Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

High-Downforce Airfoil Design for Motorsports

Author(s): P.S. Sriram, Ashok Gopalarathnam and Andrew Misenheimer


Source: SAE International Journal of Materials and Manufacturing , Vol. 5, No. 2 (June
2012), pp. 478-489
Published by: SAE International
Stable URL: https://www.jstor.org/stable/10.2307/26268482

REFERENCES
Linked references are available on JSTOR for this article:
https://www.jstor.org/stable/10.2307/26268482?seq=1&cid=pdf-
reference#references_tab_contents
You may need to log in to JSTOR to access the linked references.

JSTOR is a not-for-profit service that helps scholars, researchers, and students discover, use, and build upon a wide
range of content in a trusted digital archive. We use information technology and tools to increase productivity and
facilitate new forms of scholarship. For more information about JSTOR, please contact support@jstor.org.

Your use of the JSTOR archive indicates your acceptance of the Terms & Conditions of Use, available at
https://about.jstor.org/terms

SAE International is collaborating with JSTOR to digitize, preserve and extend access to SAE
International Journal of Materials and Manufacturing

This content downloaded from


61.68.151.141 on Sun, 23 Apr 2023 09:16:43 UTC
All use subject to https://about.jstor.org/terms
2012-01-1168
Published 04/16/2012
Copyright © 2012 SAE International
doi:10.4271/2012-01-1168
saematman.saejournals.org

High-Downforce Airfoil Design for Motorsports


P.S. Sriram, Ashok Gopalarathnam and Andrew Misenheimer
North Carolina State Univ

ABSTRACT
Using a combination of inverse airfoil design techniques, rapid interactive analysis methods, detailed computational
fluid dynamics (CFD) and wind tunnel testing, aft loading of an airfoil has been explored as a design direction for high-
downforce airfoils for race car rear wing applications while ensuring performance sustainability across a wide angle-of-
attack operating range. Unlike in aircraft oriented high-lift airfoil designs, pitching moment constraints can be
circumvented for race vehicle wing designs and this allows for further design freedom in the quest for downforce. The
PROFOIL inverse design code was used to design a candidate airfoil exhibiting downforce maximized using aft loading at
low Reynolds numbers. The resulting airfoil has a maximum lift coefficient of 2.5 at a designed Reynolds number of
300,000 and shows that aft loading on an airfoil is conducive to high-downforce requirements and is a favorable design
direction when considering airfoils for race car wing applications. Comparisons have been made with airfoils
representative of the high-lift design philosophies of Liebeck, Wortmann and Selig.

CITATION: Sriram, P., Gopalarathnam, A. and Misenheimer, A., "High-Downforce Airfoil Design for Motorsports," SAE
Int. J. Mater. Manf. 5(2):2012, doi:10.4271/2012-01-1168.
____________________________________

INTRODUCTION data at the design stage [1]. But as has been highlighted by
Agathangelou and Gascoyne [4], the front wing flow is
Downforce in motorsports has been one of the key complicated by ground effect (as a result of the close
parameters determining race vehicle performance envelopes proximity to the ground) and the close presence of the front
for over four decades now. Along with power, weight, and wheels. The rear wing, though, sees relatively ‘clean’ flow as
tires, it is one among the four most important parameters for it is mounted higher than the bodywork elements in order to
which open wheel race cars such as Formula 1 cars are gain access to free stream velocities [4]. Design of the rear
optimized [1]. Since the ground effect era of the seventies, wing and airfoils can be explored using existing aerodynamic
Formula 1 and other open-wheel race car designs have been theories.
dictated by the preferred aerodynamic layout and are The focus of this paper is to present aft loading as a
designed to work best with the wings and other elements of design direction for the design of airfoils intended for use on
the aerodynamic package [2]. The use of the Ford Cosworth rear wings or wings with reduced influence due to external
DFV eight cylinder engines by some teams in the seventies as flow field structures. A candidate high-downforce airfoil has
opposed to the considerably more powerful twelve cylinder, been designed to highlight the design methodology and
horizontally opposed engines (notably, Ferrari) is a case in underscore the downforce gain obtainable for such a design
point. The massive aerodynamic downforce benefits available direction.
from ground effect as a result of the inverted airfoil shape of The first section explains the design direction provided by
the vehicle underbody were being explored by the the aft loading and compares the prominent high-lift airfoil
aerodynamicists. Ultimately, even with a less powerful design philosophies and their respective merits and demerits
engine, the aerodynamic downforce resulted in a car that was when it comes to motorsports applications. The second
superior in vehicle dynamics and track performance [1]. section deals with the implementation of the aft loading
While the different components of an aerodynamic design philosophy applicable to high-downforce requirements
package contribute varyingly to the downforce levels and relevant to motorsports and the design and analysis methods
resulting flow fields, only the front and rear airfoils and used in this effort. Examples are included to illustrate the use
wings lend themselves to theoretical aerodynamic analysis of the multi-point inverse design method (PROFOIL) [15], to
methods and techniques for design. Other components and generate candidate airfoil shapes which were then analyzed
body shape designs still rely on experimental and numerical using the XFOIL (single element) [5] and MSES (multi

