You are on page 1of 19

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/268529297

Helicopter Rotor Blade Flexibility Simulation for Aeroelasticity and Flight


Dynamics Applications

Article  in  Journal of the American Helicopter Society · October 2014


DOI: 10.4050/JAHS.59.042006

CITATIONS READS

37 1,376

3 authors, including:

Ioannis Goulos Vassilios Pachidis


Cranfield University Cranfield University
92 PUBLICATIONS   671 CITATIONS    170 PUBLICATIONS   1,086 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

DEPART2050 View project

Variable geometry exhaust system for low specific thrust turbofans View project

All content following this page was uploaded by Ioannis Goulos on 28 July 2015.

The user has requested enhancement of the downloaded file.


JOURNAL OF THE AMERICAN HELICOPTER SOCIETY 59, 042006 (2014)

Helicopter Rotor Blade Flexibility Simulation for Aeroelasticity


and Flight Dynamics Applications

Ioannis Goulos∗ Vassilios Pachidis Pericles Pilidis


Research Fellow Senior Lecturer Professor
Centre for Propulsion, School of Engineering, Cranfield University
Bedfordshire, United Kingdom

This paper presents a mathematical model for the simulation of rotor blade flexibility in real-time helicopter flight dynamics
applications that also employs sufficient modeling fidelity for prediction of structural blade loads. A matrix/vector-based
formulation is developed for the treatment of elastic blade kinematics in the time domain. A novel, second-order-accurate,
finite-difference scheme is employed for the approximation of the blade motion derivatives. The proposed method is coupled
with a finite-state induced-flow model, a dynamic wake distortion model, and an unsteady blade element aerodynamics
model. The integrated approach is deployed to investigate trim controls, stability and control derivatives, nonlinear control
response characteristics, and structural blade loads for a hingeless rotor helicopter. It is shown that the developed methodol-
ogy exhibits modeling accuracy comparable to that of non-real-time comprehensive rotorcraft codes. The proposed method
is suitable for real-time flight simulation, with sufficient fidelity for simultaneous prediction of oscillatory blade loads.

Nomenclature flap/lag/torsion
ki,j elastic intermodal coupling coefficient between the
 flap/lag/torsion  ith and jth assumed modes
ALagrange Lagrange’s equation coefficient matrix flap/lag/torsion
mi,j inertial intermodal coupling coefficient between
[Ai ] global O to rotating reference frame Ai the ith and jth assumed modes
transformation matrix, i = 1, . . . , Nb N, Nb number of assumed mode shapes, number of blades
A(r), Ip (r), I (r) blade cross-sectional area, polar moment of inertia, p(r, t) beam element of mass dm = ρA(r)dr, located at
and area moment of inertia, respectively, m2 , m4 , beam radius r at time t
m4 Qi
flap/lag/torsion
generalized centrifugal force/moment
[Bpre ], [Zpre ] rotor blade precone and pre-sweep rotation corresponding to the ith coordinate
matrices flap/lag/torsion
qi (t) time-dependent generalized coordinate of the ith
d f̃pO , d m̃O
p beam element total differential force and moment mode shape
loads vectors {q̄ flap/lag/torsion }i ith eigenvector of Lagrange’s equation coefficient
E,G material Young’s and shear stress moduli, Pa flap/lag/torsion
matrix [ALagrange ]
e blade root/hinge offset from the hub center as a r, R local beam element and total rotor blade radius, m
fraction of rotor radius Ai
r̃p/A i
position vector of p(r, t) relative to the origin of Ai
F̃iO (t), T̃iO (t) force and moment vectors of the ith blade, r̃pb,offset airfoil center gravity offset vector from the blade
i = 1, . . . , Nb elastic axis
O fus
F̃rotor (t), F̃rotor (t) total force vector exerted by the main rotor on the O
r̃p/O position vector of p(r, t) relative to the origin of O
aircraft fuselage, N r̃Ai /O position vector of the origin of Ai
b
f̃aero , m̃baero blade element aerodynamic force and moment per [Tib ] global O to local blade element coordinate system
unit length vectors b transformation matrix
flap/lag/torsion
fi,j hub-spring intermodal coupling coefficient O
T̃i,offset (t) moment vector of the ith blade due to root/hinge
between the ith and jth assumed modes offset, i = 1, . . . , Nb
flap/lag/torsion
Gi,j overall effective stiffness intermodal coupling O
T̃i,root (t) moment vector of the ith blade due to root/hinge
coefficient between the ith and jth assumed modes spring stiffness, i = 1, . . . , Nb
flap/lag/torsion
Ii,j effective centrifugal stiffening intermodal coupling O
T̃rotor fus
(t), T̃rotor (t) total moment vector exerted by the main rotor on
coefficient between the ith and jth assumed modes the aircraft fuselage, Nm
t, t time, employed time-step during time marching, s
U, T strain and kinetic energy of the rotor blade, J
∗ Corresponding author; email: i.goulos@cranfield.ac.uk. ũ, u̇˜ translational velocity and acceleration vectors of
Manuscript received September 2012; accepted February 2014. the fuselage axes frame, = [u v w]T , [u̇ v̇ ẇ]T

DOI: 10.4050/JAHS.59.042006 042006-1 


C 2014 The American Helicopter Society
I. GOULOS JOURNAL OF THE AMERICAN HELICOPTER SOCIETY

Ai
ũp/A
A i
, ãp/A linear velocity and acceleration vectors of p(r, t) ()flap/lag/torsion referring to the blade flap, lag, and torsion degree
i i
relative to the origin of Ai of freedom, respectively
ũO b O
p , ũp , ãp , ãp
b
absolute linear velocity and acceleration vectors of ˜ fus , ()˜ O , ()˜ Ai , ()
() ˜b vector expressed in the annotated reference axes
p(r, t) system
ũAi , ãAi absolute linear velocity and acceleration vectors of
the origin of Ai
ũo , u̇˜ o translational velocity and acceleration vectors of
the global nonrotating O frame, Subscripts
= [uo vo wo ]T , [u̇o v̇o ẇo ]T
˜ /O , ()˜
() vctor relative to the origin of the annotated
V flight speed, m/s /Ai

wflap/lag (r, t) blade flap/lag transverse displacement, m reference axes system


flap/lag/torsion ˜A /
() vector referring to the origin of the rotating
Xi (r) shape for the ith mode of motion, m, rad i

Yoffset (r) airfoil center mass offset from elastic axis, positive reference frame of the ith blade Ai
˜ p/
() vector referring to beam element p(r, t)
towards the trailing edge, m
[β], [ζ ], [θtorsion ] local flap, lag, and torsion angle rotation matrices ()i,j blade or mode number indices
β(r, t), ζ (r, t) local beam element equivalent flap and lag angles,
rad
β0 , β1c , β1s rotor coning, longitudinal and lateral flapping
angles, rad Introduction
βpre , ζpre rotor blade pre-cone and pre-sweep angles, rad
β̂, ζ̂ equivalent flap and lag angles at the rotor blade Background
hub, rad
[] fuselage axes to global non-rotating reference The purpose of the main rotor on a helicopter is essentially threefold:
frame O transformation matrix to produce the required lift forces, to provide sufficient propulsion to
[0 ], [ i ] rotor blade collective pitch and azimuth angle overcome fuselage drag, and to effectively control the aircraft to sustain
rotation matrices trimmed flight or perform a designated maneuver. It is therefore recog-
θc (eR, t) cyclic pitch angle at the blade hub, rad nized that flight dynamics simulation not only must cater for accurate
θtw (r), θtorsion (r) local beam element pretwist and torsion angles, rad prediction of rotor power and thrust but must also provide precise esti-
θ0 , θ1s , θ1c collective, longitudinal cyclic, and lateral cyclic mates of the time-dependent rotor forces and moments exerted on the
pitch angles, rad aircraft fuselage. To address this requirement, it was necessary for the
[λ], [θc ], [θtw ] local sweep, cyclic pitch, and pretwist angle helicopter community to depart from the disk-like representation of the
rotation matrices main rotor and adapt methodologies treating each blade as an individual
λ(r) local beam element sweep angle, positive for entity. This not only allowed for the inclusion of more detailed rotor
backward sweep, rad dynamics models but also for the implementation of improved inflow
μ advance ratio, = V / R and aerodynamic response models. Initially developed individual–blade
ρ material density, kg/m3 rotor models would treat each blade as a rigid body. Effective flapping
σ standard deviation hinge offsets and hub springs were used to simulate the hub stiffness of
{ flap/lag/torsion (r)} assumed mode shape vector, hingeless blades. Such models include the “Genhel” model, originally
flap/lag/torsion
= [φj (r), j = 1, . . . , N]T developed by Howlett in Ref. 1 for the UH-60 Blackhawk helicopter.
φi
flap/lag/torsion
(r) assumed deformation function of the ith mode The development of more sophisticated rotor inflow (Ref. 2) and
shape, m, rad blade aerodynamics models (Ref. 3) has essentially led to the require-
ψi (t) azimuth angle of the ith blade in the global O ment for more advanced rotor dynamics models to keep up with the level
frame at time t, i = 1, . . . , Nb , rad of simulation fidelity. Brown and Houston (Ref. 4) performed a compar-
rotor speed, rad/s ative evaluation between a first-order finite-state dynamic inflow model
ωi
flap/lag/torsion
natural vibration frequency of the ith mode of (Ref. 5) and a formulation that solves the vorticity transport equation
motion, rad/s on a computational mesh enclosing the helicopter (Ref. 2). Both inflow
ω̃, ω̇˜ angular velocity and acceleration vectors of the models were coupled dynamically with a rigid–blade rotor model and a
fuselage axes frame, = [p q r]T ,[ṗ q̇ ṙ]T fuselage model. The effectiveness of both inflow models was assessed in
Ai ˜
ω̃p/Ai , ω̇p/Ai
Ai
angular velocity and acceleration vectors of p(r, t) terms of their quality of prediction of trimmed rotor blade flap, lag, and
relative to the origin of Ai feather angles, airframe attitudes, cross-coupling derivatives, response to
ω̃Ai , ω̇˜ Ai absolute angular velocity and acceleration vectors control inputs, and airframe vibration.
of the origin of Ai Brown and Houston noted that inclusion of real flow field effects is
ω̃o , ω̇˜ o angular velocity and acceleration vectors of the an important factor with regard to extending the frequency bandwidth
global nonrotating O frame, over which flight dynamics models are valid. However, they concluded
= [po qo ro ]T , [ṗo q̇o ṙo ]T that in the context of improving the overall level of flight dynamics
simulation fidelity, advances in rotor dynamics, and aerodynamics cannot
be considered in isolation from each other. Instead, they need to be
catered for simultaneously. Hence, extending the frequency bandwidth
Superscripts over which simulation models are applicable requires an integrated jump
in the complexity associated with several modeling aspects. In the case of
(), ¨ () , ()
˙ (), first and second derivatives with respect to time t, rotor dynamics, this essentially translates to dispensing with the classical
and blade radius r, respectively rigid blade model and implementation of flexible rotor blade dynamics.

