You are on page 1of 19

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/242741331

Airfoil Geometry Parameterization Through Shape Optimizer and


Computational Fluid Dynamics

Article · January 2008


DOI: 10.2514/6.2008-295

CITATIONS READS
15 651

3 authors, including:

Manujit Khurana Hadi Winarto


Chandigarh University RMIT University
10 PUBLICATIONS 163 CITATIONS 17 PUBLICATIONS 267 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Hadi Winarto on 22 February 2017.

The user has requested enhancement of the downloaded file.


Airfoil Geometry Parameterization through Shape
Optimizer and Computational Fluid Dynamics

Manas S. Khurana1, Hadi Winarto2 and Arvind K. Sinha3


The Sir Lawrence Wackett Aerospace Centre - RMIT, Melbourne, VIC, 3000

Re-Configurable Multi Mission Unmanned Aerial Vehicle (RC-MM-UAV) emerging


design concept involves morphing wings with mission segment based airfoils. To identify the
appropriate planform, the Direct Numerical Optimization (DNO) methodology is applied in
the design. The DNO comprises of two phases; a) Airfoil shape parameterization method to
develop candidate shapes; and b) Validated flow solver to compute the aerodynamic forces.
The design parameterization study involves testing the accuracy and flexibility of five shape
functions by measuring the geometrical difference between a set of benchmark airfoils and
the approximated solution through the shape functions. A Particle Swarm Optimizer (PSO)
was utilized and with the integration of a local line search algorithm, geometrical
convergence is established. The second phase of DNO examined the validity of a high fidelity
flow solver by comparing computed data against experimental solutions. A Reynolds
Average Navier Stokes (RANS) model with a multi-zonal fluid grid was developed to
duplicate flow transition, and provided lift and drag variance of 3% and 4% respectively
over an angle-of-attack range -1° to 12°. Further research will involve integration of the
above results to support the design decision of RC-MM-UAV airfoils.

Nomenclature
λ = Design Variable fi = Solution Fitness
( x / c) i = Chord Length Interval f i (x) = Shape Functions (Polynomials)
[λ1 , K , λ n ] = Cost / Fitness of Each Design Variable ƒtarget = Target Airfoil
ak = Scalar Step Length k = Iteration
c = Chord Length m = Swarm Population Size
c1, c2 = Learning Factors n = Design Variable Population
cd = Profile Drag Coefficient rand = Uniform Random Value [0,1]
cf = Coefficient of Friction vi = Velocity of Particle
cl = Profile Lift Coefficient w = Inertia Weight
cp = Coefficient of Pressure xi = Current Position of Particle
ƒapprox. = Approximated Airfoil ∆ƒ = Airfoil Geometrical Difference
(Objective Function)

1
PhD Candidate, The Sir Lawrence Wackett Aerospace Centre, 850 Lorimer Street, Port Melbourne, VIC 3000,
Australia - AIAA Student Member
2
Associate Professor, School of Aerospace, Mechanical and Manufacturing Engineering, GPO Box 2476V, 3001,
VIC, Australia - AIAA Member
3
Director Aerospace and Aviation, The Sir Lawrence Wackett Aerospace Centre, 850 Lorimer Street, Port
Melbourne, VIC 3000, Australia - AIAA Member

1
American Institute of Aeronautics and Astronautics
I. Introduction

T HE Sir Lawrence Wackett Aerospace Centre pioneered a conceptual design of Re-Configurable Multi-Mission
Unmanned Aerial Vehicle (RC-MM-UAV) for futuristic civil and military operational needs. Uni mission
UAVs presently are under various stages of design, development and trials as an alternative over manned systems to
address dull, dirty and dangerous missions. Future operational doctrines merit the development of a Multi-Mission
UAV to cover a wide mission spectrum. Though uni-mission designs address a certain section of the requirements,
the performance is limited; in regards to range/endurance and speeds. A uni-mission UAV with fixed wing geometry
design addresses a specific performance requirement of speed and endurance. RC-MM-UAV concept will
encompass long endurance/range requirements of Intelligence, Surveillance and Reconnaissance (ISR) and the
flexibility of speed for Suppression of Enemy Air Defence (SEAD) sorties in its mission profile – a typical futuristic
operational scenario. Multi-Mission profile requires variations in speeds, altitude of operation and maneuverability
to address disparate performance requirements.
UAV operations in Australia have been restricted to atmospheric monitoring and aerial photography with
mineral exploration emerging as a new sector. A detail market survey identified the requirement of a Tier of UAVs
based on future operational requirements (Table 1).2 The Australian Defence Force’s – AIR 7000 Request-for-
Proposal calls for the development of a Multi-Role UAV for introduction within the current fleet, and the JP129
project for Tactical Unmanned Aerial Vehicle for Intelligence, Surveillance, Target Acquisition and Reconnaissance
Capabilities. The Australian program on UAV acquisition covers the request of a High-Altitude Long Endurance
UAV for Reconnaissance and Surveillance, thus the development and introduction of an intelligent UAV platform is
identified as a key operational asset.
Table 1. UAV Tier Classification based on Operational Capabilities
UAV Category Designation Max. Alt [ft] Range [km] Endurance [hrs]
Tier I Interim Medium Altitude, Up to 15,000 Up to 250 5-24
Endurance
Tier II Medium Altitude, 3,000 - 25,000 900 More than 24 hrs
Endurance
Tier II Plus High Altitude, Endurance 65,000 Up to 5,000 Up to 42 hrs
Tier III Minus Low Observable – High 45,000 - 65,000 800 Up to 12 hrs
Altitude, Endurance

The pioneering RC-MM-UAV design is a viable concept to address the identified requirements in the market
survey. The re-configurable modular concept of wings and payload considered a total of 52 missions and covered
high to medium altitude long endurance of 25-42 hours and range in excess of 4,000 nm at Mach numbers 0.32-0.60.
A major section of missions (total of 32) shared commonality in flight performance and mission specific payloads
with a maximum gross take-off weight and wing area envelope of 4,000-10,000 kg and 49-68m2 respectively.
The investigation on wing extension options and its design and operational merits resulted in the morphing
concept being considered for further detailed study. Mid-flight wing planform flexibility through the use of smart
actuators / materials will provide multi-mission capabilities in the mission profile from slow speed ISR roles to
quick dash segments for SEAD operations.
In this paper a Direct Numerical Optimization (DNO) methodology to facilitate the design of mission segment
based airfoils for RC-MM-UAV is proposed. The methodology (Fig. 1) comprises of the following: a) Mathematical
shape function to represent potential airfoils; b) Flow solver validation for aerodynamic computation; and c)
Intelligent search agent for overall optimisation.

Figure 1. Direct Numerical Optimization Approach for Airfoil Analysis


2
American Institute of Aeronautics and Astronautics
II. Problem Definition
An intelligent system is required for the development of airfoils and the DNO approach is identified for design
and analysis. The sub-systems of the DNO process (Fig. 1) need to incorporate design flexibility to support disparate
requirements of multirole mission packages for the development of airfoil planforms. An airfoil function is required
to provide the flexibility to explore realistic shapes, and the flow solver to compute the aerodynamics over a wide
mission operating envelope, within the identified accuracy tolerance. The issue of airfoil development to address
mission flexibility is defined by the following approach:

A. Airfoil Shape Function


Airfoil shape development involves formulating a geometrical function to represent candidate planforms within
the DNO process. The shape function is to provide the following design flexibility:
ƒ Provide smooth and realistic airfoil shapes;
ƒ Computationally efficient with the ability to handle a large chordwise distribution for surface smoothness;
ƒ Few design variables with the ability to cover a wide design search space;
ƒ Provide independent one-to-one control for camber, thickness, leading and trailing edge angles, with the
flexibility to represent divergent trailing edge sections; and
ƒ Design variables be directly related to each geometrical feature such that design constraints can be imposed
within the optimization routine.
Several methods including Discrete, Polynomial and Spline, Conformal Mapping3 and Analytical4 methods can be
used to produce airfoil sections and are design application specific. A detail case study is required to determine a
suitable method that addresses the aforementioned points through an optimization study (section III).

