You are on page 1of 9

Aerospace Science and Technology 6 (2002) 43–51

www.elsevier.com/locate/aescte

Aerodynamic design assessment of Strato 2C and its potential for


unmanned high altitude airborne platforms

Bewertung der aerodynamischen Auslegung von Strato 2C und dessen


Potential für hochfliegende unbemannte Fluggeräte
Dirk Schawe ∗ , Claas-Hinrik Rohardt, Georg Wichmann
DLR – German Aerospace Center, Institute of Aerodynamics and Flow Technology, Lilienthalplatz 7, 38108 Braunschweig, Germany
Received 2 August 2001; revised and accepted 15 October 2001

Abstract
Currently, there is a large interest worldwide in the development of High Altitude Long Endurance (HALE) Unmanned Aerial Vehicles
(UAVs) for a number of civil and military missions, such as routine weather reconnaissance, surface reconnaissance (forest fires, etc.),
earth observation, border patrol and monitoring, fisheries and wildlife refuge management, chemical and biological agent detection, law
enforcement, disaster assistance and monitoring, telecommunications relay, movie production, agricultural surveying and control, and
provision of targeting information. Passenger and transport airplanes operate at cruising altitudes of maximum 12 000 m where the density is
about 25% compared to sea level. HALE-UVAs are foreseen to operate in the stratosphere at altitudes of 24 000 m, twice as high, where the
density drops to about 3.6% of the sea level value influencing the lift of the aircraft strongly. The environmental conditions in such altitudes
pose strong requirements for the aerodynamic layout and the power plant of an aircraft. In Europe Strato 2C – a manned civil research aircraft
– was until nowadays the only aircraft which reached altitudes above 18 000 m. In this paper Strato 2C’s aerodynamic design and propulsion
layout will be presented and critically reviewed for its suitability for these high altitudes.  2002 Éditions scientifiques et médicales Elsevier
SAS. All rights reserved.
Zusammenfassung
Derzeit gibt es weltweit ein grosses Interesse für die Entwicklung unbemannter Fluggeräte für zivile, als auch militärische Missionen,
die in grossen Höhen und über einen langen Zeitraum operieren sollen (HALE-UAVs – High Altitude Long Endurance Unmanned Aerial
Vehicles). Angedachte Missionen sind Wetterbeobachtung, Umweltüberwachung (Waldbrände, etc.), Erderkundung, Überwachung und
Schutz der Staatsgrenzen, Kontrolle von Fisch- und Tierschutzgebieten, Nachweis chemischer und biologischer Stoffe in der Atmosphäre,
Strafverfolgung, Katastrophenschutz, Telekommunikation, Filmproduktion, Kontrolle und Überwachung der Landwirtschaft und militärische
Aufklärung und Zielerfassung. Passagier- und Transportflugzeuge operieren in Reiseflughöhen von maximal 12 000 m. In diesen Höhen
hat sich die Luftdichte auf ungefähr 25% des Druckes in Meereshöhe reduziert. Die hochfliegenden unbemannten Fluggeräte sollen in
der Stratosphäre bis in Höhen von 24 000 m betrieben werden. Hier sinkt die Luftdichte auf 3.6% des Druckes in Meereshöhe, wodurch
der Auftrieb des Luftfahrzeuges stark beeinflusst wird. Die Umweltbedingungen in solchen Höhen beeinflussen damit sehr stark die
aerodynamische Auslegung des Flugkörpers und die Auswahl und Auslegung des Antriebes. Bis zum heutigen Tag war das zivile bemannte
Forschungsflugzeug Strato 2C in Europa das einzige Fluggerät, das Flughöhen oberhalb 18 000 m erreichte. In diesem Artikel werden die
aerodynamische und antriebsseitige Auslegung von Strato 2C vorgestellt and kritisch auf ihre Eignung für grosse Flughöhen überprüft.
 2002 Éditions scientifiques et médicales Elsevier SAS. All rights reserved.