042006-2
ROTOR BLADE FLEXIBILITY SIMULATION FOR AEROELASTICITY AND FLIGHT DYNAMICS APPLICATIONS 2014

Helicopter rotor blade flexibility in flight dynamics applications and pitch control angles for designated flight conditions. The complex
task of rotor design was therefore elaborately broken down within three
With regard to a hingeless rotor, treating the blade as a rigid body interdependent major steps of analysis: (1) prediction of aerodynamic
is essentially an approximation of the motion of the elastic blade and blade loads, (2) dynamic response of rotor blades to external forcing,
specifically of the first cantilever flap and lag modes. A real rotor blade and (3) aeroelastic rotor–aircraft trim.
responds to the exerted aerodynamic and inertial loads by deforming in More recently, Yeo and Johnson (Refs. 19, 20) performed an inves-
all of its modes. Shupe (Ref. 6) examined the effect of the second flap tigation of airloads and structural blade loads for five helicopter rotors
bending mode of a hingeless blade on the transmitted hub moment for operating at transition and high-speed flight conditions. They carried
a wide range of flight conditions. He supported the argument that the out extensive correlations between predictions made with CAMRAD II
more pronounced the radial nonlinearity in the once-per-rev harmonic and experimental data from wind tunnel and flight-test measurements.
forcing, the greater will be the contribution of higher modes. He therefore CAMRAD II is a comprehensive rotorcraft aeromechanics analysis code
concluded that including the influence of higher order modes is essential that combines a series of advanced technologies including unsteady non-
for the correct hub moment prediction of hingeless rotors, especially linear aerodynamics (Ref. 3), free-wake modeling (Ref. 21), multibody
considering high-speed flight conditions. dynamics, and finite element analysis (FEA) (Ref. 22). Yeo and Johnson
Lewis (Ref. 7) used a multibody dynamics formulation to investigate reported satisfactory correlations between predictions and measurements
the effect of blade flexibility on the dynamics of the UH-60 articulated considering flap bending and torsional moments for the majority of ro-
rotor and compared his results with flight-test data. He concluded that, tor models they investigated. Chord bending moment correlations were
for articulated rotor helicopters, the effect of blade flexibility is quite considered to be poor, which was attributed to unsatisfactory predictions
small in hover and increases slightly with increasing speed. However, of chordwise airloads and lag damper modeling.
Sturisky and Schrage (Ref. 8) report that for the articulated rotor of the
AH-64 helicopter, inclusion of higher order inflow dynamics coupled Scope of present work
with flexible blade modeling may substantially improve the prediction
accuracy of the off-axis rotor response to control inputs. Rotor blade aeroelasticity has been routinely modeled in non-
Turnour and Celi (Ref. 9) presented a methodology targeting the in- real-time comprehensive rotorcraft codes (UMARC, CAMRAD/JA,
clusion of coupled flap–lag–torsion (F–L–T) dynamics of elastic rotor CAMRAD II, S4, RCAS, DYMORE, and HOST (Ref. 23)) but has been
blades in flight dynamics applications. Their method included the cou- sparsely catered for in real-time flight dynamics applications. This is due
pling of the elastic rotor model used in Ref. 10 with the fuselage equations to the large computational overhead associated with high-fidelity model-
of the “UM-Genhel” flight dynamics model (Ref. 11) plus a finite-state ing methods employed by comprehensive codes, such as FEA, multibody
induced-flow model (Refs. 12, 13). The integrated formulation was sub- dynamics, and free-wake aerodynamics. This essentially prohibits the
sequently deployed to evaluate the influence of blade flexibility on the deployment of such methods in real-time flight dynamics applications,
frequency response characteristics of an articulated rotor helicopter in such as flight simulation. It is emphasized that the implementation of free-
hover and forward flight. Turnour and Celi concluded that, for the artic- wake inflow models in real-time flight simulation has been investigated
ulated rotor under investigation, including the effects of blade flexibility with encouraging results by Horn et al. in Ref. 24. However, such for-
along with higher order inflow dynamics does not improve the prediction mulations have only demonstrated real-time applicability when coupled
of the off-axis rotor response to control inputs. with rigid blade rotor models, quasi-steady blade element aerodynamics,
Theodore (Ref. 14) described an integrated, non-real-time, flight dy- and other standard assumption of flight simulation codes. Hence, they
namics simulation model, which included the effects of blade flexibility have not accounted for the effect of rotor blade flexibility and as such
with coupled F–L–T degrees of freedom and a free-wake inflow for- they are not applicable to the prediction of structural blade loads.
mulation. A baseline flight mechanics model, described in Ref. 15, was In light of the aforementioned research in the existing literature, the
coupled with the flexible blade model from Ref. 9 and the relaxation-type general scope of this work is to develop a mathematical approach for
free-wake model given in Ref. 16. Theodore used the integrated approach the simulation of rotor blade flexibility in real-time helicopter flight
to investigate the effects of inflow and blade dynamics modeling, on the dynamics applications that also employs sufficient modeling fidelity for
accuracy of the predicted control response characteristics of the Bo 105 prediction of oscillatory structural blade loads. The specific objectives
helicopter in near hover maneuvering conditions. They arrived at the of this paper are as follows:
conclusion that neither free-wake nor flexible blade models are required 1) To describe a numerical approach that enables transition from
to obtain accurate prediction of the on-axis rotor response. However, for classical rigid blade modeling to a complete framework for aeroelasticity
the given hingeless rotor system, accurate prediction of the off-axis rotor analysis, without resorting to cost-inducing FEA, multibody dynamics,
response requires sophisticated modeling and thus both free-wake and or free-wake aerodynamics.
flexible blade models must be used. 2) To assess the capability of the mathematical model to predict oscil-
latory rotor blade loads with accuracy comparable to that of established
Helicopter rotor blade flexibility in comprehensive analysis comprehensive analysis methods.
3) To identify the model’s ability to obtain accurate estimates of trim
Comprehensive rotorcraft codes are built predominantly for non-real- performance as well as stability and control (S&C) derivatives.
time rotor design applications, which entail reasonably accurate esti- 4) To investigate the effects of elastic blade and inflow dynamics
mates of oscillatory structural blade loads. The solution of the complete modeling fidelity on the accuracy of nonlinear control response simula-
aeroelastic response of helicopter rotor blades was originally tackled by tion.
Piziali and DuWaldt in Refs. 17 and 18. Their work was partially moti- 5) To evaluate the computational requirements of the proposed ap-
vated by unexplained oscillatory structural blade loads measured during proach and assess its applicability in real-time flight simulation.
flight tests. This harmonic content was subsequently attributed to highly The mathematical model for the simulation of rotor blade flexibil-
nonuniform wake-induced velocities. It was postulated in Ref. 18 that the ity presented in this paper is broadly arranged as follows: Initially a
main difficulty associated with designing a helicopter rotor system es- Lagrangian method is deployed for the estimation of natural vibration
sentially constitutes adequate estimation of performance, blade stresses, characteristics of rotating blades with nonuniform structural properties.

042006-3
I. GOULOS JOURNAL OF THE AMERICAN HELICOPTER SOCIETY

Lagrange’s equation of motion is applied to the kinematics of rotor vibration theory using the boundary conditions corresponding to the rotor
blade flap/lag bending and torsion. Modal properties obtained from the blade’s hub support. The cases of hingeless, freely hinged, and spring-
Bernoulli–Euler beam and classical torsional vibration theory are utilized hinged articulated rotor blades have been analyzed in Ref. 29. Closed-
as assumed deformation functions. Closed-form integral expressions are form expressions have been provided for the derivation of assumed modal
incorporated for each degree of freedom (DOF), describing the gen- functions for each type of hub support, including the subcases of infinite
eralized centrifugal forces and moments acting on the blade in terms as well as finite pitch control system torsional stiffness.
of normal coordinates. A matrix/vector-based formulation is developed The aforementioned process provides the N first vibration mode
flap/lag/torsion
for the treatment of three-dimensional kinematics of elastic blades. The shapes φi (r), i = 1, . . . , N for F–L–T (r being the local
proposed approach retains all nonlinear inertial terms associated with blade radius) corresponding to the idealized structure. The acquired func-
large blade deflections and fuselage angular motions. A stable, second- tions are orthogonal with each other, they satisfy the structure’s boundary
order-accurate, backward-difference scheme, originally developed for conditions, they approximate the actual mode shapes of the nonuniform–
time-accurate free-wake analysis (Ref. 25), is employed for the approx- rotating structure reasonably well (Ref. 30), and their derivatives with
imation of the blade motion derivatives in the time domain. Integral respect to r are readily available in analytical form. Hence, they are con-
expressions are provided describing the time-dependent rotor force and sidered excellent candidates for deployment as assumed deformations
moment loads, directly in the standard flight dynamics fuselage axes functions in Lagrange’s equation of motion.
frame of reference.
The developed methodology for rotor dynamics is coupled with a Lagrangian formulation for nonuniform rotating blades. Having derived
finite-state induced-flow model (Refs. 12,13), a dynamic wake distortion a series of well-conditioned assumed deformation functions, Lagrangian
model (Ref. 26), and an unsteady nonlinear blade element aerodynamics kinematics can be employed and the analysis can be extrapolated to the
model (Ref. 27). The combined aeroelastic rotor formulation is imple- nonuniform rotating case. The linearized version of Lagrange’s equation
mented in a helicopter flight dynamics code, thus devising an integrated of motion (Ref. 28) for a system whose space configuration is governed
simulation framework for helicopter aeromechanics. The integrated ap- by a finite series of time-dependent generalized coordinates qi (t), i =
proach is deployed to investigate trim performance, S&C derivatives, 1, . . . , N can be expressed as
oscillatory structural blade loads, and nonlinear control response for the  
d ∂T ∂U
MBB Bo 105 helicopter. Extensive comparisons are carried out with wind + = Qi , i = 1, . . . N (1)
dt ∂ q̇i ∂qi
tunnel/flight-test measurements and results from simulations conducted
with established comprehensive analysis and non-real-time, high-fidelity where T and U are the kinetic and strain energy of the system, whereas
flight dynamics codes. Qi signifies the generalized external force corresponding to the ith
coordinate.
Mathematical Formulation The system’s strain and kinetic energy corresponding to the F–L–
T DOFs can be expressed as functions of the assumed modal content
Rapid estimation of natural vibration characteristics of helicopter obtained from classical methods as follows:
1   flap/lag/torsion flap/lag/torsion
N N
rotor blades flap/lag/torsion
U flap/lag/torsion = k qi (t)qj (t)
2 i=1 j =1 i,j
A minimum potential energy approach based on Lagrangian kinemat-
1   flap/lag/torsion flap/lag/torsion
ics is utilized to obtain the rotor blade’s natural vibration characteristics N N
flap/lag/torsion
for F–L–T. The blade is modeled as a continuous rotating system of + f qi (t)qj (t)
2 i=1 j =1 i,j
nonuniform mass per unit length ρA(r), polar moment of inertia ρIp (r), (2)
flap/lag bending stiffness EI flap/lag (r), and torsional rigidity GJ (r).
The well-known deficiencies of previously described energy methods 1 N  N
flap/lag/torsion flap/lag/torsion flap/lag/torsion
T flap/lag/torsion = m q̇i (t)q̇j (t)
(Ref. 28) are mitigated by employing modal properties obtained from 2 i=1 j =1 i,j
classical theories as assumed deformation functions, instead of standard (3)
polynomial expressions found in the literature. This approach leads to flap/lag/torsion
considerable reduction in computational time, since modal frequencies where qi (t), i = 1, . . . , N are time-dependent generalized
converge rapidly to definitive values using only a small number of as- coordinates for F–L–T, respectively, expressing the contribution of each
sumed functions (Ref. 29). of the assumed modal functions. The intermodal coupling coefficients in
Neither inertial nor aeroelastic coupling is accounted for in the La- Eqs. (2) and (3) are defined as
grangian eigenproblem. This is due to the fact that the proposed approach  R
flap/lag flap/lag flap/lag
is designated for dynamic response analysis in the time domain. Thus, any mi,j = ρA(r)φi (r)φj (r) dr (4a)
eR
aerodynamic or nonlinear inertial loads may be treated as a time history  R
 flap/lag  flap/lag
of external forcing. Although elastically uncoupled blade motion is con- flap/lag
ki,j = EI flap/lag (r)φi (r)φj (r) dr (4b)
sidered in this paper, the modal properties of any theory can be utilized eR
flap/lag  flap/lag  flap/lag
with any level of coupling. It is emphasized that the proposed approach fi,j = K flap/lag φi (eR)φj (eR) (4c)
maintains inertial coupling between F–L–T through proper treatment of  R
the nonlinear inertial terms as external forcing. The Lagrangian method mtorsion
i,j = ρIp (r)φitorsion (r)φjtorsion (r) dr (4d)
has been extensively described and validated in Ref. 29. eR
 R
 
torsion
ki,j = GJ (r)φi torsion (r)φjtorsion (r) dr (4e)
Assumed deformation functions derivation. The theory of Lagrangian eR
kinematics requires a finite series of assumed deformation functions torsion
fi,j =K torsion
φitorsion (eR)φjtorsion (eR) (4f)
corresponding to each DOF (Ref. 28). These are obtained by initially
assuming uniform structural properties as well as nonrotating conditions where e is the root/hinge offset distance as a fraction of rotor blade

and subsequently applying classical Bernoulli–Euler beam and torsional radius R and superscripts () and ()˙ denote differentiation with respect to

042006-4
ROTOR BLADE FLEXIBILITY SIMULATION FOR AEROELASTICITY AND FLIGHT DYNAMICS APPLICATIONS 2014

flap/lag flap/lag flap/lag orthogonal with one another when using the inertial intermodal coupling
r and t, respectively. It is noted that mi,j , ki,j , and fi,j are the flap/lag/torsion
inertial, elastic, and hub-spring intermodal coupling coefficients in that coefficients mi,j for weighting.
order, between the ith and jth assumed modes of motion for flap/lag The final N first mode shapes of the nonuniform rotating blade can be
flap/lag acquired from the dot products of the assumed mode shape vectors (that
transverse displacement. The coupling term fi,j is nonzero only for
an articulated rotor employing a specifically defined hub spring with were obtained by classical methods) and the calculated eigenvectors of
flap/lag/torsion
stiffness K flap/lag . The terms mtorsion torsion
, ki,j , and fi,jtorsion
denote the corre- [A]Lagrange . This procedure can be expressed as
i,j
torsion
sponding coupling coefficients for torsional vibration. The term fi,j is
nonzero only for the case of finite pitch control system torsional stiffness flap/lag/torsion
Xi
flap/lag/torsion
(r) = { }flap/lag/torsion (r) · {q̄}i
i = 1, . . . , N ,
K torsion . (9)
Reference 29 presented a thorough kinematic analysis where closed- where { (r)} = [φj (r), j = 1, . . . , N]T are the assumed mode shape
form expressions were derived for the generalized centrifugal forces and vectors for F–L–T, respectively. It is shown in Ref. 29 that the orthogo-
torsional moments as functions of generalized coordinates corresponding flap/lag/torsion
nality condition between the eigenvectors of matrices [A]Lagrange due
to each DOF. The derived expressions give to their symmetric nature also leads to the orthogonality of the obtained
mode shapes given by Eq. (9) for the nonuniform rotating structure.
flap/lag/torsion

