actuation system is derived
and experimentally tested to investigate the initial feasibility of the piezoelectric
resonant actuation system. Finally, in the fth section, summary of this chapter
is presented.
165
6.1 Introduction
Because of their electro-mechanical coupling characteristics, piezoelectric materials
have been explored extensively for various engineering applications, where they are
often used as sensors or actuators. Some of the advantages of piezoelectric trans-
ducers include high bandwidth, high precision, compactness, and easy integration
with existing host structures to form smart structures. On the other hand, while
piezoelectric material-based actuators have shown good potential, they can only
provide limited strain. This limitation can be critical in cases where large actuator
strokes are required. The eorts to improve the piezoelectric actuator performance
have been made by researchers in developing amplication mechanisms of various
types (see Section 1.2.3 on Page 20). In general, these devices are still limited in
their performance.
The goal of this chapter is to develop a new method for actuation authority
enhancement to circumvent the aforementioned limitations of piezoelectric actu-
ators. The trailing-edge ap is used as a test bed for illustrating the concept.
The approach can be classied into two steps as outlined in the following para-
graphs. First, a selected resonant frequency of the actuation system (composed of
the piezoelectric actuator, the trailing-edge ap and the amplication mechanism,
and under the eect of unsteady aerodynamic loads) is tuned to the desired oper-
ating frequency through mechanical tailoring. It is well known that, for harmonic
control devices such as the trailing edge aps, if one can tune the natural frequen-
cies of the actuation system to the actuation frequencies, the actuation authority
can be greatly increased due to the mechanical resonance eect. In this case, one
can develop much lighter and smaller actuators to activate multiple and smaller
trailing-edge aps, each aiming at dierent operating frequencies of 3, 4, and 5/rev
(for a typical four-bladed rotor). While such a concept is indeed attractive, reso-
nant actuators could be very dicult to control and non-robust, due to its narrow
operating bandwidth. This is a critical bottleneck for realizing resonant actuation
system in practical applications.
To resolve this issue, the second step of this design process is to use electric cir-
cuitry tailoring to broaden and atten the resonant peak, so that one can achieve a
166
high authority actuation system with sucient bandwidth and robustness. In the
last decade, piezoelectric materials with electrical networks have been utilized to
create shunt damping for structural vibration suppression. It was also recognized
that such networks not only can be used for passive damping, they can also be de-
signed to amplify the actuator active authority around the tuned circuit frequency
(see see Section 1.2.4 on Page 26). Several researchers have proposed the nega-
tive capacitance concept to enhance the networks multiple mode and broadband
capabilities [129, 130]. The integration of the passive and active approach, often
referred to as an active-passive hybrid piezoelectric network, has shown to achieve
promising results in vibration control [125]. The electric network tailoring idea
proposed in this paper is built upon these previous investigations, but is based on
completely dierent design philosophy and criterion.
To demonstrate the proposed concept, the piezoelectric tube actuator devel-
oped by Centolenza et. al [116] is selected as a test bed for this study. Coupled
piezoelectric actuator, trailing-edge ap and electric network system equations are
derived. The proposed method is then analyzed numerically and veried experi-
mentally via an equivalent electric circuit based on the Van Dyke model.
167
6.2 Piezoelectric Actuation System Model
In this section, a piezoelectric actuation system model is developed for active ap
rotors. Fully coupled PZT actuator-ap-circuit system equations are derived via
the variational principle.
Figure 6.1. A piezoelectric tube actuator conguration
The PZT tube shear actuator (Figure 6.1), which uses the shear deformation
of piezoelectric materials, is assembled with the piezoelectric ceramic segments of
alternating poling signs, and then the accumulation of shear strain around the
circumference produces the angle of twist. As opposed to conventional piezoelec-
tric actuators, which are poled in the thickness direction, the PZT tube actuator
segments are poled in the length direction [116].
6.2.1 Piezoelectric Tube Actuator
The potential energy contained in piezoelectric materials is described by a function
called the electric enthalpy density function H in the linear piezoelectricity (IEEE
1998) [149]
H(S
ij
, E
i
) =
1
2
c
E
ijkl
e
kij
E
k
S
ij

1
2

S
ij
E
i
E
j
, (6.1)
and in the matrix form
H(S, E) =
1
2
S
T
[c
E
]S S
T
[e]E
1
2
E
T
[
S
]E. (6.2)
168
The constitutive equations for a linear piezoelectricity are given by
T
ij
=
H
S
ij
= c
E
ijkl
S
kl
e
ijk
E
k
, (6.3)
D
i
=
H
E
i
= e
ikl
S
kl
+
S
ik
E
k
, (6.4)
or
T = [c
E
]S [e]E, (6.5)
D = [e]
T
S + [
S
]E. (6.6)
It can also be expressed in terms of strains and electric displacements.
T = [c
D
]S [h]D, (6.7)
E = [h]
T
S + [
S
]D. (6.8)
where
[c
D
] = [c
E
] + [e][
S
][e]
T
, (6.9)
[h] = [e][
S
], [
S
] = [
S
]
1
. (6.10)
Here, nomenclature follows that of ANSI/IEEE standard 176-1987 on piezoelec-
tricity [149].
The governing equations are derived using the variational principle, assuming
no body force, as follows:
_
t
0
_
Vp
(
p
u
i
u
i
T
ij
S
ij
E
i
D
i
) dV
p
dt
+
_
t
0
_
S
p
(p
i
u
i
+
s
Q
s
) dS
p
dt = 0, (6.11)
where T
ij
is the mechanical stress component, S
ij
the strain component, E
i
is the
electric eld component, D
i
is the electric displacement component, and u
i
denotes
the mechanical displacement component. p
i
, Q
s
, and
s
represent the applied
traction, charge density, and applied electric potential on the surface, respectively.
169

p
denotes the piezoelectric material density.
For the piezoelectric tube actuator, Equation 6.11 can be written by, in the
polar coordinate system,
_
t
0
_
v
p
_

p
u

T
z
S
z
E

N
s
_
dv
p
dt
+
_
t
0
_
s
p
_
p

+
s

Q
N
s
A
s
_
ds
p
dt = 0, (6.12)
where N
s
is the number of segments of tube actuator and Q is the electric charge.
The angular displacement u

, shear strain S
z
, and the electric displacement D

are given by
u

= r
p
, S
z
= r
p

,z
, D

=
Q
A
s
, (6.13)
in which ()
,z
denotes the partial derivative with respect to z, and A
s
represents
the surface area of the PZT electrode.
The equations motion can be discretized by applying the assumed mode method.
Then the angular displacement of the piezoelectric tube actuator is given by
u

(r
p
, z; t) = r
p
(z; t) = r
p

i=1

i
(z)q
i
(t), (6.14)
where
i
(z) is the i-th mode shape function, and q
i
(t) is the i-th generalized
displacement.
Substituting Equations 6.7, 6.8, 6.13 and 6.14 into Equation 6.12 yields
_
M
p
0
0
T
0
__
q

Q
_
+
_
C
p
0
0
T
0
__
q

Q
_
+
_
K
D
p
K
c
K
T
c
K
Q
__
q
Q
_
=
_
F

V
a
_
,
(6.15)
where
M
p
=
_
l
p
0
_
2
0
_
R
o
R
i
_

p
r
2
p

j
_
r
p
dr
p
ddz, (6.16)
K
D
p
=
_
lp
0
_
2
0
_
Ro
R
i
_
c
D
55
r
2
p

i,z

j,z
_
r
p
dr
p
ddz, (6.17)
170
K
c
=
_
l
p
0
_
2
0
_
R
o
R
i
_
h
15
r
p

i,z
A
s
N
s
_
r
p
dr
p
ddz, (6.18)
K
Q
=
_
lp
0
_
2
0
_
Ro
R
i
_

S
11
A
2
s
N
2
s
_
r
p
dr
p
ddz, (6.19)
F

=
_
2
0
_
R
o
R
i
_
h
15
r
p

i,z
A
s
N
s
_
r
p
dr
p
d

z=lp
, (6.20)
V
a
=
_
s
_

s
N
s
A
s
_
ds = N
s
_

s
A
s
A
s
N
s
_
=
s
, (6.21)
in which l
p
is the length of the tube actuator, R
o
and R
i
denote the outer radius and
inner radius of the tube (see Figure 6.1). The matrix C
p
represents the structural
damping of piezoelectric tube actuator, K
Q
denotes the inverse of piezoelectric tube
actuator capacitance C
S
p
, and V
a
is the voltage across the segment of piezoelectric
tube actuator.
For the piezoelectric tube actuator, a single-mode approximation is accurate
enough to predict the actuation system response, since the operating frequencies
are far below than those of the tube actuator. The tube actuator is modeled as
a clamped-free hollow cylinder (see Figure 6.1). The rst eigenfunction of the
clamped torsion bar is given by
(z) = sin
_

2l
p
z
_
, (6.22)
and then, with assuming no external torque, Equation 6.15 can be rewritten by
_
M
p
0
0 0
__
q
t

Q
_
+
_
C
p
0
0 R
p
__
q
t

Q
_
+
_
K
D
p
K
c
K
c
K
Q
__
q
t
Q
_
=
_
0
V
a
_
, (6.23)
where q
t
represents the tip displacement of a piezoelectric tube actuator, which is
also referred to as the actuator stroke, and R
p
represents the electric resistance
that includes both the inherent resistance of the tube actuator and the external
wire resistance.
171
6.2.2 Inertial and Aerodynamic Loads
The aerodynamic model for the trailing edge ap used in this study is based on
the incompressible thin airfoil theory developed by Theodorsen [143], although
the compressible unsteady aerodynamic model, which was presented in Section
3.2.2 on Page 98, is available. Since the Theodorsens theory renders the explicit
formulation of aerodynamic loads, it is adequate for the purpose of the present
study.
The total hinge moment is comprised of the aerodynamic, inertial and cen-
trifugal propeller moments due to the rotation of blade (see Figure 6.2). Assuming
that the trailing-edge ap deection angle,
f
, is small, the inertial and propeller
moments can be expressed as [107]:
h
I
= I

f
, (6.24)
h
CF
= m
f

2
dr
I
sin(
f
) m
f

2
dr
I

f
, (6.25)
where
f
is a trailing edge ap deection angle, and is the rotation speed of
blade. I

, m
f
, d and r
I
are the ap mass moment of inertia, the ap mass unit
per length, the oset from blade elastic axis to ap hinge and the ap chordwise
c.g. after ap hinge, respectively.
Figure 6.2. Forces and moments acting on the trailing-edge ap
The aerodynamic contribution due to trailing edge ap can be obtained from
the incompressible aerodynamic model, which was presented in Section 3.2.1 on
172
Page 97. The hinge moment coecients in terms of ap deection angle
f
is given
by
C
H
=
1
2
T
12
T
10
C(k) +
1
2
(T
4
T
10
T
5
) , (6.26)
C
H

=
b
4U

T
12
T
11
C(k) +
b
4U

T
4
T
11
, (6.27)
C
H

=
b
2
2U
2

T
3
, (6.28)
where coecients, T
3,4,5
and T
10,11,12
, are the geometric parameters dened by
Theodorsen [143], U

and b are the relative wind velocity at the radial location of


the blade and the blade semi-chord, respectively. The Theodorsens lift deciency
function C(k) is a complex coecient that can be expressed in terms of Bessel
functions.
Once all the coecients are found, aerodynamic and inertial contributions can
be expressed in terms of the trailing-edge ap deection angle
f
:
F
f
(
f
; t) = M
f

f
+C
f

f
+K
f

f
, (6.29)
where
M
f
=
1
2
C
H

U
2

c
2
L
f
+h
I
, (6.30)
C
f
=
1
2
C
H

U
2

c
2
L
f
, (6.31)
K
f
=
1
2
C
H

U
2

c
2
L
f
+h
CF
, (6.32)
where c and L
f
are the blade chord and the trailing-edge ap span, respectively.
Underline terms denote inertial contributions from the ap.
6.2.3 Coupled Actuator-Flap-Circuit System
In this section, the coupled actuator-ap-circuit equations are derived to describe
the integrated actuation system. For the single-frequency electric networks (see
Figure 6.3), a series R-L-C circuit is used and the resulting piezoelectric network
173
equation can be written as
L

Q+R

Q+

K
Q
K
c
q
t
= V
c
, (6.33)
or
L

Q+R

Q+V
a
+K
a
Q
Q = V
c
, (6.34)
where L is the inductance, R is the resistance, V
a
and V
c
denote the voltage across
the PZT actuator and the control voltage, respectively. V
a
is expressed by
V
a
= K
Q
QK
c
q
t
, (6.35)
and

K
Q
is dened by

K
Q
= K
Q
+K
a
Q
, (6.36)
K
Q
=
1
C
s
p
, K
a
Q
=
1
C
add
, (6.37)
in which C
s
p
denotes the capacitance of the PZT tube actuator, and C
add
represents
the added capacitance, which can be either positive or negative.
R-L
Figure 6.3. Schematic of the PZT tube with R-L circuit and negative capacitance
In order to integrate the PZT tube actuator and the ap system, the trailing
edge ap deection angle,
f
, should be expressed in terms of the actuator stroke,
q
t
. This relation can be interpreted as the amplication mechanism, i.e., the linkage
174
from the PZT tube actuator to the ap device. The design of this linkage is not a
trivial issue, which is beyond the scope of present study. Thus, the simple fulcrum
type of amplication mechanism is used to illustrate the concept. It is assumed
that the relation between
f
and q
t
can be expressed in the form (see Figure 6.4)

f
= A
M
q
t
. (6.38)
Note that this relationship, A
M
, plays a role of unit conversion. For instance, the
trailing-edge ap deection d
f
is transformed to the actuator stroke q
t
of the PZT
actuator. On the other hand, using the same relationship, the hinge moments are
transformed into either forces or moments, depending on the actuator types. In
the tube actuator case, the hinge moment is transformed into the torque.
Substituting Equation 6.38 into Equation 6.29 and combining Equation 6.23
and Equation 6.33 yield
_
M 0
0 L
__
q
t

Q
_
+
_
C 0
0 R
__
q
t

Q
_
+
_
K
D
K
c
K
c

K
Q
__
q
t
Q
_
=
_
0
V
c
_
, (6.39)
where
M = M
p
+M
f
A
2
M
,
C = C
p
+C
f
A
2
M
, (6.40)
K
D
= K
D
p
+K
f
A
2
M
,
where the open-circuit stiness K
D
can be also expressed, in terms of the short-
circuit stiness K
E
, as:
K
E
= K
D

K
2
c
K
Q
, (6.41)
or

K
E
= K
D

K
2
c

K
Q
, (6.42)
where K
E
represents the nominal short-circuit stiness, which is normally Youngs
modulus, while

K
E
indicates the situation that the circuit is shorted between the
electrode of the PZT and that of the added capacitance.
175
PZT tube
K
f
Flap
d =
f t
q A
M
q
t
d
f
R
o
R
i
Figure 6.4. Fulcrum amplication mechanism for the PZT tube actuator
6.3 Mechanical Tuning and Electrical Tailoring
In this section, mechanical tuning and electrical tailoring are described. A mechan-
ical tuning is needed to tune the resonant frequency of actuation system to the
operating frequency. Then electrical tailoring is applied to the actuation system
to enhance the actuator operating bandwidth and phase control. After mechanical
tuning, the actuation system together with an electric network is referred to as a
resonant actuation system (RAS).
6.3.1 Mechanical Tuning
In the present study, structural resonance of the actuation system is utilized to
increase the active authority. The stiness and/or the mass of the coupled system
could be adjusted to tune the resonant frequency of the actuation system to around
the desired operating frequency.
It is assumed that the stiness and mass of the PZT tube actuator are xed,
and the modal stiness and mass of the trailing edge ap and the amplication
mechanism can be adjusted. Tuning the mass will be more eective than tuning
the stiness, because the mass of the coupled system mostly comes from the aero-
dynamic contribution (the ap mass moment of inertia in Equation 6.24). There
are two ways to adjust the mass term of the aerodynamic loads. One is to add
concentrated mass to the trailing edge ap. This is the simplest method to adjust
the mass term, but it may cause aeromechanical instability [31] due to the shift in
176
C.G. The other is to design the overhang of the trailing edge ap [84, 85]. This will
also be eective, but the design detail is beyond the scope of the present study. In
this study, the mechanical tuning is achieved by the former method.
In the present study, the simple mass tuning is adopted to achieve the mechan-
ical tuning by adding the tuning mass to the trailing-edge ap. This will aect the
ap inertial load, mainly the ap mass moment of inertia, I

, given in Equation
6.24.
6.3.2 Electrical Tailoring
It is convenient to express the actuation system equation in the nondimensional-
ized form for the derivation of electrical tailoring parameters, such as the optimal
inductance and resistance tuning ratios, which will be derived based on the transfer
functions.
From Equation 6.39 and neglecting the added capacitance (i.e., K
a
Q
= 0), the
transfer function between actuator stroke and control voltage and that between
electric charge and control voltage are, respectively,
q
t
V
c
=
K
c
(
2
M +jC +K
D
)(
2
L +jR +K
Q
) K
2
c
, (6.43)
Q
V
c
=

2
M +jC +K
D
(
2
M +jC +K
D
)(
2
L +jR +K
Q
) K
2
c
, (6.44)
and in the nondimesionalized form
q
t
V
c
=

2
(1 +
2
+ 2j
2
)(rj +
2

2
)
2

2
_
q
t
V
a
_
ST
, (6.45)
Q
V
c
=

2
(1 +
2
+ 2j +
2
)
(1 +
2
+ 2j
2
)(rj +
2

2
)
2

2
_
1
1 +
2
__
Q
V
a
_
ST
, (6.46)
where

2
E
=
K
E
M
,
2
c
=
K
Q
L
, =

E
,
2
=
K
2
c
K
E
K
Q
,
=
C
2M
E
, r =
R
L
E
, =

c

E
, (6.47)
177
and subscripts ST represents the static response that are given by
_
q
t
V
a
_
ST
=

2
K
c
, (6.48)
_
Q
V
a
_
ST
=
1 +
2
K
Q
. (6.49)
Here r and are often referred to as the resistance and inductance tuning ratios,
and is referred to as the generalized electro-mechanical coupling coecient, which
is also expressed as:

2
=
(K
E
+K
2
c
/K
Q
) K
E
K
E
=
K
D
K
E
K
E
=

2
D

2
E

2
E
. (6.50)
Now the open-circuit frequency can be expressed, in terms of the generalized
electro-mechanical coupling coecient and the short-circuit frequency, as:

2
D
= (1 +
2
)
2
E
. (6.51)
It is well known that there are optimal inductance and resistance tuning ratios
for the passive shunt circuit conguration proposed by Hagood and Flotow [118].
The optimal values for the shunt inductance and resistance can be derived via a
mechanical vibration absorber analogy. These optimal values would be dierent
from those for the best voltage driving responses when the system is used for
actuation. In this chapter, the explicit formulae of the optimal tuning ratios for
voltage driving responses are derived.
Unlike the passive shunt damping system, there is only one invariant point of
the voltage driving frequency response curve, which is at the open-circuit frequency.
In this case, the optimal inductance tuning strategy is developed by deriving the
inductance value such that the transfer function in Equation 6.45 has a station-
ary value with respect to frequency at this invariant point =
_
1 +
2
that is
the open-circuit frequency normalized with the short-circuit frequency. The opti-
mal tuning ratios for the voltage deriving response can be derived based on the
178
frequency function approach [125]. The modal damping due to the actuator and
trailing-edge ap aerodynamics is neglected in the following derivation.
The frequency response function of the piezoelectric actuator with the series
R-L-C circuit is given in Equation 6.45. This can be rewritten, without the me-
chanical damping, as follows:
q
t
V
c

