Aerospace Science and Technology: James G. Coder, Dan M. Somers

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Aerospace Science and Technology 106 (2020) 106217

Contents lists available at ScienceDirect

Aerospace Science and Technology


www.elsevier.com/locate/aescte

Design of a slotted, natural-laminar-flow airfoil for commercial


transport applications
James G. Coder a,∗,1 , Dan M. Somers b
a
The University of Tennessee, Knoxville, TN 37996, United States of America
b
Airfoils, Incorporated, Port Matilda, PA 16870, United States of America

a r t i c l e i n f o a b s t r a c t

Article history: Slotted, natural-laminar-flow (SNLF) airfoils are a novel aerodynamic concept that enable significant
Received 1 July 2020 performance improvements over conventional, single-element NLF airfoils. The S207, SNLF airfoil has
Received in revised form 27 August 2020 been designed using requirements derived from a transonic, truss-braced wing commercial aircraft
Accepted 13 September 2020
configuration. The airfoil is designed for a cruise Reynolds number of 13.2 × 106 and it exhibits a drag
Available online 22 September 2020
Communicated by Dionysios Angelidis
divergence Mach number exceeding 0.71. At these conditions, low drag is predicted to occur between lift
coefficients of 0.40 and 0.79. The S207 airfoil exhibits a range-based figure of merit nearly triple that
Keywords: of a turbulent airfoil representative of modern commercial transport aircraft. Slotted configurations also
Aerodynamics have benefits for low-speed, high-lift conditions representative of approach. The S207 airfoil exhibits a
Airfoil design low-speed, maximum lift coefficient in excess of 2.1, and is limited by compressibility effects around the
Natural laminar flow leading edge. Incorporation of this airfoil onto a transonic, truss-braced wing configuration shows strong
Boundary-layer transition potential for meeting mid- and far-term goals for reducing aircraft fuel/energy consumption.
© 2020 Elsevier Masson SAS. All rights reserved.

    
1. Introduction ML 1 Wi
Range = a∞ ln (1)
D T SFC Wf

Advanced airfoil design is a key enabling technology for achiev- simultaneous improvements to the aerodynamic performance
ing the substantial reductions in aircraft fuel/energy consumption (M L / D), propulsion (T S F C ), and structures (W i / W f ) are required.
Studies sponsored by NASA over the past 10-15 years have yielded
necessary to sustainably meet the ever-growing demands of com-
novel configurations such as the MIT/Aurora Double Bubble [2]
mercial aviation. Studies by the Air Transport Action Group, as
and the Boeing Transonic Truss-Braced Wing (TTBW) [3,4]. Both of
reported in the NASA Aeronautics Research Mission Directorate’s these concepts target transcontinental mission profiles typified by
Strategic Implementation Plan [1], indicate that improved vehicle the Boeing 737 and Airbus A320 with 150+ passenger capacities.
efficiency is a key enabler in achieving a net reduction in car- In terms of aerodynamic efficiency, both configurations anticipate
bon emissions. Pursuant to this goal, NASA has defined metrics for substantial improvements through achieving natural laminar flow
near-term (N+1), mid-term (N+2), and far-term (N+3) goals for re- (NLF) on the wing. Although this necessitates reduced wing sweep,
duction in fuel consumption relative to a 2005 best-in-class trans- the drag reduction more than offsets the reduction in cruise Mach
number. The need for this trade-off is echoed by Abbas et al. [5] in
port aircraft [1]. Achieving the far-term goals of 60-80% reduction
their review of aircraft performance improvement. An alternative
requires revolutionary changes to the aircraft configuration. If one
strategy is to use laminar-flow control [6,7], but this introduces
considers the well-known Breguet range equation, additional system-level trade-offs by requiring the addition of me-
chanical suction systems [6,8].
There are significant technical barriers in achieving large runs
of laminar flow at conditions relevant to commercial transport
aircraft in cruise. The flight conditions include both high Mach
* Corresponding author.
E-mail address: jcoder@utk.edu (J.G. Coder). numbers and high Reynolds numbers, and the aircraft configura-
1
Department of Mechanical, Aerospace and Biomedical Engineering. tions tend towards high wing loadings that drive up the section

https://doi.org/10.1016/j.ast.2020.106217
1270-9638/© 2020 Elsevier Masson SAS. All rights reserved.
J.G. Coder and D.M. Somers Aerospace Science and Technology 106 (2020) 106217

Nomenclature

a∞ free-stream speed of sound . . . . . . . . . . . . . . . . . . . . . . . . ft/s D drag . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . lbs


Cp pressure coefficient L lift . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . lbs
c airfoil chord . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ft M Mach number
cd section profile-drag coefficient Re Reynolds number based on chord and free-stream flow
cl section lift coefficient conditions
cl,ll lower limit of low-drag lift-coefficient range T SFC thrust-specific fuel consumption . . . . . . . . . . . . lbs/hr-lbs
cl,max section maximum lift coefficient t airfoil thickness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ft
cl,ul upper limit of low-drag lift-coefficient range W aircraft weight . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . lbs
cm section pitching-moment coefficient about the quarter x airfoil abscissa . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ft
chord y airfoil ordinate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ft
c m ,0 section zero-lift pitching-moment coefficient α angle of attack . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . degrees

