17.01.2014 Views

Reduction of Noise Emission of Suboptimal Operating ... - IAG

Reduction of Noise Emission of Suboptimal Operating ... - IAG

Reduction of Noise Emission of Suboptimal Operating ... - IAG

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>Reduction</strong> <strong>of</strong> <strong>Noise</strong> <strong>Emission</strong> <strong>of</strong> <strong>Suboptimal</strong> <strong>Operating</strong> Propellers<br />

A. Hirner, Th. Lutz and E. Krämer<br />

University <strong>of</strong> Stuttgart<br />

Stuttgart, D-70569<br />

Germany<br />

Abstract<br />

Propellers designed for cruising flight <strong>of</strong>ten suffer from an increased noise radiation if they are<br />

operating under <strong>of</strong>f-design conditions such as take-<strong>of</strong>f and approach as well as reverse thrust<br />

where partially high flow velocities and separations on the propeller blades can occur. This is<br />

especially true for the airship Zeppelin NT where the propellers have to cover a very large pitch<br />

angle range when used for manoeuvering on the ground. In cooperation with the project partners<br />

Zeppelin Luftschifftechnik GmbH and Steinbeis Transferzentrum a combined theoretical and experimental<br />

design approach is established which comprises modified vortex lattice methods for<br />

preliminary design as well as CFD calculations for the analysis <strong>of</strong> the aerodynamic and aeroacoustic<br />

behaviour <strong>of</strong> the propeller designs. To validate the theoretical results additional aerodynamic<br />

and aeroacoustic tests with a propeller test bench are conducted. Besides load measurements the<br />

detection <strong>of</strong> characteristic noise sources on a propeller blade is done by scanning the propeller<br />

blade with an acoustic mirror. The project aims at the development <strong>of</strong> a time-efficient process<br />

chain to reduce turn-around times in the design phase <strong>of</strong> new propellers. This comprises preliminary<br />

design codes, subsequent highly automated propeller CAD design and grid generation for<br />

the CFD calculations as well as structural analysis and milling <strong>of</strong> selected propeller designs.<br />

Nomenclature<br />

λ [mm] Wave length<br />

f [Hz] Frequency<br />

w 0<br />

[mm] Half-width <strong>of</strong> the diffraction<br />

pattern, u ∞ = 0<br />

D [m] Diameter<br />

G [dB] Effective gain factor<br />

I M [ W m<br />

] Sound intensity in the maximum<br />

2<br />

<strong>of</strong> the diffraction pattern<br />

I Ref [ W m<br />

] Sound intensity measured by<br />

2<br />

reference microphone<br />

L M [dB] Sound pressure level in the<br />

maximum <strong>of</strong> the diffraction pattern<br />

L Ref [dB] Sound pressure level measured by<br />

reference microphone<br />

1 Introduction<br />

The appearance <strong>of</strong> the aircraft turb<strong>of</strong>an came along<br />

with a decreasing interest in developing new propeller<br />

driven aircraft, especially for long-range routes and<br />

the application <strong>of</strong> the propeller was mainly restricted<br />

to small airplanes. With energy ressources becoming<br />

shorter many leading aerospace companies are<br />

currently investigating the potential <strong>of</strong> highly-loaded,<br />

many bladed propellers operating at more and more<br />

increasing flight mach numbers. Compared to turb<strong>of</strong>an<br />

designs significant fuel savings are possible at the<br />

expense <strong>of</strong> a higher noise radiation <strong>of</strong> propeller driven<br />

aircraft. This effect is undesirable both in the cabin<br />

in regard to passenger comfort as well as in the surrounding<br />

area <strong>of</strong> airports. Current propeller research<br />

is focused on understanding the noise generating mechanisms<br />

and developing highly efficient, noise reduced<br />

propeller designs. For modern turbo-prop aircrafts these<br />

efforts usually result in many bladed propellers with<br />

more or less swept blades. To gather an overview on<br />

the basic design strategies references [1] and [2] may<br />

be consulted.<br />

The operation <strong>of</strong> the propeller in reverse mode is usually<br />

restricted to a short-time period and therefore does<br />

not play a major role in the design criteria. In contrast<br />

the current propellers at the Zeppelin NT are operating<br />

under partly extremly unfavourable conditions<br />

with an extensive noise radiation when the airship is in<br />

approach or manoeuvering on the ground. Within the<br />

current project a highly automated process chain for


time-efficient design <strong>of</strong> noise reduced propellers is developed<br />

which aims at the design <strong>of</strong> a new stern thruster<br />

for the Zeppelin NT under particular consideration<br />

<strong>of</strong> <strong>of</strong>f-design conditions. To get new insights in the<br />

