Uehara Dissertation 2019
Uehara Dissertation 2019
by
Daiju Uehara
2019
The Dissertation Committee for Daiju Uehara
certifies that this is the approved version of the following dissertation:
Committee:
David B. Goldstein
Krishnaswa Ravi-Chandar
Jeffrey K. Bennighof
Mahendra J. Bhagwat
Estimation of Helicopter Rotor Loads from Blade
Structural Response
by
Daiju Uehara
DISSERTATION
Presented to the Faculty of the Graduate School of
The University of Texas at Austin
in Partial Fulfillment
of the Requirements
for the Degree of
DOCTOR OF PHILOSOPHY
v
My sincere thanks also goes to Dr. Christopher Cameron and Dr.
Anand Carpatne, who were my senior PhD students in the research group.
Chris and Anand provided significant support not only for academic, research-
related topics, but also for my private life. I do not know how many times
I asked questions to them, how many times I had drink with them, and how
many times we laughed together. I am so lucky that I had Chris and Anand
as my seniors.
vi
sisters have always supported me with love. At every turning point of my life,
including when I decided to leave my home town Okinawa and go to Tokyo
after high school, and when I decided to leave my home country Japan and go
to the United States after undergraduate, my parents just encouraged me to
do so. They did not show me how they felt in their heart at all, just gave the
push I needed. I just want to say, thank you.
vii
Estimation of Helicopter Rotor Loads from Blade
Structural Response
Publication No.
Measuring the load distribution along a helicopter rotor blade has been
one of the most challenging tasks in experimental aeromechanics. Conven-
tional loads measurements with on-blade instrumentation, such as pressure
transducers for airloads and strain gages for structural loads, require the ex-
perimentalist to overcome a large number of technical barriers; for example,
sensor integration to the rotor blade structure, sensor failure due to strong
centrifugal forces, and influence of sensor installation on rotor blade dynam-
ics. The goal of this dissertation is to develop a new, combined experimental
and theoretical methodology to estimate helicopter rotor loads without using
these conventional on-blade sensors.
viii
successfully showed its capability of measuring the three-dimensional deforma-
tion time history of a rotating blade for both a small- and a large-scale rotor
in hover. The modal properties (natural frequencies, mode shapes, damping
ratios, and modal coordinates) of the blade in the rotating-frame were then ex-
tracted from the deformation time history using Natural Excitation Technique
- Eigensystem Realization Algorithm (NExT-ERA) and Complexity Pursuit
(CP), which are operational modal analysis (OMA) algorithms. The first three
modes were identified by the OMA algorithms and well correlated with a nu-
merical model. Rotor loads were then finally estimated based on the measured
deformations and blade modal characteristics.
ix
Table of Contents
Acknowledgments v
Abstract viii
Chapter 1. Introduction 1
1.1 Problem Statement . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 State of the Art . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2.1 Optical deformation measurement . . . . . . . . . . . . 6
1.2.1.1 Grid method . . . . . . . . . . . . . . . . . . . 6
1.2.1.2 Point Tracking . . . . . . . . . . . . . . . . . . 7
1.2.1.3 Digital Image Correlation . . . . . . . . . . . . 9
1.2.2 Operational Modal Analysis . . . . . . . . . . . . . . . . 11
1.2.3 Rotor loads estimation from structural response . . . . . 15
1.3 Present approach . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.4 Contribution of the Current Research . . . . . . . . . . . . . . 19
1.5 Organization of the Dissertation . . . . . . . . . . . . . . . . . 20
x
2.3 Rotor loads estimation . . . . . . . . . . . . . . . . . . . . . . 38
2.3.1 Inertial loads . . . . . . . . . . . . . . . . . . . . . . . . 39
2.3.2 Aerodynamic loads . . . . . . . . . . . . . . . . . . . . . 42
2.3.3 Numerical experiment . . . . . . . . . . . . . . . . . . . 45
xi
Chapter 4. Results and Discussion 97
4.1 Rotor blade deformation measurement . . . . . . . . . . . . . . 97
4.1.1 Small-scale rotor . . . . . . . . . . . . . . . . . . . . . . 99
4.1.2 Large-scale rotor . . . . . . . . . . . . . . . . . . . . . . 104
4.1.2.1 Single-bladed CCR rotor (Test campaign 2) . . 105
4.1.2.2 Two-bladed single rotor (Test campaign 3) . . . 111
4.2 Rotor Blade modal characteristics identification . . . . . . . . 121
4.2.1 Small-scale rotor blade characteristics . . . . . . . . . . 122
4.2.1.1 Modal frequency . . . . . . . . . . . . . . . . . 123
4.2.1.2 Mode shapes . . . . . . . . . . . . . . . . . . . 127
4.2.1.3 Modal damping . . . . . . . . . . . . . . . . . . 136
4.2.2 Large-scale rotor blade characteristics . . . . . . . . . . 138
4.2.2.1 Random excitation (Test campaign 2) . . . . . 139
4.2.2.2 Step function excitation test (Test campaign 3) 142
4.3 Rotor loads measurement and estimation . . . . . . . . . . . . 149
4.3.1 CCR rotor loads measurement . . . . . . . . . . . . . . 150
4.3.2 Two-bladed single rotor loads measurement and estimation154
4.3.2.1 Test condition 1 . . . . . . . . . . . . . . . . . . 154
4.3.2.2 Test condition 2 . . . . . . . . . . . . . . . . . . 163
4.3.2.3 Test condition 3 . . . . . . . . . . . . . . . . . . 172
Bibliography 193
Vita 206
xii
List of Tables
xiii
List of Figures
xiv
3.9 A schematic of the AMR sensor installation on the blade grip
assembly . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
3.10 The AMR sensor characteristics . . . . . . . . . . . . . . . . . 67
3.11 A drawing of the hydraulic servo actuator assembly . . . . . . 70
3.12 Rotor pitch angle control system assembly with the hydraulic
servo actuators . . . . . . . . . . . . . . . . . . . . . . . . . . 71
3.13 A diagram of pitch angle control system for the hydraulic actu-
ation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
3.14 The setup for small-scale rotor blade deformation measurement. 75
3.15 A schematic of the small-scale DIC measurement setup . . . . 76
3.16 Unprocessed images of the rotor strobed at 16 azimuthal loca-
tions, showing complete rotor disk at 1200 RPM. . . . . . . . 79
3.17 Camera calibration target . . . . . . . . . . . . . . . . . . . . 80
3.18 Vector map processed by DIC for the 30◦ root pitch rotor, spun
at 1500RPM . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
3.19 A schematic of camera arrangement for the large-scale DIC test 83
3.20 Stereoscopic camera system for the single-bladed CCR rotor . 84
3.21 Rotor blade painted with orange fluorescent color and black dots 84
3.22 A schematic of the DIC measurement setup . . . . . . . . . . 85
3.23 Unprocessed images taken simultaneously by the high-speed
cameras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
3.24 Camera calibration setup . . . . . . . . . . . . . . . . . . . . . 86
3.25 A schematic of stereoscopic camera arrangement used during
Test campaign 3 . . . . . . . . . . . . . . . . . . . . . . . . . . 93
3.26 Pitch angle variations and the initial digital trigger timing ob-
tained from the three independent measurements . . . . . . . 95
xv
4.7 Flap bending deformation of the CCR lower rotor blade at CT /σ
= 0.09 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
4.8 Lower rotor blade tip displacement comparison between mea-
surement and prediction . . . . . . . . . . . . . . . . . . . . . 109
4.9 Maximum and minimum flap bending deformation of the lower
rotor blade as a function of blade radial station at CT /σ = 0.09 110
4.10 Interaction of blade bound circulation in the CCR rotor system 111
4.11 Recap: Two-bladed, isolated single rotor configuration . . . . 112
4.12 Out-of-plane blade deformation over the entire rotor disk . . . 113
4.13 Blade tip displacement as a function of azimuthal angle . . . . 114
4.14 3D static deformation distributed over the entire blade span . 116
4.15 Static blade deformations of each degree of freedom, along the
blade quarter-chord axis . . . . . . . . . . . . . . . . . . . . . 117
4.16 Out-of-plane blade deformation over the entire rotor disk for
the lateral pitch angle input . . . . . . . . . . . . . . . . . . . 118
4.17 Blade tip displacement as a function of azimuthal angle for the
lateral pitch angle input . . . . . . . . . . . . . . . . . . . . . 119
4.18 Blade tip deformations of each degree of freedom as a function
of time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
4.19 Rotor blades with fluorescent paint and random speckle pattern 122
4.20 Pitch angle displacement at the tip for θ0 = 30◦ , Ω = 1200 RPM125
4.21 Log-scale frequency spectrum of the out-of-plane displacement
measured at the blade tip for θ0 = 30◦ , Ω = 1200 RPM . . . . 126
4.22 Fan plot for θ0 = 0◦ , comparing natural frequencies extracted
with the CP and NExT-ERA algorithms . . . . . . . . . . . . 128
4.23 Fan plot for θ0 = 30◦ , comparing natural frequencies extracted
with the CP and NExT-ERA algorithms, along with numerical
predictions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
4.24 The first mode shape of the rotor blade for θ0 = 30◦ at 1200 RPM130
4.25 The second mode shape of the rotor blade for θ0 = 30◦ at 1200
RPM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
4.26 The first and second mode shapes along the quarter-chord axis,
for θ0 = 0◦ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
4.27 The first and second mode shapes along the quarter-chord axis,
for θ0 = 30◦ . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
xvi
4.28 Comparison of the first mode shapes estimated by the CP and
NExT-ERA as well as numerical prediction for θ0 = 30◦ at 1200
RPM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
4.29 Comparison of the second mode shapes estimated by the CP
and NExT-ERA as well as numerical prediction for θ0 = 30◦ at
1200 RPM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
4.30 Thrust measured on the single-bladed single rotor at zero mean
thrust . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
4.31 Measured variation of tip displacement on the single-bladed sin-
gle rotor at zero mean thrust . . . . . . . . . . . . . . . . . . . 141
4.32 Fan plot with measured natural frequencies (100% = 900 RPM),
compared with numerical predictions . . . . . . . . . . . . . . 142
4.33 Measured rotating mode shapes of the first flap mode at nominal
rotational speed (900 RPM) compared with numerical predictions143
4.34 Measured rotating mode shapes of the first lag mode at nominal
rotational speed (900 RPM) compared with numerical predictions143
4.35 Measured rotating mode shapes of the second flap mode at nom-
inal rotational speed (900 RPM) compared with numerical pre-
dictions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
4.36 Time history of the out-of-plane deformation measured at the
trailing-edge point of the rotor blade tip . . . . . . . . . . . . 145
4.37 Frequency spectra of the three different time periods . . . . . 146
4.38 Spectrogram of the flap response measured at the trailing edge
point of the blade tip . . . . . . . . . . . . . . . . . . . . . . . 148
4.39 Reconstructed (filtered) thrust of the CCR lower rotor, com-
pared to predictions at CT /σ = 0.09 . . . . . . . . . . . . . . 151
4.40 Reconstructed (filtered) pitch link load of the CCR lower rotor,
compared to predictions at CT /σ = 0.09 . . . . . . . . . . . . 153
4.41 Phase-averaged thrust variation as a function of blade azimuthal
location for the steady collective pitch input . . . . . . . . . . 156
4.42 Phase-averaged torque variation as a function of blade azimuthal
location for the steady collective pitch input . . . . . . . . . . 157
4.43 Phase-averaged pitching moment variation as a function of blade
azimuthal location for the steady collective pitch input . . . . 158
4.44 Phase-averaged pitch link force variation as a function of blade
azimuthal location for the steady collective pitch input . . . . 160
4.45 Phase-averaged pitch angle variation as a function of blade az-
imuthal location for the steady collective pitch input . . . . . 162
xvii
4.46 Estimated rotor loads for the steady collective pitch input . . 164
4.47 Phase-averaged thrust variation as a function of blade azimuthal
location for the lateral cyclic pitch input . . . . . . . . . . . . 166
4.48 Phase-averaged torque variation as a function of blade azimuthal
location for the lateral cyclic pitch input . . . . . . . . . . . . 167
4.49 Phase-averaged pitching moment variation as a function of blade
azimuthal location for the lateral cyclic pitch input . . . . . . 168
4.50 Phase-averaged pitch link force variation as a function of blade
azimuthal location for the lateral cyclic pitch input . . . . . . 170
4.51 Phase-averaged pitch angle variation as a function of blade az-
imuthal location for the lateral cyclic pitch input . . . . . . . 171
4.52 Estimated rotor loads for the lateral cyclic pitch input (Test
condition 2) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
4.53 Blade root pitch angle as a function of rotor revolutions during
the step pitch input operation . . . . . . . . . . . . . . . . . . 175
4.54 Rotational speed of the rotor as a function of rotor revolutions 176
4.55 Rotor thrust response to the step pitch change as a function of
rotor revolutions . . . . . . . . . . . . . . . . . . . . . . . . . 177
4.56 Rotor torque response to the step pitch change as a function of
rotor revolutions . . . . . . . . . . . . . . . . . . . . . . . . . 178
4.57 Pitch link force response to the step pitch change as a function
of rotor revolutions . . . . . . . . . . . . . . . . . . . . . . . . 179
4.58 The estimated inertial load integrated at the hub as a function
of the number of revolutions for the collective step pitch change
(Test condition 3) . . . . . . . . . . . . . . . . . . . . . . . . . 180
4.59 Comparison between the estimated and measured rotor hub
loads for the collective step pitch change (Test condition 3) . . 182
xviii
Chapter 1
Introduction
1
dynamics of the rotor blade can also be significantly altered by installation
and modifications to accommodate the pressure transducers, static pressure
taps, and strain gages.
Figure 1.1: Comparison between images taken during static display and flight
operation of the CH-53 Sea Stallion [1, 2]
2
tionship between the structural and aerodynamic loads on a helicopter rotor.
Bousman [6] provided a good summary of previous airload estimation method-
ologies, as well as his method that estimates blade aerodynamic forces based
on measured flap bending moments and modal parameters of a rotor blade.
This work showed good agreement between the estimated and directly mea-
sured spanwise aerodynamic loadings, except for regions near the blade tip.
Wang et al. [7] discussed a combined analytical-experimental approach to pre-
dict the vibratory loads in an articulated rotor. In this study, the estimated
airloads were obtained based on strain measurements as well as the lifting-line
aerodynamic and finite element blade structural models, with a satisfactory
agreement to the corresponding flight test data. These previous studies have
shown that aerodynamic estimations appeared to be theoretically robust and
feasible for prediction of rotorcraft dynamic behaviors; however, the method-
3
ologies used in these studies rely on precise identification of modal parameters
of a rotating rotor blade (i.e., mode shapes, modal amplitudes, and natural fre-
quencies), as well as data measured by pointwise on-blade sensors (i.e., strain
gages). The former would necessitate additional experiments (modal tests) or
to develop an accurate finite element model of the blade structure, while the
latter would require extensive preparation for measurements with challenges
inherent to operating in a high centrifugal force environment.
The other concern for a successful rotor load estimation, on-blade sen-
sors failure, can be resolved by introducing non-contact optical measurement
techniques that requires no troublesome preparation and installation of sen-
4
sors on a target specimen. To date, the optical methodologies have been ap-
plied to a variety of different structural engineering fields, including helicopter
rotor blade deformation and strain measurements in the rotating frame; for
example, photogrammetry, holographic interferometry, and projection moiré
interferometry (PMI). Among quite a few of these methodologies, this disser-
tation selects the Digital Image Correlation (DIC) deformation measurement
technique, since the DIC technique has shown its robustness in measuring
rotor blade deformations at different rotor configurations and scales, as dis-
cussed in [8, 9, 10, 11]. The DIC deformation measurement technique is able to
measure three-dimensional displacements over the entire blade span at a high
spatial resolution with minimal preparation of the test article (i.e., painting
the rotor blade surface with stochastic speckle patterns).
5
1.2 State of the Art
This section discusses the literature related to the three core elements of
the present study: the optical deformation measurement techniques, the oper-
ational modal analysis (OMA) algorithms, and rotor loads estimation method-
ologies.