478

This content downloaded from


61.68.151.141 on Sun, 23 Apr 2023 09:16:43 UTC
All use subject to https://about.jstor.org/terms
Sriram et al / SAE Int. J. Mater. Manf. / Volume 5, Issue 2(June 2012) 479

element) [6] codes to provide viscous predictions quickly and


efficiently and thus serve as feedback to the designer to
further refine the performance of the airfoil under
consideration. These codes allowed for rapid analysis of the
airfoils at several angles of attack, Reynolds numbers, and for
several flap configurations.
The next section deals with the results are from wind
tunnel testing and CFD simulations, which were used to
corroborate the results of the optimized airfoil shape. Surface
pressure distribution, force and moment data, and oil-flow
visualization photographs from wind tunnel tests conducted
in the NCSU subsonic wind tunnel provide comparisons with
XFOIL/MSES and the CFD predictions.
This study will be presented using a high-lift single
element airfoil configuration developed for effectiveness
even at low speeds. The design and testing was therefore
performed at low Reynolds numbers ranging from 300,000 to
600,000 to provide a realistic estimate of the feasible
aerodynamic gains even at these cornering speeds.
Computational results at higher Reynolds numbers are also Figure 1. Low Reynolds number airfoil characteristics as
considered to establish the validity of the design direction a function of pitching moment and stall type.
with respect to its applicability to higher rungs of open wheel
racing and the consequent higher speeds.
From a historical perspective, two distinct methodologies
HIGH DOWNFORCE DESIGN were developed in the quest for high lift. Several Liebeck
PHILOSOPHY airfoils are good examples of the first type where a large
rooftop suction level is employed followed by a Stratford
For a motorsports airfoil, the primary requirement is a pressure recovery (or concave pressure recovery) [12]. This
high maximum lift coefficient [7]. After this requirement is leads to hard stall characteristics and high lift with low
satisfied, various other criteria can be considered in the pitching moment. The second approach is that used by the
design to ensure proper functioning of the high lift system some of the Wortmann airfoils where the reliance on a
under various operating conditions. We will consider some of suction peak is reduced and more emphasis is placed on aft
these scenarios and examine some of the existing low loading (convex pressure recovery) in order to provide softer
Reynolds Number (LRN) high lift airfoil designs that have stall characteristics [10]. A third ‘middle ground’
been developed before for various aeronautical applications methodology is reflected by the Selig and Eppler high lift
such as low speed aircraft [8]. The distinct design airfoils where a combination of the aforementioned design
philosophies in this regime include the approaches taken by philosophies are utilized in combination to provide high lift at
Liebeck [7], Eppler [11], Wortmann [10] and Selig [8]. To LRN[8].
study the applicability of aft loading to motorsport The Liebeck airfoils rely on a Stratford boundary-layer
applications, it is necessary to understand the inverse solution whereby a pressure recovery distribution can
interdependence of various airfoil characteristics upon one be found that continuously avoids separation of the turbulent
another. It is well known[8] that as pitching moment boundary layer. It is meant to recover the maximum possible
increases, maximum lift coefficient increases along with the pressure rise in the shortest possible distance. A high rooftop
pressure recovery becoming convex, as is represented in Cp value can be specified with the desired roof top length,
Figure 1. Other observable trends from the same figure
which can then be recovered using an inverse solution that
indicate that as an airfoil tends towards a more concave
gives the Stratford distribution for that particular case of
loading, high lift is achievable along with an increase in the
rooftop dimensions [7]. This is an approach that has worked
rapidity with which stall is reached(‘fast stall’ [8]).
well for the applications it has been designed for and provides
a high lift value with low pitching moment coefficients. An
example of this type of pressure recovery is shown in Figure
2a using a Liebeck LNV-109 airfoil. The Stratford recovery
also represents the optimum distribution for low profile drag
[11] and this leads to some of the highest lift to drag ratios for
these classes of airfoils [12]. But this makes the boundary
layer on the upper surface very sensitive to surface
imperfections that may trip the flow. Bragg et al. [19] have

This content downloaded from


61.68.151.141 on Sun, 23 Apon Thu, 01 Jan 1976 12:34:56 UTC
All use subject to https://about.jstor.org/terms
480 Sriram et al / SAE Int. J. Mater. Manf. / Volume 5, Issue 2(June 2012)

studied the effect of premature flow transition on airfoils


designed with upper surface pressure distributions having
laminar rooftops and Stratford-type recovery. They
demonstrate the drastic performance drop due to the effects
of rain drops close to the leading edge. Motorsport
applications often have wings positioned close to the ground,
and this makes their surfaces susceptible to various bits of
track and tire debris. These particles can potentially act as
trips and, on airfoils with a Stratford-type pressure recovery,
can cause loss of lift. Depending on the Reynolds number, the
trips may sometimes act beneficially and prevent the
separation bubbles. But this induces an inherent uncertainty
when the aerodynamics data is to be used in performance
prediction suites such as lap simulations and other vehicle
dynamics simulations which rely on aerodynamic
performance data for a wide range of simulated operating
conditions. Figure 3 shows an XFOIL prediction of how a Figure 3. XFOIL prediction for Liebeck LNV109a airfoil
high lift Liebeck LNV-109 airfoil [20] reacts to the flow performance at Re=300,000 with free transition and

tripped at with a large drop in Clmax. High transition fixed at .


performance airfoils reliant on carefully controlled adverse
pressure gradients thus show a rapid deterioration in The range could be larger in either direction depending on
performance outside a narrow envelope [12]. the motorsports series in consideration). So these variations
Stratford recovery also results in the airfoil exhibiting can cause an increase in adverse pressure gradient which then
hard-stall which is characterized by the coefficient of lift causes a fast moving turbulent separation point. Eppler
decreasing rapidly with increasing angle of attack beyond suggested that concave pressure recoveries should be used
stall. Eppler [11] argued that the sensitivity of the turbulent but they should not be as steep as the Stratford distribution at
boundary layer in a Stratford distribution, which is on the the beginning. This forms the basis for Eppler's and Selig's
verge of separation by design, can be a cause of hard stall as high lift airfoil designs [11] where a moderated degree of
the unsteadily moving transition point can change the initial concavity is allowed into the pressure recovery along with aft
conditions of the pressure recovery such that the turbulent loading.
separation is also unsteady. A race car often sees a large Even though a motorsports wing does not see large
variation in speed across a race track which can change the changes in angle of attack during forward motion, it is
operating Reynolds number from 200,000 to 600,000 (for necessary to have as wide an operating range as possible in
Formula SAE. order to give the aerodynamicist and the vehicle dynamicist
enough options when it comes to car setup. The rear wing is
often used to balance the car (after the front wing setup has
been completed) to compensate for possible undesired
characteristics of the car endowed to it by pre-existing
handling traits [4]. In the work done by McKay and
Gopalarathnam [13], an airfoil's effect on overall lap times
was computed for the entire range of its lift curve while
accounting for wing aerodynamic considerations. The airfoils
under consideration in that study exhibited a hard stall
characteristic and based on the results of their study, it is
evident that lap times deteriorated post stall. Despite profile
drag being large in the post-stall regimes, a soft stall can
extend the range of available performance at Cl max. So one of
the requirements is that a high downforce airfoil should
possess a soft stall and should sustain high Cl for a large
angle-of-attack range near stall to provide flexibility during
Figure 2. Pressure vectors computed from XFOIL for an car set up.
angle of attack of 5 ° to show airfoil loading. Due to the very low aspect ratio high downforce systems
that race cars employ, the primary source of drag comes from
the induced component of overall drag. Profile drag values
are thus much lower than the resulting overall drag values.