noise generating mechanisms <strong>of</strong> the current 3-bladed<br />

propeller both numerical and experimental studies are<br />

conducted with the full-sale and a model propeller. In<br />

addition a second propeller with a flexible mounting <strong>of</strong><br />

the blades at the hub is examined. This configuration<br />

is said to have noise reducing effects as the blades can<br />

oscillate to a certain amount if the blade encounters<br />

unsteady inflow caused by atmospheric turbulence or<br />

inclined inflow. This effect will be studied by operating<br />

the propeller in inclined flight condition both with flexible<br />

and fixed mounting <strong>of</strong> the blades.<br />

2 Process chain for highly<br />

automated propeller design<br />

As for commercial aircraft there is a large need in increasing<br />

efficiency as well as reducing noise emission<br />

<strong>of</strong> propeller driven general aviation aircraft. This effort<br />

arises due to a lower noise acceptance <strong>of</strong> the population<br />

resulting in stricter noise regulations <strong>of</strong> local<br />

airfields. The current state <strong>of</strong> the art propeller design<br />

<strong>of</strong> small propeller manufacturers is usually limited to<br />

preliminary design methods and broadly based on experience.<br />

This approach cannot fully account for the<br />

complex three dimensional flow field about a propeller.<br />

Therefore within the current project detailed numerical<br />

studies combined with experimental work are<br />

conducted to develop new design criteria for noise reduced,<br />

efficient propellers with special attention to <strong>of</strong>fdesign<br />

conditions. To ensure short turn-around times<br />

<strong>of</strong> a propeller design there is a need for a highly automated<br />

process chain comprising both numerical as<br />

well as experimental studies. In the future parts <strong>of</strong> this<br />

process chain can be supplied to local propeller manufacturers<br />

as service work. The components <strong>of</strong> this<br />

procedure according to Figure 1 will be presented in<br />

the following.<br />

Preliminary Design<br />

XROTOR<br />

CAD, Structural Analysis<br />

CATIA<br />

Experiment<br />

Thrust, Torque, Visualisation,<br />

Acoustics<br />

New <strong>Noise</strong><br />

Reduced Design<br />

Evaluation <strong>of</strong> Design<br />

Numerical Flow Simulations<br />

FLOWer / TAU<br />

2.1 Preliminary design<br />

For preliminary design studies the propeller code<br />

XROTOR from Drela [3] is used. This code allows<br />

the interactive design and analysis <strong>of</strong> free-tip as well<br />

as ducted propellers. The propeller blade is represented<br />

by the classical lifting-line theory whereas the aerodynamic<br />

characteristics <strong>of</strong> the blade sections can<br />

be modelled by specifying arbitrary airfoil aerodynamic<br />

properties. To account for compressibility effects<br />

a Prandtl-Glauert correction is applied. The induced<br />

velocities are calculated by numerically solving for the<br />

exact potential flowfield about the helical vortex field.<br />

This approach is therefore valid for all advance ratios<br />

and blade numbers compared to classical theories<br />

which assume a small advance ratio and many blades.<br />

A first-order correction for the change in the pitch <strong>of</strong><br />

the trailing sheets is applied to account for the propeller’s<br />

self-induction which is important for highlyloaded<br />

propellers. To examine more general propeller<br />

geometries with non-radial lifting lines (raked or swept<br />

blades) an additional discrete vortex wake formulation<br />

is implemented to calculate the induced velocities.<br />

Discrete line vortices on the rigid helicoidal wake<br />

surface trailing from the lifting line into the farfield<br />

downstream are used to calculate the velocitiy on the<br />

propeller lifting line. <strong>Noise</strong> level contour plots can be<br />

calculated on a rectangular grid based on a method<br />

developped by Succi [4],[5].<br />

This code is used for time-efficient investigation <strong>of</strong> the<br />

aerodynamic and aeroacoustic behaviour <strong>of</strong> different<br />

blade geometries both under design and <strong>of</strong>f-design conditions.<br />

This approach delivers important information<br />

concerning necessary lift coefficients <strong>of</strong> the airfoils<br />

as well as Reynolds- and Mach number distribution<br />

along the blade at an early stage in the design process.<br />

The rotation <strong>of</strong> the propeller blades and the resulting<br />

centrifugal and Coriolis effects strongly influence the<br />

boundary layer development on the blade especially<br />

near the hub. Due to these effects unstalled flow can<br />

be achieved at very high angle <strong>of</strong> attack and stall itself<br />

occurs more gently. To account for these effects numerous<br />

approaches were derived in literature which either<br />

consider rotational effects by correcting calculated 2Dairfoil<br />

polars [6], [7] or directly implement a modified<br />

boundary layer model into the airfoil code [8].<br />

Most <strong>of</strong> these 2D-corrections were derived for slow turning<br />