Keywords: Airfoil; Aerodynamic design; Flow separation; UAV

Schlüsselwörter: Tragflügelprofil; Aerodynamischer Entwurf; Strömungsablösung; Drohne

* Correspondence and reprints.


E-mail address: dirk.schawe@dlr.de (D. Schawe).

1270-9638/02/$ – see front matter  2002 Éditions scientifiques et médicales Elsevier SAS. All rights reserved.
PII: S 1 2 7 0 - 9 6 3 8 ( 0 1 ) 0 1 1 2 7 - 0
44 D. Schawe et al. / Aerospace Science and Technology 6 (2002) 43–51

Nomenclature

A Aspect ratio [–] Ma Mach number [–]


cd Airfoil section drag coefficient [–] n Revolutions per minute [1 min−1 ]
CD Drag coefficient [–] p Pressure [Pa]
CD0 Zero-lift drag coefficient [–] P Power [W]
cl Airfoil section lift coefficient [–] r/R Relative propeller radius [–]
CL Lift coefficient [–] Re Reynolds number [–]
cm0 Pitching moment coefficient [–] S Wing area [m2 ]
CP Pressure coefficient [–] T Thrust [N]
d Diameter of the propeller [m] v Velocity [m s−1 ]
D Drag [N] x/c Relative chord length [–]
e Oswald’s efficiency factor [–] α Angle of attack [◦ ]
g Acceleration of gravity [9.81 m s−2 ] β Geometric blade pitch angle [◦ ]
H Altitude [m] η Propeller efficiency [–]
L Lift [N] π 3.1416 [–]
m Takeoff weight [kg] ρ Density [kg m−3 ]

1. Introduction world records, one for an altitude of 20 416 m, and one


for an endurance of 58 hours 11 minutes. Condor was an
Condor [4] rang in a new type of unmanned aircraft all-bonded composite aircraft with a 60.96 meter wingspan
flying fully autonomous from takeoff through landing in of aspect ratio 36.6 and a lift-to-drag ration (L/D) of 40
high altitudes over long endurances. On 9 October 1988 operating continuously at lift coefficients up to 1.35 with
Boeing’s Condor took off the runway for its first flight. laminar flow over 50% of the upper and lower wing surfaces
During seven more successful flights Condor achieved two at Reynolds numbers of 1 million. Condor is powered by two

Fig. 1. History of manned and unmanned high altitude and long endurance aircrafts.
D. Schawe et al. / Aerospace Science and Technology 6 (2002) 43–51 45

130.5 kW turbocharged and liquid cooled six cylinder piston Table 1


engines. Parameters of the International Standard Atmosphere for an altitude of
Condor’s as well as other aircraft designs show that aero- 24 384 m (80 000 ft)

dynamics plays an important role for operating manned or Variable Value Comp. to S.L.
unmanned aircrafts in high altitudes for extended periods of Altitude 24 384 km (80 000 ft) –
time (Fig. 1). The aerodynamic design of such a vehicle is Density 0.04353 kg m−3 3.6%
greatly complicated because of the dramatic atmospheric im- Pressure 2716.62 Pa 2.7%
Temperature 221.03 K 76.8%
plication at an altitude of e.g. 24 000 m (see Table 1). Am- Speed of sound 298.04 m s−1 87.6%
bient density and pressure are just a vanishing fraction of Kin. viscosity 3.26 × 10−4 m2 s−1 2 227%
their sea level values. The lower speed of sound increases lo-
cal Mach numbers, and higher kinematic viscosity decreases
Reynolds number. This denotes, that airfoils for HALE-
UAVs operate on fairly high lift coefficients (CL ≈ 1.2−1.5)
at relatively low Reynolds numbers (≈ 1.0 × 106 ) and rela-
tively high Mach numbers (≈ 0.4 − 0.6). For these condi-
tions we are not able to draw upon already in the past de-
veloped airfoils. Sailplane airfoil data are very near, but are
designed for operating on low Mach numbers (≈ 0.05−0.2).
They might serve as starting geometries for computational
airfoil design and optimization procedures ([1,2]).
Another important issue for operating aircraft in the
stratosphere, is the selection of the power plant. Turbofan
engines are reliable, of compact size and therefore easy to
integrate into the aircraft airframe, with high acquisition Fig. 2. Dimensions of the Strato 2C.
costs, low maintenance efforts but moderate operation costs
and high climbing performance. Unfortunately the thrust de-
creases proportional with the density, i.e. unmodified tur-
bofans are limited to about 20 000 m. On the other hand
piston engines with all their accessories (compressors, tur-
bochargers, heat exchangers, two-stage gear box, propeller,
etc.) have a low but constant performance with increasing
altitude but are much more difficult to integrate. The maxi-
mum altitude for turbocharged and liquid cooled piston en-
gines is about 26 000 m. The propeller layout for high alti-
tude operation must provide the aircraft with enough thrust
till density ratios of 1: 28.