lift coefficients. A survey of the literature highlights the difficulty This work focuses on the design of an SNLF airfoil targeted
of airfoil design to simultaneously satisfy these three requirements. toward N+3 aircraft configurations with specific design require-
For example, the redesign of the NASA Common Research Model ments derived from the Mach 0.745 variant of the Boeing Tran-
for laminar flow decreased the design lift coefficient relative to the sonic Truss-Braced Wing (TTBW). The outcome of this effort is
fully turbulent baseline of C L = 0.5 [9]. Recent experimental stud- the S207, slotted, natural-laminar flow airfoil. This airfoil was first
ies by Xu et al. [10] on a laminar-flow wing focused on aircraft documented by Somers in Ref. [23], and this paper expounds on
lift coefficients in the range of 0.38 - 0.46. A laminar-flow opti- that work by providing a more detailed, contextual overview of
mization study by Rashad and Zingg [11] included a case based the design process and additional aerodynamic results. Transonic
on the Boeing 737-800; however, they did not include a sweep NLF design considerations are presented, including identification
correction when determining the equivalent section lift coefficient. of an appropriate turbulent baseline airfoil and a discussion on
Yang et al. [12] demonstrated an inverse design strategy for hy- the aerodynamic behaviors of slotted airfoils. The design process of
the S207 airfoil including iteration with the sizing of an associated
brid laminar-flow control on a glove affixed to a vertical stabilizer,
TTBW configuration is described. Finally, theoretical aerodynamic
and while they targeted very high Reynolds numbers, the airfoil
predictions are presented in the form of section characteristics and
and glove were symmetric with small effective incidence angles.
pressure distributions, and the impact on aircraft performance is
Eggleston et al. [13] experimentally demonstrated that supercriti-
briefly described.
cal airfoils could be designed to achieve extensive runs of laminar
flow. Zhang et al. [14] successfully coupled a genetic algorithm 2. Theoretical background
with a CFD solver to optimize a supercritical laminar-flow airfoil,
but their target lift coefficients were limited to 0.55. Similarly, Zhao 2.1. Laminar-turbulent transition mechanisms
et al. [15] used a Kriging-based surrogate optimization for optimiz-
ing a supercritical natural-laminar-flow airfoil with a lift coefficient The efficacy of natural laminar flow in enabling substantial
of 0.50. Higher lift coefficients were included in the CFD-based fuel-burn reductions for commercial aircraft requires consideration
optimization work of Robitaille et al. [16] for transport-relevant of potential mechanisms for laminar-turbulent transition. After a
flight conditions, but they were able to achieve laminar flow only top-level assessment of contemporary commercial aircraft and pro-
to ∼ 10% chord. Amoignon et al. [17] used an adjoint-based ap- posed N+3 configurations, and assuming that the aircraft encoun-
proach applied to the parabolized stability equations to optimize a ters very low levels of free-stream turbulence in cruise, the domi-
transonic airfoil with realistic constraints, and while they demon- nant transition mechanisms are identified to be:
strated transition delay, they did not appreciably reduce the wave
drag. 1) Streamwise (e.g. Tollmien-Schlichting) instabilities
Slotted, natural-laminar-flow (SNLF) airfoils are a novel con- 2) Crossflow instabilities
cept that show the potential to overcome the technical barri- 3) Attachment-line contamination
ers of single-element airfoils and achieve natural-laminar-flow in 4) Separation-induced transition
5) Görtler instabilities
flight-relevant conditions. SNLF airfoils were originally proposed
6) Surface imperfections
by Somers [18], and airfoils of this type have been designed for
low-speed [19], business-jet [20], and rotorcraft applications [21].
Of these potential mechanisms, the streamwise and crossflow in-
The basic concept of SNLF airfoils, described in more detail in the
stabilities are the most prominent early in the design process.
next section, exploits the interactions of a multielement airfoil to
Streamwise instability mechanisms are the more classical path to
allow favorable pressure gradients, which stabilize the boundary
natural transition on aerodynamic lifting surfaces, and their growth
layer, to extend further aft than is possible on a single-element can be suppressed by designing the forward portion of the airfoil
airfoil without turbulent separation. Slotted airfoils have also been (and by extension, the wing) to feature favorable pressure gradi-
demonstrated to reduce wave drag via an off-surface pressure re- ents. There are two key limiting factors in the extent of laminar
covery [22]. For low-speed operations, SNLF airfoils have substan- flow that may be achieved on an airfoil. First, an adverse pressure
tially higher maximum lift coefficients than their single-element gradient is always required on the upper surface when operating
counterparts, often with cl,max > 2.0. This offers the potential ben- at positive lift coefficients. This gradient is also known as the pres-
efit of reducing wing planform area and/or permitting a simpler, sure recovery, and if it is too strong, turbulent separation occurs
and therefore lighter, high-lift system. By providing these benefits and negates the benefits of laminar flow. Second, favorable pres-
simultaneously rather than through trade-offs and compromises, sure gradients on the upper surface are associated with reduced
SNLF airfoils effectively change the rules of airfoil design and al- lift. Moreover, the pressure gradients required to control instabil-
low a designer greater degrees of freedom. ity growth become more favorable as Reynolds number increases.

2
J.G. Coder and D.M. Somers Aerospace Science and Technology 106 (2020) 106217

Both of these factors present unique challenges for realizing the


benefits of natural laminar flow at flight-relevant Reynolds num-
bers and lift coefficients.
Crossflow instabilities are prominent in swept-wing applica-
tions, as the mis-alignment of the pressure gradient and free- (a) MD 30P/30N, multielement, high-lift airfoil [32].
stream flow directions induce an inflection point in the boundary
layer. The traditional rule of thumb is that wings with less than
15◦ of leading-edge sweep are dominated by streamwise instability
mechanisms, whereas wings with greater sweep are strongly dom-
inated by crossflow [24,25]. This has profound implications on the
full aircraft design, as reducing wing sweep to mitigate crossflow (b) Original slotted, supercritical airfoil [22].
instability growth decreases the attainable cruise Mach number.
For comparison, modern commercial transports generally feature
wing sweeps in excess of 30◦ to achieve cruise Mach numbers in
excess of 0.8 [26]. Thus, reducing overall aircraft fuel burn places
a stronger burden on profile-drag reduction, as the aerodynamic (c) Slotted rotorcraft airfoil [33].
figure of merit in the Breguet range equation is M L / D, rather
than simply L / D. Further complicating the matter is that the fa-
vorable pressure gradients used to control streamwise instability
growth have the side effect of destabilizing crossflow instabilities.
Striking a balance between these two mechanisms is the basis
of the CAT-NLF design strategy proposed by Campbell and Lynde (d) S103, slotted, natural-laminar-flow airfoil [19].
[27]; however, the TTBW reference configuration features 12.5◦ of
leading-edge sweep, allowing the airfoil design to focus primarily Fig. 1. Examples of slotted airfoils for various applications.
on streamwise transition mechanisms.
Regarding the other transition mechanisms, attachment-line
contamination is a concern with rear-swept wings for which the the formation of a strong suction peak on the downstream ele-
turbulent boundary layer on the fuselage can incite transition on ment. Given that this element usually has a fresh boundary layer,
the wing. Wings with reduced sweep are less prone to attachment- the suction-peak suppression enables the airfoil to operate at much
line contamination, and swept wings may be designed to feature higher angles of attack before stall occurs.
reduced sweep on the inboard portion of the wing to prevent This concept is exploited liberally in modern high-lift systems
it [28]. Separation-induced transition encompasses both laminar- on transport-type aircraft [32]. The prototypical modern configu-
separation bubbles and shock-induced separation; however, these ration consists of a leading-edge slat and a single-slotted Fowler
may be managed and controlled through careful design of the flap, such as included on the high-lift variation of the Common
pressure gradients. Görtler instabilities occur in boundary layers on Research Model designed by Boeing [34]. The combination of a
concave surfaces [29]. This necessitates surface convexity on the slat and a slotted flap reduces the pressure gradient on the main-
airfoil upstream of the streamwise- or crossflow-dominated tran- element upper surface, enabling such airfoils to attain very high
sition location, but this constraint is relatively easy to achieve for maximum lift coefficients. The benefits of slotted, high-lift airfoils
single-element airfoils. Aft loading on the airfoil, which is typi- have also been explored for rotorcraft applications with emphasis
cal of both natural-laminar-flow and supercritical airfoils, trans- on employing slats on the outboard portion of the blade [33].
lates to a concave cove region on the lower surface. For single- Slotted airfoils have also been found to provide benefits for air-
element airfoils, however, there is an adverse pressure gradient at foils operating in the transonic regime due to the dumping-velocity
the start of the cove, resulting in transition being dominated by effect. In fact, Whitcomb’s original supercritical airfoil concept em-
streamwise mechanisms. Lastly, surface imperfections, such as con- ployed a slot towards the trailing edge to exploit this phenomenon
taminants, steps, and gaps, may incite early transition. Although [22]. The lower pressure at the main-element trailing edge en-
these features cannot be fully controlled in the airfoil design pro- ables the pressure recovery to become less severe, thereby reduc-
cess, design strategies may be employed to mitigate their behavior ing the strength of the upper-surface shockwave [22]. Ultimately,
such as designing to lower instability thresholds [30] or by sim- Whitcomb was able to incorporate the same pressure-distribution
ply assuming fully turbulent flow downstream of the imperfec- concepts into a single-element, “integral” airfoil [35]. Nevertheless,
tion. slotted configurations have significant benefits for transonic perfor-
mance, and they have garnered more recent interest by Boeing for
increasing the drag-divergence Mach number beyond that afforded
2.2. Overview of slotted airfoils
by conventional supercritical airfoils [36].
A newer development is the slotted, natural-laminar-flow
Slotted airfoils, such as those pictured in Fig. 1, have been used (SNLF) airfoil designed by Somers [18]. The first such airfoil, the
since the early days of aviation, although the primary emphasis of S103, is shown along with its on-design (Re = 9.0 × 106 , cl = 0.2,
their use has been on increasing the maximum lift coefficient over M = 0.1) pressure distribution in Fig. 2. SNLF airfoils exploit the
that of an unslotted baseline. The immediate aerodynamic impact dumping velocity and off-surface pressure recovery to overcome
of a slot is two-fold. First, it allows the element upstream of the the limitations single-element airfoils face for achieving long runs
slot to have a trailing-edge pressure lower than that of the free of laminar flow. While the entire airfoil system must recover to a
stream with the remainder of the pressure recovery occurring off positive C p at the trailing edge, this constraint does not apply to
the surface. Smith [31] also referred to this behavior as the “dump- the upstream elements. Thus, the fore element is able to support
ing velocity” due to the accelerated flow at the boundary-layer full-length favorable pressure gradients and, therefore, full-length
edge. Higher dumping velocities allow the element to carry ad- laminar flow. The aft element generates circulation and interacts
ditional load before viscous separation. Second, the slot suppresses with the fore element to sustain the dumping velocity. Because