N
flap/lag/torsion flap/lag/torsion
Qi =− Ii,j qj (t), i = 1, . . . , N (5)
j =1
Modeling of aeroelastic rotor blades for helicopter flight dynamics
flap lag torsion
where Ii,j , Ii,j , and Ii,j are the centrifugal stiffening intermodal cou-
pling coefficients for flap–lag–torsion in that order. These have been Having obtained the necessary modal content with respect to the
found to be as follows: nonuniform rotating blades, the dynamic rotor response to any pilot con-
 R  r trol inputs or external forcing can be evaluated. To capture the true aeroe-
flap  flap  flap
Ii,j = ρA(r) 2 r φi (ξ )φj (ξ )dξ dr (6a) lastic behavior of the main rotor, it is necessary that the time-dependent
eR eR aerodynamic and inertial loads exerted on the blades are accurately pre-
 R  r
lag  lag  lag dicted. The helicopter may be pitching, rolling, climbing, or turning. The
Ii,j = ρA(r) 2 r φi (ξ )φj (ξ )dξ dr
eR eR inertial velocity and acceleration terms associated with the motion of
 R the helicopter, as well as the deformation of the elastic blades, need to
lag lag
− 2
ρA(r) φi (r)φj (r)dr (6b) be accounted for. Therefore, an effective treatment of the rotor blade’s
eR
 R
nonlinear, three-dimensional kinematics is required.
torsion
Ii,j = ρA(r) 2 Yoffset
2
(r)φitorsion (r)φjtorsion (r)dr (6c)
eR
Three-dimensional kinematics of elastic blades. Figures 1(a)–1(c)
where is the shaft rotor speed and Yoffset (r) is the airfoil center mass demonstrate the standard flight dynamics reference frames, namely fuse-
offset from the elastic axis. lage axes system, rotor hub axes, and hub-wind axes system, respectively.
Substituting Eqs. (2) and (3) along with Eq. (5) into Eq. (1) and The origin of the fuselage axes frame is located at the helicopter’s center
subsequently assuming periodic solution for the generalized coordinates of gravity (CG). The corresponding unit vectors ẽxfus , ẽyfus , ẽzfus essen-
of the form qi = q̄i sin(ωi t + ψi ) leads to the following system of tially point forward, starboard, and downward in that order, as shown
equations: in Fig. 1(a). The rotor hub axes system is aligned with the shaft and
has its origin located at the center of the rotor hub. The respective unit

N
 flap/lag/torsion  flap/lag/torsion 2 flap/lag/torsion flap/lag/torsion
Gi,j − ωi mi,j q̄j = 0, vectors ẽxh , ẽyh , ẽzh are obtained by rotating the vectors of the fuselage
j =1 system as depicted in Fig. 1(b) so as to factor in the longitudinal shaft
i = 1, . . . N (7) tilt angle γs . The origin of the hub-wind axes system is identical to that
of the hub frame. The hub-wind unit vectors ẽxhw , ẽyhw , ẽzhw are acquired
flap/lag/torsion
where ωi is the natural frequency of the ith mode for flap, lag, by rotating the hub system as shown
in Fig. 1(c) to account for the rotor
and torsional vibration in that order. The parameter side-slip angle ψw = arcsin vh / u2h + vh2 . This process aligns the system
flap/lag/torsion flap/lag/torsion flap/lag/torsion flap/lag/torsion
Gi,j = ki,j + fi,j + Ii,j , may be hw

vector ẽx with the resultant edgewise rotor hub velocity component
unit
viewed as the overall effective stiffness coupling coefficient between the uh + vh .
2 2

ith and jth assumed modes of F–L–T, including elastic ki,j , hub-spring The linear and angular velocity components of the helicopter fuselage,
fi,j , and centrifugal stiffening effects Ii,j . Rearranging Eq. (7) in matrix depicted in Fig. 1(a) are classified in vector notation in the fuselage axes
notation leads to the following expression: system as follows: ũ = [u v w]T , ω̃ = [p q r]T . Identical expressions
flap/lag/torsion are used for the corresponding acceleration vectors u̇˜ and ω̇˜ . A global
[A]Lagrange {q̄}flap/lag/torsion = {0} (8) nonrotating reference frame O is defined, which is aligned with the main
where {q̄} = [q̄1 q̄2 q̄3 . . . q̄N ]T considering the cases of flap, lag, and rotor shaft. The O frame’s origin is also located at the center of the main
flap
torsional vibration, respectively. It is noted that [A]Lagrange , [A]Lagrange ,
lag rotor hub. Figure 1(c) shows that the O system’s unit vectors are related
and [A]torsion are square symmetric matrices of size N . The neces- to the hub-wind axes vectors as follows: ẽxo = −ẽxhw , ẽyo = ẽyhw , ẽzo =
Lagrange
sary condition for Eq. (8) to have nontrivial solutions dictates that −ẽzhw . The respective linear and angular velocity components, ũo =
flap/lag/torsion
det[A]Lagrange = 0. This leads to a transcendental equation that can [uo vo wo ]T and ω̃o = [po qo ro ]T , which result from the corresponding
flap/lag/torsion fuselage motions, are expressed in the global nonrotating O reference
be evaluated numerically for the first N values of ωi , i =
flap lag frame as follows:
1, . . . , N that satisfy Eq. (8). The vectors {q̄}i , {q̄}i , and {q̄}torsion i ,i =
1, . . . , N are eigenvectors of matrices [A]Lagrange
flap/lag/torsion
and are associated ⎧ ⎫
⎨ u − q · hr ⎬
flap/lag/torsion flap/lag/torsion
with ωi , respectively. It is emphasized that since [A]Lagrange ũo = [] v + p · hr − r · xcg (10)
are square symmetric matrices, their eigenvectors {q̄}flap/lag/torsion are ⎩ ⎭
w + q · xcg

042006-5
I. GOULOS JOURNAL OF THE AMERICAN HELICOPTER SOCIETY

Z
wo , ro , wo , ro
~ Ai
v, v
e~zo ez
e~yfus CG
q, q e~yo vo , q o , vo , q o
O ~
rAi / O Y
p, p u, u e~xo ψ i (t ) e~yAi
w, w r, r ~
rpA/ iAi
e~xfus
(a)
e~z fus Ai
e~zo p (r , t )
u o , po , u o , po
X e~xAi
s
e~yh
(a)
e~xh hr e~yo
O
~ Z ψ i (t ) = 0
rAi / O ζ pre Y
~h ψ i (t ) e~zAi
CG xcg ez Fuselage e~xo e~yAi e~xAi
e~zo
Ai
(b) reference line
Ai β pre
e~yhw e~yo X e~xAi O ~
rAi / O e~xo X
e~hw e~yh (b)
x

e~zAi
vh
hw Ai
+ θ c ( eR , t )
uh
e~xh
e~zh e~zhw e~zo e~xo e~yAi
e~yb
p(r , t )
dψ i (t )
+β +
(c) e~xo e~xhw , e~yo e~yhw , e~zo e~zhw e~xb b
dt
flap
w +ζ
e~zb
Fig. 1. Standard nonrotating reference frames for helicopter flight
dynamics: (a) fuselage axes system, (b) main rotor hub axes system,
and (c) main rotor hub-wind axes system.
wlag
(c) e~xAi
⎧ ⎫
⎨p ⎬ Fig. 2. Reference systems for elastic blade kinematics: (a) global
ω̃o = [] q (11) nonrotating reference system O, (b) hinge presweep and precone
⎩ ⎭
r angles definition, and (c) rotating and blade element reference frames
where [] expresses vector transformation from the fuselage axes system ( Ai and b, respectively).
to the global nonrotating O frame of reference. This transformation is
shown graphically in Fig. 1 and is defined mathematically as
⎡ ⎤⎡ ⎤ signifies the orthogonal distance between the helicopter’s CG and the
−1 0 0 cos(ψw ) sin(ψw ) 0 point on the fuselage reference line located directly below the rotor hub,
[] = ⎣ 0 1 0 ⎦ ⎣− sin(ψw ) cos(ψw ) 0⎦ positive for a forward CG location (Fig. 1(b)). Expressions identical
0 0 −1 0 0 1 to those given by Eqs. (10) and (11) are used for the corresponding
⎡ ⎤ linear and angular acceleration components: u̇˜ o = [u̇o v̇o ẇo ]T and ω̇˜ o =
cos(γs ) 0 sin(γs )
[ṗo q̇o ṙo ]T , respectively.
×⎣ 0 1 0 ⎦ (12)
A separate set of rotating reference systems is defined for the descrip-
− sin(γs ) 0 cos(γs )
tion of three-dimensional elastic blade kinematics as shown in Fig. 2.
It is noted that hr is the orthogonal distance between the rotor hub and The present formulation places each beam element p(r, t) on the blade’s
the fuselage reference line, positive upward, as shown in Fig. 1(b). xcg mean elastic axis (Fig. 2(a)). Each element p(r, t) is treated as a material

042006-6
ROTOR BLADE FLEXIBILITY SIMULATION FOR AEROELASTICITY AND FLIGHT DYNAMICS APPLICATIONS 2014

object with differential mass dm = ρA(r)dr, whose center of gravity where wflap/lag (r, t) is the time-dependent blade transverse displacement
is located on the blade chord and is offset from the elastic axis by an as regards flap and lag motion in that order. The negative sign in front
amount equal to Yoffset (r). A rotating reference system Ai , i = 1, . . . , Nb of the second element of r̃p/A
Ai
has been included due to the convention
i
is defined for each of the Nb rotor blades (Fig. 2(a)), aligned with the regarding ζ (r, t) being positive for backward blade transverse displace-
time-dependent ith blade root/hinge position. The position vector of the ment wlag (r, t), as shown in Fig. 2(c). The first and second derivatives of
Ai origin, expressed in the global nonrotating reference frame O, can be Eqs. (21) and (22) lead to the respective angular velocity and acceleration
written as r̃Ai /O = [eR cos ψi (t) eR sin ψi (t) 0]T , where ψi (t) is the components of beam element p(r, t) expressed in Ai :
azimuth angle corresponding to the root/hinge of the ith blade at time t,
in the global O system (Fig. 2(a)). Assuming rotor precone, presweep, ẇflap (r, t)
β̇(r, t) =
(23)
and collective pitch angles βpre , ζpre , and θ0 , respectively (Fig. 2(b)), a (r − eR)2 − wf lap (r, t)2
tensor operator [Ai ] is defined as ẅflap (r, t) wflap (r, t) · ẇflap (r, t)2
β̈(r, t) =
+
[Ai ] = [0 ][Bpre ][Zpre ][ i ] (13) (r − eR)2 − wflap (r, t)2 ((r − eR)2 − wflap (r, t)2 )3/2
(24)
where the individual rotation matrices [0 ], [Bpre ], [Zpre ], and [ i ] are
given by Similar expressions to Eqs. (23) and (24) can be used for the determina-
tion of ζ̇ (r, t) and ζ̈ (r, t), respectively.
⎡ ⎤ The total angular velocity and acceleration vectors of beam element
1 0 0
[0 ] = ⎣0 cos θ0 sin θ0 ⎦ (14) p(r, t) including the blade motion due to cyclic pitch can be expressed
0 − sin θ0 cos θ0 in the rotating Ai reference frame as
⎧ ⎫
⎨θ̇c (eR, t)⎬
⎡ ⎤ Ai
cos βpre 0 sin βpre ω̃p/A = −β̇(r, t) (25)
i ⎩ ⎭
[Bpre ] = ⎣ 0 1 0 ⎦ (15) −ζ̇ (r, t)
− sin βpre 0 cos βpre
⎧ ⎫
⎡ ⎤ ⎨θ̈c (eR, t)⎬
Ai
cos ζpre − sin ζpre 0 ω̇˜ p/A = −β̈(r, t) (26)
i ⎩ ⎭
[Zpre ] = ⎣ sin ζpre cos ζpre 0⎦ (16) −ζ̈ (r, t)
0 0 1
for the case that cyclic pitch is applied inboard of the blade root/hinge
⎡ ⎤ location and
cos ψi (t) sin ψi (t) 0 ⎡ ⎤−1 ⎧ ⎫
[ i ] = ⎣− sin ψi (t) cos ψi (t) 0⎦ (17) cos β̂ cos ζ̂ − cos β̂ sin ζ̂ sin β̂ ⎨θ̇c (eR, t)⎬
ω̃p/Ai = ⎣ 0 ⎦
Ai
0 0 1 sin ζ̂ cos ζ̂ 0
⎩ ⎭
− sin β̂ cos ζ̂ sin β̂ sin ζ̂ cos β̂ 0
Application of [Ai ] to any vector expressed in O will yield the corre- ⎧ ⎫
sponding expression of the same vector in Ai . Note that only collective ⎨ 0 ⎬
+ −β̇(r, t) (27)
pitch θ0 has been applied for the alignment of Ai with the blade root/hinge ⎩ ⎭
position. The effects of cyclic pitch, blade twist, and torsion will be in- −ζ̇ (r, t)
dividually accounted for within Ai . The absolute angular velocity and
acceleration vectors of the origin of Ai expressed in O can be written as ⎡ ⎤−1 ⎧ ⎫
cos β̂ cos ζ̂ − cos β̂ sin ζ̂ sin β̂ ⎨θ̈c (eR, t)⎬
ω̃Ai = [po qo (ro + )]T and ω̇˜ Ai = [ṗo ṙo (q̇o + )]˙ T , respectively. Ai
ω̇˜ p/A =⎣ sin ζ̂ cos ζ̂ 0 ⎦ 0
The absolute linear velocity and acceleration vectors of the origin of Ai i ⎩ ⎭
− sin β̂ cos ζ̂ sin β̂ sin ζ̂ cos β̂ 0
expressed in O are therefore given by ⎧ ⎫
⎨ 0 ⎬
ũAi = ũo + ω̃Ai × r̃Ai /O (18) + −β̈(r, t) (28)
⎩ ⎭
−ζ̈ (r, t)
ãAi = u̇˜ o + ω̃Ai × (ω̃Ai × r̃Ai /O ) + ω̇˜ Ai × r̃Ai /O (19)
for outboard blade pitch application, where β̂ = β(eR, t) and ζ̂ =
The position vector of each beam element p(r, t) relative to the origin of ζ (eR, t). θc (eR, t) denotes the applied cyclic pitch angle at the rotor
and expressed in the rotating Ai reference system (Fig. 2(c)) is given by hub. Equations (25) and (26) are suitable for hingeless blades with
⎧ ⎫ clamped-end boundary conditions at the rotor hub and no hinges in-
⎨(r − eR) cos β(r, t) cos ζ (r, t)⎬ volved. Equations (27) and (28) are better-suited for articulated blades
Ai
r̃p/Ai = −(r − eR) sin ζ (r, t) (20) where cyclic pitch is applied outboard of the flap/lag hinge location.
⎩ ⎭
(r − eR) sin β(r, t) Thus, they include the respective transformation of applied cyclic pitch
where β(r, t) and ζ (r, t) are the element equivalent flap and lag angles, about the local flap/lag angles at the blade hub. The negative signs of
respectively, defined in Ai as shown in Fig. 2(c). They are mathematically β̇(r, t), β̈(r, t), ζ̇ (r, t), and ζ̈ (r, t) have been included in the right-hand
expressed as side of Eqs. (25)–(28) to comply with the positive axes notation used in
the Ai frame (Fig. 2(c)). It is noted that Eqs. (27) and (28) have been
wflap (r, t) derived assuming coincident flap and lag hinge locations. When con-
β(r, t) = arcsin (21)
r − eR sidering different sequences of hinge arrangements, Eqs. (27) and (28)
have to be modified accordingly along with the employed deformation
wlag (r, t) assumed functions to reflect the kinematics of the incorporated hinge
ζ (r, t) = arcsin (22) sequence.
r − eR
042006-7
I. GOULOS JOURNAL OF THE AMERICAN HELICOPTER SOCIETY

