Sie sind auf Seite 1von 7

JOURNAL OF AIRCRAFT

Vol. 37, No. 3, May– June 2000

Validation of a Rotorcraft Mathematical


Model for Autogyro Simulation
S. S. Houston¤
University of Glasgow, Glasgow, Scotland G12 8QQ, United Kingdom

Results are presented from a study undertaken to validate a rotorcraft mathematical model for simulation
of autogyro  ight. Although this class of rotary-wing aircraft has found limited application in areas other than
sport or recreational  ying, commercial interest is increasing. A sparse contemporary literature on autogyro  ight
emphasizes the timeliness of this work, which takes advantage of  ight experiments using a fully instrumented
autogyro.Validationis based on comparisons of trim, linearized six-degree-of-freedom derivatives, and time history
response of the full nonlinear model to control inputs. The validation process is a vital ingredient in deŽ ning the
applicability of a mathematical model as an engineering tool for supporting studies into aircraft stability, control,
Downloaded by UNIV OF TEXAS AT SAN ANTONIO on February 2, 2015 | http://arc.aiaa.org | DOI: 10.2514/2.2640

and handlingqualities. The model is of generic, nonlinear, individualblade/blade element type, and its conŽ guration
as an autogyro is described. It is concluded that simulation of autogyro  ight presents no particular difŽ culties
for a generic rotorcraft model. Limitations in predictive capability are clearly delineated, although in general
comparisons between  ight and theory are good.

Nomenclature aero
Xelem , Xelem , = blade element contributions to rotor force;
inertial
A = matrix of state vector acceleration coefŽ cients; Xelem total, aerodynamic, inertial, N
linearized model system matrix X u , X w , etc. = longitudinal body axis acceleration derivatives,
A 11 , etc. = minors of linearized system matrix A 1/s, etc.
a xhinge, a hinge
z = hinge acceleration components, m/s2 x = state vector
B1 , B2 = minors of linearized control matrix x1 , x2 = subspace of linearized model state vector
C0 = apparent mass factor Yv , Y p , etc. = lateral body axis acceleration derivatives, 1/s,
I ap , Ipitch , Ilag = blade moments of inertia, Nm2 m/(rad s), etc.
L = rotor moment vector, Nm Z u , Z w , etc. = vertical body axis acceleration derivatives, 1/s,
[L ] = dynamic in ow static gain matrix etc.
L aero = aerodynamic rolling moment, Nm g c = lateral stick position, % (0% fully left)
L v , L p , etc. = rolling moment derivatives, 1/(ms), 1/s, etc. g s = longitudinal stick position, % (0% fully
Maero = aerodynamic pitching moment, Nm forward)
Mblap , Mlag bl
= blade  ap and lag moments, Nm q = air density, kg/m 3
Mu , Mw , etc. = pitching moment derivatives, 1/(ms), etc. [s ] = time constant matrix, s
Nv , N p , etc. = yawing moment derivatives, 1/(ms), 1/s, etc. u ,h = roll and pitch attitude, rad
nblades = number of blades v = wake skew angle, rad
nelem = number of blade elements w = azimuthal position on rotor disc, rad; yaw
p, q, r = angular velocity components about body axes, attitude, rad
bl bl bl
rad/s x x , x y, x z = blade angular velocities, rad/s
Q u , Q w , etc. = rotor torque derivatives, rev/min/(m), etc.
R = rotor radius, m Subscript
r = radial position on rotor disk, m
wind = wind axes
rcg = position of center of mass, body axes, m
rhinge = position of  ap, lag and feather hinge, rotor
axes, m Introduction
rhub
Taero
T1 , T2 , T3
= position of rotor hub, body axes, m
= aerodynamic thrust, N
= transformation matrices
T HE autogyrohelped to pave the way for the development of the
helicopter,introducingcyclic pitch control and blades attached
to the rotor hub by means of a hinge. However, it has found no prac-
u, v, w = translational velocity components along body tical application in commercial or military roles, although the type
axes, m/s is very popular for sport and recreational  ying. The literature on
u = control vector autogyros (or gyroplanes) is considerable,for example, Refs. 1 – 11,
u hub , v hub , = rotor hub velocities in non-rotating but this work is now primarily of historical signiŽ cance. However,
w hub reference frame, m/s a program of study was funded by the U.K. Civil Aviation Author-
v i (r, w ) = induced velocity at position (r, w ), m/s ity, following a spate of fatal accidents.12 This work has sought to
vim = momentum induced velocity, m/s establish a rigorous and consistent understanding of the factors in-
v i 0 , v 1s , v 1c = components of induced velocity, m/s  uencing autogyro stability, control, and handling qualities. A pro-
vm = wake mass  ow velocity, m/s gram of mathematical modeling, simulation, wind-tunnel measure-
vT = wake velocity, m/s ments, and  ight testing has examined the characteristicsof a single
X = rotor force vector, m type, the VPM M16 (Fig. 1). Results have been reported elsewhere;
Refs. 13– 16 separately describe factors in uencing longitudinal
Received 30 November 1998; revision received 15 October 1999; ac- stability, wind-tunnel testing of a scale-model airframe, and exper-
cepted for publication 19 November 1999. Copyright ° c 2000 by the Amer- iments to gather  ight-test data for parameter estimation exercises.
ican Institute of Aeronautics and Astronautics, Inc. All rights reserved. Developments in rotorcraft simulation have been driven exclu-
¤ Senior Lecturer, Department of Aerospace Engineering. Member AIAA. sively by the helicopter problem, and models have undergone
403
404 HOUSTON