=0
=

2
(1 +
2

2
)(rj +
2

2
)
2

2
_
q
t
V
a
_
ST
. (6.52)
This function has an invariant value at
2
= 1 +
2
with respect to the electric
resistance tuning ratio r. The optimum inductance-tuning ratio can be found if
Equation 6.52 has its stationary value at the invariant point. The derivative of
Equation Equation 6.52 with respect to the normalized frequency square
2
is
given by


2
_
q
t
V
c

=0,r=0
_


2
=1+
2
= 0, (6.53)
which yields
_
(1 +
2
)(
2

2
) (1 +
2

2
)


2
=1+
2
= 0. (6.54)
From Equation 6.54, the optimal inductance tuning ratio

can be found by

=
_
1 +
2
, (6.55)
which is the normalized open-circuit frequency and the same as that of passive
shunt damping [118].
The voltage driving frequency response function for any given resistance tuning,
r, will pass through the invariant point, =
_
1 +
2
. The optimal value of r,
which will best atten the frequency response function near the resonant frequency,
can be found by equating the magnitude of the frequency response at =

(open-circuit frequency) and that of the resonant frequency, = 1 (short-circuit


179
frequency),
_
q
t
V
c
_

=0,
2
=1+
2
=
_
q
t
V
c
_

=0,
2
=1
, (6.56)
which yields
r

2
_
1 + 0.5
2
. (6.57)
This optimal resistance tuning ratio, however, should be calculated by consid-
ering the actuation system damping, i.e. ,= 0, because the actuator stroke at
the vicinity of the resonant frequency strongly aected by the system damping.
An iterative method (e.g., a bisection method) can be used to solve the following
nonlinear equation to nd the optimal resistance tuning ratio in practice.
_
q
t
V
c
_


2
=1+
2
=
_
q
t
V
c
_


2
=1
. (6.58)
When the additional capacitance is added in series to the piezoelectric material,
and with a R-L circuit, which is presented in Equation 6.39, the transfer function
between actuator stork and control voltage is
q
t
V
c
=
K
c
(
2
M +jC +K
D
)(
2
L +jR +

K
Q
) K
2
c
, (6.59)
which can be written in the dimensionless form using non-dimensionalized param-
eters dened in Equation 6.47,
q
t
V
c
=

2

2
/
2
(1 +
2
+ 2j
2
)(rj +
2

2
)
2

2
_
q
t
V
a
_
ST
, (6.60)
where a hat

() indicates the system response with the added capacitance, the
modied coupling coecient

is dened by

2
=
K
2
c
K
E
K
Q
. (6.61)
180
The short-circuit frequency, which includes the added capacitance, is given by

2
E
= (1 +
2

2
)
2
E
. (6.62)
Then, in the same manner described for the system without added capacitance,
the optimal inductance and resistance tuning ratios are obtained, respectively, by

=
_
1 +
2
, (6.63)
r

1 +
2
0.5

2
1 +
2

2
. (6.64)
Design guidelines and physical insights for the developed resonant actuation system
will be discussed in the Section 7.1 of Chapter 7 in detail.
6.4 Equivalent Electric Circuit Model
To realize the proposed resonant actuation system and examine its feasibility for
the proof-of-concept, an equivalent electric circuit model, which is based on the
Van Dyke model, is derived and tested experimentally.
6.4.1 Van Dyke Model
The mechanical and electrical parts of piezoelectric materials are connected through
a conversion factor [150]. The electrical property of the piezoelectric actuator can
be modeled as a single capacitor, as normally assumed in piezoelectric network
analysis. The modal stiness of the resonant actuation system is governed by that
of the PZT tube actuator, although natural frequencies of the PZT tube actuator
are relatively high, while the modal mass of the resonant actuation system is gov-
erned by trailing edge ap contribution. Therefore, the mechanical stiness of the
PZT tube actuator should be considered to properly emulate the physical system.
The PZT tube actuator for trailing edge ap operates at 3/rev, 4/rev, and
181
5/rev of the main rotor, which are relatively low frequencies when compared to
natural frequencies of the PZT tube actuator. The mechanical part of the actuator
and aerodynamic and inertial loads of trailing-edge ap adding together can be
considered as a single degree of freedom system. When the negative capacitance,
C
n
, and series inductor-resistor circuit, R and L, are considered, the equivalent
electric circuit model of the resonant actuation system can be represented in Figure
6.5. The rst and second branch together can be considered as the Van Dyke model
that was introduced to model the piezoelectric materials [151].
Figure 6.5. Equivalent electric circuit model of the resonant actuation system
The dynamic equations of the equivalent electric circuit model shown in Figure
6.5 can be written as
L
m

Q
2
+R
m

Q
2
+
_
1
C
m
+C
e
p
_
Q
2

1
C
e
p
= 0, (6.65)
L

Q+R

Q+
1
C
e
p
Q
1
C
e
p
Q
2
= V
c
, (6.66)
where the electric charge Q
2
in the rst branch emulates the actuator stroke of the
PZT tube actuator. The stiness, damping, and mass of the mechanical part of the
182
PZT actuator and trailing-edge ap contributions are represented by capacitor C
m
,
resistor R
m
, and inductor L
m
, respectively. C
e
p
represents the total capacitance,
C
e
p
=
C
n
C
s
p
C
s
p
+C
n
. (6.67)
To use this equivalent electric circuit model, the coecients of the Van Dyke
model (C
m
, R
m
, L
m
) should be determined. It can be resolved by measuring the
electrical impedance of the resonant actuation system directly [150]. The equiv-
alent circuit is mainly introduced to experimentally verify the proposed resonant
actuation system that includes two active circuit components (synthetic inductor
and negative capacitor). These coecients can be determined through comparing
the resonant frequencies between the two systems. There are two types of frequen-
cies in the Van Dyke model, parallel and series frequencies (
p
,
s
) that are given
by

2
p
=
1
L
m
C
m
+C
e
p
C
m
C
e
p
=
K
D
M
=
2
D
, (6.68)

2
s
=
1
L
m
C
m
=
K
D
K
2
c
/

K
Q
M
=

K
E
M
=
2
E
. (6.69)
From the equivalence of these two frequencies, the following relationships can be
obtained.
L
m
=
1
C
e
p
1
(
2
p

2
s
)
, (6.70)
R
m
= L
m
C
M
, (6.71)
C
m
=
1
L
m

2
s
, (6.72)
K
c
=

M
C
e
p
_

2
p

2
s
_
. (6.73)
These clearly show that the coupling stiness K
c
of the resonant actuation system
is proportional to the square root of the dierence between the electro-mechanically
coupled and de-coupled frequencies, which can be increased by increasing the ef-
fective capacitance (using the negative capacitance).
183
6.4.2 Analysis and Experimental Verication
In order to evaluate the resonant actuation system for a trailing-edge ap, a Mach-
scaled helicopter rotor blade is considered as an example. The rotor has a blade
diameter of 6 feet, a blade chord of 3 inches, and a nominal rotation of 2000 RPM.
For comparison purposes, a baseline actuation system is rst dened: the length
of the baseline PZT tube actuator is 4 inches, outer radius R
o
is assumed to be
0.175 inches, the tube wall thickness is 0.1 inches, the ap length is 20% of rotor
length, 7.2 inches, and the amplication ratio is assumed to be 13. To demonstrate
the improvement of active authority, a smaller 4/rev actuator is considered for the
new resonant actuation system. The length of this smaller actuator is 2 inches,
which is one-half of the baseline actuator, the trailing edge ap length is 2.4 inches
(33% of the baseline), and the other dimensions are the same as the baseline. The
piezoelectric material properties used in this study are listed in Table 6.1.
Table 6.1. Piezoelectric material properties of PZT-5H for a Mach-scaled rotor
Piezoelectric constant Value
Charge constant, d
15
741 m/V or C/N (10
12
)
Voltage constant, g
15
26.8 V m/N (10
3
)
Elastic constant (open-circuit), c
D
55
2.3 N/m
2
(10
10
)
Dielectric constant,
T
11
2.765 F/m (10
8
)
Piezoelectric constant, h
15
6.164 V/m or N/C (10
10
)
Impermittivity,
S
11
5.865 m/F (10
7
)
When appropriate amplication ratio and mechanical tuning frequency (usu-
ally, 3, 4 or 5 /rev of the main rotor speed) have been determined, the frequency
response between trailing-edge ap deection and control voltage is evaluated,
where the trailing-edge ap deection is calculated by Equation 6.38. In the present
study, the mechanical resonant tuning is achieved by adjusting the mass moment
of inertia of the trailing edge ap.
In Figure 6.6, the voltage driving and passive damping frequency responses are
plotted, under the optimal resistance-tuning ratio for voltage driving performance.
The vertical axis denotes the trailing edge ap deections. The solid line indicates
184
the voltage driving response of the baseline actuation system. The dotted and dash-
dotted lines denote the voltage driving and passive damping responses, respectively,
with both negative capacitor (C
n
= 2.2C
s
p
) and shunt circuit. For the voltage
driving responses, an electric eld of 4 kV/cm is applied. For the passive damping
response, the excitation force is assumed to be the same moment as that produced
by the 4 kV/cm electric eld. The resonant actuation system is tuned to the
4/rev (120 Hz) by modifying the mass moment of inertia of trailing edge ap
and amplication ratio. Comparing the baseline and the new resonant actuation
systems, once can see that the active authority is signicantly increased from 3
degree to 8 degree, as shown in Figure 6.6. The ap deections around the resonant
frequency are almost constant (at plateau) and the bandwidth reasonably large
(approximately 80 Hz).
Figure 6.6. Trailing-edge ap deections of the resonant actuation system for Mach-
scaled rotor
An experimental investigation is performed on the resonant actuation system.
The equivalent electric circuit model is realized using the synthetic inductance and
the negative impdedance converter of capacitance, as shown in Figure 6.7. The
mechanical part of the PZT tube actuator and aerodynamic and inertial loads due
to the trailing edge ap are emulated by a single degree of freedom series R
m
-
185
L
m
-C
m
circuit (left branch in Figure 6.7). This branch and the middle branch
(representing the PZT tube capacitance and the negative capacitor) in Figure
6.7 form the Van Dyke model. The coecients of this model are calculated by
Equations 6.706.72, where the resonant frequency of the left branch is tuned to
4/rev (120Hz). The voltage across the capacitor C
m
is measured, and an HP35665A
dynamic signal analyzer is used to extract the frequency response.
Figure 6.7. Realization of the equivalent electric circuit for the resonant actuation
system
To create the negative capacitance, the negative impedance converter of capac-
itance [128] is used. Since the negative capacitance is non-oating, it should be
grounded. In this system, large inductance is required to emulate the trailing-edge
ap and actuator dynamics and to create the needed inductance in the R-L circuit.
Thus a synthetic inductor [122], which consists of four operational ampliers, is
used.
The piezoelectric capacitance is 8 nF, the negative capacitance is -18.4 nF (-
2.2 C
s
p
) and the capacitance C
m
and inductance L
m
of the left branch in circuit
diagram shown in Figure 8 are 6.8 nF and 285 H. The inductance of shunt circuit is
95 H. For this conguration, theoretical optimal resistance for the voltage driving
186
performance is estimated to be 83 kilo Ohms. However, only 40 kilo Ohms are
added to the circuit since there is signicant internal resistance in the actual circuit
components.
The frequency response of the equivalent electric circuit system with the well-
tuned parameters is compared to that of analytical prediction, as shown in Figure
6.8, where the experimental results (with 1 volt white noise input) are scaled
up to match with analytical results. Clearly, the test results match the theoretical
prediction quite closely, implying that the high authority robust resonant actuation
system can be realized through the proposed hardware and circuits (synthetic
inductance and negative capacitance).
Figure 6.8. Comparison of analytical and experimental results of the RAS for a Mach-
scaled rotor
187
6.5 Summary
In this chapter, a new approach is proposed to enhance the actuation authority
of piezoelectric actuation systems. The idea is to rst achieve a resonant driver
through using mechanical tuning, and then increase the bandwidth and robustness
of the resonant actuation system through electrical network tailoring. A Mach-
scaled piezoelectric tube actuator-based trailing-edge ap for helicopter vibration
control is used as an example to illustrate the proposed concept.
A coupled PZT actuator, trailing edge ap and electrical circuit dynamic model
is derived. Utilizing the model, the required electrical circuitry parameters are de-
termined. The inductance is rst tuned such that the piezoelectric shunt frequency
matches the mechanical resonant frequency. Negative capacitance is then used to
broaden the actuation authority bandwidth around the operating frequency. To
further enhance the robustness, the optimal resistance tuning ratios are derived
such that the actuation authority frequency response near the resonant frequency
can be attened.
The negative capacitance and inductance are realized and implemented us-
ing operational amplier-based circuitry. An experiment is set up using equivalent
circuit representing the integrated structure-actuator-network system. It is demon-
strated that the proposed resonant actuation system can indeed achieve both high
active authority and robustness.
In this chapter, the electric network with low voltage excitation was realized
for the case of a Mach-scaled rotor as the feasibility verication of the resonant
actuation system. For full-scale helicopter, however, a high voltage is needed to
produce the torque that should overcome strong aerodynamic hinge moments. This
issue will be addressed in the following chapter.
Chapter 7
Design and Test of Resonant
Actuation Systems
While investigations on the electro-mechanically tailored piezoelectric resonant
actuation system have shown promising results (Chapter 6), there are still research
issues to be addressed before such a concept can be realized. One of them is that
there is no systematic method for tailoring the electrical parameters of the RAS
circuitry network, such that desired actuator authority can be achieved. Thus,
in the rst section of this chapter, an analytical approach is carried out to derive
design guidelines for the RAS circuitry in dimensionless forms. A mechanically
tuned resonant actuator is analyzed based on the previous derivation in Chapter
6 and compared to an equivalent mechanical system to provide better physical
understanding. In the second section, dynamic characteristics of the RAS will
be examined. Vibration reduction performance of various ap congurations is
evaluated within the available actuation authority. In the third section, a better
method for implementing the electrical circuitry is proposed to realize the actuation
system, especially under high voltage operations. In the fourth section, the electric
power consumption of the piezoelectric RAS is quantied to evaluate the proposed
resonant actuation system. Finally, the eorts and ndings of this chapter are
summarized.
189
7.1 Design Guidelines of the RAS
In this section, an analytical approach is carried out to derive design guidelines
for the RAS circuitry in dimensionless forms. A mechanically tuned resonant
actuator is analyzed based on the previous derivation in Chapter 6 and compared
to an equivalent mechanical system to provide better physical understanding.
7.1.1 Resonant Actuators with R-L elements
Given that the electrical property of a piezoelectric actuator is similar to a capaci-
tor, Figure 7.1 illustrates the equivalent dynamics of the piezoelectric material with
a series R-L circuitry. The dynamic system equations are presented by Equation
6.45 on Page 176 in Chapter 6.
R
C
p
V
p
V
c
L
PZT
V
a
Figure 7.1. A piezoelectric network with a series R-L circuit
For this conguration, the voltage driving frequency responses of resonant
piezoelectric actuators with a series R-L circuit are shown in Figure 7.2, where
results are plotted with respect to the electro-mechanical coupling coecient .
The frequency responses under the optimal tuning ratios show a at shape around
the resonant frequency. Magnitude at the two points (short- and open-circuit fre-
quencies) under optimal tuning ratios (Equations 6.55 and 6.57 on Page 178) could
be used as the performance index of the resonant piezoelectric actuator, that is
_
q
t
V
c
_
=1, =

1+
2
=
1

2
_
q
t
V
a
_
ST
, (7.1)
190
where
_
q
t
V
a
_
ST
=

2
K
c
, (7.2)
which was presented in Equation 6.48 on Page 177.
0 0.4 0.8 1.2 1.6 2
20
10
0
10
20
30
40
50
Frequency response of q
t
/V
c
(dB)
Nondimensionalized Frequency
baseline
=0.2
=0.4
=0.6
Figure 7.2. Actuator strokes with the optimal tuning ratios and various coupling
coecients
In general, larger electro-mechanical coupling coecient, , yields better ac-
tuator performance, especially for the static case. There is, however, a trade-
o between actuator authority and bandwidth, as shown in Figure 7.2. From
Equation 7.1, one can see that the actuation authority at the operating frequency
(near resonant frequency) relative to the static value will decrease with increasing
electro-mechanical coupling coecient. On the other hand, increasing the electro-
mechanical coupling coecient will increase the bandwidth. Here, the bandwidth
is dened by
B

=

D

E
=
_
1 +
2
1 =
1
2

2
(1
1
4

2
). (7.3)
Variations of the stroke and bandwidth with dierent generalized coupling co-
191
ecients under the optimal tuning ratios are presented in Figure 7.3, where the
stroke and bandwidth are normalized with their maximum values. One can nd
that there is the generalized coupling coecient where the two curves intersect.
Coupling coecients near this value could yield the best balanced performance in
terms of both stroke and bandwidth. Thus, Equations 7.17.3 can serve as design
criteria for the resonant actuation systems.
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
0
0.2
0.4
0.6
0.8
1
Coupling coefficient,
log(q
t
/Vc)
B

Figure 7.3. Actuator stroke and bandwidth variations with coupling coecients
Figure 7.4 shows the actuator stroke with dierent resistance tuning ratios,
where the electro-mechanical coupling coecient is assumed to be 0.4. The fre-
quency response exhibits two resonant peaks when the resistor is below its optimal
value. As the resistance tuning ratio is increased, these two peaks coalesce into
a single peak which corresponds to the open-circuit frequency
D
. This can be
clearly observed considering the equivalent mechanical system (Figure 7.5).
The equations of motion for an equivalent mechanical system shown in Figure
7.5 is given by
_
M
M
0
0 m
m
__
x
e
y
e
_
+
_
C
M
0
0 c
m
__
x
e
y
e
_
+
_
K
M
+k
m
k
m
k
m
k
m
__
x
e
y
e
_
=
_
0
f
e
_
, (7.4)
192
0 0.4 0.8 1.2 1.6 2
20
10
0
10
20
30
40
50
Frequency response of q
t
/V
c
(dB)
Nondimensionalized Frequency
baseline
r=0.01
r=r
*
r=1
Figure 7.4. Actuator strokes with the optimal inductance tuning and various resistance
tuning values, = 0.4
where subscripts M and m represent the structural and electrical properties of the
piezoelectric actuation system, respectively. The capital M denotes the mechanical
modal mass (actuator and ap inertia). The small m represents the inductor in
the circuit. The displacements x
e
and y
e
correspond to the actuator stroke q
t
and
electric charge Q, respectively. The small spring k
m
represents both the electro-
mechanical coupling and the inverse of the piezoelectric actuator capacitance. C
M
and c
m
correspond to the mechanical damping and electric resistance, respectively.
f
e
represents the voltage source V
c
.
Note that the damper c
m
applied on the small mass m
m
is directly connected
to the ground, unlike the classical mechanical vibration absorber case [152]. Here,
the open- and short-circuit resonant frequencies can be explained in terms of the
electrical damping (i.e., resistor). Assuming m
m
= 0, the equivalent mechanical
system shown in Figure 7.5 becomes a one-degree-of-freedom system. If the damp-
ing c
m
is zero, the displacement x
e
is not aected by the stiness k
m
. The system
193
M
M
m
x
y
K
M
k
m
C
M
c
m
f(t)
m
m
Figure 7.5. Schematic of an equivalent mechanical system to a piezoelectric actuation
system
is analogous to the short-circuit situation, and the resonant frequency will become:

