• No results found

Rotor prediction with different downwash models

N/A
N/A
Protected

Academic year: 2021

Share "Rotor prediction with different downwash models"

Copied!
23
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

FOURTH EUROPEAN ROTORCRAFT AND POWERED LIFT AIRCRAFT FORUM

Paper No. 6

ROTOR PREDICTION WITH DIFFERENT DOWNWASH MODELS R. STRICKER, W. GRADL Messerschmitt-B6lkow-Blohm GmbH Munich, Germany September 13 - 15, 1978 STRESA ITALY

Associazione Italiana di Aeronautica ed Astronautica Associatione Industrie Aerospaziali

(2)
(3)

SUmmary

ROTOR PREDicriON WITH DIFFERENT

DOWNWASH MJDELS +)

R. Stricker,

w.

Gradl

Messerschmitt-Bolk6w-Blohm GmbH

Munich, Germany

P.O. Box

801140

Rotor induced velocities have to be calculated accurately, in order to predict rotor performance, structural limitations, vibrations, stability and acoustics.

Three typical rotor downwash models in use today are called to mind.

The fast local momentum theory i.s based on blade element - momentum

analy-sis and an empirical relation for forward flight; both developed by Glauert.

Vertex-wake models may be used to obtain better results, when the wake is

near to the rotor. For the vortex models, the blade is represented by a

lifting line and the wake is simulated by meshes of trailing and shed

vortex elements. A typical experimental-prescribed-wake model uses an experimental wake geometry taken from hovering flight measurements, but

the method may also be applied to forward flight if the geometry is

dis-torted according to the free stream velocity. A more adequate

represen-tation of the wake geometry in forward flight might be obtained from

free-wake analysis, where the free-wake elements are allowed to convect in the

velo-city field they create, until they take up positions which are consistent

with the velocity field they induce.

A semi-empirical downwash model which combines momentum theory with properties of a vortex-wake model is introduced. Local momentum theory is extended by

a

simple wake of ring-vortices to simulate wake contraction effects. Wake geometry is taken from experiments also used for prescribed-wake models. Tip losses are simulated by increase of induced velocities

following Prandtl.

For forward flight, the induced velocities at hovering

flight are used as input for a special superpositioning principle derived

from experiment. Therefore, tip vortex - blade interactions and other effects known from vortex models and experiments can also be simulated even for flight cases when the wake is near to the rotor.

Results are presented for rotor blade loads, some specific data of flight dynamics, vibrations and acoustic characteristics. Calculated

blade loads of the fast semi-empirical model are nearly identical to those

from the much more expensive vortex model in hovering flight.

In the

tran-sition flight region, the results are very similar and agreement with meas-urements is good in contrast to results from local momentum theory. As

shown by prediction of cyclic control moment and rotor phase angle, the

local momentum theory gives proper results for rotor calculations at hover

but not for the transition flight region, where the semi-empirical model is

adequate, e.g. in predicting the lateral control angle.

Prediction of rotor

hub inplane oscillatory forces and moments in hover and transition flight

also show the usefulness of the semi-empirical method.

As the oscillatory

stresses predicted by the local momentum theory increase steadily with

increasing advance ratio

~.

the semi-empirical model shows forces/moments

first increasing then decreasing and then increasing again with progressing

~.

whereby agreement with measurements is acceptable.

Finally an example

of rotor noise calculation is given for a descending helicopter, that is

when the wake is near to the rotor.

Calculated blade loads show typical

tip vortex- blade interaction effects at the advancing and retreating blade.

The predicted sound pressure level agrees well with measurements up to a

fre-quency according to the azimuthwise step used in blade loads calculation.

+)

Work sponsored by the Ministry of Defence of the Federal Republic of

(4)

NOTATION

b

number of blades

c

coefficient M

moment

m mass p force R

blade radius

r

blade

radial coordinate

thrust

t

time

v

velocity

x,y,z, rectangular coordinates

y rotor

angle of attack

e

blade angle of attack

~

advance

ratio

p

density

a

solidity

~ rotor

inflow angle

1jl rotor azimuth

angle

Q rotor

angular velocity

INDICES

G

Glauert

i

induced

ib

in board

L

lift

LMl'

local momentum theory

0

overall

T

thrust

tip

blade tip

TL

tip losses

(5)

1. INTRODUCTION

The flow around a helicopter blade is heavily affected by

the rotor downwash, especially in hovering and transition flight. The calculation of the induced velocities with sufficient accuracy is therefore a basic necessity in order to predict rotor perfor-mance, structural limitations, vibrations, stability and acoustic characteristics.

Methods for determining rotor inflow were originally deve-loped for propellers. Simple momentum theory, based on the assump-tion of a uniformly loaded act11ator disc without tip losses, was refined by introduction of blade element - momentum theory (Refer-ences 1 and 2), resulting in a better represent~tion of the radial inflow distribution. Tip losses were calculated (Reference 3) tak-ing into account the finite number of blades. For forward flight the rotor was regarded as an elliptically loaded wing to which lifting-line theory could be applied (Reference 1). The result was an empirical momentum formula and the trapezoidal downwash distri-bution, both of which are still heavily used nowadays. Modified actuator disc theories (References 4 to 6) were developed to pre-dict the rotor inflow variation versus azimuth and radius in forward flight.