Table 1 Mathematical model description

Model item Characteristics


Rotor dynamics Up to 10 individually-modeled rigid blades
(both rotors) Fully coupled  ap, lag and feather motion
Blade attachment by offset hinges and springs
Linear lag damper
Rotor loads Aerodynamic and inertial loads represented
by up to 10 elements per blade
Blade aerodynamics Lookup tables for lift and drag as function
of angle of attack and Mach number
Wake model Peters’ dynamic in ow model
Uniform and harmonic components of in ow
Rudimentary interaction with tail surfaces
ground effect
Transmission Coupled rotorspeed and engine dynamics
Fig. 1 VPM M16 autogyro.
up to 3 engines
Geared or independently controlled rotor torque
continuous development to address deŽ ciencies in their predictive Airframe Fuselage, tailplane and Ž n aerodynamics by
lookup tables or polynomial functions
ability that have been highlighted by  ight test. They have pro-
Atmosphere International Standard Atmosphere
gressed from simple forms that would be familiar to a Ž xed-wing
Downloaded by UNIV OF TEXAS AT SAN ANTONIO on February 2, 2015 | http://arc.aiaa.org | DOI: 10.2514/2.2640

provision for variation of sea-level


 ight dynamicist, that is, linearized, small perturbation derivative temperature and pressure
models, to highly nonlinear models, where each blade is individu-
ally represented.The literature contains several key examples of the
need for improving models beyond simple forms. Hansen17 high- tions will have three main rotor controls and one tail rotor control.
lighted the importance of blade  ap dynamics. Chen and Hindson18 Depending on the number of blades on each rotor, there can be up
examined the signiŽ cance of coupled  ap/in ow dynamics in hov- to 100 nonlinear,periodic ordinarydifferentialequations describing
ering  ight, and Tischler19 showed that modeling coupled roll/lag the coupled rotor/airframe behavior.Note that the matrix A contains
dynamics was necessary in the context of control system design. All off-diagonal terms not simply associated with airframe products of
three papers deal with linearized models, two in derivative form and inertia, but that also include terms associated with the dynamic in-
one in transfer function form, and the modeling uses disk, rather  ow and blade equations of motion.
than individual blade, representations of the rotors. However, the The model has been used previously for helicopter validation and
work does highlight the importance of higher-order dynamics, and simulation studies23,24 and simulation of autogyros.13 The induced
in recent years, the associated degrees of freedom have been incor- velocity model, force and moment expressions,and the blade equa-
porated in individual blade models that have been developed for tions of motion are now summarized.
piloted simulation.20¡ 22 First, the dynamic in ow representation used is taken from
Although an engineering model is a powerful tool for supporting Chen,25 although the original model development is due to Gaonkar
studies into aircraft stability, control, and handling qualities, conŽ - and Peters26 and Peters and HaQuang.27 The basic form of induced
dence in the results can only be quantiŽ ed if the model has a known velocity at any azimuth and radial station over the rotor is given by
level of validity. Formal approaches to the general problem of vali-
dation do not exist, but model Ž delity is generallydescribedin terms v i (r, w ) = v i0 + (r / R)v 1s sin w + (r / R)v 1c cos w (2)
of amplitude and frequency, and here this is addressed by compar-
ing trim, linearized small-perturbationderivatives and the response The induced velocity v i (r, w ) appears explicitly in the aerodynamic
to large-amplitude control inputs. The objective of this paper is to model, contributing to the blade element angle of attack. The three
validate a generic rotorcraft mathematical model for simulation of states v i0 , v 1s , and v 1c and are calculated from
those aspects of autogyro  ight that are important in the analysis
of stability, control, and handling qualities issues. SpeciŽ c aims of 2 3 2 3 2 3
Çv i 0 vi0 Taero
the work are, Ž rst, to demonstrate that a typical generic rotorcraft 4 5 4 5 4 5
mathematicalmodel is capable of simulating autogyro ight without [s ] Çv 1s =¡ v 1s + [L ] L aero (3)
special modiŽ cation and, second, to deŽ ne the predictive capability Çv 1c wind v 1c wind Maero wind
across a wide range of operating conditions from steady  ight to the
transientresponse to large control inputs. The paper makes a unique where
contribution to the Ž eld because there is no comparable work in the 2 3
4R ¡ R tan(v / 2)
literature that deŽ nes a level of validity for a mathematical model 6 3p v C 0 7
simulating rotorcraft  ight in autorotation. 6 T 0 12v m 7
6 7
6 64R 7
Mathematical Model [s ] = 6 0 0 7
6 45p v m (1 + cos v ) 7
6 7
Model Description 45R tan(v / 2) 64R cos v 5
The model has been in uenced strongly by the literature in the 0
8v T 45p v m (1 + cos v )
Ž eld, taking account of the signiŽ cance of research by others into
individual blade/blade element modeling and dynamic in ow mod- (4)
eling. The key features of the mathematical model used are summa- and
rized in Table 1. The model takes the form 2 3
R 15p tan(v / 2)
(1) 6 0 7
A xÇ = f (x, u) 6 2v T 64v m 7
6 7
1 6 ¡ 4 7
where the state vector x contains the airframe translational and an- L = 6 0 0 7
gular velocity, blade  ap, lag and feather angles and rates for each q p R3 6 v m (1 + cos v ) 7
6 7
blade on each rotor, the induced velocity states for each rotor wake, 4 15p tan(v / 2) ¡ 4 cos v 5
as well as the angular velocity of both rotors, and the engine torques. 0
64v T v m (1 + cos v )
Elements of the control vector u are the four controls, which vary
with aircraft type, for example, single main and tail rotor conŽ gura- (5)
HOUSTON 405