E
=
_
K
M
M
M
. (7.5)
On the other hand, if the damping c
m
is innity, the displacement x
e
is aected by
both springs K
M
and k
m
, which represents the behavior of piezoelectric actuator
under the open-circuit condition, with the resonant frequency

D
=
_
K
M
+k
m
M
M
. (7.6)
In the case that the small mass m
m
is present, the equivalent mechanical system
will have two degrees of freedom and two resonant peaks. These two peaks will
coalesce into a single resonant peak, as the damping c
m
is increased. This is the
same result as the passive shunt damping case reported in Reference [118] showing
that the two resonant frequencies in an inductance- shunt piezoelectric system
coalesce into the open-circuit frequency, as the resistance approaches innity.
Electric charges for systems with the optimal tuning ratios are presented in
Figure 7.6, where the electro-mechanical coupling coecient varies from 0.2 to
0.6. It is seen that as the electro-mechanical coupling coecient increases, electric
194
charge near the operating frequency (open-circuit frequency) becomes much lower
than that in the static case. This implies that the required electric power could
be very low near the open-circuit frequency. This open-circuit frequency can be
interpreted as the frequency that yields the minimal displacement of y
e
in Figure
7.5. This indicates that large displacement x
e
could be achieved with the smallest
excitation of f
e
at the open-circuit frequency.
0 0.4 0.8 1.2 1.6 2
40
30
20
10
0
10
20
Nondimensionalized Frequency
Amplitude of Q (dB)
=0.2
=0.4
=0.6
Figure 7.6. Electric charges with the optimal tuning ratios for various coupling coe-
cients
195
7.1.2 Resonant Actuation Systems with Additional Capac-
itance
The increase of the electro-mechanical coupling coecient, as mentioned in the
previous section, will increase the static performance. However, there is a trade-o
between the stroke amplication (relative to the static stroke) and bandwidth at
the operating frequency. For given piezoelectric materials, the nominal generalized
electro-mechanical coupling coecient can be increased by introducing a negative
capacitance (or be reduced by adding a positive capacitance) in series with the
piezoelectric transducer. The negative capacitance cannot be realized passively,
but can be achieved using an operational amplier to form a negative impedance
converter presented in Figure 6.7 on Page 185.
When the additional capacitance is added in series to the piezoelectric trans-
ducer, and with a R-L circuit (see Figure 7.7), the transfer function between ac-
tuator stroke and control voltage is given in Equation 6.59 on Page 179. Now
relationship between two generalized coecients can be obtained by

2
=
K
2
c
K
E
K
Q
=
2
K
Q

K
Q
. (7.7)
R
C
p
V
p
V
c
L
PZT
C
add
V
a
Figure 7.7. A piezoelectric network with a series R-L circuit and an additional capacitor
196
There is a limit on adding capacitance to the piezoelectric transducer. That is,
the following criterion has to be satised to ensure system stability,
K
a
Q
>
1
1 +
2
K
Q
, or

2
< 1 +
2
. (7.8)
Here it should be noted that the short-circuit condition with the added capacitance
refers to the situation that the circuit is shorted between the electrode of the PZT
and that of the added capacitance while the PZT and the added capacitance are
connected in series. Thus the short-circuit frequency will be changed to,

E
=
_
(1 +
2

2
)
E
, (7.9)
and its range is given by
0 <
E
<
D
=
_
(1 +
2
)
E
. (7.10)
This means that the new resonant frequency can be assigned to somewhere below
the open-circuit frequency
D
. This situation can be explained further by con-
sidering an equivalent mechanical system of a piezoelectric actuation system with
an additional capacitor (see Figure 7.8). Dynamic equations of the equivalent
mechanical system are given by
_
M
M
0
0 m
m
__
x
e
y
e
_
+
_
C
M
0
0 c
m
__
x
e
y
e
_
+
_
K
M
+k
m
k
m
k
m
k
m
+k
n
__
x
e
y
e
_
=
_
0
f
e
_
, (7.11)
where an added spring k
n
only aects the second equation representing the elec-
trical behavior of actuation systems. Note that here k
n
could be negative.
As shown in Figure 7.8, the spring k
n
that corresponds to K
a
Q
is directly at-
tached to the ground, so that it only aects the short-circuit frequency. It is clearly
seen that if the displacement y
e
is xed (i.e., open-circuit condition), the displace-
ment x
e
is independent of the added spring k
n
. For a short-circuit condition, the
displacement y
e
is freely moving, the displacement x
e
is now strongly aected by an
added spring. This explains that how the short-circuit frequency can be adjusted.
From the system stability point of view, there is a limit on the added spring. This
197
M
M
m
x
y
K
M
k
m
C
M
c
m
f(t)
m
m
k
n
Figure 7.8. Schematic of an equivalent mechanical system to a piezoelectric actuation
system with an additional capacitor
is similar to adding capacitance to the electro-mechanical system. The equivalent
mechanical system version of Equation 7.8 can be expressed by
k
n
>
K
M
k
m
K
M
+k
m
. (7.12)
This simply indicates that an added spring should be larger than the negative
value of the stiness of two existing springs in series, so that the overall stiness
of the system should be greater than zero.
The optimal inductance tuning is the same as that of the electro-mechanical
system without an added capacitance, as presented in Equation 6.63, while the
optimal resistance tuning given in Equation 6.57 is dierent, which can be rewritten
in terms of the coupling coecients and

as:
r

1 +
2
0.5

2
1 +
2

2
, (7.13)
where one can see that if the coupling coecient,

, approaches the nominal cou-
pling coecient, , the optimal resistance tuning ratio becomes what is shown in
Equation 6.57.
198
The performance variation due to the added capacitance is presented in Figure
7.9 for various capacitance values, where the nominal coupling coecient is = 0.5
(the coupling coecient of the piezoelectric tube actuator). The added capacitance
value can be calculated from Equation 7.7 for given coupling coecients.
^
^
^
^
Figure 7.9. Actuator strokes with additional capacitance for = 0.5
There are two important performance indices, namely the response magnitude
at the invariant point (
2
= 1 +
2
),
_
q
t
V
c
_

1+
2

2
,

1+
2
=
1

2
_
q
t
V
a
_
ST
, (7.14)
and the bandwidth,

=

D

E

E
=
_
1 +
2

_
1 +
2

2
,

1
2

2
_
1 +
1
4
(

1
2

2
)
_
, (7.15)
which can be varied by adjusting the value of the added capacitance.
199
As noted in Reference [125], the coupling coecient does not change the mag-
nitude at the invariant point (open-circuit frequency,
D
), but it does change the
bandwidth and the static stroke authority:
_
q
t
V
c
_

=0
=

2
/
2
1 +
2

2
_
q
t
V
a
_
ST
. (7.16)
The magnitude at the invariant point with respect to the static stroke can be
evaluated by
_
q
t
V
c
_

1+
2
=
1 +
2

2
_
q
t
V
c
_

=0
. (7.17)
As far as the magnitudes at the vicinity of the resonant frequency are con-
cerned, the relative magnitude at the invariant point of Equation 7.17 (i.e., stroke
amplication with respect to the static value) is an important design parame-
ter. Variations of relative magnitude and bandwidth due to added capacitance
are plotted in Figure 7.10, where the nominal coupling coecient of the piezoelec-
tric tube actuator, , is 0.5 and the modied coupling coecient,

, is assumed
to be varying from 0.1 to 0.8. The increased coupling coecient using negative
capacitances tends to increase the bandwidth, while the stroke amplication with
respect to the static stroke approaches to unity. This situation is dierent from
that shown in Figure 7.3, where the large coupling coecients near 0.8 still show
stroke amplication.
7.1.3 Summary of Design Guidelines for the RAS circuitry
For electrical tailoring of the resonant actuation system, there are three design
parameters, namely the optimal inductance tuning, Equation 6.54, the optimal
resistance tuning, Equations 6.57 and 7.13, and the capacitance tuning, Equation
7.15, which are summarized in Table 7.1. The combined inductance-resistance
tuning contributes to the bandwidth and atness of the frequency response function
near the operating frequency, while the capacitance tuning contributes to the stroke
200
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
0
0.2
0.4
0.6
0.8
1
Modified coupling coefficient,
hat
log(q
t
/Vc)
B

Figure 7.10. Relative actuator stroke and bandwidth variations with modied coupling
coecients
amplication with respect to the static stroke as well as the operating frequency
bandwidth. Such design parameters are summarized as below:
Optimal inductance tuning ratio,

, to tune the circuit resonant frequency


to the open-circuit resonant frequency of system is given by

=
_
1 +
2
for with and without added capacitance.
Optimal resistance tuning ratio, r

, to make the frequency response near the


resonant frequency as at as possible is given by
r

2
_
1 + 0.5
2
for without added capacitance, and
r

1 +
2
0.5

2
1 +
2

2
201
for with added capacitance.
Actuator stroke with respect to the static stroke is expressed by
_
q
t
V
c
_

2
=1+
2
/
_
q
t
V
a
_
ST
=
1

2
for without added capacitance, and
_
q
t
V
c
_

2
=1+
2
/
_
q
t
V
c
_
=0
=
1 +
2

2
for with added capacitance.
Operating bandwidth, B

, is expressed by
B

=
1
2

2
_
1
1
4

2
_
for without added capacitance, and

=
1
2

2
_
1 +
1
4
_

1
2

2
__
for added capacitance.
Best compromised coupling coecients, and

, in terms of actuator stroke
amplication and operating bandwidth can be found from Figures 7.3 and
7.10
Table 7.1. Design parameters for the RAS circuitry
Optimal inductance tuning,

Equation 6.54
Optimal resistance tuning, r

Equations 6.57 and 7.13


Capacitance tuning (stroke) Equations 7.17
Capacitance tuning (bandwidth) Equation 7.15
Capacitance tuning (trade-o) Figure 7.10
202
7.2 Dynamic Characteristics of the RAS in For-
ward Flight
In the previous section, it is demonstrated that the high authority actuation system
can be achieved near the resonant frequency for general piezoelectric actuation sys-
tems. Design guidelines for such systems are provided. In this section, the dynamic
characteristics of the RAS will be examined, since the aerodynamic stiness, K
f
,
presented in Section 6.2.2 is not constant. The analysis is performed for piezoelec-
tric resonant actuators in forward ight using a perturbation method. Based on
this, the actuation authority is then predicted for the nominal actuation system
(i.e., non-resonant actuation system) and the RAS. Vibration reduction perfor-
mance of various ap congurations is evaluated within the available actuation
authority, which can be realized by the actuator saturation algorithm presented in
Section 3.3.1 on Page 106.
7.2.1 A Perturbation Method
The aerodynamic contribution of the ap based on Theodorsens theory with con-
sidering the forward ight speed can be written as:
F
f
(
f
; t) = M
f

f
+C
f
[1 + sin(t)]

f
+K
f
[1 + sin(t)]
2

f
, (7.18)
where
=

r
f
, (7.19)
in which r
f
denotes the ap location along the rotor spanwise direction. M
f
, C
f
and K
f
were derived in Equation 6.29 on Page 172 in Section 6.2.2.
With this aerodynamic contribution, coupled actuator-ap-circuit system equa-
tions given in Equation 6.39 are rewritten by
_
M 0
0 L
__
q
t

Q
_
+
_
C

0
0 R
__
q
t

Q
_
+
_
K
D

K
c
K
c

K
Q
__
q
t
Q
_
=
_
0
V
c
_
, (7.20)
203
where
C

= C
p
+C
f
A
2
M
[1 + sin(t)] , (7.21)
K
D

= K
D
p
+K
f
A
2
M
[1 + sin(t)]
2
. (7.22)
For mechanical tuning, it is needed to dene the nominal stiness of the ac-
tuation system, since its stiness is a function of time. The resonant frequency
of the system in hover could be served as the reference. In this case, the relative
wind velocity is constant because the forward speed parameter becomes zero in
Equation 7.20. The resonant frequency of the system in hover is then obtained as:

h
=


K
E
M
, (7.23)
where
h
represents the resonant frequency of the RAS in hover, and

K
E
indicates
the short-circuit stiness including the added capacitance.
This frequency is tuned to one of operating frequencies that are 3, 4 and 5/rev
of the main rotor speed for four-bladed rotor systems by adjusting the ap mass
moment of inertia. The resonant frequency in forward ight will be discussed in
the next section.
It is convenient to explore the simplied version of the electro-mechanical sys-
tem without circuitry. Equation 7.20, without damping, is then rewritten as fol-
lows:
M q
t
+

K
E

(t)q
t
= F
p
(t), (7.24)
where

K
E

(t) =

K
E
+K
f
A
2
M
_
2 sin(t) +
2
sin
2
(t)

, (7.25)
F
p
(t) =
K
c

K
Q
V
c
(t), (7.26)
V
c
(t) =

V
c
cos(t), (7.27)
where

V
c
represents the magnitude of voltage signal, and denotes the excitation
frequency.
204
For the purpose of performing perturbation analysis, Equation 7.24 is expressed
in the dimensionless form. A dimensionless quantity is introduced using an azimuth
angle . Then
q

t
+
_

2
m
+
2
m
_
2 sin +
2
sin
2

_
q
t
=
m
(cos
m
), (7.28)
where

2
m
=
1

2
_

K
E
M
_
, (7.29)

2
m
=
1

2
_
K
f
A
2
M
M
_
, (7.30)

m
=

, (7.31)

m
=
1

2
M
_
K
c

K
Q
_

V
c
. (7.32)
The most elementary version of the perturbation method is to attempt a rep-
resentation of the solution of Equation 7.28 in the form of a power series in a small
parameter :
q
t
(, ) = q
t0
() + q
t1
() +
2
q
t2
() + , (7.33)
whose coecients q
ti
() are only functions of . To form equations for q
ti
(),
i = 0, 1, 2, , substitute the series Equation 7.33 into Equation 7.28:
q

t0
+q

t1
+
2
q

t2
+ +
__

2
m
+
2
m
_
2 sin +
2
sin
2

__
_
q
t0
+q
t1
+
2
q
t2
+
_
=
m
cos(
m
). (7.34)
Equating each of the coecients of
0
,
1
and
2
to zero, we have

0
: q

t0
+
2
m
q
t0
=
m
cos(
m
), (7.35)

1
: q

t1
+
2
m
q
t1
= 2
2
m
q
t0
sin , (7.36)

2
: q

t2
+
2
m
q
t2
= 2
2
m
q
t1
sin
2
m
q
t0
sin
2
, (7.37)
and so on.
205
As far as the steady-state solutions are concerned, the periodicity of 2 can be
applied. That is,
q
t
(, + 2) = q
t
(, ). (7.38)
Equations 7.35 7.35 together with the condition of Equation 7.38 are sucient
to provide the required solution. The major term in Equation 7.33 is a periodic
solution of the linearized equation ( = 0, i.e., in hover). It is therefore clear that
this process restricts us to nding the solutions of the nonlinear equation which
bifurcate from the periodic solutions.
The solution of Equation 7.28 is then obtained, which is shown in Equation
7.39. The method obviously fails if the resonant frequency
2
m
takes one of the
values
2
m
, (
m
1)
2
, (
m
2)
2
, , since certain terms would then be innite.
Such values of correspond to conditions of near-resonance.
m
=
m
is so called
a primary or main resonance. The other values of
m
correspond to nonlinear
resonances caused by the harmonics presented in forward ight, which can be
regarded as feeding back into the linear equation (hover condition) as forcing
terms [153].
q
t
(; ) =

m

2
m

2
m
cos(
m
)
+
_
ao
1

2
m
(1
m
)
2
sin(1
m
) +
bo
1

2
m
(1 +
m
)
2
sin(1 +
m
)
_
+
2
_
co
2

2
m

2
m
cos(
m
) +
ao
2

2
m
(1
m
)
2
sin(2
m
)
+
bo
2

2
m
(1
m
)
2
sin(2 +
m
)
_
, (7.39)
where
ao
1
= bo
1
=

2
m

2
m

2
m
, (7.40)
ao
2
= ao
1
_

2
m

2
m
(1
m
)
2

1
4
_
, (7.41)
bo
2
= bo
1
_

2
m

2
m
(1 +
m
)
2

1
4
_
, (7.42)
206
co
2
= ao
1
_

2
m

2
m
(1
m
)
2
+

2
m

2
m
(1 +
m
)
2

1
2
_
. (7.43)
7.2.2 Analysis of Time Responses
Time responses can be directly calculated using the state-space form of Equa-
tion 7.20. A fourth-order Runge-Kutta method is used to solve the time-varying
periodic equation, which is the built-in function ode45 in Matlab.
To investigate the time-varying characteristics of the resonant actuation system,
the piezoelectric induced-shear tube actuator developed by Centolanza et. al [116]
is selected as an example. The rotor has a blade diameter of 34 feet (MD900 class
helicopter), a blade chord of 10 inches, and a nominal rotation of 400 RPM. The
length of a piezoelectric tube actuator is 8 inches, outer radius R
o
is 0.35 inches,
and inner radius R
i
is 0.225 inches. The ap span is 6% of blade length, 12 inches,
and the amplication ratio is 5.3. For the voltage driving response, an electric eld
of 4 kV/cm (1800 V
rms
) is applied to the actuation system. The total actuator
weight of 0.775 pounds corresponds to a weight penalty of 1.5 % as MD900 blades
weight approximately 55 pounds each [117]. If the resonant actuation system is
used, the weight is approximately increased to 1.2 pounds due to the presence of
multiple aps.
Time Responses without Circuitry
In general, the frequency response functions between the ap deection and
control voltage cannot be directly derived, since there is no xed stiness of the
actuation system. Peak-to-peak ap deections are therefore used to obtain the
frequency response functions. Flap deections with several advance ratios are
shown in Figure 7.11, where the frequency of the actuation system is tuned to
4/rev (26.6 Hz) by adding mass to the ap. As mentioned earlier, there are several
resonant peaks in forward ight conditions ( =0.15, 0.35) which correspond to 2, 3,
4, 5 and 6/rev frequencies. Among them, the most signicant resonant frequencies
are the 3, 4 and 5/rev for a resonant actuator tuned to 4/rev frequency, where a
207
resonant actuator indicates the actuation system without electric networks. The
inuence of advance ratios to the major resonance of 4/rev is not signicant, while
peak-to-peak ap deections near the static condition are dierent.
0 10 20 30 40 50
0
2
4
6
8
10
12
14
Flap deflections with various flight speeds
F
l
a
p

d
e
f
l
e
c
t
i
o
n
s
,

d
e
g
r
e
e
s
Excitation frequency, Hz
hover
=0.15
=0.35
Figure 7.11. Peak-to-peak ap deections of a resonant actuator with various ight
speeds
Figure 7.12 shows the variation of instantaneous frequencies and averaged fre-
quencies. This clearly shows why the inuence of advance ratios to the major
resonant frequency is not signicant. Averaged resonant frequencies are obtained
by a constant coecient approach, and also compared to those by Floquets the-
ory. Both methods yield almost the same results. Instantaneous stiness in forward
ight is either very sti or soft. This is the reason why peak-to-peak ap deections
in forward ight are larger than those in hover, as shown in Figure 7.11.
Figure 7.13 shows the time history of ap deections in forward ight ( = 0.35)
with a 4/rev cosine voltage input. There exists large 4/rev ap motion along with
moderate 1/rev content as well as other harmonics, as would be expected by a
perturbation analysis. Maximum ap deection occurs at the fourth quadrant on
the retreating blade. Similar trends of ap motions were reported in small-scale
rotor experiments by Fulton and Ormiston [30].
208
0 90 180 270 360
0
5
10
15
20
25
30
35
40
Variation of instantaneous frequencies
F
r
e
q
u
e
n
c
y
,