3
J.G. Coder and D.M. Somers Aerospace Science and Technology 106 (2020) 106217

Fig. 3. Boeing SUGAR High TTBW model installed in the NASA Ames 11-ft Unitary
Plan Wind Tunnel.

the similarities suggest that the CRM airfoil is an appropriate tur-


bulent baseline for quantifying the benefits of natural laminar flow.
Fig. 2. Predicted, on-design pressure distribution of the S103, SNLF airfoil [19]. The NASA CRM website features a detailed description of the
airfoil used at the 65%-semispan station,2 referred to as the
the aft element has a fresh, initially laminar boundary layer, it can CRM.65 airfoil, including geometric definition and sweep correc-
maintain laminar flow on its entire lower surface and over a large tions. They report a sweep of 31.5 degrees and a lift coefficient of
portion of its upper surface. Thus, the only turbulent flow on an cl = 0.57 for a cruise Mach number of 0.85 at this station. Applying
SNLF airfoil is a small portion of the aft-element upper surface. standard sweep corrections provides reference two-dimensional
conditions of M = 0.725 and cl = 0.784. Applying the 12.5-degree
3. Reference configuration sweep of the Mach 0.745 TTBW to the 2D reference conditions
of this airfoil results in adjusted 3D conditions of M = 0.743 and
cl = 0.747. Further adjusting non-uniformity of the wing lift dis-
3.1. Transonic, truss-braced wing aircraft
tributions (i.e. how the local wing section operates at cl = 0.57
while the aircraft operates at C L = 0.5) suggests C L = 0.656 for
The reference aircraft used in this airfoil design study is the
the associated TTBW wing. These values are very close to those
Boeing Transonic Truss-Braced Wing (TTBW), shown in Fig. 3,
suggested for the Mach 0.745 TTBW configuration. Thus, the CRM
which has shown great promise for achieving mid- and far-term
airfoil is confirmed as an appropriate surrogate for the TTBW air-
performance goals [3,4,37]. Several variants of this concept have
foils for benchmarking the performance improvements offered by
been designed for various Mach numbers, different propulsion con-
SNLF airfoil technology.
figurations, and both laminar and turbulent wings. Of these, the
The pressure distribution predicted for these conditions using
‘765-095 Rev-J’, Mach 0.745 variant serves as the starting point
MSES [41] is shown in Fig. 4 with the sonic limit identified. Several
for providing realistic constraints on the airfoil design. This aircraft
key features of the design may be observed in this distribution. The
features a high-aspect-ratio wing with a relatively low leading-
flow on the upper surface is accelerated quickly around the leading
edge sweep angle of 12.5◦ . Thus, transition is assumed to be dom-
edge and undergoes a gentle net-adverse pressure gradient before
inated by streamwise mechanisms and the management of cross-
compressing through a shockwave around x/c = 0.55. Downstream
flow instabilities can be neglected during the airfoil design process.
of the shock, the flow continues to compress. Around x/c = 0.95,
Through partnerships with Boeing Research & Technology and
the final closure contribution on the upper surface is a stronger
Penn State University, supported by the NASA University Leader-
adverse gradient. The gradients on the aft portion of the upper
ship Initiative program, the system-level impacts of using an SNLF
surface are sufficiently weak that the flow remains fully attached
airfoil on a TTBW can be fully considered. Most notably, this al-
at this condition.
lows the aircraft to be resized and re-optimized as necessary to
The static pressure at the trailing edge recovers to a value
allow better airfoil/aircraft synergy and achieve the greatest possi-
higher than that of the free stream, which is a necessary condition
ble benefits.
for a closed geometry without separation. Including a divergent
trailing edge essentially opens the section and alleviates this con-
3.2. Turbulent baseline airfoil straint. The upper and lower surfaces no longer need to recover
to the same pressure on the surface, allowing the upper surface
The exact NLF airfoil used to develop the Mach 0.745 TTBW to operate at a lower static pressure. Lessening the amount of ad-
configuration remains propriety; however, computational studies verse gradient required for closure allows the design to be more
performed by NASA in support of wind-tunnel entries have been aggressive. For example, the shock may be weakened and shifted
published in Refs. [38] and [39]. In these, the wing pressure dis- further aft for the same lift coefficient and Mach number [40].
tributions are qualitatively similar to modern, turbulent transonic At a fundamental level, the aerodynamic influence of a divergent
sections with divergent trailing edges (DTEs) [40], except that the trailing edge is the same as separation ramps (cf. [42]) commonly
upper-surface pressure gradients are favorable to the shock to pro- used on natural-laminar-flow airfoils and originally proposed by
mote laminar flow. A further survey of the literature shows that F. X. Wortmann. This feature confines turbulent separation to a
the TTBW pressure distributions exhibit similarities to those of the small region near the trailing edge, and permits longer runs of
NASA Common Research Model (CRM) [26]. Granted, the CRM lacks
a divergent trailing edge, but this technology targets the aft pres-
2
sure recovery rather than the forward gradients [40]. Nevertheless, https://commonresearchmodel.larc.nasa.gov/crm-65-airfoil-sections/.