• Grid method
• Point Tracking
The grid method is one of the earliest optical methods for measurement
of structural deformations [13]. The basic procedure of the grid method is as
follows; (1) regular marking patterns are applied to the target surface of a
6
test specimen, (2) diffusive light reflection from those patterns are recorded,
(3) the differences between two grid images are used to calculate the defor-
mations of the structure. The projected grid method was used for the Higher
Harmonic Control Aeroacoustic Rotor Tests (HART) program, performed by
Kube et al. [14]. The method provided flapwise and torsional deflections along
the blade span at selected azimuthal angles, and its results were compared
to signals measured by on-blade strain gages. Electronic Speckle Pattern In-
terferometry (ESPI) and Projection Moiré Interferometry (PMI) are other
non-contact optical approaches. The former electronically observes formation
of fringes and records images of the reference (undeformed) and deformed ob-
jects [15, 16], while the latter evaluates fringe interference patterns between
the deformed projected grid and a computationally phase-shifted grid to calcu-
late out-of-plane displacements [17, 18]. Recently, Sekula [19] employed PMI
to measure blade deformations of a model-scale rotor in hover in the NASA
Langley subsonic wind tunnel and hover test facility, and successfully acquired
the out-of-plane hub height, tip height, and the flap angles of the 66.5 in (1.69
m)-radius target blade. These results showed that measuring blade deflection
using digital images is feasible.
7
are shown as follows:
8
1.2.1.3 Digital Image Correlation
9
microns [32]. Although ESPI also has the capability of measuring small struc-
tural motion on the same order of sensitivity, it is not well-suited for measuring
large deflections of a structure unlike DIC approach. Another advantage of
DIC is its ready applicability to dynamic deformation measurements. These
instantaneous measurements can be analyzed by a dynamic sequence of digital
images taken by high-speed image recording equipment. Schmidt et al. [33]
used a stereoscopic set of high-speed cameras with a pulsed YAG laser as a
light source to measure rapid deformations of a road wheel tire, achieving a
short exposure time of 7 nanoseconds.
Over the past few years, special attention has been paid to the DIC
technique not only in the field of experimental structural and solid mechanics,
but in a large number of different research and engineering fields, such as tensile
testing of a knee tendon, deformation of a frog heart, tissue studies, and testing
of biomimetic materials in microscopic biomechanics [34]. The fundamental 2D
DIC framework itself has been extensively investigated and improved towards
reducing computational cost, widening application range, and achieving high
accuracy and precision. One critical expansion of the application area was to
be capable of measuring out-of-plane, three-dimensional deformation over the
curved surface of a structure.
10
eral azimuthal locations using a phase-averaging technique, that is, the images
are captured at a specific azimuthal location and averaged over multiple rotor
revolutions. The process is repeated at a different azimuthal location, yielding
an average deformation at each azimuthal location [10, 11]. However, due to
the averaging process, these deformations are not correlated in time. Sicard
and Sirohi [9] utilized the 3D DIC technique to measure the deformation of an
extremely flexible rotor blade at a specific azimuthal location with validation
performed using laser displacement sensor measurements and an aeroelastic
numerical simulation tool. The DIC measurement technique was also used
by Cameron et al. [35] to measure the rotating modal properties of a flexible
blade in a reduced-scale coaxial helicopter rotor over a field of view limited to
one quarter of the rotor disk.
11
classical experimental modal analysis (EMA) that requires measurements of
both the input and output information. Since the beginning of its development
in civil engineering, OMA methodologies have been proposed in a wide variety
of engineering applications, particularly in situations where measuring input
forcing or exciting the structure is impractical, such as in large-scale bridges,
buildings, aircraft, and wind turbines.
12
used the ITD method to extract the modal properties of a cantilever beam and
a model-scale, conventional rotor blade under rotation. Cameron et al. [35]
identified the rotating natural frequencies and mode shapes of a rotor blade
on a reduced-scale coaxial, counter-rotating rotor system using the modified
version of ITD algorithm. Two common features in these two previous studies
are: (i) the excitation to the spinning structure was a periodic gust from a
compressed air source, and (ii) the deformation measurements were not con-
tinuous in time. Nevertheless, the ITD algorithm was able to identify the first
three flap-bending natural frequencies and mode shapes.
13
vertical and horizontal axis wind turbines, and there have been a number of
follow-up studies to this pioneering work in wind turbine research and devel-
opment [42, 43, 44].
14
optical deformation measurement techniques, although OMA in conjunction
with the non-contact optical measurement methods seems to have great po-
tential for reducing traditional difficulties of identifying rotor blade structural
dynamic characteristics. Thus, the current study in this dissertation applies
all the four methodologies to rotor blade structural responses measured using
the DIC technique, and evaluates the compatibility and applicability of each
OMA algorithm to the optically-obtained structural deformation.
15
mal modes of the rotor blade, with measured flap bending strain and blade
root motion. Bousman [6] also proposed the rotor flapwise loads estimation
technique based on a modal expansion of the flapping beam equation of mo-
tion with the aid of flap bending moment measurement. These previous studies
gave the key suggestion that in applications of their methodologies, at least
the first three modes in addition to the rigid body mode should be employed
for accurate estimation of spanwise airloads distribution.
16
YAMAHA R-50/RMAX UAV testbeds. Although the proposed methods and
processes in these studies were verified and have been shown to be useful in
practice, accurate computational models of the rotor, such as a finite element
model for the structure, aerodynamic model, and inflow distribution model,
were required to obtain the necessary information for loads estimation.
17
Pursuit (CP), in order to identify modal frequencies and mode shapes in the
rotating frame. (iii) The rotor loads including the structural and aerodynamic
forces are identified based on the measured blade deformations and the modal
parameters obtained in the previous two steps. This combined DIC-OMA ap-
proach would yield a simple and inexpensive strategy to estimate rotor loads
without using on-blade sensors and detailed computational models of the ro-
tor. It is especially valuable because the challenging and costly problem of
performing accurate surface pressure measurements could be avoided.
18
once-per-revolution frequency (cosine cyclic pitch input), and (3) step change
in rotor blade pitch angle (step function collective pitch input). Hub load (ro-
tor thrust) is then reconstructed from numerical integration of the estimated
spanwise lift distribution along the blade quarter-chord axis, and is compared
to the time history of rotor thrust directly measured by a load cell installed
on the rotor hub.
19
1.5 Organization of the Dissertation
20
Chapter 4 discusses the results of the DIC blade deformation measure-
ment, modal parameter identification, and rotor loads estimation of each ex-
perimental system. Note that the setups (i) and (ii) are used for verifying the
applicability and robustness of the combined DIC-OMA approach, while the
setup (iii) is used for verifying the entire process of the current methodology,
including the DIC, the OMA, and rotor loads estimation. These individual
measurements on the three different rotor configurations and operating con-
ditions examine whether the methodology proposed in this dissertation is a
general tool for helicopter rotor loads estimation.
21
Chapter 2
Theoretical Development
22
2.1.1 Blade coordinate system
The deformed position of the rotor blade is described using the blade-
fixed coordinate system, shown in Fig. 2.1. The origin of the coordinate system
is fixed at the center of rotation (center of the rotor hub), and the coordinate
system rotates about the positive z-axis at a constant rotational speed of Ω.
The x-axis is along the quarter-chord axis of the undeformed rotor blade.
Additionally, three degrees of freedom are used in the study: out-of-plane
(flapwise) motion, in-plane (lead-lag) motion, and torsional (feathering) mo-
tion with respect to the hub plane under the action of changing aerodynamic
lift and drag forces. Blade deformation measured using the DIC technique
shown in chapter 4 follows this definition of the blade-fixed coordinate system.
𝑧
𝑦
Ω
𝑥 Quarter-chord axis Torsion
Flap
Rotor blade
Lead-lag
23
2.1.2 2D DIC
1. Random speckle patterns are applied onto a surface of the target struc-
ture.
4. The full image is divided into a set of small interrogation windows (also
called subsets) with an area of multiple pixels such as 16×16 pixels.
7. Once the location of the target interrogation window is found in the de-
formed image, a 2D displacement vector corresponding to the motion of
the window is computed, and the search-and-compute process is repeated
for all the interrogation windows on each image.
24
To search for the same subset match in the reference and deformed im-
age (Step. 6 and 7 in the procedure), the normalized cross-correlation function
C(x0 , y0 , x1 , y1 ) is computed as follows:
P
F (x0 , y0 )G(x1 , y1 )
C(x0 , y0 , x1 , y1 ) = pP (2.1)
F (x0 , y0 )2 G(x1 , y1 )2
where F (x0 , y0 ) and G(x1 , y1 ) represent the light intensity (values of gray levels
of the pixels) inside the area of the subset from the undeformed and deformed
images, respectively. (x0 , y0 ) and (x1 , y1 ) are the coordinates of a central point
on the target subset before and after deformation, respectively. Figure 2.2
shows a schematic of the subset search process.
𝑥% , 𝑦%
𝑥# , 𝑦#
Displacement
vector
𝑦
25
Essentially, displacement computations are treated as an optimization
problem, that is, the displacement vector is calculated by searching a set of
the coordinates (x1 , y1 ) after deformation, that maximizes the cross-correlation
function C(x0 , y0 , x1 , y1 ). That is the reason why this technique is called digital
image correlation. Details of the mathematical development of 2D DIC are
available in Ref. [58].
2.1.3 3D DIC
3D DIC generally uses multiple digital cameras and arranges their rel-
ative positions in a stereoscopic manner, taking images of a target structure
from different perspectives. A commercially available software LaVision DaVis
8.4.0 [59] is used for camera control and calibration, image acquisition, and
DIC processing.
26
Image plane 1 Image plane 2
Camera 1 Camera 2
Surface height 𝑧
𝑦
𝐻 𝑥, 𝑦
Reference 𝑥
surface
27
Image plane 1 at 𝑡 Image plane 1 at 𝑡 + ∆𝑡
2D vector 1
𝑀% 𝑥, 𝑦, 𝑧
3D vector
Image plane 2 at 𝑡 Image plane 2 at 𝑡 + ∆𝑡
𝑀* 𝑥, 𝑦, 𝑧
2D vector 2
2.2.1 NExT-ERA
28
structure
Mẍ(t) + Cẋ(t) + Kx(t) = f (t) (2.2)
where M, C, and K are the mass, damping, and stiffness matrices respectively,
f (t) is the force vector, and ẍ(t), ẋ(t), and x(t) are the acceleration, velocity,
and displacement vectors at time t. Assuming that the system is excited by a
white noise, Eq. 2.2 can be written as:
where X(t) is the random response vector and F(t) is the random excitation
vector. Then multiplying Eq. 2.3 by a reference scalar random response vector
Xi (s) at a location i and time s on the system yields
29
Rx(m) xj (τ ) = Rx(m)
i xj
(τ ) (2.8)
i
where xi (t) and xj (t) are random response vectors at two different locations
(m)
i and j, and xi is the mth derivative of xi (t) with respect to time, and
(m)
Rxi xj (τ ) denotes the mth derivative of the cross-correlation function Rxi xj (τ )
with respect to τ .
With the assumption that the responses of the system are uncorrelated
to the random excitation F(t), Eq. 2.6 can now be expressed as
In Eq. 2.9, it is obvious that the cross-correlation function RXXi (τ ) satisfies the
homogenous differential equation of motion. As described by James et al. [41],
the cross-correlation function is equivalent to a sum of decaying sinusoids of
the same form as the impulse response of the original structure in time domain.
This cross-correlation function obtained with the NExT analysis is used as an
input to ERA process.
30
of the Fourier Transform of each data block in order to improve the accu-
racy of the computation. Then the cross-correlation function between the two
channels i and j is obtained using
N −1
1 X 2πkn
Rxi xj (n) = Sx x (k)exp i , n = 0, . . . , N − 1 (2.10)
N k=0 i j N
It should be noted that the assumption made for the theoretical devel-
opment, i.e., the white noise excitation is uncorrelated, is not necessarily met
for modal analysis on a real structure. Nevertheless, the operational modal
analysis with NExT has been performed successfully on real structures in past
studies [41, 62, 63] without satisfying the assumption, i.e. the excitation to
the structure is not a white noise; it is typically bandwidth limited and does
not have equal power at all frequencies. For appropriate use of this method,
it is essential for an analyst to determine the number of samples used for the
algorithm, the number of data blocks, the number of overlapping data points,
the length of the data blocks, shape of window, and reference channels for
calculating cross-correlation functions. These are all typically determined by
trial-and-error.
31
where x(t) is the state vector, u(t) is the excitation vector, y(t) is the response
vector to the excitation, and Ac , B, and C are state-space matrices in con-
tinuous time. Assuming that an impulse excitation is applied to the dynamic
system, a solution to these equations has the form
where Y(t) consists of the free responses of the system to an impulse excitation.
In discrete time, the solution is represented as
where Y(n) is the response at the nth time step, A = eAc ∆t is the discrete-
time state transition matrix and ∆t is the time step of system discretization.
The goal of ERA is to estimate the constant matrices A, B, and C from
the impulse response data Y(n) acquired from measurement. The system
modal parameters are then computed from the eigenvalues and eigenvectors
of the constructed state matrices. Note that the y(t) vector is the output of
the NExT algorithm that yields pseudo-impulse responses from the measured
structural deformation. It is important to mention that the matrices B can-
not be estimated for the present modal identification since the NExT is an
output-only modal analysis technique and the real excitation to the structure
is unknown. However, the modal parameters are obtained from the matrices
A and C.
32
System realization starts from building the generalized Hankel matrix:
Y(n) Y(n+1) ... Y(n+p)
.. ..
Y(n+1)
H(n − 1) = . .
(2.14)
..
.
Y(n+q) ... Y(n+p+q)
where Y(n) is in this case the N × 1 vector which consists of the pseudo-
impulse responses computed by NExT at the nth time step, and N is the
number of channels corresponding to the number of interrogation windows
during DIC processing. The parameters p and q correspond to the size of the
Hankel matrix. As a rule of thumb for selecting p and q, the analyst should
consider ten times the number of expected modes for the number of columns
p, and twice to three times the number of columns p for the number of rows q
in the Hankel matrix. Normally, the Hankel matrix is formed with block rows
and block columns data shifted in time by one sample from the previous block
row and column.
33
transformed to modal coordinates by multiplying with the eigenvector matrix
of A
Λ = Ψ−1 AΨ
(2.17)
Cm = CΨ
where Λ is the diagonal matrix of eigenvalues of A and Ψ is the matrix of
eigenvectors of A. After transforming back to the continuous time domain,
one can obtain the modal damping ratios and damped natural frequencies of
the system from the real and imaginary parts of the eigenvalues, respectively.
The column vectors of Cm correspond to the mode shapes.
34
where F(t) is the force vector and ẍ(t), ẋ(t), and x(t) are the acceleration,
velocity, and displacement vectors at time t. The modal transformation can
be expressed as
x(t) = Φη(t) (2.19)
where Φ and η are the modal matrix and modal coordinate vector of the
system, respectively. Note that the modes of a system form, by definition,
a linearly independent set of basis vectors [64]. Let us introduce the general
concept of BSS in mathematical form as
where y(t) is the measured displacement vector at time t, n(t) is a noise vector,
s(t) is the source vector, and A, often called “mixing” matrix, is a transfer
matrix between sensors and sources. As shown in Eq. 2.20, the usage of BSS
within the context of operational modal analysis is to estimate the mixing
matrix A and the source vector s from the measured output responses y(t),
and interpret them as the modal matrix Φ and modal coordinates η of the
structural system.
35
number of measurement points, meaning that BSS separates the measured
data into N different source signals. A number of techniques falling under the
broad spectrum of BSS can be found in the literature. These techniques share
the fact that a priori information about the sources or the input forcing is not
required.
36
expressed as weighted sums of signal values measured up to time tj−1 given by
Values of λL = 0.99 and λS = 0.5 are used in this study. Minimization of the
function G(yi ) leads to an eigenvalue problem that yields the columns Ai of
the mixing matrix and corresponding eigenvalues Λi as follows:
where Ĉ and C̄ are the short-term and long-term covariance matrices respec-
tively, calculated by:
N
X
Ĉ = (y(tj ) − ŷ(tj ))(y(tj ) − ŷ(tj ))T
j=1
N
(2.25)
X
C̄ = (y(tj ) − ȳ(tj ))(y(tj ) − ȳ(tj ))T
j=1
37
Thus, this present study selected the CP algorithm as a primary scheme to
extract rotating-frame modal properties of a rotor blade.