⎡ ⎤
The relative velocity and acceleration vectors of beam element p(r, t) 1 0 0
relative to the origin of and expressed in the Ai rotating frame are essen- [θtorsion ] = ⎣0 cos θtorsion (r, t) sin θtorsion (r, t) ⎦ (41)
tially given by 0 − sin θtorsion (r, t) cos θtorsion (r, t)
A A A
i
ũp/A i
= ω̃p/A
i
i
× r̃p/A
i
i
(29) ⎡ ⎤
Ai Ai  Ai Ai Ai Ai cos ζ (r, t) − sin ζ (r, t) 0
ãp/A i
= ω̃ p/A i
× ω̃ p/A i
× r̃p/A i
+ ω̇˜ p/A i
× r̃p/A i
(30) [ζ ] = ⎣ sin ζ (r, t) cos ζ (r, t) 0⎦ (42)
Ai 0 0 1
which along with they can be expressed in the global nonrotating
r̃p/A i
O frame of reference using the following transformations: ⎡ ⎤
cos β(r, t) 0 sin β(r, t)
[Ai ]−1 r̃p/A [β] = ⎣ ⎦
Ai
O
r̃p/A i
= i
(31) 0 1 0 (43)
ũO = [Ai ]−1 ũp/A
Ai
(32) − sin β(r, t) 0 cos β(r, t)
p/Ai i

= [Ai ]−1 ãp/A


O i A
ãp/A i i
(33) ⎡ ⎤
1 0 0
The absolute velocity and acceleration vectors of point p(r, t) ex- [θc ] = − ⎣0 cos θc (eR, t) sin θc (eR, t) ⎦ (44)
pressed in the global nonrotating reference frame O are therefore given 0 − sin θc (eR, t) cos θc (eR, t)
by
where λ(r), θtw (r), and θtorsion (r) are the local blade element sweep,
pretwist, and torsion angles, respectively. The angles θtw (r), θtorsion (r),
p = ũAi + ũp/Ai + ω̃Ai × r̃p/Ai
ũO O O
(34)
 and θc (eR, t) are positive for nose-up rotation, whereas λ(r) is taken as
ãpO = ãAi + ãp/A
O
i
+ 2 · ω̃Ai × ũO
p/Ai + ω̃Ai × ω̃Ai × r̃p/Ai
O
positive for backward blade sweep. It is noted that the applied sequence
+ω̇˜ Ai × r̃p/A
O
i
(35) of rotations expressed by Eq. (38) is valid for configurations that cyclic
pitch is applied inboard of the flap/lag hinge location. It is therefore
It is of interest to obtain the velocity and acceleration components suitable for the analysis of hingeless rotor blades. As regards the case of
expressed in a blade element fixed coordinate system. Figure 2(c) il- articulated blades where cyclic pitch is applied outboard of the flap/lag
lustrates the local blade element frame of reference corresponding to hinge, the applicable sequence of rotations gives
point p(r, t) of the ith blade. The system’s orientation is specified by ⎡ ⎤⎡ ⎤−1
unit vectors ẽxb , ẽyb , and ẽzb . The vector ẽxb is taken so as to be coincident  b 1 0 0 cos β̂ cos ζ̂ − cos β̂ sin ζ̂ sin β̂
with the blade’s deformed elastic axis, pointing radially outward. The Ti = ⎣0 −1 0 ⎦ ⎣ sin ζ̂ cos ζ̂ 0 ⎦
ẽyb vector points toward the element’s trailing edge, and ẽzb completes 0 0 −1 − sin β̂ cos ζ̂ sin β̂ sin ζ̂ cos β̂
the orthogonal system. Conversion to a locally fixed reference frame not
only provides the local flow field conditions which is necessary input for × [θc ][β][ζ ][θtorsion ][θtw ][λ][Ai ] (45)
any aerodynamic theory but also enables application of D’Alembert’s
principle with respect to the superposition of inertial loads. Finite-difference treatment of elastic blade motion in the time domain.
The global expressions given by Eqs. (34) and (35) can be converted Evaluation of Eqs. (23) and (24) requires determination of the first and
to the local blade element coordinate system using the following trans- second derivatives of wflap (r, t) and wlag (r, t) with respect to time. Ap-
formation: plicability of the developed approach in the time domain entails nu-
   merical discretization of elastic blade motion. A second-order-accurate,
ũbp = Tib ũO p (36) backward-difference approximation scheme, originally developed by
 b  O  Bhagwat and Leishman in Ref. 25 for time-accurate free-wake analy-
ãp = Ti ãp
b
(37)
sis, is deployed for the evaluation of the temporal derivatives of interest.
The [Tib ] operator is a transformation matrix which relates any vector This scheme is written as
expressed in the global nonrotating reference frame O directly to the ∂wflap/lag (r, t + t/2)
local blade element coordinate system b of point p(r, t) located on the ≈
∂t
ith blade. It is defined as the following sequence of rotations: 3wflap/lag (r, t+t)−wflap/lag (r, t)−3wflap/lag (r, t−t)+wflap/lag (r, t−2t)
⎡ ⎤ 4t
 b 1 0 0
(46)
Ti = ⎣0 −1 0 ⎦ [θc ][β][ζ ][θtorsion ][θtw ][λ][Ai ] (38)
0 0 −1 where t is the employed time-marching step. A linearized stability
analysis has been carried out in Ref. 25 regarding the properties of
where [λ], [θtw ], [θtorsion ], [ζ ], [β], and [θc ] are rotation matrices cor-
Eq. (46) where it has been shown that it is stable with positive damping.
responding to the blade element sweep, pretwist, torsion, lag, flap, and
Equation (46) provides the first temporal derivatives of wflap/lag at
cyclic pitch angle at the hub, respectively. These are defined as
the time instant corresponding to t + t/2. Application of Eq. (46) on
⎡ ⎤ the first temporal derivatives of wflap/lag taken at t + t/2, t − t/2,
cos λ(r) − sin λ(r) 0
[λ] = ⎣sin λ(r) cos λ(r) 0⎦ (39) t − 3t/2, etc., essentially provides the second temporal derivative of
0 0 1 displacement at time = t. However, the present numerical approach
requires knowledge of both first and second temporal derivatives of
blade displacement at the same time instant. Therefore, the original
⎡ ⎤
1 0 0 finite-difference approximation expressed by Eq. (46) is modified by
[θtw ] = ⎣0 cos θtw (r) sin θtw (r) ⎦
flap/lag (r,t)+w flap/lag (r,t+t)
(40) assuming wflap/lag (r, t + t/2) ≈ w 2
to provide the
0 − sin θtw (r) cos θtw (r) first temporal derivative at the time instant corresponding to time = t as

042006-8
ROTOR BLADE FLEXIBILITY SIMULATION FOR AEROELASTICITY AND FLIGHT DYNAMICS APPLICATIONS 2014

well. The modified algorithm gives i = 1, . . . , Nb . The component T̃i,offset


O
(t) is the moment that the blade
root/hinge applies about the origin O due to its offset from the center
∂wflap/lag (r, t) 3wflap/lag (r, t + t) + 2wflap/lag (r, t)
≈ O
of rotation and is given by T̃i,offset (t) = r̃Ai /O × F̃iO (t). The component
∂t 8t O
T̃i,root (t) is the moment that the blade exerts directly on its root/hinge
−4wflap/lag (r, t − t) − 2wflap/lag (r, t − 2t) + wflap/lag (r, t − 3t) due to the root/hinge spring stiffness. For a hingeless rotor blade, the
+
8t component T̃i,rootO
(t) is due to the effective stiffness of the blade root and
(47) can be expressed as
It can be shown that Eq. (47) is also second-order accurate with respect ⎧ 


⎪ GJ (eR)θtorsion (eR, t) ⎪ ⎪
to time. The second temporal derivatives of wflap/lag (r, t) are given by ⎨ ⎬
−1  flap
T̃i,root (t) = [Ai ]
O
−EI (eR)w (eR, t) , i = 1, . . . Nb (54)
flap

⎪ ⎪
∂ 2 wflap/lag (r, t)

3ẇflap/lag (r, t + t/2) − ẇflap/lag (r, t − t/2) ⎩ −EI lag (eR)w lag (eR, t) ⎪ ⎭
∂t 2 4t
−3ẇflap/lag (r, t − 3t/2)ẇflap/lag (r, t − 5t/2) O
Alternatively, T̃i,root (t) can be obtained by integrating all aerodynamic
+
4t (48) and inertial loads acting on the blade at the respective time-instant t. This
process yields
where the values of ẇflap/lag in Eq. (48) can be evaluated at their respec-  R  R
tive time instants using the original algorithm given by Eq. (46). The O
T̃i,root (t) = O
r̃p/A × df̃ O
+ dm̃O i = 1, . . . , Nb (55)
i p p,
aforementioned process is also applied for the approximation of the first eR eR
and second temporal derivatives of θtorsion (r, t). For a spring-hinged articulated rotor, the moment components at the
The superior stability characteristics of the employed finite-difference hinge are determined from the local flap/lag angles and the corresponding
schemes developed in Ref. 25 essentially translate to the stability and spring stiffness constants. For a freely hinged articulated rotor, the blade
robustness of the proposed mathematical approach. It is noted that initial cannot exert any moment at its hinge about a DOF for which there is no
versions of the formulation that incorporated typical second-order central spring or stiffness involved.
finite-differences exhibited significant instabilities, particularly during The overall rotor force and moment vectors about the helicopter’s
transient control response analysis, thus requiring very small time-steps CG (Fig. 1(a)) can now be computed directly in the fuselage axes system
for stable simulation. However, the employed schemes (Eqs. (47) and using the following expressions:
(48)) have essentially alleviated the aforementioned issues and stable
simulations can be performed with considerably larger time steps leading fus
F̃rotor (t) = []−1 F̃rotor
O
(t) (56)
to equivalent azimuth steps of the order to 15◦ −20◦ . Thus, utilization of −1
 
the particular algorithms not only results in increased simulation stability
fus
T̃rotor (t) = [] O
T̃rotor (t) + [−xcg 0 − hr ] × []−1 F̃rotor
T O
(t) (57)
but also in significant reduction of computational time. This essentially Equations (56) and (57) can be used without any modification by any
constitutes an enabler in the context of real-time flight simulation. helicopter flight dynamics application utilizing the fuselage axes ref-
erence system as illustrated in Fig. 1(a) for the calculation of the time-
Numerical integration of nonlinear aeroelastic blade loads. Concerning dependent three-dimensional rotor force and moment loads exerted on the
the process of rotor loads integration, Eq. (37) can be used for the eval- aircraft.
uation of the nonlinear acceleration vector expressed in the local blade
element coordinate system. Using the principle of superposition for the
Integrated aeroelastic rotor–aircraft model
inertial and aerodynamic loads and converting them to the global system
O yields
The developed rotor dynamics formulation is coupled with the Peters–
 −1  b 
d f̃pO = Tib f̃aero − ρA(r)ãpb dr (49) He finite-state induced-flow model (Refs. 12,13), the Leishman–Beddoes
 −1    unsteady nonlinear blade element aerodynamics model (Ref. 27), and the
d m̃O = T b
p i m̃b − ρI θ̈˜
aero − ρA(r) r̃ b,offset × ã b dr (50)
p torsion p p Zhao–Prasad dynamic wake distortion model (Ref. 26). Experimentally
derived quasi-steady airfoil data are incorporated for the calibration of
where θ̈˜torsion = [θ̈torsion (r, t) 0 0]T and r̃pb,offset = [0 Yoffset (r) 0]T . It is the Leishman–Beddoes model. Careful treatment of the induced-flow
noted that the vector r̃pb,offset denotes the airfoil center gravity offset from terms arising from shed wake circulation has been carried out, so that
b
the blade’s elastic axis. The vectors f̃aero and m̃baero signify the aerody- they are not duplicated by both induced-flow and unsteady aerodynamics
namic force and moment loads per unit length, respectively. These can be models. The modal properties obtained from the Lagrangian approach are
obtained using any aerodynamic theory. The total time-dependent forces utilized. The dynamic response of the elastic blades to external forcing
and moments exerted by the main rotor on the aircraft fuselage can be is calculated using a fifth-order-accurate numerical evaluation scheme of
expressed as the summation of contributions from the individual blades the convolution integral.
The aircraft fuselage is treated dynamically as a rigid body with six