4
J.G. Coder and D.M. Somers Aerospace Science and Technology 106 (2020) 106217

Fig. 4. Predicted pressure distribution of the CRM.65 airfoil at reference conditions. Fig. 5. Sketch of desired single-element pressure distribution.

4. Design of the S207, slotted, natural-laminar-flow airfoil


laminar flow, higher maximum lift coefficients, and docile stall
characteristics. Additionally, the presence of a separation ramp is 4.1. Overview of the design process
conducive to thickening the airfoil trailing edge for manufacturing
from the sharp, as-designed geometry. The design of a slotted, natural-laminar-flow airfoil consists of
Although significant gains in aircraft performance are antic- two stages. In the first stage, a single-element “stepping stone” air-
ipated from achieving laminar flow on the wing, the pressure foil is designed such that it exhibits the pressure gradients desired
distributions for this class of airfoil are not conducive to natural for the fore element of the slotted airfoil. The influence of the aft
laminar flow. Even in the absence of crossflow instabilities, the ad- element is included through the use of very strong aft loading.
verse pressure gradients on the upper surface will promote rapid Generally, the aft loading and associated pressure recovery of the
growth of Tollmien-Schlichting instabilities at Reynolds numbers SNLF stepping-stone airfoil will exceed the permissible limits for
typical of commercial transport aircraft in cruise. The lower surface a conventional, single-element airfoil. A sketch of a representative
would be anticipated to maintain laminar flow until approximately pressure distribution for this process is shown in Fig. 5.
mid-chord, beyond which the adverse gradients will promote rapid In the design of the S207 airfoil, the Eppler Airfoil Design and
instability growth. Granted, the design of the Mach 0.745 TTBW Analysis Code [43,44], which is a subsonic single-element code,
does not anticipate appreciable laminar flow on the lower surface was used to design the initial fore- and aft-element shapes. The
due to the use of Krueger flaps for low-speed operations [3]; how- MSES code [41], which is a multielement, transonic code, was used
ever, MSES analyses of the CRM airfoil with both free and forced to refine the shapes in the two-element configuration.
transition on the lower surface indicate an approximately 25% in-
crease in total profile drag due to early transition. 4.2. Initial design requirements
Achieving the full benefit of natural laminar flow requires a
different approach to the pressure distribution. On the upper sur- The starting point for any airfoil design is a set of well-posed
face, a favorable pressure gradient is required all the way back to design requirements. Typical design requirements may include
the shock. While such a distribution is attainable, it nevertheless maximum lift coefficient, minimum lift coefficient, the limits of the
presents a significant technical barrier in terms of performance. low-drag, lift-coefficient range, the zero-lift pitching-moment coef-
If the shock strength and location are to be preserved, the airfoil ficient, and the permissible thickness ratio. For the aerodynamic
lift coefficient must decrease, which would require a net increase coefficients, the Reynolds and Mach numbers at which they occur
in wing planform area. Recovering the lift while preserving shock must also be specified.
strength would require the total laminar run to be decreased. Al- Because of the relatively open design space for the S207 airfoil,
ternatively, the laminar run can be maintained by increasing the the strongest emphasis was placed on the maximum lift coeffi-
shock strength, but the gains from laminar flow would be wiped cient and the limits of the low-drag, lift-coefficient range. The
out by the wave drag. A notable effort in optimizing the trade- initial specified values were derived in collaboration with Boeing
offs between skin-friction and wave drags can be found in the Research & Technology, and are based on the Mach 0.745 Tran-
crossflow-attenuated NLF (CAT-NLF) design methodology of Camp- sonic Truss-Braced Wing with standard sweep corrections applied.
bell and Lynde [27]. They coupled the LASTRAC boundary-layer The values are summarized below in Table 1. The maximum lift co-
stability solver with a RANS-based computational fluid dynamics efficient was specified based on landing and approach conditions
method, and applied it to redesign the CRM wing for natural lam- to be cl,max = 2.4 at M = 0.225 and Re = 14.8 × 106 , which was
inar flow (the CRM-NLF) [28]. Their early investigations suggested an aggressive target intended to exploit the capabilities of slot-
a near-zero pressure gradient was suitable for attaining laminar ted airfoils and ideally lessen the burden on high-lift systems. The
flow at flight-relevant Mach and Reynolds numbers [27,28]. An as- lower and upper limits of the low-drag lift-coefficient range were
sessment of their predictions for the CRM-NLF indicates that this chosen based on the anticipated climb and cruise points for the
leaves very little margin for off-design conditions, and experimen- Transonic, Truss-Braced Wing, respectively. For this aircraft, these
tal validation of the concept in the National Transonic Facility (de- respective points are reversed from what one would typically ex-
tailed in Ref. [9]) has unclear results. pect [42,45,46]. The lower limit of cl,ll = 0.35 corresponds to climb,

5
J.G. Coder and D.M. Somers Aerospace Science and Technology 106 (2020) 106217

Table 1
Initial airfoil design requirements derived from the Boeing M = 0.745 TTBW.

Parameter Value Reynolds Number Mach Number


cl,max 2.4 14.8 × 106 0.225
cl,ll 0.35 24.3 × 106 0.40
cl,ul 0.71 13.1 × 106 0.745
c m ,0 Unconstrained - -
(t /c )max Unconstrained - -

even though it is a higher Reynolds number of Re = 24.3 × 106 ,


whereas the upper limit of cl,ul = 0.71 corresponds to cruise and
is a lower Reynolds number of Re = 13.1 × 106 . A consequence of
this behavior reversal is that the upper limit of the low-drag lift-
coefficient range must be achieved at a significantly higher Mach
number than the lower limit (M = 0.745 versus M = 0.40).
The zero-lift pitching-moment coefficient and the thickness ra-
Fig. 6. Progression of single-element, intermediate airfoils as a result of aircraft con-
tio were left unconstrained. figuration iterations.

4.3. Single-element design space exploration Table 2


Refined airfoil design requirements after configuration iteration.

