You are on page 1of 30

Application of CFD/CSD to Rotor Aeroelastic Stability in Forward Flight

Hyeonsoo Yeo Mark Potsdam Robert A. Ormiston


Aeroflightdynamics Directorate (AMRDEC)
U.S. Army Research, Development, and Engineering Command
Ames Research Center, Moffett Field, California

Abstract

CFD/CSD coupling was successfully applied to the rotor aeroelastic stability problem to calculate lead-lag
regressing mode damping of a hingeless rotor in forward flight. A direct time domain numerical integration of
the equations in response to suitable excitation was solved using a tight CFD/CSD coupling. Three different
excitation methods − lateral cyclic, rotating pitch, and blade tip lead-lag force excitations − were investigated
to provide suitable blade transient responses. The free decay transient response time histories are post-
processed using the moving-block method to determine the damping as a function of the rotor operating
conditions. A number of CFD/CSD implementation issues were addressed to achieve practical aeroelastic
stability results with limited computational run-time. Coupled CFD/CSD analysis results are compared
with the experimentally measured stability data obtained for a 7.5-ft dia Mach-scale hingeless rotor model
as well as stability predictions using traditional rotor aerodynamic modeling representations available in
comprehensive analysis (RCAS). The coupled RCAS/OVERFLOW predictions agreed more closely with the
experimental lead-lag damping measurements than RCAS predictions based on conventional aerodynamic
methods.

Introduction configuration topology. For rotorcraft aerodynamics, the


equally complex airloads and wake flowfield phenomena
This paper will address the development and application of have seen enormous advances in recent years. Keyed by
a new rotorcraft CFD/CSD method for the prediction of decades-long development of computational fluid dynamics
aeroelastic stability of rotor blades in forward flight. (CFD), the relatively recent development of rotorcraft
aeroelastic coupling of computational structural dynamics
Accurate prediction of rotorcraft aeroelastic and (CSD) and CFD has opened the door to consistent treatment
aeromechanical stability is essential for the rational and of the full complexity of rotorcraft aeromechanics. These
successful design of all types of rotorcraft. Rotorcraft recent developments have included the steady-state trim
history includes a number of development programs where problem using CFD/CSD loose coupling (LC) [1] and
serious schedule and cost over-runs, or even program the transient maneuver problem using the CFD/CSD tight
cancellation, resulted from the inability to anticipate and coupling (TC) [2]. These two fundamental problems, trim
accurately predict key rotorcraft instabilities. Cost effective and maneuver, treat the steady-state and transient response
and timely development of future rotorcraft will require the problems to yield airloads and structural loads in steady-state
most accurate and reliable prediction of aeroelastic stability and maneuvering flight conditions respectively.
to insure optimum mission performance unconstrained by
design compromises imposed by an inability to address The third fundamental rotorcraft aeromechanics problem,
the challenges of complex rotary wing aeroelastic and aeroelastic stability, has yet to be addressed with CFD/CSD
aeromechanical stability phenomena. methods. This broad subject encompasses rotor blade
aeroelastic stability and coupled rotor-body aeromechanical
Rotorcraft aeromechanics prediction methodology has stability including articulated rotor bending-torsion flutter,
advanced significantly in recent years. The complex hingeless/bearingless rotor flap-lag-torsion stability,
structural dynamics of a fully coupled rotorcraft with ground/air-resonance, and tiltrotor stability.
nonlinear elastic rotating blades are now amenable to
sophisticated structural analysis capabilities based on multi- The unique nature of rotorcraft introduces aerodynamic
body finite-element modeling for rotorcraft of arbitrary characteristics that are significantly different from fixed-
wing aircraft. The key difference is the radial and
Presented at the American Helicopter Society 66th Annual Forum, Phoenix,
Arizona, May 11 - 13, 2010. This material is declared a work of the U.S. azimuthal variation of velocity in hover and forward
Government and is not subject to copyright protection in the U.S. flight. This fundamentally complicates the 3-D unsteady

1
aerodynamics of rotorcraft and necessitates the use of Several transient excitation schemes are investigated for
approximate aerodynamics methods not required in fixed- effectiveness and efficiency. The free decay transient response
wing aeroelasticity. Beyond this fundamental difference, time histories are post-processed using the moving-block
rotorcraft also experience separated flow (stall), reverse flow, method to determine the damping as a function of the
and the rotor wake flowfield. rotor operating conditions. Results are compared with the
experimentally measured stability data as well as stability
Rotorcraft aeroelastic stability prediction is further predictions using traditional rotor aerodynamic modeling
complicated by the nature of the problem itself. Unlike representations available in RCAS.
the trim and maneuver problems, where the nonlinear
equations are necessarily solved by direct time domain
integration, the most useful stability predictions are obtained ADM Aeroelastic Stability
from eigensolutions of linear(ized) equations, either constant
coefficient equations for hover conditions or periodic ADM Stability Test
equations in forward flight, the latter solved by Floquet
The ADM experiments were conducted at AFDD in the 1990s
methods. The eigensolutions directly provide stability
to provide a high quality aeroelastic stability database for
characteristics, frequency and damping, however, the linear
validation of aeromechanics prediction methodology for a
approximations compromise accuracy. These approximations
typical hingeless rotor, with straight and swept-tip blades,
may be avoided by direct time domain numerical integration
operating at representative tip Mach numbers. Figure 1 shows
of the equations in response to suitable excitation, although
the two rotors installed in the wind tunnel on the Rotor Test
this requires post-processing of the transient response to
Rig (RTR), and Fig. 2 shows the straight and swept-tip rotor
provide frequency and damping.
blades.
Incorporating CFD aerodynamics modeling with a time
The experimental model was designed to represent a typical
domain stability approach should provide the most accurate
soft-inplane hingeless rotor with low torsional stiffness
method for predicting rotorcraft aeroelastic stability. While
to emphasize the fundamental aeroelastic and structural
CFD has been used for fixed-wing aeroelastic stability
couplings resulting from nonlinear bending-torsion coupling
prediction, it has not been applied to the rotorcraft stability
of hingeless rotor blades. Uniform, untwisted blades, without
problem, and there are a number of associated challenges
chordwise mass or elastic center offsets, and the use of the
that influence the accuracy and practicality of the method.
common NACA 0012 airfoil were intended to limit the focus
This paper will address these challenges and develop and
to basic hingeless rotor aeroelastic stability characteristics.
apply the CFD/CSD coupled methodology to a typical
rotorcraft aeroelastic stability problem. Predicted stability The natural frequencies of the straight and swept-tip blade
characteristics will be compared with experimental data were calculated with RCAS and are shown in Fig. 3. The
obtained for a 7.5-ft dia Mach-scale hingeless rotor model, the fundamental flap, lag and torsion frequencies calculated at the
Advanced Dynamic Model (ADM) extensively tested in hover operating speed of 1700 RPM were 1.13, 0.72 and 2.59/rev for
and in forward flight in the U.S. Army Aeroflightdynamics the straight blade and 1.12, 0.66 and 2.75/rev for the swept-tip
Directorate (AFDD) 7- by 10-Foot wind tunnel [3, 4]. blade.
During the test, regressing lag mode damping was measured
Overview of Aeroelastic Methodology Approach in hover and forward flight conditions, with a variety of shaft
angles, collective pitch angles, precones and advance ratios.
A general outline of the present CFD/CSD approach will The excitation of the regressing lag mode and its damping
provide an indication of the scope and content of the paper. measurement were achieved by using capability built into
The development will focus on the aeroelastic stability of the the test stand. The rotor stand had high-speed hydraulic
ADM, an isolated (no rotor-fuselage coupling) rotor having actuators placed in series with each of the three high authority
four hingeless (cantilevered) rotor blades flexible in bending rotor trim control electric actuators. For these tests, only
and torsion. Two blade configurations, straight and swept- one of the three hydraulic actuators was used to provide
tip, with and without precone, are studied. The structural excitation at the regressing lag mode frequency through the
and aerodynamic characteristics are modeled in detail in the fixed system swashplate. Initially, low amplitude cyclic
Army Rotorcraft Comprehensive Analysis System (RCAS) excitation was applied and the frequency was adjusted until
and the blade geometry is modeled in the OVERFLOW 2 CFD the lead-lag response reached a maximum. The amplitude
code. The two codes are coupled through a fluid structure of excitation was then adjusted to give lead-lag response just
interface for both loose and tight CFD/CSD coupling. The below the structural limit. The excitation was shut off and
rotor is trimmed in forward flight using loose coupling, 2 seconds of data were acquired in decay. The measured
and suitable maneuver excitations are applied using tight four individual blade chordwise bending moments at 12%R
coupling to initiate transient responses (maneuver analysis). were transformed into the fixed frame using the multiblade

