25.02.2013 Views

Aerodynamic design assessment of Strato 2C ... - UFSC Aerodesign

Aerodynamic design assessment of Strato 2C ... - UFSC Aerodesign

Aerodynamic design assessment of Strato 2C ... - UFSC Aerodesign

SHOW MORE
SHOW LESS

Create successful ePaper yourself

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

Aerospace Science and Technology 6 (2002) 43–51<br />

www.elsevier.com/locate/aescte<br />

<strong>Aerodynamic</strong> <strong>design</strong> <strong>assessment</strong> <strong>of</strong> <strong>Strato</strong> <strong>2C</strong> and its potential for<br />

unmanned high altitude airborne platforms<br />

Bewertung der aerodynamischen Auslegung von <strong>Strato</strong> <strong>2C</strong> und dessen<br />

Potential für hochfliegende unbemannte Fluggeräte<br />

Dirk Schawe ∗ , Claas-Hinrik Rohardt, Georg Wichmann<br />

DLR – German Aerospace Center, Institute <strong>of</strong> <strong>Aerodynamic</strong>s and Flow Technology, Lilienthalplatz 7, 38108 Braunschweig, Germany<br />

Received 2 August 2001; revised and accepted 15 October 2001<br />

Abstract<br />

Currently, there is a large interest worldwide in the development <strong>of</strong> High Altitude Long Endurance (HALE) Unmanned Aerial Vehicles<br />

(UAVs) for a number <strong>of</strong> civil and military missions, such as routine weather reconnaissance, surface reconnaissance (forest fires, etc.),<br />

earth observation, border patrol and monitoring, fisheries and wildlife refuge management, chemical and biological agent detection, law<br />

enforcement, disaster assistance and monitoring, telecommunications relay, movie production, agricultural surveying and control, and<br />

provision <strong>of</strong> targeting information. Passenger and transport airplanes operate at cruising altitudes <strong>of</strong> maximum 12 000 m where the density is<br />

about 25% compared to sea level. HALE-UVAs are foreseen to operate in the stratosphere at altitudes <strong>of</strong> 24 000 m, twice as high, where the<br />

density drops to about 3.6% <strong>of</strong> the sea level value influencing the lift <strong>of</strong> the aircraft strongly. The environmental conditions in such altitudes<br />

pose strong requirements for the aerodynamic layout and the power plant <strong>of</strong> an aircraft. In Europe <strong>Strato</strong> <strong>2C</strong> – a manned civil research aircraft<br />

– was until nowadays the only aircraft which reached altitudes above 18 000 m. In this paper <strong>Strato</strong> <strong>2C</strong>’s aerodynamic <strong>design</strong> and propulsion<br />

layout will be presented and critically reviewed for its suitability for these high altitudes. © 2002 Éditions scientifiques et médicales Elsevier<br />

SAS. All rights reserved.<br />

Zusammenfassung<br />

Derzeit gibt es weltweit ein grosses Interesse für die Entwicklung unbemannter Fluggeräte für zivile, als auch militärische Missionen,<br />

die in grossen Höhen und über einen langen Zeitraum operieren sollen (HALE-UAVs – High Altitude Long Endurance Unmanned Aerial<br />

Vehicles). Angedachte Missionen sind Wetterbeobachtung, Umweltüberwachung (Waldbrände, etc.), Erderkundung, Überwachung und<br />

Schutz der Staatsgrenzen, Kontrolle von Fisch- und Tierschutzgebieten, Nachweis chemischer und biologischer St<strong>of</strong>fe in der Atmosphäre,<br />

Strafverfolgung, Katastrophenschutz, Telekommunikation, Filmproduktion, Kontrolle und Überwachung der Landwirtschaft und militärische<br />

Aufklärung und Zielerfassung. Passagier- und Transportflugzeuge operieren in Reiseflughöhen von maximal 12 000 m. In diesen Höhen<br />

hat sich die Luftdichte auf ungefähr 25% des Druckes in Meereshöhe reduziert. Die hochfliegenden unbemannten Fluggeräte sollen in<br />

der <strong>Strato</strong>sphäre bis in Höhen von 24 000 m betrieben werden. Hier sinkt die Luftdichte auf 3.6% des Druckes in Meereshöhe, wodurch<br />

der Auftrieb des Luftfahrzeuges stark beeinflusst wird. Die Umweltbedingungen in solchen Höhen beeinflussen damit sehr stark die<br />

aerodynamische Auslegung des Flugkörpers und die Auswahl und Auslegung des Antriebes. Bis zum heutigen Tag war das zivile bemannte<br />

