You are on page 1of 11

10.1002/9780470686652.eae500, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/9780470686652.eae500 by Readcube (Labtiva Inc.), Wiley Online Library on [25/02/2023].

See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Airfoil/Wing Optimization
Thomas A. Zang
Systems Analysis and Concepts Directorate, NASA Langley Research Center, Hampton, VA, USA

time-dependent flows, most recently. Although the discus-


1 Introduction 1 sion in this chapter is confined to external flows, most of
2 Optimization Formulation 2 the discussion is equally applicable to internal flows, such
as those through aircraft engines. Readers interested in sum-
3 Shape Definition 4
maries of the important developments (and contributors) in
4 Mesh Generation 7 this field should consult review articles such as those by
5 Gradient Computation 7 Labrujère and Slooff (1993), Newman et al. (1999), Reuther
6 Optimization Using CFD 8 et al. (1999), and Mohammadi and Pironneau (2004), as well
7 Disclaimer 10 as the hundreds of references therein.
Acknowledgments 10 Figure 1 illustrates the key processes and variables for
aerodynamic shape optimization. (The gradient variables are
Notes 10
shown in parentheses, as they are only relevant for gradient-
References 10 based optimization methods.) The Optimization process
feeds a set of design variables x (a vector of length d) to
the Shape Definition component, which provides a mathe-
matical description s of the surface. The surface description
1 INTRODUCTION is used by the Mesh Generation process, which produces the
computational mesh y used by the Aerodynamic Analysis
This chapter focuses on techniques for improving the aerody- component. The analysis output is the state variable q (and
namic characteristics of an aerospace vehicle through refine- its gradient ∇q). The Objective and Constraint Evaluation
ment of the vehicle’s shape. Methods for aerodynamic shape process supplies the objective function f , the inequality and
optimization have progressed through increasingly complex
aerodynamic analysis tools – roughly speaking, for linear
aerodynamics (e.g., panel methods), transonic small distur- Optimization
x Shape s Mesh
bance equations, and full potential equations in the 1970s; for definition generation
linear aerodynamics with boundary-layer corrections in the
1980s; for Euler equations from the 1980s through the mid-
1990s; for Navier–Stokes equations on structured meshes f, g j, hk y
in the 1990s; for Euler and Navier–Stokes equations on (∇f, ∇g j ,∇hk )

unstructured meshes from the late 1990s onward; and for


Objective Aerodynamic
and constraint
q analysis
evaluation
(∇q)

Figure 1. Key processes and variables for aerodynamic shape


optimization.

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 US Government in the US and © 2010 John Wiley & Sons, Ltd in the rest of the world.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae500
10.1002/9780470686652.eae500, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/9780470686652.eae500 by Readcube (Labtiva Inc.), Wiley Online Library on [25/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
2 Aerospace System Optimization