B. Flow Solver Validation


A flow solver validated against experiment data is required within the DNO process. The drag build up over an
airfoil surface is dependent on laminar to turbulent transition, to possible flow separation and re-attachment. The
operating Reynolds and Mach number have an influence on these features, with the development of the boundary
layer dominating the drag build up over the lifting surface. Development and validation of a computational model to
accurately predict the complicated flow features over aerodynamic bodies is being addressed by NASA’s Drag
Prediction Workshop and remains a challenging task.
To capture the complex airfoil flow features, a high fidelity solver is used over conventional panel methods.
Utilization of Computational Fluid Dynamics (CFD) as a flow solver tool requires extensive validation. Thus, an
appropriate testing domain to determine the effect of grid density with wall normal distance y + , and turbulence
modeling is required to obtain computational data that is comparable with published experimental solutions over a
linear angle-of-attack range. The domain will then be applied to calculate the aerodynamic coefficients of the airfoils
produced by the shape functions and within the overall DNO process.

III. Airfoil Shape Parameterization

A. Geometrical Shape Functions


The discrete airfoil parameterization approach uses coordinates to represent shapes. Each vertex acts as a design
variable for shapes to be implemented.4 A wide search space is viable as points are individually controlled to
provide the required degree-of-freedom. The method is unsuitable for airfoil optimization, as a large population of
points is needed to generate a smooth surface, resulting in higher costs associated with design optimization
computations. The cost issue is addressed by imposing a fixed boundary condition. A set of points is static, and a
select few act as variables. The practice is ineffective for global search routines as the planforms produced is
restricted in shape variation due to the fixed degree-of-freedom, over certain regions. The generation of unrealistic
shapes is viable, particularly at the boundary where fixed and variable degree of freedom points coincide. Certain
shapes will lead to instabilities within the flow solver due to the unrealistic representation of the planform.5
The spline and polynomial approach was developed to address the issues of the discrete method. A Bezier
representation accurately represents simple shapes, though a higher-degree form is required to represent complex
shapes with higher round-off error6; defined as the geometrical difference between the approximated and actual
airfoil curve. A B-Spline method uses lower-order Bezier segments to represent a curve and demonstrated accurate
airfoil shape convergence. It is computational intense7 and does not represent implicit conic sections accurately.6 A
Non-Uniform Rational B-Spline method is readily used for shape representation. It requires a large set of control

3
American Institute of Aeronautics and Astronautics
points to generate a smooth geometry.6 Undulating sections will be present during the optimisation search, leading
to additional computational time.
Conformal mapping techniques have been previously analyzed including Joukowski and Kármán-Trefftz
transformations for airfoil development. This mapping technique requires fewer design variables to represent a
relatively large set of airfoils.3 The Joukowski mapping through two design variables controls camber and thickness-
to-chord ratio. The Kármán-Trefftz transformations are introduced with an additional variable to control the trailing
edge angle.3 The conformal mapping technique is unsuitable for multi-objective optimization problems, due to
geometrical limitations within the design space.8 The transformations are unable to control airfoil nose radius and
will not produce divergent trailing-edge airfoils, which provides improved airfoil effectiveness in terms of increased
thickness with enhanced lift capability at a lower drag performance.9 Additionally, it is limited in the chordwise
flexibility for the placement of maximum thickness and camber locations. Thus, the advantage of a lower
computation time is not justified against the limitations present in controlling important airfoil features.
Wu et al. examined the PARSEC, Hicks-Henne and mesh-point geometrical functions to represent cascade
blades through an inverse design process. A base airfoil was modified through the different shape functions to match
the pressure distribution with the benchmark pressure coefficient of a target airfoil.10 The investigation confirmed
that the selection of a method is governed by the design requirements and its subsequent application.10

B. Test Methodology
A geometrical shape parameterization investigation with an analytical approach is needed for the design of long
endurance airfoils. Methods examined include the Hicks-Henne,11 Wagner,12 Legendre,* Bernstein,† and NACA
normal modes13, 14. The polynomials represent a series of sinusoidal curves that are generated across a specified
chord length interval. The technique operates by adding a finite sum of closed shape functions to an initially
specified airfoil shape to generate a target section. The design variables ' λi ' , are multipliers to shape functions and
determine the contribution of each function to the final shape.4 Mathematically it is represented as follows:

n
y ( x / c, λi ) = y ( x / c ) initial +
airfoil
∑ λ f ( x / c)
i =1
i i (1)

Where:
λi = Design Variable f i ( x / c) = Shape Functions
n = Design Variable Population Size ( x / c) i = Chord Length Interval
A symmetrical NACA 0015 profile is used as the
Airfoil Profiles for Geometrical Shape Parameterization
benchmark. Three target sections for the shape
Base: NACA 0015
convergence test, were considered with low speed and Target 1: NASA LRN(1)-1007
0.15 Target 2: NASA LS(1)-0417Mod
long endurance, being a critical requirement for RC- Target 3: NASA NLF(1)-1015

MM-UAV. As the target sections cover a wide 0.1

geometrical layout (Fig. 2), the validation process


provides the avenue to test the flexibility and accuracy 0.05
y/c

of the shape functions. The test iterates the design


0
variables to modify the initial section, for the norm of
the residual error between the target and approximated -0.05

section is set in the acceptable tolerance limit. The


issue is an optimization routine, and the objective -0.1

function requires minimizing the cost defined as the


geometrical difference between approximated and 0 0.1 0.2 0.3 0.4 0.5
x/c
0.6 0.7 0.8 0.9 1

target airfoil. This is mathematically represented as Figure 2. Airfoils for Shape Parameterization
follows: Optimization.
⎡ ⎤
∆ ƒ min ∑ ⎢( y / c )i − ( y / c )i ⎥ (2)
⎣ approx t arg et ⎦

*
http://mathworld.wolfram.com/LegendrePolynomial.html [cited 2 May 2007]

http://mathworld.wolfram.com/BernsteinPolynomial.html [cited 2 May 2007]

4
American Institute of Aeronautics and Astronautics
Where:
∆f = Objective Function ( y / c )iapprox = Approximated Airfoil Profile ( y / c )it arg et = Target Airfoil Profile
Two search algorithms are used to minimize the objective function (Eq. 2). A conjugate gradient line search
method14 and a “probabilistic global seeking particle swarm optimizer”15-18 were applied to test and measure the
following: a) Flexibility of the shape functions by minimizing the objective function across the three target sections;
and b) Computational efficiency of the optimization models by recording the number of design iterations and clock
time required for solution convergence.
The pressure and suction surfaces are independently simulated over a design variable population size ‘n’, from 2
to 10 and then coupled to form the final shape. The testing process requires the following inputs for evaluation:
ƒ Base airfoil profile : (x / c )Base and ( y / c )Base ;
Airfoil Airfoil
ƒ Target section profile : ( y / c )T arg et in line with (x / c )Base distribution; and
Airfoil Airfoil
ƒ Design variable size :‘n’
The Hicks-Henne methodology requires additional inputs to adjust the peak magnitudes (aH-H) and location (pii)
of each contour that are prior-adjusted.6 The proposed test methodology is presented in Fig. 3.

Figure 3. Airfoil Geometrical Parameterization – Test Methodology Set up


The shape functions examined are presented in Table 2. Each method requires user defined design variable
population, with the optimizer computing the corresponding variable magnitude λi for shape convergence. The
optimizer additionally evaluates aH-H and pii points of each contour for the Hicks-Henne modulation as equated by
size n.
Table 2. Polynomial Shape Functions Definition
Polynomial Function Optimization Variables Symbol
Design Variable Population n
[(α H − H1 ,L,α H − H n )]
f i ( x) = sin α H − H (πx βi ) ; Peak Contour Abscissa
Hicks-Henne Peak Contour Magnitude
where β i = ln(0.5) / ln( p ii ) ; where i = 1, n [( pii1 ,L, piin )]
Magnitude of Design Variable/s [λ1,K, λn ]
θ + sin(θ ) ⎛θ ⎞
f1 ( x ) = − sin 2 ⎜ ⎟ ; θ = 2 sin −1 ( x )
π ⎝2⎠
Wagner Design Variable Population n
sin(nθ ) sin[(n − 1)θ ] Magnitude of Design Variable/s [λ1 , K , λ n ]
f i ( x) = + for i = 2, n
iπ π