Forschungsflugzeug <strong>Strato</strong> <strong>2C</strong> in Europa das einzige Fluggerät, das Flughöhen oberhalb 18 000 m erreichte. In diesem Artikel werden die<br />

aerodynamische und antriebsseitige Auslegung von <strong>Strato</strong> <strong>2C</strong> vorgestellt and kritisch auf ihre Eignung für grosse Flughöhen überprüft.<br />

© 2002 Éditions scientifiques et médicales Elsevier SAS. All rights reserved.<br />

Keywords: Airfoil; <strong>Aerodynamic</strong> <strong>design</strong>; Flow separation; UAV<br />

Schlüsselwörter: Tragflügelpr<strong>of</strong>il; Aerodynamischer Entwurf; Strömungsablösung; Drohne<br />

* Correspondence and reprints.<br />

E-mail address: dirk.schawe@dlr.de (D. Schawe).<br />

1270-9638/02/$ – see front matter © 2002 Éditions scientifiques et médicales Elsevier SAS. All rights reserved.<br />

PII: S1270-9638(01)01127-0


44 D. Schawe et al. / Aerospace Science and Technology 6 (2002) 43–51<br />

Nomenclature<br />

A Aspect ratio [–]<br />

cd Airfoil section drag coefficient [–]<br />

CD Drag coefficient [–]<br />

CD0 Zero-lift drag coefficient [–]<br />

cl Airfoil section lift coefficient [–]<br />

CL Lift coefficient [–]<br />

cm0 Pitching moment coefficient [–]<br />

CP Pressure coefficient [–]<br />

d Diameter <strong>of</strong> the propeller [m]<br />

D Drag [N]<br />

e Oswald’s efficiency factor [–]<br />

g Acceleration <strong>of</strong> gravity [9.81 m s−2 ]<br />

H Altitude [m]<br />

L Lift [N]<br />

m Take<strong>of</strong>f weight [kg]<br />

1. Introduction<br />

Condor [4] rang in a new type <strong>of</strong> unmanned aircraft<br />

flying fully autonomous from take<strong>of</strong>f through landing in<br />

high altitudes over long endurances. On 9 October 1988<br />

Boeing’s Condor took <strong>of</strong>f the runway for its first flight.<br />

During seven more successful flights Condor achieved two<br />

Ma Mach number [–]<br />

n Revolutions per minute [1 min −1 ]<br />

p Pressure [Pa]<br />

P Power [W]<br />

r/R Relative propeller radius [–]<br />

Re Reynolds number [–]<br />

S Wing area [m 2 ]<br />

T Thrust [N]<br />

v Velocity [m s −1 ]<br />

x/c Relative chord length [–]<br />

α Angle <strong>of</strong> attack [ ◦ ]<br />

β Geometric blade pitch angle [ ◦ ]<br />

η Propeller efficiency [–]<br />

π 3.1416 [–]<br />

ρ Density [kg m −3 ]<br />

world records, one for an altitude <strong>of</strong> 20 416 m, and one<br />

for an endurance <strong>of</strong> 58 hours 11 minutes. Condor was an<br />

all-bonded composite aircraft with a 60.96 meter wingspan<br />

<strong>of</strong> aspect ratio 36.6 and a lift-to-drag ration (L/D) <strong>of</strong> 40<br />

operating continuously at lift coefficients up to 1.35 with<br />

laminar flow over 50% <strong>of</strong> the upper and lower wing surfaces<br />

at Reynolds numbers <strong>of</strong> 1 million. Condor is powered by two<br />

Fig. 1. History <strong>of</strong> manned and unmanned high altitude and long endurance aircrafts.


130.5 kW turbocharged and liquid cooled six cylinder piston<br />

engines.<br />

Condor’s as well as other aircraft <strong>design</strong>s show that aerodynamics<br />

plays an important role for operating manned or<br />

unmanned aircrafts in high altitudes for extended periods <strong>of</strong><br />

time (Fig. 1). The aerodynamic <strong>design</strong> <strong>of</strong> such a vehicle is<br />

greatly complicated because <strong>of</strong> the dramatic atmospheric implication<br />

at an altitude <strong>of</strong> e.g. 24 000 m (see Table 1). Ambient<br />

density and pressure are just a vanishing fraction <strong>of</strong><br />

their sea level values. The lower speed <strong>of</strong> sound increases local<br />