2
coordinate transformation. Individual rotating blade and for determining rotorcraft aeroelastic and aeromechanical
non-rotating multiblade coordinate time history data were stability is to obtain eigenvalues of linearized equations to
analyzed for modal damping and frequency using the moving- directly obtain frequency and damping. By introducing
block method. coupled CFD/CSD methodology, a time-domain approach is
necessary. Since stability characteristics must be obtained
ADM Literature by post-processing transient response (maneuver) time history
The AFDD experimental investigations of the ADM rotor data, numerous rotor revolutions will typically be needed for
provided an extensive hingeless rotor aeroelastic stability accurate post-processing, and significant computational effort
database. This data set was widely used by numerous may easily result. Extensive CFD/CSD aeroelastic stability
researchers for comparison and validation of a variety investigations will, therefore, not be practical unless efficient
of specialized research and comprehensive aeromechanics excitation and post-processing methodologies are employed.
codes [3–7]. A brief overview of this research is presented There are a number of system identification methods and
here in order to place this data in context for the present techniques that have been developed and employed for
work. This paper primarily addresses the development parameter identification of flight dynamics, flutter, and
of aeroelastic stability prediction methodology, specifically aeroelastic stability characteristics of fixed- and rotary-wing
exploiting the opportunities available from accurate coupled aircraft, e.g., Ref. 8. Traditionally they were developed to
CFD/CSD modeling. The methodology will be illustrated post-process flight-test and wind-tunnel experimental data
with stability predictions of the ADM that will also be but they are becoming increasingly applied to computational
compared with experimental measurements. That said, the aeroelasticity (CAE) data. Within the rotorcraft community,
published literature on the ADM showed that the principal the moving-block method [9, 10], has been widely used
stability characteristics were generally well predicted by the for many decades since it has characteristics that are
more sophisticated prediction methods, however, several key advantageous for rotorcraft applications. In view of this
experimental characteristics were not well predicted. These quite successful experience, the moving-block method has
are discussed in detail in the literature and summarized in been adopted for the present investigation. In applications
Fig. 4 to show the lead-lag regressing mode damping as a with experimental data, measurement noise may significantly
function of advance ratio for the swept-tip blade (no precone, compromise the accuracy of damping estimates; one of the
0◦ angle of attack) and the straight blade (2◦ precone, −6◦ benefits of the present computational application is that the
angle of attack). noise in the numerical data (from several sources) is typically
The swept-tip and straight blade configurations differ far less than that encountered in even the most carefully
primarily in that the lead-lag damping of the former is designed experiments. And some effort has been devoted
generally less sensitive to operating conditions than the latter. to tailoring excitation inputs to achieve transient responses
The swept-tip blade damping exhibits a pronounced “up- suitable for accurate stability post-processing with minimum
down-up” variation in damping with advance ratio, especially CFD/CSD computational cost.
for the higher collective pitch angles. This feature was not It should be noted here that there are many other system
well predicted in previous investigations using current codes identification techniques that may be well suited to the
based on conventional rotorcraft lifting-line aerodynamics. CFD/CSD aeroelastic stability post-processing problem.
These codes typically used approximate constant coefficient Bauchau and Wang [11] surveyed a variety of techniques
or periodic Floquet analysis eigensolutions of the linearized applicable to numerically derived data and proposed refined
equations of motion. A dominant feature of the straight techniques for two related approaches. These were for
blade data is a reduction of lead-lag damping with increasing two closely related algorithms based on a partial Floquet
advance ratio and decreasing collective pitch angle, implying approach and an autoregressive approach. The methods
lead-lag instability at advance ratios slightly above the were applied to two typical rotorcraft problems, aeroelastic
experimental test conditions. Again, the large reduction in stability of a four-bladed articulated rotor and aeromechanical
damping was not well predicted with current comprehensive stability of a three-bladed wind turbine on a flexible tower.
analyses. Despite numerous attempts to explore a variety Transient data were generated with conventional aerodynamic
of structural, aerodynamic modeling, and analysis/solution methods and used to identify frequency and damping of
methods, the principal prediction deficiencies summarized a number of system modes. In a typical application of
above have not been satisfactorily resolved to date. CFD/CSD modeling to CAE for two fixed-wing aircraft
problems, McNamara and Friedmann [12] compared the
Aeroelastic Stability Identification moving-block method, a least-squares curve fitting method,
and an autoregressive system identification method. The three
Background methods gave similar frequency and damping results but the
system identification autoregressive method produced a 75%
As noted in the introduction, the preferred methodology reduction in computational cost by reducing the CFD time

3
history record length. pitch excitation. The transient time history data are analyzed
for damping estimation using the moving-block analysis
While the present investigation demonstrated the feasibility technique. Figure 6 shows a screenshot of the moving-block
of the moving-block method for applications of CFD/CSD analysis software used in the current study. The bottom figure
methods to rotorcraft aeroelastic stability, further study will is the decay response of the chord bending moments and
be needed to identify the optimum methodology for specific the top figure is the natural logarithm of the moving-block
types of problems. It is anticipated that the suitability of any function. A slope of the top curve is calculated using a least-
method will be strongly application dependent. squares fit, and damping is estimated.
Moving-Block Method
Transient Excitation and Related Dynamics
The moving-block analysis is a method of analyzing
One aspect of the computational methodologies employed
a transient time history to obtain modal damping and
in the paper is the choice of rotor excitation used to
frequency [9, 10]. A single mode’s transient response is
generate rotor blade transient responses. There are
expressed as
several considerations to be addressed including: CFD
run time, signal to noise ratio, selectivity/purity of modes
f(t) = A eσt sin(ωt + φ) excited (avoiding unwanted modes), consistency with the
experimental procedures, and compatibility with stability
where σ is damping exponent, ω is frequency, and φ is phase post-processing.
angle. The Fourier transformation of this function from τ to
For practical aeroelastic stability prediction with CFD/CSD,
τ + T is
it is clearly desirable to generate transient responses with a
Z τ+T minimum of CFD/CSD computation time. For the moving-
F(ω, τ) = A eσt sin(ωt + φ)e−iωt dt block post-processing method, accuracy increases with the
τ length of the transient record so there is a tradeoff between
record length and the quality of the transient response.
The amplitude of the function F(ω, τ) is referred to as the Experimental transient data includes noise, and the moving-
moving-block function. For small damping, the natural block accuracy typically improves when the signal-to-noise
logarithm of the moving-block function is ratio is improved. In the case of computational data,
small amounts of “CFD noise” may be present, but more
Ln F(ω, τ) = −ζωτ + 1/2sin 2(ωτ + φ) + constant importantly it is desirable that the data have a high degree
of modal purity; ideally the excitation input will generate
only the rotor mode(s) of interest. For the experimental
where σ = -ζω. Thus a plot of Ln F(ω, τ) as a function of τ
comparisons addressed herein, the mode of interest is known.
will give a straight line with a slope of -ζω and an additional
In a general application where the critical mode is unknown,
oscillating component at twice the analysis frequency.
or multiple modes are of interest, the choice of transient
The procedure to obtain damping from moving-block analysis excitation may be less obvious.
is as follows. The first step is to determine the frequency
of interest using a fast Fourier transformation on the entire As described earlier, the ADM experiment excitations were
data. Then a block size is selected and the natural logarithm fixed system swashplate oscillations and a 2-sec transient
of the moving-block function is calculated for τ = 0. The decay was recorded for post-processing. This corresponds
block is then shifted one sample point at a time and Ln F(ω, τ) to about 56 rotor revolutions which would be impractical to
recomputed. A straight line is fitted to the resulting log plot duplicate with an analogous computational approach. Several
with a least squares fit and the damping values is estimated alternative excitations were investigated to provide suitable
from the slope of the line. blade transient responses. These included rotor blade pitch
inputs and an external force input. The blade pitch inputs
A sample moving-block analysis procedure is illustrated for included a lateral cyclic pitch (swashplate) as employed in
hover with a 12◦ collective pitch angle case. Figure 5(a) the ADM experiment as well as pitch excitation individually
shows rotor blade pitch change at the pitch bearing. The applied to each blade in the rotating system. The lateral
transient analysis begins from the trimmed solution and cyclic excitation was input for one oscillation cycle at the
continues unperturbed for five rotor revolutions. A small 0.3◦ rotor lag regressing frequency − approximately three rotor
cyclic excitation is then provided for 20 revolutions. The revolutions in duration. The rotating pitch input consisted of
excitation is terminated and the analysis continues for 35 a linearly increasing oscillation at the lag frequency phased
more revolutions. Figure 5(b) shows the corresponding blade to represent a combination of lateral and longitudinal cyclic
chordwise bending moments at 12%R. The chord bending pitch at the lead-lag regressing mode frequency, simulating a
moments slowly respond to the airload changes due to the linearly increasing regressing whirl excitation.

4
Given that the blade pitch excitations produce blade lead- blade flap bending moments, and Fig. 10(c) shows the rotor
lag response indirectly as a result of the large direct blade lead-lag bending moment response expressed in the multi-
flap response, it was of interest to include an excitation that blade coordinates. These data were obtained from the
would excite blade lead-lag responses directly, with a smaller RCAS/OVERFLOW computations performed with the swept-
blade flap response. With a goal of shortening the overall tip blade in forward flight at µ = 0.3. All of the responses are
duration of the transient excitation, the objective was to the perturbation responses obtained by removing the trimmed
quickly produce large lead-lag excitations without excessive steady-state periodic response (LC analysis) from the total
flap motions. This was expected to produce a forced response transient response to the excitation input (TC analysis). Since
more closely approximating the natural regressing lead-lag the system under study is nearly linear, this may be considered
mode, thereby improving modal purity and signal-to-noise a nearly linear perturbation.
ratio of the transient decay for moving-block post-processing.
The principal difference between the three excitation inputs is
This direct excitation was accomplished by introducing an that the two blade pitch excitation inputs − the lateral cyclic
artificial external force applied to the tip of each rotor blade swashplate input and the rotating frame pitch input − produce
in the plane of the rotor disk. These lead-lag tip forces less than half the lead-lag bending moment response of the
applied to the four blades were phased azimuthally analogous blade tip force excitation (it is important to note that different
to swashplate blade pitch, that is, as three separate inputs scales are used in Figs. 10 through 12). This is because the
for collective, lateral cyclic, and longitudinal cyclic lead-lag tip force produces a direct lead-lag bending moment at the
force. The “lateral cyclic” lead-lag force was then input at blade root while the pitch inputs produce blade lift and flap
the regressing mode lead-lag frequency for one full cycle bending moments that only indirectly induce lead-lag bending
of oscillation. All of these blade excitation inputs were moments from structural and inertial coupling of flap and
implemented using standard modeling procedures in RCAS. lead-lag motion.
To more clearly illustrate the three different lead-lag One advantage of the tip force excitation is that the transient
excitations employed for the computational investigations, decay waveforms of the four different blades appear to be
Figs. 7 through 9 present the blade pitch and external force considerably more regular and consistent than in the cases
time histories. For the lateral cyclic pitch excitation (Fig. 7), of blade pitch excitation inputs even though the periodic
the first part, Fig. 7(a), shows the pitch input of all four blades signal has already been removed. This is assumed to
as a function of time (in units of revs). The second part, benefit the consistency and accuracy of the moving-block
Fig. 7(b), shows the rotor cyclic pitch input versus time. In post-processing method. Another difference between the
this case, a one-cycle oscillation of the lateral cyclic pitch at blade pitch and tip force excitation responses is the relative
the regressing lead-lag mode frequency. amplitude of the flap and lead-lag bending response. At
the point of excitation cut-off, the relatively large flap
Figure 8 for the rotating pitch excitation clearly illustrates the
bending response from the pitch inputs requires about two
linearly increasing blade pitch oscillations for the individual
revolutions to subside (reflecting the well damped flap
blades, input at the lead-lag frequency in the rotating system
bending mode) before the presumably pure regressing lead-
(Fig. 8(a)). The equivalent lateral and longitudinal cyclic pitch
lag mode emerges. In the case of the tip force excitation, the
angles are shown in Fig. 8(b). This combination of inputs
flap bending response is already very small during excitation,
leads to a lateral-longitudinal whirl excitation in the polar plot
and it might be surmised that this provides a longer period for
of Fig. 8(c). The analogous plot for Fig. 7 (not included)
post-processing a purer lead-lag regressing mode response.
would be a one-dimensional input (lateral axis alone), rather
However, it might be argued that the pitch input excitation
than a whirling excitation.
and transient decay better reflect the experimental setup.
For the cyclic lead-lag blade tip force excitation, Figs. 9(a)
and 9(b) are directly analogous to the lateral cyclic pitch The rotor MBC lead-lag bending moments (measured
excitation in Fig. 7 with the blade pitch angle replaced by the in the nonrotating, or fixed frame system) shown in
blade lead-lag force. Figs. 10(c), 11(c), and 12(c) serve to indicate the relative
regressing mode and progressing mode response as well as
The blade lead-lag and flap moment excitation and the collective and differential collective moment response
transient decay responses corresponding to the three different arising from periodic coupling of the MBC modes in
excitation inputs are presented in Figs. 10 through 12. forward flight as well as the effects of nonlinear blade
These transient responses are intended to show the relative response. The blade tip force excitation illustrates the
effectiveness of the three excitation inputs in producing “cleanest” regressing mode response of the MBC sine
transients suitable for moving-block post-processing. There and cosine modes as evidenced by transient decay at the
are three parts to each figure. For example, Fig. 10(a) shows regressing led-lag frequency (0.33/rev), without evidence of
the blade lead-lag bending moment response of primary response at the progressing mode frequency (1.67/rev). The
interest for all four blades, Fig. 10(b) shows the associated MBC collective and differential collective mode response is