H
z
Azimuth, degrees
hover
=0.15
=0.35
Averaged frequency
Hover : 26.6 Hz
=0.15 : 26.7 Hz
=0.35 : 27.3 Hz
Figure 7.12. Variations of instantaneous frequencies along the azimuth
1.05 1.0875 1.125 1.1625 1.2
15
10
5
0
5
10
15
Flap time response for a resonant actuator tuned to 4P
time, sec
F
l
a
p

d
e
f
l
e
c
t
i
o
n
s
,

d
e
g
r
e
e
s
=90 =180 =270
Figure 7.13. Time history of ap motions of the actuation system without circuitry
with 4/rev voltage excitation, = 0.35
209
Time Responses with Circuitry
To verify the optimal tuning parameters for the circuitry, the frequency re-
sponses of the RAS are investigated for forward ight conditions. The optimal
tuning parameters are obtained for the hover condition and then applied to the
forward ight conditions since the inuence of advance ratios to the major reso-
nance of 4/rev was not signicant.
Peak-to-peak ap deections with several advance ratios are shown in Figure
7.14, where the resonant frequency is mechanically tuned to the 4/rev frequency.
The actuator authority is signicantly increased from 1.25 degree to 4.5 degree
when compared to the static value (i.e., non-resonant actuator authority). The
ap deections around the operating frequency are almost constant (at plateau)
and the bandwidth reasonably large (approximately 8 Hz). These characteristics of
the RAS are also conserved in forward ight conditions (advance ratios of 0.15 and
0.35). This implies that the proposed resonant actuation system can be applied to
the forward ights as well as hovering condition.
0 10 20 30 40 50
0
2
4
6
8
Flap deflections with various flight speeds
F
l
a
p

d
e
f
l
e
c
t
i
o
n
s
,

d
e
g
r
e
e
s
Excitation frequency, Hz
hover
=0.15
=0.35
Figure 7.14. Peak-to-peak ap deections of the RAS with various ight speeds
210
Figure 7.15 shows the time history of ap deections of the RAS in forward
ight ( = 0.35) with a 4/rev cosine voltage input. There exists large 4/rev
ap motion. However, there are also large 3/rev and 5/rev ap motions because
of the mechanical resonance and aerodynamic excitation. To properly control
the resonant actuation system in forward ight, the controller that rejects the
side eects (3/rev and 4/rev) is needed, which is also needed for the nominal
actuation system (i.e., non-resonant actuation system) due to the time varying
aerodynamic excitation, as shown in Figure 7.16. Maximum ap deection occurs
at the fourth quadrant on the retreating blade, which is the same as the actuation
system without circuitry.
Figure 7.15. Time history of ap motions of the RAS with 4/rev voltage excitation,
= 0.35
211
Figure 7.16. Time history of ap motions of the nominal actuation system with 4/rev
voltage excitation, = 0.35
212
7.2.3 Vibration Reduction Within Available Actuation Au-
thority
In Chapter 5, vibratory hub loads for each ap conguration are estimated based
on the assumed actuator authority. In the previous section, it is estimated that
actuator authorities of the PZT tube actuator are 1.38 and 4.8 degrees for baseline
actuation system (normally, similar to the static performance) and the resonant
actuation systems respectively. Thus the saturation angles of the ap deection are
set to 1.38, 1.38 and 4.8 degrees for single-, dual- and multiple-ap congurations,
respectively. Flap locations and control input sequences were listed in Section
Section 5.3.3, which was determined based on parametric studies, for a four-bladed
hingeless helicopter.
Vibratory hub loads for low-speed ight are presented in Figure 7.17 showing
that vibration reductions are 35%, 34% and 91% by single-, dual- and multiple-
ap congurations, respectively. Unlike previous cases, it is shown that multiple-
ap conguration outperforms the other ones when considered available actuator
authorities. For high-speed ight case, same trends are observed as shown in
Figure 7.18 showing that vibration reductions are 56%, 57% and 79% by single-
, dual- and multiple-ap congurations, respectively. This clearly indicates that
multiple-ap conguration with the resonant actuation system can signicantly
reduce the vibration level over the wide range of ight condition as compared to
single- and dual- ap congurations.
213
Figure 7.17. Comparison of vibratory hub loads for an advance ratio of 0.15 within
the available actuator authority
Figure 7.18. Comparison of vibratory hub loads for an advance ratio of 0.35 within
the available actuator authority
214
7.3 Experimental Realization of the RAS
In Chapter 6, the RAS circuitry was realized via an equivalent electric circuit
model to investigate its feasibility. Typical operating voltages of the trailing-edge
ap for helicopter vibration controls are normally larger than 500 Volts [112, 114].
To realize the actuation system, especially under high voltage operations, a better
method for implementing the electrical circuitry is needed. The previously pro-
posed approach using Op-Amp based circuitry that includes synthetic inductance
and negative capacitance may not be realistic under such situations. In addition to
this, there is the phase variation near the operating frequency, since the mechani-
cal resonant frequency was utilized to improve the actuation authority. These two
issues should be addressed to realize the piezoelectric resonant actuation systems.
In this section, a method of implementing the electric network is realized via
digital signal processor (DSP), instead of the traditional analogue Op-Amp cir-
cuitry. this is more practical approach, especially under high voltage situations.
An adaptive feed-forward controller PD controller is designed and implemented to
track the phase variation at the vicinity of resonant frequency that is one of the op-
erating frequencies. Through these eorts, performance of the resonant actuation
system is validated experimentally.
7.3.1 Controller Design
In order to evaluate the performance of resonant piezoelectric actuators with elec-
tric circuitry, the piezoelectric induced shear tube actuator [116] is considered as an
example. There are a number of implementation problems associated with realiz-
ing the inductor and the negative capacitor. The piezoelectric actuators for certain
application, helicopter rotor trailing edge aps in this case, require high voltage
and large inductance values. This, therefore, requires that the synthetic inductor
and negative capacitor circuits constructed with high voltage operational ampli-
ers as discussed in Section 6.4.2, which may induce complexity and cause weight
and space penalties. To circumvent this problem, in this investigation, instead of
creating analogue circuitry elements, a control voltage signal is directly applied to
215
the piezoelectric actuators to emulate the circuitry functions. This signal func-
tion can be derived from the frequency response between the actuator stroke and
voltage source, which is implemented using a digital signal processor (DSP) unit.
Realization of Electric Network
From Equations 6.35, 6.39 and 6.47, the voltage across the piezoelectric actu-
ator V
a
is given as follows:
V
a
V
c
= K
Q
_

Q
V
c
_
K
c
_
q
t
V
c
_
=
_

2
_

K
Q
_

Q
V
c
_
K
c
_
q
t
V
c
_
, (7.44)
which yields
_
V
a
V
c
_
cty
=

2
(

2
/
2
)(1 + 2j
2
)
(1 +
2
+ 2j
2
)(rj +
2

2
)
2

2
, (7.45)
where subscript cty represents the voltage amplication ratio with a series R-L
circuit and the added capacitance.
The actuator stroke q
t
can now be expressed in terms of two transfer functions:
one is the transfer function of a piezoelectric resonant actuator without a circuitry,
and the other is the transfer function between the voltage across the PZT and the
voltage source.
q
t
=
_
q
t
V
a
_ _
V
a
V
c
_
cty
V
c
. (7.46)
Here it should be noted that the voltage across the piezoelectric actuator, V
a
,
should not exceed the de-poling limit. This is another constraint for adding ca-
pacitances to the piezoelectric RAS (especially for negative capacitance).
To incorporate the electric network in high voltage piezoelectric resonant actu-
ation systems, the transfer function of Equation 7.45 is implemented via the DSP
system (Matlab and dSPACE) as a feed-forward controller (Figure 7.19).
216
Piezoelectric
Actuator
Flap with
tuning mass
Aerodynamics
Va/Vc Phaser V
ref
V
signal
DSP
RAS
Figure 7.19. Controller diagram of the resonant actuation system
Phase Controller
When a periodic input signal with period To is applied to a resonant actuation
system, the output has the same period as the input but is shifted in phase. In this
case, one can use a phaser that represents apparent phase shift between input
and output [154]. The phaser can be viewed as the counter part of a gain which
modies the magnitude of an input signal but not its phase. The simple phaser
based on the Hilbert transform can be rewritten as the PD controller form in the
time domain.

f
(t) = cos


ref
f
(t) +
sin

f
ref
(t). (7.47)
A similar strategy was used experimentally in Reference [154]. In this paper, the
phase angle

f
is adaptively corrected through the feedback of the output signal
(see Figure 7.20).
217
Figure 7.20. Diagram of an adaptive phase controller based on Matlab/Simulink
218
7.3.2 Bench Top Testing
A resonant actuation system using piezoelectric tube actuator developed by Cen-
tolanza et. al [116] is evaluated analytically and experimentally. The length of the
piezoelectric tube is 8 inches, its diameter is 0.7 inches, and its mass is 0.5 lbs,
which is originally designed for the 12 inches ap of MD900 class helicopter rotors.
The trailing edge ap mass is 0.15 lbs.
A bench top test is conducted to examine the phase controller and the resonant
piezoelectric actuation system. Aerodynamic loads, which could be simulated by
a physical spring, are not considered in this study. A diagram of the experimental
set-up and equipment used in the experiment are shown in Figures 7.21 and 7.22.
The displacement and voltage frequency responses are measured using a Polytec
laser vibrometer and a HP dynamic signal analyzer (35665A). The coupled circuit
dynamics is realized by a feed-forward type controller of Equation 7.45, which
is implemented together with an adaptive phaser controller via a digital signal
processor (DSP, dSPACE ds1102).
Figure 7.21. Experimental set-up for the resonant actuation system
219
Figure 7.22. Equipments used in the experiment
The resonant frequency of the actuation system (actuator and ap) is 41Hz.
This is tailored to the 4/rev frequency (26.6Hz) by adding a mass of 0.1 lbs to the
ap. Frequency responses before and after mechanical tuning are shown in Figure
7.23, where the ap responses are normalized by their static values.
Analytical and experimental results are presented in Figures 7.24 and 7.25 with
two dierent modied electro-mechanical coupling coecients (

= 0.5 and 0.6).


Experimental results show good agreement with analytical predictions near the
resonant frequency. It is observed that resonant peaks are reduced due to friction
force at the ap hinge. This could be improved by implementing a thrust bearing
design [108]. Actuator authority at the tuned frequency (26.6Hz) is increased
about 2.1 3.5 times when compared to the static deection (which would be
produced by the original actuation system without resonant tuning). There is a
220
0 10 20 30 40 50 60
0
1
2
3
4
5
6
Frequency, Hz
F
l
a
p

d
e
f
l
e
c
t
i
o
n

r
a
t
i
o
s
Mechanical tuning 4/rev (26.6Hz)
Baseline (Experiment)
Baseline (Analysis)
Mechanically tuned system (Experiment)
Mechanically tuned system (Analysis)
Figure 7.23. Frequency responses of the PZT tube actuator before and after mechanical
tuning
trade o between bandwidth and stroke as illustrated in the gures.
Phase variations of a RAS (with the voltage signal function) shows more grad-
ual change than those of pure mechanically tuned system around the resonant
frequency, as is shown in Figures 7.24 and 7.25. This can help to track the phase
variation at the vicinity of resonant frequency. Required inductance values in these
cases are more than 1000 H, which is hard to implement using physical circuitry
components due to the high voltage across the inductor.
221
0 10 20 30 40 50
0
1
2
3
4
5
F
l
a
p

d
e
f
l
e
c
t
i
o
n

r
a
t
i
o
s
Analytical predictions
0 10 20 30 40 50
180
90
0
90
180
Frequency, Hz
P
h
a
s
e
,

d
e
g
Figure 7.24. Analytical predictions of a resonant actuation system: , tuned system
w/o the voltage signal function; , RAS with

= 0.5; , RAS with

= 0.6
0 10 20 30 40 50
0
1
2
3
4
5
F
l
a
p

d
e
f
l
e
c
t
i
o
n

r
a
t
i
o
s
Expreimental results
0 10 20 30 40 50
180
90
0
90
180
Frequency, Hz
P
h
a
s
e
,

d
e
g
Figure 7.25. Experimental results of a resonant actuation system: , tuned system
w/o the voltage signal function; , RAS with

= 0.5; , RAS with

= 0.6
222
While the voltage signal function, which emulating the electrical network, can
help to broaden and atten the resonant peak of actuation system, the phase near
a resonant frequency varies with the control input frequencies. To track the phase
variation, an adaptive phaser is implemented and tested experimentally. Signal
errors and ap signals before/after implementing the controller with a 24Hz input
are shown in Figure 7.26, where the solid and dash-dotted lines represent the
ap response and reference input, respectively. It is shown that the implemented
adaptive controller is able to accurately follow the reference input.
0 5 10 15 20
2
0
2
S
i
g
n
a
l

e
r
r
o
r
s
24 Hz input signal
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
2
0
2
B
e
f
o
r
e

c
o
n
t
r
o
l
10 10.05 10.1 10.15 10.2 10.25 10.3 10.35 10.4
2
0
2
A
f
t
e
r

c
o
n
t
r
o
l
time, sec
reference
output
Figure 7.26. Time responses of ap deection signal before/after phase control with
24Hz input signal
Reference inputs of 26Hz and 29Hz are also tested in the experiment. Time
responses with 26Hz input signal are shown in Figure 7.27. It is illustrated that
the controller works well, implying that the it can adapt for the phase variation
near the operating frequency. Such a controller together with the voltage signal
function yields a robust and high performance ap actuation system.
223
0 5 10 15 20
2
0
2
S
i
g
n
a
l

e
r
r
o
r
s
26 Hz input signal
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
2
0
2
B
e
f
o
r
e

c
o
n
t
r
o
l
10 10.05 10.1 10.15 10.2 10.25 10.3 10.35 10.4
2
0
2
A
f
t
e
r

c
o
n
t
r
o
l
time, sec
reference
output
Figure 7.27. Time responses of ap deection signal before/after phase control with
26Hz input signal
224
7.4 Power Consumption of Piezoelectric Reso-
nant Actuation Systems
In previous sections, it is demonstrated, analytically and experimentally, that the
high actuation authority with a wide bandwidth can be achieved near the operating
frequency for piezoelectric resonant actuation systems. In this section, the electric
power consumption of the present RAS is evaluated and compared to that of
piezoelectric resonant actuators without circuitry.
7.4.1 Piezoelectric resonant actuators without circuitry
The electrical property of a piezoelectric actuator is similar to a capacitor. The
electrical power consumption for the case of the non-ideal capacitor can be esti-
mated as [155, 156]:
P = V I cos
c
(7.48)
where V and I represent the absolute values of voltage and current respectively,

c
is the phase angle between voltage V and current I, and cos
c
is known as the
power factor.
The power factor of a piezoelectric actuator is approximately zero when the
operating frequency is far away from the resonance frequency. In order to increase
the eciency, it is desirable to increase the power factor and make it as close
to unity as possible [155]. In our case, this is achieved through the mechanical
resonance tailoring. That is, by tuning the actuation system natural frequency
to the operating frequency, the power factor becomes unity at two points in the
frequency response function, which are the short- and open-circuit frequencies.
From Equation 6.23 on Page 170, the frequency response between the current
and voltage source can be evaluated as:
I
V
a
=
_
Q
V
a
_
,
_
I
V
a
_
=
_
Q
V
a
_
+ 90
o
. (7.49)
225
The frequency response and phase plot of a piezoelectric resonant actuator for
= 0.5 with dierent resistances are shown in Figure 7.28, where r
p
=
E
R
p
/K
Q
.
The resonant and anti-resonant frequencies correspond to the short- and open-
circuit frequencies, respectively. It is seen that the apparent electric power is
minimal at the open-circuit frequency. This is obvious since the open-circuit con-
dition is to have constant electric displacement which includes zero electric charge.
The optimal operating frequency would then be the open-circuit frequency
D
be-
cause the electric power to the actuators is minimal. As far as the power factor is
concerned, short- and open-circuit frequencies are the best operating frequencies
for electric power eciency, in which the power factor is unity, (Since the phase,

c
, between current and voltage is zero at these frequencies, as shown in the phase
plot in Figure 7.28). The electric resistance can help to vary the phase gradu-
ally around the resonant frequency, but not the phase around the anti-resonant
frequency.
Figure 7.28. Frequency response of current and phase variation for = 0.5 with
dierent resistances
226
7.4.2 Piezoelectric resonant actuators with circuitry
When a series R-L circuit is connected to the PZT, the frequency response between
the current and voltage source is given by, from Equation 6.46 on Page 176,
I
V
c
=
_
Q
V
c
_
,
_
I
V
c
_
=
_
Q
V
c
_
+ 90
o
. (7.50)
The frequency response and phase plot of a piezoelectric resonant actuator with
a series R-L circuit for = 0.5 are shown in Figure 7.29, where results are plotted
with the optimal inductance tuning. There are now three points where the power
factor is unity (i.e., the phase angle,
c
, is zero), as shown in the second plot
of Figure 7.29. One is the open-circuit frequency, and the others are at the two
resonant frequencies. The voltage across the PZT should be considered to evaluate
the power consumption since it is not constant in this case.
Figure 7.29. Frequency response of current and its phase with circuitry for = 0.5
227
The voltage across the piezoelectric actuator, which is similar to Equation 7.45,
can be evaluated as:
_
V
a
V
c
_
=

