Koning CR-2019-220236 - FINAL

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

NASA/CR—2019–220236

Airfoil Selection for Mars Rotor Applications


Witold J. F. Koning
Science and Technology Corporation
Ames Research Center, Moffett Field, California

Click here: Press F1 key (Windows) or Help key (Mac) for help

July 2019
This page is required and contains approved text that cannot be
changed.
NASA STI Program ... in Profile

Since its founding, NASA has been dedicated  CONFERENCE PUBLICATION.


to the advancement of aeronautics and space Collected papers from scientific and
science. The NASA scientific and technical technical conferences, symposia, seminars,
information (STI) program plays a key part in or other meetings sponsored or co-
helping NASA maintain this important role. sponsored by NASA.

The NASA STI program operates under the  SPECIAL PUBLICATION. Scientific,
auspices of the Agency Chief Information technical, or historical information from
Officer. It collects, organizes, provides for NASA programs, projects, and missions,
archiving, and disseminates NASA’s STI. The often concerned with subjects having
NASA STI program provides access to the NTRS substantial public interest.
Registered and its public interface, the NASA
Technical Reports Server, thus providing one of  TECHNICAL TRANSLATION.
the largest collections of aeronautical and space English-language translations of foreign
science STI in the world. Results are published in scientific and technical material pertinent to
both non-NASA channels and by NASA in the NASA’s mission.
NASA STI Report Series, which includes the
following report types: Specialized services also include organizing
and publishing research results, distributing
 TECHNICAL PUBLICATION. Reports of specialized research announcements and feeds,
completed research or a major significant providing information desk and personal search
phase of research that present the results of support, and enabling data exchange services.
NASA Programs and include extensive data
or theoretical analysis. Includes compilations For more information about the NASA STI
of significant scientific and technical data program, see the following:
and information deemed to be of continuing
reference value. NASA counterpart of peer-  Access the NASA STI program home page
reviewed formal professional papers but has at http://www.sti.nasa.gov
less stringent limitations on manuscript length
and extent of graphic presentations.  E-mail your question to [email protected]

 TECHNICAL MEMORANDUM.  Phone the NASA STI Information Desk at


Scientific and technical findings that are 757-864-9658
preliminary or of specialized interest,
e.g., quick release reports, working  Write to:
papers, and bibliographies that contain NASA STI Information Desk
minimal annotation. Does not contain Mail Stop 148
extensive analysis. NASA Langley Research Center
Hampton, VA 23681-2199
 CONTRACTOR REPORT. Scientific and
technical findings by NASA-sponsored
contractors and grantees.
NASA/CR—2019–220236

Airfoil Selection for Mars Rotor Applications


Witold J. F. Koning
Science and Technology Corporation
Ames Research Center, Moffett Field, California

National Aeronautics and


Space Administration

Ames Research Center


Moffett Field, CA 94035-1000

July 2019
ACKNOWLEDGMENTS

The author thanks Wayne Johnson, Larry Young, and Alan Wadcock for their helpful
discussions while writing this paper.

Available from:

NASA STI Support Services National Technical Information Service


Mail Stop 148 5301 Shawnee Road
NASA Langley Research Center Alexandria, VA 22312
Hampton, VA 23681-2199 [email protected]
757-864-9658 703-605-6000

This report is also available in electronic form at


http://ntrs.nasa.gov
TABLE OF CONTENTS

List of Figures ................................................................................................................................ iv


List of Tables ................................................................................................................................. iv
Introduction ......................................................................................................................................1
Low-Reynolds-Number Airfoil Performance ..................................................................................1
Performance Evaluation of Unconventional Airfoil Shapes ............................................................2
Conventional Airfoil ..................................................................................................................3
Tripped Airfoil ...........................................................................................................................3
Cambered Plate ..........................................................................................................................4
Corrugated Airfoil ......................................................................................................................4
Thickness Variation ...................................................................................................................5
Shape Optimization Efforts..............................................................................................................6
References ........................................................................................................................................7

iii
LIST OF FIGURES

Figure 1. Maximum lift-to-drag ratio versus Reynolds number. .................................................... 2