Nb
O
F̃rotor (t) = F̃iO (t) (51) DOF (three translations and three rotations) using the inertial tensors
i=1 given in Ref. 31. Experimentally derived lookup tables from Ref. 31 are
employed that describe the fuselage force and moment coefficients as

Nb
O
T̃rotor (t) = T̃iO (t) (52) functions of incidence and sideslip angles. Two-dimensional character-
i=1 istics are utilized for the prediction of the aerodynamic behavior of the
 R horizontal and vertical stabilizers. Steady-state airfoil data along with
F̃iO (t) = d f̃pO , i = 1, . . . , Nb (53) first-order dynamic inflow (Ref. 5) are used for the prediction of tailrotor
eR
performance.
The overall moment that a rotor blade with a root/hinge offset A globally convergent Newton–Raphson method is incorporated to
exercises about the origin of O is given by T̃iO (t) = T̃i,offset
O
(t) + T̃i,root
O
(t), obtain rotor trim, for either wind tunnel simulation with prescribed hub

042006-9
I. GOULOS JOURNAL OF THE AMERICAN HELICOPTER SOCIETY

moments, or for a designated set of flight conditions. The numerical 1st flap Lagr. 2nd flap Lagr.
scheme is used for the solution of Euler’s nonlinear differential equations 3rd flap Lagr. 4th flap Lagr.
for rigid body kinematics, applied to the aircraft fuselage (Ref. 31). The 1st lag Lagr. 2nd lag Lagr.
main and tail rotors are marched simultaneously in time using the initial 1st torsion Lagr. 2nd torsion Lagr.
condition of nonexistent circulatory wake. The time-marching process 1st lag Y-71 1st flap Y-71
continues until a once-per-rev (1P) periodic condition is achieved with 2nd flap Y-71 2nd lag Y-71
respect to the main rotor flap/lag mean and first harmonic multiblade
3rd flap Y-71 1st torsion Y-71
coordinates β0 , β1c , β1s , ζ0 , ζ1c , and ζ1s . Having achieved periodicity
2nd torsion Y-71
in the aeroelastic rotor response, a finite series of rotations is carried out 450
where the main rotor forces and moments are averaged in time to acquire 8P
mean representative values to be used for trim. 400

Mode frequency (rad/s)


Estimation of S&C derivatives is conducted by numerically perturb- 7P
350 4F
ing the integrated model around an acquired trim condition. Extraction 6P
of perturbation forces and moments is carried out when 1P periodicity is 300 2T
achieved with respect to the main rotor flap/lag multiblade coordinates. 3F 5P
250
The corresponding S&C derivatives are subsequently calculated using 2L 4P
second-order central finite differences. Simulation of nonlinear control 200
response is carried out after obtaining rotor trim by applying a predefined 150 1T 3P
schedule of control inputs. The fourth-order-accurate Runge–Kutta inte- 2F 2P
100
gration scheme is employed to integrate Euler’s differential equations for
1F
rigid body kinematics applied to the helicopter fuselage in the respective 50 1P
frame of reference. 1L
0
0 10 20 30 40 50
Results and Discussion
Rotor speed (rad/s)
The integrated simulation model has been deployed for the investiga-
Fig. 3. Resonance chart calculated for the hingeless rotor blade of
tion of trim, stability, control response, and oscillatory structural blade
the MBB Bo 105 helicopter: comparison with calculations performed
loads of a full-scale helicopter modeled after the MBB Bo 105. The
using Boeing–Vertol’s Y-71 computer code, reported in Ref. 32.
design characteristics of the preconed hingeless rotor of the Bo 105 he-
licopter have been extensively analyzed in Ref. 32. Simulations have
been performed for a wide range of advance ratios (μ = V / R). Unless
indicated otherwise, a total of eight inflow harmonics are deployed for with quasi-steady, two-dimensional airfoil characteristics. Unsteady flow
the simulation results presented in this paper, to capture the effect of corrections are employed coupled with an empirical dynamic stall model.
higher harmonic loading on rotor inflow response. The corresponding Empirical corrections are also applied for three-dimensional flow effects
radial inflow distribution polynomials used reach up to the tenth power including yawed flow, tip-relief, and blade sweep (Ref. 3). A free-wake
of nondimensional rotor radius. The particular model configuration es- model is deployed to predict the rotor wake induced velocities (Ref. 21).
sentially leads to a dynamic-wake representation with 62 states. A total Thus, the self-induced distorted nature of the wake due to the rollup
of six F–L–T vibration modes are employed unless stated otherwise. of the tip vortices downstream of the main rotor blades is adequately
Extensive comparisons of predicted oscillatory structural blade loads modeled.
have been carried out with experimental data from wind tunnel and
flight-test measurements reported in Ref. 33. Correlations have also been Rotor blade modal analysis
conducted with flight-test identification results reported in Refs. 34–36,
in terms of trim controls, S&C derivatives, and nonlinear aircraft response Figure 3 presents the calculated resonance chart for the full-scale,
to pilot controls. With reference to the structural blade loads predictions, hingeless rotor blade of the Bo 105 helicopter. Results from nonlinear
the main rotor is numerically trimmed to prescribed values of thrust and simulations conducted using Boeing–Vertol’s computer code Y-71 in
hub moments corresponding to a specific advance ratio and shaft angle Ref. 32, for flap, lag, and torsion resonant frequencies at nominal rotor
relative to the freestream flow. This is done in order for the simulations speed ( = 44.4 rad/s) are also included for comparison. Boeing–Vertol
to comply with the flight and wind tunnel testing procedures reported in Y-71 essentially calculates elastically and inertially coupled modal prop-
Ref. 33. erties for F–L–T. The Lagrangian approach treats each DOF separately
Predictions made with the comprehensive code CAMRAD/JA (Com- due to the fact that nonlinear inertial coupling terms are included as ex-
prehensive Analytical Model of Rotorcraft Aerodynamics and Dynamics, ternal forcing functions in the corresponding equations of motion during
Johnson Aeronautics (Ref. 37)) of oscillatory structural blade loads corre- a dynamic response analysis (Eqs. (34) and (35)). It is noted that the
sponding to the wind tunnel test conditions are compared with the present solid and broken lines denote Lagrangian predictions (Lagr.), whereas
modeling approach. All CAMRAD/JA results have been extracted from the markers signify Y-71 calculations.
Ref. 33. The rotor dynamics model of CAMRAD/JA is based on en- Very good agreement between the Lagrangian method and the anal-
gineering beam theory for rotating wings with large pitch and pretwist ysis conducted with Boeing–Vertol Y-71 can be observed with respect
angles (Ref. 22). The rotor blade is modeled as a cantilever beam with to the predicted resonant frequencies at nominal rotor speed. A small
a straight, undeformed elastic axis. Free-vibration F–L–T modes are uti- deviation with regard to the frequency of the second torsion mode can be
lized that are elastically coupled in flap/lag bending motion but uncoupled observed. This is attributed to the absence of elastic and inertial modal
in torsion. Inertial and aerodynamic coupling is retained through proper coupling in the predictions carried out with the Lagrangian approach.
treatment of the inertial velocity and acceleration terms in the employed It is noted that the total computational time required for modal analysis
equations of motion. Second-order lifting-line theory is utilized along using the Lagrangian method is of the order of half a second on a modern

042006-10
ROTOR BLADE FLEXIBILITY SIMULATION FOR AEROELASTICITY AND FLIGHT DYNAMICS APPLICATIONS 2014

Simulation ability of trim controls indicates the sound estimation of the aeroelas-
Flight test tic rotor response to 1P harmonic excitation. Figure 4 presents rotor
trim control angles and power requirement predictions for the Bo 105
Rotor power required (kw)

Free-wake/FEA
600 helicopter. Comparisons with flight-test data (Ref. 34) and simulation
results from a high-fidelity flight dynamics simulation code (Ref. 14) are
500
also included for comparison. It is noted that the simulation method of
400 Ref. 14 is a non-real-time formulation employing advanced technolo-
gies. Elastically and inertially coupled modal properties obtained from
300 nonlinear FEA are used for the structural representation of the elastic
blades. A relaxation-type free-wake model is deployed for the prediction
200 of wake-induced flow, thus accounting for the rollup of the tip vortices
0 0.1 0.2 0.3 0.4
downstream of the main rotor blades.
(a) Advance ratio
Figures 4(a) and 4(b) present trim values of main rotor power re-
20 quirement and collective pitch angle (θ0 ), respectively. Good correlation
Collective pitch angle

18 can be observed between the present approach and flight-test data re-
garding both trim outputs for μ ≤ 0.2. However, the agreement is not as
16 good considering higher advance ratios. This is attributed to a potential
(deg)

14 underprediction of aerodynamic fuselage drag in the employed lookup


tables. Very good correlation can be observed between predictions made
12 with the present approach and the flight dynamics model from Ref. 14,
10 considering higher values of advance ratio for which flight-test data are
0 0.1 0.2 0.3 0.4 not available. The collective pitch angle correlation between the two
Advance ratio simulation models is excellent for μ ≥ 0.15, where the flow field in the
(b) vicinity of the main rotor begins to be dominated by the freestream flow.
4 Reference 14 showed that implementation of a simple fuselage drag
Longitudinal cyclic

model with an equivalent flat-plate area of 20 ft2 may improve correla-


pitch angle (deg)

2
tions of power requirement and collective pitch angle significantly.
0 Figures 4(c) and 4(d) present trim values of longitudinal (θ1s ) and
−2
lateral (θ1c ) cyclic pitch control angles, respectively. Good agreement is
observed between predictions made with the present model and flight-test
−4 data considering both cyclic control inputs. Rotor wake impingement on
the horizontal stabilizer leads to underprediction of longitudinal cyclic
−6
0 0.1 0.2 0.3 0.4 pitch (θ1s ) for μ ≤ 0.02. θ1s is also underpredicted for μ ≥ 0.15. This is
Advance ratio attributed to the aforementioned underprediction of airframe drag along
(c)
with a potential underestimation of the aerodynamic fuselage pitching
4 moment. A noticeable deviation between the predicted and measured
Lateral cyclic pitch

3 lateral cyclic (θ1c ) for μ ≈ 0.05 reveals the requirement for more so-
angle (deg)

phisticated wake modeling considering the specific flight region. The


2 strong influence of the wake rollup on induced flow during transition
1 from hover is not accounted for by the Peters–He inflow model, which
assumes a skewed cylindrical wake. This deficiency essentially leads to
0 the underprediction of θ1c for μ ≤ 0.1.
−1 The predictive qualities exhibited by the developed aeroelastic rotor
0 0.1 0.2 0.3 0.4 model indicate its sound ability to estimate the quasi-steady aeroelastic
(d) Advance ratio rotor response to pilot-induced 1P harmonic excitation. Hence, the model
can be utilized for prediction of oscillatory structural blade loads. This
Fig. 4. Trim results for the MBB Bo 105 helicopter: comparison
conclusion is valid for advance ratios above μ ≈ 0.1, where the wake
with flight-test data and free-wake/FEA simulation results extracted
roll-up effect on rotor-induced flow is reduced substantially.
from Ref. 14. (a) Rotor power required, (b) collective pitch angle
θ0 , (c) longitudinal cyclic pitch angle θ1s , and (d) lateral cyclic pitch
angle θ1c .
Oscillatory structural rotor blade loads: Comparison with
experiment and comprehensive analysis

personal computer. The obtained modal content is utilized in this paper Figures 5 and 6 present correlations between predictions made with
for the estimation of the rotor blade’s dynamic response to nonlinear the developed simulation model (Sim.) and wind tunnel (WT) as well
external forcing. as flight-test (FT) measurements of oscillatory structural blade loads for
the full-scale hingeless rotor of the Bo 105 helicopter. Results from
Aeroelastic rotor–aircraft trim performance simulations carried out with the comprehensive code CAMRAD/JA
(CMRD/JA) for the wind tunnel test conditions are also included for
Before proceeding to correlations of oscillatory structural blade loads comparison. All wind tunnel and flight-test data as well as CAMRAD/JA
with experimental data, it is essential that the trim procedure is validated predictions have been extracted from Ref. 33. The correlations pre-
in terms of rotor trim control angles. The reason is that good predictive sented correspond to μ ≈ 0.197. It is noted that there are substantial

042006-11
I. GOULOS JOURNAL OF THE AMERICAN HELICOPTER SOCIETY

FT Sim. WT Sim. FT Sim. WT Sim.