5
virtually negligible. The situation for the two blade pitch systems, and aerodynamics. RCAS includes multiple
excitations is different. Although the MBC sine and cosine aerodynamic options for airloads, wake induced flowfields,
modes appear to reflect nearly exclusively regressing mode and component aerodynamic interference. Airloads models
without evidence of the progressing mode, there is significant include 2-D airfoil and lifting-line models for rotor blade,
collective and differential mode response at the lead-lag mode wings, or empennages and 3-D airloads for bodies.
frequency (0.67/rev). The collective lead-lag mode response
may be explained as follows: rotor flapping response to CFD calculations use the complex geometry Navier-Stokes
lateral cyclic produces rotor disk angle of attack oscillations, CFD solver OVERFLOW 2.1ad. Capabilities for loose (delta)
causing rotor thrust and torque responses, which in turn excite coupling (periodic) and tight coupling (maneuver) have
collective lead-lag mode response. When the cyclic pitch been added to the NASA release version based on original
input ends, the collective lead-lag mode transient decays at the developments under the DoD CHSSI Portfolio, Collaborative
collective lead-lag mode frequency. In fact, it is the collective Simulation and Testing (CST-05) [14, 15]. OVERFLOW
mode that is primarily responsible for the variation among the computes solutions on structured, overset grids using a near-
individual blade responses in Figs. 10(a) and 11(a). and off-body discretization paradigm. The near-body grids
surround the solid surfaces and capture the viscous effects
In order to arrive at the excitations illustrated in Figs. 7 with highly stretched curvilinear meshes. They extend out
through 12, considerable experimentation was carried out to a suitable distance (approximately one chord) from the rotor
tailor the blade pitch and tip force excitations and optimize blade. Automatically-generated Cartesian off-body grids
the quality of the transient decay responses, including the surround the near-body grids and capture the wake. They
following. It was found desirable to avoid any pitch extend out to the far field boundary, with geometrically
excitation inputs that began or ended discontinuously. For increasing spacing. Time-accurate simulations of complex
either regressing or progressing mode excitations, it was aircraft configurations with aeroelastic bodies in relative
desirable to tune the excitation frequency to most closely motion can be efficiently computed on parallel processors
match the blade lead-lag frequency in the rotating system. using the overset methodology.
Even numbers of excitation cycles were desirable; otherwise
mixed combinations of progressing and regressing mode RCAS ADM Model
response were produced. Finally, whirling excitations using
The ADM structural and aerodynamic characteristics were
a combination of lateral and longitudinal cyclic pitch or
modeled with RCAS for the baseline aeroelastic stability
tip force excitations produced significantly larger lead-lag
predictions. The RCAS structural model was used in
bending moment response for a given number of cycles,
conjunction with OVERFLOW 2 for the coupled CFD/CSD
suggesting a means of reducing CFD computation time for
predictions. Two distinct structural models for the straight
future applications of the present methodology.
and swept-tip rotor blades both consisted of nine nonlinear
beam elements for the flexible portions of the four identical
RCAS and CFD Modeling cantilever blades, three rigid bars for the hub, a rigid body
mass for the swept-tip mass balance, and rotational hinges
RCAS and OVERFLOW 2 Codes to represent the blade pitch bearing. The models were
idealized as an isolated rotor and no rotor support flexibility
The Rotorcraft Comprehensive Analysis System (RCAS) is a was provided. Given the measured properties of the rotor
comprehensive multi-disciplinary, computer software system test apparatus as discussed in Ref. 3, this is appropriate for
for predicting rotorcraft aerodynamics, performance, stability the regressing lead-lag mode addressed in this study. The
and control, aeroelastic stability, loads, and vibration [13]. blades were nominally untwisted but small twist angles were
RCAS was developed by the Aerofightdynamics Directorate, included for the straight blades. The measured structural,
US Army Aviation and Missile Research, Development, and geometric, and mass properties [3], were tabulated as a
Engineering Center (RDECOM), to provide state-of-the-art function of blade length for use in RCAS. The main portion of
rotorcraft modeling and analysis technology for government, the blades were essentially uniform in mass and stiffness with
industry, and academia. The RCAS is capable of modeling the elastic, tension, and mass center aligned with the quarter
a wide range of complex rotorcraft configurations operating chord line and the blade pitch axis.
in hover, forward flight, and maneuvering conditions. The
RCAS structural model employs a hierarchical, finite element, The RCAS aerodynamic model is based on the conventional
multibody dynamics formulation for coupled rotor-body lifting-line approach for rotorcraft. The blades were
systems. It includes a library of primitive elements including discretized into 14 spanwise aerosegments. Airload
nonlinear beams, rigid body mass, rigid bar, spring, damper, calculations were based on airfoil tables for the NACA 0012
and mechanical applied load as well as hinges and slides airfoil section, providing wind-tunnel measured lift, drag,
to build arbitrarily complex models. Rotor and fuselage and pitching moment coefficients as a function of angle of
modeling is fully integrated with engines, drivetrain, control attack and Mach number. Airfoil unsteady aerodynamics was

6
modeled with simple linear quasi-steady Theodorsen theory, Coupling is on a per revolution basis due to the assumption
C(k)=1.0, and the rotor wake was modeled with both uniform of periodicity. Motions (3 rotations and 3 translations of the
inflow and a dynamic inflow model. Nonlinear dynamic stall, airfoil sections) and airloads (3 component − section normal
yawed flow effects, and radial drag were not included. force, chord force, and pitching moment) are exchanged.
Fully-automated coupling between OVERFLOW and RCAS
Standard RCAS solution options were used to calculate is accomplished via simultaneous execution of the two codes.
baseline trim and stability characteristics. The modal analysis The codes “communicate” via flag files written to disk,
option (20 modes) was used in place of the full finite waiting for each other to finish their tasks. In loose coupling
element structural model, and modal damping was used this task is the analysis of a complete (trimmed) rotor
to represent structural damping of the experimental model. revolution. Consistent inputs to each solver synchronize
Conventional constant coefficient equation eigenanalysis was the coupling. Computations for the two disciplines are not
used to determine baseline stability characteristics in hover overlapped. Tight coupling (TC) for the maneuver transient
and as an approximation for periodic coefficient Floquet response portion of the analysis is similarly performed with
solutions in forward flight. Since the moving-block method the codes executing simultaneously and waiting for each
for post-processing transient responses was adopted as the other to finish, signaled by a flag file [16]. For TC the
principal method for stability analysis in the paper, the use of communication takes place every time step. The TC algorithm
Floquet analysis was not included; such results for the ADM implemented in OVERFLOW/RCAS uses a lagged scheme,
stability predictions may be found in the literature [4, 5]. with no subiterating between modules. This coupling is only
OVERFLOW 2 ADM CFD Model formally first-order accurate [17]. As is always the case in
CFD/CSD coupling, it is necessary to ensure that the CFD and
Conventional structured overset grids were developed for CSD models are consistent with each other, particularly with
the ADM configurations. The surface and volume grid respect to reference coordinate systems. This was particularly
configuration is shown in Figs. 13 and 14. The rotor is true for the preconed rotor.
faithfully duplicated up to the hub attachment, including
the rectangular cross-section spar and transition region. An For especially tight loose coupling trim convergence, 18
approximation to the hub was developed from drawings, and coupling iterations are used, with 1/4 rev (4-bladed rotor)
is shown in the figure. The swept-tip and straight blade grid between each coupling. Therefore, a coupled solution
systems are identical. Each blade is made up of 3 grids (root requires 5 complete rotor revolutions, including an initial
cap, main blade, and tip cap). The main blade is an O-grid starting revolution. For the TC transient maneuver, a
topology. The dimensions of the blade grid in the chordwise, small number of rotor revolutions (< 3.5), as determined
spanwise, and normal directions are 265 x 215 x 63. There by excitation method, are applied, followed by ∼10 rotor
are 25 points on the blunt trailing edge of the NACA 0012. revolutions of transient decay without excitation. Longer
The first normal spacing is chosen to obtain y+ less than 1. decay transients did not modify the damping results.
Chordwise leading and trailing edge spacings are .001c and
The time-accurate calculations use a 4th-order central
.0001c, respectively. The grid contains 29.7 million points
difference spatial discretization with added 4th-difference
(46% off-body) with 8.8% mean chord wake spacing (0.3
scalar (near-body and coarse off-body) and matrix (fine off-
inches) in the level 1 off-body grid that surrounds the rotor
body) artificial dissipation, resulting in a 3rd-order scheme. A
and captures the wake. The level 1 wake grid extends 3 chords
2nd-order temporal backward difference scheme with iterative
above and below the rotor. There are 4.0 million points per
dual time stepping is used for time advancement along with a
blade. A coarse mesh has been extracted from this fine mesh
penta-diagonal left-hand side scheme. Twelve subiterations
and is generally an every-other-point subset. Each coarse grid
are used on the fine grid and 6 subiterations on the coarse
blade contains 517,000 points. For the coarse mesh, the level
grid, typically resulting in 1.5 − 2.0 orders of magnitude
1 off-body grid spacing is 17.6% chord. The total number of
reduction in the main blade grid residuals. Quarter degree
points is 3.9 million (47% off-body). Numerous checks on
(0.25◦) time steps are used (1440 steps per rotor revolution).
both swept-tip and straight blade configurations indicated that
The Spalart-Allmaras turbulence model is employed in the
the airloads, structural response, and stability obtained from
near-body grids. The off-body wake grids are inviscid.
the coarse grid was in excellent agreement with the fine grid
at a fraction of the cost. Almost all CFD calculations in this For the coarse mesh, periodic loose coupling solutions
work use the coarse grid with fine grid analyses as noted as a are obtained on 64 processors and require 1.4 hours per
check. rotor revolution on a Cray XT5. A trim solution can be
RCAS/OVERFLOW 2 CFD/CSD Coupling obtained in 7 hours. For transient maneuver analysis 128
processors are used and require 1.1 hours per rotor revoluti on.
CFD/CSD coupling for the steady-state trim portion of the Transient analyses with excitation inputs were run for 13
analysis is performed using a conventional (for rotorcraft) rotor revolutions and required 14.3 hours, resulting in stability
loose coupling (LC) incremental “delta” formulation [1]. calculations being obtained overnight. For the fine mesh,