Figure 2. Minimum section drag coefficient versus Reynolds number. ......................................... 2
Figure 3. Various representative airfoil shapes versus Reynolds number. ..................................... 3
Figure 4. Performance of various airfoil shapes versus Reynolds number. .................................... 4

LIST OF TABLES

Table 1. Overview of candidate airfoil shapes evaluated for Rec = O(103-104) ..............................5

iv
AIRFOIL SELECTION FOR MARS ROTOR APPLICATIONS

Witold J. F. Koning1

Ames Research Center

INTRODUCTION

This paper provides an overview of design considerations for airfoil choices with rotor
applications in the Martian atmosphere, at very low chord-based Reynolds number flows, around
Rec = O (103-104). The low Reynolds number typical of rotorcraft operation in the Martian
atmosphere reduces the rotor lifting force and efficiency, which is only partially compensated for
by a lower gravity on Mars compared to Earth. Additionally, the low temperature and largely
CO2-based atmosphere of Mars compound the overall aerodynamic problem by resulting in a
lower speed of sound, further constraining rotor operation in the Martian atmosphere by limiting
the maximum rotor tip speed possible so as not to exceed an acceptable tip Mach number.

In light of the expected reduced rotor efficiency, evaluation of airfoils for compressible, low-
Reynolds-number Mars rotor applications is key. Prior research on airfoil optimization and
performance evaluation at low Reynolds numbers, especially in the compressible regime, is
scarce and further investigation is needed. Specifically, the proposed goal stemming from this
overview is to develop airfoils tailored to the unique demands of the second generation of Mars
rotorcraft, i.e. the Mars Science Helicopter (MSH).

This research focuses on the airfoil performance at low Reynolds numbers and hopes to add to
the work performed by Kroo et al.,1 Kunz and Kroo,2 Oyama and Fujii,3 Anyoji et al.,4–6 and
others.

LOW-REYNOLDS-NUMBER AIRFOIL PERFORMANCE

McMasters and Henderson7 provide a summary of attainable airfoil lift-to-drag ratios over a
wide Reynolds number range in Figure 1. Results are collected from a wide variety of
experiments, mostly with conventional airfoil geometries.

At the Reynolds number range under consideration, Rec = O (103-104), the boundary layer can be
fully laminar up to the point of separation without subsequent (turbulent) flow reattachment or
on-body transition.

1
Science and Technology Corporation, NASA Research Park, Moffett Field, CA 94035.

1
Figure 1. Maximum lift-to-drag ratio versus Reynolds number.7

Figure 2. Minimum section drag coefficient versus Reynolds number.7

The Reynolds number effect on the minimum airfoil drag coefficient is presented in Figure 2.
The flow state in absence of laminar-to-turbulent transition is called subcritical and derives its
relatively low efficiency because of (a) the increased pressure drag component from early
separation, and (b), to lesser extent, reduced lift due to an effective camber reduction.

The Reynolds number at which laminar flow over an airfoil begins to exhibit turbulent features
(either due to on-body transition or turbulent reattachment) is called the critical Reynolds
number. Reynolds numbers where turbulent transition always occurs before laminar separation,
or during/after reattachment, are referred to as supercritical.

Finally, compressible flow—versus incompressible airfoil flow—is not well understood for low-
Reynolds-number airfoils. Limited experimental and computational work in the literature—and
performed previously by the author8—suggests that conventional airfoil geometries exhibit
Mach-number sensitivities whereas cambered, flat-plate airfoils seem to be insensitive to Mach
number.4,6

PERFORMANCE EVALUATION OF UNCONVENTIONAL AIRFOIL SHAPES

Overall airfoil performance changes with Reynolds number, especially the relatively poor
performance (in terms of lift-to-drag ratio,7 minimum section drag,7,9 and maximum section
lift10) up to Rec = O (105), serves as the main motivation for (unconventional) airfoil optimization
in this regime. An example of the dramatic change in efficient airfoil shapes crossing this
“barrier” becomes clear in the overview presented in Figure 3, by Lissaman.11 It should be noted
that the dragonfly and pigeon wing airfoil profiles are used in highly unsteady “flapping-wing”
applications; their applicability to relatively steady “rotating-wing” operation is not necessarily
ensured.