Following the early development of the simple actuator disc and blade element - momentum methods, emphasis was placed on vortex theory, where the improvements of the methods largely paralleled the progress in the development of high-speed computers. Besides simple theories, using vortex cylinders and 'rectangularizations' of the vortex wake (e.g. References 7 and 8), the undistorted wake from each blade is modelled by straight trailing and shed vortex elements (e.g. References 9 to 11), The constant vorticity of the elements varies in accordance with the variation of the blade bound circulation in the azimuth and radial directions. The geometry of the wake is defined by the rotational and translational velocities of the rotor in planes parallel to the tip path plane and by the mean flow velocity in the axial direction, leading to the term

'prescribed-classical-wake models'.

The realization that the sensitivity of blade inflow and associated airloading to wake distortions can be significant, led to 'prescribed-contracted-wake models'. The roll-up of the spiral wake was taken into account by truncating the mesh of trailing and shed vortex elements behind the blade, at a given wake azimuth, and by providing thereafter for two concentrated vortex filaments to represent a rolled-up tip vortex and root vortex (References 11 and 12). Wake contraction has been considered analytically by many authors (e.g. References 13 and 14), but initially experimental investigations employing model rotors and flow visualization tech-niques, resulted in a sufficient description of wake characteristics in hovering flight (Reference 15) and in the •prescribed-empirical-wake models' (e.g. References 15 and 16). For the 'free-wake models'

(6)

(e.g. References 17 and 18), the vortex elements are allowed to

convect in the velocity field they create. The iteration either

starts with an initial distribution of the vortex wake, or a process similar to the start-up of a rotor in a free stream is

set up.

The vortex elements move until they take up positions

which are consistent with the velocity field they induce. As

might be expected, the computing time for such calculations is

prodigious.

The primary objectives of this paper are

to compare typical rotor downwash models, such as local

momentum theory, prescribed-contracted-wake analysis, and a free-wake model,

to present a simple semi-empirical model which is based on

local momentum theory but promises results, especially in

transition flight, similar to those calculated with more

expensive vortex models, and

to describe results from these models calculating blade loads

and specific flight dynamic data, rotor vibrations and acoustic

characteristics.

2. TYPICAL

ROTOR DOWNWASH MODELS

IN

USE TODAY

Treating the helicopter rotor as an integrated aerodynamic/

dynamic system, rotor downwash has to be calculated rapidly with

sufficient accuracy.

A

basic theory is the local momentum model

as shown in Figure 1.

In hovering flight it is identical to the

blade element- momentum theory of Glauert (References 1 and 2).

The radial variation of inflow is found by neglecting contraction

of the wake and considering the rotor disc to be divided into

con-centric ring-segments, according to the number of blades. The

thrust produced by each ring-segment is determined by blade element

theory.

By equating it to the overall momentum change in the air

flow through each ring-segment, and assuming that the inflow is

constant over the element, the inflow can then be determined,

For

forward flight conditions the simple linear representation of the

fore-aft variation of inflow over the rotor (provided by Glauert,

Reference 1) is added to the result from blade element- momentum

theory.

In this form, the local momentum theory may be used to

calculate rotor performance and stability as well as the first

harmonics of structural stresses in hovering flight and for advance

ratios greater than about 0,15 with sufficient accuracy at low

computing costs.

To calculate higher harmonics of aerodynamic loads in the

transition flight region as well, vortex-wake models are in use.

As the computing times are nearly equal for rigid-wake and

pre-scribed-wake analyses, models such as the

experimental-prescribed-contracted-wake model shown in Figure 2 are preferred (References

11 and 16), The blade is represented by a lifting line and the

(7)

the near wake, meshes of trailed and shed vortices form the in-board section, whereas a bundle of trailing vortices forms the tip vortex. In the mid wake the shed vortices are omitted and in the far wake there are only two single tip and root vortices. Wake geometry is taken from experimental investigations of Land-2rebe 1Reference 15), Figure 2. The radial and axial coordinates

r . , z . of the tip vortex and the axial coordinates z,b' r=O;

_t>p t>p "

Zib' r=l of the plane of the inboard section are given as func-tions of rotor thrust CT, solidity

a,

blade twist

ev,

number of blades b, and wake azimuth angle

w.

Though this wake geometry was evaluated only for ho'Jering flight, i t can also be used for forward flight, if the geometry is distorted according to the free stream velocity.