This content downloaded from


61.68.151.141 on Sun, 23 Apr 2023 09:16:43 UTC
All use subject to https://about.jstor.org/terms
Sriram et al / SAE Int. J. Mater. Manf. / Volume 5, Issue 2(June 2012) 481

This allows the design to focus on downforce as opposed to was for used rapid interactive design by specifying the
airfoil L/D. Therefore the chief concern in motorsports airfoil inviscid velocity distributions and analyzing the resulting
design is not one of profile drag reduction [7, 13]. Instead it is candidate airfoils in codes with viscous analysis capabilities
a maximization of downforce and the ability of the designed such as XFOIL and MSES [6, 5]. PROFOIL was used with a
airfoil to sustain the highest possible levels of downforce MATLAB-based graphical user interface (GUI) [16] which
across a wide range of physical and aerodynamic adversities. helped execute the various elements of the design code
Hence a highly concave pressure recovery employing a interactively and concurrently plot the resulting airfoil with
Stratford distribution is not the ideal solution for a its constraints and the specified velocity distributions. Many
motorsports airfoil design. Rather the goal is to maximize features of PROFOIL are well suited to designing for
downforce and retain high levels of performance across a motorsports and maximum downforce. For instance, it
broad range of operating conditions. permits control over the design of the transition ramp and this
Wortmann's approach with the FX-63-137 consisted of aft can be used to influence the characteristics of the laminar
loading with more gradual initial gradients. The design separation bubble. The turbulent boundary layer development
approach with this airfoil was to increase Cl max primarily by can also be prescribed to avoid separation by a certain design
adding pitching moment [8]. Wortmann argued that in the margin [15].
case of a concave pressure distribution, initial thickness The PROFOIL code consists of an inverse design
effects on the turbulent boundary layer are much stronger methodology with an integral boundary-layer method for
than for pressure rises with smaller initial gradients [10]. This rapid analysis at the design points. It allows the designer to
gives the FX63-137 a convex pressure distribution, as seen in divide both surfaces of the airfoil chord into segments along
Figure 2b, along with an increase in length of the each of which the velocity distributions can be prescribed as
representative pressure vectors on the lower surface at the aft functions [14]. A design angle of attack, α*, is specified for
portion of the airfoil, thus indicating aft loading. Eppler each of these segments to tailor the velocity distributions. For
showed that the lift of an airfoil with concave recovery could a segment, α* is the angle of attack relative to the zero lift, αz
be improved using aft loading and this was meant to espouse at which the segment has zero velocity gradient. So if the αz
both concave pressure recovery and aft loading as a means to of the airfoil is greater than the α* for a particular segment on
enhance high lift performance. An example of this design the upper surface, then that particular segment will
direction is the Wortmann FX74- CL5-140[Figure 2d], which experience an adverse pressure and vice versa. So increasing
is a high lift design that was tailored for high lift performance or decreasing α* can change which parts of the airfoil
at a higher Reynolds number than those considered here. It experience adverse gradients at various angles of attack.
uses a more gradual initial pressure recovery compared to The design process consisted of various candidate designs
Stratford recovery airfoils along with aft loading, as shown in being produced and compared against a backdrop of the
Figure 2d. Selig adapted concave recovery and aft loading to required parameters. Every instance where the candidate
produce airfoils optimized for high lift at LRN. The S1223, airfoil failed to meet the specified design goals, the
shown in Figure 2c, produces the maximum lift currently for experience gleaned from that particular iteration was useful in
airfoils operating in this regime. redesigning the airfoil to facilitate a convergence onto the
The Eppler, Wortmann and Selig approaches have so far desired performance specifications. This iterative process
been effective in generating airfoils with high Cl max values continued until a successful airfoil meeting the pre-set
for this regime. But due to their constraints born out of performance goals was generated. Despite the computational
adhering to aeronautical considerations, it is felt that an advances in optimization and inverse design, it is still not
approach more tailored to high downforce generation for possible to fully automate the airfoil design procedure and it
motorsports can yield higher Cl max values and satisfy still remains a sophisticated cut-and-try procedure that is
requirements such as performance sustainability across a reliant on the designer's judgment to provide the right
large range of angles-of-attack, soft stall characteristics and a direction [17].
relative insensitivity to adverse surface roughness effects on The α* values were individually manipulated and kept
the performance characteristics of the airfoil. This approach high over the upper surface to reduce the severity of the
eliminates any pitching moment constraints imposed in recovery gradient adversity in order to provide soft stall and
previous designs and attempts to use aft loading as the chief ensure that Cl max, or values close to it, were available over a
driver towards maximizing downforce while maintaining a large angle of attack range. The leading edge α* values were
rudimentary level of concave pressure recovery that has been set higher than 30° to ensure that even at high angles of
kept gradual to ensure airfoil stability under varying attack, the velocity gradient isn't very adverse, so as to reduce
operational conditions. the reliance on suction peak related performance and the
associated fast movement of the turbulent separation point at
DESIGN IMPLEMENTATION high angles of attack. Along with these α* manipulations,
The design implementation was done using the PROFOIL several constraints were used to achieve the desired airfoil
multi point inverse airfoil design code [14, 15]. PROFOIL characteristics. A thickness constraint was used to change