2
Figure 3. Various representative airfoil shapes versus Reynolds number.11

To identify future avenues for Mars rotor airfoil optimization research the following basic types
of airfoil are examined: a conventional airfoil, a tripped/rough airfoil, a cambered plate, and a
corrugated airfoil. Important factors in examining such airfoils are (a) airfoil sensitivity to
operating conditions (with regard to laminar-turbulent transition), (b) possible hysteresis
behavior with operating condition, and (c) relative technology readiness level of
(unconventional) airfoil geometries and clear understanding of their aerodynamic behavior and
analytic predictability.

Conventional Airfoil
A conventional airfoil is presumed to be relatively impractical for the transition region between
subcritical and supercritical flow states. This “transition region” between the two flow states is
difficult to analyze and reliably predict12 because of a number of factors including the possible
contribution of external influences such as free-stream turbulence (FST) levels, vibrations, and
surface roughness on boundary layer transition.13,14 Another factor is possible flow hysteresis
(thought to stem mostly from highly unstable laminar separation bubble behavior with changing
angle of attack).12 Finally, unsteady laminar separation bubble features or transient boundary
layer transition behavior, in turn, can give rise to unpredictable rotary-wing flight dynamics.15 If
the airfoils are operating only in subcritical mode, it is possible that cambered-plate airfoils can
attain higher performance than conventional low-Reynolds-number airfoils, such as the Eppler
193, as shown in Figure 3.7 The Jet Propulsion Laboratory (JPL) Mars Helicopter (MH) rotor
airfoils are likely to operate fully subcritical in hover.16 Previous work also indicates the
competitiveness of cambered plates versus airfoils for rotor performance of the JPL Mars
Helicopter technology demonstrator (MHTD).8

Tripped Airfoil
Conventional airfoils with a trip device (or an entirely “rough-surfaced” airfoil) can have lower
Reynolds number sensitivity because of forcing or “fixing” transition.7 The forcing of transition
allows relatively good performance down to lower Reynolds numbers compared to smooth
conventional airfoils.17 However, ensuring a trip is functional—and, thereby, forcing transition—
at very low Reynolds numbers is troublesome, and may be impossible below Rec  30,000. If
transition cannot be guaranteed for the complete operational domain, the applicability of airfoil
trips for rotary-wing application is unlikely because of the resulting unpredictable rotor
performance and flight dynamic characteristics.

3
Cambered Plate
In contrast to a conventional airfoil, a cambered plate (also known as a circular-arc airfoil) is
shown to be almost Reynolds number insensitive because of the resulting small, fixed leading-
edge separation bubble that forms for all but a small (near zero) angle-of-attack range. The
leading-edge flow separation will fix the separation bubble location and effectively limit
Reynolds number sensitivity. This, in turn, reduces the possibility of hysteresis in lift, drag, and
pitching moment as a function of angle of attack. It also reduces airfoil performance sensitivity
to FST and other transient effects stemming from free-stream velocity variation and rotor blade
pitching and flapping motion.

Prior research found that the cambered plate in this regime potentially outperforms conventional
airfoils (in terms of minimum drag, maximum lift-to-drag ratio, and possibly maximum lift
coefficient)7,9,12,14,15,18,19 but the geometry variation for cambered plates in references is
limited.7,12,14,15,20,21

Schmitz12 lists the “advantageous cooperation of tangential incident flow at the leading edge at
large angles of attack with the turbulence effect of the small nose radius”, “the strongly concave
underside, which shares significantly in the lift generation”, “and the comparatively small
camber of the airfoils top side, causing the flow to remain largely attached” as main reasons for
the competitive performance of the cambered plate at these Reynolds numbers.