( )
i
1 ⎛ d ⎞ 2 i
Legendre f i ( x) = ⎜ ⎟ x − 1 ; where i = 1, n Design Variable Population n
2 i i! ⎝ dx ⎠ Magnitude of Design Variable/s [λ1 , K , λ n ]
⎛n⎞
f n,i ( x ) = ⎜⎜ ⎟⎟ x i (1 − x) n −i ;
⎝i ⎠ Design Variable Population n
Bernstein
⎛n⎞
where ⎜⎜ ⎟⎟ =
n!
, i = 0, n ; where i = 1, n
Magnitude of Design Variable/s [λ1 , K , λ n ]
⎝ i ⎠ i!(n − i)!
f1 ( x) = x − x
Design Variable Population n
f 2 ( x) = x(1 − x)
NACA Magnitude of Design Variable/s [λ1 , K , λ n ]
f i +1 ( x) = x i (1 − x), for i = 2,3,4,5

5
American Institute of Aeronautics and Astronautics
C. Airfoil Shape Optimization
The two optimization models proposed operate with unique search strategies. Gradient methods are
computationally efficient and provide rapid solution convergence if the starting region is well defined. When a
function requires adjustment of multiple variables, local methods settle into a local minima. Hence, evolutionary
programming model needs to be adopted.
A stand-alone PSO model is used for shape convergence across the five functions proposed and convergence
performance was measured against the line search method for Wagner, Legendre, Bernstein and NACA normal
modes (Table 3). As the Hicks-Henne method involves computing multiple functions (Table 2), a PSO model is
used to obtain a starting point. The results are used as an input to the line search and a hybrid methodology is
proposed (Table 3). The test evaluates which polynomial is most appropriate across the three target sections and the
effectiveness of the gradient, global, and the combination of both global and local search agents based on
geometrical and aerodynamic convergence is evaluated.
Table 3. Airfoil Shape Optimization Tools
Polynomial Optimizer Utilized Type
ƒ PSO ƒ Global
Hicks-Henne
ƒ Hybrid – PSO / Line Search ƒ Global / Local
ƒ PSO ƒ Global
Wagner
ƒ Line Search ƒ Local
ƒ PSO ƒ Global
Legendre
ƒ Line Search ƒ Local
ƒ PSO ƒ Global
Bernstein
ƒ Line Search ƒ Local
ƒ PSO ƒ Global
NACA
ƒ Line Search ƒ Local

1) Line Search Method


The line search methodology from MATLAB calculates the gradient of the objective function at each iteration k,
through a higher order Quasi-Newton method.14 The convergence criterion is based on a constant objective function
over a 30 iteration count with 10-6 as the degree of precision. The algorithm uses the partial derivative of the
objective function ∇f ( x * ), (Eq.4) where approximations to the Hessian H, are based on the Quasi-Newton method.14

∇f ( x * ) = Hx * + c ≈ 0 (4)

Where:
x ∗ = Optimal Solution c = Constant Vector H = − x* c
−1

The gradient of the objective function through the perturbation of design variables is obtained with a finite
differences method, to refine the search towards a minimum solution. The line search algorithm uses a pre-defined
scalar step length a k , to direct the search p k , with a decreasing objective function, after each iteration x k . The
iterative process is based on Wolfe’s method which operates along the line x k + a k p k , until the convergence criteria
is established.14 The line search algorithm and the search mode are expressed as follows:

x k +1 = x k + a k p k (5)

p k = − H k−1 .∇f ( x k ) (6)

6
American Institute of Aeronautics and Astronautics
2) Particle Swarm Optimization
The PSO methodology is based on the evolution of social behavior where the solutions are represented by a set
of particles that heuristically navigate through a design space.15 Unlike Genetic Algorithms, the effectiveness of a
potential solution is not dependent on crossovers and mutations. It is updated based on an information sharing
methodology with the population of particles.
Each particle in swarm of size m, is a potential solution of the objective function (Eq. 2) and is represented by
the relative position xi , and velocity vi . The optimization in initiated through the random dispersion of the particles
with each variable n, set within a domain of minimum and maximum limit (Fig. 4). The
position [ xi = ( xi1, xi 2,..., xin ), i = 1,2,..., m] and velocity [vi = (v i1, vi 2,..., vin ), i = 1,2,..., m] of the individual particles is
recorded and solution fitness evaluated (Eq. 2) and stored [ Pgbest = ( Pg1, Pg 2,..., Pgn )] (Fig. 4). A particle with the
‘best’ global solution, ' Pgbest ' , from the population, is recorded and the remaining particles update their position and
velocity to follow Pgbest over the subsequent iterations until solution convergence (Fig. 4).19
At each iteration ' k ' , the search direction is refined by updating the position, ' xi (k + 1)' , and velocity, ' vi (k + 1)' ,
(Eqs. 7-8); where the velocity rate of change is a function of user pre-defined learning factors as follows: a)
Cognitive c1 and b) Social c 2 parameters that influence local and global search patterns;17 and c) Constriction weight
factor (Eq. 9) for solution convergence, by neutralizing a balance between the local and global search regions (Fig.
4).15,18

xi (k + 1) = x i (k ) + v i (k + 1) (7)

v i (k + 1) = w.v i (k ) + c1 .rand .( Pi − x i (k )) + c 2 .rand .( Pg − x i (k )) (8)

2
w= ; where ϕ = c1 + c 2 (9)
2 − ϕ − ϕ 2 − 4ϕ

The solution dimensional space is governed by the constraints of the design variables. The velocity of the
particles are controlled, through a scalar multiplier (Eq. 8) to limit the maximum velocity and avoid particles
overshooting the computational domain.19 Despite the velocity restrictions, particles conjugate outside the design
domain and the application of boundary conditions is needed to re-direct the particles within the search space.19 The
absorbing, random and reflecting conditions are used for such purposes. A global search pattern is preferred to cover
a wide search region. The absorbing technique offers higher local search capabilities, as the particles are imposed
into the boundary of the computational domain. It will offer rapid convergence for instances where the solution is
within the wall region. Conversely, if a local minimum is present, the particles may be trapped within the boundary,
and will ultimately diverge over a progression of iterations, leading to an increase in computational requirements.
The random initialization condition, disperses solution space exceeded particles within the search domain, thus re-
instating a heuristic search process. The reflecting wall technique will promote a linear search pattern, as the
particles are reflected in the opposite direction to the original search path. This method offers rapid convergence if
the global minimum lies in the original search path and is ignored due to the overshooting effect of the particles and
high velocity. The random and reflecting conditions were independently simulated and tested in this paper, to
determine which method is appropriate for airfoil shape convergence.
The inputs to the PSO algorithm define the solution search behavior and together with the imposed wall
boundary conditions, the robustness of the model is defined (Table 4). The boundary wall condition test is simulated
over a single shape function and covered in section IV of this paper.

7
American Institute of Aeronautics and Astronautics
Table 4. PSO Inputs and Constraints
Input / Parameters Value
Fixed Variables: Minimum Maximum
Cognitive - c1 ---- 25 ----
Social - c2 ---- 10 ----
Swarm Population - m ---- 2,000 ----
Velocity Scalar Multiplier ---- 0.1 ----
Continuous Variables:
Shape Function Population Size - n 2 10
Shape Function Design Variables Limit - λn -0.05 0.05
Hicks-Henne Geometrical Features
αH-H 0 5
pii 0 1
Algorithm Set-Up
Boundary Conditions: a) Random Initialization
b) Reflected Wall
Convergence Criteria:
Constant Gradient Run 15
Degree of Precision 10-6
Maximum Iterations 400

The optimization definition in Table 4, is presented in the following flowchart (Fig. 4). The cognitive c1 and
social c2 parameters, together with swarm size are user defined and required to initiate the search process. Swarm
population effects convergence rate, with high values resulting in greater computation efforts, while may lead to
improved solution fitness. A valid combination of these inputs, allows for an overall effective search algorithm.

Figure 4. Shape Optimization Flowchart

* For Bernstein, Legendre, NACA and Wagner polynomials, the final PSO solution is stored in array Pbestg, as
[λ1 , K , λ n ] . The solution to the Hicks-Henne function is stored in Pbestg as [(α H − H1 , L , α H − H n ), ( p ii1 , L , p iin )] and
used as an input into the line search algorithm for an initial estimate to derive the solution [λ1 , K , λ n ]

8
American Institute of Aeronautics and Astronautics
IV. Geometrical Shape Optimization Results and Discussion
This section covers convergence data by the PSO algorithm with random and reflected boundary wall conditions,
line search and hybrid method (Table 3). The total cost presented, is related to the fitness of the objective function
(Eq. 2), and is the geometrical difference between target and approximated planform. The cost magnitude
incorporates modeling upper and lower airfoil surfaces over a variable size of 2-10. A complete airfoil will
incorporate twice the original variables of 4-20.