It was discovered early in the design process that the design re- Parameter Value Reynolds Number Mach Number
quirements derived from the Boeing Mach 0.745 TTBW configura- cl,max 2.3 16.0 × 106 0.225
tion were too aggressive. That is, a reasonable natural-laminar-flow cl,ll 0.39 13.2 × 106 ≥ 0.66 (0.70 actual)
cl,ul 0.65 13.2 × 106 ≥ 0.66 (0.70 actual)
airfoil that satisfied the design constraints could not be found. This
c m ,0 Unconstrained - -
actually exposes a dangerous pitfall of designing an airfoil against a (t /c )max Unconstrained - -
frozen aircraft configuration: the initial airfoil design requirements
were based on aircraft performance optima, which in turn were
strong functions of the original airfoil. Consequently, the aircraft
will likely work best with the original airfoil, and this combination
will be a local optimum. Finding a global fuel-burn optimum, or at
least a better local optimum, therefore requires the cruise condi-
tions, wing planform, and airfoil to be designed simultaneously in Fig. 7. S207, slotted, natural-laminar-flow airfoil.
a coupled manner.
One of the biggest challenges in this design, as outlined in the 4.4. Final geometry
previous section of this paper, is achieving natural laminar flow
at both high lift coefficient and high Mach number. It is more- With the design requirements frozen, a final single-element
or-less an “area under the curve” limitation. Without a valid and airfoil was designed and used as the basis of a slotted, natural-
suitable NLF airfoil as a starting place, the coupled airfoil/aircraft laminar-flow airfoil. The new, SNLF airfoil, termed the S207, is
design process cannot proceed. Thus, it was deemed advantageous pictured in Fig. 7. It features a thickness ratio of t /c = 13.49%, a
to relax the requirement for the upper limit of the low-drag lift- fore element chord that is 81.3% of the total airfoil chord, and an
coefficient range to cl,ul = 0.65 and M = 0.65 to find this starting aft element chord that is 29.7% of the total chord.
place. It was discovered early that both the high Reynolds number
associated with cl,ll , which primarily influences the lower surface, 5. Predicted aerodynamic characteristics
excessively constrained the design space, and drove the airfoil to
unacceptably large thicknesses approaching t /c = 20%. Thus, satis- 5.1. Theoretical methods
fying the original climb constraint results in a severe limitation on
cruise Mach number, which was deemed unacceptable. Overcom- Aerodynamic characteristics of the S207 airfoil have been pre-
ing this required the conditions on the lower surface to be relaxed, dicted using the MSES code [41]. MSES is a coupled Euler/integral-
and after system-level assessments, it was deemed more appropri- boundary-layer solver well-suited for analyzing multielement air-
ate to define the cl,ll conditions to be those of cruise. With these foils in transonic conditions. It features a three-equation boundary-
adjustments in place, an initial, viable, single-element airfoil was layer model. Two of the equations predict the displacement and
obtained. Theoretical section characteristics were provided as in- momentum thicknesses. The third equation is used for transition
puts for full aircraft performance predictions with a configuration prediction in the laminar boundary layer and a lag parameter in
based on the Mach 0.745 TTBW. Resizing of the TTBW aircraft by the turbulent boundary layer. The code is capable of predicting
Penn State University using the single-element airfoils and iterat- transition using either a full e N method [47] or the approximate
ing on the cruise conditions allowed the final design constraints envelope method of Drela and Giles [48]. For the cases analyzed
to be defined. The progression of single-element airfoils is shown here, the code was run using the approximate-envelope method
in Fig. 6, and the design requirements for the slotted configura- and a specified critical amplification factor of 9. For transition-
tion are detailed in Table 2. It was found during the design of the free cases, a boundary-layer trip was specified at 98% of the fore-
slotted configuration that a higher drag-divergence Mach number element upper surface to improve solver convergence. Without a
was attainable. As a result, limits of the low-drag, lift-coefficient trip, this location experiences a laminar separation bubble even at
range were designed to occur for M = 0.70, and drag divergence high-Reynolds number conditions. Additional transition-fixed cases
was found to occur beyond M = 0.71. were run including specifying transition at the leading edge of

6
J.G. Coder and D.M. Somers Aerospace Science and Technology 106 (2020) 106217

each element as well as at the entrance to the slot on the fore-


element lower surface at x/c = 0.70. This latter scenario is in-
tended to model the potential for Görtler instabilities to induce
transition in this region.

5.2. High-speed theoretical results

5.2.1. Section characteristics


Predicted section characteristics for the S207 airfoil with free
transition for multiple Mach numbers all with Re = 13.2 × 106 are
plotted in Fig. 8. The lowest Mach number shown is M = 0.66,
which is the lower limit for acceptable cruise Mach numbers. Very
low levels of drag (cd < 30 counts, where 1 count is 0.0001) are
achieved in the low-drag lift coefficient range. The lower limit of
the low-drag range meets the design requirement, but the upper
limit is around cl = 0.55, which is lower than the target value. The
fore element is completely subcritical in the low-drag range, while
there is a small pocket of supersonic flow in the slot that recov-
ers without a shock. For M = 0.70, even lower values of drag are (a) Lift and Pitching Moment
achieved (∼ 27 counts) and the upper limit is shifted upward to
cl = 0.73, which exceeds the target value. This behavior is to the
benefit of the airfoil, as cl,ul was one of the values compromised
upon during the design-space exploration with the single-element
airfoil. Wave drag accounts for approximately 0.1 counts of drag
in the low-drag range at this Mach number. Further increasing
the Mach number to 0.71 introduces a small amount of additional
wave drag that gets stronger as cl increases, and the upper limit
shifts to cl = 0.8. By M = 0.72, drag divergence occurs and wave
drag accounts for a significant portion of the drag even in the low-
drag range.
The performance improvements offered by SNLF airfoils may be
quantified using M L / D as a first-order measure, as this quantity
is the aerodynamic figure of merit in the Breguet range equation.
The predicted M L / D values for the S207 with free transition at
Re = 13.2 × 106 are compared with those of the CRM.65 airfoil for
the same Reynolds number in Fig. 9. Two curves are included for
the CRM airfoil: one with free transition at M = 0.740, which was
found to be the peak performance of the airfoil, and one with tran-
sition fixed at the leading edge to simulate fully turbulent flow at (b) Profile Drag
the design Mach number of 0.725. The fully turbulent CRM airfoil
achieves a maximum M L / D of 64, while permitting free transition Fig. 8. S207 section characteristics for Re = 13.2 × 106 and various Mach numbers.
increases this to 123. This contrasts strongly with the performance
of the S207 airfoil, which achieves a maximum M L / D of 178 at
M = 0.70. At M = 0.71, which is below drag divergence by stan-
dard metrics, the maximum value of M L / D drops to 167; how-
ever, further increases to Mach number result in a rapid dropoff
in performance. These predictions suggest that the S207 airfoil can
achieve a 178% increase in aerodynamic performance over the fully
turbulent CRM baseline, and a 44% increase when free transition is
permitted for the baseline airfoil.