This content downloaded from


61.68.151.141 on Sun, 23 Apr 2023 09:16:43 UTC
All use subject to https://about.jstor.org/terms
482 Sriram et al / SAE Int. J. Mater. Manf. / Volume 5, Issue 2(June 2012)

upper surface α* values to maintain a 13% (of chord) RESULTS


thickness. A camber constraint was used to vary lower
surface α* to maintain camber at 13.81% and a constraint was The comparative analysis and computational validation
placed on the Cm c/4 (pitching moment at quarter chord) to was performed using XFOIL [6]. The XFOIL code solves the
panel method equations coupled to an integral boundary-layer
maintain it at −0.490. These two values were determined to
be adequately robust and prevent the inverse design routine formulation using a global Newton iteration scheme. The en
from producing very thin trailing edges, which are difficult to transition model used in XFOIL has been shown to be
manufacture, or physically unrealizable airfoils with crossed- reliable in predicting various airfoil related flow phenomena
over surfaces. Additionally, a thickness constraint was placed such as LSB formations and transition locations accurately.
on the trailing edge area to ensure sufficient thickness to ease XFOIL has also been validated against wind tunnel results for
fabrication. This generated the single-element MSHD (Motor other high lift airfoils, NLF airfoils and multiple flap
Sports High Downforce) airfoil, seen in Figure 4. Apart from configurations [8, 21]. However, it is also known to over
this, an additional trailing edge gap thickness was used to predict Cl and L/D at post stall α values. Since the study here
produce a blunt trailing edge. Since most motorsport is concerned partly with extending the available performance
governing bodies have a requirement for blunt trailing edges envelope before stall and maximizing downforce
or some form of finite radius trailing edge for safety reasons, performance before stall, it was decided that the potential
this aspect of the inverse design was considered important. post-stall inaccuracies inherent in XFOIL solutions can be
The inclusion of a blunt trailing edge required adjustments to ignored for the time being, especially since the onset of stall
some α* values in order to ensure that the change in the is predicted reasonably accurately by XFOIL. Post-stall over-
trailing edge geometry maintains high levels of downforce prediction aside, it was decided that XFOIL would be useful
and ensures adherence to the other desired characteristics. All to compare the performance of the MSHD airfoil with the
the α* manipulations and constraints were aimed at achieving other high lift airfoils in consideration, primarily due to the
the following design targets: ease of setting up and running multiple angle of attack sweep
• High extent of aft loading for increased downforce, cases readily and expediently.
• Soft stall characteristics by promoting longer turbulent
pressure recovery regions with reduced amounts of concavity,
BASE AIRFOIL PERFORMANCE
• Large leading edge radius in order to prevent formation of The polar plot for the MSHD airfoil is shown in Figure 5.
long LSBs even at large angles of attack and reduce suction As seen, the Cl max is 2.5 at an α of 20°. Beyond this, there is
peak dependence, a region of decreasing Cl right up to 25°. This is indicative of
• Manipulation of maximum thickness in order to begin a very soft stall and the Cl at 25° is still at 2.4. The airfoil has
adverse gradient further forward on the airfoil chord and a large range of high lift values beginning from α= 1° and
minimize dependence on laminar boundary-layer, Cl=1.5 to α= 25° and Cl=2.4. The pitching moment values are
• Trailing edge stall by utilizing leading edge radius and very high. This is a result of the various geometrical
camber, concessions for high downforce gain and the relaxation of the
• High levels of downforce even with trailing edge gaps, pitching moment constraint during inverse design. As this is
• Performance retention despite the presence of debris and not a factor for race car wing downforce, the fact that it is as
trips, and high as it is bears no consequence to the prospect of
• Similarity of performance and high downforce successful downforce generation. But is does help extract
characteristics across the span of target speeds and Reynolds large amounts of downforce from early in the angle of attack
numbers. range: Cl values cross 2 at a modest 4° angle of attack. One of
the targets during the design process was to instill a relative
behavioral insensitivity to changes in speed or changes in
Reynolds number. In other words, the airfoil needed to
exhibit the same characteristics and maintain similar
performance across a large low Reynolds number range. This
was essential from the vehicle dynamics point of view as the
stability of an aerodynamic set-up is very important when
various speed regimes are considered. Usually, the sensitivity
to the Reynolds number can be mitigated by extending the
instability range, i.e., extending the range of the turbulent
boundary layer [11]. This prevents increases in adverse
pressure gradients which cause fast moving transition points
that can change the initial conditions of the pressure recovery
Figure 4. Geometry of the MSHD airfoil. and result in an unsteady turbulent separation. Figure 6 shows
the lift curves for the MSHD airfoil for Reynolds numbers of

This content downloaded from


61.68.151.141 on Sun, 23 Apr 2023 09:16:43 UTC
All use subject to https://about.jstor.org/terms
Sriram et al / SAE Int. J. Mater. Manf. / Volume 5, Issue 2(June 2012) 483

Figure 5. Performance polar for the MSHD at Re=300,000 computed using XFOIL.