7
solutions require 3.8 hours per rotor revolution. Because When restarting the tight coupling maneuver inputs from the
of the overhead associated with the tight coupling exchange periodic solution using RCAS and OVERFLOW restart files,
and the extra CSD execution time, equal loose and tight it was necessary to ensure that a well converged trim condition
coupling cost per revolution on this mesh requires 256 and was obtained. Some minor differences were noted in the
320 processors, respectively. LC periodic signals compared with a TC transient analysis
without control inputs (null maneuver). This was potentially
CFD Convergence and Numerical Issues
due to different time steps used in RCAS between LC and TC
Since the development and widespread application of (5◦ versus 0.25◦), as well as other algorithmic and minor trim
CFD/CSD methods to rotorcraft airloads and blade loads differences.
in recent years, many investigators have found that the
During transient excitation of the rotor, it is important that
success of such applications is dependent on many variables
the resulting rotor blade motions are appropriate for the
involving the coupling of CFD and CSD codes. Achieving
CFD simulation. It was found that for the straight blade,
accurate and robust solutions has not always been simple
large pitch inputs produced unnecessarily large blade flapping
or straightforward. The present investigation for the novel
motions as shown in Fig. 19. While not necessarily physically
application of CFD/CSD stability analysis experienced a
erroneous, this occasionally occurred inadvertently, resulting
number of issues before satisfactory results were achieved.
in a solution with poor domain connectivity (i.e. orphan
Some details of these issues will be presented in this section.
points) when the near-body (NB) rotor blade grids moved
CFD/CSD Trim Convergence beyond the enclosing level 1 (L1) off-body grids. In
OVERFLOW, near-body grids are not able to communicate
An example LC trim convergence plot of CFD/CSD cyclic
with level 2 (L2) and higher Cartesian grids. The remedy in
pitch control angles for the straight blade across a range of
such cases was to reduce the blade pitch excitation amplitude.
advance ratios is shown in Fig. 15. Convergence is quite tight
except for the 0.15 advance ratio. It should be noted that in the In general, no other robustness issues were encountered
experiment, the higher advance ratios (µ > 0.35) approached with the CFD solutions, even at the high advance ratios.
regressing lead-lag mode instability. If RCAS accurately No problems were encountered with the blade tip force
reproduced these trends, it might not be possible to trim to excitation inputs, despite the relatively large lead-lag motions.
these high advance ratio conditions using a time marching One indication of the validity of the dynamic responses for
solution algorithm. CFD thrust and torque convergence for determination of aeroelastic stability characteristics is the
the straight blade at 0.30 advance ratio are shown in Fig. 16. degree of linearity of transient response data. The linearity
of the CFD solution was checked using different cyclic
Transient Response Numerical Sensitivities and Issues
excitation input amplitudes. Figure 20 illustrates the degree of
Due to the sensitivity of the aeroelastic stability analysis linearity for two cases with 2◦ and 3◦ lateral cyclic excitation
results, numerous numerical parameters were investigated inputs. By scaling the chordwise bending moment responses
to ensure that accurate and converged solutions were being by a constant factor (1.49 in this case), the results are seen to
obtained during the LC trim and TC maneuver transient be linear with excitation amplitude.
portions of the CFD/CSD analysis. A grid convergence study
Transient Response Noise
was performed on both the straight and swept-tip blades. The
major difference between the coarse and fine grids appeared Even though noise sources are typically associated with
to be a shift in chordwise root bending moment for the experimental data, numerical methods can also generate
LC trim periodic solution (Fig. 17(a)). However, when this “noise” and undesired oscillations in the solution [11].
periodic signal is subtracted from the TC transient response, Such computational noise may cause difficulty in the post-
the resulting perturbation signal is in excellent agreement processing stability analysis. Noise sources from the
between grids (Fig. 17(b)). These two signals produce a CFD calculations may include wake oscillations and non-
difference of 0.01 in damping value. Based on these and periodicity, rotor hub turbulence, non-smooth excitation
similar results for the straight blade, the coarse grid is used inputs, inexact numerical convergence, and rotor blade
throughout this work. A fast Fourier transformation analysis aerodynamic non-linearities. The wake shed from the hub
of the transient response in Fig. 17(b), determined that the and rectangular cross-section spar were found to produce non-
higher frequency oscillations are the second lead-lag mode, at periodic stochastic turbulence. Figure 21 compares wake
a frequency of 5.4/rev. Increased number of subiterations (15) visualizations (Q criteria) with and without the hub for the
and smaller time step (0.125◦) were also investigated against two blade configurations. The turbulence resulted in degraded
baseline values (6◦ and 0.25◦) for a non-smooth rotating trim convergence and some variation in the structural loads.
pitch excitation input. Again the results show no noticeable The rotor thrust and torque were also affected (Fig. 22),
difference (Fig. 18) in the perturbation signal. Damping of the particularly with respect to irregular variations in the cycle
three cases varies by only 0.015 and indicates that the coarse peak amplitudes. The hub was subsequently removed from
grid results are accurate as well as efficient. the ADM physical model for all CFD calculations.

8
During development and refinement of transient excitations, total response and perturbed response are compared for each
blade pitch inputs that contained step changes produced very excitation method. For the blade tip force excitation, there
noisy transient responses. These oscillations, typically due to is virtually no difference between the total and perturbed
high frequency structural response, usually dissipated quickly responses because the blade tip force excitation provides
following cessation of the excitation. An example of the a purer high amplitude lead-lag regressing mode response,
non-smooth excitation inputs is shown in Fig 18. In the as was discussed earlier. For the lateral cyclic excitation,
final results, depending on the excitation methodologies, the significant oscillation in the natural logarithm of the moving-
excitation inputs were timed to apply continuous control block function was observed with the total response. The
inputs on both entry and exit from the maneuver. For the least-squares fit of this signal hampers the accurate estimation
rotating pitch input, this went even farther by ramping up the of blade damping. The perturbed response, however, provides
input amplitude from zero to the final value (Fig 17). For an almost linear signal.
the blade inplane tip force and lateral cyclic excitation inputs,
an extra revolution of periodic, unexcited response was run
at the beginning of the maneuver (Fig. 20 − lateral cyclic Results and Discussion
excitation).
This section of the paper will present results of the
Blade-to-Blade Differences and Moving-Block Sensitivity stability analyses performed with the CFD/CSD methodology
developed in this paper. Baseline results from conventional
In the CFD and CSD models all four blades are identical, methods and ADM experimental results will be included for
unlike the experimental blades. As noted above, apparent comparison. A few additional airloads and trim results will
differences in responses between blades were found to also be included for completeness.
arise due to numerical issues, the CFD/CSD coupling
algorithm, CFD turbulent flow physics, etc. Another apparent Hover Stability RCAS Results
manifestation of blade-to-blade differences arises from the
Figures 26 and 27 compare the calculated regressing lag
dynamics of the excitation itself. Figure 23(a) shows the
mode damping for the swept-tip rotor blades with the
perturbation lead-lag bending moment response of the straight
measured values in hover with precone angle of 0 ◦ and 2◦ ,
blade configuration to lateral cyclic pitch excitation. Here
respectively. Both constant coefficient eigenanalysis and
the amplitude of the four blades differs significantly. These
moving-block results are presented as a function of collective
differences are similar but much more pronounced than the
pitch. Uniform inflow was used for the constant coefficient
example shown in Fig. 10(a) for the responses of the four
eigenanalysis, and both uniform inflow and dynamic inflow
swept-tip blades to similar excitation. Figure 23(b) shows
were used for the transient response analysis with moving-
the rotor lead-lag bending moments transformed into multi-
block method. The dynamic inflow used is a three-state
blade coordinates, and it is evident that there is significant
Peters-He model [18] whose highest power of the radial terms
collective mode response present that is responsible for the
is one and whose maximum number of harmonics is one,
apparent differences in response of the four blades. The cause
denoted as 1×1, which is essentially the Pitt-Peters dynamic
was described earlier and is due to the rotor thrust and torque
inflow model [19].
response of the rotor to the lateral cyclic pitch excitation. The
effect is more pronounced for the straight blade rotor since Dynamic inflow reduces damping compared to uniform inflow
the torsional response of the swept-tip blades opposes thrust for both 0◦ and 2◦ precone angles, especially at high collective
changes due to blade pitch and rotor disk angle of attack and pitch angles. The tip loss has a significant influence on the
minimizes the effect. damping prediction. Including the tip loss resulted in much
higher damping values, especially at high collective pitch
It was found that the CFD noise in the tight coupling transient angles. The analysis without the tip loss agrees well with
responses could be mitigated if the periodic responses from
the measured values. There is no difference between the
the loose coupling trim analysis were subtracted from the
constant coefficient eigenanalysis and the transient response
transients to yield the perturbation transients as discussed analysis with moving-block method. This shows that the
earlier. The moving-block post-processing of the total
transient response analysis with moving-block technique
response was often unsatisfactory, while post-processing of works properly.
the perturbation transient usually produced a more acceptable
result. Noise in the coupled CFD/CSD solutions nonetheless Forward Flight CFD/CSD and RCAS Trim Results
did produce some blade-to-blade differences in transient
responses and the post-processed damping results. Numerous studies have shown that CFD/CSD airloads
predictions are more accurate in most cases than lifting-
Figures 24 and 25 show screenshots of the moving-block line aerodynamics results from comprehensive codes, even
analysis for the swept-tip blade at µ = 0.3 with blade tip force with the use of free wake models [1]. An example airloads
excitation and lateral cyclic excitation, respectively. Both comparison between coupled OVERFLOW/RCAS and RCAS