2. Strato 2C – introduction
Fig. 3. Strato 2C on its maiden flight on 31 March 1995.
Strato 2C is designed as an instrument for ozone and
climate research which was a programme of the German
Ministry of Education and Technology (BMBF). The project • observation of land, ocean and polar icecaps from high
coordinator was DLR (German Aerospace Center) who was altitude.
responsible for the project management, the flight operations
and the scientific instrumentation and mission preparation. The aircraft was designed for altitudes up to 24 000 m
The construction was commissioned 1992 by the German and long endurance and range operations (18 000 km) in the
company Grob. stratosphere. It was supposed to carry a scientific payload of
Its main capabilities have been devoted to: 800 to 1 000 kg depending on the flight mission. Strato 2C is
fully built of fiber composite materials with a takeoff weight
• research of the dynamics and chemistry of the atmo- of about 12 000 kg and a wingspan of 56.5 m and a length
sphere; and height of 23.98 m and 7.76 m, respectively (see Fig. 2).
• environmental research (e.g. pollution produced by air The first successful flight was on 31 March 1995 (see
traffic); Fig. 3). 29 test flights were scheduled successfully until
• exchange process between troposphere and strato- August 1995. At the last flight Strato 2C reached its
sphere; maximum ceiling at 18 500 m.
46 D. Schawe et al. / Aerospace Science and Technology 6 (2002) 43–51

Fig. 4. (a) Altitude design mission; and (b) long endurance mission.