Equation (3) is transformed from wind to body axes in this appli- teeter bar, suspended from a teeter bolt. The blades are untwisted,
cation. The reader is referred to the original literature for detail and and no cyclic pitch can be applied. This hub assembly is mounted
history of the development of these equations. on a spindle, about 200 mm long, and this spindle pivots about its
Second, the rotor forces and moments are given by lower end to tilt the entire rotor fore and aft and laterally to effect
X
nblades
± (
X
nelem
)! pitch and roll control, respectively. The rotor system is linked me-
chanically by rods to the pilot’s control column, and a conventional
X = T1¡ 1 T2¡ 1 T3¡ 1 Xelem (6) rudder is operated by a cable assembly from foot pedals.
j =1 i =1 Two aspects of this aircraft require consideration in relation to

X
nblades
± " ± (
X
nelem
)!#!
! the model formulation outlined earlier. First is autorotationbecause
the literature suggests that no consideration has been given to the
L = T1¡ 1 T2¡ 1 ¡ rhinge x T3¡ 1 Xelem use of the dynamic in ow model in this mode. Consider v T , v m , and
j =1 i =1 v . By the neglecting of the contribution of the longitudinal tilt of
the rotor disk, which simulation results have shown to be at least an
+ (rhub ¡ rcg )xX (7) order of magnitude smaller than the rotor angle of attack, then
q ¡ ¢2
where
vT = u 2hub + v hub
2 + v im ¡ w hub (13)
aero
Xelem = Xelem + inertial
Xelem (8)
© ¡ ¢£ ¤ª
v m = u 2hub + v hub
2
+ v i m ¡ w hub 2v im ¡ w hub ê v T (14)
The matrices T1 , T2 , and T3 transform quantities from blade ele-
ment to rotor and nonrotating airframe axes. Development of these p
Downloaded by UNIV OF TEXAS AT SAN ANTONIO on February 2, 2015 | http://arc.aiaa.org | DOI: 10.2514/2.2640