• Resampling
• Windowing
• Averaging NExT CP
• Ref. channel
selection
• Size of Hankel
matrix Spectrum
• Peak-picking
• # of poles ERA analysis
• # of reference
channels
• CMI
• Smoothing User User
mode shapes evaluation evaluation
Modal Modal
parameters parameters
38
out-of-plane rotor loads.
The vertical hub reaction force, which is essentially the rotor thrust, is
the net shear at the root of the rotating blade and is theoretically obtained by
integrating the sectional inertial and aerodynamic forces on the blade. Con-
sidering a rigid blade with no hinge offset as shown in Fig. 2.6, the vertical
shear Sz at the blade root is expressed as
Z R Z R
Sz = Fz dr − mz̈ dr (2.26)
0 0
Aerodynamic force
$%
Ω
Centrifugal
force
Hub reaction
force &% Inertial force
Blade #
!#̈
(
39
of Eq. 2.26. The transverse acceleration z̈ can be obtained by numerically
differentiating measured out-of-plane displacements z in space, since the DIC
technique measures three-dimensional, full-field, time-resolved deformation of
the blade structure. With the mass properties of rotor blades calculated from
a CAD model, one can compute the distributed inertial forces on the blade as
a function of time (azimuthal angle) and space (radial location).
where y(x) is a continuous function that exactly describes the given data
trend. The first term of the equation above is to quantify the goodness of fit
of a ”smooth” function ŷ(x) to the trend y(x), whereas the second term of
40
the equation is to determine the ”roughness” of ŷ(x). The second derivative
of ŷ(x) corresponds to the curvature of the function. Since large curvature
of a function means that the function changes rapidly, constraining the term
to be small makes the desired curve smooth. The goal of the regularization
is to find a function ŷ(x) such that the objective function Q(ŷ) is minimized
with a weighting factor λ. Further details of the mathematical development
as well as several methods for selecting an appropriate value of λ can be found
in Ref. [66].
41
140 140
Original Original
120 Regularized, =5 120 Regularized, = 50
Z displacement, [mm]
Z displacement, [mm]
100 100
80 80
60 60
40 40
20 20
0 0
-20 -20
0 10 20 30 40 50 60 70 80 90 100 110 0 10 20 30 40 50 60 70 80 90 100 110
Radial location, [%R] Radial location, [%R]
(a) λ = 5 (b) λ = 50
on this result, a 3rd-order polynomial curve fit was selected to smoothen the
deformation measured by DIC in this dissertation. Note that the inertial
properties of a rotor blade used in this study are known from an accurate
CAD model, and its details are described in chapter 3.
42
140
Original
120 3rd order polynomial
Z displacement, [mm]
100
80
60
40
20
0
-20
0 10 20 30 40 50 60 70 80 90 100 110
Radial location, [%R]
where E is the modulus of elasticity of the blade section, I is the area moment
about the chordwise principal axis, and Ω is the rotational speed. Now consider
the free vibration of the rotating rotor blade at frequency ν. The solution for
the homogeneous partial differential equation can be written as
z = η(r)eiνt (2.29)
Substituting Eq. 2.29 to 2.28 and making the right hand side of Eq. 2.28 zero,
the result is
Z R
d2 d2 η
d 2 dη
EI 2 − mΩ ρdρ − ν 2 mη = 0 (2.30)
dr2 dr dr r dr
which can be viewed as the equation for vibration in a vacuum. This modal
equation yields an eigenvalue problem for the natural frequency ν and mode
shape η(r) and there exists a series of eigensolutions ηk (r) and corresponding
43
eigenvalues νk of the k-th mode. The out-of-plane deflection z(r, t) is thus
expanded as a series of mode shapes describing the spanwise deformation as
∞
X
z(r, t) = ηk (r)qk (t) (2.31)
k=1
where ηk (r) and qk (t) correspond to the k th mode shape and modal coordinate
of the rotating rotor blade, respectively. Substituting Eq. 2.31 into 2.28 yields
X d2 d2 ηk Z R
d 2 dηk X
EI − mΩ ρdρ q k + mηk q̈k = Fz (2.32)
k
dr2 dr2 dr r dr k
The terms in the first summation in Eq. 2.32 can be replaced by Eq. 2.30,
yielding
X
(νk2 mηk qk + mηk q̈k ) = Fz . (2.33)
k
44
DIC OMA Load estimation Verification
Differentiation in time
Noise-reduced Distributed
3rd-order polynomial displacement inertial forces
curve fitting
CP Natural frequencies
Distributed Comparison
Mode shapes
aerodynamic forces
Modal amplitudes Modal expansion
approach Measured rotor thrust
Before the load estimation based on the measured data was examined,
a numerical experiment was carried out using a finite element model of a one
dimensional (1D) rotating cantilever beam with an arbitrary external forcing.
The numerically-derived blade elastic response and modal parameters were
used as an input to the present framework of rotor loads estimation approach,
and the known external force was compared to that estimated by the present
approach. The purposes of the numerical experiment were; (i) to examine if
the approach proposed in this dissertation was feasible to estimate rotor loads
distribution based on identified modal parameters and blade deformation and
(ii) to evaluate the influence of the participating number of modes on accuracy
of rotor loads estimation.
Figure 2.12 shows the original and estimated force distribution along
the blade span, obtained from the numerical experiment on the 40-element
(80 degrees of freedom) 1D rotating cantilever beam spun at 900 RPM. The
45
shape of the original excitation to the system followed a typical spanwise lift
distribution along a helicopter blade, that is, the lift force increases as the
local flow velocity increases towards the blade tip and the sharp drop occurs
at proximity of the blade tip due to the strong downwash from the trailed
blade tip vortices [68].
As can be seen in Fig. 2.10a, the first three modes were not enough
to reconstruct the unique shape of the typical airload over the blade span,
especially at regions near the blade tip. As the number of participating modes
increased, the accuracy of the load estimation was improved and the over-
all shape was satisfactorily captured with 10 modes, as shown in Fig. 2.10c.
From Fig. 2.10d to 2.10f, there was little improvement after the number of
participating modes went beyond 15 modes.
46
350 350
Original Original
Force per unit length, [N/m]
250 250
200 200
150 150
100 100
50 50
0 0
0 10 20 30 40 50 60 70 80 90 100 0 10 20 30 40 50 60 70 80 90 100
Radial location, [%R] Radial location, [%R]
250 250
200 200
150 150
100 100
50 50
0 0
0 10 20 30 40 50 60 70 80 90 100 0 10 20 30 40 50 60 70 80 90 100
Radial location, [%R] Radial location, [%R]
250 250
200 200
150 150
100 100
50 50
0 0
0 10 20 30 40 50 60 70 80 90 100 0 10 20 30 40 50 60 70 80 90 100
Radial location, [%R] Radial location, [%R]
Figure 2.10: Comparison between the original and estimated lift distribution
over the finite element 1D rotating beam
47
350
Original force
Force per unit length, [N/m] 300 Reconstructed force with 3 modes
250
200
150
100
50
0
0 10 20 30 40 50 60 70 80 90 100
Radial location, [%R]
Figure 2.11: Integrated force comparison between the original and recon-
structed loads
48
The area difference normalized by the mean original hub load and the RMS
error are plotted in percentage as a function of the number of modes partic-
ipating in the estimation process, as shown in Fig. 2.12a and 2.12b. Both
plots show the common trend that the difference dramatically decreases from
the first three to ten modes participation, and gradually converges towards
zero. The conclusion from the qualitative analysis, that is, hub loads might be
accurate even with a small number of modes participation, is now reinforced
by the fact that the hub load estimated from the first three modes falls within
approximately 2% deviation from the original value.
49
2.5
2
Area difference [%]
1.5
0.5
0
3 5 10 15 20 40
Number of modes
(a) Integrated loads difference
30
25
20
RMS
15
10
0
3 5 10 15 20 40
Number of modes
(b) RMS error of the spanwise lift distribution
Figure 2.12: Comparison between the original and estimated loads as a func-
tion of the number of participating modes
50
Chapter 3
The rotor loads estimation starts with measuring the deformation time history
of the rotor blade using the time-resolved DIC, which requires one to take a
sequence of digital images over an entire rotor disk. The diameter of the rotor
is an essential measurement parameter that determines the whole setup of
the DIC measurement, including camera positioning and optical equipment
arrangement. To show the scalability and flexibility of the methodology, the
51
current study performed the time-resolved DIC measurements on the small-
and large-scale rotor blade.
Using these hover test rigs, three test campaigns were conducted. Test
campaign 1 was performed on the small-scale, extremely flexible rotor hover
test stand, and Test campaigns 2 and 3 were performed on the large-scale,
2 m-diameter rotor test stand at different rotor configurations and operating
conditions. This chapter is divided into three sections. In § 3.1, the design,
specification, and instrumentation of each rotor test stand are documented.
In § 3.2, the camera arrangement, optical settings, and other details of DIC
rotor blade deformation measurement on each rotor test stand are presented.
In § 3.3, the measurement envelope and procedure of the three test campaigns
are summarized.
52
and makes it an ideal testbed for the DIC measurement. This subsection
describes the key components of this extremely flexible rotor hover test stand.
228 mm
0.45
9” m
Flexible blade
Fixed-pitch hub
Flexible blade Fixed-pitch hub
DCmotor
DC motor
Shaft
Shaft encoder
encoder
The two-bladed rotor test article, shown in Fig. 3.1, consists of two
flexible blades and a fixed-pitch rigid hub. The blade structure comprises two
plies of ±45◦ AS4/3501-6 prepreg and a tungsten rod (25.4 mm long, and 2.4
mm diameter) held at the blade tip in a chordwise direction (see Fig. 3.2) to
provide centrifugal stiffening and passive stabilization. A rectangular brass
plate (51 µm thick, 7.6 mm length in chordwise, and 15.2 mm length in span-
53
wise direction) is inserted between the two plies of the composite laminate
at regions of the leading edge, and a brass cylinder is directly soldered onto
the plate. The brass structure with a small amount of epoxy is used to hold
the tungsten mass at the blade tip. The blade is nominally untwisted with a
constant thin circular arc airfoil profile; the thin, open section profile results
in low bending stiffness and negligible torsional stiffness, both dominated by
centrifugal force.
The fixed-pitch hub was designed in CAD and fabricated using a rapid-
prototyping plastic 3D printer. Four different root pitch angles were selected
for hover testing; 0◦ , 10◦ , 20◦ , and 30◦ . The rotor parameters are summarized
in Table 3.1; note the low thickness to chord ratio of the rotor blades. The
whole rotor assembly is placed onto a six-component strain gage load cell (ATI
Mini-40) to measure hub loads. The rotor is driven by a brushless DC motor
(Hacker A50-16S) and the rotational speed is measured by a 1024/rev optical
incremental encoder (US Digital E5), as shown in Fig. 3.1.
54
Tip mass
Blade grip Leading edge
Trailing edge
202 mm
Figure 3.2: The hub and the flexible blade with the fluorescent paint and
random speckle pattern
55
measurement) at integer multiples of the rotational frequency. The one-per-
revolution pulse from the Z-index was used to synchronously start analog data
logging and image acquisition.
56
Transmission Counter weight
Hydraulic motor
1.016 m
Two-bladed
single rotor
Swashplate assembly
A-shaped frame
Transmission
Hydraulic motor
Outer shaft rotor
The CCR rotor system is designed to replicate the X2TD rotor sys-
tem [69], featuring a rigid hub and a closely-spaced CCR rotor. The specifi-
cations of the rotor system are summarized in Table 3.2. For the CCR rotor
system, the blade passages occur at the top and bottom of the rotor disk, i.e.,
58
at azimuthal angles of 0◦ and 180◦ . In single-bladed rotor configurations, one
rotor blade and a counter-weight are attached to each rotor, whereas in two-
bladed rotor configurations two identical blades are attached to each rotor. A
counter-weight is attached to the hingeless hub to balance the location of the
center of mass (CM) of each rotor. The counter-weight assembly consists of a
cylindrical-shaped steel solid piece, a threaded rod, and a rod fixture on the
hub. To adjust the location of CM of the entire rotor assembly, the location
of the steel piece can be readily changed by rotating the threaded rod.
Table 3.2: Summary of rotor parameters for the large-scale hover test stand
Rotor blades used on the large-scale test stand have the following fea-
tures: no taper, constant chord, uniform VR-12 airfoil section including a
5% trailing edge tab. The blade structure is made of a foam core and for-
ward D spar, wrapped with carbon-epoxy composite materials (a plain-weave
59
AS4/3501-6 prepreg and an IM7/3501-6 uni-directional carbon fiber tape).
The foam core in the D-spar is machined to place tungsten weights along the
leading edge, for adjusting the location of the blade center of gravity. The
±45◦ orientation plain-weave fabric provides increased torsional stiffness. Ro-
hacell IG31-F closed cell foam is used for the core and FM-300K film adhesive
is used to bond the carbon fiber prepreg to the form core. At the blade root, a
rectangular aluminum insert is placed onto the form core to provide compres-
sive strength where the rotor blade is clamped in the blade grip of the hub.
Figure 3.5 shows a breakdown of the blade laminate lay-up.
Carbon-epoxy
composite
Tungsten leading
edge mass
Aluminum blade
root insert
60
3.1.2.4 Rotor hub
61
Blade grip assembly
Load cell
cies of ±0.5 Nm. While the isolated load cell has a natural frequency above
1000 Hz, the addition of the hub, blade, and counter-weight significantly mod-
ifies the frequency response. An in-situ dynamic calibration was performed
on the vertical (FZ ) component of each load cell using a calibrated impact
hammer. The resulting transfer functions for the upper and lower rotors are
shown in Fig. 3.8. The load cell response remains flat until about 140 Hz, or
7-per-revolution at a rotor speed of 1200 RPM. These transfer functions were
62
Needle bearing
Thrust bearing
Blade grip
Bearing carrier
used to correct the measured FZ hub loads presented throughout the study.
Two different blade root pitch angle sensors were tested: A linear
Hall effect sensor (Honeywell SS495A1) and an Anisotropic Magneto-Resistive
(AMR) position sensor (KMZ60).
The Hall effect sensor is mounted onto a stationary fixture inside the
central hub structure, and two Neodymium magnets are bonded to the bottom
surface of the blade grip. This root pitch measurement system operates based
on the Hall effect: the sensor measures the variation of magnetic field as the
two Neodymium magnets of opposite polarity rotate with the pitching motion
of the rotor blade grip. Details of the instrumentation and specification of the
Hall effect pitch angle sensor are available in Ref. [70]. The Hall effect sensor
63
Figure 3.8: Dynamic response of upper and lower rotor load cells
was used for initial tests, however, it was found that the sensor often needed
re-calibration.
Fig. 3.9 shows a schematic of the AMR sensor installation on the blade
grip assembly. The AMR sensor generally consists of a thin film of ferromag-
netic metals such as Ni and Fe. The resistance of the thin film varies according
to the strength of the applied magnetic field along a specific direction. The sen-
sor uses the variation of resistance, which results in voltage change. The AMR
sensor is held on a stationary plate and a Neodymium magnet is bonded to the
face of the blade grip, as shown in Fig. 3.9. The stationary plate is mounted
on to the blade grip carrier in a way such that the feathering axis of the blade
64
grip is aligned to the coordinate system of the AMR sensor. Then by placing
the magnet at the center of the blade grip face, the AMR sensor can detect
the variation of the magnetic field associated with the feathering motion of
the rotor blade.