Flight test WT test Flight test WT test

Flapwise bending moment


CMRD/JA
Flapwise bending moment

250 CMRD/JA
100
75
150
50

(Nm)
50 25
(Nm)

0
−50 −25
−50
−150 −75
−100
−250
0 90 180 270 360
0 90 180 270 360
(a) Azimuth angle (deg)
(a) Azimuth angle (deg)
300 200
Chordwise bending

Chordwise bending
150
200
moment (Nm)

moment (Nm)
100
100 50
0
0 −50
−100
−100
−150
−200 −200
0 90 180 270 360 0 90 180 270 360
(b) (b) Azimuth angle (deg)
Azimuth angle (deg)
40 40
Torsional moment (Nm)
Torsional moment (Nm)

30 30
20 20
10 10
0 0
−10 −10
−20 −20
−30
−30
−40
−40
0 90 180 270 360
0 90 180 270 360 (c) Azimuth angle (deg)
(c) Azimuth angle (deg)
Fig. 6. Oscillatory structural blade loads for μ = 0.197: comparison
Fig. 5. Oscillatory structural blade loads for μ = 0.197: comparison
with flight tests, wind tunnel experiments, and CAMRAD/JA simu-
with flight tests, wind tunnel experiments, and CAMRAD/JA simu-
lations reported in Ref. 33. (a) Flapwise bending moment for r/R =
lations reported in Ref. 33. (a) Flapwise bending moment for r/R =
0.57, (b) chordwise bending moment for r/R = 0.45 (FT) and- r/R =
0.144, (b) chordwise bending moment for r/R = 0.144, (c) torsional
0.57 (WT), and (c) torsional moment for r/R = 0.57.
moment for r/R = 0.40.

variations in rotor hub moments and shaft angle between flight-tests and Good agreement is observed between results obtained from simu-
wind tunnel experiments, which have been summarized in Ref. 33. These lations performed with both rotor models for all DOFs, considering
variations essentially lead to considerable differences in the correspond- the wind tunnel test conditions. Figure 5(a) shows that the correlation
ing rotor trim states and are thoroughly accounted for by the employed between the present modeling approach and experimental data con-
trim procedure. sidering the 1P component of flap bending moment at the blade root
Considering the wind tunnel correlations, comparisons of oscillatory (r/R = 0.144) is very good. This is due to the fact that the quasi-steady
flapwise and chordwise bending moments are presented for radial sta- hub moments have been essentially prescribed by the trim procedure. It
tions corresponding to the effective blade root (r/R = 0.144) and near is noted that CAMRAD/JA as well as the present model both overpredict
midspan (r/R = 0.57) locations. Torsional loading is presented for two the 2P and 3P components of oscillatory flap loading at the blade root
radial stations located near midspan at r/R = 0.40 and 0.57, respec- compared to the wind tunnel measurements. Good agreement is exhib-
tively. For the correlations with flight tests, comparisons of flapwise and ited between simulations carried out with the present rotor model and
chordwise loading are carried out at different midspan radial locations to flight-test data. The respective 2P and 3P oscillatory loading components
maintain consistency with the measured data. Flapwise bending moment residing in the measured waveform have been adequately captured by
results near midspan correspond to r/R = 0.57, whereas chordwise load- the proposed approach.
ing is presented for r/R = 0.45. Both flap and chord bending moment Good agreement is shown in Fig. 6(a) for r/R = 0.57 between CAM-
predictions at the blade root are extracted for r/R = 0.144. Compar- RAD/JA and the developed simulation model, considering the behavior
isons of torsional moments are only presented for one radial location of flap bending moment. Both formulations have succeeded in capturing
corresponding to r/R = 0.40 due to lack of available flight-test data. the amplitude of the measured 3P moment component, but they both

042006-12
ROTOR BLADE FLEXIBILITY SIMULATION FOR AEROELASTICITY AND FLIGHT DYNAMICS APPLICATIONS 2014

underpredict the impact of 2P loading. Neither rotor model has captured Figure 6(c) presents the respective torsional moment correlations for
the 8P oscillatory component of flap moment residing in both wind tunnel r/R = 0.57. Comparisons are carried out with wind tunnel experiments
and flight-test measurements. Good correlation in terms of 1P, 2P, and 3P only, due to lack of available flight-test data for the particular radial
oscillatory flap loading is observed between predictions made with the location. It can be noticed that the 8P moment component present in
present model and flight-test data. Both simulation models agree very the wind tunnel measurements has not been captured by neither simula-
well with the oscillatory loading measured during flight tests, despite the tion model. However, good agreement can be observed between predic-
fact that CAMRAD/JA predictions have been carried out for the wind tions made with the proposed modeling approach and the comprehensive
tunnel test conditions. CAMRAD/JA.
It is noted that the amplitude of higher harmonic content residing in Thus, it has been shown that the rotor model of this paper and the
the oscillatory loading waveforms measured in the wind tunnel is signif- comprehensive code CAMRAD/JA exhibit similar levels of accuracy for
icantly higher compared to that obtained from flight tests. Reference 33 oscillatory structural blade loads prediction. The two simulation models
attributed this high-frequency content to aerodynamic interaction effects have performed along similar lines due the fact that they both account
between the main rotor and the test apparatus. This difference in higher for the dominant dynamic and aerodynamic effects that influence rotor
harmonic content between flight-test data and wind tunnel measurements performance and loads. Both models incorporate modal characteristics
is evident throughout the remaining correlations presented in this paper. for the structural representation of elastic blades and account for the non-
Figure 5(b) shows that, with respect to the effective blade root location linear inertial terms in the respective equations of motion. The absence
(r/R = 0.144), both simulation models have succeeded in capturing the of elastic coupling in the modes employed for the simulations reported
basic waveform of chordwise loading as measured in the wind tunnel. in this paper does not seem to have affected the overall model accuracy.
Considering the respective correlations with flight-test data, the basic CAMRAD/JA employs a free-wake inflow formulation that models
waveform shape of chordwise bending moment has also been adequately the wake’s self-induced distortion using a discretized three-dimensional
captured. The higher harmonic components of chordwise moment have representation of the vortex wake. As a result, the particular approach is
been accurately predicted, both in terms of amplitude and phase, up to prone to discretization errors and real-time execution is prohibited. The
the third harmonic component (3P) of oscillatory loading. The developed developed aeroelastic rotor model incorporates the finite-state induced-
simulation model has successfully predicted the significant change in the flow model by Peters et al. (Refs. 12, 13), which employs a continuous
phase angle observed in the 1P chordwise loading between wind tunnel wake representation with no associated discretization errors. However,
measurements and flight-test data. This phase angle difference stems the rotor wake is modeled as a skewed cylinder with no wake rollup. The
from the corresponding change in trim conditions. Neither CAMRAD/JA induced flow at the rotor disk is expressed as a function of a number of
nor the present model has been successful in predicting the 8P harmonic flow states determined by the solution of a first-order ordinary matrix
content residing in the waveforms obtained from wind tunnel experiments differential equation. Hence, owing to its simplicity, the particular ap-
and flight tests. proach effectively lends itself to real-time execution. The overall model
Figure 6(b) shows that for the near midspan location (r/R = 0.57), accuracy however is dependent on the employed number of flow states.
neither CAMRAD/JA nor the developed simulation model has accu- Although not included in this paper, a comparative evaluation has
rately predicted the qualitative nature of chordwise loading as measured been carried out to determine the effect of number of flow states on the
in the wind tunnel. Very good correlation is exhibited, however, be- predicted oscillatory blade loads. Simulations have been performed with
tween CAMRAD/JA and the modeling approach of this paper. This is a 15, 33, 47, and 96 flow states. The respective inflow models employ
strong indication that both rotor models essentially exhibit similar lev- aerodynamic terms reaching up to the third, fourth, sixth, and eighth
els of accuracy when it comes to the prediction of oscillatory structural harmonic of oscillatory loading in that order. It has been found that the
blade loads. This observation is valid for higher advance ratios where the predicted loading waveforms converge after the inclusion of only 33 flow
influence of the wake rollup on induced flow is substantially reduced. states. The 8P harmonic content found in the measured data has not been
The observed discrepancies between flight tests and wind tunnel mea- captured, not even with the inclusion of 96 flow states. Hence, since the
surements are attributed to the aforementioned aerodynamic interactions proposed model focuses on harmonic content with frequency up to 4P, a
with the rotor test apparatus. As regards the respective correlations with total of 33 flow states is sufficient to capture the aeroelastic response of
flight tests, Fig. 6(b) suggests fair agreement between simulation and the rotor blades with reasonable accuracy.
measured data considering the harmonic content reaching up to third The structural blade loads correlations presented in this paper have
harmonic of chordwise loading. been carried out for μ ≥ 0.1 where the wake is quickly swept away
Figures 5 and 6(c) present correlations of oscillatory torsional mo- from the rotor by the freestream flow and the roll-up effect is much less
ment. Good agreement is exhibited in Fig. 5(c) for r/R = 0.40 between pronounced. Both rotor models have therefore yielded similar results
predictions carried out with both simulation models and wind tunnel for the test conditions presented in this paper. It can thus be concluded
test results. It can be observed that torsional moment at the particular that the Peters–He induced-flow model can be successfully deployed for
radial location comprises predominantly 1P harmonic content, which the estimation of oscillatory structural blade loads considering operating
has been adequately captured by both rotor models. A noticeable differ- regimes where the wake roll-up effect is diminished and the assumption
ence is observed between wind tunnel experiments and flight-test data, of a skewed cylindrical wake is valid (typically μ ≥ 0.1).
predominantly considering the 1P and 2P loading components. Refer-
ence 33 attributed this discrepancy to the different torsional stiffness Low-frequency aeroelastic rotor–aircraft dynamics
of the pitch control system incorporated by the Bo 105 rotor employed
for the flight tests. This difference in torsional stiffness may have affected This section aims to evaluate the ability of the developed approach
the blade’s natural torsion frequencies. Although the harmonic content to obtain reasonably accurate estimates of S&C derivatives for the low-
suggested by flight tests has been captured in terms of amplitude, there frequency unsteady motion of the Bo 105 helicopter. Analysis has been
is an underprediction in the phase angle of 1P loading which reaches carried out for straight and level flight at V ≈ 41.15 m/s (80 kt) cor-
approximately 60◦ . This may be due to the aforementioned change in responding to an advance ratio of μ ≈ 0.189. Reference 35 reports
torsional frequencies which is not accounted for by the present modeling on the application of system identification methods for the extrac-
approach. tion of six-DOF models in terms of S&C derivatives from flight-test

042006-13
I. GOULOS JOURNAL OF THE AMERICAN HELICOPTER SOCIETY

Table 1. Comparison of estimated stability and control derivatives with values identified
from flight tests