5.2.2. Pressure distributions


Pressure distributions for three separate lift coefficients at M =
0.71 and Re = 13.2 × 106 are plotted in Fig. 10. The lift coefficients
are cl = 0.39, which is near the lower limit of the low-drag range,
cl = 0.65, which was the original target value associated with
cruise, and cl = 0.78, which is near the upper limit of the low-
drag range. The fore element exhibits expected behaviors where
the pressure gradient on the lower surface becomes more favorable
as the lift coefficient increases while the upper-surface gradient
becomes less favorable. As is typical of SNLF configurations, the
pressure distribution on the aft element shows very little sensitiv-
ity to airfoil lift coefficient.
There are prominent Mach number effects observed in the
section characteristics. Of particular note is that the upper cor-
ner of the low-drag range shifts to higher lift coefficients as the Fig. 9. Aerodynamic figure of merit of the S207 and CRM.65 airfoils.

7
J.G. Coder and D.M. Somers Aerospace Science and Technology 106 (2020) 106217

Fig. 10. Effect of lift coefficient on predicted pressure distribution for M = 0.71 and Fig. 12. Effect of Mach number on predicted pressure distribution for cl = 0.70 and
Re = 13.2 × 106 . Re = 13.2 × 106 .

pressure gradient to be supported on the fore element in the slot.


The side effect of the slot becoming a supersonic nozzle is that
a strong shock forms over the aft-element upper surface, causing
both shock-induced laminar separation and wave drag.

5.2.3. Potential impact of Görtler instabilities


The fore element of the S207 airfoil features a concave region
on the lower surface through the slot, introducing the potential
for Görtler instabilities [29] in the boundary layer. This possibil-
ity was recognized in the first wind-tunnel investigation of an
SNLF airfoil, the S103, as affecting the total drag values [19]. While
a detailed investigation of the Görtler instabilities in the slot is
beyond the scope of the current study, the potential impact on
aerodynamic performance is assessed by forcing transition at the
slot entrance on the lower surface of the fore element. The sec-
tion characteristics with the tripped lower surface for M = 0.71
and Re = 13.2 × 106 are plotted in Fig. 13. The limits of the low-
drag range exhibit very little sensitivity, whereas the value of drag
increases by approximately 6 counts as a result of the increased
Fig. 11. Effect of Mach number on predicted pressure distribution for α = −1.5◦ and turbulent flow.
Re = 13.2 × 106 . Pressure distributions for both transition free and simulated
Görtler instabilities at fixed α = −1◦ are plotted in Fig. 14. The re-
Mach number increases. The origin of this behavior can be seen gion primarily affected by the premature transition is the slot. An
in the pressure distributions at M = 0.66 and M = 0.70 for fixed abrupt adverse pressure gradient forms on both elements, and this
α = −1.5◦ , which are plotted in Fig. 11. On the fore-element lower is attributed to the increased thickness turbulent boundary layer
surface, there is little influence of Mach number except at the which results in additional blockage through the slot.
slot entrance, whereas increasing free-stream Mach number pro-
motes additional flow acceleration on the upper surface. This sta- 5.3. Theoretical low-speed results
bilizes the boundary layer and delays transition, which enables the
upward shift of the upper limit of the low-drag range. The aft- The target maximum lift coefficient of the S207 airfoil and the
element pressure distribution is mostly unchanged with increasing associated Mach and Reynolds numbers of 0.225 and 16.0 × 106 ,
Mach number save for a small region on the upper surface down- respectively, reflect anticipated approach and landing conditions.
stream of the slot. Analysis of the S207 airfoil for these conditions with transition
Increasing the Mach number past M = 0.71 results in the for- free using MSES shows a maximum lift coefficient, cl,max , of ap-
mation of drag-producing shockwaves. Pressure distributions for proximately 2.11. The maximum lift coefficient does not fully meet
fixed α = −1◦ , fixed Re = 13.2 × 106 , and various Mach num- the design objective, and was found to be limited by compressibil-
bers are plotted in Fig. 12. From M = 0.70 to M = 0.71, the most ity rather than purely by viscous effects. There is a small pocket
significant effect is strengthening of the favorable gradient on the of supersonic flow that forms around the leading edge and recom-
fore-element upper surface. For both of these Mach numbers, flow presses through a shockwave. The influence of compressibility is
through the slot is supersonic towards the aft element but sub- confirmed through analysis at a slightly lower Mach number of
sonic towards the fore element. Increasing to M = 0.72 results in 0.2. At this condition, and for the same Reynolds number, the max-
additional strengthening of the upper-surface gradient as well as imum lift coefficient is predicted to be in excess of 2.23, which is
fully supersonic flow at the slot exit. This allows a more favorable an increase of at least 5%. The S207 airfoil was also analyzed with

8
J.G. Coder and D.M. Somers Aerospace Science and Technology 106 (2020) 106217

(a) Lift and Pitching Moment (a) Lift and Pitching Moment

(b) Profile Drag (b) Profile Drag

Fig. 13. Possible effect of Görtler instabilities on predicted section characteristics for Fig. 15. Predicted low-speed section characteristics for Re = 16.0 × 106 .
Re = 13.2 × 106 and M = 0.71.

transition fixed at the leading-edge of the fore element to simu-


late potential leading-edge contamination. This was found to have
a negligible influence on the lift curve with ∼ 1% decrease in cl,max ,
although there is an expected increase in profile-drag coefficient.
The lift, pitching-moment, and profile-drag curves for these three
cases are plotted in Fig. 15.
Pressure distributions for the M = 0.225, transition-free results
are plotted in Fig. 16 for angles of attack of 5◦ , 10◦ , and 15◦ .
As was observed with the high-speed results, the change in lift
with change in angle of attack is carried much more strongly by
the fore element than it is by the aft element. The aft element
exhibits very little change in pressure distribution across the angle-
of-attack range plotted, with less variation between 10◦ and 15◦
than there was between 5◦ and 10◦ . A dumping-velocity effect
[31] may be observed in the pressure distributions in which the
fore element recovers to C p ≈ −0.6, which is lower than the free-
stream pressure. This allows the fore element to support much
higher velocities on its upper surface than would be permissible
on a single-element airfoil without boundary-layer separation. This
Fig. 14. Possible effect of Görtler instabilities on predicted pressure distributions for is a key feature of the SNLF concept for increasing cl,max while also
α = −1◦ , Re = 13.2 × 106 , and M = 0.71. enabling very low cruise profile-drag coefficients.

9
J.G. Coder and D.M. Somers Aerospace Science and Technology 106 (2020) 106217

Funding sources

This work was supported by the National Aeronautics and Space


Administration (NASA) University Leadership Initiative (ULI) “Ad-
vanced Aerodynamic Design Center for Ultra-Efficient Commercial
Vehicles” (Award NNX17AJ95A).