Mach numbers, and higher kinematic viscosity decreases<br />

Reynolds number. This denotes, that airfoils for HALE-<br />

UAVs operate on fairly high lift coefficients (CL ≈ 1.2−1.5)<br />

at relatively low Reynolds numbers (≈ 1.0 × 10 6 )andrelatively<br />

high Mach numbers (≈ 0.4 − 0.6). For these conditions<br />

we are not able to draw upon already in the past developed<br />

airfoils. Sailplane airfoil data are very near, but are<br />

<strong>design</strong>ed for operating on low Mach numbers (≈ 0.05−0.2).<br />

They might serve as starting geometries for computational<br />

airfoil <strong>design</strong> and optimization procedures ([1,2]).<br />

Another important issue for operating aircraft in the<br />

stratosphere, is the selection <strong>of</strong> the power plant. Turb<strong>of</strong>an<br />

engines are reliable, <strong>of</strong> compact size and therefore easy to<br />

integrate into the aircraft airframe, with high acquisition<br />

costs, low maintenance efforts but moderate operation costs<br />

and high climbing performance. Unfortunately the thrust decreases<br />

proportional with the density, i.e. unmodified turb<strong>of</strong>ans<br />

are limited to about 20 000 m. On the other hand<br />

piston engines with all their accessories (compressors, turbochargers,<br />

heat exchangers, two-stage gear box, propeller,<br />

etc.) have a low but constant performance with increasing<br />

altitude but are much more difficult to integrate. The maximum<br />

altitude for turbocharged and liquid cooled piston engines<br />

is about 26 000 m. The propeller layout for high altitude<br />

operation must provide the aircraft with enough thrust<br />

till density ratios <strong>of</strong> 1: 28.<br />

2. <strong>Strato</strong> <strong>2C</strong> – introduction<br />

<strong>Strato</strong> <strong>2C</strong> is <strong>design</strong>ed as an instrument for ozone and<br />

climate research which was a programme <strong>of</strong> the German<br />

Ministry <strong>of</strong> Education and Technology (BMBF). The project<br />

coordinator was DLR (German Aerospace Center) who was<br />

responsible for the project management, the flight operations<br />

and the scientific instrumentation and mission preparation.<br />

The construction was commissioned 1992 by the German<br />

company Grob.<br />

Its main capabilities have been devoted to:<br />

• research <strong>of</strong> the dynamics and chemistry <strong>of</strong> the atmosphere;<br />

• environmental research (e.g. pollution produced by air<br />

traffic);<br />

• exchange process between troposphere and stratosphere;<br />

D. Schawe et al. / Aerospace Science and Technology 6 (2002) 43–51 45<br />

Table 1<br />

Parameters <strong>of</strong> the International Standard Atmosphere for an altitude <strong>of</strong><br />

24 384 m (80 000 ft)<br />

Variable Value Comp. to S.L.<br />

Altitude 24 384 km (80 000 ft) –<br />

Density 0.04353 kg m−3 3.6%<br />

Pressure 2716.62 Pa 2.7%<br />

Temperature 221.03 K 76.8%<br />

Speed <strong>of</strong> sound 298.04 m s−1 87.6%<br />

Kin. viscosity 3.26 × 10−4 m2 s−1 2 227%<br />

Fig. 2. Dimensions <strong>of</strong> the <strong>Strato</strong> <strong>2C</strong>.<br />

Fig. 3. <strong>Strato</strong> <strong>2C</strong> on its maiden flight on 31 March 1995.<br />

• observation <strong>of</strong> land, ocean and polar icecaps from high<br />

altitude.<br />

The aircraft was <strong>design</strong>ed for altitudes up to 24 000 m<br />

and long endurance and range operations (18 000 km) in the<br />

stratosphere. It was supposed to carry a scientific payload <strong>of</strong><br />

800 to 1 000 kg depending on the flight mission. <strong>Strato</strong> <strong>2C</strong> is<br />

fully built <strong>of</strong> fiber composite materials with a take<strong>of</strong>f weight<br />

<strong>of</strong> about 12 000 kg and a wingspan <strong>of</strong> 56.5 m and a length<br />

and height <strong>of</strong> 23.98 m and 7.76 m, respectively (see Fig. 2).<br />

The first successful flight was on 31 March 1995 (see<br />

Fig. 3). 29 test flights were scheduled successfully until<br />

August 1995. At the last flight <strong>Strato</strong> <strong>2C</strong> reached its<br />

maximum ceiling at 18 500 m.