Corrugated Airfoil
The performance of a corrugated airfoil is potentially Reynolds number independent because of
forcing fixed location(s) of separation. With the same reasoning applied to flat-plate airfoils,
sensitivity to FST and other operating conditions is expected to be low. Research currently
available for steady operation of corrugated airfoils is limited. Performance is likely to only be
competitive at the lower end of the Reynolds number range under investigation, Rec  1,000.21–23
It is also speculated that rotor blades incorporating corrugated airfoils might provide needed
structural bending-moment and torsional stiffness (compared to the cambered-plate airfoils) in
the inboard rotor region, where very low Reynolds numbers occur. Levy and Seifert investigated
dragonfly airfoils, both using computational fluid dynamics (CFD) and experiments in steady
free-stream flow, and found relatively promising performance figures in the range of Rec  2,000
to 8,000.22
The applicability of these various airfoil types in the Reynolds number range under consideration
is assessed based on various published sources in the literature and summarized in Figure 4.

Figure 4. Performance of various airfoil shapes versus Reynolds number.7,11,12,15,24,25

4
Thickness Variation
The effect of a thickness distribution on the cambered plate and discontinuous type airfoils
should be investigated. This could yield a higher structural bending-moment and torsional
stiffness for a set of geometries that are inherently weak in that respect. Thickness distribution
for the discontinuous or corrugated airfoils opens up a new domain—a family of polygonal
airfoils. These can include triangular or diamond shape airfoils. Munday et al.26 investigated a
triangular airfoil in low-Reynolds-number compressible flow.

Table 1 provides an overview of the airfoil geometries and discusses their applicability in the
current low-Reynolds-number regime.

Table 1. Overview of candidate airfoil shapes evaluated for Rec = O (103-104).


Airfoil Re and FST Hysteresis Demonstrated
Geometry Sensitivity With Condition Concept Comments
Conventional Large Hysteresis If outside of critical Can work reliably if
airfoil sensitivity possible Reynolds number Reynolds number is too
possible (laminar region; used for low for boundary layer
separation small unmanned transition throughout
bubble induced) aerial vehicles operational regime like
(UAVs) for the MHTD
Tripped airfoil, If transition is Hysteresis Difficult to ensure Transition needs to
rough airfoil fixed, sensitivity possible if trip works below be guaranteed for all
is minimized bubble occurs Re = 30,000; conditions otherwise
before trip uncertain at higher unpredictable flight
Re < 100,000 dynamics can ensue
Cambered Leading-edge Hysteresis less Used for small Possible stiffness
plate, separation of likely because UAVs or manned issues due to low
curved plate large angle-of- of majority of aerial vehicles thickness/chord ratio
attack range operating (MAVs) (t/c)
reduces conditions with
sensitivity leading-edge
separation
Corrugated Separation at Hysteresis less No rotary-wing Performance only
airfoil corrugation likely because experiments using competitive at lower
features likely of separation corrugated airfoils Re < 10,000
to reduce at corrugation known
sensitivity features
Polygonal Separation at Hysteresis less No rotary-wing Possible mediation of
airfoil corrugation likely because experiments using stiffness issues due to
features likely of separation polygonal airfoils increased t/c compared
to reduce at corrugation known to corrugated airfoil
sensitivity features

5
SHAPE OPTIMIZATION EFFORTS

Figure 4 shows that the geometry of the cambered-plate-type airfoil seems to be a logical avenue
for further research, as its optimal aerodynamic performance is roughly at the expected rotor
chord-based Reynolds number range for Mars rotor applications. The geometry variation for
cambered plates in literature is limited,7,12,14,15,20,21 implying cambered-plate airfoil performance
optimization for plates (wherein nonlinear camber lines, chordwise plate thickness distributions,
and even local corrugation features can be examined) could yield performance improvements
over the “simple cambered plate” often presented in literature. The Reynolds number influence
on optimal shape and high Mach/compressibility effects on optimal performance (for desired
cruise speeds) can be subsequently evaluated.