wind turbines. Within the current project rotational<br />

effects on the airfoil characteristics will be considered<br />

by evaluating both CFD and experimental data<br />

gathered with a propeller test bench (see Figure8).<br />

Grid Generation<br />

GRIDGEN<br />

Figure 1: Flowchart <strong>of</strong> process chain<br />

2.2 CAD design and structural analysis<br />

The propeller geometry output from the XROTOR code<br />

is imported by the 3D-CAD-CAM s<strong>of</strong>tware package<br />

CATIA by means <strong>of</strong> a macro. This procedure ensu-


es a time-efficient generation <strong>of</strong> a watertight propeller<br />

surface which can be used as a basis for further<br />

grid generation. As the subsequent grid generation for<br />

the numerical flow simulations is script based as well,<br />

it’s advantageous when each propeller geometry has<br />

the same layout to be uniquely identified by the mesh<br />

generating s<strong>of</strong>tware independent <strong>of</strong> the detailed blade<br />

geometry. The propeller surface is exported in IGESformat<br />

which is the basis for mesh generation as well<br />

as for milling the moulds <strong>of</strong> selected propeller designs<br />

for further experimental studies. In addition CATIA is<br />

used for the structural analysis <strong>of</strong> the propeller blades,<br />

propeller shaft and clamping hub.<br />

2.3 Grid generation<br />

The computational mesh for the numerical flow simulations<br />

is based on the IGES surface data <strong>of</strong> the propeller<br />

blades and generated with the meshing s<strong>of</strong>tware<br />

GRIDGEN from POINTWISE [9]. For time-efficient<br />

grid generation <strong>of</strong> multiple different blade geometries<br />

and blade numbers this procedure was realised within<br />

a script based approach. For first preliminary numerical<br />

studies including interference effects <strong>of</strong> the propeller<br />

blades with a motor cowling an unstructered<br />

approach was realised first. In addition, a completely<br />

structered procedure is persecuted. As no wall function<br />

is applied in the CFD calculations a y + -value <strong>of</strong><br />

about 1 is anticipated for the first cell layer. In the<br />

grid 30 cell layers are extruded from the blade surface<br />

to represent the boundary layer on the blades. In addition<br />

the cell distribution on the propeller blades and<br />

the farfield and its extent are easily adaptable within<br />

the script. This procedure is necessary as the numerical<br />

flow simulations are the basis for the calculation<br />

<strong>of</strong> noise levels both in the vicinity <strong>of</strong> the blade as well<br />

as in the near farfield by evaluating the pressure fluctuations.<br />

As no special CAA method is used it must<br />

be granted that no erroneous sound levels result due<br />

to inappropriate discretization <strong>of</strong> the farfield.<br />

2.4 Numerical flow simulations<br />

The numerical flow simulations are conducted with the<br />

flow solvers TAU [10] and FLOWer, both developed by<br />

the DLR.<br />

TAU is a three-dimensional, parallel, hybrid, multigrid<br />

code. It implements a finite volume scheme for solving<br />

the Reynolds-averaged Navier-Stokes (RANS) equations.<br />

The code uses explicit time stepping, the multistep<br />

Runge-Kutta scheme and optionally an implicit<br />

time stepping with a LU-time scheme. For accelerating<br />

the convergence to steady state a local time-stepping<br />

concept, different residual smoothing algorithms and<br />

a geometrical multigrid method are implemented.<br />

At the beginning the main focus <strong>of</strong> the project is to<br />

examine different propeller configurations in regard to<br />

noise emission and aerodynamic behaviour. In combination<br />

with the preliminary studies with XROTOR<br />

these numerical flow simulations <strong>of</strong>fer detailed output<br />

for the selection or design <strong>of</strong> adapted propeller airfoil<br />

sections operating without separation under <strong>of</strong>fdesign<br />

conditions <strong>of</strong> the propeller. At this stage propeller<br />

operation is restricted to axial inflow. Therefore<br />

time-consuming unsteady calculations and the application<br />

<strong>of</strong> the chimera technique are not necessary as a<br />

quasi-steady approach is adequate. To consider interference<br />

effects the cowling is modelled as an axisymmetric<br />

body. To prevent a rotating boundary layer on<br />

the cowling EULER boundary conditions are applied<br />

to this surface whereas the propeller blades are calculated<br />

fully viscous.<br />

2.5 Experimental investigations<br />

The design loop is closed by experimental studies conducted<br />

with a propeller test bench where selected propeller<br />

designs can be further tested. This facility <strong>of</strong>fers<br />

the possibility to thoroughly investigate the predicted<br />

performance from the previous flow simulations<br />

and serves simultaneously to validate the used codes.<br />

This is especially important as the project is aimed<br />

at the reduction <strong>of</strong> noise under <strong>of</strong>f-design conditions<br />