A more adequate representation of the wake geometry in forward flight might be obtained by using free-wake analysis such as the free-wake model of Sadler, see Figure 3 and Reference 18. The blade is again represented by a lifting line and the wake is calculated by a process similar to the start-up of a rotor in a free stream. An array of discrete trailing and shed vortices is generated with vortex strengths corresponding to stepwise radial and azimuth blade circulations. The array of shed and trailing vortices is limited to the near •,.,rake whereas an arbitrary number of trailing vortices form the far wake. The end points of the vortex elements are allowed to be transported by the resultant velocity of the free stream and vortex-induced velocities. Cal-culation is terminated when the vortices trailing from the blade during the first azimuthwise step no longer influence points of interest near the rotor disc. The results are wake geometry and wake flow data in addition to wake influence coefficients that may be used to calculate blade loads and blade response.

From the theoretical point of view, the three downwash models described above should satisfy the rotor designer's needs, but in standard rotor calculation of rotor performance, structural limitations, vibrations, stability, acoustics etc., the fast local momentum theory often gives poor results in the higher harmonics. Also computing time for vortex models, not to mention that for free-wake analysis, is excessive. Therefore, a semi-empirical downwash model, based on blade element-momentum theory and on an

idea from Wood and Hermes (Reference 6) for induced flow build-up in forward flight, will be described below. The model might be an efficient tool for standard rotor calculations, especially in the transition flight region, both in terms of computing time and accuracy.

(8)

3, A SEMI-EMPIRICAL DOWNWASH MODEL

As shown in Reference 19 and Figure 4, the build-up of the

induced velocities of a hovering rotor, following a rapid

collec-tive pitch increase of 200 deg./sec., will take about 1.2 sec.

The curve of the induced velocities versus time may be

approxi-mated by an empirical relationship.

From blade element - momentum

theory, e.g. the induced velocity V. (

)

versus blade radius r

~ r,to:o

at a

timet=~,,

is known.

So the local induced velocity Vi(r,t)

for t ~ o:o can be read from the start-up approximation. The local

downwash may then be calculated by summing up the portions

~v

1

,

~v

2

,....

Of course, the result in hovering flight is identical

to the value from blade element - momentum theory.

Transferring these relationships to forward flight, the

build-up of the local induced velocities can be seen from Figure 5

for a blade at an azimuth angle of 180 deg., i.e. in flight

direc-tion for simplicity.

In region A there is no induced velocity,

Region B is affected by the induced velocity

~v

1

that was built

up during the time interval

~t

=

2•n/~•b

by the preceding blade.

For region C, the induced velocities

~v

1

from the preceding blade

and

~v

2

from the pre-preceding blade have to be added.

The induced

velocity at region D however, is calculated by summing up the

por-tions

~v

1

, ~v

2

and

~v

3

,

all calculated in accordance with their

time intervals and the start-up approximation shown in Figure 4.

For a given blade element P

(r,~)

at any radial station r

and azimuth station

~

we have to sum up the induced velocity

por-tions

~vi,

from all blades that have passed this point before.

The azimuth angle Wi+l of the actual preceding blade may be

calcu-lated from an implicit equation. The time interval

~ti+

1

depends

on rotor angular velocity

~~

blade number

angles Wi and

~i+

1

as shown in Figure 5,

b and blade azimuth

~Vi

has to be calculated

from the start-up approximation using

~ti+

1

and the local induced

velocity vi(

,,,

) for t =

~

ri+1' "'i+1

from blade element - momentum

analysis as input. Summation of

~Vi,~Vi+

1

.•• is terminated when

the terms of the sum become small, either because the start-up

approximation reaches the final value, or ri+

1

becomes greater than

the rotor radius, at which station Vi from blade element - momentum

theory vanishes.

Local momentum theory, see Figure 1, is used as the basic

inflow model to calculate the Vi at t =

~.

However, in the momentum

equation of Glauert, the inplane component of the free stream

velo-city is omitted in order to calculate quasi-hovering cases in

(9)

fnrwA.rrl fl iCJht, sP.e Figure 6. This is clone because the empirical momentum 0'1uation of Glac.ert for forward flight (Reference 1) is reploced by the 0mpirical principle of superpositioning, shown in F:Lgures 4 anrl ).

WAke cnntr~ctinn is simulated by a wake of up to 4 ring vort l.cr.::os, see F.i qure f), r,..•here the wake geometry is taken from the

experimenta1-pre.sc:ri.beri-c-ontracted wake model of Landgrebe, see Reference 1:::; 'mrl. F'iqure 2. The vorticity of the wAke is cA.lcu-lated in accr;rrln.nce with the maximum blade circulation.

Tip lr,ss0s c·ne simulated by an increase of induced velocity ViTL tn proclucc: ,, qi.ven thrust, see Figure 6. Following Prandtl

(see RP.fP.rPnce :~0), V. may be calculated from the number of 1.TL

blades b, r~rlial station x, and inflow angle

¢.

Then the induced velocity portions, from modified blade element - momentum theory, from wake cs·ntrn.cti 0n simulation and from tip loss calculation, when added, wi ll a i t·e thE local induced velocity.