Airfoil geometry optimization can, unfortunately, produce a solution space with various local
extremes, making gradient-based optimization techniques less applicable because of the
criticality of finding an appropriate starting location. Alternatively, a genetic algorithm for airfoil
optimization allows for a potentially more robust exploration of the solution space for a wide
variety of airfoil shapes, providing greater insight into the aerodynamic performance of various
airfoil shapes.

A custom airfoil design genetic algorithm has been written in Python and is able to optimize
airfoil shapes using a preset number of design variables. The algorithm performs OVERFLOW
grid generation and case execution, variation of airfoil geometry within set constraints, and post
processing of OVERFLOW output. The algorithm has been demonstrated on the NASA Ames
Pleiades supercomputer, and can queue run cases on different nodes and central processing units
(CPUs) depending on the population size being evaluated. Currently the algorithm is operating as
a single-objective optimization (SOO) at fixed alpha or lift coefficient. The optimized airfoils
can be used to generate airfoil C81 input decks to evaluate rotor performance using
comprehensive rotorcraft analyses as done previously for the JPL MHTD development effort.27

Future enhancement of the genetic algorithm airfoil optimization will include the ability to
perform multi-objective optimization (MOO), evaluation of rotor blade thickness/stiffness
spanwise distributions, and ultimately, coupling to the Comprehensive Analytical Model of
Rotorcraft Aerodynamics and Dynamics (CAMRAD) comprehensive rotor analysis software
tool.

The progress is aimed at increasing the understanding of low-Reynolds-number airfoil


performance and developing airfoils tailored to the unique demands of second generation Mars
rotorcraft, i.e. the MSH.

Care must be taken in airfoil selection; direct consequences on various rotor design parameters
such as blade stiffness and structural frequencies, blade chordwise center-of-gravity placement,
and possible aeroelastic effects such as blade flutter or “live twist” must be considered.

6
REFERENCES

1 Kroo, I., Prinz, F., Shantz, M., Kunz, P., Fay, G., Cheng, S., Fabian, T., and Partridge, C.:
The Mesicopter: A Miniature Rotorcraft Concept Phase II Interim Report, Stanford
University, Palo Alto, CA, 2000.
2 Kunz, P. J., and Kroo, I.: Analysis and Design of Airfoils for Use at Ultra-Low Reynolds
Numbers, Fixed and Flapping Wing Aerodynamics for Micro Air Vehicle Applications,
vol. 195, 2001, pp. 35–59.
3 Oyama, A., and Fujii, K.: A Study on Airfoil Design for Future Mars Airplane, 44th AIAA
Aerospace Sciences Meeting and Exhibit, Reno, NV, 2006, p. 1484.
4 Anyoji, M., Nose, K., Ida, S., Numata, D., Nagai, H., and Asai, K.: Low Reynolds Number
Airfoil Testing in a Mars Wind Tunnel, 40th Fluid Dynamics Conference and Exhibit,
Chicago, IL, 2010, p. 4627.
5 Anyoji, M., Nonomura, T., Aono, H., Oyama, A., Fujii, K., Nagai, H., and Asai, K.:
Computational and Experimental Analysis of a High-Performance Airfoil Under Low-
Reynolds-Number Flow Condition, Journal of Aircraft, vol. 51, 2014, pp. 1864–1872.
6 Anyoji, M., Numata, D., Nagai, H., and Asai, K.: Effects of Mach Number and Specific
Heat Ratio on Low-Reynolds-Number Airfoil Flows, AIAA Journal, vol. 53, Oct. 2014,
pp. 1640–1654.
7 McMasters, J., and Henderson, M.: Low-Speed Single-Element Airfoil Synthesis, Technical
Soaring, vol. 6, 1980, pp. 1–21.
8 Koning, W. J. F., Romander, E. A., and Johnson, W.: Low Reynolds Number Airfoil
Evaluation for the Mars Helicopter Rotor, AHS International 74th Annual Forum &
Technology Display, Phoenix, AZ, 2018.
9 Hoerner, S. F.: Fluid-Dynamic Drag: Practical Information on Aerodynamic Drag and
Hydrodynamic Resistance, Hoerner Fluid Dynamics, 1965.
10 Hoerner, S. F., and Borst, H. V.: Fluid-Dynamic Lift: Practical Information on
Aerodynamic and Hydrodynamic Lift, Hoerner Fluid Dynamics, 1985.
11 Lissaman, P. B. S.: Low-Reynolds-Number Airfoils, Annual Review of Fluid Mechanics,
vol. 15, 1983, pp. 223–239.
12 Schmitz, F. W.: Aerodynamics of the Model Airplane. Part 1 - Airfoil Measurements,
Huntsville, AL, 1967.
13 Van Ingen, J.: The eN Method for Transition Prediction. Historical Review of Work at TU
Delft, 38th Fluid Dynamics Conference and Exhibit, 2008, Seattle, WA, p. 3830.
14 Laitone, E. V.: Wind Tunnel Tests of Wings at Reynolds Numbers Below 70 000,
Experiments in Fluids, vol. 23, Nov. 1997, pp. 405–409.
15 Carmichael, B. H.: Low Reynolds Number Airfoil Survey, Volume 1, Capistrano Beach, CA,
1981.