equality constraints g and h (perhaps along with their gradi- bounds on the i-th design variable. The design variables x
ents) back to the Optimization process. have been added to the argument list of the state variable to
Those components of the optimization process that are of emphasize its parametric dependence upon them. Likewise,
special importance to aerodynamic shape optimization are the we now write the state equation (1) as
optimization formulation (objectives and constraints), shape
definition, mesh generation, and gradient computation (for r(q(x̂; x), x̂; x) = 0 (6)
those methods that utilize gradient information). These top-
ics are covered in that order. The initial expository material We necessarily have hk = rk for k = 1, . . . , dimension of r,
on the optimization formulation is given in the context of as each component of r yields a constraint. Because the state
linear aerodynamics. Thereafter, the emphasis is on the im- vector is defined implicitly, as in equation (6), the objective
portant considerations for aerodynamic shape optimization and constraint functions in equations (2)–(3) have a depen-
using the Euler or Navier–Stokes equations, that is, what are dence upon q and x̂. For aerodynamic shape optimization, the
commonly called computational fluid dynamics (CFD) tools. design variables characterize the shape of the airfoil, wing or
The chapter concludes with some considerations for the use aircraft.
of CFD in aerodynamic shape optimization. The discussion To illustrate the basic principles, we consider optimization
is restricted to use of optimization for improving an existing of airfoils in the four-digit NACA series (see, e.g., Abbott and
design, that is, optimization that starts from an existing base- von Doenhoff, 1959, Chapter 6). These airfoils are described
line design and is constrained to maintain the same topology by the maximum camber (m), the distance of the location
for the shape. of the maximum camber (p) from the leading edge, and the
maximum thickness (t); all of these are given as fractions of
the chord c. These parameters are illustrated in Figure 2a.
2 OPTIMIZATION FORMULATION For a four-digit NACA airfoil, say, a NACA d1 d2 d3 d4 airfoil,
the first digit d1 is m (in hundredths), the second digit is p
The state variable q is a function of the spatial coordinate, (in tenths), and the last two digits are t (in hundredths). Thus,
which is denoted by the vector x̂. (The spatial variable has a NACA 2416 airfoil has a maximum camber of 0.20c located
the unusual hat because in this volume the symbol x – with- at x = 0.10c, and a maximum thickness of 0.16c, where c is
out the hat – is reserved for the vector of design variables.) the airfoil chord. Figure 2b is indicative of internal structures
The state variable is a solution of the governing equations for that constrain the geometry; such important constraints are
an appropriate conceptual model for the flow (compressible ignored in the present illustration.
or incompressible, viscous or inviscid, turbulent or laminar, The conceptual model used for the airfoil analysis in this
nonlinear or linear). We write the governing equations gener- example is inviscid, irrotational, incompressible, constant-
ically as density flow. The velocity potential φ can be used as the
(scalar) state variable. The mathematical model is given by
r(q(x̂), x̂) = 0 (1) the Laplace equation

The optimization problems that are considered in this r = ∇ˆ · ∇φ


ˆ (7)
chapter have the form
where ∇ˆ denotes the gradient operator (with respect to x̂).
min f (x; q(x̂; x), x̂) (2) The Laplace equation is subject to a homogeneous Neumann
x boundary condition (∇φˆ · n = 0, where n is unit vector nor-
mal to the surface) on the airfoil surface, an homogeneous
subject to Dirichlet condition (φ = 0) at infinity, and a Kutta condition
(upper and lower surface pressures match) at the trailing edge
gj (x; q(x̂; x), x̂) ≤ 0 , j = 1, . . . , m (3) The computational model is based upon a panel (boundary-
element) method1 . This very simple model ignores many im-
hk (x; q(x̂; x), x̂) = 0 , k = 1, . . . , p (4) portant effects such as viscosity, compressibility, vorticity,
and nonlinearity but provides a useful illustration of some
xiL ≤ xi ≤ xiU , i = 1, . . . , n (5) basic concepts of aerodynamic shape optimization.
For this example, the design variables x = (m, p, t), with
where x is the vector of design variables, f is the objec- these three NACA airfoil parameters treated as continuous
tive function, gj and hk are inequality and equality constraint variables; their upper and lower bounds are taken to be
functions, respectively, and xiL and xiU are upper and lower xL = (0.0, 0.1, 0.0) and xU = (0.1, 0.5, 0.2), respectively. In
Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 US Government in the US and © 2010 John Wiley & Sons, Ltd in the rest of the world.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae500
10.1002/9780470686652.eae500, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/9780470686652.eae500 by Readcube (Labtiva Inc.), Wiley Online Library on [25/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Airfoil/Wing Optimization 3

Fuel tank Stiffeners and ribs



Mean camber
line Chord Trailing
x̂ line
Leading edge edge
tc
mc

pc
(a) Chord c (b) Spars

Figure 2. (a) National Advisory Committee for Aeronautics (NACA) four-digit airfoil parameters; (b) representative internal structures
(courtesy of J. A. Samareh).