Isometric view
𝜃
Section A
Section B
AMR sensor
Section B
Rotor hub
𝜃
AMR sensor
Sensor holder
Figure 3.9: A schematic of the AMR sensor installation on the blade grip
assembly
The AMR sensor provides ratiometric sine and cosine analog output
signals and its calibration is performed using an absolute optical inclinometer
(US A2T series). Figure 3.10a shows the variation of the two output voltage
signals from the AMR sensor as a function of rotational angle measured by
the inclinometer. A rotating magnetic field delivers the two sinusoidal output
signals with the double frequency of the mechanical angle between the sensor
and magnetic field direction. The angle can be calculated using the following
65
equation:
arctan (VSIN /VCOS )
α= (3.1)
2
where α is the angle between the sensor and magnetic field direction, VSIN
and VCOS are sine and cosine analog output signals, respectively. A sinusoidal
function f (x) = A sin (Bx + C) + D curve-fitted to the sensor output over the
angle range of interest is shown in Fig. 3.10b and the curve-fitting coefficients
with 95% confidence are summarized in Table 3.3. The uncertainties based
on the standard deviation calculated from three individual repeated measure-
ments are shown as the highlighted area in Fig. 3.10c. The measurement
uncertainty was computed to be ± 19 mV and corresponds to approximately
0.38 deg. Note that the current study used only the sine signal due to the
number of measurement channels available in the data acquisition system.
Table 3.3: Summary of the curve-fit coefficients to the AMR sensor output
signal
66
5 5
Output 1
Output 2 Output signal
4 4 Curve fit
Output amplitude [V]
2 2
Curve fit model
f(x) = A sin(Bx+C)+D
1 1
0 0
-180 -120 -60 0 60 120 180 -40 -20 0 20 40 60
Angle [deg.] Angle [deg.]
(a) Sine and cosine output signals from (b) Curve-fit to the voltage variation
the AMR sensor from the AMR sensor
3
2.75
Output Voltage, [V]
2.5
2.25
1.75
1.5
-10 -5 0 5 10 15
Angle [deg.]
(c) Output voltage signals from the AMR
sensor with measurement uncertainty
67
pitch linkage. Due to the unique load path and load cell configuration of the
rotor assembly, pitch link loads must be measured and subtracted from load
cell measurements for recovering the true rotor hub loads. As such, tension-
compression load cells with a full-scale range of ±250 N and an accuracy of
±0.5 N are installed in each pitch link to measure the variation of pitch link
forces as a function of blade azimuthal angle and appropriately correct the
hub load cell measurements.
• A lead-screw actuator
• A hydraulic actuator
The first design, a lead screw servo actuator, uses a lead-screw rod, driven by
a brushed DC motor through a gear reduction system. The main design target
of this servo actuator was to maintain or change the blade pitch angle under
severe rotor load conditions with high servo resolution and low compliance.
Table 3.4 summarizes the specification of the lead-screw servo actuator, and
further details can be found in Ref. [70].
68
Table 3.4: Specifications of the lead-screw servo actuator
piston and a cylinder body) was designed in-house and fabricated by an exter-
nal manufacturer (SMC Corporation of America). The hydraulic servo valve
(KNR Systems) has a capability of operating 5.5 L/min at pressure of 70 bar at
a bandwidth above 60 Hz. The piston position feedback sensor (SICK MPS-T
position sensor) is mounted on a T-slot outside the cylinder body and con-
tinuously detects the variation of the magnetic field generated from a magnet
installed on the piston. The whole rotor control assembly with the hydraulic
actuation system is shown in Fig. 3.12
A block diagram of the pitch angle control system for the hydraulic ac-
tuation is shown in Fig. 3.13. In the pitch control signal line, input commands
(collective and cyclic pitch angle position) are first provided to a central con-
trol panel. These PWM signals are then sent to a signal distribution panel,
which takes 12 V power from a power supply, and the power and signals are
distributed to three micro-controllers (Pololu jrk 21v3). Each micro-controller
takes a feedback signal from the position sensor and drives the hydraulic servo
valve of each actuator for closed-loop control of the piston position.
69
M6 rod end
Piston
Return
Pressure
M6 rod end
tems) control software. First, users set a pressure level and hydraulic flow rate
on the software and these parameters are sent to a mobile Hydraulic Power
Unit (mHPU) that supplies the hydraulic flow at the specified pressure and
rate. The fluid flow is distributed by a hydraulic manifold to the three indi-
vidual actuator assemblies. This hydraulic actuation system allows dynamic
variation of rotor blade pitch angle, as compared to the lead-screw servo ac-
tuation system.
70
Swashplate assembly
Hydraulic line
Pushrod
Rotor hub
Pitch horn
Swashplate
Figure 3.12: Rotor pitch angle control system assembly with the hydraulic
servo actuators
71
Hydraulic line Pitch angle Pressure and flow rate
Signal line User input User input
Power line
User control panel Pump control software
Power supply
Servo valve Position sensor Servo valve Position sensor Servo valve Position sensor
Actuator 1 Actuator 2 Actuator 3
Figure 3.13: A diagram of pitch angle control system for the hydraulic actua-
tion
72
transmission for monitoring the fixed-frame vibration. Data from the ac-
celerometer is also used to analyze the dynamic imbalance of the rotor system
before testing, and to monitor dynamic unsteady loads observed during ac-
tual hover testing. Four precision integrated-circuit temperature sensors with
an output voltage linearly-proportional to the Centigrade temperature (Texas
Instruments LM35) are installed onto four bearing housings (two for the main
shaft, and two for the drive shaft) to monitor bearing health.
73
3.1.2.8 Data acquisition
DIC blade deformation measurement on the small scale rotor hover test
stand (see § 3.1.1) was performed in the test chamber shown in Fig. 3.14 and its
schematic is shown in Fig. 3.15. To acquire the time history of the rotor blade
deformation, a sequence of digital images over the whole rotor disk must be
taken. This was accomplished by mounting high-speed digital cameras on lin-
ear camera rails above the rotor plane with their axes oriented approximately
45◦ with respect to the rotor disk, as shown in Fig. 3.15. The stereoscopic ar-
74
rangement of two cameras (Phantom Miro M310 with Nikon NIKKOR f/1.8D
35 mm lens) enables measurements of three-dimensional displacement fields
on the entire flexible blade. Knowing the positions of two cameras relative to
each other and the magnifications of lenses, the stereoscopic DIC algorithm
can calculate the absolute three-dimensional coordinates of any point on the
blade surface.
Figure 3.14: The setup for small-scale rotor blade deformation measurement.
75
Mirror
Camera 2
Camera 1
Diffuser
Beam Rotor
Motor
Laser
Mirror
76
of the structure and the environmental background; however, this approach
may cause out-of-focus images due to shallow depth of focus inherent to a large
aperture setting. Increasing exposure time while maintaining a small aperture
value is an alternate approach, however, increased exposure time would cause
more blurry images. Thus, the present study makes novel use of two techniques
for DIC measurement: fluorescence and pulsed laser illumination.
77
surface. A Photonics dual-head Nd:YLF DM-30 pulsed laser strobe with a
wavelength of 527 nm and a pulse energy of 30 mJ/pulse routed through
an engineered diffuser was used to illuminate the entire rotor disk at any
desired instant of time (as shown in Fig. 3.15). The short duration pulsed
laser illumination enabled a high f-number (f/8.0D) and a short exposure time
(10µs) for all the images captured for DIC. The optical incremental shaft
encoder on the rotor shaft was used to generate a 16-per-revolution pulse
train, which triggered laser strobing and image acquisition at 16 evenly-spaced
azimuthal locations. Since the maximum rotational speed was set at 25 Hz,
the maximum sampling rate was 400 Hz. At each test condition, images were
taken over 400 rotor revolutions, yielding 6400 images per camera over 16
seconds. Note that reference (undeformed blade) images for DIC calculations
were captured at the same azimuthal locations by rotating the rotor shaft
slowly by hand, i.e. at a negligibly small rotational speed, thereby eliminating
any aerodynamic forcing. Figure 3.16 shows the unprocessed images of the
rotor at 1200 RPM acquired by the high-speed camera at 16-per-revolution,
overlaid on each other to illustrate the complete rotor disk; a typical DIC
interrogation window is also shown for reference.
78
Interrogation
window
selected as 19 x 19 pixels and the window shift was set to be 5 pixels for map-
ping the interrogation windows to the whole image, resulting in 54% overlap.
The camera scaling factor was approximately 2.24 pixel/mm determined by
the geometry of the test setup as well as the camera lens characteristics. De-
formations at a total of 988 locations (13 chordwise and 76 spanwise locations)
were calculated at each azimuth.
79
of the calibration plate with respect to the rotor plane enables calculation of
a mapping function between the global coordinates and camera image plane
coordinates. With the help of mapping functions calculated during the camera
calibration process, the two two-dimensional vectors from each camera are
combined to yield a three-dimensional vector. The accuracy of the vectors
depends on the size of the interrogation window and is estimated to be 0.01%R
(0.023 mm), while the spatial resolution of the displacement field is 0.2%R
(0.45 mm). A sample processed result for the 30◦ root pitch rotor, rotating at
1500 RPM, is shown in figure 3.18.
80
10
Blade root 9
2
Blade tip
1
Figure 3.18: Vector map processed by DIC for the 30◦ root pitch rotor, spun
at 1500RPM
81
imaging of the whole rotor disk as shown in Fig. 3.3. Other modifications are
discussed in this subsection.
Figure 3.19 and 3.20 show one of the stereoscopic camera arrangements
used for the large-scale DIC measurement. In the setup, three high-speed
cameras were used to image the entire rotor disk. The first and second cam-
eras (Phantom Miro M310 with Nikon NIKKOR f/1.8D 35 mm lenses) were
mounted so that each camera captured the entire rotor disk, and the third
camera was mounted on a tripod located between the first and the second
camera. At some azimuthal locations, images taken by the left and right cam-
eras (Camera No. 1 and Camera No. 3) were distorted due to large parallax.
These images were corrected using images taken by the middle camera (Cam-
era No. 2).
As was the case for the small-scale DIC measurement, orange fluores-
cent paint with stochastic speckle patterns was applied to the blade surface as
shown in Fig. 3.21, and the LIF strobing was achieved by a digital pulse train
generated from the optical encoder installed onto the drive shaft (see § 3.1.2).
A schematic of the whole DIC measurement setup is shown in Fig. 3.22. Cam-
era exposure time was maintained at 10 µs. The short exposure time with
laser strobing enables a high contrast (almost pure black and white), blur-free
image acquisition. Figure 3.23 shows unprocessed images taken by the three
cameras that were triggered at a certain azimuthal angle.
82
Figure 3.19: A schematic of camera arrangement for the large-scale DIC test
and camera image plane coordinate system. To cover the larger field of interest
than that for the small-scale DIC, a large calibration plate was used, as shown
in Fig. 3.24. The large calibration plate was moved around to cover the whole
area of the rotor disk. The displacement accuracy for the large-scale rotor
blade was estimated to be 0.01%R (0.1 mm), and the spatial resolution of the
displacement field was estimated to be 0.2%R (2 mm).
83
Rotor Camera 3
Camera 1
Camera 2
Figure 3.20: Stereoscopic camera system for the single-bladed CCR rotor
1.016 m
0.122 m
𝒚
𝒙
Figure 3.21: Rotor blade painted with orange fluorescent color and black dots
84
Rotor Wake direction
Camera 1, 3
Diffuser
Camera 2
Laser Mirror
Beam
It should be highlighted here that output data obtained from the com-
mercial DIC software must be post-processed by the analyst to recover the
continuous time history of the rotor blade deformation. There are two fac-
tors that must be addressed: First, a whole area of each digital image is put
into a Cartesian grid and is divided into a number of interrogation windows.
Thus, the windows placed over the rotor blade structure do not correspond
to the blade-fixed coordinate system, that is, the measurement points are not
85
Left image Right image
Center image
Calibration target
Rotor blade
86
uniformly distributed in either chordwise or spanwise direction over the blade
structure. Second, output data structure from the software is not organized
in time order because the DIC processing in the software is not performed
between images taken at different azimuthal locations but performed between
images taken at a certain azimuthal location at different time steps. This is
because of the fact that in-plane displacements between different azimuthal
locations are too large to compute with the DIC algorithm.
With the experimental setups described in § 3.1 and § 3.2, three indi-
vidual test campaigns were performed in the present study. Test campaign 1
was performed on the small-scale rotor hover test stand, whereas Test cam-
87
paigns 2 and 3 were performed on the large-scale hover test stand at different
rotor configurations and operating conditions. Table 3.5 summarizes the mea-
surement setups of the three test campaigns, and details of each campaign are
described in the following subsections.
The collective pitch angle was first set by selecting the fixed-pitch hub,
and the rotor with the extremely flexible blades was spun at a target rotational
speed. Having synchronized the timing of two high-speed cameras and laser
88
Table 3.6: Summary of the test matrix for Test campaign 1
89
operational modal parameter identification. Then, the single-bladed isolated
rotor was tested at a moderate thrust, where the rotor produced 70 N corre-
sponding to a blade loading coefficient of CT /σ = 0.07. For all test conditions,
the rotational speed Ω was set at 900 RPM.
90
Table 3.7: Summary of operating conditions for Test campaign 2
While the single-bladed, single or CCR rotor with the lead-screw pitch
actuation system was used during the second test campaign, the two-bladed,
isolated single rotor with the hydraulic pitch actuation system was used dur-
ing Test campaign 3. The two-bladed rotor stand and its hydraulic actuation
system are shown in Fig. 3.3b and 3.12, respectively. The hydraulic servo actu-
ators enabled three different rotor pitch control inputs: (1) a steady collective
pitch input at a blade loading (CT /σ) of 0.125, (2) a lateral cyclic pitch input
(θ1C = 2◦ ), and (3) a collective step pitch input (2◦ step function). For all
of these conditions, the rotor was spun at 900 RPM. As was the case of Test
campaign 2, the blade deformation time history and rotor loads were simul-
91
taneously measured at each operating condition. These three test conditions
are summarized in Table. 3.8.
92
Camera 1
Top view
Diffuser
45°
Rotor Mirror
A-shaped frame
Camera 2
Side view
Camera 1, 2
Wake direction
Beam
Mirror Laser
Optical bench
On the other hand, three independent data sets were obtained for the
test condition 3, each recorded with the digital pulse triggering the step pitch
93
actuation. The initial intention was to use the trigger as a reference point for
phase-averaging data sets taken from multiple individual experiments; how-
ever, the pitch angle responses were not consistent over the three experiments.
Figure 3.26 shows the pitch angle variations, measured by the AMR root pitch
angle sensors, for three independent measurements. The rise time of the col-
lective increase and the time lag between the trigger and blade pitch increase
were calculated to be approximately 200 and 75 milliseconds for all the three
different experiments, respectively. However, the magnitudes of the pitch an-
gle response were not consistently maintained; for example, the blade pitch
was changed from 9.5◦ to 11.5◦ during the first experiment, whereas the pitch
◦
angle was changed from 9.5 to 11.0◦ during the second experiment as shown
in Fig. 3.26. Hence, phase-averaging was not performed on the measured quan-
tities for the test condition 3, and results from the first experiment at which
the blade pitch successfully responded to the 2◦ collective step pitch input will
be primarily discussed in this dissertation.
Table 3.9: Summary of the design parameters for digital low-pass filters
94
12 12 12
Blade 1
Pitch angle, [deg] Blade 2
11 11 11
10 10 10
9 9 9
8 8 8
27 28 29 30 31 32 35 36 37 38 39 40 24 25 26 27 28 29
1 1 1
Digital pulse
Digital pulse
Digital pulse
0 0 0
27 28 29 30 31 32 35 36 37 38 39 40 24 25 26 27 28 29
N [revs] N [revs] N [revs]
Figure 3.26: Pitch angle variations and the initial digital trigger timing ob-
tained from the three independent measurements
cies were selected for each measured quantity, thus, the actual frequencies
used for this filtering process will be described along with experimental results
in the next chapter. Since an IIR low-pass filter introduced a frequency-
dependent delay and phase shift between input and output signals, a specific
post-processing algorithm was applied to the output signal in time domain to
compensate for the delays. The algorithm essentially applies the digital low-
pass filter to the input data in both the forward and backward directions in
time to remove the delay.
In summary, this chapter covered the details of the hover test stands,
95
the blade deformation measurements using the time-resolved DIC, the rotor
loads measurements, and the test envelopes and procedures for the small- and
large-scale rotor configurations.