9
lifting-line table-lookup with dynamic inflow is shown in Forward Flight CFD/CSD and RCAS Transient Response
Fig. 28 for the swept-tip blade at 0.30 advance ratio (αs Time Histories
= 0◦ , θ0 = 6◦ ). The comprehensive analysis is unable
to accurately capture the wake induced oscillations in the Before presenting the post-processed lead-lag regressing
airloads, particularly at 90 ◦ and 270◦ azimuth. Overall the mode damping results, it is of interest to compare the transient
results are of the same magnitude but much smoother with excitation response and decay of the baseline RCAS analysis
dynamic inflow. Similar discrepancies might be expected at with the coupled RCAS/OVERFLOW analysis. The straight
high advance ratios, especially on the straight blade. This blade with the tip force excitation is used for this example,
undoubtedly has some affect on aeroelastic stability. and the response is first shown pictorially in Fig. 31 as a
stop-motion sequence of images depicting the blade motion
Because the initial trim solution is critical to the lead-lag amplitude. Figure 32 shows the time history lead-lag bending
stability, Fig. 29 compares the control angles calculated with moment response at µ = 0.3 in the rotating system including 1)
RCAS and RCAS/OVERFLOW against the measured values the trim periodic response for the first revolution, 2) followed
for the straight blade, 6◦ collective, −6◦ shaft angle for a by one cycle of regressing frequency tip force (about 3.5
range of advance ratios. The trends are in good agreement but revolutions), and 3) another 8.5 revolutions of transient decay.
the longitudinal cyclic angles are offset by as much as 1.5◦ for Note that the time history is the total transient response
RCAS. The measured collective pitch at µ = 0.1 appears to be rather than the perturbation component. The trim periodic
an outlier. The coupled CFD/CSD analysis shows consistently response and initial excitation are in close agreement for both
better correlation than RCAS alone analysis. the baseline RCAS and the RCAS/OVERFLOW calculations,
To help provide a basis for the aeroelastic stability results however, as the transient decay progresses, the lower damping
to follow, Fig. 30 is included to illustrate how the rotor of the CFD/CSD transient is apparent. It is interesting to note
thrust varies as a function of advance ratio for the different that the lead-lag damping is a very sensitive indicator of the
operating conditions of the straight and swept-tip blades. aeroelastic response, given that the two time history responses
These conditions represent a subset of the test conditions look about the same. As will be seen in the comparisons with
of the two configurations summarized at the beginning of the experimental damping measurements presented below,
the paper in Fig. 4. It is noted that the rotor thrust was CFD/CSD represents the more accurate result.
not measured during the wind tunnel testing and, therefore,
Forward Flight Stability Blade-to-Blade Sensitivity
cannot be compared with these results. However, the primary
purpose is not for correlation but to relate rotor thrust to One of the key points of interest about the methodology
stability characteristics and, thereby, help interpret the results. presented in this paper is the accuracy and robustness of the
Basic flap-lag stability of hingeless rotor blades may be CFD/CSD transient response calculations and the moving-
related to the trimmed flapping deflection of the rotor block post-processing method. A number of results were
blade and the effective pitch-lag aeroelastic couplings presented to illustrate some of the issues encountered for the
arising from perturbation lead-lag deflections of torsionally CFD/CSD calculations, and several transient excitations were
flexible blades [20, 21]. Essentially, the equilibrium blade examined for suitability in generating well behaved transient
flap deflection accompanying increased thrust will produce responses for stability post-processing. Results using the
stabilizing pitch-lag coupling that will increase lead-lag baseline RCAS calculations in the hover condition validated
damping in soft-inplane hingeless rotor blades like the ADM. the moving-block results by comparing the regressing lead-
This is observed in the hover results in Figs. 26 and 27 at both lag damping to the eigensolution results.
positive and negative thrust (positive and negative collective
For the CFD/CSD results in forward flight, the moving-block
pitch). The destabilizing effect of precone arises from its
results were further assessed by comparing the damping of
effect on equilibrium flap deflection.
all four blades for the swept and straight blades and the three
Figure 30 shows that, in forward flight, the thrust of the different excitation methods as a function of advance ratio.
swept-tip blade at zero rotor angle of attack varies only The results are presented in Figs. 33 and 34 for the swept-
moderately with advance ratio, while the thrust of the straight tip and straight blades, respectively. The agreement with the
blade, at 6◦ negative angle of attack, diminishes significantly experimental data is shown to be generally very good and will
with advance ratio due to the increasing free stream velocity be discussed in more detail in the next section. Of interest
component normal to the rotor disk. The thrust becomes here is the sensitivity of the results for the different blades.
negative at µ < 0.3 for the 3◦ collective pitch. Analogous As noted earlier, the blades were modeled with identical
to the hover result at zero collective pitch, the damping in properties for the calculated results, so the differences in
forward flight would be expected to be a minimum under the results are due to variations in the calculated and post-
these conditions. Whether the regressing lag mode would processing results for the four blades. In general, the swept-
become unstable (as the trend in Fig. 4(b) suggests) or not tip blades show more consistency than the straight blades,
would depend on the effects of forward flight and precone. and this may be related to the overall lower sensitivity of the

10
swept-tip blades to changes in operating condition. The blade the constant coefficient eigenanalysis at higher advance ratios.
tip force excitation method appears to give more consistent The coupled RCAS/OVERFLOW moving-block transient
results than the two blade pitch excitation methods, and this is analysis shows that the three different excitation methods
likely due to the larger lead-lag motions relative to blade pitch produce almost identical damping values. It should be noted
and flap motions as discussed earlier. that damping was calculated at µ = 0.45 with the lateral cyclic
excitation and at µ = 0.4 and 0.5 with inplane force and
Isolated anomalies in the blade sensitivity results, such as rotating pitch excitations. As with the 3◦ collective case, the
the higher damping of Blades 2 and 4 in Fig. 33(a) or the coupled CFD/CSD analysis shows increased overall damping
slight differences between the three excitation methods, are compared to the RCAS alone analysis. The “up-down-up”
unexplained. Overall, however, the predicted damping results variations with advance ratio are well captured only with the
appear to be reasonably consistent and generally agree well coupled CFD/CSD analysis.
with the measured results.
Figure 37 shows the calculated regressing lag mode damping
Forward Flight Stability Results
for the straight blades with 3◦ collective. RCAS constant
This section compares results of the forward flight stability coefficient eigenanalysis shows reasonably good correlation
analyses with experimental results. Blade regressing lag mode up to µ = 0.2, but significantly overpredicts after that. RCAS
stability was calculated using three different methods: 1) moving-block transient analysis shows that all three methods
RCAS constant coefficient eigenanalysis, 2) RCAS transient show somewhat similar damping values, and the predicted
response analysis using the moving-block method, and damping values are almost same as those with constant
3) coupled RCAS/OVERFLOW transient response analysis coefficient eigenanalysis. The coupled RCAS/OVERFLOW
using the moving-block method. For the RCAS alone moving-block transient analysis with inplane force excitation
analysis, a fifteen-state Peters-He dynamic inflow model [18], better captures the damping reductions at high speed.
whose highest power of the radial terms is four and whose The damping values predicted with the coupled CFD/CSD
maximum number of harmonics is four, was used. It should analysis decrease at µ = 0.30 and 0.35 compared to those with
mentioned that the moving-block results for the coupled RCAS analysis, and thus the correlation is improved.
CFD/CSD analysis in the following figures were averaged
Figure 38 shows the calculated regressing lag mode damping
from all four blades.
for the straight blades with 6◦ collective. RCAS constant
Figure 35 compares the calculated regressing lag mode coefficient eigenanalysis overpredicts the damping. RCAS
damping for the swept-tip rotor blades with 3 ◦ collective. moving-block transient analysis reduces damping values at
RCAS constant coefficient eigenanalysis shows good high advance ratios compared to the constant coefficient
agreement with the test data up to µ = 0.35, but overpredicts eigenanalysis results. Differences among the three excitations
at higher advance ratios. RCAS moving-block transient are larger at this collective angle compared to the 3 ◦
analysis shows that the three different excitation methods collective results. The coupled CFD/CSD analysis captures
show almost identical damping values, and the predicted the decreased damping with advance ratio and significantly
damping values are slightly reduced at high advance ratios improves the correlation with the test data, but also shows
compared to the constant coefficient eigenanalysis results. differences among the three excitations.
For the coupled RCAS/OVERFLOW analysis, only inplane
tip force excitation was used. The coupled transient response Taken together, the forward flight regressing lead-lag damping
analysis with moving-block method increased damping at all results for the swept-tip and straight blades presented in
advance ratios except µ = 0.45 compared to the RCAS alone Figs. 35 through 38 show that the coupled CFD/CSD
analysis. The increased damping trend with advance ratio is predictions produce a significant improvement compared with
noticeably reduced. the RCAS predictions based on conventional aerodynamics
methods. In particular, the key features in the damping
Figure 36 compares the calculated regressing lag mode trends that have eluded numerous ADM correlation studies,
damping for the swept-tip rotor blades with 6 ◦ collective. namely the “up-down-up” trend of the swept-tip blade and
RCAS constant coefficient eigenanalysis shows that the the “decreasing damping” trend of the straight blade, now
damping increases almost linearly, thus underpredicts the appear to have been reasonably well captured. Furthermore,
regressing lag mode damping at low advance ratios, but the moving-block transient response results based on
overpredicts at high advance ratios. RCAS moving-block conventional RCAS showed a small improvement compared
transient analysis shows that the predicted damping values to the constant coefficient eigensolution approximation,
agree well between the inplane tip force excitation and consistent with previously published results based on periodic
rotating pitch excitation methods, however, cyclic pitch coefficient Floquet eigensolutions. This implies that the
excitation results show larger damping. The predicted principal source of the improved damping is a result of
damping values are almost constant up to around µ = 0.3 and introducing more accurate aerodynamics from the CFD
then slowly increase. In general, correlation is better than modeling.