equations is given in Ref. 23. u 2hub + v hub


2
tan v = (15)
Finally, the blade equations of motion are based on the derivation v im ¡ w hub
given in Ref. 28 and are given by
In autorotation, w hub > v i m . Therefore the velocities v T and v m will
I ap x Ç bl
x +x bl
yx
bl
z
bl hinge
¡ m bl ycg az = Mblap be positive, and for v T in particular this means that v i 0 > 0, that is,
the  ow slows down on passing through the rotor, as it should in
bl bl bl bl
Ipitch x Ç y ¡ x x x z = Mpitch autorotation.Note that v < 0 due to w hub > v i m is geometricallycon-
sistent with  ow in autorotation, but it does mean that those terms
Ilag x Ç bl
z
+x bl
x x
bl
y
+ m bl ycg
bl hinge
ax bl
= Mlag (9) in Eq. (5) that couple v i 0 and v 1c change sign. The signiŽ cance in
relation to the physical attributes of the real  ow is not clear, but
Again, expressionsfor the angular velocitiesx bl bl bl the autogyro simulation results are similar to the helicopter case in
x , x y , and x z are de-
rived in Ref. 23. Hinge acceleration terms a xhinge, and ahinge
z are de- that there is an upwash component at the front of the rotor disk and
rived using standard kinematic principles.Note that the applied mo- downwash at the rear. This is consistent with the intuitive result for
ment terms Mblap and Mlagbl
include spring restraint terms, to be used circular wings, irrespective of powered or autorotative operation.
if appropriate, and the lag degree of freedom embodies a very rudi- Inspection of these equations indicates that the model will emulate
mentary lag damper term. physical aspects of autorotative behavior, but its efŽ cacy for simu-
The model is, therefore, a very conventional individual blade/ lation of autorotation will be deŽ ned by the validation results. For
blade element representationof a generic two-rotor aircraft. The ro- example, the lateral stick position is an indirect validation of v 1c . It
tor module is called twice in the simulation code, each rotor being is the case that boundariesdeŽ ning the various rotor operatingstates
discriminatedby data that specify its location and orientationon the for axial  ight have been determined using simple momentum the-
airframe and its characteristics in terms of blade mass distribution, ory, for example, Refs. 28 and 29. Figure 2 shows the location on
hinge offset and restraint, etc. Trim and linearization is performed such a diagram, taken from Ref. 29, of the 14 autogyro conŽ gura-
using the procedure described in Ref. 23. The autogyro conŽ gura- tions simulated here. The arrow indicates progression from low to
tion, however, requires the linearizationprocess to reduce the model high airspeed. It is perhaps misleading to represent these autogyro
to the form results on such a chart, which is strictly applicable only to axial, as
³ ´ ³ ´³ ´ ³ ´ opposed to edgewise,  ight. Whatever the applicability, Fig. 2 does
xÇ 1 A 11 A12 x1 B1 serve to represent these autogyro simulations in the context of the
= + u (10)
very limited literature available on autorotation.
xÇ 2 A 21 A22 x2 B2
The second aspect of this autogyro conŽ guration requiring con-
where the conventional six-degree-of-freedom(6-DOF), nine-state siderationis setting the hinge offset term in the model to zero, which
model structure is represented by can satisfactorily emulate the  ight mechanics of teetering rotor
systems.29 In effect, the simulation then models a rotor system with
xÇ 1 = A11 x1 + B1 u, x1 = [u, v, w, p, q, r, u h w ]T (11) two centrally hinged blades (although the coning mode of a two-
bladed articulated rotor is dissimilar to that of a teetering rotor).
and

Çx2 = A 21 x1 + A 22 x2 + B2 u, x2 = [X ] (12)

is the rotorspeed DOF. The minor A 12 then couples rotor speed into
the conventional rigid-body DOF. Note that the treatment of X is
in accordance with convention for derivative calculation, that is, it
is held constant for calculation of elements in A11 , A21 , B1 , and B2
and only perturbed for deriving the elements in A22 and A12 .

ConŽ guration of the Model as an Autogyro


The aircraft used in this study was the VPM M16 gyroplane
(Fig. 1). It is of Italian origin, produced in kit form for assembly
by the owner. The maximum all-up mass is 450 kg. The aircraft is
powered by a four-cylindertwo-stroke engine driving a three-bladed
Ž xed-pitch propeller. For helicopter engineers not familiar with gy-
roplanes, the rotor system is of an interesting conŽ guration, typical Fig. 2 Model operating condition in context of typical vertical descent
of this class of aircraft. The two main rotor blades are bolted to a  ow states.
406 HOUSTON

Once again,validationwill determine how appropriatethis approach some extent rotor speed. Care is required in assessing the source
is. The simulation is sufŽ ciently generic to allow other aspects of of discrepancies in rotor speed because it is very sensitive to a va-
the rotor system and propeller geometry to be accommodated by riety of parameters such as blade twist (which may manifest itself
appropriate data entries or software switches to enable or disable aeroelastically in  ight and is not incorporated in the rigid-blade
particular features. model), blade incidence angle relative to the hub, and section drag
Althoughautogyroshave been simulatedpreviously,30,31 the mod- characteristics. The pitch attitude discrepancy could be associated
els and the results obtained were severely limited in their applicabil- with airframe lift and pitching moment. However, wind-tunnel data
ity. For example, simple disk representationsof the main rotor were for this airframe were obtained for a wide range of angle of attack
used and analytical expressions for linearized stability and control and sideslip.14 The contribution of the airframe is weak in relation
derivatives were formulated. Reference 31 was further limited by to the rotor and propeller, especially so at the lower speeds, which
using data for a generic aircraft type, that is, a speciŽ c aircraft was tends to suggest that the source of pitch attitude modeling errors
not modeled. No validation was attempted in either case. lies elsewhere. For example, a simple interpretation of rotor angle
of attack
Results
The simulation model was conŽ gured with data appropriateto the a rotor =g s +h +b 1c (16)
 ight conditionfor each test point. Mass and inertia informationwas
available from ground-based measurements. The aircraft mass for tends to suggest that in ow modeling at low speed may be the source
each test point was in uenced only by fuel state, which could not be of this particular discrepancy. This is because g s is accurately pre-
measured accurately.Accordingly, simulation results were obtained dicted, and b 1c is typically small (less than 2 deg across the speed
range), and so the deŽ ciency in pitch attitude at low speed is most
Downloaded by UNIV OF TEXAS AT SAN ANTONIO on February 2, 2015 | http://arc.aiaa.org | DOI: 10.2514/2.2640