when partially high flow velocities and flow separation<br />

on the propeller blades have to be expected.<br />

Propeller test bench<br />

The propeller test bench was constructed and built by<br />

Steinbeis-Transferzentrum [11]. The propeller is driven<br />

by a compact air-cooled servo motor. Propeller thrust<br />

is measured with a load cell mounted in-between the<br />

rear side <strong>of</strong> the motor and a rib in the motor cowling.<br />

The motor is mounted rotatable about its length axis<br />

so that propeller torque can be measured with a second<br />

lateral mounted load cell (Figure 2). With an additional<br />

turnable rig two prandtl sensors can be positioned<br />

near the propeller plane, one sensor being positioned<br />

ahead and the second one behind the propeller plane.<br />

Evaluation <strong>of</strong> the static and dynamic pressure at these<br />

two locations enables the determination <strong>of</strong> ciculation<br />

at a specific radial position on the blade and therefore<br />

<strong>of</strong>fers additional information on possible separation in<br />

addition to paint coatings. Within the current project<br />

the propeller diameters to be examined are about 1m<br />

und the rotational speed about 3250U/min. This rotational<br />

speed at the test bench results from the chosen<br />

scale factor between the current reference propeller <strong>of</strong><br />

the Zeppelin NT and the model propeller to guarantee<br />

the same blade tip mach number <strong>of</strong> 0.5 between original<br />

and model propeller which is essential to conserve<br />

aeroacoustic similarity. Reynolds number effects between<br />

both propeller scales can be determined by inve-


stigating the original propeller at a second test bench<br />

with identical propulsion units from the Zeppelin NT.<br />

velocities <strong>of</strong> about 7 m/s at the outside location <strong>of</strong> the<br />

test bench.<br />

Acoustic measurements<br />

Figure 2: Propeller test bench for model propellers<br />

Test facility Gust Wind Tunnel<br />

The experimental studies are conducted at the Gust<br />

Wind Tunnel <strong>of</strong> the institute. This wind tunnel was<br />

concipated to simulate gusts on wind turbines by suddenly<br />

opening and closing sails at the inlet <strong>of</strong> the tunnel.<br />

The tunnel itself has a diameter <strong>of</strong> 6.3m and the<br />

length <strong>of</strong> the test section is about 8m with a maximum<br />

flow velocity <strong>of</strong> 18m/s which corresponds to the effective<br />

inflow velocity at the stern propeller <strong>of</strong> the Zeppelin<br />

NT. For acoustic measurements the test bench is<br />

operated outside the test section to avoid disturbances<br />

from the tunnel walls. For the Zeppelin NT noise levels<br />

are highest when the airship is manoeuvering on the<br />

ground and the propeller operates under nearly static<br />

thrust conditions. The inflow area <strong>of</strong> the wind tunnel<br />

was extended with a mobile frame construction with<br />

nets <strong>of</strong> varying porosity fastened between the spars<br />

(see Figure 8). Because <strong>of</strong> the missing contraction <strong>of</strong><br />

the wind tunnel itself this construction guarantees a<br />

more laminar and steady flow in the test section and<br />

the sensitivity <strong>of</strong> the tunnel concerning atmospheric<br />

turbulence is reduced. On the other hand, the canalization<br />

<strong>of</strong> the flow due to this device still <strong>of</strong>fers flow<br />

Acoustic noise level contour maps dependent on flow<br />

veloctity and blade angle <strong>of</strong> the propeller will be captured<br />

by ground microphones. This approach delivers<br />

the total sound pressure level at a specific location but<br />

<strong>of</strong>fers no indication <strong>of</strong> the noise source distribution on<br />

the blades. To clarify sound generating mechanisms<br />

and to localize the dominating noise sources along a<br />

propeller blade especially under <strong>of</strong>f-design conditions<br />

an acoustic mirror with elliptical shape was designed<br />

which serves as a highly directional microphone system.<br />

The characteristic <strong>of</strong> an elliptic mirror is to focus<br />

sound waves emanating from a volume element<br />

<strong>of</strong> a source distribution at the far focus upon the near<br />

focus point where a microphone is located. Due to<br />

constructive superposition <strong>of</strong> sound waves at the microphone<br />