96
Chapter 4
This chapter discusses the results of the rotor blade deformation mea-
surements, the operational modal identification, and the rotor loads estimation
for several different rotor systems and operating conditions. First, the results
of the time-resolved DIC deformation measurements performed on the small-
and large-scale rotor test rigs (see chapter 3) are discussed in § 4.1. Second, the
rotor blade modal characteristics identification with several OMA algorithms
based on the blade deformation time history are described in § 4.2. Finally,
the results of the rotor loads measurements and estimation are summarized
in § 4.3. A block diagram shows the organization of this chapter in Fig. 4.1.
Portions of this chapter were previously published as “Blade Passage Loads and De-
formation of a Coaxial Rotor System in Hover” [57] in the AIAA Journal of Aircraft and
“Full-field optical deformation measurement and operational modal analysis of a flexible
rotor blade” [75] in the Journal of Mechanical Systems and Signal Processing. All writ-
ing and figures included in this chapter are the original work of the author, incorporating
computational data from Dr. Roland Feil and Dr. Juergen Rauleder, with editing by Dr.
Jayant Sirohi.
97
Section 4.1 Subsection
Rotor loads estimation and measurement • Single-bladed, single or CCR rotor (large-scale test rig)
• Two-bladed, single rotor (large-scale test rig)
two-bladed, extremely flexible rotor in § 4.1.1 and (2) the large-scale DIC mea-
surement on the 2.0 m-diameter, single-bladed or two-bladed, single or coaxial
rotor in § 4.1.2. The goals of these measurements were to verify and vali-
date if the time-resolved DIC technique with laser strobing and fluorescence
paint, developed in the dissertation, could be applied to different measurement
scales, rotor configurations, and operating conditions. Note that the correct-
ness of the 3D DIC deformation in static, non-rotating frame has been proved
by comparing results with those obtained by a laser displacement sensor, as
thoroughly described in [70].
98
4.1.1 Small-scale rotor
Figure 4.2 and Table 4.1 shows a recap of the rotor setup and test
matrix used for this small-scale DIC test, respectively. The blade deformation
measurements were performed at four different blade root pitch angles and
rotational speeds during Test campaign 1 (see in § 3.3).
228 mm
0.45
9” m
Flexible blade
Fixed-pitch hub
Flexible blade Fixed-pitch hub
DCmotor
DC motor
Shaft
Shaft encoder
encoder
Table 4.1: Recap: Summary of the test matrix for Test campaign 1
Figure 4.3 shows the flap bending deformation of the rotor blade at 30◦
root pitch angle spun at 1200 RPM over one revolution. Note that images were
continuously taken over 400 revolutions at 16 evenly-spaced azimuthal loca-
tions, yielding 6400 images per camera at each test condition. The variation
of the blade flap bending over one revolution shown in Fig. 4.3 was obtained
99
by three steps: (i) The time history of blade deformation over 400 revolutions
was phase-averaged to yield the mean blade bending deformation as a function
of radial station at each of the 16 azimuthal locations. (ii) Displacement vec-
tors along the quarter-chord axis were extracted from the phase-averaged data
set. (iii) Data was interpolated between adjacent azimuthal angles, yielding
the out-of-plane bending deformation over the entire rotor disk; note that the
phase-averaging process eliminates most of the stochastic noise in the mea-
surement.
Figure 4.3: Flap bending deformation over the rotor disk at θ0 = 30◦
The rotor blade at a root pitch angle of 30◦ experienced a positive flap
bending deformation, increasing along the radial direction, and reached ap-
proximately 7 mm at the blade tip. The contour plot shown in Fig. 4.3 clearly
100
presents the azimuthal variation of the blade flapping motion, regardless of
the fact that blade pitch angle was fixed at the root. Over the range between
Ψ = 0◦ and 180◦ , the blade continuously flapped down from 7 mm to 4.5 mm
measured at the tip and started flapping up during the other half of the rev-
olution. This 1/rev motion was possibly caused by the combination of three
major factors: (i) the extreme torsional flexibility of the rotor blade structure,
(ii) no active stabilization mechanism, and (iii) some mean asymmetric air flow
in the test chamber due to flow recirculation. Although there was a passive
stabilization control system (i.e., a blade tip mass inducing propeller moment
as shown in Fig. 3.2), the rotor blade experienced significant elastic twist due
to the low blade torsional rigidity, resulting in azimuthal variation of blade
sectional angle of attack.
101
spectively. The negative twist along the rotor blade increased with rotor speed
due to the centrifugal moments on the tip mass; at 600 RPM, the blade had
a spanwise nose-down twist of about 1◦ , whereas at 1500 RPM it increased to
a nose-down twist of nearly 10◦ . The increase in pitch angle at the blade tip
is due to the presence of the tip mass. The shaded area in Fig. 4.4a to 4.4c
represents the estimated DIC measurement uncertainty based on the standard
deviation and the size of interrogation window as specified in the software
documentation.
102
8 1
600 RPM
600 RPM
7 900 RPM
900 RPM
0 1200 RPM
1200 RPM 1500 RPM
6 1500 RPM
Deflection [mm]
Deflection [mm]
-1
5
4 -2
3
-3
2
-4
1
0 -5
0 20 40 60 80 100 120 140 160 180 200 0 20 40 60 80 100 120 140 160 180 200
Radial location [mm] Radial location [mm]
30
Blade pitch angle [deg]
28
26
24 600 RPM
900 RPM
1200 RPM
22 1500 RPM
20
0 20 40 60 80 100 120 140 160 180 200
Radial location [mm]
Figure 4.4: Static blade deformations of each degree of freedom, along the
blade quarter-chord axis at 30◦ root pitch and different rotational speeds
103
0
= 0°
30 30 = 10°
= 10° 0
0 = 20°
25 = 20° 25 0
0 = 30°
20 0
= 30° 20 0
15 15
10 10
[deg]
[deg]
5 5
0 0
-5 -5
-10 -10
-15 -15
-20 -20
-25 -25
-30 -30
0 0.1 0.2 0.3 0.4 0.5 0 0.05 0.1 0.15 0.2
time [s] time [s]
(a) Rotor spun at 600 RPM (b) Rotor spun at 1500 RPM
Figure 4.5: Time history of the pitch angle displacement at the blade tip for
different root pitch angles
104
deformation measurements, by comparing the experimental results with the
numerical predictions. Note that the blade structural properties (the bending
and torsional stiffness) used in the numerical modeling were obtained from
multiple DIC deformation measurements in the non-rotating frame, resulting
from static loadings in out-of-plane, in-plane, and torsional degrees of freedom.
Further details of the structural characteristics identification can be found in
Ref. [70].
The single-bladed CCR rotor system, shown in Fig. 4.6 as a recap, was
tested at a mean thrust of the lower rotor of 90 N (CT /σ = 0.09). The DIC
deformation measurements were carried out only on the lower rotor of CCR
system (the orange-painted blade shown in Fig. 4.6), due to the geometrical
limitation of experimental setup. This rotor was spun at 900 RPM and the
coaxial system was trimmed to achieve torque balance, that is, the total yaw
moment was trimmed to be zero.
105
Transmission Counter weight
Hydraulic motor
1.016 m
106
(a) Measured CCR lower rotor blade flap bending de-
formation
180°
150° 210° 80
70
120° 240° 60
Deflections, [mm]
50
0.2 270° 40
90° 0.6 0.4
30
60° 300° 20
10
30° 330° 0
= 0°
Figure 4.7: Flap bending deformation of the CCR lower rotor blade at CT /σ
= 0.09
107
Figure 4.8 shows the measured and predicted CCR lower rotor blade
tip displacement extracted at the blade quarter chord location as a function
of azimuthal angle at CT /σ = 0.09. The trend revealed the significant 2/rev
characteristics, that are associated with the blade crossings located at ψ =
0◦ and 180◦ . This 2/rev characteristic reflects the lower rotor blade response
to the transient loads generated by the upper–lower rotor blade crossing. For
this investigated thrust level (i.e., a blade loading coefficient of CT /σ = 0.09),
vibrations around the mean deflection of 69.6 mm occurred with a maximum
deflection of 74.6 mm and a minimum deflection of 66.4 mm. The maximum
peaks of blade tip deflection were seen at approximately ψ = 30◦ and 210◦ ,
whereas the minimum peaks were observed at approximately ψ = 150◦ and
290◦ . The 2/rev half-peak-to-peak magnitudes of 4.1 mm corresponded to
5.9% of the mean deflection. Predicted results correlated satisfactorily with
these measurements in both magnitude and phase. The numerical predictions
showed that the upper rotor blade had a greater mean deformation, which
is consistent with its greater thrust share compared to the lower rotor in a
torque-balanced trim condition.
Figure 4.9 shows the measured and predicted blade deformation of the
lower rotor in the coaxial system along the blade quarter-chord axis as a func-
tion of radial station. The two lines correspond to the azimuthal locations of
maximum and minimum tip deflection that were previously determined from
Fig. 4.8. There was excellent agreement between the measured and predicted
flap bending (out-of-plane) deflection, with the largest discrepancy limited to
108
90 Coaxial lower measurement
Coaxial lower prediction
Coaxial upper prediction
85
80
75
70
65
60
0 60 120 180 240 300 360
Figure 4.8: Lower rotor blade tip displacement comparison between measure-
ment and prediction
less than 3% (2.5mm) at the approximately 55% of the radial station on the
minimum deflection curve. Those errors were also more apparent over the
inboard portion due to smaller magnitude deflections.
The phase lag between the maximum and minimum deflection and the
blade crossings (i.e., ψ = 0◦ and 180◦ ), shown in Fig. 4.8, can be explained by
the influence of rotor blade bound circulation. As the upper and lower rotor
blades approach each other, the bound vortex on one blade induces interac-
tional upwash on the other blade, resulting in an increase in effective angle
of attack of the other blade, as shown in Fig. 4.10. The increase in angle of
attack causes a 2/rev flap bending deflection of the rotor blade. Aerodynamic
109
80
Measurement
Prediction
60
max.
40 min.
20
0
0 0.2 0.4 0.6 0.8 1
Figure 4.9: Maximum and minimum flap bending deformation of the lower
rotor blade as a function of blade radial station at CT /σ = 0.09
110
Before crossing After crossing
Γ"##$% Γ"##$%
Upwash Downwash
Upper
Γ&'($% Γ&'($%
Free stream
velocity
Airflow induced by
upper rotor
Airflow induced by
lower rotor
Figure 4.10: Interaction of blade bound circulation in the CCR rotor system
due to the blade crossings in the CCR rotor system, which could only be
identified by analyzing the continuous time history of the blade deformation.
111
Two-bladed
single rotor
Swashplate assembly
A-shaped frame
Transmission
The first operating condition tested on the two-bladed single rotor was
a steady collective pitch input (θ0 = 9◦ ) at a blade loading CT /σ of 0.125.
The deformation measurement was carried out on one of the two rotor blades
during this test campaign. Note that the deformation data shown in the
current section were computed based on the averaged reference image taken
at the blade root pitch angle of 0◦ , spun at the same rotational speed of 900
RPM.
First, the out-of-plane blade deformation over the entire rotor disk is
shown in Fig. 4.12. The flap bending deflection reached nearly 90 mm at the
blade tip with little azimuthal variation. The axisymmetric trend represents
112
the fact that the constant collective pitch angle of 9◦ was properly applied
to the rotor and steady-state loading over the rotor disk was achieved. The
phase-averaged blade tip displacement over one revolution was then extracted
from the contour plot and is shown in Fig. 4.13, representing the data both
from the 32 measurement points and from the spline interpolation. A slight
fluctuation over the entire range was clearly observed in Fig. 4.13, especially
over the range from 60◦ to 180◦ where a large rise-and-fall between 85 mm and
95 mm occurred. One possible cause of this dip is that one of three hydraulic
servo actuators might have had non-ideal free-play in the linkage path, such
as a rod end, resulting in the ± 5 mm fluctuation that periodically occurred
over the range.
Figure 4.12: Out-of-plane blade deformation over the entire rotor disk
113
110
32 data points
105 Interpolation
100
Tip path, [mm]
95
90
85
80
75
70
0 60 120 180 240 300 360
Azimuth, [deg]
Figure 4.13: Blade tip displacement as a function of azimuthal angle
Recall that one of the advantages of the DIC technique is that it can
measure the three-dimensional (3D) full-field deformation on the surface of
a target structure. Figure 4.14 shows the 3D static deformation distributed
over the blade span. This 3D deformation with high spatial resolution cannot
be obtained if conventional on-blade sensors, that can only perform pointwise
measurements, are used. From the 3D plot, deformations of each degree of free-
dom (flap, lead-lag, and torsion) along the quarter-chord axis were extracted
and plotted in Fig. 4.15.
The spanwise variations of the blade flap and lag deformation shown in
Fig. 4.15a and 4.15b represent the out-of-plane and in-plane static deflection at
114
the given rotor pitch angle and rotational speed, respectively. One can clearly
see a flap-up and lagging motion corresponding to increase in lift and drag
forces as the blade angle of attack increases. While the flapwise and chordwise
deformation along the blade quarter-chord axis was directly interpolated from
the 3D distributed plot, the sectional blade pitch angle, shown in Fig. 4.15c,
was obtained based on the 1st-order polynomial fitting to the 3D surface height
distribution at each spanwise station. The relationship between the sectional
pitch angle and the slope of the fitting is as follows:
wi (yi ) = P1 yi + P0
(4.1)
θi = arctan P1
where wi , yi , and θi are out-of-plane displacement, chordwise location, and
local torsional deformation at a spanwise station i, respectively. P1 and P0
represent the slope and y-intercept of the 1st-order polynomial fit.
As can be seen in Fig. 4.15c, the elastic twist of the rotor blade was
small, considering the fact that the blade root pitch was trimmed to the
mean collective angle θ0 = 9◦ measured in real time using the AMR sensor
(see § 3.1.2). There was a dip observed in regions of the blade tip, however,
its magnitude was within a fraction of a degree and on the same order of
measurement uncertainty. Note that highlighted areas in these plots represent
uncertainties of each measured quantity.
115
Figure 4.14: 3D static deformation distributed over the entire blade span
The collective pitch was adjusted so that the mean blade loading of this test
condition remained the same as for the first operating condition (θ0 = 9◦ ).
The lateral cyclic pitch θ1C was set to be ±2◦ , by introducing a swashplate in-
clination with the hydraulic-driven actuators. Rotor track was achieved using
the real-time hub loads measurement, specifically hub pitching and rotating
moment in the rotating frame.
The contour plot in Fig. 4.16 shows the phase-averaged flap bending
deformation over one revolution. Due to the lateral pitch input and the gy-
roscopic phase shift, the whole rotor disk tilted in the lateral direction. This
behavior was more pronounced in the blade tip displacement, as shown in
Fig. 4.17. The sinusoidal motion over one revolution as well as the phase shift
116
110 0
100
-2.5
80
70 -7.5
60
-10
50
40 -12.5
30 -15
20
-17.5
10
0 -20
0 10 20 30 40 50 60 70 80 90 100 0 10 20 30 40 50 60 70 80 90 100
Radial location, [%R] Radial location, [%R]
(a) Out-of-plane deformation (b) In-plane deformation
10.5
Torsional deformation, [deg.]
10
9.5
8.5
8
20 30 40 50 60 70 80 90 100
Radial location, [%R]
(c) Torsional deformation
Figure 4.15: Static blade deformations of each degree of freedom, along the
blade quarter-chord axis
In Fig. 4.17, the maximum flap bending deflection occurred at the az-
imuthal angle of 250◦ , whereas the minimum occurred at 60◦ . The tip deflec-
tion difference between the maximum and minimum reached approximately 75
mm. In other words, it was demonstrated that this large sinusoidal oscillatory
117
Figure 4.16: Out-of-plane blade deformation over the entire rotor disk for the
lateral pitch angle input
motion of the 2 m-diameter rotor at 1/rev frequency (15 Hz) was successfully
captured by the DIC technique.