A block diaqram for the rotor calculation using the semi-empirical downwash model is given in Figure 7. Input data are flight conditions and rotor data. Under the rotor trim loop and blade dynamics iteration loop, the rotor calculation is performed. Computation (._)f induccr:l VP locity for a blade element is done under

consideration 0f. local mcmenturn, forward speed superposition, and the influence ~,f ,.,n_ke c;-;r traction and tip losses. The blade radial loop is folL>I·.'<?lJ t--·y n::::aJ culation of the wake vorticity at the actual azimuth angle, .,.nd the r< tor azimuth loop terminates the rotor cal-culation. Results are tr.e rotor forces and moments, blade motion and loads, as ~-·P ll as r-otor inflow and wake data.

4. RESULTS AND DISCUSSION

Comparison of computing times for the downwash models pre-sented, is shown in Figure 8. Local momentum theory provides results in less than 0.2 minutes for a rotor trim case shown in Figure 7. A single blade dynamics loop for a rigid wake model, taking into account a wake of 3 rotor revolutions, takes about 1 minute of computing time. Free-wake analysis using only 5 trailing vortices and calculating a single blade dynamics loop uses about 2 to 20 minutes of computer time for~ = 0.3 and 0.1 respectively. The semi-empirical model provides results for a rotor trim loop in about 1 to 0.2 minutes between hover and~= 0.3. However, computing time for hover is given for a disturbed flight and the value for an ideal hovering flight case is equal to the time for u = 0.3. Therefore in the low speed region, computing time for a rotor trim loop using the semi-empirical model is nearly equal to the time needed by the vortex model for a single blade dynamics loop. For high speed flight, computing times for the semi-empirical model and the local momentum theory are nearly equal. Further reductions of calculation costs for the semi-empirical model may be attainable by refinement of the numerical calculation procedure.

(10)

Calculated blade loads versus radius and azimuth-angle for

hovering and transition flight of the S 58, are shown in Figures 9 to 11 and compared with measurements (Reference 21). Results are presented versus radius for azimuth angles

W

=

o,

90, 180 and 270 deg. and versus azimuth angle for radial stations r

=

0.95 and 0.85. For the hovering case, see Figure 9, the local momentum model cannot reproduce the important wake-contraction effects near the blade tips, whereas the results from the semi-empirical model and from the prescribed-contracted-wake analysis are almost identical and agreement with measurements is very good considering

the difficulty in finding undisturbed hovering flights.

For the transitional advance ratio of ~

=

0.064, blade loads are shown in Figure 10. At the advancing and retreating blade, the tip loading is increased by the tip vortices, whereas in the fore-aft position, blade loading is rather triangular. Plots of load versus azimuth also show typical tip vortex - blade interaction effects for the semi-empirical model and the prescribed-contracted-wake analysis, whereas the local momentum theory provides only average values. In spite of the fact that the semi-empirical model may overestimate the tip vortex effects for some azimuth angles, agreement with measurements is as good as for the prescribed-contracted-wake analysis.

Blade loads for the moderate advance ratio~= 0.112 are

presented in Figure 11.

Again the typical blade tip loading effects

owing to the tip vortices at the advancing and retreating blade can be noticed. Results of the semi-empirical model in general are

simi-lar to those of prescribed-contracted-wake analysis.

The yields of

20 minutes of computing time invested in a free-wake model are also

shown in Figure 11,

In spite of the sophisticated model, the results

are rather poor compared to measurements and results of the more

simple methods.

This may be due to the rough discretization of the

wake, using only 5 trailing vortices and 12 steps around the azimuth.

Trends of calculated and measured blade load harmonics for

the S 58 at advance ratios

of~ =

0.064, 0.112 and 0.229 are

presen-ted in Figure 12.

For low and moderate

~.

the higher harmonics are

underestimated by the local momentum theory and to some extent also

by the rigid-wake analysis.

For a typical forward flight condition

(~

=

0.229) the local momentum theory as well as the semi-empirical

model and the rigid-wake analysis nearly give the same results

show-ing an extremum for the second harmonic. The increasing and

decreas-ing of the higher harmonics with increasdecreas-ing

~

in the transition flight

region can only be read from the results of the semi-empirical model.

Some typical results from rigid rotor flight mechanics are

given in comparison to measurement in Figures 13 and 14.

As shown

in Figure 13 (Reference 22) the cyclic control moment and the phase

angle versus rotor thrust in hovering flight can only

be

predicted

using the local momentum theory (i.e. the basis for the semi-empirical

model) instead of the simple momentum equation. The local momentum

theory shows the typical decrease of cyclic pitch effectiveness and

the increase of longitudinal/lateral coupling with decreasing rotor

(11)

thrust in hovering flight. Disregarding the longitudinal control angle, which in the first place is affected by horizontal tail effectiveness, the lateral control angle in transition flight is shown for the BO 105 in Figure 14. Typical lateral control angle versus forward flight speed is predicted by the semi-empirical model much better than by the local momentum theory.