2
(1 + 2j
2
)
(1 +
2
+ 2j
2
)(rj +
2

2
)
2

2
, (7.51)
and the apparent electric power is then evaluated,
_
P
V
c
_
=
_
V
a
V
c
__
Q
V
c
_
V
c
, (7.52)
which has two poles at the short- and open-circuit frequencies. This is clearly
illustrated in Figure 7.30 showing the apparent electric power of a piezoelectric
resonant actuator with circuitry for = 0.5. Results are obtained under the
optimal inductance tuning and compared to the baseline (without circuitry).
Figure 7.30. Apparent electric power of a piezoelectric actuator with circuitry for
= 0.5
There is now the banded frequency range between the short- and open-circuit
frequencies. This bandwidth can be evaluated by Equation 7.3 on Page 190 that
is closely related to the electro-mechanical coupling coecient . Corresponding
228
actuator strokes are presented in Figure 7.31. From these gures, one can conclude
that a piezoelectric resonant actuation system with proper tuning parameters not
only yields high authority with broad bandwidth but also is electrically ecient
when compared to that without circuitry.
Figure 7.31. Actuator strokes under the optimal tuning ratios for = 0.5
229
7.5 Summary
The investigation of this chapter addresses research issues related to the synthesis
and realization of a piezoelectric resonant actuation system described in Chapter
6. The eorts and ndings can be summarized as follows:
Design guidelines for the RAS circuitry (inductance, resistance and addi-
tional capacitance) are provided based on the nondimensionalized formulae.
The RAS system is also compared to an equivalent mechanical system to
provide better physical understanding.
Time varying characteristics of the RAS in forward ights are investigated.
The analysis shows that the actuator authority is signicantly increase from
1.25 to 4.5 degrees when compared to the static value. The main characteris-
tics of the RAS are conserved in forward ight conditions with the mechanical
tuning and electrical tailoring for hover condition.
Within the available actuation authority, multiple-ap conguration with
resonant actuation systems can signicantly reduce the vibration level over
the wide range of ight condition as compared to single- and dual- ap cong-
urations. The multiple-ap conguration reduces the vibration by 91% and
79%, in low- and high-speed ight conditions, while single-ap conguration
does it by 35% and 56%.
A method is proposed to emulate the circuitry functions by directly applying
a voltage signal function to the actuator. This will avoid the diculties
of implementing high voltage analogue electrical components. The voltage
signal function is derived and implemented using a digital signal processor
(DSP) system, forming a feed-forward type controller. An adaptive phaser
is also implemented to track the phase variation near the resonant frequency.
The concepts presented are applied to the piezoelectric induced-shear tube
actuator for MD900 class helicopters, and veried experimentally. The the-
oretical predictions match well with the experimental ndings. Promising
results are illustrated in this study. The actuator stroke is increased 2 3.5
times compared to its static value with wide bandwidth of 5 to 10 Hz.
230
In addition to the actuator authority and bandwidth, the electric power
consumption is also investigated to truly evaluate the performance of the
RAS. It is found that the electric power consumption of the RAS is low
between the short- and open-circuit frequencies, where the required voltage
is very low at the short-circuit frequency and the required current is very low
at the open-circuit frequency. Based on this investigation, it is concluded that
the proposed piezoelectric resonant actuation system (RAS) with properly
tuned circuitry parameters not only yields a high actuation authority with a
wide bandwidth, but also is electrically very ecient.
Chapter 8
Conclusions and Recommendations
In this thesis, active control methods utilizing the multiple trailing-edge aps
conguration and the piezoelectric resonant actuation system are proposed for ro-
torcraft vibration suppression and blade loads control, and such resonant actuation
systems are developed and tested. The general conclusions of this dissertation are
summarized, and recommendations for future work are presented in this chapter.
232
8.1 Summary of Research Eorts and Achieve-
ments
In Chapter 2, an aeroelastic analysis has been formulated to examine the per-
formance and loads of rotorcraft with elastically coupled composite blades. The
analysis includes the free-wake model developed by Tauzsig and Gandhi [132] to
accurately predict the blades loads. A nonlinear composite beam theory, which
includes the cross-section model based on Vlasov theory, is employed to obtain the
blade structural model. The derivation of strain measure used in this formulation
is presented in Appendix A. The explicit forms of strain and kinetic energy con-
tributions used in the nite elements implementation are provided in Appendix B.
Blade loads calculation and vehicle propulsive trim procedure are also provided.
In Chapter 3, the trailing-edge ap model and optimal controller for both
vibration and blade loads have been developed. The inertial loads of trailing-
edge aps are derived based on the previous work [69]. The incompressible and
compressible aerodynamic models were presented to predict the aerodynamic loads
generated by ap motions. The optimal controller with the actuator saturation
algorithm were developed to reduce the vibration and blade loads. The hybrid
design approach developed by Zhang [57] was briey reviewed.
Based on the formulations presented in Chapter 2 and Chapter 3, the feasibility
of multiple trailing-edge aps for the simultaneous reductions of vibration and
blade loads was investigated in Chapter 4. The concept involved straightening the
blade by introducing dual trailing edge aps in a conventional articulated rotor
blade. An active loads control strategy was numerically tested for the steady-state
forward ight condition, = 0.35. It was shown that straightening the rotor blade
using dual ap congurations can reduces both vibratory hub loads and bending
moments without a signicant change of control settings for trim. Only 1/rev
control input was required to reduce vibratory hub loads with the present method,
which is well suited for resonant actuation systems. The proposed active loads
control method was shown that the apwise bending moment and the vibratory
hub loads can be reduced by 32% and 57%, respectively, through dynamically
233
straightening the blade with the 1/rev control input. The required maximum
ap deection was reduced by 25% when compared to retrot design using the
composite tailoring of pitch-ap coupling stiness via a hybrid design approach.
In Chapter 5, an active control method for multiple trailing-edge aps congu-
ration was developed. The concept involved deecting the trailing-edge aps by in-
troducing the resonant actuation, in which each ap operates at a single frequency,
in a four-bladed hingeless rotor system. A proposed active vibration control strat-
egy was simulated for the steady-state forward ight conditions ( = 0.15, 0.35).
The nite wing eect of the ap was not signicant in the multiple trailing-edge
ap conguration, since the controller was able to compensate for changes in ap
capability by adjusting the ap deections and phases. For the low-speed ight,
the vibration level was reduced by 56 60 % with a 4/rev control input. It was
shown that multiple-ap conguration can reduce the peak-to-peak hinge moments
by 49.8% and 50.7% at in- and out-board aps when compared to the single-ap
approach. It was also shown that all ap congurations can reduce the vibration
level by 80 85% for both low- and high- speed ight conditions with multi-cyclic
control inputs. For high-speed ight, however, the peak-to-peak hinge moments
were reduced by 37.5% to 61.2% in each individual actuator compared to single-ap
conguration.
In Chapter 6, a new approach was developed to enhance the actuation author-
ity of piezoelectric actuation systems. The idea was to rst achieve a resonant
driver through using mechanical tuning, and then increase the bandwidth and ro-
bustness of the resonant actuation system through electrical network tailoring. A
coupled PZT actuator, trailing edge ap and electrical circuit dynamic model was
derived. Utilizing the developed model, the required electrical circuitry parameters
were systematically determined. A Mach-scaled piezoelectric tube actuator-based
trailing-edge ap for helicopter vibration control was used as an example to illus-
trate the proposed concept. The negative capacitance and inductance were real-
ized and implemented using operational amplier-based circuitry. An experiment
were set up using equivalent circuit representing the integrated structure-actuator-
network system. It was demonstrated that the proposed resonant actuation system
234
can indeed achieve both high active authority and robustness.
While investigations on the electro-mechanically tailored piezoelectric resonant
actuation system have shown promising results in Chapter 6, there were still re-
search issues to be addressed before such a concept can be realized. In Chapter
7, such issues related to the synthesis and realization of a piezoelectric resonant
actuation system were addressed.
First, design guidelines for the RAS circuitry (inductance, resistance and addi-
tional capacitance) were provided based on the nondimensionalized formulae. The
RAS system were also compared to an equivalent mechanical system to provide
better physical understanding.
Second, time varying characteristics of the RAS in forward ights were inves-
tigated. It was shown that the actuator authority is signicantly increase from
1.25 to 4.5 degrees when compared to the static value. The main characteristics of
the RAS were conserved in forward ight conditions with the mechanical tuning
and electrical tailoring for hover condition. It was also demonstrated that within
the available actuation authority prediced, the multiple-ap conguration with
resonant actuation systems can signicantly reduce the vibration level over the
wide range of ight condition. The multiple-ap conguration reduces the vibra-
tion by 91% and 79%, in low- and high-speed ight conditions, while single-ap
conguration does it by 35% and 56%.
Third, a method was developed to emulate the circuitry functions by directly
applying a voltage signal function to the actuator. The voltage signal function was
derived and implemented using a digital signal processor (DSP) system, forming
a feed-forward type controller. An adaptive phaser was also implemented to
track the phase variation near the resonant frequency. The concept presented
were applied to the piezoelectric induced-shear tube actuator for MD900 class
helicopters, and veried experimentally. The theoretical predictions matched well
with the experimental ndings. Promising results were illustrated in this study.
The actuator stroke was increased 2 3.5 times compared to its static value with
wide bandwidth of 5 to 10 Hz.
Fourth, in addition to the actuator authority and bandwidth, the electric power
consumption is also investigated to truly evaluate the performance of the RAS.
235
Based on this investigation, it is concluded that the proposed piezoelectric resonant
actuation system (RAS) with properly tuned circuitry parameters not only yields
a high actuation authority with a wide bandwidth, but also is electrically very
ecient.
8.2 Recommendations for Future Work
The work described in this dissertation is to reduce the helicopter vibration and
blade loads using multiple trailing-edge aps together with resonant actuation
systems and to develop such resonant actuation systems. Although the present
research has demonstrated that the multiple-trailing edge ap conguration with
resonant actuation systems can achieve the vibration reduction and blade loads
reduction within the available actuation authority, there are still room to improve
and extend the present research. Several suggestions for future work are described
below.
1. Trailing-Edge Flap Dynamics in Comprehensive Rotor Analysis:
The rst recommendation for future study is that it is needed to consider the
trailing-edge ap dynamic model in an aeroelastic comprehensive rotor anal-
ysis. The trailing-edge ap response is strongly aected by the time-varying
aerodynamic loads as shown in Section 7.2.2. Thus implementing accurate
dynamic model of the actuation system is important for more realistic anal-
ysis, especially for the case of using the resonant actuation concept.
2. Active Loads Control for Heavy Lift Class Helicopter:
A heavy lift class helicopter suers from the excessive blade loads. Dynami-
cally straightening the blade could yield low blade loads as well as low vibra-
tions using dual ap conguration. In this research, the blade loads control
was explored only for the high speed ight. To consider the low speed ight,
the wake-model should be used in this case. Multiple aps conguration can
be also applied to reduce the blade loads via the resonant actuation concept.
It is also needed to reduce the pitch-link loads and lag bending moment as
well as the ap bending moment.
236
3. Realistic Adaptive Controller for Helicopter Vibration and Blade Loads Con-
trols:
It is needed to design the realistic controller that performs the actuator dy-
namic compensation as well as system variations (e.g. air turbulence)
4. Further Investigation of Composite Tailoring:
Composite tailoring was considered in active loads control using dual ap
conguration in Section 4. To reduce both vibration and blade loads, com-
posite tailoring can be used in multiple-trailing edge ap conguration. This
can be further extended to reduce the pitch-link loads together with dynamic
shock absorber located at the pitch-link place.
5. Swashplateless Helicopter Control Using Multiple Flaps:
As shown in Chapter 4 and Chapter 5, the resonant actuation concept is well
suited for the multiple ap conguration. Using only 1/rev control input can
reduce both blade loads and vibration. This can be further extended for
swashplateless helicopter control, since it requires 1/rev control as the pri-
mary vehicle control. Multiple aps together with 1/rev resonant actuation
could improve the swashplateless helicopter control.
6. Nonlinear Modeling of Piezoelectric Resonant Actuators:
The present model of piezoelectric resonant actuation system was based on
the linear piezoelectricity. There are nonlinearities in piezoelectric materials
due to the high voltage excitation, high dynamic stress, and temperature. In
order to consider these nonlinearities, nonlinear modeling is needed. Thus
one can achieve better physical understandings and appropriate controller
design for the piezoelectric resonant actuation system.
7. Controller for Hysteresis and Nonlinear Resonances in Resonant Actuation
Systems:
There are two types of nonlinearities in piezoelectric materials, such as hys-
teresis and material nonlinearity. These two nonlinearities are important in
designing the controller of piezoelectric actuators, especially for the resonant
type actuators. In the case of the resonant actuation systems for active ap
rotors, they are under both high voltage and high stroke driving operations.
237
8. Wind Tunnel Test of Resonant Actuation Systems:
The resonant actuation system developed was veried experimentally only
for the hovering condition. It is needed to test it for forward ight conditions
for the investigation of aerodynamic damping and structural damping that
are important parameters in trailing-ap actuation performance. This can
be achieved through the wind tunnel test.
Appendix A
Derivation of Strain Measure
The strain-displacement relations reect the nonlinearity due to moderately large
rotation of the exible beam. Additional terms are also visible due to blade pitch,

o
. This pitch is generally due to combinations of elastic torsion, pretwist, and
pitch control inputs.
The strain measure with small strains but moderately large rotation was de-
veloped by Hodges and Dowell [134]. They developed the strain measure based
on Eulerian formulation. The cross-section of beam, however, was treated as a
solid section. In this section, the strain measure is re-derived for a thin-walled
cross-section. The strain measure is given by

xx
= u

+
v
2
2
+
w
2
2

T

+ (
2
+
2
)
_

+

2
2
_
yv

zw

, (A.1)
2
x
=
x
= ( +
T,
)

, (A.2)
2
x
=
x
= (
T
,
)

, (A.3)
where
T
is a warping function of beam cross-section, ()
,x
and ()

are equivalent
to
()
x
,
y = cos
1
sin
1
, z = sin
1
+ cos
1
, (A.4)
and
x
and
x
are engineering shear strains.
Here, note that the coordinate system e
c
i
is based on the undeformed coordinate
system e
u
i
but rotated by pitch input angle
o
due to rigid input pitch
75
and linear
239
V
h
V
h
s
n
c
q
c
P
P
) cos(
) sin(
) cos( ) sin( ) , ( ) , (
) , (
3 2
3 2
c
c
c
c
c
c c
c c
c
n
n
e n e n P P
e e P
q z V
q h h
q q z h V h
z h z h
- =
+ =
- + =
+ =
r r
r
r r
r
c
e
2
r
c
e
3
r
h
z
Figure A.1. Thin-walled cross section of rotor blade
twist angles
tw
. This relation can be obtained as:
e
c
i
= T
cu
ij
e
u
j
, T
cu
=
_

_
1 0 0
0 cos
o
sin
o
0 sin
o
cos
o
_

_
(T
du
)
o
. (A.5)
Figure A.1 shows the coordinate systems of the thin-walled cross section of rotor
blade. The contour line of the cross section is represented by the local coordinate
system, s and n. The distance from torsion center to contour line can be expressed
by and

, and relations between (, ) and ( ,

) are given by
= +n sin
c
, =

n cos
c
, (A.6)

T
=

T
n ( cos
c
+

sin
c
), (A.7)
where the overbar

() represents quantities on the contour line. These relationships
can be found in Reference [137].
Note that
c
and
1
are dierent angles. The former indicates the relative
orientation of the two coordinate systems that are the local coordinate system
(, , ) and the undeformed coordinate system e
c
i
after rotated by rigid pitch
input and linear twist angle
o
. The latter represents the total blade pitch angles.
240
A.1 Coordinate Transformation
For convenience, the strains based on (x, , ) coordinates are transformed into
the strains based on (x, s, n) coordinates. It is assumed that these two coordinate
systems are on the same plane, which is consistent with Euler-Bernoulli hypothesis.
Then shear strains are given as follows:

xs
=
x
cos
c
+
x
sin
c
, (A.8)

xn
=
x
sin
c

x
cos
c
, (A.9)
in which
xs
represents the shear strain due to torsion, while
xn
represents the
shear strain due to exure. The shear strain
xn
may be cancelled out because
it was developed under Euler-Bernoulli hypothesis, which means that the cross
section remains rigid and perpendicular to the elastic axis.
The shear strains,
xs
and
xn
, can be expressed in the coordinate system ( ,

, n) as follows:

xs
= (r

T
,s
)

+n 2

=
t
xs
+n k
xs
, (A.10)

xn
= q

T
,
sin
c

T
,
cos
c
)

0, (A.11)
where
r = sin
c

cos
c
, q = cos
c
+

sin
c
. (A.12)
As shown in Equation A.11, the shear strains due to exure
xn
turns out to
be zero because
_
q ds = 0 , cos
c
=
d
ds
, sin
c
=
d

ds
. (A.13)
Thus q term can be neglected because it does not aect the resultant force. This
is consistent with Reference [137]. Then the warping function of contour line

T
is simplied by

T
=
_
_
r
t
xs
_
ds

T,s
= r
t
xs
,
t
xs
=
t
xs

, (A.14)
in which
t
xs
denotes the shear strains due to torsion in the beam with the closed
241
cross section. In case of the open cross section, this strain may be vanished, which
is Vlasovs assumption. The various denition of
t
xs
can be found in References
[136, 138]. Note that the sign of warping function is dierent from References
because of a dierent sign convention. Finally, the shear strains can be obtained
as follows:

xs
=
t
xs
+n 2

. (A.15)
Now normal strain (Equation A.1) and shear strain (Equation A.15) are ex-
pressed in the coordinate system (x, s, n). It can be achieved by substituting the
total pitch input (Equation 2.6) and elastic torsion relations (Equation 2.8) into
the strain components (Equations A.1 and A.15). Then the following linear and
nonlinear strain components are obtained.

xx
=
L
xx
+
NL
xx
+n k
L
xx
=
o
xx
+n k
xx
, (A.16)

xs
=
L
xs
+
NL
xs
+n k
L
xs
=
o
xs
+n k
xs
, (A.17)
in which

L
xx
= u

y
o
v

z
o
w

, (A.18)

NL
xx
=
1
2
(v
2
+w
2
)
+ (
2
+

2
)
_

o
(

+w

) +
1
2

2
+

_
+

( z
o
v

y
o
w

), (A.19)
k
L
xx
= w

cos

v

sin

+q

, (A.20)

L
xs
=
t
xs

,
NL
xs
=
t
xs
w

, k
L
xs
= 2

, (A.21)
where
y
o
= cos
o

sin
o
, z
o
= sin
o
+

cos
o
,

=
o
+
c
. (A.22)
Superscripts L and NL denote linear and nonlinear terms, respectively. Nonlinear
terms have been neglected in the curvature equations, because the nonlinear cur-
vature is physically very small and higher order O(
5
). Some unusual linear terms
242
have been treated as nonlinear term. The linear strain components are the same
as those in Reference [137] except the terms related to the pitch angle
o
.
A.2 Foreshortening Term
The axial deection u is represented by two separate components; an elastic ax-
ial deection, u
e
, and a kinematic axial deection, u
F
, due to foreshortening or
moderately large rotation. That is,
u = u
e
u
F
, u
F
=
1
2
_
x
0
(v
2
+w
2
)dx, (A.23)
As mentioned in section 2.1.2, the base vector of displacements u, v, w is based
on the undeformed coordinate system, e
u
i
, which is the rectangular cartesian co-
ordinate system. In this section, some explanation for replacing u with u
e
will be
given.
For simplicity, let us consider the two-dimensional beam deection problem
with moderately large rotation (see Figure A.1). In developing the strain mea-
sure, the undeformed coordinate system was used in Equation A.1. The deformed
position in terms of displacement, in the undeformed coordinate system, can be
expressed as:
r
1
= (x +u)e
u
1
+we
u
3
, r
o
= xe
u
1
, (A.24)
and the innitesimal increment of position vector can be obtained as,
dr
1
= (1 +u