The last operating condition tested during Test campaign 3 was a col-
lective step change input. The rotor pitch angle was first set to be at a con-
stant collective angle θ0 = 9◦ , corresponding to the blade loading (CT /σ) of
0.125. Once the rotor reached a steady-state condition, the collective pitch was
rapidly increased by approximately 2◦ to examine the rotor transient responses
to the dynamic pitch control from an aeroelastic perspective. This type of step
(practically ramp) pitch input excites the structure such that the responses at
different modal frequencies could be studied, and may help to extract higher
118
140
32 data points
Tip displacement, [mm] Interpolation
120
100
80
60
40
0 60 120 180 240 300 360
Azimuth, [deg]
Figure 4.17: Blade tip displacement as a function of azimuthal angle for the
lateral pitch angle input
modes with the operational modal analysis. As for the other test conditions,
the time history of the rotor blade deformation during the collective step pitch
input was measured using the time-resolved DIC technique.
119
were initiated during the 28th revolution and the rise time of the blade flap,
lag, and torsion was computed to be about 280 milliseconds (4.2 revolutions)
from Fig 4.18a to 4.18c.
Due to the step change, the resultant flapwise deflection at the blade
tip increased by 20 mm and reached nearly 115 mm in Fig. 4.18a. The step
response in chordwise degree of freedom due to the pitch angle change was also
observed, and the change was approximately 10 mm according to Fig. 4.18b.
Additionally, the local pitch angle at the blade tip was computed using Eq. 4.1
and is shown in Fig. 4.18c as a function of the number of revolution. First, the
local twist at the blade tip was settled at approximately 9◦ , then increased to
11.5◦ at maximum immediately after the step change, and stabilized at nearly
11◦ .
120
130 -10
Out-of-plane deformation, [mm]
110 -20
100 -25
90 -30
80 -35
20 25 30 35 40 45 50 55 60 65 70 20 25 30 35 40 45 50 55 60 65 70
N [rev] N [rev]
(a) Out-of-plane deformation (b) In-plane deformation
13
Sectional pitch angle, [deg.]
12
11
10
8
20 25 30 35 40 45 50 55 60 65 70
N [revs]
(c) Torsional deformation
121
large-scale rotor blade, respectively. Results of the modal parameters iden-
tification on the 0.46 m-diameter rotor blade are first presented in § 4.2.1,
and results of the OMA analysis on the large-scale rotor blade are discussed
in § 4.2.2.
Tip mass
Blade grip Leading edge
Trailing edge
202 mm
0.122 m
𝒚
𝒙
Figure 4.19: Rotor blades with fluorescent paint and random speckle pattern
122
scale rotor blade characteristics were the combined NExT-ERA approach and
CP algorithm. Each parameter of the small-scale rotor blade at different op-
erating conditions extracted by OMA processing will be separately addressed
in this section.
The primary objectives of this process were: (1) to examine how many
modes can be identified using the OMA algorithms based on deformation data
measured by the time-resolved DIC technique, (2) to compare the results iden-
tified by the OMA algorithms, with numerical simulations, and (3) to evaluate
the NExT-ERA and CP algorithm regarding accuracy, usability, and robust-
ness of each identification process.
123
and frequency spectrum of the local pitch angle displacement measured at
the blade tip for θ0 = 30◦ case, rotating at 1200 RPM, were first plotted in
Fig. 4.20a and 4.20b. The time history of the pitch angle variation at the tip
shows periodic motion, corresponding to the oscillation at 1/rev frequency.
It can be seen that in this operating condition the spectrum peaks are clear
enough to identify which peaks are associated with operating frequencies or
modal frequencies. The first peak right next to the 1/rev and second peak
between 2/rev and 3/rev harmonics correspond to the first and second modal
frequencies as shown in Figure 4.20b. Even if the excitation to the system is
random, there are always 1/rev harmonics on any rotating system.
Figure 4.21 shows the frequency spectrum of the blade tip out-of-
plane displacement in frequency domain (non-dimensionalized by the rotor
frequency), in log scale. This is basically another way of looking at the data
taken at 30◦ root pitch angle spun at 1200 RPM. It appears that the tip dis-
placement variation was a combination of random and periodic motion as also
observed in both Fig. 4.20b and 4.21. There were obvious peaks observed in
the frequency spectrum, corresponding to the harmonics of rotational speed
(i.e., 1200 RPM) as well as the rotating natural frequencies of the rotor blade.
Note that the presence of harmonic components might be an issue if the eigen-
frequencies of the test specimen are quite close (on the order of 0.1 Hz) to the
harmonic excitation, which is not the case in this study. If the frequency of
input harmonic excitation is close enough to a natural frequency of the struc-
ture to cause failure of the OMA-based modal identification, one would need
124
10
Figure 4.20: Pitch angle displacement at the tip for θ0 = 30◦ , Ω = 1200 RPM
125
special treatments to account for the harmonic components; see Mohanty and
Rixen [77].
101
Tip displacement [mm]
10-1
10-3
10-5
0 1 2 3 4 5 6 7 8
N/rev
Figure 4.21: Log-scale frequency spectrum of the out-of-plane displacement
measured at the blade tip for θ0 = 30◦ , Ω = 1200 RPM
Figure 4.22 and 4.23 show the modal frequencies identified by the OMA
algorithms as a function of rotor speed; these plots are known as fan plots.
In addition, the fan plot for θ0 = 30◦ shows the comparison between these
experimental results and predictions from the aeroelastic numerical analysis.
Note that these fan plots are presented based on measured data in air, i.e.,
aerodynamic damping is included. Predictions for θ0 = 0◦ were not available
because the numerical model could not converge for this root pitch angle. One
126
possible explanation for the divergence is instability of rotor blade motion with
zero-root pitch angle; during measurements, unstable behaviors of the rotor
blade were observed for all the rotational speeds. However, instability analysis
is beyond the scope of the current study.
Figure 4.26 and 4.27 shows the components of each degree of freedom
along the blade quarter-chord line for the first and second mode shapes, for
θ0 = 0◦ and θ0 = 30◦ root pitch angles at 1200 RPM, respectively. The
127
100
4/rev
1st mode CP
90 2nd mode CP
1st mode NExT/ERA
80
2nd mode NExT/ERA
3/rev
70
Frequency [Hz]
60
50
2/rev
40
30
1/rev
20
10
0
0 300 600 900 1200 1500
[RPM]
amplitudes of the mode shapes are normalized by the maximum value among
the flap, lead-lag, and torsional degrees of freedom, to identify the contribution
of each degree of freedom to each mode.
128
100
1st mode CP 4/rev
90 2nd mode CP
1st mode NExT/ERA
80 2nd mode NExT/ERA
70 1st mode prediction 3/rev
2nd mode prediction
Frequency [Hz]
60
50
2/rev
40
30
1/rev
20
10
0
0 300 600 900 1200 1500
[RPM]
Figure 4.23: Fan plot for θ0 = 30◦ , comparing natural frequencies extracted
with the CP and NExT-ERA algorithms, along with numerical predictions
The mode shapes extracted using the two different OMA algorithms
are compared to the numerically-predicted results for θ0 = 30◦ , operated at
129
Figure 4.24: The first mode shape of the rotor blade for θ0 = 30◦ at 1200 RPM
1200RPM. Figures 4.28 and 4.29 show the first and second mode shapes of all
three degrees of freedom, identified by CP, NExT-ERA, and the computational
prediction tool, respectively. For the first mode shown in Fig. 4.28, the lead-
lag and torsional mode shapes identified by the CP algorithm appear to be
in excellent agreement with those predicted by the numerical model; however,
there is a minor discrepancy between the flap mode shape obtained by CP
and the computational tool. On the other hand, the mode shapes obtained
by the NExT-ERA algorithm are quite different from those predicted by the
numerical model for all the degrees of freedom. For the second mode shown in
130
Figure 4.25: The second mode shape of the rotor blade for θ0 = 30◦ at 1200
RPM
Fig. 4.29, the CP-based mode shapes of all three degrees of freedom agree quite
well with the numerical results, as was observed for the first mode comparison.
However, the mode shapes obtained by the NExT-ERA process are not well
correlated with the numerical prediction, except for the lead-lag degree of
freedom.
131
1 1st mode
2nd mode
Flap
0
-1
0 0.25 0.5 0.75 1
1
Lead-lag
-1
0 0.25 0.5 0.75 1
1
Torsion
-1
0 0.25 0.5 0.75 1
Radial location
Figure 4.26: The first and second mode shapes along the quarter-chord axis,
for θ0 = 0◦
the most common tools for the quantitative comparison of two modal vectors
and is defined in terms of the normalized scalar product of the two vectors
{ϕ1 } and {ϕ2 } as:
The MAC is calculated for two types of comparisons: (1) between the
NExT-ERA and the numerical model, and (2) between the CP and the numer-
ical model. Table 4.2 summarizes the MAC values calculated from the mode
132
1
1st mode
Flap
2nd mode
0
-1
0 0.25 0.5 0.75 1
1
Lead-lag
-1
0 0.25 0.5 0.75 1
1
Torsion
-1
0 0.25 0.5 0.75 1
Radial location
Figure 4.27: The first and second mode shapes along the quarter-chord axis,
for θ0 = 30◦
shapes identified by the OMA algorithms and the mode shapes predicted by
the numerical model. These mode shapes were obtained from the operating
condition where the blade root pitch angle was 30◦ . Based on the MAC val-
ues, the mode shapes extracted by the CP algorithm correlated quite well
with the predicted mode shapes (MAC values larger than 0.9 indicate con-
sistent correspondence between two modal vectors). However, the low MAC
values between mode shapes extracted by NExT-ERA and predicted mode
shapes indicate poor resemblance of these two mode shapes, except for the
133
CP
1 NExT/ERA
Prediction
Flap
0
-1
0 0.25 0.5 0.75 1
1
Lead-lag
-1
0 0.25 0.5 0.75 1
1
Torsion
-1
0 0.25 0.5 0.75 1
Radial location
Figure 4.28: Comparison of the first mode shapes estimated by the CP and
NExT-ERA as well as numerical prediction for θ0 = 30◦ at 1200 RPM
1st and 2nd lag modes. Note that this study uses Partial Modal Assurance
Criterion (PMAC) where only a selected degree-of-freedom from a complete
modal vector is used to compute the consistency of identified modes. In this
way, linearity of a mode shape of each degree-of-freedom can be evaluated.
134
Table 4.2: Summary of the MAC values calculated from the measured and
predicted mode shapes, obtained at θ0 = 30◦
1
Flap
-1
0 0.25 0.5 0.75 1
1
Lead-lag
-1
0 0.25 0.5 0.75 1
1
Torsion
-1
0 0.25 0.5 0.75 1
Radial location
Figure 4.29: Comparison of the second mode shapes estimated by the CP and
NExT-ERA as well as numerical prediction for θ0 = 30◦ at 1200 RPM
135
quency domain that the real mode could not be properly extracted. Another
possible reason is that the mode shape extraction by means of NExT-ERA
requires a number of parameters and post-processing steps that must be cho-
sen by an analyst, such as applying polynomial fits or removing outliers; these
post-processing manipulations could result in poor mode shape identification.
A block diagram in Fig. 2.5 shown in § 2.2 summarizes the advantages and
disadvantages of these two OMA algorithms. In contrast, the mode shapes
identified by CP algorithm without any post-processing showed good correla-
tions with the numerical prediction results.
Modal damping has been reported to be the most difficult value among
the modal parameters to be estimated with OMA approaches [79, 80]. The
present study calculated the damping ratios using a curve fit based on a single-
degree-of-freedom impulse response to the auto-correlation function of the sep-
arated source signal. Table 4.3 summarizes the estimated modal damping of
the rotor blade at a root pitch angle of θ0 = 0◦ for different rotational speeds.
For all the cases, the values of modal damping were too small if considering
the influence of aerodynamic damping on the rotor blade motion. Table 4.3
also shows the estimated damping values by NExT-ERA [81]. All the damping
values from the NExT-ERA analysis were also less than 3% as is the case of
the CP analysis.
136
of harmonic excitations on a target structure, as discussed by Mohanty and
Rixen [77]. If the harmonic frequency is very close to the modal natural
frequency, the OMA algorithm may select this harmonic component of the
structural response as one of the eigenvalues. In the current study, however,
each mode was selected after excluding the sources with frequencies that were
integer multiples of once-per-revolution frequency, as well as the corresponding
mode shape. Thus, the identified mode should not be associated with the
harmonic excitations, and the cause for the low modal damping is still unclear.
These results imply that the modal damping identification process must be
further refined to improve accuracy.
Table 4.3: Summary of modal damping ratios for different rotational speeds
at 0◦ root pitch angle
In summary, there are three key conclusions from this modal identifi-
cation processing based on the small-scale rotor blade deformation measure-
ment: (1) The first two modes were identified including the natural frequen-
cies, mode shapes, and modal damping ratios. (2) Modal parameters identified
by CP algorithm showed good agreement with the results provided from the
numerical model. (3) Identification process with CP algorithm was more ef-
ficient and user-friendly, as compared to that with the combined NExT-ERA
algorithm. Only the first two modes were identified; this could be because
137
higher-frequency deformations tend to be small and might approach the noise
floor of the DIC technique itself. Increasing the signal-to-noise ratio of the
DIC measurements is essential to enable identification of higher modes. The
conclusions (2) and (3) are somewhat related, that is, the complicated manual
steps with NExT-ERA algorithm resulted in poor quality of modal parameter
identification.
138
compares the results of each test case with the numerical model.
The tip displacement in time and frequency domain, shown in Figs. 4.31a
and 4.31b, also exhibited a trend similar to that for the rotor thrust response.
The harmonic components of the rotating frequency were the significant con-
139
25 1
10
20
0
15 10
10
10-1
Thrust [N]
Thrust [N]
5
0 10-2
-5
-10 10-3
-15
10-4
-20
-25
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Time [s] N/rev
Figure 4.30: Thrust measured on the single-bladed single rotor at zero mean
thrust
Now, the CP algorithm was used to extract the rotating natural fre-
quencies and mode shapes of the rotor blade. The fan plot in Fig. 4.32 shows
the natural frequencies identified by the CP algorithm (at the nominal rota-
tional frequency of 900 RPM) along with numerical predictions. Note that
the natural frequencies at 300 RPM, 600 RPM, and 1200 RPM were obtained
in previous experiments by Cameron et al. [35], who used a different modal
identification algorithm on the same experimental setup. It can be seen that
the extracted natural frequencies agree well with the numerical predictions as
well as with previous measurements over different rotational frequencies.
140
6 2
10
5
4 101
Tip displacement [mm]
Figure 4.33 to 4.35 shows the first flap, the first lag, and the second
flap rotating mode shapes compared to predicted results, respectively. There
was good agreement between the numerically predicted characteristics and the
results obtained by the CP algorithm for all three mode shapes. The MAC
was also used for the results of the large-scale rotor blade modal parameter
identification, to quantify the differences between the measured and predicted
mode shapes; see Eq. 4.2. Table 4.4 summarizes the MAC values of the mode
shapes as well as the rotating frame natural frequencies for both identified and
predicted parameters. This excellent agreement between the CP-derived and
numerically-obtained modal parameters indicates that the DIC deformation
measurement in conjunction with the CP algorithm can also be used to identify
the rotating frame natural frequencies and modes shapes of the large-scale
rotor blade.
141
200
10/rev 9/rev
Measurement
180 Cameron et al. [35] 8/rev
Prediction
160 3rd flap
7/rev
140
6/rev
Frequency, Hz
120
5/rev
100
4/rev
80
2nd flap 3/rev
60
1st lag
40
2/rev
20 1/rev
1st flap
0
0 25 50 75 100 125 150
Rotational Speed, % nominal
Figure 4.32: Fan plot with measured natural frequencies (100% = 900 RPM),
compared with numerical predictions
Table 4.4: Summary of the rotating natural frequencies and MAC values cal-
culated from the measured and predicted mode shapes.
142
1
Prediction
0.8
Modal displacement
Measurements
0.6
0.4
0.2
-0.2
0 0.2 0.4 0.6 0.8 1
Radial station r/R
Figure 4.33: Measured rotating mode shapes of the first flap mode at nominal
rotational speed (900 RPM) compared with numerical predictions
0.8
Modal displacement
0.6
0.4
0.2
-0.2
0 0.2 0.4 0.6 0.8 1
Radial station r/R
Figure 4.34: Measured rotating mode shapes of the first lag mode at nominal
rotational speed (900 RPM) compared with numerical predictions
the first lag, and the second flap bending mode. This was the motivation of
performing modal parameter identification based on another data set, which
was the blade response to the step function excitation.