46 D. Schawe et al. / Aerospace Science and Technology 6 (2002) 43–51<br />

Fig. 4. (a) Altitude <strong>design</strong> mission; and (b) long endurance mission.<br />

3. <strong>Strato</strong>’s missions<br />

The aircraft was <strong>design</strong>ed with regard to two main<br />

mission pr<strong>of</strong>iles. The high altitude mission intended for<br />

stratospheric research has a range <strong>of</strong> 7 000 km with a cruise<br />

endurance <strong>of</strong> 8 hours at an altitude <strong>of</strong> 24 000 m and 800 kg<br />

mission payload (see Fig. 4(a)). The long endurance mission<br />

has a total time <strong>of</strong> 48 hours over a range <strong>of</strong> 18 100 km in<br />

18 000 m ceiling (see Fig. 4(b)).<br />

4. <strong>Aerodynamic</strong>s <strong>of</strong> <strong>Strato</strong> <strong>2C</strong><br />

For <strong>Strato</strong> <strong>2C</strong> an airfoil for operating in high altitudes was<br />

<strong>design</strong>ed without disregarding the requirements for start,<br />

climb, descent and landing. The <strong>design</strong> Reynolds number<br />

<strong>of</strong> 1.0 × 10 6 combined with a fairly high lift coefficient at<br />

relatively high Mach number <strong>of</strong> Ma = 0.43 causes normally<br />

laminar separation bubbles with turbulent reattachments on<br />

smooth surfaces which produce additional pressure drag.<br />

These unusual <strong>design</strong> criteria yield to a laminar airfoil –<br />

named LH37 [7] – <strong>of</strong> 17.5% maximum thickness located<br />

at x/c = 0.354 and a maximum camber <strong>of</strong> 0.040 located at<br />

x/c = 0.385. The pitching moment is cm0 =−0.13, and for<br />

lift coefficients between 0.3 and 1.5 low drag is achieved.<br />

The relatively high airfoil thickness affects the structural<br />

weight positively and enlarges the torsional stiffness <strong>of</strong> the<br />

high aspect ratio wing.<br />

The main goal <strong>of</strong> the airfoil <strong>design</strong> was to achieve high<br />

lift coefficients while the drag remains rather low. High lift<br />

occurs from high pressure differences between the upper<br />

and lower surface <strong>of</strong> the airfoil, i.e. the upper surface<br />

<strong>of</strong> the airfoil features high suction and the lower surface<br />

overpressure over the whole chordlength. A low friction drag<br />

will be reached by large laminar extends even at high lift<br />

coefficients. The drawback <strong>of</strong> such a <strong>design</strong> goal is that<br />

the airfoil will respond very sensitively on a poor finish<br />

qualtiy and/or slightest contamination <strong>of</strong> the wing. This<br />

effect is demonstrated in Fig. 5. The subsonic wind-tunnel<br />

MUB located at the DLR in Braunschweig (Germany) was<br />

equipped with a smooth and clean LH37-wing <strong>of</strong> constant<br />

Fig. 5. Experimentally determined lift coefficient versus angle <strong>of</strong> attack for<br />

Ma = 0.39 and Re = 1.1 × 10 6 .<br />

chordlength in order to realize quasi-2D conditions. The<br />

upper curve shows the lift coefficient distribution versus<br />

the angle <strong>of</strong> attack for the unchanged wing. In a second<br />

experiment the point <strong>of</strong> transition was set to 7% <strong>of</strong> the<br />

chordlength on the upper and lower surface in order to<br />

simulate a poor quality and/or contaminated wing. The result<br />

was a significantly reduced performance <strong>of</strong> the airfoil caused<br />

by fully turbulent flow and additionally by the beginning <strong>of</strong> a<br />

trailing edge separation for angles <strong>of</strong> attack <strong>of</strong> α = 7.5 ◦ and<br />

higher. The maximum difference gained by laminar flow in<br />

units <strong>of</strong> the lift coefficient is 0.4 at an angles <strong>of</strong> attack <strong>of</strong><br />

α = 10 ◦ . The airfoil is non-critical because for increasing<br />

angles <strong>of</strong> attack the lift does not breakdown.<br />

The performance <strong>of</strong> the airfoil was calculated with the<br />

two-dimensional viscous aerodynamic <strong>design</strong> and analysis<br />

code ISES <strong>of</strong> Drela/Giles [1] and experimentally verified in<br />

the subsonic wind-tunnel MUB. For a Mach number <strong>of</strong> 0.3,<br />

cl = 1.5 and 7.5 ◦ angle <strong>of</strong> attack we received a very good<br />

conformity in the cp-distribution <strong>of</strong> computer simulation and<br />

experimental measurement. In this case transition through a<br />

laminar separation bubble occurred on the upper surface <strong>of</strong><br />

the airfoil at x/c = 0.38 and on the lower surface at x/c =<br />

0.79 (see Fig. 6). Keeping laminar flow on the airfoil’s upper<br />

surface till 38% chordlength at such high cl’s is a reasonable<br />

value. If the lift coefficient will be reduced to cl = 1.25 the<br />

laminar extend could be held up till 45% chordlength.<br />

Airfoils where the flow remains laminar over a wide<br />

chordlength at high lift coefficients produce low drag over<br />

a wide cl-range as shown in Fig. 7. The drag increase<br />

<strong>of</strong> the airfoil LH37 received from wind-tunnel experiment<br />

compared with a computer simulation for Ma = 0.39 and


Re = 1.2 × 10 6 is pretty small until cl = 1.5. For higher<br />

lift coefficients the drag increases significantly but without<br />

cl-breakdown, which indicates too that the stall behaviour<br />

<strong>of</strong> the airfoil is noncritical. The shapes <strong>of</strong> both polars are<br />