Results from rotor vibration calculations in the transition flight region are presented in Figures 15 and 16. Rotor hub in-plane oscillatory forces for the BO 105 at advance ratios of ~ = o to ~ = 0.268 are shown in Figure 15. As the forces predicted by the local momentum theory increa£e steadily with increasing ~~ the semi-empirical model shows oscillat0ry forces first increasing and then decreasing with progressing ~ in the transition flight region. For hover and for forward flight at ~

=

0.268, the "esults from both theories are nearly identical. Comparison with measurement (Refer-ence 25) is made for the rotorhub inplane oscillatory moments of the BO 105 in transition flight in Figure 16. The semi-empirical model again shows the increase and decrease of the moments versus ~.

As the local momentum theory underestimates the moments for small advance ratios, the agreement of the results predicted by the semi-empirical model with measurements is ac.ceptable with regard to the amplitude and phase of the moments.

An

example of rotor noise calculation is given in Figure 17 for the BO 105 during partial power descent (Reference 23). The plots of the calculated blade loads for several radial stations versus azimuth-angle show typical tip vortex - blade interaction effects at the advancing and retreating blade. Using the calcu-lated aerodynamic blade loads as input data, the sound pressure can be estimated (Reference 24). The plots of calculated and measured sound pressure versus time show acceptable agreement, neglecting the extreme pressure peaks in the measurement which result from main rotor - tail rotor interference that is not in-cluded in the analysis. Predicted and measured sound pressure levels versus frequency, also coincide very well up to a frequency of about 375 Hz, which corresponds approximately to the 12th blade harmonic. Higher harmonics may be predicted correctly using smaller azimuthwise steps in the blade load calculations.

5. CONCLUSIONS

Treating the rotor as an integrated aerodynamic/dynamic system in standard rotor calculations, the induced velocities in the rotor disc have to be calculated accurately, especially when they are not small compared with the free stream velocity. Typical flight cases are hover and transition flight as well as flare and descent, when the wake is near to the rotor.

(12)

Rotor downwash analysis in standard calculations today is

based either on momentum theory or on vortex-wake models.

Typical

momentum models, such as the local momentum analysis, can predict

induced velocities accurately and reasonably in undisturbed

hover-ing flight (neglecthover-ing wake contraction) and in all cases when the

wake is far from the rotor, i.e.

~ >

0.15 at nose-down angles of

the rotor disc.

For all other cases when the wake is near to the

rotor, vortex models, such as an empirical-prescribed-wake model

or free-wake analysis, have to be consulted.

Typical

prescribed-wake models, using an experimental. prescribed-wake geometry taken from

hover-ing flight, may also be applied to the transition flight region if

the geometry is distorted according to the free stream velocity and

computing time is tolerable for use in special cases.

However,

computing time for free-wake models, in transition flight cases, is

too formidable to use these models in standard rotor calculations.

A semi-empirical downwash model which combines momentum

theory with the properties of a vortex-wake model, may be an

effi-cient tool for standard rotor calculations both in terms of computing

time and accuracy.

Local momentum theory can be extended to simulate

wake contraction and tip losses in order to give nearly the same

results in hovering flight as an experimental-prescribed-wake model.

A special superpositioning principle can also simulate tip vortex

-blade interactions and other effects known from vortex models and

experiments in the transition flight region.

Therefore, results of

the fast semi-empirical downwash model are very similar to those

calculated by much more expensive vortex-wake analysis.

The usefulness of the semi-empirical downwash model in

stan-dard rotor calculations, in order to predict rotor performance,

structural limitations, vibrations and acousticsF can be

demon-strated by predicting, e.g. blade loads, rotor hub inplane

oscil-latory stresses and rotor noise, with acceptable accuracy for flight

cases where results from local momentum theory are inadequate.

6 .

REFERENCES

1)

H. Glauert, A General Theory of the Autogyro, Aeronautical

Research Council (Great Britain), R&M No, 1111, (1926)

2)

H. Glauert, on the vertical Ascent of a Helicopter,

Aero-nautical Research Council (Great Britain), R&M No. 1132,

(1927)

3)

s. Goldstein, On the Vortex Theory of Screw Propellers,

Royal Society Proceedings, Ser. A 123, (1929)

4)

K.W. Mangler and H.B. Squire, The Induced Velocity Field

of a Rotor, Aeronautical Research Council (Great Britain),

R&M No. 2642, (1953)

(13)

5) R.A. Ormiston, An Actuator Disc Theory for Rotor Wake Induced Velocities, ~-CPP-111, (1972)

6) E.R. Wood and M.E. Hermes, Rotor Induced Velocities in Forward Flight by Momentum Theory, AIAA/AHS VTOL Research, Design, and Operations Meeting1 AIAA Paper No. 69-244,

( 1969)

7) H.H. Heyson and S. Katzoff, Induced Velocities Near a Lifting Rotor with Nonuniform Disc Loading, NACA TR 1319,

(1957)

8) M.A.P. Willmer, The Loading of Helicopter Rotor Blades in Forward Flight, Aeronautical Research Council (Great Britain), R&M No, 3318, (1963)