Declaration of competing interest

The authors declare that they have no known competing finan-


cial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Acknowledgements

The authors would like to acknowledge P. Camacho, L. Chou,


K. Pham, N. Harrison, and A. Khodadoust from Boeing Research &
Technology and L. Metkowski and M. Maughmer from Penn State
University for their inputs on the airfoil design requirements.
Fig. 16. Predicted pressure distributions for M = 0.225 and Re = 16.0 × 106 .
References

5.4. Impact on aircraft performance [1] NASA Aeronautics Strategic Implementation Plan - 2019 Update, NASA, Wash-
ington, DC, 2019.
[2] E.M. Greitzer, P.A. Bonnefoy, E.D. la Rosa Blanco, C.S. Dorbian, M. Drela, D.K.
Section characteristics of the S207 airfoil, predicted using a Hall, R.J. Hansman, J.I. Hileman, R.H. Liebeck, J. Lovegren, P. Mody, J.A. Per-
transitional RANS-based method, were provided to Boeing Re- tuze, S. Sato, Z.S. Spakovszky, C.S. Tan, J.S. Hollman, J.E. Duda, N. Fitzgerald, J.
Houghton, J.L. Kerrebrock, G.F. Kiwada, D. Kordonowy, J.C. Parrish, J. Tylko, E.A.
search & Technology for an assessment on the TTBW configuration. Wen, W.K. Lord, N+3 Aircraft Concept Designs and Trade Studies, Final Report:
After a full resizing of the aircraft against the FAA NextGen Mission Volume 1, NASA/CR-2010–216794/VOL1, 2010.
Profile, it was found that including SNLF technology enables an ap- [3] M.K. Bradley, C.K. Droney, Subsonic Ultra Green Aircraft Research: Phase I Final
proximately 58% decrease in block fuel per seat compared to the Report, NASA/CR-2011-216847, 2011.
[4] M.K. Bradley, C.K. Droney, Subsonic Ultra Green Aircraft Research Phase II: N+4
2008 “SUGAR Free” baseline, and with a cruise Mach number of
Advanced Concept Development, NASA/CR-2012-217556, 2012.
M = 0.727. More details on this study may be found in Ref. [49]. [5] A. Abbas, J. de Vicente, E. Valero, Aerodynamic technologies to improve aircraft
performance, Aerosp. Sci. Technol. 28 (2013) 100–132, https://doi.org/10.1016/
j.ast.2012.10.008.
6. Conclusion [6] G. Schrauf, Status and perspectives of laminar flow, Aeronaut. J. 109 (1102)
(2005) 639–644, https://doi.org/10.1017/S000192400000097X.
[7] R.D. Joslin, Aircraft laminar flow control, Annu. Rev. Fluid Mech. 30 (1) (1998)
The S207, slotted, natural-laminar-flow airfoil for commercial
1–29, https://doi.org/10.1146/annurev.fluid.30.1.1.
transport applications was successfully designed using constraints [8] K.S.G. Krishnan, O. Bertran, O. Seibel, Review of hybrid laminar flow control
based on a Boeing Transonic Truss-Braced Wing aircraft. The slot- systems, Prog. Aerosp. Sci. 93 (2017) 24–52, https://doi.org/10.1016/j.paerosci.
ted configuration enables extensive runs of laminar flow, confining 2017.05.005.
turbulent flow to the aft portion of the aft-element upper sur- [9] M.B. Rivers, M. Lynde, R. Campbell, S. Viken, D. Chan, A.N. Watkins, S. Goodliff,
Experimental investigation of the NASA common research model with a natural
face. Theoretical predictions show extremely low values of profile laminar flow wing in the NASA langley national transonic facility, AIAA Paper
drag at relatively high lift coefficients, and a drag-divergence Mach 2019-2189, in: AIAA SciTech 2019 Forum, San Diego, CA, January 2019, 2019.
number in excess of 0.71. Using M L / D as a figure of merit, the [10] J. Xu, Z. Fu, J. Bai, Y. Zhang, Z. Duan, Y. Zhang, Study of boundary layer tran-
S207 airfoil offers nearly triple the performance over a typical fully sition on supercritical natural laminar flow wing at high Reynolds number
through wind tunnel experiment, Aerosp. Sci. Technol. 80 (2018) 221–231,
turbulent airfoil designed for similar applications. Low-speed pre- https://doi.org/10.1016/j.ast.2018.07.007.
dictions show a maximum lift coefficient in excess of 2.1, with stall [11] R. Rashad, D.W. Zingg, Aerodynamic shape optimization for natural laminar
driven primarily through compressibility effects. flow using a discrete-adjoint approach, AIAA J. 54 (11) (2016) 3321–3337,
During the design process, the airfoil specifications were iter- https://doi.org/10.2514/1.J054940.
[12] Y. Yang, J. Bai, L. Li, T. Yang, H. Wang, S. Ma, Y. Zhang, An inverse design method
ated with the aircraft sizing to provide better synergy between the
with aerodynamic design optimization for wing glove with hybrid laminar flow
two. It was found that achieving extensive laminar flow on the control, Aerosp. Sci. Technol. 95 (2019) 105493, https://doi.org/10.1016/j.ast.
lower surface was one of the more challenging constraints. The 2015.105493.
design Reynolds number for this surface had a strong influence [13] B. Eggleston, R.J.D. Poole, D.J. Jones, M. Khalid, Thick supercritical airfoils with
low drag and natural laminar flow, J. Aircr. 24 (6) (1987) 405–411, https://
on airfoil thickness and, in turn, drag-divergence Mach number.
doi.org/10.2514/3.45460.
This necessitated relaxing of the constraints and resizing of the [14] Y. Zhang, X. Fang, H. Chen, S. Fu, Z. Duan, Y. Zhang, Supercritical natural lam-
aircraft multiple times through the design process. With the S207- inar flow airfoil optimization for regional aircraft wing design, Aerosp. Sci.
TTBW combination, mid-term fuel/energy consumption goals are Technol. 43 (2015) 152–164, https://doi.org/10.1016/j.ast.2015.02.024.
fully achieved, and it nearly meets the far-term goals. [15] K. Zhao, Z. hong Gao, J. tao Huang, Robust design of natural laminar flow
supercritical airfoil by multi-objective evolution method, Appl. Math. Mech.
Further studies, both theoretical and experimental, of the S207 35 (2) (2014) 191–202, https://doi.org/10.1007/s10483-014-1783-6.
airfoil are necessary to explore and validate its aerodynamic ca- [16] M. Robitaille, A. Mosahebi, E. Laurendeau, Design of adaptive transonic laminar
pabilities. A transonic, wind-tunnel experiment using this airfoil airfoils using the γ − Re ˜ θ t transition model, Aerosp. Sci. Technol. 46 (2015)
is anticipated to occur in early 2022. Three-dimensional compu- 60–71, https://doi.org/10.1016/j.ast.2015.06.027.
[17] O. Amoignon, J. Pralits, A. Hanifi, M. Berggren, D. Henningson, Shape optimiza-
tational fluid dynamics studies are planned to both evaluate the
tion for delay of laminar-turbulent transition, AIAA J. 44 (5) (2006) 1009–1024,
behavior of an SNLF airfoil on a finite wing and also to provide https://doi.org/10.2514/1.12431.
higher-fidelity estimates of the TTBW performance improvements. [18] D.M. Somers, Laminar-Flow Airfoil, US Patent 6,905,092 B2, June 14, 2005.