11
It is of interest to understand the specific reasons for the principal reason for the improved damping predictions of
improvements in aeroelastic stability and lead-lag damping the coupled CFD/CSD method is a consequence of the more
due to CFD aerodynamics. The RCAS and OVERFLOW accurate CFD unsteady aerodynamics.
airloads comparisons in Fig. 28 implied some of the
limitations of conventional aerodynamics. The improved trim 4) The moving-block transient response method was validated
solutions in Fig. 29 were largely due to the more accurate by showing conventional RCAS lead-lag damping results for
CFD rotor wake inflow distributions, and since the trim hover in agreement with RCAS constant coefficient equation
solution directly influences the blade equilibrium solution eigensolutions, as should be the case.
through the effective aeroelastic couplings, this would account
5) A number of CFD/CSD implementation issues were
for some of the damping improvement. More accurate
addressed to achieve practical aeroelastic stability results with
unsteady aerodynamics are also likely to be a significant
limited computational run-time. It was found that relatively
part of the improvement since time-dependent perturbation
short excitation inputs would yield transient responses
airloads ultimately govern the damping of the transient blade
suitable for moving-block stability post-processing if the
motions. Exactly which aspects of CFD fluid mechanics are
inputs were tuned to the blade lead-lag frequency, and were
most important − compressibility, unsteady aerodynamics,
not initiated or terminated discontinuously.
or three-dimensional effects − will take more detailed
investigation. To consider only one of these possibilities, it 6) Of the several transient excitation methods investigated
is well known [20] that the airfoil profile drag coefficient is in this work, the blade tip force excitation was considerably
a basic determinant of simple flap-lag stability in hover; how more effective than the cyclic pitch and rotating blade pitch
airfoil drag varies for complex unsteady motion and reversed inputs in generating blade lead-lag responses for stability
flow in forward flight is presently poorly understood, but CFD post-processing.
analysis may reveal its importance in determining lead-lag
damping trends at high advance ratios. 7) The three different excitation methods gave generally
consistent lead-lag damping results but some unexplained
differences and variations between blades were observed.
Summary and Conclusions Typically these differences were more apparent in the
CFD/CSD results than for the RCAS alone results.
The CFD/CSD coupling method for the rotor aeroelastic
stability problem was successfully developed. The method
Acknowledgments
was applied to calculate lead-lag regressing mode damping
of both straight and swept-tip hingeless rotor blades and
compared with experimental data in forward flight. The The authors would like to thank Thomas Maier at U.S.
principal findings of this work are summarized here. Army AFDD for sharing his experience of the experiments,
Dr. Hossein Saberi at Advanced Rotorcraft Technology
1) For isolated rotor lead-lag mode aeroelastic stability, Inc. for his assistance in the RCAS analysis, and Dr.
the CFD/CSD method based on loose coupling trim, Mahendra Bhagwat at U.S. Army AFDD for his assistance
tight coupling transient response, and moving-block post- with CFD/CSD tight coupling analyses. Dr. Pieter Buning
processing was shown to be reasonably practical and robust at NASA performed the integration of rotorcraft capabilities
when transient excitations were tailored to minimize the into OVERFLOW 2.1. Computing resources under a DoD
number of rotor revolutions needed. HPCMP Challenge Project at the Navy DSRC are gratefully
acknowledged.
2) The RCAS/OVERFLOW coupled CFD/CSD lead-lag
damping predictions in forward flight agreed more closely
with the experimental measurements than conventional RCAS References
predictions. In particular, previously elusive ADM stability
characteristics − the swept-tip blade “up-down-up” variation [1] Potsdam, M., Yeo, H., and Johnson, W., “Rotor Airloads
in damping with advance ratio and the straight blade Prediction Using Loose Aerodynamic/Structural
“decreasing damping” trend were reasonably well captured in Coupling,” Journal of Aircraft, Vol. 43, No. 3, May-
the present CFD/CSD results. June 2006.

3) Conventional RCAS damping results using the moving- [2] Bhagwat, M. J., Ormiston, R. A., Saberi, H. A.,
block transient analysis showed a modest improvement and Xin, H., “Application of CFD/CSD Coupling for
(increasing with collective pitch and advance ratio) over the Analysis of Rotorcraft Airloads and Blade Loads in
approximate constant coefficient eigensolutions, as might be Maneuvering Flight,” American Helicopter Society 62rd
expected. However, the relatively small differences indicate Annual Forum Proceedings, Virginia Beach, VA, May 1-
the periodic coefficient effects are not large and that the 3, 2007.

12
[3] Maier, T. A., Sharp, D. L., and Lim, J. W., “Fundamental [14] Strawn, R., Nygaard, T., Bhagwat, M., Dimanlig, A.,
Investigation of Hingeless Rotor Aeroelastic Stability, Saberi, H., Ormiston, R., and Potsdam, M., “Integrated
Test Data and Correlation,” American Helicopter Computational Fluid and Structural Dynamics Analyses
Society 51st Annual Forum Proceedings, Forth Worth, for Comprehensive Rotorcraft Analysis,” AIAA Paper
TX, May 9-11, 1995. 2007-6575, AIAA Atmospheric Flight Mechanics
Conference, Hilton Head, SC, August 20-23, 2007.
[4] Maier, T. A., Sharp, D. L., and Abrego, A. I.,
“Aeroelastic Stability for Straight and Swept-Tip Rotor [15] Nygaard, T., Ormiston, R. O., Potsdam, M., Saberi,
Blades in Hover and Forward Flight,” American H., and Strawn, R., “CFD and CSD Coupling
Helicopter Society 55th Annual Forum Proceedings, Algorithms and Fluid Structure Interface for Rotorcraft
Montreal, Canada, May 25-27, 1999. Aeromechanics in Steady and Transient Flight
Conditions,” American Helicopter Society 62nd
[5] Maier, T. A., and Abrego, A. I., “Analytical Model Annual Forum, Phoenix, AZ, May 9-11, 2006.
Sensitivity Study for Aeroelastic Stability of Straight
and Swept-Tip Rotor Blades,” The 26th European [16] Bhagwat, M., and Ormiston, R. O., “Examination
Rotorcraft Forum Proceedings, The Hague, Netherlands, of Rotor Aerodynamics in Steady and Maneuvering
September 26-29, 2000. Flight Using CFD and Conventional Methods,” AHS
Specialists’ Conference on Aeromechanics, San
[6] Subramanian, S., Ma, G., Gaonkar, G. H., and Maier, Francisco, CA, January 23-25, 2008.
T. A., “Correlation of Several Aerodynamic Models and
Measurements of Hingeless-Rotor Trim and Stability,” [17] Silbaugh, B., and Baeder, J., “Coupled CFD/CSD
Journal of the American Helicopter Society, Vol. 45, No. Analysis of a Maneuvering Rotor Using Staggered and
2, April 2000, pp. 106-117. Time-Accurate Coupling Schemes,” AHS Specialists’
Conference on Aeromechanics, San Francisco, CA,
[7] Cafarelli, I., Hopkins, A. S., Truong, K, V., and Maier, January 23-25, 2008.
T. A., “Aeroelastic Stability Analysis of Two Hingeless
Rotors,” 33rd European Rotorcraft Forum Proceedings, [18] Peters, D. A, and He, C. J., “Correlation of Measured
Kazan, Russia, September 11-13, 2000. Induced Velocities with a Finite-State Wake Model,”
Journal of the American Helicopter Society, Vol. 36, No.
[8] Anon, “Flutter Testing Techniques,” Conference at 3, July 1991.
Dryden Flight Research Center, Edwards, CA, October
9-10, 1975, NASA SP-415, 1976. [19] Pitt, D. M., and Peters, D. A, “Theoretical Prediction
of Dynamic-Inflow Derivatives,” Sixth European
[9] Hammond, C. E., and Doggett, R. V., Jr., “Determination Rotorcraft and Powered Lift Aircraft Forum
of Subcritical Damping by Moving-Block/Randomdec Proceedings, Bristol, England, September 16-19,
Application,” NASA SP-415, October 1975. 1980.
[10] Bousman, W. G., and Winkler, D. J., “Application of the [20] Ormiston, R. A., and Hodges, D. H., “Linear Flap-
Moving-Block Analysis,” AIAA/ASME/ASCE/AHS Lag Dynamics of Hingeless Helicopter Rotor Blades
22nd Structures, Structural Dynamics, and Materials in Hover,” Journal of the American Helicopter Society,
Conference Proceedings, Atlanta, GA, April 6-8, 1981 Vol. 17, No. 2, April 1972, pp. 2-14.
[11] Bauchau, O. A., and Wang, J., “Efficient and [21] Hodges, D. H., and Ormiston, R. A., “Stability of Elastic
Robust Approaches for Rotorcraft Stability Analysis,” Bending and Torsion of Uniform Cantilever Rotor
American Helicopter Society 63rd Annual Forum and Blades in Hover with Variable Structural Coupling,”
Technology Display, Virginia Beach, VA, May 1-3, NASA TN D-8192, April 1976.
2007.