1) Particle Swarm Optimization


As PSO is a probabilistic search agent due to the random initialization of the particles, 5 runs were performed for
each case study and the minimum geometrical cost recorded. The effect on solution convergence for random
initialization and reflected wall conditions is examined. To determine the applicability of the methods, an
independent case study on the convergence of the Hicks-Henne over the three target airfoils (Fig. 2) was performed.
The study examined the effect of boundary condition as a function of solution dimensional space through a
population size of n = 4 and 20. The NASA LRN(1)-1007 convergence evolution over the pressure and suction
surfaces is presented in Fig. 5-6 and the inset illustrates the final approximated section. It indicates (Fig. 5) that for a
small dimensional space, the reflected condition yields a lower cost over both surfaces in comparison to random
initialization, at a higher computational expense. As the solution sub-set is restricted with lower bounds n, the
probability of locating the minima is improved through the oscillation of particles between the outermost boundary
region and the inner space of the solution space. Due to the overshoot response of the particles, the reflected
boundary condition reflects the particles back into the solution space, and consequent iterations tend to oscillate
about this point for convergence. Thus, it leads to a higher computational expense. Conversely, for a larger solution
space, random initialization methodology is superior (Fig. 6) as it encourages a broader search routine with the
added degree of freedom to cover a wide design space. As the search space increases, the probability of the solution
being within the wall boundary is significantly less than for a restricted solution set. Consequently, random
initialization of particles is feasible, to explore a wide region and not restrict the movement within the boundary.
Similarly, the reflected boundary condition (Fig. 6), exhibits a larger computational expense due to the oscillation of
particles within the restricted boundary region.
Effect of Boundary Conditions on Geometrical Convergence Effect of Boundary Conditions on Geometrical Convergence
0.7
0.1

0.09 0.1
0.09
0.09
0.08 0.09
0.6 0.08
0.07
0.08 0.25 0.08
0.07
0.07
0.06
0.07
0.06
0.06
0.05
0.5
y/c

0.06
y/c

0.05 0.05
0.04
y/c

0.2
y/c

0.05
Objective Function(∆ f)

0.04
Objective Function(∆ f)

0.03 0.04

0.03 0.04
0.02 0.03
0.4 0.02 0.03
0.01 0.02
Target Target
0.01
0
Approx.
0.15 0.01
0.02
Approx.
Target Target
0
-0.01 Approx. 0.01 Approx.
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0.3 -0.01 x/c
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 x/c
x/c 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
x/c

0.1
Random: Pressure Surface
0.2
Random: Suction Surface Random: Pressure Surface
Random: Suction Surface
Reflected: Pressure Surface
Reflected: Pressure Surface
Reflected: Suction Surface 0.05
0.1 Reflected: Suction Surface

0 0
0 20 40 60 80 100 120 140 160 0 50 100 150 200 250 300 350 400 450
Iteration Iteration

Figure 5. NASA LRN(1)-1007: Effect of Boundary Figure 6. NASA LRN(1)-1007: Effect of Boundary
Condition on Objective Function; n = 4 Condition on Objective Function; n = 20
The convergence evolution of the three target sections in Fig. 2, over maximum and minimum solution set, n is
presented in Table 5 through the Hicks-Henne shape functions, with boundary wall condition variations. The shaded
regions indicate lower costs between the two conditions examined. In a solution set with 4 variables, the reflected
condition is superior for all three airfoils over the suction side and provides lower cost distribution on the pressure
side for only one of the three airfoils. Hence, the reflected condition indicates superior convergence over the random
methodology for this test case. The search process is not compromised with local search pattern, an attribute of the
reflecting methodology due to the restricted search space. The method is computationally inefficient across the
examined solution space, due to the possible oscillation of particles between the inner and outer boundaries which
only settle when the convergence criteria is satisfied. For large search domain (n = 20), the random methodology
exhibits superior performance in all cases (Table 5). A global search process is required for multi-dimensional
solution domains, and the random initialization of particles, promotes this search pattern. A linear/local search

9
American Institute of Aeronautics and Astronautics
process through the reflecting condition, was ineffective in this case, thus suggesting the requirement of
implementing the random methodology, for optimization studies with large population of n.
Table 5. Effect of Boundary Condition on Hicks-Henne Geometrical Convergence
∆f Random ∆f Re flected ∆f Random ∆f Re flected k Random k Re flected k Random k Re flected
Airfoil Suction Suction Pr essure Pr essure Suction Suction Pr essure Pr essure

Design Variable Population Size; n = 4


NASA LRN(1)-1007 0.0996 0.0389 0.0655 0.0170 45 78 65 158
NASA LS(1)-0417 Mod 0.0330 0.0272 0.0219 0.0409 83 344 51 59
NASA NLF(1)-1015 0.0613 0.0186 0.0451 0.0619 70 77 83 88
Design Variable Population Size; n = 20
NASA LRN(1)-1007 0.0122 0.0146 0.0084 0.0154 256 401 124 231
NASA LS(1)-0417 Mod 0.0242 0.0276 0.0161 0.0278 203 204 79 401
NASA NLF(1)-1015 0.0111 0.0129 0.0116 0.0125 401 401 175 384

The reflected wall boundary condition is 7


Geometrical Shape Convergence - PSO Algorithim

effective for a limited solution set (Fig. 5-6). It is NASA LRN(1)-1007


NASA LS(1)-0417 Mod
restricted in design flexibility due to the failure in 6 NASA NLF(1)-1015

providing lower cost in comparison to the random


initialization scheme for large search domains 5 Objective Function ( ∆f)

(Table 5). Thus, the convergence test of the


4
remaining shape functions through the PSO
methodology will be based on the random
3
initialization condition, since a balance between
design accuracy and flexibility is required and the 2
random methodology provides this freedom.
The fitness evaluation over n = 4,6,8...20, of the 1

target airfoils (Bernstein, Hicks-Henne, NACA and


Wagner shape functions) based on the random 0
Bernstein Hicks-Henne NACA Wagner
initialization methodology is presented in Fig. 7. Shape Functions

The Hicks-Henne provides the lowest fitness for the Figure 7. Geometrical Cost Magnitude – PSO
three airfoils over the balance parameterization Random Initialization Methodology
methods. The Legendre was omitted from the analysis due to the computational expense attributed to the derivation
of a factorial and derivative term per iteration. This included small design variable population size of n =2, for
which total convergence time of 8 and 14 minutes (NASA LRN(1)-1007 and NASA LS(1)-0417 Mod airfoil) was
recorded respectively on a CPU running at 2.20GHz with 1.0 GB of RAM.

2) Line Search / Particle Swarm Optimization


Geometrical Shape Convergence - Line Search Cost Evaluation
The test methodology through the line search 8

algorithm was tested to study the effect of utilizing NASA LRN(1)-1007


NASA LS(1)-0417Mod
7
gradient search methods. Solution convergence across NLF(1)-1015

the five shape functions is presented in Fig. 8. The results 6


Hybrid:
Objective Function ( ∆f)

indicate a similar convergence pattern, with no


5 Line Search /
significant fitness reduction between the two optimizers
PSO
for the Bernstein, NACA and Wagner polynomials. The 4

Hicks-Henne model with the line search/PSO technique 3


(Fig. 8), indicate performance improvement in
comparison to a stand-alone PSO model (Fig. 7). 2

The improved performance of the hybrid optimization 1


method with Hicks-Henne function, is attributed to the
adjustment of constants αH-H and pii with PSO (Table 2). 0
Bernstein Hicks-Henne Legendre NACA Wagner
Shape Functions
Due to the presence of multiple variables, the solution
Figure 8. Geometrical Cost Magnitude – Line
space is complex for gradient search optimizers. As a
Search Methodology
result, the robustness of the search path for non-linear
solutions is influenced by the location of the initial guess. Due to the presence of multiple-minima’s, the probability