300,000 and 600,000, showing a high level of insensitivity to larger angle of attack range than the other two airfoils. Cl max
Reynolds number. is considerably higher and occurs at a much higher angle of
attack (20° as compared to roughly 12° for the other two
airfoils in question). This can give a lot more potential for
adjustability and provides a large range of high downforce
values for the aerodynamicists and the vehicle dynamicists to
use. Even at α = 0°, it is seen that the Cl ≈1.5 and is
considerably higher than those for the other two airfoils. High
downforce is available even beyond Cl max and the airfoil
stalls very softly compared to the other two in consideration,
which have also been designed to have soft stall. So a large
range of angles of attack with high downforce are available
up to α = 25°, whereas it is seen that the FX74-CL5-140 and
S1223 stall before α = 15°. For angles of attack less than 0,
there is a sudden drop in the values predicted by XFOIL for
the MSHD airfoil. This maybe a result of the highly separated
flow that the airfoil maybe encountering at negative angles of
attack due to the large concavity in the lower surface
geometry. This may be indicative of the fact that the MSHD
Figure 6. Performance comparison from XFOIL airfoil experiences a ‘hard’ negative α stall. Figure 8 shows
predictions at varying Reynolds numbers. the Cp distributions for the three airfoils at an angle of attack
of 10°. It is seen that the S1223 has the largest suction while
the FX74-CL5- 140 has the least suction. The MSHD has a
PERFORMANCE COMPARISON suction peak that is in the middle of both these values. It
The performance of the MSHD airfoil has been compared shows hardly any concavity in the recovery when compared
here with the S1223 and the FX74-CL5-140. Figure 7 shows to the recovery on the S1223. Also seen is the extent of aft
the lift curve for the MSHD, S1223 and the FX74-CL5-140 loading which is the highest on the MSHD. The middle and
airfoils at a Reynolds number of 300,000. It is seen that the aft portion on the MSHD show more pronounced loading
Wortmann FX74-CL5-140 and the Selig S1223 show very than that of the other two airfoils and is reflective of the
similar behavior. They have both been designed on similar aftloading employed in this exercise.
principles, although historically, the Wortmann airfoil was
designed for a higher Reynolds number close to 1,000,000
and the S1223 was designed for an operating Reynolds
number range very similar to the MSHD's design conditions:
between 200,000 to about 800,000. Also noticeable from the
same figure, is the fact that the overall downforce
performance of the MSHD airfoil is sustained across a much

This content downloaded from


61.68.151.141 on Sun, 23 Apr 2023 09:16:43 UTC
All use subject to https://about.jstor.org/terms
484 Sriram et al / SAE Int. J. Mater. Manf. / Volume 5, Issue 2(June 2012)

varied up to a maximum speed of approximately 40 m/s at a


dynamic pressure of 720 Pa. This maximum tunnel speed
results in a maximum chord Reynolds number of
approximately 0.8 million for a 10-inch chord airfoil model.
It is recognized that faster speeds may be required in order to
evaluate the airfoil for other racing series but for the purpose
of testing for Formula SAE applications, the existing tunnel
setup is well suited.
Clean Airfoil Results
Wind tunnel testing was conducted for clean-airfoil cases
and for cases where the airfoil was tripped at various
locations. The airfoil used in wind tunnel testing was the
MSHD airfoil with a 0.5% blunt trailing edge. This was done
Figure 7. Cl vs.α (in degrees) curve comparison from in order to provide flexibility during fabrication and also
XFOIL prediction at Re=300,000. since most motorsports governing bodies require blunt
trailing edges.
The following paragraphs will review the results of the
clean wind tunnel test in comparison with the XFOIL
predictions for the same. The ncrit parameter in the en model
used in XFOIL can be changed to replicate various flow
conditions. The ncrit parameter is the critical value for n,
which is the log of the amplification factor of the most
amplified frequency which triggers transition. In other words,
it changes the transition characteristics based on a model of
the ambient disturbance level in which the airfoil operates, as
suggested by the user. In order to accurately replicate the
wind tunnel conditions, an ncrit of 4 was used, which signifies
a small, turbulent wind tunnel setup [22]. Lift data was
obtained from surface pressure measurements taken from
pressure taps placed along the upper and lower surface of the
airfoil model. Drag measurements were not obtained in this
effort. Also, as was pointed out earlier, drag reduction for
race car airfoils is less important than the downforce. Owing
to the fact that drag of the highly-loaded, small-span wings
Figure 8. Comparison of pressure profiles.
used in motorsports is dominated by induced drag, measuring
drag values would not have added significant insight into the
design methodology for a high downforce airfoil.
WIND TUNNEL RESULTS The results are shown in Figures9 and 10 for Reynolds
NCSU Subsonic Wind Tunnel numbers of 300,000 and 400,000. As is clear from the
The NCSU subsonic wind tunnel is a closed-circuit tunnel figures, there is a good match between the predicted and
with a 0.81 m high, 1.14 m wide, and 1.17 m long test experimental values. The Cl max values correspond very
section. Upstream of the test section and forward of the closely between the two. The main difference occurs in the
contraction section is a settling chamber consisting of an region of the soft stall where the wind tunnel results are
aluminum honeycomb screen followed by two stainless steel slightly lower than the predicted values. The trends still show
anti-turbulence screens. Turbulence levels have been that there is an adherence to the required soft stall
determined to be less than 0.33% [18].The contraction section characteristic and Cl values in the region are still above 2.
is composed of four sides of identical curvature and the test There is a slight over-prediction from XFOIL in this region,
section walls diverge slightly to allow for boundary layer as is known to happen. The predicted values and the
growth. The two vertical sides of the test section are made of experimental results are in excellent agreement over the
Plexiglas hinged at the top for easy access and visibility. In entire positive range of the airfoil operation. The negative
order to ventilate the tunnel to room pressure, a breather is range, on the other hand, shows extremely large differences.
located downstream of the test section. This is due to the inadequacy of XFOIL in accurately
The tunnel fan is equipped with variable pitch blades predicting largely separated and highly vortical flows. The
allowing the velocity in the test section to be continuously wind-tunnel results for the two Reynolds numbers, compared

This content downloaded from


61.68.151.141 on Sun, 23 Apr 2023 09:16:43 UTC
All use subject to https://about.jstor.org/terms
Sriram et al / SAE Int. J. Mater. Manf. / Volume 5, Issue 2(June 2012) 485

in Figure 11, show that the earlier stated design goal of


having similar performance at differing speeds has been
achieved to some extent. Most values between the two speeds
show excellent correlation, including the negative angles of
attack. For a large positive angle-of-attack range, the values
of Cl for the Re=300,000 case are marginally higher than for
the Re=400,000 case. This is an effect of the ‘fast stalling’
trailing edge and has been found to occur on airfoils with a
reasonably high aft loading [9].