)dxe
u
1
+w

dxe
u
3
, dr
o
= dxe
u
1
, (A.25)
2
ij
(dx)
2
= dr
1
dr
1
dr
o
dr
o
(2u

+w
2
)(dx)
2
. (A.26)
This strain measure is well known as von Karman partial nonlinearity, which
causes a foreshortening of the beam. In moderately large rotation case, the axial
displacement u is no longer just an elastic axial displacement. It includes a fore-
shortening term, and its magnitude is greater than the elastic axial displacement.
This makes the system stier, as the rotation increases. Eventually, a lateral de-
ection w will be restricted to certain small value. The added foreshortening term
243
w
2
reduces the strain, u

, compared to the strain in an undeformed blade. The


strain energy of the system is therefore smaller, and the system is more exible,
which allows larger deections.
It is possible to dene the deformed position vector based on the orthogonal
curvilinear coordinate system. Then the deformed and undeformed position vector
can be written as
r
1
= (r +u
e
) e
d
1
, r
o
= xe
u
1
, (A.27)
Note that the deformed covariant base vector e
d
1
already includes all information
related to a deformation, and the displacement is measured along the deformed
elastic axis e
d
1
. Another dierence is that e
d
1
is not a constant but a function of x.
The increment of position vector for this case can be expressed by
dr
1
= (dr +du
e
) e
d
1
+ (r +u
e
) de
d
1
, dr
o
= dx e
u
1
, (A.28)
where the underline term represents the curvature eect of the deformed elastic
axis. The strain measure based on Equation A.28 is given by
2
ij
(dx)
2
= dr
1
dr
1
dr
o
dr
o
= (1 + u
+
e
)
2
(dr)
2
(dx)
2
2u
+
e
(dx)
2
, (A.29)
where superscript + denotes
()
r
. In second order approximation, ()
+
becomes
equivalent to ()

.
Assuming the system undergoes moderately large deformations, the two strain
terms presented by Equations A.26 and A.29 should be equivalent and independent
of the coordinate systems, for a second order approximation. Then, the following
relationship can be obtained as
u

e
= u

+
1
2
w
2
, (A.30)
In case of three-dimensional problem, Equation A.30 can be rewritten by
u

e
= u

+
1
2
(v
2
+w
2
) , or u

= u

1
2
(v
2
+w
2
), (A.31)
As mentioned in the above, we can see that u

already includes a foreshortening


term. Now the deformed position vector based on the undeformed coordinate
244
system can be rewritten by
r =
_
x +u
e

1
2
_
x
0
(v
2
+w
2
)dx
_
e
u
1
+we
u
3
. (A.32)
Then we obtain the same strain energy as the linear strain energy, while we keep
the foreshortening term in the position vector (Equation A.32). This explicit form
of position vector does not produce an additional energy due to the foreshortening
term, which comes from a moderately large blade deformation, but aects the
kinetic energy.
A.3 Deformed Coordinate System
In this section, the strain measure is expressed in the mixed coordinate system.
The deformed coordinate in e
d
i
and the undeformed coordinates y and z in e
u
i
are combined together to get the simple expression.
Among the strain measure, the only underline terms of the axial strain com-
ponent
o
xx
is altered (see Equation A.19). Referring to the previous sections, the
strain measure in the mixed coordinate system is given by

=
o

+n k

,
o

=
L

+
NL

(A.33)

s
=
o
s
+n k
s
,
o
s
=
L
s
+
NL
s
(A.34)
where

= u

e
y
o
v

z
o
w

(A.35)

NL

= (
2
+

2
)
_

o
(

+w

) +
1
2

2
+

_
+

( z
o
v

y
o
w

) (A.36)
k
L

= w

cos

v

sin

+q

(A.37)

L
s
=
t
s

(A.38)

NL
s
=
t
s
w

(A.39)
k
L
s
= 2

(A.40)
245
where some approximation has been made using ordering scheme of O(
2
). Note
that the axial displacement u
e
is measured in the direction of e
d
1
, while the lateral
displacements v and w are measured in the directions of e
u
2
and e
u
3
, respectively.
In Chapter 2, the strain measure or strains represent Equations A.33-A.40,
where overbar

() in and

will be dropped.
Appendix B
Rotor System Matrices and Force
Vectors
Total system matrices and force vectors are dened as follows:
[M] = [M]
k
+ [M]
a
(B.1)
[C] = [C]
k
+ [C]
a
(B.2)
[K] = [K]
s
+ [K]
k
+ [K]
a
(B.3)
F = F
s
+F
k
+F
a
+F
f
(B.4)
where the superscript s and k represent the contributions from strain energy and
kinetic energy of rotor blades, respectively. The superscript a denotes the con-
tribution from aerodynamic loads on the rotor. The superscript f indicates the
contribution from inertial and aerodynamic loads on trailing-edge aps.
In this study, the aerodynamic and inertial contributions of trailing-edge aps
are treated as an additional load vector. This is discussed in Chapter 3. In this
Appendix B, among others, the system matrices and force vectors by strain energy
and kinetic energy of rotor blades will be given . The underline terms indicate
terms kept in Reference [76], throughout this Appendix B.
247
B.1 Strain Energy of Rotor Blades
The system matrices and force vectors due to the strain energy of rotor blades are
dened in this section.
B.1.1 Stiness Coecients of Composite Beam
The stiness coecients of composite beam are dened as
K
11
=
_
A
11
ds (B.5)
K
12
=
_
(A
11
+B
11
cos
c
) ds (B.6)
K
13
=
_
(A
11
B
11
sin
c
) ds (B.7)
K
14
=
_
(
T
A
11
+qB
11
) ds (B.8)
K
15
=
_
_

t
s
A
16
+ 2B
16
_
ds (B.9)
K
22
=
_
_

2
A
11
2B
11
cos
c
+D
11
cos
2

c
_
ds (B.10)
K
23
=
_
(A
11
B
11
cos
c
+B
11
sin
c
D
11
cos
c
sin
c
) ds (B.11)
K
24
=
_
(
T
A
11
qB
11

T
B
11
cos
c
+qD
11
cos
c
) ds (B.12)
K
25
=
_
_

t
s
A
16
2B
16
+
t
s
B
16
cos
c
+ 2D
16
cos
c
_
ds (B.13)
K
33
=
_
_

2
A
11
+ 2B
11
sin
c
+D
11
sin
2

c
_
ds (B.14)
K
34
=
_
(
T
A
11
qB
11
+
T
B
11
sin
c
qD
11
sin
c
) ds (B.15)
K
35
=
_
_

t
s
A
16
2B
16

t
s
B
16
sin
c
2D
16
sin
c
_
ds (B.16)
K
44
=
_
_

T

t
s
A
16
+q
t
s
B
16
2
T
B
16
+ 2qD
16
_
ds (B.17)
K
45
=
_
_
(
t
s
)
2
A
66
+ 4
t
s
B
66
+ 4D
66
_
ds (B.18)
K
55
=
_
_
q
t
s
A
66
+
t
s

T
A
16
+ 2qB
66
+ 2
T
B
16
_
ds (B.19)
248
B.1.2 Stiness Matrices and Force Vectors
The linear stiness matrices, [K]
s
, are given by
[K
uu
] =
_
K
11
H
T
u

H
u

dx (B.20)
[K
uw
] =
_
K
13
sin
o
H
T
u

H
w

dx (B.21)
[K
uv
] =
_
K
13
cos
o
H
T
u

H
v

dx (B.22)
[K
up
] =
_
_
K
15
H
T
u

H
p

+Ak
2
11
H
T
u

p
_
dx (B.23)
[K
ww
] =
_
(K
22
cos
2

o
+K
33
sin
2

o
) H
T
w

H
w

dx (B.24)
[K
wv
] =
_
(K
33
K
22
) cos
o
sin
o
H
T
w

H
v

dx (B.25)
[K
wp
] =
_
(K
25
cos
o
+K
35
sin
o
)
+ (AB
1
sin
o
+AB
2
cos
o
)

o
H
T
w

p
dx (B.26)
[K
vv
] =
_
(K
33
cos
2

o
+K
22
sin
2

o
) H
T
v

H
v

dx (B.27)
[K
vp
] =
_
(K
35
cos
o
K
25
sin
o
)
+ (AB
1
cos
o
AB
2
sin
o
)

o
H
T
v

p
dx (B.28)
[K
pp
] =
_
(K
44
H
T
p

H
p

+K
55
H
T
p

H
p

) dx (B.29)
and the nonlinear forces, F
s
, are given by
F
u

NL
=
_ _
Ak
2
11
(
1
2

2
+

o
v

)
+

K
13
(w

cos
o
v

sin
o
) +

K
15
v

_
H
T
u

dx (B.30)
F
w

NL
=
_
_
(

K
33


K
22
)(v

cos 2
o
+w

sin 2
o
) +

K
13
cos
o
u

+

K
25
(v

cos
o

sin
o
) +

K
35
(v

sin
o
+

cos
o
)
_
H
T
w

dx

_
_

K
25
(v

cos
o
v
2
sin
o
) +

K
35
(v

sin
o
+v
2
cos
o
)
+

K
55
v

+Ak
2
11

o
u

e
v

+

K
15
u

e
v

_
H
T
w

dx (B.31)
249
F
v

NL
=
_
_
(

K
33


K
22
)(w

cos 2
o
v

sin 2
o
)

K
13
sin
o
u

+

K
55
w

+Ak
2
11

o
u

e
w

+

K
15
u

e
w

+

K
25
_
(w

) cos
o
2w

sin
o
_
+

K
35
_
(w

) sin
o
+ 2w

cos
o
_ _
H
T
v

dx (B.32)
F
p

NL
=
_
_

K
44
w

H
T
p

+
_

K
35
(w

cos
o
v

sin
o
)


K
25
(w

sin
o
+v

cos
o
) +

K
55
w

+Ak
2
11
u

_
H
T
p

+
_
(

K
33


K
22
)
_
v

cos 2
o
+ (w
2
v
2
) sin
o
cos
o
_
+

K
13
(w

e
cos
o
v

e
sin
o
) +

K
35
(w

cos
o
v

sin
o
)


K
25
(w

sin
o
+v

cos
o
)
_
H
T
p
_
dx (B.33)
B.2 Kinetic Energy of Rotor Blades
The system matrices and force vectors due to the kinetic energy of rotor blades
are dened in this section. Note that, derivatives with respect to the azimuth are
represented by

(), instead of using ()

, in this section.
B.2.1 Stiness Matrix
The linear stiness matrices, [K]
k
, are given by
[K
uu
] =
_
mH
T
u
H
u
dx (B.34)
[K
uw
] =
_
_
me
g
sin
o
H
T
u
H
w

+m
p
H
T
u
H
w
_
dx
+
_
me
g
_
(

2
o
+ 2

o
) sin
o

o
cos
o
_
H
T
u
H
w

dx (B.35)
[K
uv
] =
_
me
g
cos
o
H
T
u
H
v

dx
+
_
me
g
(

2
o
cos
o
+

o
sin
o
) H
u
T
H
v

dx (B.36)
[K
up
] =
_
2me
g

o
cos
o
H
T
u
H
p
dx (B.37)
[K
wu
] = 0 (B.38)
250
[K
ww
] =
_
F
A
(x) H
T
w

H
w

dx

_
m(k
2
m1
cos
2

o
+k
2
m2
sin
2

o
) (1 +

2
o
)H
T
w

H
w

dx
+
_
mcos
o
sin
o
(k
2
m2
k
2
m1
)

o
H
T
w

H
w

dx (B.39)
[K
wv
] =
_
m(k
2
m1
cos
2

o
+k
2
m2
sin
2

o
)

o
H
T
w

H
v

dx
+
_
mcos
o
sin
o
(k
2
m1
k
2
m2
) (1 +

2
o
)H
T
w

H
v

dx (B.40)
[K
wp
] =
_
me
g
_

2
o
+

o
_
H
T
w
H
p
dx
+
_
m
_
e
g
x cos
o
+ 4

o
sin
o
cos
o
(k
2
m1
k
2
m2
)
_
H
T
w

H
p
dx (B.41)
[K
vu
] = 0 (B.42)
[K
vw
] =
_
2me
g

o
cos
o
H
T
v
H

w
dx
+
_
(k
2
m1
sin
2

o
+k
2
m2
cos
2

o
)

o
H
T
v

w
dx
+
_
cos
o
sin
o
(k
2
m1
k
2
m2
)(1 +

2
o
)H
T
v

w
dx (B.43)
[K
vv
] =
_
F
A
(x) H
T
v

H
v

dx

_
_
mH
T
v
H
v
+ 2me
g

o
sin
o
H
T
v
H
v

_
dx

_
(k
2
m1
sin
2

o
+k
2
m2
cos
2

o
)(1 +

2
o
)H
T
v

v
dx
+
_
cos
o
sin
o
(k
2
m1
k
2
m2
)

o
H
T
v

v
dx (B.44)
[K
vp
] =
_
me
g
_
sin
o
+

2
o
sin
o

o
cos
o
_
H
T
v
H
p
dx
+
_
me
g
_
2

o
(k
2
m1
k
2
m2
) cos 2
o
x sin
o
_
H
T
v

H
p
dx (B.45)
[K
pu
] = 0 (B.46)
[K
pw
] =
_
m
_
xe
g
cos
o
+ sin 2
o
(k
2
m2
k
2
m1
)

o
_
H
T
p
H
w

dx (B.47)
[K
pv
] =
_
me
g
_
sin
o
H
T
p
H
v
x sin
o
H
T
p
H
v

_
dx (B.48)
[K
pp
] =
_
m(k
2
m2
k
2
m1
) cos 2
o
H
T
p
H
p
dx (B.49)
251
where F
A
(x) denotes the centrifugal force and has been dened as
F
A
(x) =
_
1
x
mdx (B.50)
B.2.2 Mass Matrix
The linear mass matrices, [M]
k
, are given by
[M
uu
] =
_
mL
T
u
L
u
dx (B.51)
[M
uw
] =
_
me
g
sin
o
L
T
u
L
w

dx (B.52)
[M
uv
] =
_
me
g
cos
o
L
T
u
L
v

dx (B.53)
[M
up
] = 0 (B.54)
[M
wu
] = 0 (B.55)
[M
ww
] =
_
_
mL
T
w
L
w
+m(k
2
m1
cos
2

o
+k
2
m2
sin
2

o
)H
T
w

H
w

_
dx (B.56)
[M
wv
] =
_
m(k
2
m2
k
2
m1
) sin
o
cos
o
H
T
w

H
v

dx (B.57)
[M
wp
] =
_
me
g
cos
o
H
T
w
H
p
dx (B.58)
[M
vu
] = 0 (B.59)
[M
vw
] =
_
m(k
2
m2
k
2
m1
) sin
o
cos
o
H
T
v

H
w

dx (B.60)
[M
vv
] =
_
_
mH
T
v
H
v
+m(k
2
m1
sin
2

o
+k
2
m2
cos
2

o
)H
T
v

H
v

_
dx (B.61)
[M
vp
] =
_
me
g
sin
o
H
T
v
H
p
dx (B.62)
[M
pu
] = 0 (B.63)
[M
pw
] =
_
me
g
cos
o
H
T
p
H
w
dx (B.64)
[M
pv
] =
_
me
g
sin
o
H
T
p
H
v
dx (B.65)
[M
pp
] =
_
mk
2
m
H
T
p
H
p
dx (B.66)
252
B.2.3 Damping Matrix
The linear mass matrices, [C]
k
, are given by
[C
uu
] = 0 (B.67)
[C
uw
] =
_
2me
g

o
cos
o
H
T
u
H
w

dx (B.68)
[C
uv
] =
_
_
2mH
T
u
H
v
+ 2me
g

o
sin
o
H
T
u
H
v

_
dx (B.69)
[C
up
] =
_
2me
g
sin
o
H
T
u
H
p
dx (B.70)
[C
wu
] = 0 (B.71)
[C
ww
] =
_
m

o
(k
2
m2
k
2
m1
) sin 2
o
H
T
w

H
w

dx (B.72)
[C
wv
] =
_
_
2m
p
H
T
w
H
v
+ 2me
g
sin
o
H
T
w

H
v
2m

o
(k
2
m1
cos
2

o
+k
2
m2
sin
2

o
)H
T
w

H
v

_
dx (B.73)
[C
wp
] =
_
_
2me
g

o
sin
o
H
T
w
H
p
+ 2m(k
2
m1
cos
2

o
+k
2
m2
sin
2

o
)H
T
w

H
v

_
dx (B.74)
[C
vu
] =
_
2mH
T
v
H
u
dx (B.75)
[C
vw
] =
_
_
2m
p
H
T
v
H
w
2me
g
sin
o
H
T
v
H
w

+ 2m

o
(k
2
m1
sin
2

o
+k
2
m2
cos
2

o
)H
T
v

H
w

_
dx (B.76)
[C
vv
] =
_
_
2me
g
cos
o
H
T
v
H
v

+ 2me
g
cos
o
H
T
v

H
v
+ 2m

o
(k
2
m1
k
2
m2
) sin
o
cos
o
H
T
v

H
v

_
dx (B.77)
[C
vp
] =
_
_
2me
g

o
cos
o
H
T
v
H
p
+ msin 2
o
(k
2
m1
k
2
m2
)H
T
v

H
p
_
dx (B.78)
[C
pu
] = 0 (B.79)
[C
pw
] =
_
2m(k
2
m1
cos
2

o
+k
2
m2
sin
2

o
)H
T
p
H
w

dx (B.80)
[C
pv
] =
_
_
2me
g

p
cos
o
H
T
p
H
v
253
2m(k
2
m1
k
2
m2
) sin
o
cos
o
H
T
p
H
v

_
dx (B.81)
[C
pp
] = 0 (B.82)
B.2.4 Force Vectors
The constant force vectors, F
k
, are dened as
F
k
= F
k
o
+F
k
NL
(B.83)
where the subscript o denotes the constant terms and NL denotes the nonlinear
terms, and
F
u

o
=
_
m
_
x e
g
(
p
+ 2

) sin
o
_
H
T
u
dx (B.84)
F
w

o
=
_
m
_
x
p
e
g

o
cos
o
+e
g
sin
o
(

2
o
+ 2

p
)
_
H
T
w
dx
+
_
m
_
x sin
o
+ (k
2
m1
cos

o
+k
2
m2
sin
2

o
)(
p
+ 2

o
)
_
H
T
w

dx (B.85)
F
v

o
=
_
m
_
e
g
cos
o
(1 + 2

p
+

2
o
) + e
g

o
sin
o
_
H
T
v
dx
+
_
m
_
e
g
x cos
o
+ cos
o
sin
o
(
p
+ 2

o
)
_
H
T
v

dx (B.86)
F
p

o
=
_
m
_
(k
2
m1
k
2
m2
) sin
o
cos
o
k
2
m

o
e
g
x
p
cos
o
_
H
T
p
dx (B.87)
and the nonlinear force vector terms are given as follows:
F
u