a gain is achievable dependent on the shape<br />

<strong>of</strong> the mirror. Provided that there is no significant flow<br />

between sound source and mirror both spatial resolution<br />

and gain factor <strong>of</strong> the mirror system are mainly<br />

limited by diffraction <strong>of</strong> the sound waves at the edge<br />

<strong>of</strong> the mirror.<br />

As mentioned before the acoustic measurements will<br />

be conducted outside the test section <strong>of</strong> the wind tunnel.<br />

In this case the test bench is mounted on an additional<br />

framework to keep the propeller axis centered<br />

in the tunnel axis. The acoustic mirror is positioned<br />

ahead or behind the propeller and can be moved parallel<br />

to the propeller rotating plane with a linear drive<br />

unit. The focus point is adjusted scarcely above the<br />

blade surface <strong>of</strong> a horizontally levelled blade. By moving<br />

the mirror parallel to the propeller plane from one<br />

blade tip to the other, the noise distribution <strong>of</strong> sources<br />

emanating from the vicinity <strong>of</strong> the blade surface can be<br />

captured. By positioning the mirror ahead or behind<br />

the propeller noise sources emanating rather from the<br />

upper or lower side can be detected (Figure 3). As the<br />

distance <strong>of</strong> the focus point to the blade surface slightly<br />

varies dependent on blade angle and airfoil thickness<br />

distribution it has to be studied if a significant difference<br />

between both positions can be detected at all.<br />

In the design <strong>of</strong> the mirror some geometric restrictions<br />

were given by the dimensions <strong>of</strong> the wind tunnel. The<br />

diameter <strong>of</strong> the mirror should be kept small enough<br />

to avoid any blocking effects on the propeller flow but<br />

large enough to capture low frequencies in the range<br />

<strong>of</strong> the blade passing frequency. The focal length <strong>of</strong> the<br />

mirror is influenced by the desire <strong>of</strong> positioning the<br />

mirror both in front and behind the propeller plane in<br />

the inflow section <strong>of</strong> the tunnel (Figure 3).


Figure 3: Alignment <strong>of</strong> acoustic mirror<br />

Theoretical relations were derived both for the gain<br />

and the spatial resolution <strong>of</strong> an acoustic mirror by<br />

Grosche [12]. The effective gain factor G according to<br />

equation (1) describes the enhancement <strong>of</strong> the sound<br />

intensity caused by the focusing effect <strong>of</strong> the elliptic<br />

mirror at the mirror microphone compared to its free<br />

field value measured by a reference microphone at the<br />

same distance from the sound source. The ability <strong>of</strong> the<br />

acoustic mirror to enhance sound intensities at the far<br />

focus and therefore discriminate background noise makes<br />

this system a highly directional sound microphone<br />

system with the ability to resolve sound sources in all<br />

directions normal to the mirror axis.<br />

( )<br />

IM<br />

G = L M − L Ref = 10 log<br />

I Ref<br />

)<br />

(1)<br />

= 10 log<br />

(0.5 D4<br />

λ 2 B 2<br />

Whereas the focusing effect decreases relatively slowly<br />

in the direction <strong>of</strong> the mirror axis the maximum<br />

in vertical and lateral direction is much more sharper.<br />

Using the theory <strong>of</strong> diffraction <strong>of</strong> waves by a circular<br />

aperture provides an approximate relation for the halfwidth<br />

w 0 <strong>of</strong> the diffraction pattern <strong>of</strong> a point source.<br />

According to Figure 4 w 0 is the distance from the maximum<br />

within which the sound intensity decreases by<br />

50%.<br />

(2) w 0 = 0.54λ S D<br />

In consideration <strong>of</strong> all parameters the final diameter<br />

D <strong>of</strong> the acoustic mirror was set to 2.45m and the<br />

focal length S to 4.5m with the distance <strong>of</strong> the near<br />

focus B chosen to 1.06m.<br />

The calibration <strong>of</strong> the acoustic mirror was performed<br />

using a Kenwood KA 3080 R preamplifier and a broadband<br />

loudspeaker vifa 10 BGS 119/8 which provides a<br />

balanced frequency response. Both for the mirror and<br />

reference microphone 1/2” Free-field Microphone Type<br />

4190 from Brüel & Kjæer were used. For lack <strong>of</strong><br />

Figure 4: Principle <strong>of</strong> acoustic mirror [12]<br />

a special acoustic room various problems due to reflections<br />

and diffraction emerged during the beginning<br />

<strong>of</strong> the calibration process. For this reason the principle<br />

<strong>of</strong> the impulse response measurement <strong>of</strong> the s<strong>of</strong>tware<br />