3. Strato’s missions

The aircraft was designed with regard to two main


mission profiles. The high altitude mission intended for
stratospheric research has a range of 7 000 km with a cruise
endurance of 8 hours at an altitude of 24 000 m and 800 kg
mission payload (see Fig. 4(a)). The long endurance mission
Fig. 5. Experimentally determined lift coefficient versus angle of attack for
has a total time of 48 hours over a range of 18 100 km in Ma = 0.39 and Re = 1.1 × 106 .
18 000 m ceiling (see Fig. 4(b)).
chordlength in order to realize quasi-2D conditions. The
4. Aerodynamics of Strato 2C upper curve shows the lift coefficient distribution versus
the angle of attack for the unchanged wing. In a second
For Strato 2C an airfoil for operating in high altitudes was experiment the point of transition was set to 7% of the
designed without disregarding the requirements for start, chordlength on the upper and lower surface in order to
climb, descent and landing. The design Reynolds number simulate a poor quality and/or contaminated wing. The result
of 1.0 × 106 combined with a fairly high lift coefficient at was a significantly reduced performance of the airfoil caused
relatively high Mach number of Ma = 0.43 causes normally by fully turbulent flow and additionally by the beginning of a
laminar separation bubbles with turbulent reattachments on trailing edge separation for angles of attack of α = 7.5◦ and
smooth surfaces which produce additional pressure drag. higher. The maximum difference gained by laminar flow in
These unusual design criteria yield to a laminar airfoil – units of the lift coefficient is 0.4 at an angles of attack of
named LH37 [7] – of 17.5% maximum thickness located α = 10◦ . The airfoil is non-critical because for increasing
at x/c = 0.354 and a maximum camber of 0.040 located at angles of attack the lift does not breakdown.
x/c = 0.385. The pitching moment is cm0 = −0.13, and for The performance of the airfoil was calculated with the
lift coefficients between 0.3 and 1.5 low drag is achieved. two-dimensional viscous aerodynamic design and analysis
The relatively high airfoil thickness affects the structural code ISES of Drela/Giles [1] and experimentally verified in
weight positively and enlarges the torsional stiffness of the the subsonic wind-tunnel MUB. For a Mach number of 0.3,
high aspect ratio wing. cl = 1.5 and 7.5◦ angle of attack we received a very good
The main goal of the airfoil design was to achieve high conformity in the cp -distribution of computer simulation and
lift coefficients while the drag remains rather low. High lift experimental measurement. In this case transition through a
occurs from high pressure differences between the upper laminar separation bubble occurred on the upper surface of
and lower surface of the airfoil, i.e. the upper surface the airfoil at x/c = 0.38 and on the lower surface at x/c =
of the airfoil features high suction and the lower surface 0.79 (see Fig. 6). Keeping laminar flow on the airfoil’s upper
overpressure over the whole chordlength. A low friction drag surface till 38% chordlength at such high cl ’s is a reasonable
will be reached by large laminar extends even at high lift value. If the lift coefficient will be reduced to cl = 1.25 the
coefficients. The drawback of such a design goal is that laminar extend could be held up till 45% chordlength.
the airfoil will respond very sensitively on a poor finish Airfoils where the flow remains laminar over a wide
qualtiy and/or slightest contamination of the wing. This chordlength at high lift coefficients produce low drag over
effect is demonstrated in Fig. 5. The subsonic wind-tunnel a wide cl -range as shown in Fig. 7. The drag increase
MUB located at the DLR in Braunschweig (Germany) was of the airfoil LH37 received from wind-tunnel experiment
equipped with a smooth and clean LH37-wing of constant compared with a computer simulation for Ma = 0.39 and
D. Schawe et al. / Aerospace Science and Technology 6 (2002) 43–51 47

Re = 1.2 × 106 is pretty small until cl = 1.5. For higher Strato 2C has a slight dihedral (1◦ ) triple swept wing with
lift coefficients the drag increases significantly but without a straight trailing edge and is equipped with an aileron and a
cl -breakdown, which indicates too that the stall behaviour conventional Schempp–Hirth flap.
of the airfoil is noncritical. The shapes of both polars are For the empennage symmetric laminar flow Eppler air-
very similar although the simulated one achieved lower drag. foils were selected.
The lower drag values of the simulated one is caused by the
restricted mathematical description of the flow physics on 4.1. Wing-engine nacelle interference
which the computer program is based.
The engine-nacelle is bluntly mounted on top of the wing
without fillets/fairings. In such regions flow separations
occur because of an appearing cross-flow caused by the
pressure differences between the flow around the nacelle
and the suction side of the airfoil. Due to the pressure
gradients a cross-flow from nacelle to wing is caused. The
boundary layer of the wing in the near region of the nacelle
will be thickened and therefore susceptible to separation,
especially when the wing operates at high lift coefficients.
Furthermore a vortex developes in the intersection of the
wing and nacelle. The separation area is wedge shaped,
running on the upper surface of the wing from the pressure
minimum under an angle of 45◦ spanwise to the trailing edge
(see Fig. 8).
Several methods could possibly weaken or dampen the
affinity of separation. Most methods provoke increasing
the energy in the boundary layer against positive pressure
gradients:

• cowling the wing-nacelle intersection with fairings which


will reduce the pressure gradient after the maximum air-
Fig. 6. Theoretically and experimentally determined pressure distribution foil thickness;
for Ma = 0.3, cl = 1.5, Re = 1.0 × 106 and α = 7.5◦ . • modifying the power plant integration by positioning
the nacelle with the engine inlet on the underwing
(see Fig. 9(a)), where the pressure gradients are much
lower, and therefore the boundary layer responds less
sensitive on disturbances. An additional advantage is
the possibility of integrating the landing gear into the
nacelle (see Fig. 9(b)), and thereby avoiding the vortex
generating belly fairing;