Figure 11. Comparison of performance for the MSHD


airfoil in the wind tunnel at Re=300,000 and
Re=400,000.

Tripped Airfoil Results


In order to verify that the downforce from the MSHD
airfoil is not significantly affected when tripped, the airfoil
was tested with a trip placed at one of the three locations on
the upper surface: 0.1c, 0.2c and 0.3c. The trip consisted of
Figure 9. Cl vs. α comparison between wind tunnel and
two layers of tape pasted on top of each other to give a
XFOIL calculated results for the MSHD airfoil at cumulative height of roughly 1 mm. This sizing was found to
Re=300,000. be effective after it was tested using flow visualization (as
will be seen in the next subsection) and it was apparent that
the boundary layer was getting tripped to a turbulent state and
preventing the formation of a laminar separation bubble. The
length of the trip was such that it spanned the entire length of
the wing model. An example arrangement of a trip placed at
0.1c is shown in Figure 12.

Figure 10. Cl vs.α comparison between wind tunnel and


XFOIL calculated results for the MSHD airfoil at
Re=400,000.

Figure 12. Airfoil model in the wind tunnel, with

boundary layer trip (seen as a yellow stripe) at .

This content downloaded from


61.68.151.141 on Sun, 23 Apr 2023 09:16:43 UTC
All use subject to https://about.jstor.org/terms
486 Sriram et al / SAE Int. J. Mater. Manf. / Volume 5, Issue 2(June 2012)

Since tripping the flow closer to the leading edge


produces more adverse effects than flow tripped further
downstream, only results from the 0.1c trip case will be
considered here to exhibit the MSHD high downforce
characteristic despite the presence of trips. This is due to the
fact that the case shown here (0.1c trip) exhibits the desired
characteristics and the other cases only adhere further to the
clean airfoil characteristics.
The wind tunnel results for the tripped airfoil are
compared with the XFOIL predictions with fixed transition in
Figure 13. The comparisons with XFOIL show that the values
predicted by XFOIL for the initial 10° positive angles of
attack are lower than that obtained in the wind tunnel. But
from there on, the values from both remain close to each
other up to 25°. When compared to the clean airfoil run
(Figure 14), it can be seen that the negative angles of attack Figure 14. Wind tunnel comparison between clean
yield the same results for both the clean and tripped cases. airfoil and airfoil tripped at 0.1c; Re=300,000.
But from 0° on to about 22°, the clean airfoil shows higher Cl
values consistently. This is because of the early transition that
Flow Visualization
is forced onto the airfoil by the presence of the trip. After this
point, separation starts to affect the performance of the clean Oil flow visualization was conducted in order to obtain an
airfoil and there is a dip in Cl values below that of the tripped insight into the flow behavior under various operating
conditions. For these tests, the airfoil was positioned at the
case. The tripped case shows a sustained increase in Cl due to
desired angle of attack and an oil mixture (SAE 20W motor
the early initiation of the turbulent boundary layer, thus oil mixed with white titanium dioxide power) was applied on
leading to increased resistance to stalled flow. It was its surface before the wind tunnel was switched on. The wind
observed from all three cases that at high angles of attack, the tunnel was allowed to run for a few minutes to let the flow
airfoil constantly produces a Cl between 2 and 2.2 at low patterns develop and influence the flow of the oil mixture on
Reynolds number conditions. It is clear from these results that the surface [23].
airfoil still maintains its high downforce despite the presence The flow visualization exercise can be used to confirm the
of trips. This is a useful characteristic in a wing that operates presence of short laminar separation bubbles (LSBs) because,
in proximity to the ground and is likely to get leading-edge as has been mentioned earlier, this is one of the aspects of the
contamination during operation. design methodology required for achieving high downforce
across a broad range of operating conditions. Pictures were
taken (some of which are shown in to study the flow
structures for angles of attack ranging from 0° to 25° in
intervals of 5° in order to corroborate trends. Figure 15 shows
the upper-surface oil flow patterns for angles of attack of 0°
and 10°. The oil accumulation line indicating the start of the
separation bubble was seen to be moving closer to the leading
edge with increasing angle of attack, thus indicating that the
bubble is moving forward along the airfoil with increasing
angle of attack. Another facet of the airfoil's characteristics
that can be seen from the photographs is the fact that the
bubble shows a reduction in chordwise size with increasing
angle of attack (Figure 15, for α = 0° and α=10°). The
chordwise extent of the bubble reduces from 0.1c at α=0° to
0.07c at α=25°. This gradual reduction associated with the
Figure 13. Comparison of performance with XFOIL with presence of the bubble even at high angles of attack indicates
airfoil tripped at 0.1c; Re=300,000. a short bubble [11].This can be concluded from the
observations above because a large bubble which affects the
shape of the boundary layer would result in separation behind
the bubble at high angles of attack [11]. In the photographs
for high angles of attack, however, separation bubbles are
present with reattachment regions clearly visible. This is
exhibited clearly in Figure 15 in the picture showing the

This content downloaded from


61.68.151.141 on Sun, 23 Apr 2023 09:16:43 UTC
All use subject to https://about.jstor.org/terms
Sriram et al / SAE Int. J. Mater. Manf. / Volume 5, Issue 2(June 2012) 487

airfoil at 20° (both clean and with a 0.3c trip), where an LSB Gauss-Seidel iterative matrix solution scheme was used.
has formed and the flow has reattached before the trip, thus Spatial and temporal accuracy were both set to second order
confirming that the bubble is a short one. to improve accuracy. No turbulent wall functions were used
and this necessitated the finely resolved y+ mentioned earlier
in this section. Other turbulence models that were tried
include the D.M.S.A and the k - ωmodels, both of which are
two equation turbulence models. Convergence was, however,
not as good (even with preconditioning) as with the Spallart-
Allmaras model. Thus it was decided that the Spallart-
Allmaras model would be used for all the computations. The
velocity and turbulence level was setup to replicate the wind
tunnel test conditions.