very similar although the simulated one achieved lower drag.<br />

The lower drag values <strong>of</strong> the simulated one is caused by the<br />

restricted mathematical description <strong>of</strong> the flow physics on<br />

which the computer program is based.<br />

Fig. 6. Theoretically and experimentally determined pressure distribution<br />

for Ma = 0.3, cl = 1.5, Re = 1.0 × 10 6 and α = 7.5 ◦ .<br />

Fig. 7. Drag polar for the airfoil LH37 for Ma = 0.39 and Re = 1.2 × 10 6 .<br />

D. Schawe et al. / Aerospace Science and Technology 6 (2002) 43–51 47<br />

<strong>Strato</strong> <strong>2C</strong> has a slight dihedral (1 ◦ ) triple swept wing with<br />

a straight trailing edge and is equipped with an aileron and a<br />

conventional Schempp–Hirth flap.<br />

For the empennage symmetric laminar flow Eppler airfoils<br />

were selected.<br />

4.1. Wing-engine nacelle interference<br />

The engine-nacelle is bluntly mounted on top <strong>of</strong> the wing<br />

without fillets/fairings. In such regions flow separations<br />

occur because <strong>of</strong> an appearing cross-flow caused by the<br />

pressure differences between the flow around the nacelle<br />

and the suction side <strong>of</strong> the airfoil. Due to the pressure<br />

gradients a cross-flow from nacelle to wing is caused. The<br />

boundary layer <strong>of</strong> the wing in the near region <strong>of</strong> the nacelle<br />

will be thickened and therefore susceptible to separation,<br />

especially when the wing operates at high lift coefficients.<br />

Furthermore a vortex developes in the intersection <strong>of</strong> the<br />

wing and nacelle. The separation area is wedge shaped,<br />

running on the upper surface <strong>of</strong> the wing from the pressure<br />

minimum under an angle <strong>of</strong> 45 ◦ spanwise to the trailing edge<br />

(see Fig. 8).<br />

Several methods could possibly weaken or dampen the<br />

affinity <strong>of</strong> separation. Most methods provoke increasing<br />

the energy in the boundary layer against positive pressure<br />

gradients:<br />

• cowling the wing-nacelle intersection with fairings which<br />

will reduce the pressure gradient after the maximum airfoil<br />

thickness;<br />

• modifying the power plant integration by positioning<br />

the nacelle with the engine inlet on the underwing<br />

(see Fig. 9(a)), where the pressure gradients are much<br />

lower, and therefore the boundary layer responds less<br />

sensitive on disturbances. An additional advantage is<br />

the possibility <strong>of</strong> integrating the landing gear into the<br />

nacelle (see Fig. 9(b)), and thereby avoiding the vortex<br />

generating belly fairing;<br />

Fig. 8. Flow separation area in the wing-nacelle intersection.


48 D. Schawe et al. / Aerospace Science and Technology 6 (2002) 43–51<br />

(a)<br />

(b)<br />

Fig. 9. (a) Modified power plant integration; (b) integration <strong>of</strong> the landing<br />

gear into the nacelle (proposed by IABG).<br />

• feeding the decelerated fluid particles <strong>of</strong> the boundary<br />

layer with additional energy by blowing air into flow<br />

direction. The flow velocity will increase and thereby<br />

the separation risk reduced;<br />

• the boundary-layer suction could be performed in the<br />

area <strong>of</strong> increasing pressure, where the retarded flow will<br />

be sucked <strong>of</strong>f, and the non-retarded flow layers will<br />

build the new boundary layer which is more resistent<br />

against flow separation;<br />

• influencing the boundary layer with vortex generators<br />

in order to exchange high-energy fluid particles <strong>of</strong> the<br />

outside flow with the boundary layer.<br />

Three <strong>of</strong> these five methods – fairings, vortex generators<br />

and blowing out – for avoiding flow separation in the wingnacelle<br />

intersection were investigated in the subsonic windtunnel<br />

MUB on a model with scale 1: 22.4.<br />

The surveyed vortex generators could not prevent the flow<br />

from separation. A more successful and less expensive approach<br />

was the modification <strong>of</strong> the wing-nacelle intersection<br />

with different fairings. These experiments made obvious<br />

that an important parameter for avoiding separations completely<br />

is to lengthen the fairing behind the wing trailing<br />

edge. The fairing shown in Fig. 10 eliminated the separation<br />

completely. Another promising solution was blowing air into<br />

the boundary layer at x/c = 0.4 with a velocity <strong>of</strong> 80 m s −1<br />