Fig. 7. Drag polar for the airfoil LH37 for Ma = 0.39 and Re = 1.2 × 106 . Fig. 8. Flow separation area in the wing-nacelle intersection.
48 D. Schawe et al. / Aerospace Science and Technology 6 (2002) 43–51

(a)

Fig. 10. Modified wing-nacelle intersection with an optimized fairing.

vanished completely. For the real aircraft the exhaust gases


of the engine could be blown into the wing-nacelle intersec-
tion with a volume flow rate of 30 ls−1 .
A more detailed investigation about the wing-nacelle in-
terferences including wind-tunnel experiments and methods
(b) to reduce their drag contribution are given in [5].
Fig. 9. (a) Modified power plant integration; (b) integration of the landing
gear into the nacelle (proposed by IABG).
5. Performance evaluation of Strato 2C’s propeller
• feeding the decelerated fluid particles of the boundary
layer with additional energy by blowing air into flow The propulsion system (see Fig. 11) is a combination of a
direction. The flow velocity will increase and thereby 300 kW piston engine with a gas generator which serves as
the separation risk reduced; a turbocharger and provides an additional jet thrust of about
• the boundary-layer suction could be performed in the 12% of the propeller thrust at design altitude (24 000 m).
area of increasing pressure, where the retarded flow will Between each compressor stage the air is cooled by inter-
be sucked off, and the non-retarded flow layers will coolers. The compressor has an overall compression ratio of
build the new boundary layer which is more resistent 32 : 1 and provides charge air for the engine’s manifold inlet
against flow separation; at 24 000 m of sea level condition.
• influencing the boundary layer with vortex generators Each of the two engines mounted in large nacelles on
in order to exchange high-energy fluid particles of the top of the wing drives a wooden Kevlar composite coated
outside flow with the boundary layer. variable-pitch five-blade, constant speed and six meter
diameter propeller, working in pusher mode. The propeller
Three of these five methods – fairings, vortex generators is abnormally strongly twisted from β = 43◦ at the propeller
and blowing out – for avoiding flow separation in the wing- hub to −10◦ at the tip.
nacelle intersection were investigated in the subsonic wind-
tunnel MUB on a model with scale 1: 22.4.
The surveyed vortex generators could not prevent the flow
from separation. A more successful and less expensive ap-
proach was the modification of the wing-nacelle intersec-
tion with different fairings. These experiments made obvious
that an important parameter for avoiding separations com-
pletely is to lengthen the fairing behind the wing trailing
edge. The fairing shown in Fig. 10 eliminated the separation
completely. Another promising solution was blowing air into
the boundary layer at x/c = 0.4 with a velocity of 80 m s−1
(p = 1.5 × 105 Pa). For this configuration the separation Fig. 11. Propulsion system of Strato 2C.
D. Schawe et al. / Aerospace Science and Technology 6 (2002) 43–51 49

aerodynamic forces in order to design and analyse an opti-


mal propeller for the selected application.
The main assessment criterion for propellers is whether
they are capable of providing the airplane with sufficient
thrust for level flight at these operation altitudes, and
furthermore if they have an adequate climbing performance
reserve in order to reach the maximum altitude in a finite
period of time. The basis for calculating the propeller
performance are the drag polars in various slices, especially
between 60% and 90% of the propeller radius where the
main thrust is produced. In Fig. 14 we observe that for
altitudes above operation altitude 2 (18 500 m, see also
Table 2) the drag for the 60% slice experienced a strong
increase, which indicates flow separation. In the 80% slice
the flow is still attached, but separates nearby operation
altitude III (22 000 m). These observations indicate that
Fig. 12. Strongly twisted propeller of Strato 2C. The four airfoil sections
are taken into account for the evaluation.
the propeller is limited to altitudes between 18 500 m and
22 000 m. This presumption will be confirmed if we look at
the thrust of the propeller, the maximum power of the engine,
and the drag produced by the aircraft.
This minimal thrust necessary for level flight is approx-
imately equal to the drag D of the airplane flying at veloc-
ity v:
ρ 2 CL2
D= v SCD , with CD = CD0 + , (1)
2 πAe
where S = 150 m2 .