for the two limiting conditions of full fuel (maximum weight) and
zero fuel (minimum weight). probablydue to simulation of rotor angle of attack. Because the deŽ -
ciency is speed dependent,it is argued that this points to a weakness
Equilibrium Flight that is related to physical modeling, rather than inappropriate data
such as blade section lift characteristics.
Comparison of the model with measurements taken in steady,
It is also possible to speculate that an in ow modeling deŽ -
level  ight is shown in Fig. 3. Generally, there is good agreement
ciency may be responsiblefor the lateral stick position result, where
between model and  ight. Simulation of longitudinal stick position
the model underestimates  ight by 5 – 10% across the speed range.
compares very favorably with the  ight measurement throughout
Figure 4 shows the impact of setting the harmonic components of
the speed range, indicating that it gives a good prediction of speed
induced velocity (v 1s and v 1c ) to zero. (Note that v 1s has a negligible
stability.Although lateral stick position suffers from a uniform error
impact on lateral stick position, but was set to zero for consistency.)
of about 5% across the speed range, this is less than 0.5 in. of
This has little impact above 50 mph, but does improve the result
stick displacement. Roll attitude is simulated to within 0.5 deg.
in the lower-half of the speed range. However, a perusal of Chen’s
The trend in simulated pitch attitude with speed is consistent with
survey paper25 indicates that, at least for powered rotors, there is no
 ight, although this trend is more accentuated in the simulation
evidence that would tend to support the postulate that v 1c might be
giving rise to relatively poor prediction at low speed. Rotor speed
zero.
trend with speed is simulated accurately by the model, although
uniformly in error by about 20– 25 rpm. This is considered a good
result because rotor speed is not governed to a given value as in Small-Perturbation Flight: Longitudinal
helicopter applications. Figure 5 shows comparisonsof key longitudinalstability and con-
The deŽ ciencies in validity can be summarized as pitch attitude trol derivatives obtained using the linearization approach outlined
at low speeds, lateral stick position across the speed range, and to earlier and those from  ight test.15 The  ight-derived values are
presented as 95% conŽ dence bounds, whereas the model results are
those for minimum and maximum mass.
Some simulation and  ight derivatives show a consistent trend
with speed. Simulation values can lie within, or overlap, the un-
certainty bounds of the  ight results. It is argued that in such cases,
model and  ightcan be said to be in agreement,within the limitations
of derivativeidentiŽ ability and modeling uncertainty.However, cer-
tain cases do point to modeling inadequacy that can be classed
as moderate to severe. For example, the drag damping term X u ,
which is of paramount importance in phugoid mode damping, is
very poorly predicted by the model. Similarly, heave damping Z w
plays an important role in short-period mode damping, but is sub-
stantially overestimatedby the simulation model. Of less concern is
the mismatch in Mw . The model results appear simply shifted rela-
tive to the  ight values; signiŽ cantly, the model captures the unusual
result for unaugmented rotorcraft, that Mw < 0. A physical reason-
ing for this result is given in Ref. 15. Coupling of rotor speed with

Fig. 4 Effect of remov-


ing harmonic components
of induced velocity on
lateral stick position: £ ,
 ight data; –n – , model,
minimum weight; –u – ,
model, maximum weight;
- -n - -, model, minimum
weight no in ow harmon-
ics; and - -u - -, model,
Fig. 3 Flight and simulation comparisons of key trim parameters: maximum weight no in-
£ ,  ight data; n , model, minimum weight; and u , model, maximum  ow harmonics.
weight.
HOUSTON 407

Table 2 Comparison of  ight and simulation


lateral/directional derivatives
Flight Model
Lower Upper Minimum Maximum
Parameter bound bound weight weight
Yv ¡ 0.133 ¡ 0.044 ¡ 0.194 ¡ 0.179
Lp ¡ 2.727 ¡ 2.377 ¡ 2.714 ¡ 2.822
Lg c 0.060 0.067 0.078 0.085
Lv 0.028 0.041 0.172 0.142
Nv 0.055 0.063 0.064 0.060
Nr ¡ 0.996 ¡ 0.812 ¡ 0.104 ¡ 0.091
Downloaded by UNIV OF TEXAS AT SAN ANTONIO on February 2, 2015 | http://arc.aiaa.org | DOI: 10.2514/2.2640

Fig. 6 Flight and simulation comparisons of key rotor torque deriva-


tives: £ ,  ight data; n , model, minimum weight; and u , model, maxi-
mum weight.

Fig. 5 Flight and simulation comparisons of key longitudinal deriva-


tives: £ ,  ight data; n , model, minimum weight; and u , model, maxi-
mum weight.