(�p = 1.5 × 10 5 Pa). For this configuration the separation<br />

Fig. 10. Modified wing-nacelle intersection with an optimized fairing.<br />

vanished completely. For the real aircraft the exhaust gases<br />

<strong>of</strong> the engine could be blown into the wing-nacelle intersection<br />

with a volume flow rate <strong>of</strong> 30 ls −1 .<br />

A more detailed investigation about the wing-nacelle interferences<br />

including wind-tunnel experiments and methods<br />

to reduce their drag contribution are given in [5].<br />

5. Performance evaluation <strong>of</strong> <strong>Strato</strong> <strong>2C</strong>’s propeller<br />

The propulsion system (see Fig. 11) is a combination <strong>of</strong> a<br />

300 kW piston engine with a gas generator which serves as<br />

a turbocharger and provides an additional jet thrust <strong>of</strong> about<br />

12% <strong>of</strong> the propeller thrust at <strong>design</strong> altitude (24 000 m).<br />

Between each compressor stage the air is cooled by intercoolers.<br />

The compressor has an overall compression ratio <strong>of</strong><br />

32 : 1 and provides charge air for the engine’s manifold inlet<br />

at 24 000 m <strong>of</strong> sea level condition.<br />

Each <strong>of</strong> the two engines mounted in large nacelles on<br />

top <strong>of</strong> the wing drives a wooden Kevlar composite coated<br />

variable-pitch five-blade, constant speed and six meter<br />

diameter propeller, working in pusher mode. The propeller<br />

is abnormally strongly twisted from β = 43 ◦ at the propeller<br />

hub to −10 ◦ at the tip.<br />

Fig. 11. Propulsion system <strong>of</strong> <strong>Strato</strong> <strong>2C</strong>.


Fig. 12. Strongly twisted propeller <strong>of</strong> <strong>Strato</strong> <strong>2C</strong>. The four airfoil sections<br />

are taken into account for the evaluation.<br />

Fig. 13. Reynolds number distribution <strong>of</strong> one propeller blade at four<br />

altitudes.<br />

The propeller was <strong>design</strong>ed for operating in the stratosphere<br />

from 12 000 m up to the maximum mission altitude<br />

<strong>of</strong> 24 000 m. In other words, the propeller must cover a remarkable<br />

Reynolds number range from 1.8 × 10 5 to far beyond<br />

1.0 × 10 6 (see Fig. 13). For the performance evaluation<br />

four discrete flight levels 12 000 m, 18 500 m, 22 000 m<br />

and 24 000 m have been selected. Therefore one propeller<br />

blade was discretized into four airfoil sections (see Fig. 12).<br />

The pressure distributions and the drag polars for these four<br />

airfoils were calculated with the two-dimensional viscous<br />

aerodynamic <strong>design</strong> and analysis code ISES <strong>of</strong> Drela and<br />

Giles [1] and published in [6]. The problem <strong>of</strong> calculating<br />

the drag polars <strong>of</strong> various propeller slices are the very low<br />

Reynolds numbers below 200 000 at relatively high Mach<br />

numbers between 0.6 and 0.8. Up to now ISES still <strong>of</strong>fers<br />

the highest accuracy for such flow conditions. But its error<br />

should not be underestimated, and 20% to 30% error should<br />

be taken into account. The performance characteristics <strong>of</strong> the<br />

propeller are calculated on the basis <strong>of</strong> the afore determined<br />

aerodynamic characteristics with a computer program developed<br />

by Hepperle [3]. This computer code is based on the<br />

combined blade-element and momentum theory which accounts<br />

for details <strong>of</strong> the propeller airfoil geometry and their<br />

D. Schawe et al. / Aerospace Science and Technology 6 (2002) 43–51 49<br />

aerodynamic forces in order to <strong>design</strong> and analyse an optimal<br />

propeller for the selected application.<br />

The main <strong>assessment</strong> criterion for propellers is whether<br />

they are capable <strong>of</strong> providing the airplane with sufficient<br />