7
16 Koning, W. J. F., Johnson, W., and Allan, B. G.: Generation of Mars Helicopter Rotor
Model for Comprehensive Analyses, AHS Aeromechanics Design for Transformative
Vertical Flight, San Francisco, CA, 2018.
17 McMasters, J. H., and Henderson, M. L.: Low-Speed Single-Element Airfoil Synthesis,
Third International Symposium on the Science and Technology of Low Speed and Motorless
Flight, Seattle, WA, 1979, pp. 19, 23.
18 McArthur, J.: Aerodynamics of Wings at Low Reynolds Numbers: Boundary Layer
Separation and Reattachment, University of Southern California, Los Angeles, CA, 2008.
19 Winslow, J., Otsuka, H., Govindarajan, B., and Chopra, I.: Basic Understanding of Airfoil
Characteristics at Low Reynolds Numbers (104–105), Journal of Aircraft, Dec. 2017,
pp. 1–12.
20 Wazzan, A. R., Okamura, T. T., and Smith, A. M. O.: Spatial and Temporal Stability Charts
for the Falkner-Skan Boundary-Layer Profiles, 1968.
21 Okamoto, M., Yasuda, K., and Azuma, A.: Aerodynamic Characteristics of the Wings and
Body of a Dragonfly, The Journal of Experimental Biology, vol. 199, Feb. 1996, p. 281 LP-
294.
22 Levy, D.-E., and Seifert, A.: Simplified Dragonfly Airfoil Aerodynamics at Reynolds
Numbers Below 8000, Physics of Fluids, vol. 21, 2009, p. 71901.
23 Kesel, A. B.: Aerodynamic Characteristics of Dragonfly Wing Sections Compared With
Technical Aerofoils, Journal of Experimental Biology, vol. 203, 2000, pp. 3125–3135.
24 Mueller, T. J., and DeLaurier, J. D.: Aerodynamics of Small Vehicles, Annual Review of
Fluid Mechanics, vol. 35, Jan. 2003, pp. 89–111.
25 Young, L. A., Aiken, E., Lee, P., and Briggs, G.: Mars Rotorcraft: Possibilities, Limitations,
and Implications for Human/Robotic Exploration, IEEE Aerospace Conference, Big Sky,
MT, 2005.
26 Munday, P. M., Taira, K., Suwa, T., Numata, D., and Asai, K.: Nonlinear Lift on a
Triangular Airfoil in Low-Reynolds-Number Compressible Flow, Journal of Aircraft, vol.
52, 2014, pp. 924–931.
27 Koning, W. J. F., Johnson, W., and Grip, H. F.: Improved Mars Helicopter Aerodynamic
Rotor Model for Comprehensive Analyses, 44th European Rotorcraft Forum, Delft, the
Netherlands, 2018.

You might also like