[12] McNamara, J. J., and Friedmann, P. P., “Flutter-


Boundary Identification for Time-Domain
Computational Aeroelasticity,” AIAA Journal, Vol.
45, No. 7, July 2007.

[13] Saberi, H. A., Khoshlahjeh, M., Ormiston, R. A., and


Rutkowski, M. J., “RCAS Overview and Application
to Advanced Rotorcraft Problems,” AHS Fourth
Decennial Specialists’ Conference on Aeromechanics,
San Francisco, CA, January 21-23, 2004.

13
Fig. 1: Advanced Dynamic Model (ADM) straight and swept-tip rotor models installed on the Rotor Test Rig (RTR) in the
AFDD 7-by 10-Foot wind tunnel.

Fig. 2: Planform of straight and swept-tip rotor models

14
8.0 8.0
T2
T2

F3
6.0 F3 6.0
Frequency, per rev

Frequency, per rev


F4
L2
L2

4.0 4.0
F2
F2

T1 T1
2.0 2.0
F1 F1

L1 L1
0.0 0.0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2
Rotor speed (Ω/Ω ) Rotor speed (Ω/Ω )
o o
(a) Straight blade. (b) Swept-tip blade.

Fig. 3: Fan plots for straight and swept-tip rotor blades with 0◦ collective.

2.0 3 deg collective 3.0


3 deg collective
4 deg collective
5 deg collective 4 deg collective
6 deg collective 5 deg collective
7 deg collective
6 deg collective
σ, rad/sec

σ, rad/sec

1.5 2.0
Damping -σ

Damping -σ

1.0 1.0

0.5 0.0
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
Advance ratio Advance ratio
(a) Swept-tip blade, αs = 0◦ , βp = 0◦ (b) Straight blade, αs = −6◦ , βp = 2◦

Fig. 4: ADM measured damping showing key trends for the swept-tip and straight blades.

15
12.6

Pitch angle, deg


12.3

12.0

11.7

11.4

0 0.5 1 1.5 2 2.5

Time, sec
(a) Pitch angle.

8.0
CBM @12%R, ft-lb

4.0

0.0

-4.0

-8.0

-12.0

0 0.5 1 1.5 2 2.5

Time, sec
(b) Chord bending moment @12%R.

Fig. 5: Representative pitch excitation and blade lad-lag structural response in hover.

Fig. 6: Moving-block transient response analysis for typical transient decay in hover.

16
4 4
Blade 1 MBC_Coll
Blade 2 MBC_Diff

Perturbatation pitch angle, deg

Perturbatation pitch angle, deg


Blade 3 MBC_Cos
Blade 4 MBC_Sin
2 2

0 0

-2 -2

-4 -4
0 1 2 3 4 5 6 7 8 0 1 2 3 4 5 6 7 8
Rotor revolution Rotor revolution
(a) Rotating frame (b) Fixed frame

Fig. 7: Lateral cyclic pitch excitation with regressing lag frequency.

3 3 1.0
Blade 1 MBC_Coll
Blade 2 MBC_Diff
Perturbatation pitch angle, deg

Perturbatation pitch angle, deg

2 Blade 3 2 MBC_Cos 0.5


Blade 4 MBC_Sin

Sine cyclic pitch


1 1 0.0

0 0 -0.5

-1 -1 -1.0

-2 -2 -1.5

-3 -3 -2.0
0 1 2 3 4 5 6 7 8 0 1 2 3 4 5 6 7 8 -1.0 -0.5 0.0 0.5 1.0 1.5 2.0
Rotor revolution Rotor revolution Cosine cyclic pitch
(a) Rotating frame (b) Fixed frame (c) Polar plot of fixed frame

Fig. 8: Rotating system blade pitch excitation with lead-lag frequency in the rotating system.

12 12
Blade 1 MBC_Coll
Blade 2 MBC_Diff
Blade 3 MBC_Cos
Lateral cyclic tip force, lbs

Lateral cyclic tip force, lbs

Blade 4 MBC_Sin
6 6

0 0

-6 -6

-12 -12
0 1 2 3 4 5 6 7 8 0 1 2 3 4 5 6 7 8
Rotor revolution Rotor revolution
(a) Rotating frame (b) Fixed frame

Fig. 9: Blade tip lead-lag force excitation with regressing lag frequency.

17
16 16 16
Chord bending moment @12%R, ft-lb

Blade 1 Blade 1 MBC_Coll

Flap bending moment @12%R, ft-lb


Blade 2 Blade 2 MBC_Diff

Fixed frame CBM @12%R, ft-lb


Blade 3 Blade 3 MBC_Cos
Blade 4 Blade 4 MBC_Sin
8 8 8

0 0 0

-8 -8 -8

-16 -16 -16


0 1 2 3 4 5 6 7 8 0 1 2 3 4 5 6 7 8 0 1 2 3 4 5 6 7 8
Rotor revolution Rotor revolution Rotor revolution
(a) Blade chord bending moment @12%R (b) Blade flap bending moment@12%R (c) Blade chord bending moment @12%R in
fixed frame

Fig. 10: Blade perturbation chord and flap moment response to lateral cyclic pitch excitation, RCAS/OVERFLOW, swept-tip
blade, µ = 0.3, αs = 0◦ , βp = 0◦ , θ0 = 6◦ .

30 30 30
Chord bending moment @12%R, ft-lb

Blade 1 Blade 1 MBC_Coll


Flap bending moment @12%R, ft-lb

Blade 2 Blade 2 MBC_Diff

Fixed frame CBM @12%R, ft-lb


Blade 3 Blade 3 MBC_Cos
Blade 4 Blade 4 MBC_Sin
15 15 15

0 0 0

-15 -15 -15

-30 -30 -30


0 1 2 3 4 5 6 7 8 0 1 2 3 4 5 6 7 8 0 1 2 3 4 5 6 7 8
Rotor revolution Rotor revolution Rotor revolution
(a) Blade chord bending moment @12%R (b) Blade flap bending moment@12%R (c) Blade chord bending moment @12%R in
fixed frame

Fig. 11: Blade perturbation chord and flap moment response to rotating system blade pitch excitation, RCAS/OVERFLOW,
swept-tip blade, µ = 0.3, αs = 0◦ , βp = 0◦ , θ0 = 6◦ .

80 10 80
Chord bending moment @12%R, ft-lb

Blade 1 Blade 1 MBC_Coll


Flap bending moment @12%R, ft-lb

Blade 2 Blade 2 MBC_Diff


Fixed frame CBM @12%R, ft-lb

Blade 3 Blade 3 MBC_Cos


Blade 4 Blade 4 MBC_Sin
40 5 40

0 0 0

-40 -5 -40

-80 -10 -80


0 1 2 3 4 5 6 7 8 0 1 2 3 4 5 6 7 8 0 1 2 3 4 5 6 7 8
Rotor revolution Rotor revolution Rotor revolution
(a) Blade chord bending moment @12%R (b) Blade flap bending moment@12%R (c) Blade chord bending moment @12%R in
fixed frame

Fig. 12: Blade perturbation chord and flap moment response to blade tip force excitation, RCAS/OVERFLOW, swept-tip blade,
µ = 0.3, αs = 0◦ , βp = 0◦ , θ0 = 6◦ .

18
µ = 0.10
µ = 0.15
µ = 0.20
2 µ = 0.25
µ = 0.30
µ = 0.35
µ = 0.40
µ = 0.45
1

cyclic angles, deg


0

-1

-2

-3

-4
5 10 15
coupling iteration

Fig. 15: Loose coupling trim convergence, straight blade,


Fig. 13: ADM swept-tip blade surface grids with hub. range of advance ratios, αs = −6◦ , βp = 2◦ , θ0 = 6◦ .

thrust
convergence / final value

torque
0.95

0.9

0.85

5 10 15
coupling convergence

Fig. 14: ADM swept-tip blade volume grids with hub. Fig. 16: Loose coupling thrust and torque convergence,
straight blade, µ = 0.30, αs = −6◦ , βp = 2◦ , θ0 = 6◦ .

19
5 coarse grid
fine grid

CBM @12%R, ft-lb


0

-5

-10

0 2 4 6 8 10 12

Rotor revolutions
(a) Steady state trim periodic response

15 coarse grid
fine grid
10
CBM @12%R, ft-lb

-5

-10

-15

0 2 4 6 8 10 12

Rotor revolutions
(b) Perturbation transient response

Fig. 17: CFD grid convergence with rotating pitch excitation, swept-tip blade, µ = 0.30, αs = 0◦ , βp = 0◦ , θ0 = 6◦ .

20 ∆ψ = 0.25°, subiterations = 6
∆ψ = 0.125°
subiterations = 15
CBM @ 12%R, ft-lb

10

-10

-20
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17
rotor revolutions

Fig. 18: Numerical parameter investigation: effect of timestep and number of CFD subiterations, swept-tip blade, µ = 0.30,
αs = 0◦ , βp = 0◦ , θ0 = 6◦ , non-smooth rotating pitch excitation.

20
Fig. 19: Excessive flapping from rotating pitch excitation with 4◦ pitch amplitude, straight blade, µ = 0.40, αs = −6◦ , βp = 2◦ ,
θ0 = 6◦ .

θ1 = 2°
20 θ1 = 3°
θ1 = 2°, scaled
CBM @ 12%R, ft-lb

10

-10

-20
0 1 2 3 4 5 6 7 8 9 10 11 12 13
rotor revolutions

Fig. 20: Effect of lateral cyclic pitch input excitation amplitude on perturbation response, straight blade, µ = 0.30, αs = −6◦ ,
βp = 2◦ , θ0 = 6◦ .

(a) swept-tip blade (αs = 0◦ ) (b) straight blade (αs = −6◦ )

Fig. 21: CFD visualization of wake, comparing hub effects, fine grid, µ = 0.30.