the pitch DOF manifested in terms of MX is reasonably represented


in the simulation model, except at high speed.
The rotor torque derivatives [Eq. (12)] are unique to rotors in au-
torotation and are, therefore, of major importance to simulation of
Fig. 7 Comparison of  ight-identiŽ ed and model eigenvalues: j ,
autogyros. In addition, given the sensitivity of rotor speed to a mul-
tiplicity of factors, good modeling of these terms gives conŽ dence ²
shortperiod; , rotorspeed; , phugoid; ¦ , phugoid, model; ±, rotor-
speed, model; and u , shortperiod, model.
that the physical modeling is accurate. Some of the key derivatives
in Eq. (12) are presented in Fig. 6, which shows that the simulation
model results exhibit excellent agreement with  ight. is consistent with Glauert’s elementary theory.1 Paradoxically, in
The deŽ ciencies in X u and Z w and to a lesser extent Mw have a Ref. 15 it is argued that the  ight result is consistent with observed
salient impact on the prediction of the modes of motion. Figure 7 aircraft behavior. It is difŽ cult, therefore, even to speculate as to
shows a comparison between  ight and simulation model longitudi- the source of this discrepancy without further  ight tests on aircraft
nal eigenvalues for 30 and 70 mph. Arrows indicate the progression dissimilar to that used for the present study.
from low to high speed. In general terms, the short-period mode
damping is greater in the simulation model because of the over- Small-Perturbation Flight Lateral/Directional
estimate in Z w , which could be attributed to errors in modeling Table 2 compares key lateral/directional derivatives from simu-
rotorspeed. Blade torsion can have a pronounced effect on thrust, lation and  ight. The  ight results are taken from Ref. 16, which
and this may also have a role to play. The slight difference in short- focused on 70 mph. The rolling moment derivatives L p and L g c
period mode frequency is due to the offset in M w across the speed are dominated by the rotor system, and the simulation displays fair
range. This could be due to airframe aerodynamics, but previous correlation with  ight. As in the case of the pitch axis, the model
work13 has shown this derivative to be sensitive to the vertical lo- overestimates the control derivative L g c . The rotor makes a weak
cation of the center of mass in relation to the propeller thrust line. contribution to L v , which is strongly in uenced by airframe aero-
The phugoid mode damping is of much greater concern. Grossly dynamics. The  ight results indicate that the aircraft displays lack
overestimated by the model, this is directly attributable to the poor of dihedral effect, which is because the airframe side area is low rel-
prediction of X u . In the simulation, airframe drag characteristics ative to the center of mass. The simulation overpredicts the severity
account for about 10% of X u , the bulk coming from the rotor. This of this conŽ guration effect.
408 HOUSTON

The directional derivatives are dominated by airframe aerody-


namics, which will also be in uenced by propeller slipstream. The
simulation model, therefore, depends crucially on the quality of the
wind-tunnel test data if good correlation between simulation and
 ight is to be obtained. It is clear that correlation varies from rea-
sonable to extremely poor. Weathercock effect N v is reasonably
predicted by the model, but correlation in yaw damping Nr is ex-
tremely poor.

Response to Large-Amplitude Inputs


Figures 8 and 9 show comparisons between  ight and the full
nonlinear,individualblade/blade element model, for lateral and lon-
gitudinal stick inputs at a nominal 70 mph. Note that the simulation
displays the high-frequency periodicity associated with modeling
individual blades.
The Ž rst case considered is shown in Fig. 8. The lateral control
stick input measured in  ight was used to drive the full nonlinear,
individual blade/blade element model. The shape and magnitude of
the primary response in roll rate is modeled well by simulation up
Downloaded by UNIV OF TEXAS AT SAN ANTONIO on February 2, 2015 | http://arc.aiaa.org | DOI: 10.2514/2.2640

to about 10 s. This is consistent with the result in Table 2 for L p


and L g c , which have a dominant effect on short-term roll response.
After the Ž rst response peak, the model responds more sharply than
the real aircraft. This is not manifestation of a lag due to blade  ap
dynamics because the interval is too great and in any case such
a lag would be apparent during the initial response. The phasing
of the off-axis response in pitch rate is surprisingly well modeled
up to 10 s as well. Cross coupling remains one of the particular
challenges in helicopter modeling, and so why the coupling is so
well predicted with the model in the autogyro mode is not clear.
Beyond 10 s, the real aircraftis very well damped,whereas the model
enters a lightly damped Dutch roll oscillation. This is consistent
with comparisons in Table 2, where two derivatives that directly
in uence the Dutch roll, L v and Nr , compare poorly with  ight. Fig. 9 Response to longitudinal stick, comparison of  ight and simu-
lation: – – – ,  ight data, and ——, simulation.
The rotor speed response correlationis not as poor as it Ž rst appears;
there is little short-term response, and this is re ected by the model.
Thereafter, the incorrectly modeled Dutch roll mode dominates the
simulation result, whereas the real aircraft, as already noted, is well
damped.
Figure 9 shows comparisons of response to a longitudinal stick
input. The longer-term rotorspeed response, that is, greater than
10 s, is more heavily damped in the simulation model than in- ight.
This is entirely consistent with linearized model validation, which
revealedthat the real aircraft has little drag damping and, as a result,
little phugoid or long-term damping. The opposite is true with the
linearized simulation model. The pitch rate response is simulated
fairly well, the only anomalous period in time being around 10 s,
consistent with the mismatch in rotor speed. However, the off-axis
response in roll rate displays behavior familiar to helicopter  ight
dynamicists, where the amplitude is reasonably predicted in mag-
nitude, but not in phase.