CL is calculated using
mg
Fig. 13. Reynolds number distribution of one propeller blade at four CL = ρ 2
altitudes. 2v S
with m = 12 000 kg.
The propeller was designed for operating in the strato- e is the Oswald’s efficiency factor which depends
sphere from 12 000 m up to the maximum mission altitude only on the Mach number. A value of e = 1 indicates
of 24 000 m. In other words, the propeller must cover a re- an elliptical lift distribution. For the calculations e =
markable Reynolds number range from 1.8 × 105 to far be- 0.865 = const, A = 21.28.
yond 1.0 × 106 (see Fig. 13). For the performance evalua-
tion four discrete flight levels 12 000 m, 18 500 m, 22 000 m However, in the incompressible theory for positive pro-
and 24 000 m have been selected. Therefore one propeller peller thrust TP it will be assumed that the flow through the
blade was discretized into four airfoil sections (see Fig. 12). propeller disk is incompressible and irrotational. From this
The pressure distributions and the drag polars for these four theory we are able to calculate the perfect/ideal propeller ef-
airfoils were calculated with the two-dimensional viscous ficiency ηid as follows:
aerodynamic design and analysis code ISES of Drela and  1/3
Giles [1] and published in [6]. The problem of calculating 2P
v = ηid , (2)
the drag polars of various propeller slices are the very low πρd 2 (1 − ηid )
Reynolds numbers below 200 000 at relatively high Mach with P the power output and d the diameter of the propeller.
numbers between 0.6 and 0.8. Up to now ISES still offers Here, the thrust of a perfectly working propeller is
the highest accuracy for such flow conditions. But its error
ηid P
should not be underestimated, and 20% to 30% error should Tid = . (3)
be taken into account. The performance characteristics of the v
propeller are calculated on the basis of the afore determined If we pursue the propeller thrust and the aircraft drag (re-
aerodynamic characteristics with a computer program devel- mark: propeller thrust given for one engine and aircraft drag
oped by Hepperle [3]. This computer code is based on the refers only to the half aircraft. The additional jet thrust of
combined blade-element and momentum theory which ac- about 12% of the gas generator is not considered.) for main-
counts for details of the propeller airfoil geometry and their taining level flight with increasing altitude in Fig. 15 and
50 D. Schawe et al. / Aerospace Science and Technology 6 (2002) 43–51

Fig. 14. Drag polar of the propeller airfoil section at 60% and 80% radius. The roman numbers (I, II, III, IV) indicate the four operation altitudes which are
specified in Table 2.

Table 2
Physical properties for four discrete operation altitudes. Comparison of the aircraft drag in level flight with the thrust provided
by the propeller. (Remark: Drag, Thrust and Power are referred to one engine without considering the 12% additional thrust of
the gas generator.)
Operation altitude I II III IV
Altitude H [m] 12 000 18 500 22 000 24 000
Density ρ [kg m−3 ] 0.30 0.11 0.064 0.047
Flight velocity v [m s−1 ] 62.2 101.3 131.3 153.8
Mach number Ma [–] 0.21 0.34 0.45 0.52
Lift coefficient CL [–] 1.35 1.37 1.42 1.41
Zero–lift drag coefficient CD0 [–] 0.023 0.025 0.026 0.027
Drag coefficient CD [–] 0.054 0.057 0.061 0.062
Twist angle β75 [◦ ] 26.0 43.5 48.5 64.0
Advance ratio v/n d [–] 1.088 1.772 2.066 2.418
RPM n [ 1/min] 572 572 636 636
Power output P [kW] 179 298 300 300
Efficiency η [–] 0.87 0.87 0.84 0.64
Perfect efficiency ηid 0.96 0.96 0.97 0.97
η/ηid 0.91 0.91 0.87 0.66
Comparison aircraft drag in level flight D with propeller thrust TP
D [N] 2 376.9 2 468.1 2 524.3 2 563.3
TP [N] 2 502 2 556 1 909 1 252

the aircraft weight multiplied with the sine of the flight-path


angle (= climbing velocity/flight velocity) is necessary. The
engine however, already operates, at 18 500 m, at its power
limit of 300 kW. The relative efficiency of the propeller
η/ηid which indicates the performance rate of a perfectly
working propeller, drops rapidly from 91% in 18 500 m to
66% in 24 000 m.