9) R.A. Piziali and F.A. DuWaldt, Computation of Rotary Wing Harmonic Airloads and Comparison With Experimental Results, Proceedings 18th Annual Forum, American Helicopter Society,

( 1962)

10) M.P. Scully, Approximate Solutions for Computing Helicopter Harmonic Airloads, Massachusetts Institute of Technology, TR 123-2, (1965)

11) G. Daske, I.A. Simons, Rotorberechnung mit BerUcksichtigung der ungleichf6rmigen Verteilung der induzierten Geschwindig-keiten, MBB-Report OF 70, (1967)

12) R.A. Piziali, Method for the Solution of the Aeroelastic Response Problem for Rotating Wings, Jl. Sound and Vibration, Vol. 4 , No. 3 , ( 1966)

13) T.T. Theodorsen, Theory of Static Propellers and Helicopter Rotors, 25th Annual Forum, American Helicopter Society, No. 326, (1969)

14) M. Jogelar and R. Loewy, An Actuator-Disc Analysis of Helicopter Wake Geometry and the Corresponding Blade Response, USAAVLABS Technical Report 69-66, (1970)

15) A.J. Landgrebe, An Analytical and Experimental Investigation of Helicopter Rotor Hover Performance and Wake Geometry Characteristics, USAAMRDL Technical Report 71-24, Eustis Directorate, u.s. Army Air Mobility Research and Development

16) R. Stricker, W. Gradl, G. Polz, Aerodynamische Arbeitsgrund-lagen fUr zukUnftige Hubschrauberentwicklungen, MBB-Report UD-159-75, (1975)

17) A.J. Landgrebe, An Analytical Method for Predicting Rotor Wake Geometry, Jl. American Helicopter Society, Vol. 14, No. 4, (1969)

(14)

18) S.G. Sadler, Development and Application of a method for predicting rotor free wake positions and resulting rotor blade air loads, NASA CR-1911, (1971)

19) P.J. Carpenter, B. Fridovich, Effect of a Rapid Blade-Pitch Increase on the Thrust and Induced Velocity Response of a Full-Scale Helicopter Rotor, NACA TN 3044, (1953)

20) A.R.S. Bramwell, Helicopter Dynamics, Edward Arnold, London, (1976)

21)

J.

Scheiman, A Tabulation of Helicopter Rotor-Blade

Differential Pressures, Stresses, and Motions as Measured in Flight, ~ TM X-952, (1964)

22) H. Huber, V. Langenbucher, H.J. Dahl, E. Laudien, Theore-tische Grundlagen und Untersuchungen zur Flugmechanik von Drehfluglern, ~-Report UD-131-74, (1974)

23) E. Laudien, Untersuchungen zur Ldrmminderung an Hubschraubern, MBB-Report UD-197-76, (1976)

24) K. Heinig, Untersuchungen zur Ldrmminderung bei Hubschraubern, MBB-Report TNA 012-28/70, (1970)

25) G. Stiller, Analyse der BO 105 - Getriebevibrationen im Hinblick auf SchwingungsisolationsmaBnahmen, ~-Report

(15)

FORWARD fl !GUT (GI A!:ERD

v

1G •

v

10 · ( 1 • K · x · cos !J,tl

l OtA! !IDHf]Il!M THEORY

vil..Mf viG ~ vio • vi

[IInnrr11Til]

0

r~T

[g:-m_

.([1

t

I D

--Figure 1: Principle of local momentum theory

ROOT VORTEX TIP VORTEX NEAR WAKE MID WAKE FAR WAKE 0 LO

i'-,

f---+---L

I

. '<

~

l

~B.R

• 0 ' I"-. ZTIP~- . . i .."_~, R • l 1 I

---T

Z, ~ "F 1 1

cr<· ev . .;,

b l

-i

I ~T':~ I

}

/ " I

!

I 0 180 l60

WAY£ AZIMUTH AIIGLE • DEG Figure 2: Wake structure and geometry of Landgrebe for

the prescribed-contracted-wake model 'UNDUFL'

NEAR WAI(£

Figure 3:

Wake structure and program flow diagram

for the free - wake model of Sadler

(16)

1 ~ ~

g

~ ~ Q

~

:!!: u

Yt Fi!J/1 BLADE EI'£1£NT • llll£llllll THEORY

0

o.s

BLADE RAD I US

Figure 4:

Hovering flight

lllADE AT <jl • 180 DEG

D C B A

0 1.0

lUilE IIAD I US

Figure 5:

Forward flight

BASIC !NFI OK MQI'fl

LOCAL tiii'I:NTll1 THEORY

'

N .?;

>

"'

~ ~ ~ ~

"'