143
1
-0.5
-1
0 0.2 0.4 0.6 0.8 1
Radial station r/R
Figure 4.35: Measured rotating mode shapes of the second flap mode at nom-
inal rotational speed (900 RPM) compared with numerical predictions
Figure 4.37 shows the frequency spectrum of the three different time
periods up to 120 Hz (8/rev). The overall trend commonly observed in the
spectra was the dominant periodic peaks corresponding to the integer multi-
ples of rotational frequency. There were some smaller peaks between those
multiples of rotational frequency, however, it was challenging to distinguish
144
130
110
(i)
100
90
80
20 25 30 35 40 45 50 55 60 65 70
N [rev]
Figure 4.36: Time history of the out-of-plane deformation measured at the
trailing-edge point of the rotor blade tip
which peaks were associated with fictitious noise or the actual rotating frame
natural frequencies. Since the time period (iii) had the largest number of sam-
ples, the frequency spectrum (iii) in Fig. 4.37 has a higher resolution in the
frequency domain than in the spectra in cases (i) and (ii).
Noise level of each time period seems to be different in Fig. 4.37, how-
ever, that is due to the number of sample sizes used for the Fourier Transform.
Since the time period (iii) had a largest amount of samples, the frequency
spectrum (iii) appears to contain more noise than the other two sets of data.
145
102
( i ) Before step pitch input
0
10
10-2
10-4
0 10 20 30 40 50 60 70 80 90 100 110 120
102
z-displacement [mm]
100
10-2
-4
10
0 10 20 30 40 50 60 70 80 90 100 110 120
102
( iii ) After the response settled down
to the steady-state condition
100
10-2
-4
10
0 10 20 30 40 50 60 70 80 90 100 110 120
Frequency [Hz]
main, called spectrogram, was used and applied to the same data set. A
spectrogram is defined as a visual representation of frequency spectrum of a
time-dependent signal and can estimate the time-localized frequency contents
146
using the short-time Fourier transform. The signal is typically divided into
several segments for windowing, so that one can see how frequency compo-
nents in the signal change in time. Figure 4.38 shows the spectrogram of
the same signal analyzed in Fig. 4.37. The parameters used for the spec-
trum are summarized in Table 4.5. Note that the sampling rate of the DIC
measurement was 15 Hz (rotational speed) × 32 azimuthal resolution = 480
Hz. At regions where the step change occurred (around 2.5 s) in the spectro-
gram, wide-bandwidth vibration was propagated in the power contour and it
immediately died out after the step change. This frequency content can be
interpreted as a proof of the successful excitation to the blade structure. On
the other hand, the periodic harmonic components attributed to the integer
multiples of the rotational frequency (15 Hz) remained constant over the entire
measurement period.
147
120
20
105
Power/frequency (dB/Hz)
90 0
Frequency [Hz]
75
-20
60
45 -40
30
-60
15
0 -80
0.5 1.5 2.5 3.5 4.5 5.5 6.5 7.5
Time [s]
Figure 4.38: Spectrogram of the flap response measured at the trailing edge
point of the blade tip
flap modes, which was less than the number of modes identified during the
Test campaign 2. Additionally, the MAC for the second flap mode identified
during Test campaign 3 was low, indicating that the second mode shape was
not well correlated with either the numerical prediction or the mode shape ex-
tracted during Test campaign 2. The original intention of this test condition
was to excite the rotor blade structure by the rapid collective pitch change
(step function as an excitation source), so that the resultant response might
contain a large number of structural modes at higher amplitudes as compared
to the response to the random, turbulence-driven, ambient aerodynamic load-
148
ing achieved during Test campaign 2.
Table 4.6: Summary of the experimentally and numerically obtained modal
frequencies and MAC values calculated from the measured and predicted mode
shapes for Test campaign 2 and 3
There could be several reasons why the OMA processing of the data
measured in this test condition failed to extract the higher structural modes:
First, the sampling rate was not large enough to capture the step-induced os-
cillation whose decay rates were so large as shown in the spectrogram (see in
Fig. 4.38). In other words, the decay of the transient response to the collective
step change was to rapid to capture with the current sampling rate (480 Hz).
Second, the magnitude of the step change (corresponding to 2◦ increase) was
not strong enough to excite the higher modes of the blade structure during
rotation. These results of the modal parameter identification must be investi-
gated further in future work.
149
In § 4.3.1, measured thrust and pitch link force of the CCR lower rotor are
compared with the numerical model provided from Ref. [57], and measured
rotor loads are compared with results of loads estimation process for the three
different operating conditions (a constant collective input, a periodic 1/rev,
and a collective step input) in § 4.3.2.
150
25
Measurement
CAMRAD II
20
Thrust [N] 15
10
0
1 2 3 4 5 6 7 8 9
n/rev
(a) Frequency domain
140
Experiment
Prediction
120
Thrust [N]
100
80
60
0 90 180 270 360
[deg]
(b) Azimuthal domain
Figure 4.39: Reconstructed (filtered) thrust of the CCR lower rotor, compared
to predictions at CT /σ = 0.09
increased the effective angle of attack of the lower rotor blade, resulting in the
increase in rotor thrust, and vice versa. While the magnitude of the predic-
151
tions correlated well with the measurements, there was some discrepancy in
the phase. For the first blade passage, peaks occurred at about ψ = 15◦ and
ψ = 30◦ for prediction and measurement, respectively, whereas for the second
blade passage, peaks occurred at about ψ = 195◦ and ψ = 210◦ for prediction
and measurement, respectively. Recall that the blades crossed each other at
ψ = 0◦ and ψ = 180◦ .
Overall, the vibratory thrust was approximately 10% of the mean thrust,
which is consistent with previous studies by Cameron et al. [35, 82] where the
vibratory thrust was 11% of the mean thrust in hover. More significantly, the
vibratory pitch link load was found to be nearly 30 % of the mean pitch link
load, which can have a profound impact on the design of the rotor control
system. Thus, for the large-scale rotor hover test rig, it is concluded that the
three major components of measured quantities for the rotor loads estimation
methodology proposed in the dissertation, i.e., blade deformation, blade modal
properties, and rotor loads, were captured well by the numerical prediction.
152
20
Measurement
Prediction
10
1 2 3 4 5 6 7 8 9
n/rev
(a) Frequency domain
10
Experiment
Prediction
0
Pitch link load [N]
-10
-20
-30
-40
-50
0 90 180 270 360
[deg]
(b) Time domain
Figure 4.40: Reconstructed (filtered) pitch link load of the CCR lower rotor,
compared to predictions at CT /σ = 0.09
153
4.3.2 Two-bladed single rotor loads measurement and estimation
The first operating condition tested on the two-bladed single rotor was
a constant collective pitch input at a blade loading CT /σ of 0.125. Measured
quantities include: hub loads, pitch link loads, and blade root pitch angles.
154
and phase-averaged over 250 revolutions, is shown in time (azimuthal) and
frequency (non-dimensionalized by the rotational frequency, 15 Hz) domain
in Fig. 4.41a and 4.41b, respectively. Note that all the loads variations in
this section are measured in the rotating frame. The frequency spectrum
clearly represents the general relationship between the mean and vibratory
components of hub loads, that is, vibratory loads of a single rotor in hover
are much smaller than the steady mean component of rotor loads. In this test
case, there was little vibration in hub loads (less than approximately 1% of
the mean thrust), since a constant collective pitch input was applied to the
rotor.
A similar trend can be seen in the rotor torque; the variation in time
and frequency domain are shown in Fig. 4.42a and 4.42b, respectively, with
an oscillatory behavior at the 3/rev harmonic visible in both domains. The
magnitude ratio of the 3/rev component to the mean torque is approximately
8.5%, which is relatively large compared to a typical vibration level in hover.
The 3/rev variation can also be seen in more profound manner in the rotating
frame rolling moment, as shown in Fig. 4.43. One possible reason for this large
3/rev vibration is the proximity of the 1st lag mode frequency to the 3/rev
blade crossing (see the fan plot shown in Fig. 4.32).
155
260
250
240
240
210
180
Thrust [N]
150
120
90
60
30
0
Mean 1 2 3 4 5 6 7 8 9
N/rev
(b) Frequency domain
156
25
Torque
22.5
Torque, [Nm] 20
17.5
15
12.5
10
0 60 120 180 240 300 360
Azimuth, [deg]
(a) Time domain
20
17.5
15
Torque [Nm]
12.5
10
7.5
5
2.5
0
Mean 1 2 3 4 5 6 7 8 9
N/rev
(b) Frequency domain
157
7
6 My
5
4
My, [Nm] 3
2
1
0
-1
-2
-3
-4
0 60 120 180 240 300 360
Azimuth, [deg]
(a) Time domain
2.5
2
My [Nm]
1.5
0.5
0
Mean 1 2 3 4 5 6 7 8 9
N/rev
(b) Frequency domain
158
shows the azimuthal variation and its frequency spectrum of pitch link loads
for both blades. In Fig. 4.44a, there was a clear phase difference of nearly
180◦ between the pitch link (blade) 1 and 2, implying that there might exist a
slight inclination of the swashplate that caused the out-of-phase pitch moment
variation for the two blades. Most of the vibratory components were contained
in the first three harmonics for both blades, with a 1/rev component as the
largest contributor to the pitch link force variation as shown in Fig. 4.44b.
A possible source of the slight magnitude difference of the 1/rev harmonic
component was the presence of viscous damping in the bearing mechanism
(grease lubrication) for rotor feathering motion.
Blade root pitch angles for both rotor blades were measured using the
Anisotropic Magneto-Resistive (AMR) sensors installed on the blade grips
(see § 3.1.2 for more details). The real-time measurement of root pitch an-
gles enables trimming the rotor at a target operating condition during hover
testing.
Figure 4.45 shows the phase-averaged root pitch angle of both blades as
a function of azimuthal position over one revolution. The overall trend of the
blade pitch 1 corresponds to the trim target of this test condition 1 (a steady,
mean collective pitch θ0 = 9◦ ). However, there was a large fluctuation in the
blade pitch 2 measurement over the range of 60 to 180 deg, possibly due to
the free play of the corresponding pitch link or needle bearings that hold the
blade grip. It is interesting to note that the mean component of root blade
pitch 1 (about 9.27◦ from Fig. 4.45a) was nearly the same as the torsional
159
25
Pitch link 1
20 Pitch link 2
9
Pitch link 1
8 Pitch link 2
7
Pitch link force [N]
6
5
4
3
2
1
0
Mean 1 2 3 4 5 6 7 8 9
N/rev
(b) Frequency domain
160
displacement over the blade span calculated using the DIC results shown in
Fig. 4.15c. This result implies that the rotor blade used for this hover testing
experienced little elastic twist at this blade loading condition, and the DIC
measurement agrees well with the results measured by the AMR pitch angle
sensor.
The frequency spectrum of the pitch angle variations is shown in Fig 4.45b,
revealing that the observed fluctuation on the pitch angle measurement 2 was
associated with the 1/rev frequency component. The normalized amplitude of
the vibratory component was less than 2% of the mean collective pitch, thus
it is fair to say that the intended rotor trim control was achieved for the test
case A.
Now the results of rotor loads estimation for the test condition 1 are
discussed. Rotor loads include the spanwise lift distribution along the quarter-
chord blade axis, and the hub loads (thrust) obtained by numerical integration
of the lift distribution in spanwise direction. For rotor load estimation, the
first and second flap modes identified from Test campaign 2 (see § 4.2.2) were
used.
161
10
Pitch angle 1
Pitch angle 2
9.5
Pitch angle, [deg]
9
8.5
8
0 60 120 180 240 300 360
Azimuth, [deg]
(a) Time domain
11
10 Pitch angle 1
Pitch angle 2
9
Pitch angle [degree]
8
7
6
5
4
3
2
1
0
Mean 1 2 3 4 5 6 7 8 9
N/rev
(b) Frequency domain
162
imen for the DIC deformation measurement). As expected from the numerical
experiment, the sectional load monotonically increased up to the blade tip
without capturing the decrease (tip loss) due to the influence of tip vortices
(see Fig. 2.12). The lack of participating modes (only two modes) was the
primary cause for the poor estimation at the blade tip. On the other hand,
the integrated hub load, shown in Fig. 4.46b appeared to be well correlated
to the directly-measured thrust, especially for the steady, mean component of
the thrust variation. Although the vibratory loads seemed to have a slight
discrepancy, the mean thrust difference was observed within 5%.
For the test condition 2, the rotor was trimmed at a constant collective
plus a lateral cyclic pitch angle. The collective pitch θ0 was set to be 9◦ so
that the mean blade loading remained the same as for the test condition 1.
The periodic 1/rev lateral pitch θ1C was set to be ±2◦ . Measured quantities
are the same as the previous test condition, including hub loads, pitch link
loads, and root pitch angle variation.
163
(a) Estimated lift distribution over one revolution
260
Measured
250 Estimated
240
230
Thrust, [N]
220
210
200
190
180
170
0 60 120 180 240 300 360
Azimuth, [deg]
(b) Comparison between the estimated and measured
rotor hub loads
Figure 4.46: Estimated rotor loads for the steady collective pitch input
164
and 4.48b show the torque variation, revealing the large 3/rev vibratory com-
ponent (relative amplitude of 11%) that was also observed for the test condi-
tion 1. The presence of the 3/rev harmonics is consistent with both the test
condition 1 and 2, indicating that the source of this vibration might depend on
the rotor blade modal parameters, as discussed in the previous test condition.
The primary difference between the test condition 1 and 2 was found
in the rotating frame rolling moment My as shown in Fig. 4.49. Due to the
asymmetric lift distribution over the rotor disk, the rotor experienced the large
variation in the rolling moment over one revolution. The clean sinusoidal trend
at the 1/rev frequency corresponds to the lateral cyclic pitch input (θ1C = 2◦ )
to the rotor control. As can be seen in Fig 4.16, the tilting motion of the
rotor disk was consistently associated with the trend of the rolling moment,
meaning that the tilting direction of the rotor disk appropriately reflected the
sign change of the rolling moment.
The pitch link force variation for both rotor blades is shown in Fig 4.50.
165
270
260
250
Thrust, [N] 240
230
220
210
200
190
0 60 120 180 240 300 360
Azimuth, [deg]
(a) Time domain
250
225
200
175
Thrust [N]
150
125
100
75
50
25
0
Mean 1 2 3 4 5 6 7 8 9
N/rev
(b) Frequency domain
166
25
Torque
22.5
Torque, [Nm] 20
17.5
15
12.5
10
0 60 120 180 240 300 360
Azimuth, [deg]
(a) Time domain
20
17.5
15
Torque [Nm]
12.5
10
7.5
5
2.5
0
Mean 1 2 3 4 5 6 7 8 9
N/rev
(b) Frequency domain
167
30
25 My
20
15
My, [Nm] 10
5
0
-5
-10
-15
-20
-25
-30
0 60 120 180 240 300 360
Azimuth, [deg]
(a) Time domain
25
22.5
20
17.5
My [Nm]
15
12.5
10
7.5
5
2.5
0
Mean 1 2 3 4 5 6 7 8 9
N/rev
(b) Frequency domain
168
Figure 4.50a presents a significant 1/rev sinusoidal trend for both pitch links
and the clear azimuthal phase offset of approximately 180◦ between the two
rotor blades. The 1/rev vibratory component directly corresponds to the vari-
ation of the blade pitching moment due to the swashplate inclination corre-
sponding to the lateral cyclic pitch. On the other hand, Fig. 4.50b compares
the vibratory components of the two pitch link loads in frequency domain. The
magnitudes of the 1/rev component for the pitch link 1 and 2 were respectively
19.4 N and 15.7 N, which were the dominant component consistently for both
linkages; however, the values were shown to have a discrepancy of approxi-
mately 5 N. The mean, steady component seems to have a difference on the
same order of magnitude as well. The constant offset between two linkages
can also be seen in Fig. 4.44 for the test case condition 1, indicating that
there must be some common source of this difference, regardless of rotor trim
condition, such as free-play in rod ends or grease lubrication in the feathering
bearings.