6. Conclusion
Fig. 15. Propeller thrust and aircraft drag versus altitude. (Remark: both
forces refer to one Propeller or in lieu of the drag, of the half aircraft.) Airborne platforms for altitudes above 20 000 m are until
nowadays a challenge for the aerodynamic design. Each
Table 2 we observe that for 18 500 m level flight with only a meter altitude for a given payload is hard-won through
very modest climbing capability could be ensured. However, optimizing each detail of the aircraft. This is aggravated
above 19 000 m the drag exceeds the propeller thrust. For by the fact that these details are interdependent. The main
further climbing to higher altitudes additional power equal design challenges are:
D. Schawe et al. / Aerospace Science and Technology 6 (2002) 43–51 51

• a laminar wing, producing drag as low as possible at velocities 18 500 m was reached. Further climbing is
high lift coefficients; possible but not practical because of the vanishing
• a drag minimized fuselage; climbing rate.
• a flow optimized wing-fuselage intersection with only
low pressure gradients in order to avoid flow separation; The preliminary layout of such aircrafts is still afflicted
• the selection of a suitable power plant, and its integra- with errors which should not be underestimated. The reasons
tion into the airframe without interfering the flow too are the lack of suitable and accurate tools on the theoretical
much; and side (CFD codes) as well on the practical side (wind-
• if the power plant is a piston engine the propeller should tunnels) to determine the aerodynamics of a body flying at
be adjusted optimally to the shaft power of the engine low Reynolds numbers with high lift coefficients and high
and the altitude range to be covered. Mach numbers. Nevertheless the available tools point out the
trends with 20–30% accuracy.
The research aircraft Strato 2C accomplished these re-
quirements for its design altitude of 24 000 m only with re-
gard to the wing design based on the airfoil LH37 and the References
selection of a turbocharged piston engine as power plant.
The other points need improvement. The specific problems [1] M. Drela, M.B. Giles, ISES: A two-dimensional viscous aerodynamic
in summary were: design and analysis code, AIAA paper 87-0424, 1987.
[2] R. Eppler, D. Somers, A Computer Program for the Design and
• the belly fairing for the landing gear at the fuselage Analysis of Low-Speed Airfoils, NASA-TM-80210, 1980.
[3] M. Hepperle, Ein Computerprogramm für den Entwurf und Analyse
generates high drag vortical flow; von Propellern, Institut für Flugzeugbau der Universität Stuttgart, 1984.
• the wing was mounted in high wing configuration [4] R. Johnstone, N. Arntz, CONDOR – High altitude long endurance
without suitable fairings; (HALE) automatically piloted vehicle (APV), AIAA paper 90-3279,
• the large nacelles are just put on the wing with sharp 1990.
intersections. No additional devices influencing the flow [5] Ph. Tölke, A. Quast, Untersuchungen zur Ablösung im Bereich
der Flügel-Gondel-Verschneidung am Forschungsflugzeug Strato 2C,
in the intersection are envisioned. Large flow separation
DLR-IB 129-96/36, 1996.
areas were the result;
[6] G. Wichmann, H. Köster, Leistungsnachrechnung des Propellers des
• the propeller together with the 300 kW piston engine Höhenforschungsflugzeugs Strato 2C, DLR-IB 129-96/31, 1996.
is not capable to bring Strato 2C in its current layout [7] G. Wichmann, C.H. Rohardt, P. Hirt, Kenndaten für Profile: Profil DLR-
into altitudes of 24 000 m. With reasonable climbing LH37, Luftfahrttechnisches Handbuch – LTH, Band Aerodynamik AD
41102-24, 1998.

You might also like