package ARTA [13] was utilized to generate a<br />

swept sine impulse <strong>of</strong> about 5 ms short enough to avoid<br />

ground reflections at the microphones. The corresponding<br />

ARTA measurement setup for single channel<br />

measurement setup is shown in Figure 5. Due to the<br />

single channel setup the response <strong>of</strong> mirror and reference<br />

microphone were not measured simultaneously<br />

but in two subsequent measurements.<br />

Figure 5: Single channel measurement setup for acoustical<br />

measurements with ARTA s<strong>of</strong>tware<br />

[13]<br />

The resulting gain factor <strong>of</strong> the elliptical mirror is<br />

shown in Figure 6 and the spatial resolution in Figure<br />

7. The effective gain factor shows a good agreement<br />

compared to the theoretical relation according to eq.<br />

(1) derived by Grosche [12]. In the relevant frequency<br />

range a gain <strong>of</strong> about 10 to 40dB is attainable.<br />

Above frequencies <strong>of</strong> about 10kHz a decrease in the<br />

measured gain factor <strong>of</strong> the mirror can be detected<br />

which is caused by the non-omnidirectional directivity<br />

<strong>of</strong> the loudspeaker. These effects have to be corrected


y measuring the directivity pattern <strong>of</strong> the loudspeaker.<br />

The measured spatial resolution <strong>of</strong> the mirror widely<br />

agrees with the linear trend <strong>of</strong> equation (2) but<br />

shows slightly better values.<br />

40<br />

Reference Microphone<br />

Acoustic Mirror Microphone<br />

Theoretical Gain<br />

Measured Gain<br />

20<br />

SPL [dB]<br />

0<br />

-20<br />

Figure 8: Measurement setup<br />

-40<br />

-60<br />

10 3 10 4<br />

f [Hz]<br />

Figure 9 shows a closer view <strong>of</strong> the described servo<br />

controlled rig to measure static and dynamic pressures<br />

in front and behind the propeller plane.<br />

Figure 6: Effective gain factor <strong>of</strong> the elliptic mirror as<br />

function <strong>of</strong> sound frequency<br />

100<br />

Theoretical values<br />

Measured values<br />

80<br />

w 0<br />

[mm]<br />

60<br />

40<br />

20<br />

20 40 60 80 100<br />

λ [mm]<br />

Figure 7: Width <strong>of</strong> diffraction pattern<br />

Measurement setup<br />

Figure 8 shows a typical measurement setup for acoustic<br />

measurements in front <strong>of</strong> the wind tunnel. The<br />

acoustic mirror is positioned in front <strong>of</strong> the propeller<br />

with the far focus adjusted in the vicinity <strong>of</strong> a horizontally<br />

levelled blade. The mirror can be moved parallel<br />

to the propeller plane using a linear drive unit. In addition<br />

a ground microphone is placed on a metal plate in<br />

front <strong>of</strong> the propeller. To registrate possible atmospheric<br />

turbulences a sonic anemometer is located besides<br />

the propeller to measure local wind speeds.<br />

Figure 9: Rig for measuring static and dynamic pressures<br />

at the propeller plane


3 Results<br />

In the following two characteristic operating conditions<br />

<strong>of</strong> the reference propeller will be presented to illustrate<br />

general effects <strong>of</strong> aerodynamic phenomenons<br />

and noise radiation <strong>of</strong> the propeller. The measurements<br />

were conducted outside the wind tunnel with<br />

the measurement setup according to Figure 8. The<br />

propeller was operating under static thrust conditions<br />

(u ∞ = 0). Some preliminary results from the aeroacoustic<br />

measurements are given in noise plots where<br />

the sound pressure level is shown dependent <strong>of</strong> the<br />

sound frequency and the dimensionless radial position<br />

on the blade focused by the mirror. Positive values <strong>of</strong><br />

the radial position mean a position where the blade is<br />

moving downward and negative values a position where<br />

the blade is moving upward. In the analysis <strong>of</strong> the<br />

acoustic data the doppler effect caused by the rotation<br />

<strong>of</strong> the blades and the inclined adjustment <strong>of</strong> the mirror<br />

is not yet considered. This effect leads to a shift<br />

<strong>of</strong> the fequency to lower values for the downward moving<br />

blade and a shift to larger values for the upward<br />

moving blade while sound pressure levels remain unchanged.<br />

The increment <strong>of</strong> the acoustic mirror while<br />

scanning the propeller blades was set to 50mm which<br />

corresponds to a dimensionless propeller radius increment<br />

<strong>of</strong> 0.1. The sample rate <strong>of</strong> the mirror microphone<br />

was 44kHz. The acoustic data was postprocessed<br />

with a FFT analysis with a bandwidth <strong>of</strong> 3Hz. The<br />

sound pressure levels <strong>of</strong> the acoustic mirror represent<br />

the measured values at the mirror microphone. If the<br />

sound pressure level in the vicinity <strong>of</strong> the blade surface<br />

is <strong>of</strong> interest, these values have to be corrected according<br />