NL
=
_
m
__
x
0
( v

+v

+ w

+w

) d
_
H
T
u
dx (B.88)
F
w

NL
=
_ _
2w

_
1
x
m vd
_
H
T
w

dx (B.89)
F
v

NL
=
_ _
2m
_
x
0
(v

+w

)d
_
H
T
v
dx

_ _
2v

_
1
x
m vd
_
H
T
v

dx (B.90)
Appendix C
Formulations using Mathematica
In this Appendix, the Mathematica 4.0 programs used to derive the rotor strain
energy, rotor kinetic energy, rotor quasi-aerodynamic loads, and trailing-edge aps
inertial loads are presented. Some functions, which are modied based on the func-
tions developed by Milgram [69], are provided to implement the ordering scheme.
The rst three sections describe the strain energy, kinetic energy, quasi-steady
aerodynamic contributions of rotor blade based on the ordering scheme presented
in Table 2.2. In the fourth section, the trailing-edge aps inertial contributions
are derived using the ordering schemes presented in Tables 2.2 and 3.1.
255
C.1 Rotor Strain Energy
256
257
258
259
C.2 Rotor Kinetic Energy
260
261
262
263
264
265
266
C.3 Rotor Quasi-steady Aerodynamic Loads
267
268
269
270
271
272
273
C.4 Trailing-Edge Flaps Inertial Loads
274
275
276
277
278
Bibliography
[1] Laing, E. M., Army Helicopter and Vibration Survey Methods and Results,
Journal of the American Helicopter Society, Vol. 19, No. 2, 1974, pp. 430.
[2] Wilkie, W.K., Anisotropic Piezoelectric Twist Actuation of Helicopter Ro-
tor Blades: Aeroelastic Analysis and Design Optimization, Ph.D dissertation,
Uiniversity of Colorado, Department of Aerospace Engineering Sciences, De-
cember 1997.
[3] Shin, S., Integral Twist Actuation of Helicopter Rotor Blades for Vibration
Reduction, Ph.D dissertation, MIT, Department of Aeronautics and Astro-
nautics, August 2001.
[4] Anusonti-Inthra, P. Semi-Active Control of Helicopter Vibration using Con-
trollable Stiness and Damping Devices, Ph.D Dissertation, The Pennsylvania
State University, Department of Aerospace Engineering, August 2002.
[5] Padeld, G.D., Helicopter Flight Dynamics: The Theory and Application of
Flying Qualities and Simulation Modeling, AIAA Education Series, AIAA Inc.,
1996.
[6] Johnson, W., Helicopter Theory, Dover Publications, Inc., New York, 1980.
[7] Shaw, J. and Albion, N., Active Control of Rotor Blade Pitch for Vibration
Reduction: A Wind Tunnel Demonstration, Vertica, Vol. 4, 1980, pp. 311.
[8] Miao, W., Kottapalli, S.B.R. and Frye, H. M., Flight Demonstration of
Higher Harmonic Control (HHC) on S-76, Proceedings of the 42nd AHS Fo-
rum, Washington D.C., June 24 1986, pp. 777791.
[9] Walsh, D. M., Flight Test of an Open Loop Higher Harmonic Control System
on an S-76A Helicopter, Proceedings of the 42nd AHS Forum, Washington
D.C., June 24, 1986, pp. 831843.
280
[10] Nguyen, K. and Chopra, I., Application of Higher Harmonic Control to Rotor
Operating at High Speed and Thrust, Journal of the American Helicopter
Society, July 1990, pp. 7889.
[11] Robinson, L. H. and Friedmann, P. P., A Study of Fundamental Issues in
Higher Harmonic Control Using Aeroelastic Simulation, Journal of the Amer-
ican Helicopter Society, April. 1991, pp. 3243.
[12] Ham, N. D., A Simple System for Helicopter Individual Blade Control Using
Modal Decomposition, Vertica, Vol. 4, 1980, pp. 2328.
[13] Guinn, K. F., Individual Blade Control Independent of a Swashplate, Jour-
nal of the American Helicopter Society, July 1982, pp. 2531.
[14] Ham, N. D., Helicopter Individual-Blade-Control and Its Applications, Pro-
ceedings of the 39th AHS Forum, St. Louis, MO, May 911, 1983, pp. 613623.
[15] Jacklin, S. A., Leyland, J. A. and Blaas, A., Full-Scale Wind Tunnel In-
vestigation of a Helicopter Individual Blade Control System, Proceedings of
34th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and
Materials Conference, AIAA-93-1361-CP, Lajolla, CA April 1922, 1993, pp.
576586.
[16] Welsh, W. A., Von Hardenberg, P. C. and Staple, A. E., Test and Evaluation
of Fuselage Vibration Utilizing Active Control of structural Response (ACSR)
Optimized to ADS-27, Proceedings of the 46th AHS Forum, Washington, DC,
May 2123, 1990, pp. 2137.
[17] Staple, A. E., Evaluation of Active Control of Structural Responses as a
Means of Reducing Helicopter Vibration, Proceedings of the 46th AHS Fo-
rum, Washington, DC, May 2123, 1990, pp. 317.
[18] Welsh, W., Fredrickson, C., Rauch, C. and Lyndon, L., Flight Test of an
Active Vibration Control System on the UH-60 Black Hawk Helicopter, Pro-
ceedings of the 51st AHS Forum, Fort Worth, TX, May 911,1995, pp. 393
402.
[19] Chiu, T., and Friedmann, P. P., Vibration Suppression in Helicopter Ro-
tor/Flexible Fuselage System Using the ACSR Approach With Disturbance
Rejection, Proceedings of the 52nd AHS Forum, Washington D.C., June 46,
1996, pp. 736757.
[20] Goodman, R. K. and Millott, T. A., Design, Development, and Flight Testing
of the Active Vibration Control System for the Sikorsky S92, Proceedings of
the 56th AHS Forum, Virginia Beach, Virginia, May 24, 2000.
281
[21] Cribbs, R., and Friedmann, P. P., Vibration Reduction in Ro-
torcraft using an Enhanced ACSR Model, Proceedings of 41st
AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and
Materials Conference, AIAA-2000-1687, Atlanta, GA, April 36, 2000.
[22] Chen, P.C., and Chopra, I., Hover Testing of Smart Rotor with Induced-
Strain Actuation of Blade Twist, AIAA Journal, Vol. 35. No. 1, 1997, pp.
616.
[23] Cesnik, C.E.S., Shin, S., and Wilbur, M.L., Dynamic Response of Active
Twist Rotor Blades, Smart Materials and Structures, Vol. 10, No. 1, 2001,
pp. 6276.
[24] Shin, S., and Cesnik, C.E.S., Helicopter Vibration Reduction in For-
ward Flight using Blade Integral Twist Control, Proceedings of 43rd
AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and Mate-
rials Conference, Vol. 3, 2002, pp. 19881997.
[25] Sekula, M.K., Wilbur, M.L., Yeager Jr., W.T., Aerodynamic Design Study of
an Advanced Active Twist Rotor, AHS 4th Decennial Specialists Conference
on Aeromechanics, 2004, pp. 4960.
[26] McCloud III, J.L., An Analytical Study of a Multicyclic Controllable Twist
Rotor, Proceedings of the 31st AHS forum, May 1975.
[27] Millott, T. A., and Friedmann, P. P., Vibration Reduction in Helicopter
Rotors Using an Active Control Surface located on the Blade, Proceedings of
33rd AIAA/ASME/ASCE/AHS/ASC/ Structures, Structural Dynamics and
Materials Conference, AIAA-92-2451-CP, Dallas, TX, April 1315, 1992, pp.
19751988.
[28] Milgram, J.H., and Chopra, I., Helicopter Vibration Reduction with Trailing-
Edge Flaps, Proceedings of the 36th AIAA/ASME/ASCE/AHS/ASC Struc-
tures, Structural Dynamics and Materials Conference, New Orleans, LA, April
1995, pp. 601612.
[29] Straub, F. K. and Hassan, A. A., Aeromechanic Considerations in the Design
of a Rotor with Smart Material actuated Trailing Edge Flaps, Proceedings
of the 52nd AHS Forum, Washington D.C., June 46, 1996, pp. 704714.
[30] Fulton, M. V., and Ormiston, R., Small-Scale Rotor Experiments with On-
Blade Elevons to Reduce Blade Vibratory Loads in Forward Flight, Pro-
ceedings of the 54th AHS Forum, Washington, D.C., May 2022 1998, pp.
433451.
282
[31] Straub, F. K. and Charles, B. D., Aeroelastic Analysis Of Rotors With Trail-
ing Edge Flaps Using Comprehensive Codes, Journal of the American Heli-
copter Society, Vol. 46, No. 3, July 2001, pp. 192199.
[32] Friedmann, P. P., Rotary-Wing Aeroelasticity - Current Status and Future
Trends, Proceedings of the 39th AIAA Aerospace Sciences Meeting and Ex-
hibit, AIAA paper No. 2001-0427, Reno, Nev., January 2001.
[33] Friedmann, P. P., Vibration Reduction in Rotorcraft Using Actively Con-
trolled Flaps - From Theoretical Concept To Flight Ready Hardware, Pro-
ceedings of the AHS 4th Decennial Specialists Conference on Aeromechanics,
San Francisco, CA, January 2123, 2004.
[34] Chopra, I., Status of Application of Smart Structures Technology to Rotor-
craft Systems, Journal of the American Helicopter Society, Vol. 45, No. 4,
2000, pp. 228252.
[35] Loewy, R., Helicopter Vibrations: A Technological Perspective, Journal of
the American Helicopter Society, October 1984, pp. 430.
[36] Bielawa, R., Rotary Wing Structural Dynamics and Aeroelasticity, AIAA Ed-
ucation Series, AIAA Inc., 1992.
[37] Hamouda, N. H. and Pierce, G. A., Helicopter Vibration Suppression Using
Simple Pendulum Absorbers on the Rotor Blade, Journal of the American
Helicopter Society, July 1984, pp. 1929.
[38] Kottapalli, S., Sources Of Helicopter Rotor Hub Inplane Shears, Proceed-
ings of 34th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynam-
ics, and Materials Conference, La Jolla, CA, Apr 1922 1993.
[39] Schuett, E., Application of Passive Rotor Isolation for Alleviation of Rotor
Induced Vibration, Journal of the American Helicopter Society, April. 1969,
pp. 3448.
[40] Viswanathan, S. P. and Myers, A. W., Reduction of Helicopter Vibration
Through Control of Hub-Impedance, Journal of the American Helicopter
Society, October 1980, pp. 312.
[41] Braun, D., Development of Antiresonant Force Isolators for Helicopter Vibra-
tion Reduction, Journal of the American Helicopter Society, October 1982,
pp. 3744.
[42] Friedmann, P. P. and Shanthakumaran, P., Optimum Design of Rotor Blades
for Vibration Reduction in Forward Flight, Journal of the American Heli-
copter Society, October 1984, pp. 7080.
283
[43] Davis, M. W. and Weller, W. H., Application of Design Optimization Tech-
niques to Rotor Dynamics Problems, Journal of the American Helicopter
Society, July 1988, pp. 4250.
[44] Lim, J. W., Aeroelastic Optimization of a Helicopter Rotor, Ph.D Dissertation,
University of Maryland, College Park, Department of Aerospace Engineering,
1988.
[45] Lim, J. W. and Chopra, I., Aeroelastic Optimization of a Helicopter Rotor,
Journal of the American Helicopter Society, January 1989, pp. 5262.
[46] Friedmann, P. P., Helicopter Vibration Reduction Using Structural Opti-
mization with Aeroelastic/ Multidisciplinary Constraints - A Survey, Journal
of Aircraft, Vol. 28, No. 1, January 1991, pp. 821.
[47] Young, D. K. and Tarzanin, F. J. Jr., Structural Optimization and Mach
Scale Test Validation of a Low Vibration Rotor, Proceedings of the 47th
AHS Forum, Phoenix, Arizona, May 68, 1991, pp. 955968.
[48] Ganguli, R., and Chopra, I., Aeroelastic Optimization of An Advanced Ge-
ometry Composite Helicopter Rotor, Proceedings of the 51st AHS Forum,
Fort Worth, TX, May 911 1995, pp. 965984.
[49] Yuan, K. A., and Friedmann P. P., Structural Optimization for Vibration
Reduction of Composite Helicopter Rotor Blades with Advanced Geometry
Tips Subject to Multidisciplinary Constraints, Proceedings of the 51st AHS
Forum, Fort Worth, TX, May 911,1995, pp. 937956.
[50] Tarzanin, F. J., Young, D.K., Panda, B. Advanced Aeroelastic Optimization
Applied to an Improved Performance, Low Vibration Rotor, Proceedings of
the 55th AHS Forum, Montreal, Canada, May 2527, 1999, pp. 183194.
[51] Bao, J., Chopra, I., Bernhard, A.P.F. and Nagaraj, V.T., Design and hover
test of low vibration mach scale rotor with twisted composite tailored blades,
Proceedings of the 44th AIAA/ASME/ASCE/AHS/ASC Structures, Struc-
tural Dynamics and Materials Conference, Norfolk, VA, Apr 710 2003.
[52] Friedmann, P. P., Vibration Reduction in Helicopter Rotors Using Actively
Controlled Trailing Edge Flaps, The Eighth ARO Workshop on Aeroelasticity
of Rotorcraft Systems, The Nittany Lion Inn, State College, Pennsylvania,
October 1820, 1999.
[53] Bent, A.A. and Hagood, N.W., Piezoelectric Fiber Composites with Inter-
digitated Electrodes, Journal of Intelligent Material Systems and Structures,
Vol. 8, 1997, pp. 903919.
284
[54] Wilkie, K.W., Bryant, R.G., High, J.W., Fox, R.L., Hellbaum, Jalink, Jr.A.,
Little, B.D., and Mirick, P.H., Low-Cost Piezocomposite Actuator for Struc-
tural Control Applications, Proceedings of SPIE Smart Structures and Ma-
terials Conference, Newport Beach, CA, March 59, 2000.
[55] Zhang, J., Smith, E. C. , and Wang, K. W., An Active-Passive Hybrid
Method for Optimization of Rotor Blade with Trailing Edge Flaps, Pro-
ceedings of 40th AIAA/ ASME/ASCE/AHS/ASC Structures, Structural Dy-
namics and Materials Conference, AIAA-99-1222, St. Louis, Mo. April 1214,
1999.
[56] Zhang, J., Smith, E. C. , and Wang, K. W., Active Passive Hybrid Opti-
mization of Rotor Blades with Trailing Edge Flaps, Proceedings of the 56th
AHS Forum, Virginia Beach, VA, May 24, 2000.
[57] Zhang, J., Smith, E. C. , and Wang, K. W., Active Passive Hybrid Opti-
mization of Rotor Blades with Trailing Edge Flaps, Journal of the American
Helicopter Society, Vol. 49, No. 1, January 2004, pp. 5465.
[58] Anusonti-Inthra, P., and Gandhi, F., Helicopter Vibration Reduction
Through Cyclic Variations in Blade Root Stiness, Journal of Intelligent
Material Systems and Structures, Vol. 11, No. 2, Feb 2000, pp. 153166.
[59] Anusonti-Inthra, P., and Gandhi, F., Optimal Control of Helicopter Vibra-
tion Through Cyclic Variations in Blade Root Stiness, Smart Materials and
Structures, Vol. 10, No. 1, Feb 2001, pp. 8695.
[60] Payne, P. R., Higher Harmonic Rotor Control, Aircraft Engineering, Vol.
30, No. 354, August 1958, pp. 222226.
[61] McCloud III, J. L., and Weisbrich, A. L., Wind Tunnel Test Results of a
Full-Scale Multicyclic Controllable Twist Rotor, Proceedings of the 34th AHS
Forum, Washington, DC, May 1978.
[62] Millott, T. A., and Friedmann, P. P., The Practical Implementation of An
Actively Controlled Flap to Reduce Vibration in Helicopter Rotors, Pro-
ceedings of the 49th AHS Forum, St. Louis, Missouri, May 1921 1993, pp.
10791092.
[63] Millott, T. A., Vibration Reduction in Helicopter Rotors Using an Active Con-
trolled Partial Span Trailing Edge Flap Located on the Blade, Ph.D Disser-
tation, University of California, Department of Aerospace Engineering, Los
Angeles, CA, 1993.
285
[64] Millott, T. A., and Friedmann, P. P., Vibration Reduction in Hingeless Rotors
Using an Actively Controlled Trailing Edge Flap: Implementation and Time
Domain Simulation, Proceedings of 35th AIAA/ASME/ASCE/AHS/ASC/
Structures, Structural Dynamics and Materials Conference, AIAA-94-1306-
CP, Hilton Head, SC, April 1820, 1994, pp. 822.
[65] Friedmann, P. P. and Millott,T. A., Vibration Reduction in Rotorcraft Using
Active Control: A Comparison of Various Approaches, Journal of Guidance,
Control and Dynamics, Vol. 18, No. 4, August 1995, pp. 664673.
[66] Dawson, S., and Straub, F. K., Design, Validation, and Test of a Model
Rotor with Tip Mounted Active Flaps, Proceedings of the 50th AHS Forum,
Washington, DC, May 1113, 1994, pp. 361372.
[67] Dawson, S., Hassan, A., Straub, F.K., and Tadghighi,H., Wind Tunnel Test
of an Active Flap Rotor: BVI Noise and Vibration Reduction, Proceedings
of the 51st AHS Forum, Fort Worth, TX, May 911, 1995.
[68] Straub, F. K., Active Flap Control for Vibration Reduction and Performance
Improvement, Proceedings of the 51st AHS Forum, Fort Worth, TX, May 9
11, 1995, pp. 381392.
[69] Milgram, J.H., A Comprehensive Aeroelastic Analysis of Helicopter Main Ro-
tors with Trailing Edge Flaps for Vibration Reduction, Ph.D dissertation, Uni-
versity of Maryland, Department of Aerospace Engineering, College Park, MD,
1997.
[70] Milgram, J., Chopra, I., and Straub, F., Parametric Design Study For Ac-
tively Controlled Trailing Edge Flaps, Journal of American Helicopter Soci-
ety, Vol. 43, No. 2, April 1998, pp. 110119.
[71] Milgram, J., Chopra, I., and Straub, F., Rotors with Trailing Edge Flaps:
Analysis and Comparison With Experimental Data, Journal of American
Helicopter Society, Vol. 43, No. 4, October 1998, pp. 319332.
[72] Hariharan, N., Unsteady Aerodynamics of a Flapped Airfoil in Subsonic Flow
Using Indicial Concepts, M.S. Thesis, University of Maryland, Department of
Aerospace Engineering, College Park, MD, June 1995.
[73] Hariharan, N., and Leishman, J., Unsteady Aerodynamics Of A Flapped
Airfoil In Subsonic Flow By Indicial Concepts, Journal of Aircraft, Vol. 33,
No. 5, Sep-Oct, 1996, pp. 855868.
[74] Bagai, A., Contributions to the Mathematical Modeling of Rotor Flow-Fields
Using a Pseudo-Implicit Free-Wake Analysis, Ph.D Dissertation, University
of Maryland, Department of Aerospace Engineering, College Park, MD, 1995.
286
[75] Bagai, A., and Leishman, J., Rotor Free-Wake Modeling Using A Pseudo-
Implicit Technique - Including Comparisons With Experimental Data, Jour-