10
American Institute of Aeronautics and Astronautics
Hicks-Henne Polynomials: Effect of Optimization Methodology on Shape Convergence
1.4
NASA LRN(1)-1007
of the initial solution starting the search phase within a local NASA LS(1)-0417 Mod
NASA NLF(1)-1015
1.2
region is possible, and is influenced by the location of initial
guess. To address this, the PSO through the validated random 1

Objective Function (∆f)


initialization boundary condition, was applied to obtain an initial 0.8

estimate of αH-H and pii. These parameters were the inputs into the
0.6
line search from Fig. 4, to reduce the function from multi-to-
single variable, hence deriving the magnitude of the design 0.4

variables. The PSO provides a valid staring point for the line 0.2

search method, that operates by further reducing the cost as the


starting point is well defined. 0
Hybrid: Line Search / PSO
Optimizer
PSO

The analysis has shown that Bernstein, NACA and Wagner Figure 9. Comparison of Uni & Hybrid Based
functions encounter shape convergence issues (Fig. 7-8). The Optimization Technique for Hicks-Henne Function
Legendre polynomials were computationally inefficient, within
NASA LRN(1)-1007: Cost Function History through PSO and Hyrbid Methodology
the line search model, hence was not considered in the exhaustive 1
10
PSO: Suction Surface
search process of a PSO algorithm. The Hicks-Henne model, PSO / Line Search: Suction Surface
PSO: Pressure Surface
provided the lowest fitness of all the shape polynomials tested PSO / Line Search: Pressure Surface

(Fig. 7-8). Performance comparison between a stand-alone PSO 10


0

Objective Function ( ∆ f)
model and a hybrid technique indicate performance improvement
with the hybrid technique due to a lower cost distribution (Fig. 9).
This provides the avenue for further analysis. -1
10

ƒ Convergence Comparison between PSO and Hybrid-Line


Search / PSO Model
-2

Comparison between the independent PSO model and the 10

hybrid technique with the Hicks-Henne model, over the three 0 50 100 150 200 250
Iteration
benchmark airfoils is presented in Fig. 10-12. Convergence of the Figure 10. Hicks-Henne Fitness Convergence through the
NASA LRN(1)-1007 airfoil (Fig. 10), indicates the hybrid PSO & Line Search Model: NASA LRN(1)-1007
technique as computationally efficient with a lower cost and
NASA LS(1)-0417 Mod: Cost Function History through PSO and Hybrid Methodology
iteration count over the suction side. The hybrid technique 10
1

PSO: Suction Surface

provides a 3% improvement in fitness and 70% reduction in total PSO / Line Search: Suction Surface
PSO: Pressure Surface
PSO / Line Search: Pressure Surface
iteration count on the upper surface. On the lower surface, a
fitness reduction of 27% is observed with the hybrid technique, at 0
O bjective Function ( ∆ f)

10
a convergence rate approximately 20% greater than the standard
PSO run. The pressure side converges at a slow rate and is
attributed to strong geometric variations within the trailing edge.
The geometrical functions encounter complications in providing 10
-1

sufficient shape closure, thus reducing the speed of convergence.


Solution oscillations on both upper and lower airfoils surfaces are
present within the hybrid methodology in comparison to the -2
10
standard PSO run. This indicates the tendency of the line search 0 20 40 60 80 100
Iteration
120 140 160 180 200

method populating at local minimum. This is interrupted by the Figure 11. Hicks-Henne Fitness Convergence through the
PSO model that works to re-direct the search towards a global PSO & Line Search Model: NASA LS(1)-0417 Mod
solution, resulting in a fluctuating phase. Hence, the hybrid model
NASA NLF(1)-1015: Cost Function History through PSO and Hyrbid Methodology
is efficient in terms of fitness and iteration time for the NASA 2
10
PSO: Suction Surface
LRN section in comparison to the standard PSO model. PSO / Line Search: Suction Surface
PSO: Pressure Surface
The solution plot for the NASA LS(1)-0417 Mod airfoil (Fig. 10
1
PSO / Line Search: Pressure Surface

11) indicates quicker convergence in comparison to the analysis


Objective Function ( ∆ f)

presented in Fig. 10. The hybrid technique provides a lower


fitness over the two surfaces in comparison to a stand-alone PSO 10
0

model. The total iteration count between the two optimization


models is similar, with the exception of the suction surface with 10
-1

the PSO model, which converged later. The oscillation phase


between iterations is also dormant, thus the airfoil shape is
satisfactorily captured by the Hicks-Henne method. The fitness 10
-2

distribution of the airfoil is considerably lower than the 0 50 100 150 200
Iteration
250 300 350 400

magnitude observed over the low Reynolds number airfoil (Fig. Figure 12. Hicks-Henne Fitness Convergence through the
PSO & Line Search Model: NASA NLF(1)-1016
11
American Institute of Aeronautics and Astronautics
10). This is due to convergence evolution over the trailing edge region. Thus, the hybrid technique indicated lower
fitness in modeling the NASA LS airfoil. Iteration count between the two optimizers was also negligible.
The hybrid technique provides accurate geometrical convergence on the suction side of the NLF section (Fig.
12). A 25% reduction in fitness and 78% less computation iteration time is required with the hybrid method over the
PSO run in modeling the suction side (Fig. 12). Similarly, the pressure side is also modeled with fewer iterations
(37%), and a fitness reduction of 20% in comparison to a PSO model. Iteration fluctuations are similar to the trend
in Fig. 10, due to the presence of local minima’s within the solution set. Thus, the hybrid optimizer produces higher
fitness and iteration count in comparison to the analysis presented in Fig. 10-11.
The results presented in this section provides the following:
ƒ A PSO model with random initialization of particles boundary condition, provided accurate and rapid shape
convergence over the reflecting method (Fig. 5-6);
ƒ In single variable based functions (Wagner, Legendre, Bernstein, and NACA normal modes), gradient and global
search agents provided similar geometrical convergence (Fig. 7-8), with the line search methodology proving to
be computationally efficient;
ƒ The Hicks-Henne method provided the lowest fitness of all the shape functions tested, regardless of the
optimization technique utilized (Fig. 7-8);
ƒ A hybrid technique with a PSO model was used to obtain an initial estimate of αH-H and pii within the Hicks-
Henne function (Table 2). The parameters are then integrated into a line search model, that provided lower
fitness in comparison to a standard PSO algorithm, over the three target airfoils (Fig. 9); and
ƒ The benefit of the hybrid optimization technique, with the Hicks-Henne model, is shown in Fig. 10-12 across
the three benchmark airfoils, with lower fitness and rapid convergence over both upper and lower surfaces.

V. Aerodynamic Convergence
In this section the second unit of the DNO process is presented. It involves selection and validation of a suitable
flow solver. A high fidelity CFD solver was implemented in this study. The pre-processing analysis is in GAMBIT
and the computation process is in FLUENT. Both pre and post processing systems require a balance between
computational efficiency and solution accuracy. It is a compromise between coarse and fine mesh and choice of
turbulence modeling to achieve the required results.

A. Computational Model Set Up


A structured C-type grid was used in this study with the airfoil
chord c, normalized to 0.61m based on experimental set up.1 Grid
study involved a total of four grids ranging from a coarse and fine
mesh with 32,000 and 96,000 cells respectively. A distance of
1.0 × 10 −3 , between airfoil and first cell was set based on chord
length and flow Re, for the finest of the four meshes which provided
y + ≈ 150 (Table 6). A boundary layer consisting of 30 cells with a
growth factor of 1.05 was attached with cell aspect ratio of 5,
resulting in 160 nodes on the airfoil surface. The grid outside the
boundary layer was stretched by a ratio of 1.035 towards the
outward normal boundary, defined by a distance of 10c. The
pressure outlet far-field boundary was set to 18c with a stretch ratio
of 1.2 and 40 nodes and the distance from the velocity inlet to the
leading edge was 10c. A separate fluid zone was simulated at the
point of boundary layer transition at 0.075c on both pressure and
suction surfaces.1 High aspect ratios cells causes solution Figure 13. Computational Grid Set Up
divergence, particularly for higher angels of attack, when the flow is
not aligned with the cells. Thus, the aspect ratio of the cells within this transition region were reduced to 3, which
resulted in a total of 200 gird points on the nose region. The grid study involved varying the stretching ratio and the
total number of nodes within the computational domain in all directions (Table 6). These grids were used to compute
the flowfield about the airfoil and solutions compared to experimental data.
Airfoil aerodynamics is computed in FLUENT with a segregated solver coupled with a second-order upwind
finite differencing method. The continuity, momentum and turbulence intensity residuals are normalized to be less
then 10-5 for solution convergence. The pressure-velocity coupling is accounted through the SIMPLE algorithm and

12
American Institute of Aeronautics and Astronautics
the PRESTO interpolation method applied to calculate the face pressure within the segregated solver.20 Turbulence
modeling with the Spalart-Allmaras (S-A)21 and Menter’s Shear Stress Transport (SST) κ-ω22, 23 with enhanced wall
treatment, were independently applied in this study. These models were previously tested with success in predicting
the aerodynamics at pre-stall angles of attack, and the SST models were applied to compute post-stall trends.24-26
The velocity inlet boundary condition was defined with a turbulence intensity and viscosity ratio of 0.5% and 5
respectively, and the pressure outlet specified with a backflow turbulent intensity of 2% and a viscosity ratio of 20.