The spanwise distribution and integrated hub load for the test condition
2 were estimated and compared to the measured data. For the test condition
2, the rotor was trimmed at the steady collective pitch θ0 = 9◦ + the lateral
169
50
Pitch link 1
40 Pitch link 2
20
Pitch link 1
17.5 Pitch link 2
Pitch link force [N]
15
12.5
10
7.5
5
2.5
0
Mean 1 2 3 4 5 6 7 8 9
N/rev
(b) Frequency domain
170
12
Pitch angle 1
Pitch angle 2
11
6
0 60 120 180 240 300 360
Azimuth, [deg]
(a) Time domain
11
10 Pitch angle 1
Pitch angle 2
9
Pitch angle [degree]
8
7
6
5
4
3
2
1
0
Mean 1 2 3 4 5 6 7 8 9
N/rev
(b) Frequency domain
171
cosine cyclic input θ1C = 2◦ . Figure 4.52a shows the sectional lift distribution
based on the phase-averaged blade flap bending variation over the entire rotor
disk. The contour clearly represents the aerodynamic force variation at once-
per-revolution frequency with the consistent phase, however, the lift tip loss
was not captured for this case, either. Figure 4.52b compares the integrated
hub loads for the lateral cyclic pitch input, revealing the pronounced 2/rev
vibratory content in the estimated thrust variation. The 2/rev component of
the estimated thrust agreed well with that of the measured thrust in terms of
both magnitude and phase. There was a mean thrust discrepancy of nearly
5%, which was also observed in Fig. 4.46 for the test condition 1.
172
(a) Estimated lift distribution over one revolution
260
Measured
250 Estimated
240
Thrust, [N]
230
220
210
200
190
180
0 60 120 180 240 300 360
Azimuth, [deg]
(b) Comparison between the estimated and measured
rotor hub loads
Figure 4.52: Estimated rotor loads for the lateral cyclic pitch input (Test
condition 2)
173
Table 4.8: Summary of cutoff frequencies for low-pass filtering process for Test
campaign 3
To clearly show the pitch variation, the low-pass filter was applied with a
3 dB corner frequency of 13 Hz. In Fig. 4.53, the blade root pitch angle
started increasing from approximately 9.5 degree during the 28th revolution,
and reached the steady state value of about 11.5 degree after completing the
35th revolution.
174
13
12
Pitch angle, [deg]
11
10
Pitch angle 1
9 Pitch angle 2
8
20 25 30 35 40 45 50 55 60 65 70
N [revs]
Figure 4.53: Blade root pitch angle as a function of rotor revolutions during
the step pitch input operation
Now the rotor thrust and torque responses to the step pitch increase
are examined. The time history of the low-pass filtered thrust and torque
are shown in Fig. 4.55 and 4.56, respectively. The increase in thrust started
during the 28th revolution and maintained the rate of change up until the
30th revolution, then the slope changed to a lower value. This slope transition
was consistent to the variation of other quantities including the pitch angle
variation and blade tip deflection. The rotor mean thrust was first at about
230 N at θ0 = 9◦ , then increased by approximately 40 N due to the 2◦ collective
175
920
900
890
880
870
20 25 30 35 40 45 50 55 60 65 70
N [revs]
Figure 4.54: Rotational speed of the rotor as a function of rotor revolutions
On the other hand, the trend of the rotor torque was slightly different
from that of the rotor thrust. Figure 4.56 shows the torque variation as a
function of time, low-pass filtered at the 3dB cutoff frequency of 25 Hz. The
rate of change of the rotor torque variation appeared to be monotonic unlike
the trend of the thrust variation, and the torque reached its peak immediately
after passing the 35th revolution and died out at around the 45th revolution.
A possible reason for this trend difference is that there was a close correlation
between the torque and the rotational speed variation and these two quantities
had a pronounced influence on each other. The low-frequency rotor speed
176
300
280
Thrust, [N]
260
240
220
200
20 25 30 35 40 45 50 55 60 65 70
N [revs]
Figure 4.55: Rotor thrust response to the step pitch change as a function of
rotor revolutions
ringing hence appeared more in the torque response, than in the responses of
other measured quantities.
177
30
27.5
25
Torque, [Nm]
22.5
20
17.5
15
20 25 30 35 40 45 50 55 60 65 70
N [revs]
Figure 4.56: Rotor torque response to the step pitch change as a function of
rotor revolutions
increase the blade pitch angle. Then, the pitch link force started increasing
after passing the 30th revolution and overshot at about the 35th revolution,
because of the increase in blade pitch moment in nose-down direction (tension).
Hence, the trends observed in Fig. 4.57 adequately represent the blade pitch
moment (aerodynamic and inertial) attributed to the dynamic collective step
change.
178
15
Pitch link 1
Pitch link force, [N] 12.5 Pitch link 2
10
7.5
5
2.5
0
-2.5
-5
20 25 30 35 40 45 50 55 60 65 70
N [revs]
Figure 4.57: Pitch link force response to the step pitch change as a function
of rotor revolutions
The time history of the integrated hub load for the test condition 3
179
was estimated and compared to the measured thrust variation, as shown in
Fig. 4.59. The resultant increase in rotor thrust due to the collective step
change was found to be consistently 40 N in both the measured and estimated
profile; however, there was a constant magnitude offset of approximately 15
N and a slight time lag of step response between the measured and estimated
thrust. The former concern was consistently observed for all three operating
conditions (the constant collective pitch input, the 1/rev periodic pitch input,
and this collective step pitch input) on the same order of magnitude (10-15
N), which could possibly be explained by the area difference under the original
10
7.5
5
Inertial load [N]
2.5
0
-2.5
-5
-7.5
-10
20 25 30 35 40 45 50 55 60 65 70
N [revs]
Figure 4.58: The estimated inertial load integrated at the hub as a function of
the number of revolutions for the collective step pitch change (Test condition
3)
180
and estimated spanwise lift curve; (see Fig. 2.11). The latter concern could be
because of the nature of this load estimation methodology, i.e., the estimated
loads were purely based on rotor blade structural deformation measured by
the non-contact optical technique (DIC), whereas the hub loads measured by
the load cell contained a large number of experiment-specific factors, such as
linkage mechanism, free-play and lubrication in bearings, non-ideal aerody-
namic interaction with the experimental setup, or the recirculation effect in
the hover test chamber. These factors might cause the constant magnitude off-
set and time lag between the measured and estimated thrust response; further
investigation must be performed as future work.
181
tip. Nevertheless, it must be emphasized that this rotor load estimation was
not made by complicated, troublesome on-blade sensors but achieved by pro-
cessing the sequence of digital images taken during the hover tests, and the
only preparation on the blade structure was the orange fluorescent paint with
random speckle patterns using a spray lacquer and a black-ink marker.
300
280
260
Thrust, [N]
240
220
Measured thrust
Estimated thrust
200
180
20 25 30 35 40 45 50 55 60 65 70
N [revs]
Figure 4.59: Comparison between the estimated and measured rotor hub loads
for the collective step pitch change (Test condition 3)
182
Chapter 5
183
DIC technique at four different root pitch angles (0◦ -30◦ ) and rotational speeds
(600-1500 RPM). The time-resolved DIC was achieved by (i) taking a sequence
of digital images over the whole rotor disk by stereoscopic high-speed cameras,
(ii) illuminating the surface of the rotor blade painted with fluorescent orange
and stochastic speckle pattern by laser light, and (iii) triggering image ac-
quisition and laser strobe timing with a downsampled (16/rev) digital pulse
train generated from an optical encoder. Blade deformation measurements
consisted of the flap, lead-lag, and torsional deformation over 400 revolutions
at approximately 900 measurement channels over the entire rotor span at 16
evenly-spaced azimuthal locations. The measurements revealed that the time-
resolved DIC technique developed in this dissertation can be used to measure
the rotating blade deformation on the order of a few millimeters.
184
times larger than the tip deflection of the small-scale rotor. The time-resolved
DIC successfully demonstrated its capability of measuring the deformation
time history of a rotating blade at different measurement scales.
185
fied using these two OMA algorithms. To quantify the differences between the
measured and predicted mode shapes, Modal Assurance Criterion (MAC) was
used as a measure of agreement between two vectors. The modal frequencies
and mode shapes identified by the CP algorithm agreed well with numerical
predictions, however, there was some discrepancy between mode shapes iden-
tified by NExT-ERA and numerical prediction. All the modal damping values
extracted by the NExT-ERA analysis were less than 3% as was the case of the
CP analysis. While the NExT-ERA required a number of steps and choices of
parameters based on experience or trial-and-error, the CP algorithm was quite
simple to implement and did not require any tuning of parameters. Thus, the
CP algorithm was selected as a primary OMA algorithm for modal identifica-
tion of the large-scale rotor blade.
For the large-scale rotor blade modal identification, two different types
of excitation to the rotor blade were tested: (i) random excitation at zero mean
thrust and (ii) collective step input excitation. Overall, the modal extraction
algorithm based on the random response successfully identified the first flap,
first lag, and second flap frequencies and mode shapes. Using the MAC, it
was demonstrated that there was excellent agreement between mode shapes
extracted by the CP algorithm and the numerical model. On the other hand,
the CP analysis extracted only the first two flap modes from the response to
the collective step change. The spectrogram of the same data set showed that
wide-bandwidth vibration was propagated when the step change occurred,
however, this response immediately died out after the step. Hence, the first
186
three modes identified in the test case (i) (random excitation) were used for
rotor loads estimation.
Finally, rotor loads were estimated based on the measured blade de-
formation and modal parameters identified by the CP algorithm. Rotor loads
estimation includes the spanwise lift distribution along the blade quarter-chord
axis and hub loads (thrust) obtained by numerical integration of the lift distri-
bution in the spanwise direction. Note that the modal parameters for the first
and second flap modes were used for rotor loads estimation because the lag
mode does not contribute to the out-of-plane lift force reconstruction. Data
sets taken at (1) a constant pitch, (2) a 1/rev periodic pitch, and (3) a col-
lective step pitch on the large-scale, two-bladed rotor were used. For all the
test conditions, the estimated sectional lift distribution failed to capture the
typical trend of lift loss in regions of the blade tip induced by the trailed tip
187
vortices, due to the limited number of participating modes for load reconstruc-
tion. However, the integrated thrust agrees well with the directly-measured
value for the three test cases, possibly due to the fact that the areas under
the estimated and original lift curve are close to each other even if only a
few modes are used for the estimation. In conclusion, the rotor loads estima-
tion methodology is a powerful tool to obtain a rough estimate of the rotor
hub loads without using on-blade instrumentation. To accurately estimate the
spanwise lift distribution, a larger number of modes must be included.
188
formation measurement sometimes requires an analyst to deal with an unnec-
essarily large number of data points in space over the entire blade span. To
effectively utilize the significant amount of measured information, spatial av-
eraging over the neighborhood is suggested; instead of tracking one particular
measurement point that corresponds to a single interrogation window, aver-
aging 3D deformations over the neighborhood points in both spanwise and
chordwise directions might effectively eliminate noise and outliers in raw data.
189
motion tends to die out rapidly due to the high damping. Thus, the estimation
of damping remains the most difficult part of modal analysis in helicopter rotor
blade dynamics.
Regarding the rotor loads estimation, the tip lift loss induced by the
190
trailed tip vortices was not quite captured due to the lack of participating
modes in the estimation process. However, the tip loss might be relatively
easily corrected by introducing an additional well-known shape function, such
as the Prandtl tip loss function, to the estimation process, instead of struggling
to identify the higher-order structural modes. The tip loss correction could
not only help one to accurately reconstruct the spanwise lift distribution, but
also address the consistently-observed 5% mean value offset in the estimated
hub loads estimation.
191
copter rotor loads will provide a foundation to develop methods to mitigate
vibration in future helicopter design.
192
Bibliography
[1] An image of a static display of the CH-53 Sea Stallion. Downloaded from
https://www.military.com/equipment/ch-53d-sea-stallion.
[2] An image of the CH-53 Sea Stallion taken under operation. Downloaded
from https://commons.wikimedia.org/wiki/File:CH-53 Sea Stallion -
Side View (Balikatan 2016).JPG.
193
[7] G. Wang, A. Abhishek, I. Chopra, N. Phan, and D. Liebschutz. Ro-
tor Loads Prediction Using Estimated Modal Participation from Sensors.
39th European Rotorcraft Forum Proceedings, Paris, France, Sept. 9-11,
2010.
[10] G. C. Cameron, D. Uehara, and J. Sirohi. Transient Hub Loads and Blade
Deformation of a Mach-Scale Coaxial Rotor in Hover. 56th AIAA/ASCE/
AHS/ASC Structures, Structural Dynamics, and Materials Conference,
Kissimmee, Florida, Jan. 5-9, 2015.
194
[13] J. P. Sevenhuijsen. The photonical, pure grid method. Optics and Lasers
in Engineering, 18(3), pp. 173–194, 1993.
195
[19] M. K. Sekula. The Development and Hover Test Application of a Pro-
jection Moiré Interferometry Blade Displacement Measurement System.
Proceedings of the 68th Annual Forum of the American Helicopter Society,
Fort Worth, TX, May 1-3, 2012.
196
[25] T. Lundstrom, J. Baqersad, and C. Niezrecki. Monitoring the Dynam-
ics of a Helicopter Main Rotor with High-Speed Stereophotogrammetry.
Experimental Techniques, 40(3), pp. 907–919, 2016.
197
pp. 13–20, 1997.
[37] S. R. Ibrahim and R. S. Pappa. Large Modal Survey Testing Using the
Ibrahim Time Domain Identification Technique. Journal of Spacecraft
and Rocket, 19(5), pp. 459–465, 1982.
198
[38] J. N. Juang and R. S. Pappa. An Eigensystem Realization Algorithm
for Modal Parameter Identification and Model Reduction. Journal of
Guidance, Control, and Dynamics, 8(5), pp. 620–627, 1985.
[40] Q. Qin, H. B. Li, L. Z. Qian, and C.-K. Lau. Modal identification of tsing
ma bridge by using improved eigensystem realization algorithm. Journal
of Sound and Vibration, 247(2), pp. 325–341, 2001.
[43] T. G. Carne and G. H. James III. The inception of OMA in the develop-
ment of modal testing technology for wind turbines. Mechanical Systems
and Signal Processing, 24(5), pp. 1213–1226, 2010.
199
modal analysis. Structural Health Monitoring, 15(3), pp. 289–301, 2016.
200
[51] Y. Yang and S. Nagarajaiah. Blind Modal Identification of Output-only
Structures in Time-domain Based on Complexity Pursuit. Earthquake
Engineering Structural Dynamics, 42, pp. 1885–1905, 2013.
201
[57] D. Uehara, J. Sirohi, R. Feil, and J. Rauleder. Blade Passage Loads and
Deformation of a Coaxial Rotor System in Hover. Journal of Aircraft,
2019.
[59] LaVision 2017 Strain Master 3D, Software Package Version 8.4. LaVi-
sion GmbH, Göttingen, Germany.
[61] P. D. Welch. The Use of Fast Fourier Transform for the Estimation of
Power Spectra: A Method Based on Time Averaging Over Short, Modified
Periodograms. IEEE Transactions on Audio and Electroacoustics, 15, pp.
70–73, 1967.
202
[64] L. Meirovitch. Fundamentals of Vibrations. McGraw-Hill Book, New
York, NY, 2000.
203
[72] T. W. Jones, A. A. Dorrington, P. L. Brittman, and P. M. Danehy. Laser
Induced Fluorescence for Photogrammetric Measurement of Transparent
or Reflective Aerospace Structures. 49th Annual International Instru-
mentation Symposium, Orlando, FL, May, 2003.
204
[79] D. Tcherniak, S. Chauhan, M. Rossetti, I. Font, J. Basurko, and O. Sal-
gado. Output-only Modal Analysis on Operating Wind Turbines: Appli-
cation to Simulated Data. European Wind Energy Conference, Warsaw,
Poland, April 20-23, 2010.
205
Vita
Contact: udaiju9@gmail.com
† A
LT EX is a document preparation system developed by Leslie Lamport as a special
version of Donald Knuth’s TEX Program.
206