Derivative Unit AFDD (σ) DLR (σ) NAE (σ) Analysis


Xu 1/s −0.038 (0.0036) −0.059 (0.0006) −0.050 (0.0094) −0.0300
Yv 1/s −0.221 (0.011) −0.170 (0.0030) −0.279 (0.005) −0.1084
Zw 1/s −1.187 (0.0555) −0.998 (0.0072) −1.106 (0.013) −0.9441
Mu rad/(m · s) – 0.029 (0.0004) 0.0078 (0.001) 0.0021
Mw rad/(m · s) 0.096 (0.0064) 0.053 (0.0011) 0.0696 (0.0022) 0.1350
Lv rad/(m · s) −0.207 (0.021) −0.271 (0.004) −0.270 (0.0071) −0.2562
Lw rad/(m · s) 0.168 (0.014) 0.116 (0.0035) 0.116 (0.007) 0.1207
Mp 1/s −0.998 (0.066) −0.419 (0.038) −1.414 (0.066) −0.9433
Mq 1/s −4.493 (0.235) −3.496 (0.047) −2.992 (0.074) −4.2184
Lp 1/s −8.779 (0.641) −8.501 (0.011) −7.048 (0.201) −8.2558
Lq 1/s 3.182 (0.624) 3.037 (0.125) 4.454 (0.222) 3.1646
Z θ0 m/(s2 · %) −0.388 (0.02) −0.349 (0.002) −0.337 (0.005) −0.366
M θ1s rad/(s2 · %) 0.098 (0.0043) 0.093 (0.001) 0.0787 (0.0015) 0.078
L θ1c rad/(s2 · %) 0.179 (0.0139) 0.185 (0.0023) 0.1361 (0.0036) 0.074
M θ1c rad/(s2 · %) – −0.009 (0.0008) 0.0106 (0.0013) 0.011
L θ1s rad/(s2 · %) 0.073 (0.0064) 0.024 (0.0025) −0.005 (0.0044) −0.057

measurements. Results have been obtained from various working group blades due to the pitch–roll angular motion of the hub, as well as by the
members that include the U.S Army Aeroflightdynamics Directorate in-plane lift and drag forces generated by aerodynamic loads (Ref. 30).
(AFDD), the German aerospace center DLR (Deutsches Zentrum für These components are thoroughly accounted for through inclusion of
Luft– und Raumfahrt), and the Canadian National Aeronautical Estab- the respective nonlinear forcing terms in the corresponding equations
lishment (NAE; now National Research Council). (Eqs. (35) and (49)). It can be observed that the developed model has
Table 1 presents correlations for some of the most important yielded pitch and roll damping derivatives that are in very good agreement
S&C derivatives of the Bo 105 helicopter, between predictions made with the ones identified by the various working group members.
with the present approach and the respective values identified from flight Accurate prediction of the pitch–roll cross-coupling derivatives (Mp
tests. Comparisons are carried out with identification results reported by and Lq ) from first principles is an issue of great current interest to
AFDD, DLR, and NAE in Ref. 35. The respective standard deviations (σ ) the helicopter flight dynamics community. The contributing factors are
of the identified derivatives are provided in parentheses. These represent predominantly aerodynamic in nature and are related to the effective cur-
the theoretically lowest achievable standard deviation for each identified vature of the wake during pitching or rolling motion. The Zhao–Prasad
derivative (Cramẽr–Rao lower bounds). According to Ref. 35, these val- dynamic wake distortion model, which caters for the wake curvature
ues of standard deviation are unrealistically small and it is recommended effect, has been utilized for the predictions reported in this paper. The
to multiply them by a factor of 5–10 to obtain more reasonable estimates. effectiveness of this approach is evident from the fact that the associ-
It can be observed from the results presented in Table 1 that there ated pitch–roll cross-coupling derivatives (Mp and Lq ) have been fairly
are significant differences between the identified values for most of the accurately predicted. It is noted that simulations carried out without ac-
derivatives of interest. Fairly accurate prediction of the direct force damp- counting for the effects of wake curvature yielded an Lq value of almost
ing derivatives Xu and Yv is demonstrated, especially considering the twice the magnitude and of opposite sign.
differences noted in the identification results along with the respective The control derivatives in Table 1 are given as fractions of the respec-
standard deviations. It is emphasized that Ref. 35 attributed the larger tive control displacements, with 100% of displacement corresponding
values of Yv identified by NAE and AFDD to a lateral speed measurement to full control travel (Ref. 35). Good agreement can be noted between
problem. Good agreement is exhibited between analysis and identifica- analysis and identification considering the thrust derivative Zθ0 . It is also
tion results for the heave damping derivative Zw . evident that the on-axis longitudinal cyclic control derivative (Mθ1s ) has
It can be observed that the speed and incidence static stability deriva- been fairly well estimated. However, the derivative for on-axis lateral
tives, Mu and Mw , are quite under- and overpredicted, respectively. Fur- cyclic control (Lθ1c ) appears to be quite underestimated compared to the
ther investigation on the source of these discrepancies revealed that these identified values. Reference 35 elaborated on the sensitivity of the iden-
derivatives are significantly influenced by the contribution of the aero- tified control derivatives to the selection of equivalent time delays for the
dynamic fuselage pitching moment. Thus, the modeling uncertainties actuator and rotor dynamics. The control derivatives have been identified
described in the section Aeroelastic rotor–aircraft trim performance with assuming a single time delay for each control input that produces the best
regard to the aerodynamic behavior of the fuselage have essentially prop- match for all fuselage responses. However, the present approach extracts
agated to the stability analysis of this section. It is important to notice the the perturbation forces and moments when quasi-steady periodic condi-
good correlation between analysis and identification for the respective tions are obtained with respect to the main rotor multiblade coordinates.
rolling moment derivatives, Lv and Lw . It is noted that the analysis as- Hence, different time delays are assumed not only in the identification
sumes negligible aerodynamic fuselage rolling moment. Therefore, the results yielded by the various group members but also in the process of
aforementioned modeling uncertainties have not affected the predicted analytical derivation. This leads to exacerbated differences between the
values of Lv and Lw , hence the good correlation with the identified identified values themselves, as well as in comparison with the analysis.
values. It is noted that Ref. 35 contains identification results from other working
The pitch and roll damping derivatives, Mq and Lp respectively, are group members that have yielded values for Lθ1c that are very close to
influenced by inertial loads related to the gyroscopic precession of the that estimated analytically.

042006-14
ROTOR BLADE FLEXIBILITY SIMULATION FOR AEROELASTICITY AND FLIGHT DYNAMICS APPLICATIONS 2014

Δθ1c Δθ0 Δθ1s Δθ0t 6 F-L-T 1 F-L-T


10 3 F-L-T Flight Test
Control perturbations

Free-wake/FEA
5 20
from trim (%)

15

Roll rate (deg/s)


0 10
5
−5 0
−5
−10 −10
0 1 2 3 4 5 −15
Time (s) −20
Fig. 7. Control perturbations from trim for the selected near-hover 0 1 2 3 4 5
maneuver; μ ≈ 0.040. (a) Time (s)
15
Even larger differences are observed considering the identified off-

Pitch rate (deg/s)


10
axis control derivatives Mθ1c and Lθ1s . It is worth noticing that derivatives
of different signs are identified by certain group members. This indicates 5
a difficulty in identifying the off-axis control derivatives which is partially
due to the fact that they are highly correlated with their corresponding 0
pitch-roll stability cross-coupling counterparts Mp and Lq (Ref. 35). It
is noted that the predicted value of Mθ1c agrees well in terms of both −5
magnitude and sign with the corresponding value identified by NAE. As −10
regards Lθ1s , although it is agreement with the value identified by NAE 0 1 2 3 4 5
in terms of sign, it is considerably larger in terms of magnitude. This is (b) Time (s)
also attributed to the dependence of the identified control derivatives to
the selection of equivalent time delays. Fig. 8. Influence of number of F–L–T mode shapes on the predicted
Thus, it has been shown that the proposed approach can be used response of the Bo 105 helicopter for the selected control maneuver,
for the extraction of six-DOF flight dynamics models in terms of S&C μ ≈ 0.040: (a) fuselage roll rate (on-axis response) and (b) fuselage
derivatives from first principles, with no empirical corrections involved. pitch rate (off-axis response).

Nonlinear response to pilot control inputs


helicopter to the selected control input schedule. Results are presented
This section aims to assess the ability of the proposed model to for the fuselage roll (on-axis response) and pitch rates (off-axis response)
predict the nonlinear control response characteristics of the Bo 105 heli- in Figs. 8(a) and 8(b), respectively. Simulations have been carried out
copter. The effects of modeling fidelity in terms of modal representation incorporating one, three, and six elastically uncoupled modes for each
of elastic blades and finite-state inflow dynamics are also investigated. DOF. A total of 62 flow states are used in the respective dynamic wake
The obtained predictions are compared with flight-test data reported in model.
Ref. 36, as well as with results from simulations carried out in Ref. 14 Very good correlation in terms of both amplitude and phase can be
(free wake/FEA) using a high-fidelity, non-real-time flight dynamics observed in Fig. 8(a) between predictions and measured data, consid-
simulation code. ering the on-axis fuselage response. Correlation is best for simulations
It is once again noted that the simulation methodology of Ref. 14 performed employing six F–L–T modes in the rotor dynamics model.
utilizes elastically and inertially coupled modal properties obtained from Certain discrepancies are observed for t ≥ 4 s between predictions car-
nonlinear FEA. A relaxation-type free-wake inflow model is deployed ried out utilizing different levels of modeling fidelity in the employed
(Ref. 16), which accounts for the effects of wake curvature due to maneu- modal content. Excellent agreement is also demonstrated between the
vering flight. The present approach does not account for elastic coupling present modeling approach and the high-fidelity flight dynamics code
in the employed modal properties. However, inertial coupling is thor- of Ref. 14 when utilizing three F–L–T mode shapes. This is due to the
oughly retained through proper inclusion of the corresponding forcing fact that the respective modal content used in Ref. 14 also reaches up to
terms in the kinematic equations for elastic blade motion (Eqs. (34) and the third mode of elastic blade flap motion. The aforementioned results
(35)). The proposed approach caters for wake curvature effects due to essentially indicate that inclusion of high-frequency modal content in
maneuvering flight through augmenting the employed Peters–He finite- rotor dynamics may considerably improve the prediction of the on-axis
state inflow model (Refs. 12,13) with the dynamic wake distortion model response when one also accounts for high-frequency aerodynamic terms
developed by Zhao–Prasad in Ref. 26. in the employed forcing functions.
A representative maneuver has been selected to be simulated at near- Figure 8(b) demonstrates that the developed modeling approach has
hover conditions (μ ≈ 0.040). Figure 7 presents the time variations also been successful in predicting the off-axis response of the fuselage
of control deflections from trim as percentages of full control travel. with reasonable accuracy. Good correlation can be observed between
The respective control schedule comprises predominantly deflections in flight-test data and predictions carried out using six F–L–T modes.
lateral cyclic θ1c . The specific maneuver is part of a larger series of flight Good agreement can also be noted between the present model and the
tests conducted for system identification purposes (Ref. 36). flight dynamics code of Ref. 14. Significant deviations are observed for
Figure 8 presents the effect of rotor dynamics modeling fidelity in t ≥ 2 s between predictions carried out utilizing different levels of mod-
terms of employed modal content, on the predicted response of the Bo 105 eling fidelity. Once again, agreement with flight-test data is best when

042006-15
I. GOULOS JOURNAL OF THE AMERICAN HELICOPTER SOCIETY

47 states 15 states 1 F-L-T 3 F-L-T 6 F-L-T

Computational time ratio (nd)


96 states No distortion 5
Flight test Free-wake/FEA
20 4 40 BEs
15 20 BEs
Roll rate (deg/s)

3
10
5 2
0
−5 1
−10 Real-time threshold
0
−15 0 15 30 45 60 75 90
−20 Number of flow states
0 1 2 3 4 5
Fig. 10. Influence of number of flow states, F–L–T modes and num-
(a) Time (s) ber of blade elements (BEs) on the computational time ratio for a
15 personal computer with a 2.3-GHz CPU and 4 GB of RAM.
Pitch rate (deg/s)

10
considering the off-axis response as demonstrated in Fig. 9(b). The
5
predictions from simulations carried out using over 33 flow states are
0 considered practically identical. It can therefore be concluded that sole
inclusion of higher harmonic inflow content does not result in improved
−5 off-axis response predictions. This is in accordance with the observations
reported in Ref. 9.
−10 The significance of including dynamic wake distortion effects when
0 1 2 3 4 5 it comes to accurate prediction of the fuselage off-axis response is evi-
(b) Time (s) dent from the results presented in Fig. 9(b). Considering the simulations
conducted excluding the aforementioned effects, the predicted off-axis
Fig. 9. Influence of number of flow states on the predicted response of
response correlates poorly with flight-test data. This observation is in
the Bo 105 helicopter for the selected control maneuver, μ ≈ 0.040:
agreement with typical trends reported by other investigators (Ref. 26).
(a) fuselage roll rate (on-axis response) and (b) fuselage pitch rate
A small improvement in the on-axis response can also be observed in
(off-axis response).
Fig. 9(a), when accounting for wake distortion. It is believed that im-
provement of the off-axis response has inevitably affected the predicted
on-axis behavior in a favorable manner due the substantial cross-coupling
incorporating higher frequency modal content. The observed trends sug- involved (Ref. 35). This effect is also evident in Fig. 8 where the improved
gest that the off-axis fuselage response is a highly coupled aeroelastic off-axis behavior due to inclusion of higher harmonic modal content has
problem, particularly for the case of the Bo 105 helicopter. As such, ac- also resulted in better prediction of the on-axis response for t ≥ 4 s.
curate predictions essentially entail inclusion of higher frequency modal
content in rotor dynamics. This observation is in accordance with the con- Assessment of computational time requirements
clusions reached from previous investigations on the nonlinear control
response of the Bo 105 helicopter (Ref. 14). To evaluate the suitability of the proposed model for real-time flight
It is important to notice that, with regard to the first second of flight simulation, the computational time requirements associated with nonlin-
simulation where the control deflections from trim are minimum, the ear control response analysis have been thoroughly assessed. An effective
respective roll and pitch rates are practically zero. This indicates that the parameter is defined, namely the computational time ratio, as the com-
employed trim procedure has been successful in obtaining a trim state putational time required per control response simulation time. Hence, a
which is compatible with nonlinear control response simulation. This computational time ratio less that unity (real-time threshold) essentially
confirms the validity of the numerical approach employed for finding allows for the nonlinear mathematical model to be deployed in real time.
rotor trim. However, this is not the case for the flight dynamics code Figure 10 presents the influence of number of flow states, F–L–
of Ref. 14 where substantial deviations from trim are observed with T modes, and number of blade elements on the computational time
respect to the first second of simulation. This has been attributed to a ratio. Time measurements have been conducted on a personal computer
deficiency of the corresponding approach for finding rotor trim and has equipped with 4 gigabytes (GB) of random access memory (RAM) and
been elaborated in Ref. 14. a central processing unit (CPU) operating at 2.3 GHz. A time step of
Figure 9 presents the effect of finite-state inflow modeling fidelity in t = 2 × 10−3 s has been employed for time-marching resulting in an
terms of employed number of flow states on the predicted response of the azimuth step of roughly 5◦ , which is sufficient for accurate estimation of
Bo 105 helicopter. Simulations have been carried out employing 15, 33, aeroelastic blade loads.
47, and 96 flow states. A total of six F–L–T mode shapes are incorporated It can be observed that computational times change slightly with the
in the respective rotor dynamics model. Simulations are also presented inclusion of up to six deformation modes for each DOF. Inclusion of even
for 96 flow states, where the effects of dynamic wake curvature due to higher harmonic content is unlikely to improve correlations of oscillatory
maneuvering flight conditions have been excluded (no distortion). blade loads or nonlinear control response. The influence of the number
Figure 9(a) shows that, with respect to the predicted on-axis response, of flow states is relatively moderate if one considers that computational
there are no observable differences between simulations carried out uti- time almost triples for increasing the number of flow states by a factor
lizing different levels of modeling fidelity. This observation is also valid of 30. Figure 10 also shows that the computational time ratio is almost