nal of the American Helicopter Society, Vol. 40, No. 3, July, 1995, pp. 2941.
[76] Anonymous, University of Maryland Advanced Rotorcraft Code (UMARC)
Theory Manual, UM-AERO 94-18, The University of Maryland, 1994.
[77] Myrtle, T.F. and Friedmann, P.P., New Comprehensive Time Domain Un-
steady Aerodynamics for Flapped Airfoils and Its Application to Rotor Vibra-
tion Reduction Using Active Control, Proceedings of the 53rd AHS Forum,
Virginia Beach, VA, April 29 May 1, 1997.
[78] Myrtle, T.F. and Friedmann, P.P., Vibration Reduction in Rotorcraft Using
the Actively Controlled Trailing Edge Flap and Issues Related to Practical
Implementation, Proceedings of the 54th AHS Forum, Washington, D.C.,
May 2022, 1998.
[79] Fulton, M. V., and Ormiston, R., Hover Testing of a Small-Scale Rotor with
On-Blade Elevons, Proceedings of the 53rd AHS Forum, Virginia Beach, VA,
April 29May 1, 1997, pp. 249273.
[80] de Terlizzi, M., and Friedmann, P. P., Active Control of BVI Induced Vibra-
tions Using a Rened Aerodynamic Model and Experimental Correlation,
Proceedings of the 55th AHS Forum, Montreal, Canada, May 2527, 1999, pp.
599618.
[81] Johnson, W., Comprehensive Analytical Model of Rotorcraft Aerodynamics
and Dynamics, Johnson Aeronautics, Palo Alto, CA, 1988.
[82] Johnson, W., CAMRAD II Nonlinear Substructure Aeromechanics Analysis,
Johnson Aeronautics, Palo Alto, CA, Version 3.1, 1998.
[83] Shen, J., and Chopra, I., Aeroelastic Modeling Of Trailing-Edge
Flaps with Smart Material Actuators, Proceedings of the 36th
AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and
Materials Conference, Atlanta, GA, AIAA-2000-1622, April 36, 2000.
[84] Shen, J., and Chopra, I., Aeroelastic Modeling Of Trailing-Edge-Flap He-
licopter Rotors Including Actuator Dynamics, Journal of Aircraft, Vol. 41,
No. 6, November/December, 2004, pp. 14651472.
[85] Shen, J., and Chopra, I., Aeroelastic Stability Of Trailing-Edge Flap Heli-
copter Rotors, Journal of the American Helicopter Society, Vol. 48, No. 4,
October 2003, pp. 236243. 14651472.
287
[86] Roget, B, and Chopra, I., Robust Individual Blade Control Algorithm For A
Dissimilar Rotor, Journal of Guidance, Control, and Dynamics, Vol. 25, No.
5, September/October 2002, pp. 915923.
[87] Roget, B, and Chopra, I., Individual Blade Control Methodology For A Rotor
With Dissimilar Blades, Journal of the American Helicopter Society, Vol. 48,
No. 3, July 2003, pp. 176185.
[88] Fink, D.A., Hawkey, T.J., and Gaudreau, M.P.J., An Electromagnetic Ac-
tuator for Individual Blade Control, Proceedings of the 56th AHS Forum,
Viginia Beach, VA, May 24, 2000.
[89] Phillips, N.B., and Merkley, D.J., BHTIs Technical Assessment of Advanced
Rotor and Control Concepts, AHS Vertical Lift Aircraft Design Conference,
San Francisco, CA, January 1719, 1990.
[90] Shen, J., and Chopra, I., A Parametric Design Study for a Swashplateless He-
licopter Rotor with Trailing-Edge Flaps, Journal of the American Helicopter
Society, Vol. 49, No. 1, January, 2004, pp. 4353.
[91] Shen, J., and Chopra, I., Swashplateless Helicopter Rotor with Trailing-Edge
Flaps, Journal of Aircraft, Vol. 41, No. 2, March/April, 2004, pp. 208214.
[92] Shen, J., and Chopra, I., Actuation Requirements Of Swashplateless Trailing-
Edge Flap Helicopter Rotor In Maneuvering And Autorotation Flights, Pro-
ceedings of the 60th AHS Forum, Vol. 2, 2004, pp. 16231631.
[93] Cribbs, R. and Friedmann, P.P., Actuator Saturation and its Inuence on
Vibration Reduction by Actively Controlled Flaps, Proceedings of the 42th
AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and Mate-
rial Conference, Seattle, WA, April 2001.
[94] Depailler, G. and Friedmann, P.P., Alleviation of Dynamic Stall Induced
Vibrations using Actively Controlled Flaps, Proceedings of the 58th AHS
Forum, Montreal, Canada, June 1113, 2002.
[95] Liu, L., Patt, D., and Friedmann, P.P., Simultaneous Vibration And Noise
Reduction In Rotorcraft Using Aeroelastic Simulation, Proceedings of the
60th AHS Forum, Vol. 2, Baltimore, MD, 2004, pp. 16481672.
[96] Patt, D., Liu, L., and Friedmann, P.P., Rotorcraft Vibration Reduction And
Noise Prediction Using A Unied Aeroelastic Response Simulation, Journal
of the American Helicopter Society, Vol. 50, No. 1, January 2005, pp. 95106.
[97] Brenter, K., A Computer Program Incorporating Realistic Blade Motions and
Advanced Acoustic Formulation, NASA Technical Memorandum, Vol. 87721,
1986.
288
[98] Loewy, R.G., Recent Developments in Smart Structures with Aeronautical
Applications, Proceedings of the 37th Israel Annual Conference on Aerospace
Sciences, Feburary 2627, 1997.
[99] Friedmann, P.P., The Promise of Adaptive Materials for Alleviating Aeroe-
lastic Problems and Some Concerns, Proceedings of Innovation in Rotorcraft
Technology, London, United Kingdom, June 2425, 1997, pp. 10.110.16
[100] Giurgiutiu, V., Recent Advances in Smart-Material Rotor Control Actua-
tion, Proceedings of the 41st AIAA/ASME /ASCE/AHS /ASC Structures,
Structural Dynamics and Materials Conference, Atlanta, Georgia, April 36,
2000.
[101] Barrett, R., Intelligent Rotor Blade Structures Development using Direction-
ally Attached Piezoelectric Crystals, MS thesis, University of Maryland, De-
partment of Aerospace Engineering, College Park, MD, 1990.
[102] Rodgers, J. P. and Hagood, N. W., Preliminary Mach-scaled Hover Testing
of an Integral Twisted-Actuated Rotor Blade, Proceedings of the SPIE on
Smart Structures and Materials, San Diego, CA, March, 1998.
[103] Wilkie, W.K., Wilbur, M.L., Mirick, P.H., Cesnik, C.E.S., and Shin, S.,
Aeroelastic Analysis of the NASA/ARMY/MIT Active Twist Rotor, Pro-
ceedings of the 55th AHS Forum, Montreal, Canada, May 2527, 1999.
[104] Cesnik, C.E.S., and Shin, S., On the Modeling of Integrally Actuated Heli-
copter Blades, International Journal of Solids and Structures, Vol. 38, 2001,
pp. 17651789.
[105] Samak, D.K., and Chopra, I., A Feasibility Study to Build A Smart Rotor:
Trailing Edge Flap Actuator, SPIE Conference on Smart Structures and
Materials, Feb 1993.
[106] Ben-Zeev, O. and Chopra, I., Advances in the Development of Intelligent
Helicopter Rotors Employing Smart Trailing Edge Flaps, Smart Materials
and Structures, Vol. 5, No. 1, Feb 1996, pp. 1125.
[107] Koratkar, N.A., and Chopra, I., Analysis And Testing Of Mach-Scaled Rotor
With Trailing-Edge Flaps, AIAA Journal, Vol. 38, No. 7, July, 2000, pp.
11131124.
[108] Koratkar, N.A., Spencer, M.G., and Chopra, I., Wind Tunnel Testing Of
A Smart Rotor Model With Trailing-Edge Flaps, Journal of the American
Helicopter Society, Vol. 47, No. 4, October 2002, pp 263272.
289
[109] Lee, T. and Chopra, I., Design Of Piezostack-Driven Trailing-Edge Flap
Actuator For Helicopter Rotors, Smart Materials and Structures, Vol. 10,
No. 1, Feb 2001, pp. 1524.
[110] Lee, T., and Chopra, I., Design of a Bidirectional Piezoelectric Actuator
for Blade Trailing-Edge Flap, SPIE Conference on Smart Structures and
Materials, Vol. 4327, 2001.
[111] Prechtl, E.F and Hall, S.R., An X-Frame Actuator Servoap Actuation Sys-
tem for Rotor Control, SPIEs Smart Structures and Materials Symposium,
San Diego, March 1998.
[112] Straub, F., Whirl Tower Test of the SMART Active Flap Rotor, AHS 4th
Decennial Specialists Conference on Aeromechanics, San Francisco, Jan 21,
2004.
[113] Hall, S.R., Tzianetopoulou, T., Straub, F.K., and Ngo, H.T., Design and
Testing of a Double X-Frame Piezoelectric Actuator, SPIEs Smart Struc-
tures and Materials Symposium, Newport Beach, March 2000.
[114] Enenkl, B. , Kloppel, V., and Preibler, D., Full Scale Rotor with Piezoelec-
tric Actuated Blade Flaps, 28th European Rotorcraft Forum, Bristol, United
Kingdom, Sep 1719, 2002.
[115] Clement, J.W., Brei, D., Moskalik, A.J., Barrett, R., Bench-Top Character-
ization of an active Rotor Blade Flap system Incorporating C-Block Actua-
tors, Proceedings of 39th AIAA/ASME/ASCE/AHS/ASC/Structures, Struc-
tural Dynamics and Materials Conference, Long Beach, CA, AIAA-98-2108,
April 1998.
[116] Centolanza, L.R., Smith, E.C., and Munsky, B., Induced-shear Piezoelec-
tric Actuators for Rotor Blade Trailing Edge Flaps, Smart Materials and
Structures, Vol. 11, Feb. 2002, pp. 2435.
[117] Centolanza, L.R., Induced Shear Piezoelectric Actuators for Smart Rotor
Blades, Ph.D dissertation, The Pennsylvania State University, Department
of Aerospace Engineering, August 2001.
[118] Hagood, N.W. and Von Flotow, A. Damping of Structural Vibrations with
Piezoelectric Materials and Passive Electrical Networks, Journal of Sound
and Vibration, Vol. 146, No. 2, 1991, pp. 243268.
[119] Lesieutre, G.A., and Davis, C.L., Can a Coupling Coecient of a Piezo-
electric Device Be Higher Than Those of Its Active Material, Proceedings of
SPIE, Smart Structures and Materials, Vol. 3041, 1997, pp. 281292.
290
[120] Agnes, G.S., Active/Passive Piezoelectric Vibration Suppression, Proceed-
ings of SPIE, Smart Structures and Materials, Vol. 2193, 1994, pp. 2434.
[121] Morgan, R.A., Adaptive Active-Passive Piezoelectric Vibration Absorbers For
Variable Frequency Harmonic Excitations, The Pennsylvania State University,
Department of Mechanical and Nuclear Engineering, December 2001.
[122] Agnes, G.S., Development of a Modal Model for Simultaneous Active and
Passive Piezoelectric Vibration Suppression, Journal of Intelligent Material
Systems and Structures, Vol. 6, No. 4, 1995, pp. 482487.
[123] Khan, S.P., and Wang, K.W., Structural Vibration Controls via Piezo-
electric Materials with Active-Passive Hybrid Networks, Proceedings of the
ASME IMECE, 1994, pp. 187194.
[124] Tsai, M.S., and Wang, K.W., On the Structural Damping Characteristics
of Active Piezoelectric Actuators with Passive Shunt, Journal of Sound and
Vibration, Vol. 221, No. 1, 1999, pp. 122.
[125] Tang, J. and Wang, K.W., Active-Passive Hybrid Piezoelectric Networks
for Vibration Control: Comparisons and Improvement, Smart Structures and
Materials, Vol. 10, No. 4, 2001, pp. 794806.
[126] Tsai, M.S., and Wang, K.W., Control of a Ring Structure with Multiple
Active-Passive Hybrid Piezoelectric Networks, Smart Materials and Struc-
tures, Vol. 5, No. 5, Oct 1996, pp. 695703.
[127] Tsai, M.S., and Wang, K.W., Integrating Active-Passive Hybrid Piezoelec-
tric Networks with Active Constrained Layer Treatments for Structural Damp-
ing, Proceedings of the ASME IMECE, 1997, pp. 1324.
[128] Horowitz, P. and Hill, W., The Art of Electronics, Cambridge University
Press, 1989.
[129] Bruneau, H., Le Letty, R., Claeyssen, F., Barillot, F., Lhermet, N., and
Bouchilloux, P., Semi-Passive And Semi-Active Vibration Control Using New
Amplied Piezoelectric Actuators, SPIEs Smart Structures and Materials,
1999, pp. 814821.
[130] Behrens, S., Fleming, A.J., and Moheimani, S.O.R., New Method For
Multiple-Mode Shunt Damping Of Structural Vibration Using A Single Piezo-
electric Transducer, SPIEs Smart Structures and Materials, Vol. 4331, 2001,
pp. 239250.
[131] Tauzsig, L., Numerical Detection and Characterization of Blade Vortex In-
teractions Using a Free Wake Analysis, M.S. Thesis, Department of Aerospace
Engineering, The Pennsylvania State University, 1998.
291
[132] Tauzsig, L., Gandhi, F, Numerical Detection of Blade Vortex Interactions
Using a Free Wake Analysis - A Comparison of various Approaches, Proceed-
ings of the 24th European Rotorcraft Forum, Marseilles, France, September
1998.
[133] Floros, M.W., Elastically Tailored Composite Rotor Blades for Stall Allevia-
tion and Vibration Reduction, The Pennsylvania State University, Department
of Aerospace Engineering, December 2000.
[134] Hodges,D.H. and Dowell, E.H. Nonlinear Equations of Motion for the Elas-
tic Bending and Torsion of Twisted Nonuniform Blades, NASA TN D-7818,
Dec 1974.
[135] Zhang, J., Active-Passive Hybrid Optimization of Rotor Blades With Trail-
ing Edge Flaps, The Pennsylvania State University, Department of Aerospace
Engineering, May 2001.
[136] Hodges, D.H., Atilgan, A.R., Fulton, M.V., and Reheld, L.W., Free-
Vibration Analysis of Composite Beams, Journal of the American Helicopter
Society, Vol. 36, No. 3, July 1991, pp. 3657.
[137] Gjelsvik, A., The Theory of Thin Walled Bars, John Wiley and Sons, Inc.,
1981.
[138] Song, O., and Librescu, L., Structural Modeling and Free Vibration Anal-
ysis of Rotating Composite Thin-walled Beams, Journal of the American
Helicopter Society, Vol. 39, October 1997, pp. 36-57.
[139] Fung, Y.C., An introduction to the theory of aeroelasticity, Dover publica-
tions, Inc., 1993.
[140] Reddy, J.N., Mechanics of Laminated Composite Plates Theory and Analysis,
CRC Press, Boca Raton, FL, 1996.
[141] Leishman, J.G., Principles of Helicopter Aerodynamics, Cambridge Univer-
sity Press, 2000.
[142] Greenwood, D.T., Principles of Dynamics, Prentice-Hall, Inc., 1988.
[143] Theodorson, T., General Theory of Aerodynamic Instability and the Mecha-
nism of Flutter, NACA Report No. 496, 1935.
[144] Johnson, W., Self-Tuning Regulators for Multicyclic Control of Helicopter
Vibration, NASA TP 1996, March 1982.
292
[145] Landgrebe, A.J., An Analytical Method for Predicting Rotor Wake Geom-
etry, Journal of the American Helicopter Society, Vol. 14, No. 4, 1969, pp.
2032
[146] DOT User Manual, VR&D, 1995.
[147] McCormick, B.W., Aerodynamics of V/STROL Flight, Dover Publications,
Inc., New York, 1998.
[148] Jones, R.T., The Unsteady Lift of a Wing of Finite Aspect Ratio, NACA
Report No. 681, 1940.
[149] Standards Committee of the IEEE Ultrasonics, Ferroelectric, and Frequency
Control Society, An American National Standard: IEEE Standard on Piezo-
electricity, The Institute of Electrical and Electronics Engineers, ANSI/IEEE
Std 176, New York, 1987.
[150] Kim, J., Ryu, Y.-H. and Choi, S.-B., New Shunting Parameter Tuning
Method for Piezoelectric Damping Based on Measured Electrical Impedance,
Smart Structures and Materials, Vol. 9, 2000, pp. 868877.
[151] Rosenbaum, J.F., Bulk Acoustic Wave Theory and Devices, Artech House,
Inc., 1988.
[152] Den Hartog, J.P., Mechanical Vibrations, McGraw-Hill Inc., 1956.
[153] Jordan, D.W., and Smith, P., Nonlinear Ordinary Dierential Equations, 3rd
edition, Oxford University Press, 1999.
[154] Cruz-Hernandez, J.M., and Hayward, V., Phase Control Approach to Hys-
teresis Reduction, IEEE Transaction on Control Systems Technology, Vol. 9,
No. 1, Jan. 2001, pp. 1726.
[155] Niezrecki, C. and Cudney, H.H., Improving the Power Consumption Char-
acteristics of Piezoelectric Actuators, Journal of Intelligent Materials and
Systems, Vol. 5, 1994, pp. 522529.
[156] Sirohi, J. and Chopra, I., Actuator Power Reduction using L-C Oscillator
Circuits, Proceedings of 41st AIAA/ASME/ASCE/AHS/ASC/Structures,
Structural Dynamics and Materials Conference, AIAA-2000-1971, Atlanta,
GA, 2000.
Vita
Jun-Sik Kim
Education
The Pennsylvania State University University Park, PA 20002005
Ph.D. in Aerospace Engineering
Area of Specialization: Rotorcraft Dynamics and Smart Structures
Inha University Inchon, Korea 19941996
M.S. in Aerospace Engineering
Area of Specialization: Solid Mechanics and Finite Element Method
Inha University Inchon, Korea 19901994
B.S. in Aerospace Engineering
Awards and Honors
PSU Rotorcraft Center Fellowship 20002001
Research Experience
Research Assistant The Pennsylvania State University 20012005
Research Project sponsored by NRTC
Research Assistant Inha University 1996, 1999
Teaching Experience
Part-time Instructor Inha Technical College, Inchon, Korea 19992000
The strength of materials in undergraduate level.
Selected Publications
J.-S. Kim, K. W. Wang, and E. C. Smith, Design and Analysis of Piezoelec-
tric Transducer Based Resonant Actuation Systems, Proceedings of the ASME
IMECE, Orlando, FL, Nov 511, 2005.
J.-S. Kim, K. W. Wang, and E. C. Smith, High Authority Piezoelectric
System Synthesis through Mechanical Resonance and Electrical Tailoring, Journal
of Intelligent Material Systems and Structures, Vol. 16, No. 1, 2005, pp. 2132.
J.-S. Kim, E. C. Smith, and K. W. Wang, Helicopter Vibration Suppression
via Multiple Trailing Edge Flaps Controlled by Resonance Actuation System,
Proceedings of the 60th AHS Forum, Bltimore, Maryland, June 810, 2004.
J.-S. Kim, Reconstruction of First-order Shear Deformation Theory for Lam-
inated and Sandwich Shells, AIAA Journal, Vol. 41, No. 5, 2004, pp. 16851697.

You might also like