thrust for level flight at these operation altitudes, and<br />

furthermore if they have an adequate climbing performance<br />

reserve in order to reach the maximum altitude in a finite<br />

period <strong>of</strong> time. The basis for calculating the propeller<br />

performance are the drag polars in various slices, especially<br />

between 60% and 90% <strong>of</strong> the propeller radius where the<br />

main thrust is produced. In Fig. 14 we observe that for<br />

altitudes above operation altitude 2 (18 500 m, see also<br />

Table 2) the drag for the 60% slice experienced a strong<br />

increase, which indicates flow separation. In the 80% slice<br />

the flow is still attached, but separates nearby operation<br />

altitude III (22 000 m). These observations indicate that<br />

the propeller is limited to altitudes between 18 500 m and<br />

22 000 m. This presumption will be confirmed if we look at<br />

the thrust <strong>of</strong> the propeller, the maximum power <strong>of</strong> the engine,<br />

and the drag produced by the aircraft.<br />

This minimal thrust necessary for level flight is approximately<br />

equal to the drag D <strong>of</strong> the airplane flying at velocity<br />

v:<br />

D = ρ<br />

2 v2SCD, with CD = CD0 + C2 L<br />

, (1)<br />

πAe<br />

where S = 150 m2 .<br />

CL is calculated using<br />

CL = mg<br />

ρ<br />

2 v2S with m = 12 000 kg.<br />

e is the Oswald’s efficiency factor which depends<br />

only on the Mach number. A value <strong>of</strong> e = 1 indicates<br />

an elliptical lift distribution. For the calculations e =<br />

0.865 = const, A = 21.28.<br />

However, in the incompressible theory for positive propeller<br />

thrust TP it will be assumed that the flow through the<br />

propeller disk is incompressible and irrotational. From this<br />

theory we are able to calculate the perfect/ideal propeller efficiency<br />

ηid as follows:<br />

�<br />

2P<br />

v = ηid<br />

πρd2 �1/3 , (2)<br />

(1 − ηid)<br />

with P the power output and d the diameter <strong>of</strong> the propeller.<br />

Here, the thrust <strong>of</strong> a perfectly working propeller is<br />

Tid = ηidP<br />

. (3)<br />

v<br />

If we pursue the propeller thrust and the aircraft drag (remark:<br />

propeller thrust given for one engine and aircraft drag<br />

refers only to the half aircraft. The additional jet thrust <strong>of</strong><br />

about 12% <strong>of</strong> the gas generator is not considered.) for maintaining<br />

level flight with increasing altitude in Fig. 15 and


50 D. Schawe et al. / Aerospace Science and Technology 6 (2002) 43–51<br />

Fig. 14. Drag polar <strong>of</strong> the propeller airfoil section at 60% and 80% radius. The roman numbers (I, II, III, IV) indicate the four operation altitudes which are<br />

specified in Table 2.<br />

Table 2<br />

Physical properties for four discrete operation altitudes. Comparison <strong>of</strong> the aircraft drag in level flight with the thrust provided<br />

by the propeller. (Remark: Drag, Thrust and Power are referred to one engine without considering the 12% additional thrust <strong>of</strong><br />

the gas generator.)<br />

Operation altitude I II III IV<br />

Altitude H [m] 12 000 18 500 22 000 24 000<br />

Density ρ [kg m−3 ] 0.30 0.11 0.064 0.047<br />

Flight velocity v [m s−1 ] 62.2 101.3 131.3 153.8<br />

Mach number Ma [–] 0.21 0.34 0.45 0.52<br />

Lift coefficient CL [–] 1.35 1.37 1.42 1.41<br />

Zero–lift drag coefficient CD [–] 0 0.023 0.025 0.026 0.027<br />

Drag coefficient CD [–]<br />

Twist angle β75 [<br />

0.054 0.057 0.061 0.062<br />

◦ ]<br />

Advance ratio<br />

26.0 43.5 48.5 64.0<br />

v/nd[–] 1.088 1.772 2.066 2.418<br />

RPM n [ 1/min] 572 572 636 636<br />

Power output P [kW] 179 298 300 300<br />

Efficiency η [–] 0.87 0.87 0.84 0.64<br />

Perfect efficiency ηid 0.96 0.96 0.97 0.97<br />

η/ηid 0.91 0.91 0.87 0.66<br />

Comparison aircraft drag in level flight D with propeller thrust TP<br />

D [N] 2 376.9 2 468.1 2 524.3 2 563.3<br />

TP [N] 2 502 2 556 1 909 1 252<br />

Fig. 15. Propeller thrust and aircraft drag versus altitude. (Remark: both<br />

forces refer to one Propeller or in lieu <strong>of</strong> the drag, <strong>of</strong> the half aircraft.)<br />

Table 2 we observe that for 18 500 m level flight with only a<br />

very modest climbing capability could be ensured. However,<br />

above 19 000 m the drag exceeds the propeller thrust. For<br />

further climbing to higher altitudes additional power equal<br />

the aircraft weight multiplied with the sine <strong>of</strong> the flight-path<br />

angle (= climbing velocity/flight velocity) is necessary. The<br />

engine however, already operates, at 18 500 m, at its power<br />

limit <strong>of</strong> 300 kW. The relative efficiency <strong>of</strong> the propeller<br />