B. Validation Lift Curve Slope Validation


In this study, experimental results for the Modified 17- 2
Exp.
Percent-Thick Low-Speed Airfoil Section (NASA LS(1)-0417 1.8 Transition - Spalart-Allmaras
Transition - κ-ω
Mod)1 at Reynolds and Mach number condition of 6.0 × 10 6 and 1.6 Fully Turbulent (κ-ω )

0.32 respectively were used for validation. Experimental data 1.4


includes natural and forced boundary layer transitions; where the
1.2
later condition is simulated within CFD for the following reasons:

L
C
a) Flow solver is used to compute and compare the aerodynamic 1

forces of the approximated airfoils against the theoretical 0.8


solution; and b) Failure to capture laminar-to-turbulent transition,
0.6
will result in discrepancies between experimental and
computational data. Flow transition was manually induced in the 0.4

experiment1 and with the transition location known and 0.2


-2 0 2 4 6 8 10 12
implemented within CFD, this source of error was addressed. Angle of Attack α (° )

Manually tripping of the boundary layer provides an opportunity Figure 14. Lift Curve Slope Validation;
to test the suitability of adopting RANS within the overall DNO R e = 6.0 × 10 6 ; Mach = 0.32
process of airfoil design for a RC-MM-UAV. Drag Polar Curve Validation
The four grids generated were used in the validation process 2

over three separate case studies as follows: a) The κ-ω SST 1.8

model, with a fully turbulent flow condition; b) The κ-ω SST and 1.6

S-A model through a two-zonal fluid generation system based on 1.4


the tripping of the boundary layer according to experimental set
up.1 1.2 Exp.
Transition - Spalart-Allmaras
L
C

Transition - κ-ω
Data presented in Fig. 14-16, is based on grid that provided 1 Fully Turbulent(κ-ω)

minimum percentage difference of cl and c d over an angle-of- 0.8

attack domain of -1° to 12° in comparison to experimental data. 0.6

The lift curve evolution (Fig. 9) indicates the effectiveness of a 0.4


multi-zonal fluid grid where the two tripping models provide
almost identical operating performance and in line with 0.2
0.008 0.01 0.012 0.014 0.016 0.018 0.02 0.022 0.024 0.026 0.028

experimental solutions. The κ-ω SST model with transition


C
D

modeling, approximates lift that is within 3% of experimental Figure 15. Drag Polar Curve Validation;
data. The fully turbulent model exhibits lower lift and the onset of R e = 6.0 × 10 6 ; Mach = 0.32
flow separation is at higher angles of attack. The effectiveness of Effect of Turbulence Modelling on Coefficient of Pressure Convergence
-4
the three case studies was established from the drag polar (Fig. Exp.
-3.5 Transition: Spalart-Allmaras
15), where drag was within 4% of experimental solutions for the Fully Turbulent: κ-ω
κ-ω SST flow transition methodology over the testing angle-of- -3 Transition: κ-ω

attack range. The full turbulent condition over estimates drag by -2.5

approximately 26% and undermines lift by 5% in comparison to -2

experiment data. The wall pressure coefficient on suction and


P

-1.5
C

pressure sides is plotted in Fig. 16. The fully turbulent model -1


exhibits the lowest pressure peak at the leading edge of the
-0.5
suction side where the flow is laminar. Consequently the tripping
of the boundary layers through the two models generates a peak 0

pressure that is comparable to experimental solution. 0.5

The c p evolution of the computed models is slightly under 1


0 0.1 0.2 0.3 0.4 0.5 0.6
x/c
predicted aft of the leading edge on the suction side, up to Figure 16. Coefficient of Pressure Validation;
x / c = 0.4, with the two transition models generating almost
R e = 6.0 × 10 6 ; Mach = 0.32; α = 10°

13
American Institute of Aeronautics and Astronautics
identical results and in line with experimental solutions. Downstream of x / c = 0.4 and towards the trailing edge, the
tripping models provide acceptable solution agreement in comparison to the fully turbulent condition which
underestimates the aft loading of the section.
Based on the analysis over the three case studies, the κ-ω SST transition modeling scheme is most effective
through grid 4 as shaded in Table 6. The lift coefficient was underestimated across the four grids with the magnitude
of percentage difference remaining negligible, while the evolution of c d was within acceptable agreement. The total
average percentage difference over the testing angle-of-attack domain for lift and
drag, %∆clα [−1,0,K,12] and %∆c dα [−1,0,K,12] further indicated similar cl trend over the four grids, and c d difference was
minimal.
Table 6. κ-ω Grid Lift and Drag Coefficient Evaluation

Total Cell %∆ Cl %∆ Cd
Grid y+ Cl [α=10°] Cd [α=10°]
Count [α -1,0,…,12] [α -1,0,…,12]
Exp.1 ---- ---- 1.67 ---- 0.0155 ----
1 31,580 ≈ 200-450 1.59 (≈5% ↓) 2.4 0.0147 (≈5% ↓) 5.8
2 48,120 ≈ 200-450 1.60 (≈4% ↓) 2.3 0.0143 (≈8% ↓) 7.1
3 55,160 ≈ 50-160 1.58 (≈5% ↓) 2.6 0.0163 (≈5% ↑) 6.4
4 96,200 ≈ 150-275 1.57 (≈6% ↓) 2.7 0.0157 (≈1% ↑) 4.3

The onset of transition is modeled by visualizing skin 0.02


Effect of Turbulence Modelling on Coefficient of Friction

friction coefficient c f , and the mean velocity component in Transition: Spalart-Allmaras


0.018 Transition: κ-ω
coordinates locally normal v / U ∞ to the airfoil surface on the 0.016
Fully Turbulent: κ-ω

suction side. Experimental data1 was limited 0.014


to c l , c d , and c p evolution, thus the computed friction
0.012

coefficient and local velocity components was not evaluated


f

0.01
C

against solid data. The effect of boundary layer transition in


0.008
comparison to a total turbulent condition is evident (Fig. 17)
through the implementation of a two zonal fluid system. The 0.006

fully turbulent model exhibits a very sharp and sudden rise 0.004

of c f in comparison to the tripping models which indicate a 0.002

region of low and constant distribution from the leading 0


0 0.1 0.2 0.3 0.4 0.5 0.6
edge and downstream to ≈ x / c = 0.05; recalling that the x/c

boundary layer was manually tripped at x / c = 0.04575. 1 Figure 17. Suction Side Skin Friction Coefficient
Following the laminar region, the onset of turbulent flow is evident with the sudden increase in skin friction, where
the peak of the S-A model is in-line with the fully turbulent condition. The κ-ω SST model indicates an increase of
c f due to the onset of turbulent flow, although the gradient to the peak is considerably low and delayed. A decrease
in c f indicates a region of flow separation and further reattachment downstream, hence the delay. Upon
reattachment, c f rises to its maximum which is considerably lower than the two models presented.
The c f evolution of the three conditions is in-line with the computed lift and drag coefficients (Table 6). The
fully turbulent condition provides discrepancy in lift and drag based on experimental solution, while the tripping
term within S-A induces slight improvement due to the implementation of a laminar region (Table. 7). Drag
prediction with S-A remains high and is related to the c f evolution, which is similar to the fully turbulent condition
(Fig. 17).
Table 7. Effect of Turbulence Modeling on Lift and Drag Coefficients; α=10°