Figure 15. Flow visualization for the MSHD airfoil.

CFD RESULTS Figure 16. Complete grid for the MSHD airfoil CFD
The grid is an unstructured grid with 124, 354 elements. computations.
In order to eliminate any wall interference effects, the airfoil
is placed in the center with all four walls 20 chord lengths
away from the airfoil. This is shown in Figure 16, where the
airfoil is visible as a small dark point in the middle of the
grid. Further details of the grid around the airfoil can be seen
in Figure 17. For the prismatic grid elements, 15 inflation
layers were provided on the airfoil surfaces with a 1:1 grid
aspect ratio on the final elements of the inflation layer. The
inflation layers can be seen around the airfoil in Figure 18,
seen as a thick line around the airfoil. A detailed picture of
the inflation layer structure is shown in Figure 18 where an
aspect of ratio of 1:1 can be seen for the final layer. The
minimum element size of the prismatic layers was 1e−5 m
and was derived using a y+ of 1. These layers are essential for Figure 17. Regions of localized mesh density around
effective boundary layer resolution and ensuring solution airfoil and wake.
accuracy. Open boundary conditions were applied to the top
and bottom walls to allow free-stream conditions to prevail
unhindered. Velocity inlet and pressure outlet boundary
conditions were imposed on the front and back walls
spanning the element thickness. The walls adjoining the
airfoil on either side were assigned symmetry boundary
conditions. The entire grid generation was done using the
ICEM CFD package at Corvid Technologies. The simulation
was run using the Raven CFD solver, developed in-house at
Corvid Technologies. The simulations were run using the
Spallart-Allmaras one equation turbulence model. Due to
very low Mach numbers (0.04 M to 0.1 M) and the highly
vortical flows encountered, the algorithm needed pre-
conditioning in order to improve convergence. Temporal
damping was also employed to improve convergence. The

This content downloaded from


61.68.151.141 on Sun, 23 Apr 2023 09:16:43 UTC
All use subject to https://about.jstor.org/terms
488 Sriram et al / SAE Int. J. Mater. Manf. / Volume 5, Issue 2(June 2012)

Figure 18. Prismatic inflation layer elements around


airfoil surface.
Figure 19. Comparison between wind tunnel and CFD
results for the MSHD airfoil; Re=300,000.
As seen in Figure 19, CFD predictions match that of the
wind tunnel in the positive angle of attack range up to α=15°.
Beyond this, the CFD predictions are higher than that of the
wind tunnel and this may be due to highly unsteady separated
flow at higher angles of attack. This is best seen at angles of
attack beyond 20°. For negative angles of attack, CFD
predictions are similar to those seen in XFOIL, where the
highly cambered lower surface of the airfoil is inducing
vortical flow and may be preventing an accurate
computational result. The general characteristics of the airfoil
are evident from the plot and the computations reflect the
expected behavior of the airfoil where it shows no noticeable
stall in the positive angle of attack range (very soft stall) and
shows a distinct stall region in the negative angle of attack
regime. Figures 20 and 21 illustrate the flow physics on the
airfoil and help providing further confirmation that trailing
edge separation is prevalent. Even at zero degrees angle of
attack, the airfoil shows slight separation on the underside Figure 20. Eddy viscosity plots showing vortices for the
and minor separation near the trailing edge. This was also MSHD airfoil at different angles of attack; Re=300,000.
seen in flow visualization. For the higher angle of attack
flows, large regions of separated flow can be seen can be seen
over the aft portions of the airfoil with the forward portions
near the leading edge showing attached flow. It is this
phenomenon that allows a very soft stall behavior and
enables the generation of high downforce deep into the
positive angle of attack range, well beyond the stall angle of
other high lift airfoils. The computations are also useful in
explaining the behavior of the airfoil for negative angles of
attack. As can be seen for the plots at zero degrees, there are
small regions of separation on the lower surface, which is
exacerbated at lower negative angles of attack, thus leading to
an abrupt stall in the negative angle of attack regime.

Figure 21. Velocity magnitude plots showing extents of


separation for the MSHD airfoil at different angles of
attack; Re=300,000.

This content downloaded from


61.68.151.141 on Sun, 23 Apr 2023 09:16:43 UTC
All use subject to https://about.jstor.org/terms
Sriram et al / SAE Int. J. Mater. Manf. / Volume 5, Issue 2(June 2012) 489