to the distance law under consideration <strong>of</strong> the<br />

focal length <strong>of</strong> the mirror according to Figure 3 and<br />

4. The ground microphome was covered with a wind<br />

shield which has not been calibrated yet. All in all the<br />

following noise pressure values are intended to illustrate<br />

general effects between two characteristic test cases<br />

and not absolute values at this stage <strong>of</strong> the project.<br />

3.1 <strong>Operating</strong> point with positive<br />

propeller thrust<br />

The first testcase demonstrates an operating point<br />

when the propeller is generating positive thrust. According<br />

to the CFD calculations and paint coatings<br />

there are no separations on the blades. Figure 10 illustrates<br />

the flow field calculated by TAU in the vicinity<br />

<strong>of</strong> the propeller under static conditions. The resulting<br />

vortex about 4 propeller radii downstream and the resuction<br />

<strong>of</strong> air lead to a divergent wake.<br />

Figure 11 shows a three-dimensional plot <strong>of</strong> the noise<br />

spectrum for this condition. The locations <strong>of</strong> dominant<br />

noise sources can be clearly seen as peaks. Figure 12<br />

shows the corresponding projection <strong>of</strong> this plot. Three<br />

main locations <strong>of</strong> higher noise levels are apparent. As<br />

the focusing effect in the direction <strong>of</strong> the mirror axis<br />

decreases slowly the noise radiation <strong>of</strong> the servo motor<br />

can still be clearly seen in the hub region. A second<br />

region with higher noise levels is located between 50<br />

to 80% <strong>of</strong> the propeller blade. This corresponds to the<br />

region where propeller circulation is highest and most<br />

<strong>of</strong> the thrust is generated for normal operating conditions.<br />

A third peak is found at the blade tips.<br />

Figure 10: Flow field about reference propeller generating<br />

positive thrust, u ∞ = 0<br />

SPL [dB]<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

-1 -0.5 0 0.5 1<br />

r/R [-]<br />

0<br />

5000<br />

10000<br />

15000<br />

20000<br />

f [Hz]<br />

SPL [dB]<br />

65<br />

60<br />

55<br />

50<br />

45<br />

40<br />

35<br />

30<br />

25<br />

Figure 11: Sound pressure level 1/6-octave, positive<br />

thrust<br />

It’s striking that the noise peaks are not exactly symmetric<br />

to the location <strong>of</strong> the hub. The symmetry plane<br />

is slightly shifted to the side <strong>of</strong> the upward moving blade.<br />

Although the acoustic tests were done under static<br />

conditions bearing Figure 3 and 8 in mind propeller<br />

induced flow velocities in the inside <strong>of</strong> the acoustic


Figure 12: Sound pressure level 1/6-octave, positive<br />

thrust<br />

Figure 14: Flow field <strong>of</strong> reference propeller generating<br />

reverse thrust, u ∞ = 0<br />

SPL [dB]<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

for this operating point and the airfoils do not deliver<br />

high negative lift coefficients. As the propeller was originally<br />

designed for forward flight conditions the lift<br />

coefficient curve is inflated at the inboard sections <strong>of</strong><br />

the propeller blade. Changing the blade angle to lower<br />

pitch angles causes a nearly vertical shift <strong>of</strong> the lift<br />

coefficient curve. Under reverse thrust conditions this<br />

fact is responsible that at the inboard sections <strong>of</strong> the<br />

blade still positive lift coefficients prevail. According to<br />

Figure 14 this effect causes a recirculation area at the<br />

inboard sections <strong>of</strong> the blade in front <strong>of</strong> the propeller.<br />

10<br />

0<br />

0 1 2 3 4 5<br />

f [kHz]<br />

Figure 13: <strong>Noise</strong> spectrum <strong>of</strong> ground microphone, positive<br />

thrust<br />

mirror can be responsible for an apparent shift <strong>of</strong> the<br />

noise sources.<br />

Figure 13 shows the noise spectrum <strong>of</strong> the ground microphone<br />

located in front <strong>of</strong> the propeller. Tonal noise<br />

components are dominating as far as about 5kHz<br />

and the harmonics <strong>of</strong> the blade passing fequency <strong>of</strong><br />

162.5Hz can be clearly seen.<br />

3.2 <strong>Operating</strong> point with reverse propeller<br />

thrust<br />

Manoeuvering on the ground the Zeppelin NT also<br />

uses the variable pitch stern thruster to generate reverse<br />

thrust although the used propeller was not designed<br />

Figure 15: Sound pressure level 1/6-octave, reverse<br />

thrust<br />

Compared to the previous operation case the sound<br />

pressure levels for reverse thrust are much higher (Fi-


gure 15) with the maxima shifted to the tip region.<br />

This effect is caused by separations on the lower side<br />

<strong>of</strong> the propeller blade in this region. Again the maxima<br />

are not symmetric to the hub region. For this operating<br />

point the symmetry axis is shifted to the side <strong>of</strong><br />

the downward moving blade.<br />

The noise spectrum <strong>of</strong> the ground microphone supports<br />

the assumption <strong>of</strong> separations in the tip region<br />

as broadband noise components dominate and tonal<br />

noise components are restricted to the first two harmonics<br />

<strong>of</strong> the blade passing frequency.<br />

SPL [dB]<br />

100<br />

80<br />

60<br />

40<br />

20<br />

f 0<br />

f 1<br />

0 1 2 3 4 5<br />

f [kHz]<br />

Figure 16: <strong>Noise</strong> spectrum <strong>of</strong> ground microphone, reverse<br />