η/ηid which indicates the performance rate <strong>of</strong> a perfectly<br />

working propeller, drops rapidly from 91% in 18 500 m to<br />

66% in 24 000 m.<br />

6. Conclusion<br />

Airborne platforms for altitudes above 20 000 m are until<br />

nowadays a challenge for the aerodynamic <strong>design</strong>. Each<br />

meter altitude for a given payload is hard-won through<br />

optimizing each detail <strong>of</strong> the aircraft. This is aggravated<br />

by the fact that these details are interdependent. The main<br />

<strong>design</strong> challenges are:


• a laminar wing, producing drag as low as possible at<br />

high lift coefficients;<br />

• a drag minimized fuselage;<br />

• a flow optimized wing-fuselage intersection with only<br />

low pressure gradients in order to avoid flow separation;<br />

• the selection <strong>of</strong> a suitable power plant, and its integration<br />

into the airframe without interfering the flow too<br />

much; and<br />

• if the power plant is a piston engine the propeller should<br />

be adjusted optimally to the shaft power <strong>of</strong> the engine<br />

and the altitude range to be covered.<br />

The research aircraft <strong>Strato</strong> <strong>2C</strong> accomplished these requirements<br />

for its <strong>design</strong> altitude <strong>of</strong> 24 000 m only with regard<br />

to the wing <strong>design</strong> based on the airfoil LH37 and the<br />

selection <strong>of</strong> a turbocharged piston engine as power plant.<br />

The other points need improvement. The specific problems<br />

in summary were:<br />

• the belly fairing for the landing gear at the fuselage<br />

generates high drag vortical flow;<br />

• the wing was mounted in high wing configuration<br />

without suitable fairings;<br />

• the large nacelles are just put on the wing with sharp<br />

intersections. No additional devices influencing the flow<br />

in the intersection are envisioned. Large flow separation<br />

areas were the result;<br />

• the propeller together with the 300 kW piston engine<br />

is not capable to bring <strong>Strato</strong> <strong>2C</strong> in its current layout<br />

into altitudes <strong>of</strong> 24 000 m. With reasonable climbing<br />

D. Schawe et al. / Aerospace Science and Technology 6 (2002) 43–51 51<br />

velocities 18 500 m was reached. Further climbing is<br />

possible but not practical because <strong>of</strong> the vanishing<br />

climbing rate.<br />

The preliminary layout <strong>of</strong> such aircrafts is still afflicted<br />

with errors which should not be underestimated. The reasons<br />

are the lack <strong>of</strong> suitable and accurate tools on the theoretical<br />

side (CFD codes) as well on the practical side (windtunnels)<br />

to determine the aerodynamics <strong>of</strong> a body flying at<br />

low Reynolds numbers with high lift coefficients and high<br />

Mach numbers. Nevertheless the available tools point out the<br />

trends with 20–30% accuracy.<br />

References<br />

[1] M. Drela, M.B. Giles, ISES: A two-dimensional viscous aerodynamic<br />

<strong>design</strong> and analysis code, AIAA paper 87-0424, 1987.<br />

[2] R. Eppler, D. Somers, A Computer Program for the Design and<br />

Analysis <strong>of</strong> Low-Speed Airfoils, NASA-TM-80210, 1980.<br />

[3] M. Hepperle, Ein Computerprogramm für den Entwurf und Analyse<br />

von Propellern, Institut für Flugzeugbau der Universität Stuttgart, 1984.<br />

[4] R. Johnstone, N. Arntz, CONDOR – High altitude long endurance<br />

(HALE) automatically piloted vehicle (APV), AIAA paper 90-3279,<br />

1990.<br />

[5] Ph. Tölke, A. Quast, Untersuchungen zur Ablösung im Bereich<br />

der Flügel-Gondel-Verschneidung am Forschungsflugzeug <strong>Strato</strong> <strong>2C</strong>,<br />

DLR-IB 129-96/36, 1996.<br />

[6] G. Wichmann, H. Köster, Leistungsnachrechnung des Propellers des<br />

Höhenforschungsflugzeugs <strong>Strato</strong> <strong>2C</strong>, DLR-IB 129-96/31, 1996.<br />

[7] G. Wichmann, C.H. Rohardt, P. Hirt, Kenndaten für Pr<strong>of</strong>ile: Pr<strong>of</strong>il DLR-<br />

LH37, Luftfahrttechnisches Handbuch – LTH, Band Aerodynamik AD<br />

41102-24, 1998.

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

Saved successfully!

Ooh no, something went wrong!