Turbulence Model Cl Cd
Exp.1 1.67 0.0155
Fully Turbulent: κ-ω SST 1.50 (≈10% ↓) 0.0214 (≈38% ↑)
Tripping: Spalart-Allmaras 1.59 (≈5% ↓) 0.0169 (≈9% ↑)
Tripping: κ-ω SST 1.57 (≈6% ↓) 0.0157 (≈1% ↑)

14
American Institute of Aeronautics and Astronautics
Wall Normal Velocity Magnitude
Local normal velocity component v / U ∞ , at four wall 2
x/c=0.03

stations x / c = 0.03, 0.05, 0.3, and 0.5 for the κ-ω SST model 1.8 x/c=0.05
x/c=0.3
x/c=0.5
is presented (Fig. 18) to show the effect of flow transition 1.6

and eventual separation. Under the influence of adverse 1.4

pressure gradient and with an increase 1.2

in c f at ≈ x / c = 0.05, the mean normal velocity is lower in

y/c
1

comparison to the profile near the leading edge 0.8

at x / c = 0.03 where the flow is undisturbed and laminar. 0.6

The tripping of the flow at ≈ x / c = 0.05 further induces low 0.4


velocity magnitude which continues to decrease
0.2
downstream of the leading edge until eventual flow reversal
0
with negative v / U ∞ . Thus, the effect of transition and -0.2 -0.1 0 0.1 0.2 0.3 0.4 0.5
v/U

separation are related to a decrease in velocity magnitude, Figure 18. Wall Normal Velocity Magnitude
and the onset of flow reversal far downstream in the vicinity over Suction Side
of the trailing edge is visible.

C. Aerodynamic Convergence of Geometrical Shape Functions


Airfoil geometry is sensitive to surface perturbations. To quantify geometrical variations against aerodynamic
performance, a c p convergence analysis was performed based on the computational model described above. Hence,
the c p profile of the approximated section is evaluated against the performance of the benchmark profile, to
quantitatively equate the effect of geometrical variations between target and approximated section, on the
aerodynamic performance.
The initial NACA 0015, benchmark and approximated profiles with the lowest fitness through n = 20, for each
shape function are presented in Fig. 19-21. Convergence between benchmark and approximated sections for each
shape function is evident across the three airfoils. Of the five shape functions examined, the Hicks-Henne provided
the lowest cost over the 3 target airfoils investigated. The key sources of errors are isolated within the trailing edge
region where sharp and rapid geometrical changes occur. By isolating the analysis to the trailing edge portion, the
Bernstein Polynomials provided the lowest fitness over the last five chord length intervals leading up to the trailing
edge. Hence, it is effective in converging to regions with sharp wedge angles.
The geometrical convergence is evaluated against the c p distribution (Fig. 22-24), based on airfoils with the
lowest geometrical variations. The NASA LRN(1)-1007 c p evolution (Fig. 22) indicates performance differences
between final and benchmark shape. Differences exist at the trailing and leading edge region where flow recovery
magnitude is underestimated. In general, unsteadiness of c p at the suction side indicates surface perturbations that
have failed to converge to the benchmark profile. The low geometrical variation fitness value (Fig. 19), does not
adequately provide an aerodynamically converged shape. The airfoil was tested at a Reynolds number that exceeds
the intended operational limit of 40,000 – 150,00027. The aerodynamic analysis of the final and benchmark sections
was tested under identical flow conditions, thus similar performance metric between the two plots was expected.
Since the final shape failed to converge, this indicates that low Reynolds number airfoils are extremely sensitive to
surface perturbations and the Hicks-Henne modulation, with a low geometrical cost, has failed to provide suitable
aerodynamic convergence. Hence, based on the test methodology, the Hicks-Henne is considered unsuitable for the
design of low Reynolds airfoils.
The aerodynamic evolution of the NASA LS(1)-0417 Mod planform indicates acceptable aerodynamic
performance in comparison to the benchmark profile (Fig. 23). The cost magnitude in this case, is greater in
comparison to the converged state of the NASA LRN(1)-1007 profile. Despite the larger fitness, satisfactory
aerodynamic matching between target and approximated sections is observed (Fig. 23). The NASA LS(1)-0417 Mod
planform was designed specially for chord Reynolds number range of 2-4 million1. A geometric convergence fitness
of 0.0218 equates to conclusive aerodynamic agreement between target and approximated planforms. Hence, this
test has shown that Hicks-Henne functions, can implement airfoils at Reynolds number in excess of 4 million.
Geometrical cost is also not overly sensitive towards aerodynamic convergence as was the case with low Reynolds
number scenarios.
Aerodynamic convergence of the NASA NLF(1)-1015 airfoil which was primarily designed for flight Reynolds
number of 4.0 million,28 although has been used in the 0.7 – 2.0 million29 range, is provided in Fig. 24. At first

15
American Institute of Aeronautics and Astronautics
c p , fluctuations are evident that indicate the geometrical function has not converged. The performance mismatch for
the NASA LRN section (Fig. 22), was attributed to the inadequacy of the Hicks-Henne in providing sufficient
geometrical resolution. The performance fluctuations in this case (Fig. 24) is attributed to an issue within the flow
solver as both benchmark and final sections undergo a similar oscillation phase. A fitness magnitude of 0.0175 is
almost identical to 0.0179 for the low Reynolds number airfoil (Fig. 19), though the aerodynamic convergence
between the two differs considerably in Fig. 22 and Fig. 24 respectively. This further validates that the cost
magnitude is sensitive in the design of low Reynolds number airfoils.
NASA LRN(1)-1007 Shape Convergence NASA LRN(1)-1007 Coefficient of Pressure Convergence
0.1
-1 Initial
Benchmark
0.08 Final (Hybrid: Hicks-Henne)
-0.8

0.06 -0.6

0.04 -0.4

-0.2
0.02
y/c

CP
Initial 0
0 Benchmark
BP(∆f)=0.0527 0.2
-0.02 Hicks-Henne(∆f)=0.0179
Legendre(∆f)=0.1123 0.4

-0.04 NACA(∆f)=0.0274
0.6
Wagner(∆f)=0.0197
-0.06
0.8

-0.08 1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6
x/c x/c

Figure 19. NASA LRN(1)-1007 Airfoil Shape Figure 22. NASA LRN(1)-1007 Cp Convergence
Function Convergence
NASA LS(1)-0417 Mod Shape Convergence NASA LS(1)0417 Mod Coefficient of Pressure Convergence
0.15
-1.5 Initial
Benchmark
Final (Hybrid: Hicks-Henne)
0.1
-1

0.05 Initial
-0.5
Benchmark
y/c

BP(∆f)=0.0246
C

Hicks-Henne(∆f)=0.0238
0 Legendre(∆f)=0.025 0
NACA(∆f)=0.0412
Wagner(∆f)=0.032

-0.05 0.5

-0.1 1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6
x/c x/c
Figure 20. NASA LS(1)-0417 Mod Airfoil Shape Figure 23. NASA LS(1)-0417 Mod Cp
Function Convergence Convergence
NASA NLF(1)-1015 Shape Convergence NASA NLF(1)-1015 Coefficient of Pressure Convergence