thrust<br />

4 Conclusion and Outlook<br />

The development <strong>of</strong> a process chain for highly automated<br />

propeller design along with first experimental<br />

results conducted with a propeller test bench were presented.<br />

The aeroacoustic experiments show significant<br />

separations in the tip region <strong>of</strong> the propeller blade if<br />

the propeller operates in reverse mode. One reason for<br />

the current asymmetry in regard to the hub in the<br />

acoustic plots are caused by propeller induced flow<br />

velocities in the acoustic mirror. This effect will be<br />

pretty hard to account for as the flow field is strongly<br />

dependent on the operating point <strong>of</strong> the propeller and<br />

the flow field <strong>of</strong> the wind tunnel. From the beginning<br />

the use <strong>of</strong> the acoustic mirror was intended to show<br />

relative differences between different blade angles or<br />

propeller designs rather than absolute noise levels. Nonetheless<br />

this effect will be studied furthermore with<br />

the intention to find appropriate corrections. Another<br />

aspect in regard to the asymmetry between downward<br />

and upward moving blade is caused by the directivity<br />

<strong>of</strong> the trailing edge noise and convective amplification<br />

which has been observed in acoustic array measurements<br />

on a wind turbine as well [14]. These effects have<br />

to be considered in the further analysis <strong>of</strong> the acoustic<br />

data. In additon, further tests have to be done concerning<br />

the repeatability <strong>of</strong> the noise measurements<br />

especially under static thrust condition when the flow<br />

about the blade tips is quite sensitive to external disturbances.<br />

References<br />

[1] F. Bruce Metzger: An Assessment <strong>of</strong> Propeller Aircraft<br />

<strong>Noise</strong> <strong>Reduction</strong> Technology. NASA Contractor<br />

Report 198237, 1995.<br />

[2] H. Patrick, R.W. Finn and C.K. Stich: Two and<br />

Three-bladed Propeller Design for the <strong>Reduction</strong> <strong>of</strong><br />

Radiated <strong>Noise</strong>. AIAA-97-1710-CP, 1997.<br />

[3] M. Drela: XROTOR User Guide. MIT, August<br />

1996.<br />

[4] G.P. Succi: Design <strong>of</strong> Quiet Efficient Propellers. SAE<br />

Paper 790584, 1979.<br />

[5] G. Succi, D. Munro, J. Zimmer: Experimental Verification<br />

<strong>of</strong> Propeller <strong>Noise</strong> Prediction. AIAA-80-0994.<br />

[6] D.V. Uhlig and M.S. Selig: Post Stall Propeller<br />

Behavior at Low Reynolds Numbers. 6th AIAA Aerospace<br />

Sciences Meeting and Exhibit, 7 - 10 January<br />

2008, Reno, Nevada AIAA 2008-407<br />

[7] O. Gur and A. Rosen: Propeller Performance at<br />

Low Advance Ratio. Journal <strong>of</strong> Aircraft, Vol. 42, No. 2,<br />

March-April 2005.<br />

[8] B.O.G. Montgomerie, A.J. Brand, J. Bosschers,<br />

R.P.J.O.M. van Rooij: Three-Dimensional Effects<br />

in Stall. ECN-C–96-079, 1997.<br />

[9] CFD Meshing S<strong>of</strong>tware GRIDGEN. Homepage:<br />

http://www.pointwise.com<br />

[10] Technical Documentation <strong>of</strong> the DLR TAU-Code.<br />

Technical Report, Deutsches Zentrum für Luft- und<br />

Raumfahrt e. V.<br />

[11] Steinbeis Flugzeug- und Leichtbau GmbH. Homepage:<br />

http://www.sfl-gmbh.de<br />

[12] F.-R. Grosche, G. Schneider, H. Stiewitt: Wind<br />

Tunnel Experiments on Airframe <strong>Noise</strong> Sources <strong>of</strong><br />

Transport Aircraft. AIAA-97-1642-CP, 1997.<br />

[13] I. Mateljan: ARTA Program for Impulse Rresponse<br />

Measurement and Real Time Analysis <strong>of</strong> Spectrum<br />

and Frequency Response User Manual. Electroacoustics<br />

Laboratory, Faculty <strong>of</strong> electrical engineering,<br />

R. Boskovica bb., 21000 Split, Croatia.<br />

[14] S. Oerlemans, P. Sijtsma, B. Méndez López: Location<br />

and quantification <strong>of</strong> noise sources on a wind<br />

turbine. Journal <strong>of</strong> Sound and Vibration 299 (2007)<br />

869-883.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!