042006-16
ROTOR BLADE FLEXIBILITY SIMULATION FOR AEROELASTICITY AND FLIGHT DYNAMICS APPLICATIONS 2014

proportional to the number of blade elements. It is emphasized that a content as well as wake curvature effects. Sole inclusion of higher har-
number of 20–25 blade elements have been found sufficient with regard monic inflow dynamics does not improve off-axis response predictions.
to both flight simulation as well as prediction of oscillatory structural 8) The developed mathematical approach is applicable to real-time
blade loads. flight simulation on a modern personal computer, employing sufficient
Figure 10 shows that, regarding the machine used for this work, real- fidelity for simultaneous prediction of oscillatory blade loads.
time simulation can be achieved using up to 33 flow states, 3 F–L–T Thus, the proposed mathematical formulation enables transition from
modes, and 20 blade elements. This results in an aeroelastic rotor model classical rigid blade modeling to a complete framework for real-time
with sufficient fidelity for nonlinear control response analysis as well as aeroelasticity simulation, without resorting to comprehensive analysis
prediction of oscillatory structural blade loads. The developed approach methods. It is compatible with every rotor inflow and blade element
is expected to run much faster on higher-end machines, thus allowing aerodynamics theory that can operate in the time domain. It is readily
real-time execution using model configurations with substantially higher available for implementation in existing helicopter flight dynamics codes
fidelity. utilizing the standard fuselage axes reference frame, without any change
in the offered expressions.

Summary and Conclusions


References
A mathematical model has been offered for the simulation of rotor
1
blade flexibility in real-time helicopter flight dynamics applications, that Howlett, J. J., “UH-60A Black Hawk Engineering Simulation Pro-
also employs sufficient fidelity for prediction of oscillatory structural gram: Volume I–Mathematical Model,” NASA CR 166309, 1981.
2
blade loads. Lagrangian kinematics have been utilized for the estima- Brown, R. E., “Rotor Wake Modeling for Flight Dynamic Simulation
tion of natural vibration characteristics of rotor blades. A matrix/vector- of Helicopters,” AIAA Journal, Vol. 38, (1), January 2000, pp. 57–63.
3
based formulation has been developed for the treatment of elastic blade Johnson, W., “Rotorcraft Aerodynamics Models for a Comprehen-
kinematics in the time domain. The developed approach has been cou- sive Analysis,” American Helicopter Society 54th Annual Forum Pro-
pled with a finite-state induced-flow model, a dynamic wake distortion ceedings, Washington, DC, May 20–22, 1998.
4
model, and an unsteady blade element aerodynamics model. The inte- Brown, R. E., and Houston, S. S., “Comparison of Induced Velocity
grated methodology has been deployed to investigate trim performance, Models for Helicopter Flight Mechanics,” Journal of Aircraft, Vol. 37,
stability and control derivatives, oscillatory blade loads, and nonlinear (4), August 2000, pp. 623–629.
5
control response for the Bo 105 helicopter. Peters, D. A., and HaQuang, N., “Dynamic Inflow for Practical
The general conclusions arising from the completion of this work can Applications,” Journal of the American Helicopter Society, Vol. 33, (4),
be summarized as follows: October 1988, pp. 64–68.
6
1) Good agreement has been found between Lagrangian predictions Shupe, N., “A Study on the Dynamic Motions of Hingeless Rotored
and nonlinear analysis carried out using Boeing–Vertol Y-71, considering Helicopters,” U.S. Army Electronics Command, TR ECOM-3323, 1970.
7
the rotor blade’s resonant frequencies at nominal rotor speed. Lewis, W. D., “An Aeroelastic Model Structure Investigation for a
2) Good predictive qualities have been found with regard to trim con- Manned Real-Time Rotorcraft Simulation,” American Helicopter Soci-
trol angles for straight and level flight. A small underprediction of lateral ety 49th Annual Forum Proceedings, St. Louis, MO, May 19–21, 1993.
8
cyclic has been noted at very low advance ratios that has been attributed Sturisky, S. H., and Schrage, D. P., “System Identification Valida-
to the absence of wake rollup in the Peters–He model. Rotor power re- tion of an AH-64 Aeroelastic Simulation Model,” American Helicopter
quirement and collective pitch angle have been found to be considerably Society 49th Annual Forum Proceedings, St. Louis, MO, May 19–21,
underestimated at higher advance ratios, that has been attributed to the 1993.
9
underprediction of aerodynamic fuselage drag. Turnour, S. R., and Celi, R., “Modeling of Flexible Rotor Blades
3) The devised formulation has been successful in predicting the for Helicopter Flight Dynamics Applications,” Journal of the American
low-frequency components of oscillatory structural blade loads as mea- Helicopter Society, Vol. 41, (1), January 1996, pp. 52–66.
10
sured in wind tunnel and flight tests. The model has been unable to Spence, A. M., and Celi, R., “Coupled Rotor–Fuselage Dynamics
capture the higher frequency components (8P) residing the measured and Aeroelasticity in Turning Flight,” Journal of the American Helicopter
loading waveforms. The proposed approach and the comprehensive code Society, Vol. 40, (1), January 1995, pp. 47–58.
11
CAMRAD/JA have essentially exhibited similar levels of accuracy. Kim, F. D., Celi, R., and Tischler, M. B., “Forward Flight Trim
4) It has been shown that the Peters–He finite-state induced-flow and Frequency Response Validation of a Helicopter Simulation Model,”
model can be successfully deployed for the estimation of oscillatory Journal of Aircraft, Vol. 30, (6), November 1993, pp. 854–863.
12
structural blade loads considering operating regimes where the wake roll- Peters, D. A., Boyd, D. D., and He, C. J., “Finite-State Induced-Flow
up effect is less pronounced and the assumption of a skewed cylindrical Model for Rotors in Hover and Forward Flight,” Journal of the American
wake is valid. Helicopter Society, Vol. 34, (4), October 1989, pp. 5–17.
13
5) The present modeling approach has been successful in obtaining Peters, D. A., and He, C. J., “Correlation of Measured Induced
reasonably accurate estimates of stability and control derivatives from Velocities with a Finite-State Wake Model,” Journal of the American
first principles. Certain discrepancies have been observed in the estimated Helicopter Society, Vol. 36, (3), July 1991, pp. 59–70.
14
values of the speed and incidence static stability derivatives, that have Theodore, C., “ Helicopter Flight Dynamics Simulation with Refined
been attributed to modeling uncertainties related to the aerodynamic Aerodynamic Modeling,” Ph.D. Thesis, University of Maryland, College
behavior of the fuselage. Park, MD, 2000.
15
6) The proposed approach has successfully predicted the on-axis as Kim, F. D., Celi, R., and Tischler, M. B., “High-Order State Space
well as off-axis response of the Bo 105 helicopter to pilot control inputs Simulation Models of Helicopter Flight Mechanics,” Journal of the
for a representative maneuver at near-hover conditions. American Helicopter Society, Vol. 38, (4), October 1993, pp. 16–27.
16
7) For the case of the Bo 105 helicopter, accurate prediction of the Bagai, A., and Leishman, J. G., “Rotor Free-Wake Modeling Using a
off-axis fuselage response entails inclusion of higher frequency modal Pseudo-Implicit Technique - Including Comparisons with Experimental

042006-17
I. GOULOS JOURNAL OF THE AMERICAN HELICOPTER SOCIETY

Data,” Journal of the American Helicopter Society, Vol. 40, (3), April 27
Leishman, J. G., and Beddoes, T. S., “A Semi-Empirical Model for
1995, pp. 29–41. Dynamic Stall,” Journal of the American Helicopter Society, Vol. 34,
17
Piziali, R. A., and DuWaldt, F. A., “A Method for Computing Rotary (3), July 1989, pp. 3–17.
Wing Airload Distribution in Forward Flight,” TCREC TR 62-44, 1962. 28
Bisplinghoff, R. L., Ashley, H., and Halfman, R. L., Aeroelasticity,
18
Piziali, R., “Method for the Solution of the Aeroelastic Response Dover Publications, Mineola, NY, 1996.
Problem for Rotating Wings,” Journal of Sound and Vibration, Vol. 4, 29
Goulos, I., “Simulation Framework Development for the Multidisci-
(3), November 1966, pp. 445–489. plinary Optimization of Rotorcraft,” Ph.D. thesis, School of Engineering,
19
Yeo, H., and Johnson, W., “Assessment of Comprehensive Analy- Cranfield University, Bedfordshire, UK, 2012.
sis Calculation of Airloads on Helicopter Rotors,” Journal of Aircraft, 30
Bramwell, A. R. S., Done, G., and Balmford, D., Bramwell’s Heli-
Vol. 42, (5), September 2005, pp. 1218–1228. copter Dynamics, Butterworth-Heinemann, Oxford, UK, 2001.
20 31
Yeo, H., and Johnson, W., “Prediction of Rotor Structural Loads with Padfield, G. D., Helicopter Flight Dynamics, Blackwell Publishing,
Comprehensive Analysis,” Journal of the American Helicopter Society, Oxford, UK, 2007.
Vol. 53, (2), April 2008, pp. 193–209. 32
Staley, J. A., “Validation of Rotorcraft Flight Simulation Program
21
Johnson, W., “General Free Wake Geometry Calculation for Wings through Correlation with Flight Data for Soft-in-Plane Hingeless Rotor,”
and Rotors,” American Helicopter Society 51st Annual Forum Proceed- USAAMRDL-TR-75-50, 1976.
ings, Fort Worth, TX, May 9–11, 1995. 33
Peterson, R. L., Maier, T., Langer, H. J., and Tränapp, N., “Corre-
22
Johnson, W., “Rotorcraft Dynamics Models for a Comprehensive lation of Wind Tunnel and Flight Test Results of a Full-Scale Hingeless
Analysis,” American Helicopter Society 54th Annual Forum Proceed- Rotor,” Proceedings of the American Helicopter Society Aeromechanics
ings, Washington, DC, May 20–22, 1998. Specialist Conference, San Francisco, CA, January 19–21, 1994.
23 34
Johnson, W., “A History of Rotorcraft Comprehensive Analyses,” Padfield, G. D., Basset, P. M., Dequin, A. M., von Grunhagen, W.,
American Helicopter Society 69th Annual Forum Proceedings, Phoenix, Haddon, D., Haverdings, H., Kampa, K., and McCallum, A. T., “Predict-
AZ, May 21–23, 2013. ing Rotorcraft Flying Qualities through Simulation Modelling. A Review
24
Horn, J. F., Bridges, D. O., A.Wachspress, D., and Rani, S. L., “Im- of Key Results from Garteur AG06,” Proceedings of the 22nd European
plementation of a Free-Vortex Wake Model in Real-Time Simulation of Rotorcraft Forum, Brighton, UK, September 17–19, 1996.
Rotorcraft,” Journal of Aerospace Computing, Information, and Com- 35
AGARD, “Rotorcraft System Identification,” Advisory Group for
munication, Vol. 3, (3), March 2006, pp. 93–114. Aerospace Research and Development, AGARD LS 178, 1991.
25 36
Bhagwat, M. J., and Leishman, J. G., “Stability, Consistency and Kaletka, J., and Gimonet, B., “Identification of Extended Models
Convergence of Time-Marching Free-Vortex Rotor Wake Algorithms,” from BO-105 Flight Test Data for Hover Flight Condition,” Proceed-
Journal of the American Helicopter Society, Vol. 46, (1), January 2001, ings of the 21st European Rotorcraft Forum, Saint Petersburg, Russia,
pp. 59–71. August 30–September 1, 1995.
26 37
Zhao, J., Prasad, J. V. R., and Peters, D. A., “Rotor Dynamic Wake Johnson, W., CAMRAD/JA, “A Comprehensive Analytical Model of
Distortion Model for Helicopter Maneuvering Flight,” Journal of the Rotorcraft Aerodynamics and Dynamics Vol. I: Theory Manual,” John-
American Helicopter Society, Vol. 49, (4), October 2004, pp. 414–424. son Aeronautics, Palo Alto, CA, 1988.

042006-18

View publication stats

You might also like