21
0.0058 hub
no hub
0.0056

0.0054

T
0.0052

C
0.0050

0.0048

0.0046

0 1 2 3 4 5

Rotor revolutions
(a) Rotor thrust

0.00028 hub
no hub

0.00027
Q
C

0.00026

0.00025

0.00024

0 1 2 3 4 5

Rotor revolutions
(b) Rotor torque

Fig. 22: Effect of hub on thrust and torque, swept-tip blade, CFD periodic solutions with no excitation, µ = 0.30, αs = 0◦ ,
βp = 0◦ , θ0 = 6◦ .
Blade 1
15 Blade 2
Blade 3
10
CBM @12%R, ft-lb

Blade 4
5

-5

-10

-15

0 2 4 6 8 10 12

Rotor revolutions
(a) Perturbation transient responses for all four blades

15
MBC_Coll
MBC_Diff
10
CBM @12%R, ft-lb

MBC_Cos
5 MBC_Sin

-5

-10

-15

0 2 4 6 8 10 12

Rotor revolutions
(b) Perturbation MBC transient response in the fixed frame

Fig. 23: RCAS/OVERFLOW transient response, µ = 0.30, αs = −6◦ , βp = 2◦ .

22
(a) Total response

(b) Perturbed response

Fig. 24: Moving-block analysis of swept-tip blade with blade tip force excitation, µ = 0.30, αs = 0◦ , βp = 0◦ , θ0 = 6◦ .

23
(a) Total response

(b) Perturbed response

Fig. 25: Moving-block analysis of swept-tip blade with lateral cyclic excitation, µ = 0.30, αs = 0◦ , βp = 0◦ , θ0 = 6◦ .

24
4.0 4.0 Test
Test
Uniform inflow - Tip loss 0.98
Uniform inflow - Tip loss 0.98
Dynamic inflow, 1x1 - Tip loss 0.98
Uniform inflow - No tip loss Uniform inflow - No tip loss
Dynamic inflow, 1x1 - No tip loss
3.0 3.0
σ, rad/sec

σ, rad/sec
Damping -σ

Damping -σ
2.0 2.0

1.0 1.0

0.0 0.0
-15 -10 -5 0 5 10 15 -15 -10 -5 0 5 10 15
Collective pitch, deg Collective pitch, deg
(a) Constant coefficient eigenanalysis. (b) Transient response analysis with moving-block technique.

Fig. 26: Effect of dynamic inflow and tip loss on damping in hover of swept-tip blades with 0◦ precone

4.0 Test 4.0 Test


Uniform inflow - Tip loss 0.98
Uniform inflow - Tip loss 0.98 Dynamic inflow, 1x1 - Tip loss 0.98
Uniform inflow - No tip loss Uniform inflow - No tip loss
Dynamic inflow, 1x1 - No tip loss
3.0 3.0
σ, rad/sec

σ, rad/sec
Damping -σ

Damping -σ

2.0 2.0

1.0 1.0

0.0 0.0
-15 -10 -5 0 5 10 15 -15 -10 -5 0 5 10 15
Collective pitch, deg Collective pitch, deg
(a) Constant coefficient eigenanalysis. (b) Transient response analysis with moving-block technique.

Fig. 27: Effect of dynamic inflow and tip loss on damping in hover of swept-tip blades with 2◦ precone

25
0.12
OVERFLOW/RCAS
RCAS 0.003 r/R = 0.93 0.01

0.002
0.1 0.005
M 2cn

0.001

M 2cc
0.08 0

M 2cm
0

0.06 -0.005
-0.001

0.04 -0.002 -0.01


0 90 180 270 360 0 90 180 270 360 0 90 180 270 360
azimuth angle, deg azimuth angle, deg azimuth angle, deg
(a) Normal force (b) Pitching moment (c) Chord force (+ve leading edge)

Fig. 28: steady state periodic airload comparisons at 93% R, swept-tip blade, µ = 0.30, αs = 0◦ , θ0 = 6◦ .

8.0 4.0 Test


Test RCAS
RCAS RCAS/OVERFLOW 2
Test
RCAS/OVERFLOW 2 2.0 RCAS
7.0 RCAS/OVERFLOW 2
lateral
Cyclic angles, deg
Collective, deg

0.0

6.0

-2.0 longitudinal

5.0
-4.0

4.0 -6.0
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
Advance ratio Advance ratio
(a) Collective angle (b) Cyclic angles

Fig. 29: Trim control angles, straight blade, αs = 0◦ , βp = 0◦ .

0.006

θ=6
o
0.004

swept
(α =0 , β =0 )
o o
T
C

s p

0.002
θ=3
o

0.000
straight
(α =-6 , β =2 )
o o
s p

-0.002
0 0.1 0.2 0.3 0.4 0.5
Advance ratio

Fig. 30: Variation of trim rotor thrust coefficient with collective pitch and rotor angle of attack as a function of advance ratio
based on RCAS/OVERFLOW 2 calculations.

26
Fig. 31: Lead-lag motions from blade tip lead-lag force excitation at different revolutions (red, blue, green) compared with
undeflected (gray), straight blade, µ = 0.30, αs = −6◦ , βp = 2◦ , θ0 = 6◦ , blades at 90◦ and 270◦.

80 RCAS
RCAS/OVERFLOW 2
53
CBM @12%R, ft-lb

27

-27

-53

-80

0 2 4 6 8 10 12

Rotor revolutions

Fig. 32: Comparison of transient response with blade tip lead-lag force excitation between RCAS and RCAS/OVERFLOW 2,
straight blade, µ = 0.30, αs = −6◦ , βp = 2◦ , θ0 = 6◦ .

27
3.0 3.0
Test Test
Blade 1 Blade 1
Blade 2 Blade 2
Blade 3 Blade 3
σ, rad/sec

σ, rad/sec
Blade 4 Blade 4
2.0 2.0
Damping -σ

Damping -σ
1.0 1.0

0.0 0.0
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
Advance ratio Advance ratio
(a) Lateral cyclic excitation (a) Lateral cyclic excitation
3.0 3.0
Test Test
Blade 1 Blade 2
Blade 2
Blade 3 Blade 3
σ, rad/sec

σ, rad/sec
Blade 4 Blade 4
2.0 2.0
Damping -σ

Damping -σ

1.0 1.0

0.0 0.0
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
Advance ratio Advance ratio
(b) Rotating pitch excitation (b) Rotating pitch excitation
3.0 3.0
Test Test
Blade 1 Blade 1
Blade 2 Blade 2
Blade 3 Blade 3
σ, rad/sec

σ, rad/sec

Blade 4 Blade 4
2.0 2.0
Damping -σ

Damping -σ

1.0 1.0

0.0 0.0
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
Advance ratio Advance ratio
(c) Blade tip force excitation (c) Blade tip force excitation

Fig. 33: RCAS/OVERFLOW 2 stability analysis of swept-tip Fig. 34: RCAS/OVERFLOW 2 stability analysis of straight
blade showing damping from moving-block transient analysis blade showing damping from moving-block transient analysis
of different blades, αs = 0◦ , βp = 0◦ , θ0 = 6◦ . of different blades, αs = −6◦ , βp = 2◦ , θ0 = 6◦ .

28
3.0
3.0 Test
Test
Constant coefficient
Constant coefficient

σ, rad/sec
σ, rad/sec

2.0
2.0

Damping -σ
Damping -σ

1.0
1.0

0.0
0.0
0 0.1 0.2 0.3 0.4 0.5
0 0.1 0.2 0.3 0.4 0.5
Advance ratio Advance ratio
(a) RCAS constant coefficient eigenanalysis
(a) RCAS constant coefficient eigenanalysis
3.0
3.0 Test
Test Lateral cyclic
Lateral cyclic pitch Rotating pitch
Rotating pitch Blade tip force
Blade tip force

σ, rad/sec
σ, rad/sec

2.0
2.0 Damping -σ
Damping -σ

1.0
1.0

0.0
0.0
0 0.1 0.2 0.3 0.4 0.5
0 0.1 0.2 0.3 0.4 0.5
Advance ratio Advance ratio
(b) RCAS moving-block transient response analysis
(b) RCAS moving-block transient response analysis
3.0
3.0 Test
Test Lateral cyclic
Blade tip force Rotating pitch
Blade tip force
σ, rad/sec
σ, rad/sec

2.0
2.0
Damping -σ
Damping -σ

1.0
1.0

0.0
0.0
0 0.1 0.2 0.3 0.4 0.5
0 0.1 0.2 0.3 0.4 0.5
Advance ratio Advance ratio
(c) Coupled RCAS/OVERFLOW moving-block transient
(c) Coupled RCAS/OVERFLOW transient response analysis
analysis

Fig. 35: Stability analysis of swept-tip blade, αs = 0◦ , Fig. 36: Stability analysis of swept-tip blade, αs = 0◦ ,
βp = 0◦ , θ0 = 3◦ . βp = 0◦ , θ0 = 6◦ .

29
3.0 3.0 Test
Test
Constant coefficient Constant Coefficient
σ, rad/sec

σ, rad/sec
2.0 2.0
Damping -σ

Damping -σ
1.0 1.0

0.0 0.0
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
Advance ratio Advance ratio
(a) RCAS constant coefficient eigenanalysis (a) RCAS constant coefficient eigenanalysis
3.0 3.0
Test Test
Lateral cyclic pitch Lateral cyclic pitch
Rotating pitch Rotating pitch
Blade tip force Blade tip force
σ, rad/sec

σ, rad/sec
2.0 2.0
Damping -σ

Damping -σ

1.0 1.0

0.0 0.0
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
Advance ratio Advance ratio
(b) RCAS moving-block transient analysis (b) RCAS moving-block transient analysis
3.0 3.0
Test Test
Lateral cyclic pitch
Blade tip force
Rotating pitch
Blade tip force
σ, rad/sec

σ, rad/sec

2.0 2.0
Damping -σ

Damping -σ

1.0 1.0

0.0 0.0
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
Advance ratio Advance ratio
(c) Coupled RCAS/OVERFLOW moving-block transient (c) Coupled RCAS/OVERFLOW moving-block transient
analysis analysis

Fig. 37: Stability analysis of straight blade, αs = −6◦ , Fig. 38: Stability analysis of straight blade, αs = −6◦ ,
βp = 2◦ , θ0 = 3◦ . βp = 2◦ , θ0 = 6◦ .

30

You might also like