Discussion
The criteria used to assess whether or not the rotorcraftmodel can
adequately simulate the autogyro are necessarilysubjective, relying
on engineering judgment because there are no formal criteria for
autogyro simulation. However, comparisons with Federal Aviation
Administrationlevel D simulator requirements for helicopters32 can
be instructive.They are expressed in terms of trim and time response
comparisons, and the model generally satisŽ es the requirements
for control position prediction to be within 5% and attitude to be
within 1.5 deg. Primary axis time responseswould fall within accept-
able envelopes, although the cross coupling would not. The criteria
for rotorspeed simulation in autorotation is not pertinent to the
autogyro.
Although there are no criteria for derivative comparisons,
Tischler33 is developing criteria for frequency-responsevalidation.
It is accepted that the linearized 6-DOF model structure is of limited
utility in the context of frequency-responsemethods, being applica-
Fig. 8 Response to lateral stick, comparison of  ight and simulation: ble to a much narrowerbandwidth.However, the validationenvelope
– – – ,  ight data, and ——, simulation. deŽ ned by the approach taken in this paper is consistent with pilot
HOUSTON 409

control strategy for autogyros, which is essentially low frequency. 8 Wheatley, J. B., and Hood, M. J., “Full-Scale Wind-Tunnel Tests of a

In addition, the derivative allows a high degree of insight into the PCA-2 Autogyro Rotor,” NACA TR 515, 1935.
9 Wheatley, J. B., “An Analytical and Experimental Study of the Effect
causes of speciŽ c modeling deŽ ciencies,such as the heavilydamped
phugoid X u or under-dampedDutch roll, L v and Nr , which is absent of Periodic Blade Twist on the Thrust, Torque and Flapping Motion of an
with the aggregate presentation of a frequency response. Autogyro Rotor,” NACA TR 591, 1937.
10 Schad, J. L., “Small Autogyro Performance,” Journal of the American
Notwithstanding the lack of formal criteria, it is argued that vali- Helicopter Society, Vol. 10, No. 3, 1965, pp. 39– 43.
dation discrepanciesare typical of that obtained for helicopter mod- 11 McKillip, R. M., and Chih, M. H., “Instrumented Blade Experiments
els. For example, the poorly predicted directional characteristicsare Using a Light Autogyro,” Proceedings of the 16th European Rotorcraft Fo-
determinedlargely by airframe aerodynamicsthat are obtainedfrom rum, Glasgow, Scotland, 1990, pp. II.8.4.1– II.8.4.8.
wind-tunnel tests. However, discrepancies in the drag and heave 12 “Airworthiness Review of Air Command Gyroplanes,” Air Accidents

damping derivatives X u and Z w could be directly attributed to mod- Investigation Branch Rept., Sept. 1991.
13 Houston, S. S., “Longitudinal Stability of Gyroplanes,” Aeronautical
eling of physicalphenomenaunique to the autogyro,speciŽ cally the
induced velocity in autorotation. However, the functionality of the Journal, Vol. 100, No. 991, 1996, pp. 1 – 6.
14 Coton, F., Smrcek, L., and Patek, Z., “Aerodynamic Characteristics
Peters wake dynamic in ow model for the autogyro appears accept-
of a Gyroplane ConŽ guration,” Journal of Aircraft, Vol. 35, No. 2, 1998,
able, and given that no previousautogyromodel validationhas taken pp. 274– 279.
place, a deŽ nitive statement cannot be made until the predictiveabil- 15 Houston, S. S., “IdentiŽ cation of Autogyro Longitudinal Stability and
ity of the model for simulating other types reveals whether or not Control Characteristics,” Journal of Guidance, Control, and Dynamics,
errors in X u and Z w , in this case, are indicative of a generic problem Vol. 21, No. 3, 1998, pp. 391– 399.
16
or are type speciŽ c. Note that if the model can simulate autogyro Houston, S. S., “IdentiŽ cation of Gyroplane Lateral/Directional Stabil-
Downloaded by UNIV OF TEXAS AT SAN ANTONIO on February 2, 2015 | http://arc.aiaa.org | DOI: 10.2514/2.2640

 ight, it should in principle be capable of simulating helicopters in ity and Control Characteristics from Flight Test,” Proceedings of the Institute
this  ight regime as well. It is, therefore, important that it is capable of Mechanical Engineers, Pt. G, Vol. 212, No. G4, 1998, pp. 1 – 15.
17 Hansen, R. S., “Toward a Better Understanding of Helicopter Stability
of being used without major modiŽ cation to the underlyingphysical
Derivatives,” Journal of the American Helicopter Society, Vol. 29, No. 1,
modeling.
1984, pp. 15– 24.
18 Chen, R. T. N., and Hindson, W. S., “In uence of Dynamic In ow on
Conclusions
the Helicopter Vertical Response,” Vertica, Vol. 11, Nos. 1 and 2, 1987, pp.
It would appear that generic rotorcraft simulation models can be 77– 91.
used to simulate autogyro  ight without any modiŽ cation to the 19 Tischler, M. B., “Digital Control of Highly Augmented Combat Rotor-
physical form of the governing equations, although further experi- craft,” NASA TM-88346, 1987.
mental work with different autogyros is required before a deŽ nitive 20 Lehmann, G., Oertel, C.-H., and Gelhaar, B., “A New Approach in Heli-