Am!JlX!MATJO!f• ·5!2 ·4 t Vj/VjQ)•1·.0,Sit • t J 1.2 ~ 0.8

v

1/

~.M

~

--- 1'£ASURH£NT-\ APPROXIMTIO!I

oV

~ M,

1-0 O,q

o.s

1.2 Til'£ • SEC FLIGHT DIRECTION . . . .

• IHPLAHE FREE STREAII VELOCITY CIJIPW:HT HEGL£C1!11

WW' CfWT8ACT! t:l

SI~LAT£0 BY RIIIG VORTICES • WAKE GEO/f:TRY OF LAIIDGR£BE

• VORTICITY FI!JIIIIAXI~~ BLADE CIRCULATION liP IQSSES IPRMUTLl

SI~LATI:D BY INCREASE OF

INDUCED IUOCITY ViTL

8£.SUll vi • Vilxr •

vi we • vi

TL

~

riTL

~j

~

0 1 JUIE • • •

(17)

FLIGHT CONDIT IONS BLADE ELEi1ENT THEORY:

ROTOR DATA CALCULATION OF INDUCED VELOCITY

l

UNDER CONSIDERATION OF

r l

TRIM PROCEDURE - LOCAL MOI·IENTUM

~

- FORWARD SPEED SUPERPOSITION

rl

BLADE DYNAMICS

-

VORTEX 1/AKE

-

TIP LOSSES

I

BLADE DYNA1~ I C LOOP BLADE RADIAL LOOP

I

WAKE VORTI C 1TY

I

ROTOR INTEGRATION

I

I

L-..j TRIMM LOOP ROTOR AZIMUTH-WISE

LOOP

I

ROTOR FORCES Ai/0 r10MENTS, BLIJJE ~.OTION AND LOADS. FLOW DATA. WAKE DATA

Figure 7: Rotor calculation block diagram using the semi - empirical downwash model

'-

..

10

'-...

z

;::

FREE - HAKE ANALYSIS UBDU '•

~

I

"'-.,_ ~

"'

~ ;:: l

i

IGID-HAKE ANALYSIS (JBDU . ·,

-

-

--

---- ---

-

---- oM I' I R I CAL.-::-:::-:-" ' I MODEL (JRTLJ

=-..lli

L MOMENTUM .A.!ffi Y..ill,fTQ_

-~-O.l

0

0.1

0.2

ADVANCE RATIO 14

Figure 8: Computing time vs. advance ratio for

selected downwash models

(18)

f-00 600 q {da1Ntml

1~

0

on

I

"

~

400 ~p ~

v

~

200

/

ljJ • \80

~

.}¥"'"

ljJ = 0 l

0

1 ljJ = 270 1 0 1 q {daN/ml 0 0 0 0 0 0 0 0 ~ y

"

0

-

---

---

--o

0 0

400

o o MEASUREMENT. SCHE!MAII - - - PRESCRIBED-CONTRACTED- X • 0,95

200

-LOCAL MOMEMTIJII WAKE ANALYSIS

rLYS!S

1

T

SEN!

I-

ENPIRTL MODEL

0 9() 180

270

360 600 0 0 4(.() 0 0 )( • 0,85

---

- r

---

--

---

- - - "'0""'1 ....a.-

---0 0 0 200 0 90 180

270

360 <I> {dog!

Figure 9:

Blade loads vs. radius and azimuth-angle

for the S58, m

=

5240 kg,

~

=

o

(19)

600

T

0 '"'0 0 q{daN/ml

h

1/;~

0

qoo

k':f;;

r~

t:--..

/

I

K',

,:.---I

--...

200

'

~

~"'

~

v

<I! " 180 <1!"0 ' 1 <I! "270 1 200 I 0 600 --LOCAL MOr£HTUI1 ANPLYSI S 90 4(11)

1--...

_...---..._

o'\,

L

/

---::-=-t-./o 0 200 0 0 PRESCRIBED-CONTRACTEDI--'-..t-+----! WAKE ANALYSIS

SEMI - EMPIRICAL MODEL

180 270 360 ),-...Q., 0 0 'l.. J1

,_

#

~-~~

0

--

--~'ol

--o-, "-/ o o ~lfASUREfiEN(' SCHE lfltN 1 1 X • 0,95 X • 0,85 0 90 180 270 360 'l> ldtgl

Figure 10: Blade loads vs. radius and azimuth-angle for the S58, m

=

5140 kg, ~

=

0.064

(20)

1 --~-.---,---r600 --LOCAL MOMENTUM 0 I ~ q!daN/m I 1 MAlYSIS

1 ' - - -PRESCR I BEU- q 00

-\--_:;=:-"----"-:;...::c=.:...;:="--+-"i-P\1

/ ''-., CGIHRACTED-WAKE ANALYSIS 1 I ' " ·

I

i - - ""'

~.

200 I ' , <jJ ' 270 0 0 ---- PRESCRIDED-COIITRACTED-WAKE f,NAL YS IS 0

f- ...

_ / / '

'·"

I ' \ /

l\.

/

\

'

bZ.