10
J.G. Coder and D.M. Somers Aerospace Science and Technology 106 (2020) 106217

[19] D.M. Somers, An Exploratory Investigation of a Slotted, Natural-Laminar-Flow 54th AIAA Aerospace Sciences Meeting, AIAA Paper 2016-0308, San Diego, CA,
Airfoil, NASA/CR–2012-217560, 2012. January 2016.
[20] D.M. Somers, Design of a Slotted, Natural-Laminar-Flow Airfoil for Business Jet [35] R.T. Whitcomb, Review of NASA Supercritical Airfoils, ICAS Paper No. 74-10,
Applications, NASA/CR–2012-217559, 2012. August 1974.
[21] J.G. Coder, M.D. Maughmer, D.M. Somers, Theoretical and experimental re- [36] J. Vassberg, L. Gea, J. McLean, D. Witkowski, S. Krist, R. Campbell, Slotted Air-
sults for the S414, slotted, natural-laminar-flow airfoil, J. Aircr. 52 (6) (2014) craft Wing, US Patent 7,048,235 B2, May 23 2006.
1883–1890, https://doi.org/10.2514/1.C032566. [37] N.A. Harrison, M.D. Beyar, E.D. Dickey, K. Hoffman, G.M. Gatlin, S.A. Viken, De-
[22] R.T. Whitcomb, L.R. Clark, An airfoil shape for efficient flight at supercritical velopment of an efficient Mach=0.80 transonic truss-braced wing aircraft, in:
Mach numbers, in: NASA TM X-1109, 1965. AIAA SciTech 2020 Forum, AIAA Paper 2020-0011, Orlando, FL, January 2020.
[23] D.M. Somers, Design of a Slotted, Natural-Laminar-Flow Airfoil for a Transport [38] J. Xiong, J. Fugate, N. Nguyen, Investigation of transonic truss-braced wing
Aircraft, NASA/CR–2019-220403, 2019. aircraft wing-strut interference effects using FUN3D, in: AIAA Aviation 2019
[24] D.I.A. Poll, Some observations of the transition process on the windward face Forum, AIAA Paper 2019-3026, Dallas, TX, June 2019.
of a long yawed cylinder, J. Fluid Mech. 150 (1985) 329–356. [39] D. Maldonado, J.A. Housman, J.C. Duensing, J.C. Jensen, C.C. Kiris, S.A. Viken,
[25] H.L. Reed, W.S. Saric, Stability of three-dimensional boundary layers, Annu. Rev. C.A. Hunter, N.T. Frink, S.N. McMillin, Computational simulations of a Mach
Fluid Mech. 21 (1989) 235–284, https://doi.org/10.1146/annurev.fl.21.010189. 0.745 transonic truss-braced wind design, in: AIAA SciTech 2020 Forum, AIAA
001315. Paper 2020-1649, Orlando, FL, January 2020.
[26] J.C. Vassberg, M.A. DeHaan, S.M. Rivers, R.A. Wahls, Development of a com- [40] P.A. Henne, R.D. Gregg, New airfoil design concept, J. Aircr. 28 (5) (1991)
mon research model for applied CFD validation studies, in: 26th AIAA Applied 300–311, https://doi.org/10.2514/3.46028.
Aerodynamics Conference, AIAA Paper 2008-6919, Honolulu, HI, August 2008. [41] M. Drela, A User’s Guide to MSES 3.05, MIT Department of Aeronautics and
[27] R.L. Campbell, M.N. Lynde, Natural laminar flow design for wings with moder- Astronautics, July 2007.
ate sweep, in: 34th AIAA Applied Aerodynamics Conference, AIAA Paper 2016- [42] M.D. Maughmer, D.M. Somers, Design and experimental results for a high-
4326, Washington, DC, June 2016. altitude, long-endurance airfoil, J. Aircr. 26 (2) (1989) 148–153, https://doi.org/
[28] M.N. Lynde, R.L. Campbell, Computational design and analysis of a transonic 10.2514/3.45736.
natural laminar flow wing for a wind tunnel model, in: 35th AIAA Applied [43] R. Eppler, Airfoil Design and Data, Springer-Verlag, Berlin, 1990.
Aerodynamics Conference, AIAA Paper 2017-3058, Denver, CO, June 2017. [44] R. Eppler, Airfoil Program System “PROFIL11” User’s Guide, 2011.
[29] H. Görtler, On the Three-Dimensional Instability of Laminar Boundary Layers [45] M.S. Selig, M.D. Maughmer, D.M. Somers, Natural-laminar-flow airfoil for
on Concave Walls, NACA TM 1375, 1954. general-aviation applications, J. Aircr. 32 (4) (1995) 710–715.
[30] J.D. Crouch, V.S. Kosorygin, Surface step effects on boundary-layer transition [46] M. Fujino, Y. Yoshizaki, Y. Kawamura, Natural-laminar-flow airfoil development
dominated by Tollmien-Schlichting instability, AIAA J. Preprint (2020), https:// for a lightweight business jet, J. Aircr. 40 (4) (2003) 609–615.
doi.org/10.2514/1.J058518. [47] M. Drela, Implicit implementation of the full e N transition criterion, in: 21st
[31] A.M.O. Smith, High-lift aerodynamics, J. Aircr. 12 (6) (1975) 501–530, https:// AIAA Applied Aerodynamics Conference, AIAA Paper 2003-4066, Orlando, FL,
doi.org/10.2514/3.59830. June 2003.
[32] A. Bertelrud, Transition Documentation on a Three-Element High-lift Configu- [48] M. Drela, M.B. Giles, Viscous-inviscid analysis of transonic and low-Reynolds
ration at High Reynolds Numbers - Analysis, NASA/CR-2002-211438 2002. number airfoils, AIAA J. 25 (10) (1987) 1347–1355, https://doi.org/10.2514/3.
[33] K.W. Noonan, W.T. Yeager, J.D. Singleton, M.L. Wilbur, P.H. Mirick, Wind Tunnel 9789.
Evaluation of a Model Helicopter Main-Rotor Blade with Slotted Airfoils at the [49] P. Camacho, K. Pham, L. Chou, N. Harrison, A. Khodadoust, Progress on aero-
Tip, NASA/TP-2001-211260, 2001. dynamic performance analysis of SNLF transonic truss-braced wing, in: AIAA
[34] D.S. Lacy, A.J. Sclafani, Development of the high lift common research model SciTech 2020 Forum, AIAA Paper 2020-1025, Orlando, FL, January 2020.
(HL-CRM): a representative high lift configuration for transonic transports, in:

11

You might also like