statement can be made. This is because the validation against a copter Real-Time Simulation,” Proceedings of the 15th European Rotorcraft
speciŽ c aircraft type over a range of amplitude and frequency has Forum, Amsterdam, 1989, pp. 62.1– 62.17.
21 DuVal, R. W., “A Real-Time Blade-Element Helicopter Simulation for
revealed limitations in Ž delity that may be unique to the type in
question, rather than the autogyro problem in general. However, Handling Qualities Analysis,” Proceedings of the 15th European Rotorcraft
Forum, Amsterdam, 1989, pp. 59.1– 59.18.
in terms of trim and rotor torque derivatives, the model emulates 22 Meerwijk, L., and Brouwer, W., “Real-Time Helicopter Simulation Us-
satisfactorily the variability in rotor speed that is a consequence of ing the Blade-Element Method,” Proceedings of the 17th European Rotor-
operation in autorotation. The results are, therefore, of wider rele- craft Forum, DGLR, Berlin, 1989, pp. 599– 616.
vance and signiŽ cance, speciŽ cally to the simulation of helicopters 23 Houston, S. S., “Validation of a Non-Linear Individual Blade Rotor-
in autorotation. craft Flight Dynamics Model Using a Perturbation Method,” Aeronautical
Journal, Vol. 98, No. 977, 1994, pp. 260– 266.
Acknowledgment 24 Houston, S. S., “Validation of a Blade-Element Helicopter Model for

This work was conducted for the U.K. Civil Aviation Author- Large-Amplitude Manoeuvres,” Aeronautical Journal, Vol. 101, No. 1001,
1997, pp. 1 – 7.
ity under Research Contract 7D/S/1125. The Project Manager was 25 Chen, R. T. N., “A Survey of Non-Uniform In ow Models for Rotorcraft
David Howson. Flight Dynamics and Control Applications,” Vertica, Vol. 14, No. 2, 1990,
pp. 147– 184.
References 26 Gaonkar, G. H., and Peters, D. A., “A Review of Dynamic In ow Mod-
1 Glauert, H., “A General Theory of the Autogyro,” Aeronautical Research
eling for Rotorcraft Flight Dynamics,” Vertica, Vol. 12, No. 3, 1988, pp. 213–
Committee Repts. and Memoranda 1111, London, Nov. 1926. 242.
2 Lock, C. N. H., “Further Development of Autogyro Theory Parts I and II,” 27 Peters, D. A., and HaQuang, N., “Dynamic In ow for Practical Appli-
Aeronautical Research Committee Repts. and Memoranda 1127, London, cations,” Journal of the American Helicopter Society, Vol. 33, No. 4, 1988,
March 1927. pp. 64– 68.
3 Glauert, H., “Lift and Torque of an Autogyro on the Ground,” Aeronau- 28 Bramwell, A. R. S., Helicopter Dynamics, Arnold, London, 1976, pp.
tical Research Committee Repts. and Memoranda 1131, London, July 1927. 1 – 13.
4 Lock, C. N. H., and Townend, H. C. H., “Wind Tunnel Experiments on 29 Johnson, W., Helicopter Theory, Princeton Univ. Press, Princeton, NJ,
a Model Autogyro at Small Angles of Incidence,” Aeronautical Research 1980, pp. 237, 238.
Committee Repts. and Memoranda 1154, London, March 1927. 30 Brotherhood, P., “Stability and Control of the Wallis W117 Autogyro,”
5 Glauert, H., and Lock, C. N. H., “A Summary of the Experimental and
Royal Aircraft Establishment, TM Aero. 1461, Dec. 1972.
Theoretical Investigations of the Characteristics of an Autogyro,” Aero- 31 Arnold, U. T. P., “Untersuchungen zur Flugmechanik eines Tragschrau-
nautical Research Committee Repts. and Memoranda 1162, London, April bers,” Technische Univ. Braunschweig Diplomarbeit 88-5D, Brunswick,
1928. Germany, March 1988.
6 Wheatley, J. B., “Wing Pressure Distribution and Rotor-Blade Motion 32 “Helicopter Simulator QualiŽ cation,” U.S. Dept. of Transportation,
of an Autogyro as Determined in Flight,” NACA TR 475, 1933. Federal Aviation Administration Advisory Circular AC 120-63,Washington,
7 Wheatley, J. B., “An Aerodynamic Analysis of the Autogyro Rotor with
Nov. 1994.
a Comparison Between Calculated and Experimental Results,” NACA TR 33 Tischler, M. B., “Digital Control of Highly Augmented Combat Rotor-
487, 1934. craft,” NASA TM-88346, May 1987.

Das könnte Ihnen auch gefallen