~

/

--¥

--0

\"

0 0 ----LOCAL MOMENTUM ANALYSIS 180 2l0 I/~'\

/

-- ~-' -~~~

/

·~-

-00

'--··

-SEMI - EMPirAL '·'\, - - FRE 1

E-WAKE ANrLYSIS MO?EL

u

90 180

2io

Figure 11: Blade loads vs, radius and azimuth-angle for the S58, m ~ 5180 kg, ~ ~ 0.112 X • 0,95 360 0~ 0 /

V/

X • 0,85 . /

/

3c

0 <jJ ld•gl

(21)

--~CALl MO~EN

1

ru~

--- RIGID-WAKE ANAL

!.'

YSI

o o MEASUREMENT

-

JEMI

~

JPIJR/cl f--rDEL -ANALYSIS

!!;!

0,1 5

0.01

I / /

\

\

1

2

Figure 12:

0.03

al"

SCHEIMAN ~ • o.oGq 0.1

/

?\

'

0.1

~~

'

'

-

I -\ I I I

'

•\

I

\

I

\

'

.

I

1\

I

\ -

\

'

1

0,01

o.o

'

.

l.l.

0.112

f-l.l•

0.229

q 6 8 1 1 NUMBER OF HARMONICS

Trends of blade load harmonics for the 858

at u

=

0.064, 0.112, and 0.229 0

--/

'

60

~

'

--'-I

'

\

l\

'

'

I

'

0

-$!?. ·~0 0.02 !2

-

0 0 o, ~!OMENTUM ANALYSIS-' 40

i

9{

I to ---- LOCAL Mllf.IEIITUM E u

;;; 0.01

-0

2

0.01 0.02

B01r5•1.l"f 6n w ~20 il:

4

6

8

10

0

2

0.05

0,07

O.Ol 0.02

COLLECTIVE PITCH - DEG THRUST COEFFECIENT

0

Figure 13:

Cyclic control moment and phase angle

for 80 lOS at hover

ANALYSIS o MEASUREI'ENT

I

4 6 8

0.05

0.07

6 10

(22)

- I

LOCAL

rui:HTt~~IIIALYdis

- - S(l11 • EMPIRICAl/101{L 0 0 I'EASUREIEHT 0 BO lOS

I

0 0 m • 1720 kg 0

(/

' - - -

...

~

0

/

10 20 30

fLIGiiT SPEED • MIS

Figure 14:

Lateral control angle in transition flight

for the BO lOS

'lc

!daN) - - - LOCAL IU"''HTIII 5 AltALYSIS

+

- - SEI\I·EI\PIRICAL P ki:!N) MODEL 5 ~·0 ~· O.Otil

++

~ • O,J3q ~. 0.201 ~. 0.268

Figure lS:

Rotorhub inplane oscillatory forces vs.

~

for

the BO lOS, m

=

1720 kg

- - LOCAL IOOTIII A/tAL !SIS - SE/11 • EJ1PIRICAl ~l{L • o I£ASUR!:IENT 10

"'r

lmdaNl ~. 0.146

(23)

E 0

z

0 0 ..--< "0

j_

C/)

1

Cl «: 0 ---' w Cl «: ---'

""

"'

Cl r w

"'

-

1--

:5

=> L> ---' «: u (J w

MEASUREMENT

""

=>

0.94

C/) C/) w

""

"-Cl

ANALYSIS

z:

0.87

=> 0 C/)

0.81

z:

0.74

0 1--«:

MEASUREflENT

1--C/)

0.68

---'

80

«: Cl

;:2

""

0.61

-u

60 .

---'

ANALYSIS

w

40

>

·-so

105-0.55

w ---'

Y

~

6 deg

"-

Vx = 60kts_Vz = 3.5m/s

20

C/)

0.48

m = 2300 kg DIST. =lOOm

18U

360

250

soo

AZIMUTH - ANGLE - DEG

FREQENCY - Hz

Figure 17: Rotor noise of BO 105 during partial power descent

Referenties

GERELATEERDE DOCUMENTEN

Patients with an isolated type II endo- leak did show a higher incidence of aneurysm growth during longer term follow-up, but rupture may have occurred only once in the isolated type

Voor de scriptie naar de tuniek APM 16.388, een rode wollen tuniek uit het depot van het Allard Pierson Museum, zullen verschillende onderzoeken uitgevoerd worden om de conditie

Voorafgaand aan het onderzoek werd verwacht dat consumenten die blootgesteld werden aan een ‘slanke’ verpakkingsvorm een positievere product attitude, hogere koopintentie en een

Since we concluded that the main effect of service type on perceived quality is significant and since there is a difference in level of personalization between hairstylist and

with the power spectrum of the input signal. From simulations and anaiytical results, it appears that the influence of the smoothing constant fl on the conver- gence behavior

Adaptatie biedt kansen aan nieuwe sectoren Verruiming van het projectgebied en het betrekken van diverse, voor het gebied vaak nieuwe sectoren bij gebiedsontwikkeling bieden

We did expect that in these multi-generational households, caregivers would experience relatively less burden on account of sharing of roles between family members; however, since

Dat de gezaghebbende ouder(s) toestemming moet geven voor een jeugdige onder de 12 jaar of voor een jeugdige ouder dan 12 die niet instaat is tot een redelijke waardering van zijn