7. Liebeck, R., “Subsonic Airfoil Design,” in Applied Computational


CONCLUSION Aerodynamics, Henne, P. A. (ed.), Vol.125, Progress in Astronautics
and Aeronautics, American Institute of Aeronautics and Astronautics,
The aim of this research was to present a high downforce Washington, DC, 1990, pp.501-530.
design philosophy for race car rear wings. The motivation for 8. Selig, M.S., Guglielmo, J.J., “High-Lift Low Reynolds Number Airfoil
Design,” Journal of Aircraft, Vol.34, No.1, 1997, pp.72-79.
this research was the fact that race car rear wings are one of 9. Wortmann, F.X., “A Critical Review of the Physical Aspects of Airfoil
the few aerodynamic components that can be designed and Design at Low Mach Numbers,” Motorless Flight Research, edited by
Nash-Weber, J.L., NASA CR-2315, Nov. 1973, pp.179-196.
optimized for high downforce using theoretical aerodynamic 10. Wortmann, F.X., “The Quest for High Lift,” Proceedings of the
methods such as inverse design. Front wings and other AIAA/MIT/SSA Second International Symposium of the Technology and
Science of Low-Speed and Motorless Flight, Soaring Society of
aerodynamic components cannot be designed and optimized America, Los Angeles, CA, September 1974.
purely for high downforce due to the flow constraints on their 11. Eppler, R., “Turbulent Airfoils for General Aviation,” Journal of
Aircraft, Vol. 15, No.2, 1978, pp. 93-99.
wake developments. This also leads to the preclusion of their 12. Gad-El-Hak, M., “Control of Low-Speed Airfoil Aerodynamics,” AIAA
design using standard aerodynamic methods and requires Journal, Vol. 28, No.9, 1990, pp. 1537-1552.
13. Mckay*, N. and Gopalarathnam†, A., “The Effects of Wing
extensive numerical solutions from the beginning of the Aerodynamics on Race Vehicle Performance,” SAE Technical Paper
design phase. Using the PROFOIL inverse design tool, the 2002-01-3294, 2002, doi:10.4271/2002-01-3294.
14. Selig, M.S. and Maughmer, M.D., “A Multi-Point Inverse Airfoil
MSHD airfoil was developed to highlight the results of the Design Method Based on Conformal Mapping,” AIAA Journal, Vol. 30,
proposed design philosophy and its motorsports-specific No.5,1992, pp. 1162-1170.
15. Selig, M.S. and Maughmer, M.D., “Generalized Multi-Point Inverse
merits when compared to existing high lift design Airfoil Design,” AIAA Journal, Vol. 30, No.11,1992,pp. 2618-2625.
methodologies. Since other high lift design methodologies 16. Gopalarathnam, A. and Selig, M.S., “Design of High-Lift Airfoils for
Low Aspect Ratio Wings with Endplates,” Paper No. 97- 2232, 15th
were proposed for aeronautical applications, the current AIAA Applied Aerodynamics Conference, Atlanta, GA, 1997.
approach sought to tailor airfoil design to achieve the 17. Drela, M., “Elements of Airfoil Design Methodology,” in Applied
Computational Aerodynamics, Henne, P. A. (ed.), Vol.125, Progress in
requisite characteristics for motorsports by eliminating the Astronautics and Aeronautics, American Institute of Aeronautics and
constraints required for aeronautical designs. For instance, Astronautics, Washington, DC, 1990, pp.501-530.
18. Gittner, N.M., An investigation of the Effects of Aft Blowing on a 3.0
the pitching moment criterion was relaxed during inverse Caliber Tangent Ogive Body at High Angles of Attack, Master's thesis,
design to permit large amounts of aft loading, in an effort to North Carolina State University, 1992.
19. Bragg, M.B., Broeren, A.P., Blumenthal, L., “Iced-Airfoil
increase downforce and enable the airfoil to have a very soft Aerodynamics,” Progress in Aerospace Sciences, 41 (2005) pp.
stall. The design philosophy resulted in a single-element 323-362.
20. UIUC, “UIUC Airfoil Database,” http://www.ae.illinois.edu/m-
airfoil with a Cl max value of 2.5 at a Reynolds number of selig/ads/coord_database.html, UIUC Applied Aerodynamics Group,
300,000, with a large range of angles of attack with Cl> 2. 1995-2011.
21. Gopalarathnam, A., and Selig, M.S., “Low-Speed Natural-Laminar-
Wind tunnel testing and computational data were compared Flow Airfoils: Case Study in Inverse Airfoil Design,” Journal of
for the MSHD airfoil and were found to show a good match. Aircraft, Vol. 38, No. 1, January-February 2001, pp. 57-63.
22. Drela, M., “XFOIL User's Guide,” http://web.mit.edu/drela/Public/web/
Flow visualization was conducted on the airfoil in the wind xfoil/.
tunnel to study various airfoil characteristics and confirm 23. Barlow, J.B., Rae, H.R. and Pope, A., Low-Speed Wind Tunnel Testing,
John Wiley and Sons, Inc., New York, 1999.
adherence to the design goals. These collective testing
measures corroborate the efficacy of the design route and CONTACT INFORMATION
will, it is hoped, be useful for guiding future rear wing design
P.S.Sriram: Cell Ph. No.: 216-571-6054
across motorsport series and classes.
Work Ph. No.:704-799-6944 (x 141)
sriram042000@gmail.com
ACKNOWLEDGEMENTS
The authors wish to acknowledge the computational
expertise and assistance provided by Greg McGowan and
Corvid Technologies.

REFERENCES
1. Wright, P. and Matthews, T., “Formula 1 Technology,” Society of
Automotive Engineers, Inc., Warrendale, PA, ISBN 978-0-7680-0234-8,
2001.
2. Wright, P.G., “The influence of aerodynamics on the design of Formula
one racing cars,” International Journal of Vehicle Design, Vol.3 Issue 4,
1982, pp. 383-397.
3. Katz, J., Race Car Aerodynamics, Robert Bentley Publishers,
Cambridge, MA, 1995.
4. Agathangelou, B. and Gascoyne, M., “Aerodynamic Design
Considerations of a Formula 1 Racing Car,” SAE Technical Paper
980399, 1998, doi: 10.4271/980399.
5. Drela, M., “Design and Optimization Method for Multi-Element
Airfoils,” AIAA Paper 93-0969, Feb. 1993.
6. Drela, M., “XFOIL: An analysis and Design System for Low Reynolds
Number Airfoils,” Lecture Notes in Engineering: Low Reynolds Number
Aerodynamics, Mueller, T. J. (ed.), Vol.54, Springer-Verlag, New York,
June 1989.

This content downloaded from


61.68.151.141 on Sun, 23 Apr 2023 09:16:43 UTC
All use subject to https://about.jstor.org/terms

You might also like