0.12 Initial
-1 Benchmark
Final (Hybrid: Hicks-Henne)
0.1
-0.8
0.08
-0.6

0.06
Initial -0.4

0.04 Benchmark
BP(∆f)=0.0297 -0.2
y/c

CP

0.02 Hicks-Henne(∆f)=0.0175
Legendre(∆f)=0.0558 0

0 NACA(∆f)=0.0271
0.2
Wagner(∆f)=0.0197
-0.02
0.4

-0.04
0.6

-0.06 0.8

-0.08 1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6
x/c x/c

Figure 21. NASA NLF(1)-1015 Airfoil Shape Figure 24. NASA NLF(1)-1015 Cp Convergence
Function Convergence
16
American Institute of Aeronautics and Astronautics
VI. Concluding Remarks
The DNO test methodology developed a validated architecture for the design of RC-MM-UAV airfoils. A shape
parameterization study examined the analytical approach by measuring the accuracy and flexibility of five geometric
shape functions in converging to a set of benchmark airfoils. The line search and PSO optimizers were
independently simulated over the three target airfoils, with the exception of Hicks-Henne that required integration of
the two optimizers; as multiple variables were to be resolved. It was concluded, that a multi-variant Hicks-Henne
method was superior over single function based models due to the lower geometrical cost and quicker solution
convergence. A hybrid optimization model was needed to address multi-variant functions. The PSO was used to
reduce the design space from multi to single based function and the line search used this initial estimate to guide the
search towards a minimum state. Aerodynamic convergence study indicated that low Reynolds number airfoils,
primarily used in the design of micro UAVs are highly sensitive to surface perturbations and the adopted test
methodology was not adequate in achieving sufficient convergence. Acceptable geometric and aerodynamic
convergence was established for airfoils operating at Reynolds number in excess of 2.0 million. This test case
presented its suitability for application in the design of RC-MM-UAV project.
Further work is needed to develop a relationship between design variable population size and the aerodynamic
coefficients through a statistical analysis to quantitatively present the relationship between the two terms. A data-
mining study through self-organizing maps similar to the PARSEC representation,30 will further illustrate the
relationship between each design variable and the overall aerodynamics. This will provide the avenue to eliminate
the variables that have marginal or no effect on the aerodynamics.
Within the overall DNO process, a flow solver validation study provided adequate lift and drag convergence in
comparison to experimental data. The study was limited since the location of the tripping point was manually
specified within the solver. A RANS solution through a two-equation turbulence model was seen to provide
acceptable agreement for a linear angle-of-attack range. Further studies will require establishing a finer grid with
y + ≈ 1 to enhance the accuracy of the computation process. The issue of boundary layer transition and modeling of
high angles-of-attack aerodynamics, is critical for the overall DNO process. The fully turbulent flow simulation
provided results that were not comparable to experimental solutions, thus illustrating the importance of
incorporating a flow transition model. Present research focuses on testing the effectiveness of a Detached Eddy
Simulation to predict high-lift flows. The boundary layer transition calculations based on the e n method are also an
area of active investigation for eventual implementation within the flow solver.

Acknowledgment
Research candidate Omar Ilaya – The Sir Lawrence Wackett Aerospace Centre, for his technical guidance with
the Particle Swarm Optimizer.

17
American Institute of Aeronautics and Astronautics
References
1
McGhee, R. J. and Beasley, W. D., "Wind-Tunnel Results for a Modified 17-Percent-Thick Low-Speed Airfoil Section,"
NASA, 1981.
2
Wong, D. K. C., "Aerospace Industry Opportunities in Australia - Unmanned Aerial Vehicles (UAVs) - Are They Ready
This Time? Are We?," 1997, pp. 13.
3
Khurana, M., Sinha, A. and Winarto, H., "Multi-Mission Re-Configurable UAV - Airfoil Analysis through Shape
Transformation and Computational Fluid Dynamics," Twelfth Australian International Aerospace Congress - Second
Australasian Unmanned Air Vehicles Conference, 2007, pp. 1-22.
4
Samareh, J. A., "A Survey of Shape Parameterization Techniques," CEAS/AIAA/ICASE/NASA Langley International Forum
on Aeroelasticity and Structural Dynamics, 1999, pp. 333-343.
5
Keane, A. J. and Nair, P. B., Computational Approaches for Aerospace Design: The Pursuit of Excellence, Wiley, John &
Sons, Incorporated, 2005.
6
Namgoog, H., "Airfoil Optimization of Morphing Aircraft," PhD, Aerospace Engineering, Purdue, Indiana, 2005.
7
Song, W. and Keane, A. J., "A Study of Shape Paramaterisation Methods for Airfoil Optimisation," AIAA/ISSMO
Multidisciplinary Analysis and Optimization Conference, 2004.
8
Khurana, M., Sinha, A. and Winarto, H., "Multi-Mission Re-Configurable UAV - Airfoil Shape Parameterisation Study,"
22nd International Unmanned Air Vehicle Systems Conference, 2007, pp. 1-15.
9
Henne, P. A. and Gregg, I., Robert D, McDonnell Douglas Corporation (Long Beach, CA), USA, Patent Application for a
"Divergent Trailing-Edge Airfoil," Docket No. 056250, filed 22 August 1989
10
Wu, H.-Y., Yang, S. and Liu, F., "Comparison of Three Geometric Representations of Airfoils for Aerodynamic
Optimisation," 16th AIAA Computational Fluid Dynamics Conference, 2003.
11
Hicks, R. M. and Henne, P. A., "Wing Design By Numerical Optimisation," AIAA Aircraft Systems & Technology Meeting,
1977, pp. pp 8.
12
Ramamoorthy, P., Dwarakanath, G. S. and Narayana, C. L., "Wagner Functions," National Aeronautical Laboratory, 1969.
13
Chang, I.-C., Torres, F. J. and Tung, C., "Geometric Analysis of Wing Sections," NASA Technical Memorandum, 1995.
14
Mathworks, "Optimization Toolbox User's Guide," 2005.
15
Eberhart, R. and Kennedy, J., "A New Optimizer Using Particle Swarm Theory," International Symposium on Micro
Machine and Human Science, 1995.
16
Eberhart, R. C. and Shi, Y., "Comparing Inertia Weights and Constriction Factors in Particle Swarm Optimization,"
Congress on Evolutionary Computation, Vol. 1, 2000, pp. 84-88.
17
Shi, Y. and Eberhart, R., "A Modified Particle Swarm Optimizer," IEEE World Congress on Computation Intelligence,
1998, pp. 69-73.
18
Yong-ling, Ma, L.-h., Zhang, L.-y. and Qian, J.-x., "Empirical Study of Particle Swarm Optimizer with an Increasing
Inertia Weight," 2003.
19
Robinson, J. and Rahmat-Samii, Y., "Particle Swarm Optimization in Electromagnetics," IEEE Transactions on Antennas
and Propagation, Vol. 52, No. 2, 2004, pp. 397-407.
20
FLUENT, "Introductory Fluent Notes," FLUENT, 2005.
21
Spalart, P. R. and Allmaras, S. R., "A one-equation turbulence model for aerodynamic flows," 30th AIAA Aerospace
Sciences Meeting & Exhibit, 1992,
22
Menter, F. R., "Performance of Popular Turbulence Models for Attached and Separated Adverse Pressure Gradient Flows,"
AIAA Journal, Vol. 30, No. 8, 1992, pp. 2066-2072.
23
Menter, F. R., "Two Equations Eddy-Viscosity Turbulence Models for Engineering Applications," AIAA Journal, Vol. 32,
No. 8, 1994, pp. 1598-1605.
24
Kern, S., "Evaluation of Turbulence Models for High-Lift Military Airfoil Flowfields," 34th Aerospace Sciences Meeting &
Exhibit, 1996.
25
Fuhrmann, H., "Design Optimisation of a Class of Low Reynolds, High Mach Number Airfoils For Use in the Martian
Atmosphere," 23rd AIAA Applied Aerodynamics Conference, 2005.
26
Kotapati-Apparao, R. B., Squires, K. D. and Forsythe, J. R., "Prediction of the Flow over an Airfoil at Maximum Lift,"
42nd AIAA Aerospace Sciences Meeting and Exhibit, 2004, pp. 1-13.
27
Evangelista, R., Pfenninger, W., Mangalam, S. M. and Bar-Sever, A., "Design and Wind Tunnel Test of a High
Performance Low Reynolds Number Airfoil," AIAA Applied Aerodynamics Conference, Vol. 5th, 1987, pp. 175-185.
28
Somers, D. S., "Design and Experimental Results for a Natural-Laminar-Flow Airfoil for General Aviation Applications,"
Langley Research Center, 1981.
29
Maughmer, M. and Somers, D. M., "An Airfoil Designed for a High-Altitude, Long Endurance Remotely Piloted Vehicle,"
AIAA Applied Aerodynamics Conference, Vol. 5th, 1987, pp. 539-545.
30
Jeong, S., Chiba, K. and Obayashi, S., "Data Mining for Aerodynamic Design Space," Journal of Aerospace Computing,
Information and Communication, Vol. 2, 2005, pp. 452-469.

18
American Institute of Aeronautics and Astronautics

View publication stats

You might also like