all cases, the angle of attack is chosen to be α = 1◦ . The key


performance outputs are the (section) lift, drag, and pitching 0.2
moment coefficients, cl , cd and cm , respectively. (The mo-
ment reference center is the quarter-chord point.) Four rep- 0.1
resentative cases are considered: (i) maximize lift (f = −cl )
with no constraints, (ii) maximize lift with inequality con- ẑ 0
straints on the pitching moment (g1 = −cm − 0.04) and drag
(g2 = cd − 0.02), (iii) minimize drag (f = cd ) with inequal- max c l
−0.1
ity constraints on the pitching moment (g1 = −cm − 0.04) max c l [gi : cm and cd ]
and lift (g2 = cm − 0.30), and (iv) minimize lift/drag (f = min cd [gi : cm and cl ]
−0.2
−cl /cd ) with inequality constraints on the pitching moment min cl / cd [gi : cm and c l ]

(g1 = −cm − 0.04) and lift (g2 = cl − 0.30. (There are no 0 0.2 0.4 0.6 0.8 1
equality constraints in these examples.) The gradient-based x̂
optimizations employ Sequential Quadratic Programming,
using finite differences to compute the gradients and the Figure 3. Optimal shapes for the NACA four-digit airfoil example.
Broyden–Fletcher–Goldfarb–Shanno (BFGS) method to ap-
proximate the Hessian2 . The starting point for the optimiza- Another widely used objective function is the difference
tions is x0 = (xL + xU )/2. (in the least squares sense) between the surface pressure
The optimal solutions to these four problems are given in distribution and a target pressure distribution:
Table 1. The pitching moment constraint is active in cases (ii)

and (iv). The drag constraint is also active in case (ii). The lift
constraint is active in case (iii). The corresponding airfoils are f (x) = || p(x̂; x) − ptarget (x̂) ||2 (8)
S
shown in Figure 3. The key performance parameters are also
included in the table. The impact of constraints is especially where the integral is taken over the airfoil or wing surface
dramatic in the contrast between the results for the first two S of interest, p is the pressure in the flow produced by an
cases (maximization of lift). The conceptual model for this airfoil described by the design variables x, and ptarget is the
simple example has neglected several important physical ef- target pressure. The basic concept is illustrated in Figure 4.
fects, most noticeably those of viscosity and compressibility. The solid line is the pressure coefficient (Cp ) for the present
The resulting “optimal” designs are by no means representa- (baseline) airfoil. The dotted line is the desired (target) pres-
tive of airfoils that would be useful in practice. sure distribution. Many rules have been built up over the

Table 1. Design variables and optimal solutions for the airfoil examples.

Case f gi m p t min f cl cd cm

1 −cl — 0.10 0.50 0.20 −1.53 1.529 0.0030 −0.321


2 −cl cm and cd 0.029 0.10 0.15 −0.44 0.436 0.0020 −0.040
3 cd cm and cl 0.016 0.30 0.10 0.0010 0.300 0.0019 −0.036
4 −cl /cd cm and cl 0.028 0.11 0.10 −369 0.403 0.0011 −0.040

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 US Government in the US and © 2010 John Wiley & Sons, Ltd in the rest of the world.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae500
10.1002/9780470686652.eae500, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/9780470686652.eae500 by Readcube (Labtiva Inc.), Wiley Online Library on [25/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
4 Aerospace System Optimization

3 SHAPE DEFINITION

Aerodynamic shape optimization methods using nonlinear


analysis tools are faced with significant challenges in shape
parameterization, volume mesh generation, and sensitivity
analysis. Although these components are routine for many
optimization problems in other disciplines, they are nontriv-
ial for aerodynamic shape optimization methods that yield
realistic shapes robustly and efficiently. The surface of an
aircraft, a wing, or even an airfoil is an infinite-dimensional
object, but it must be parameterized as a finite-dimensional
object. (The parameters of the shape correspond to the design
variables for aerodynamic shape optimization.) The shape
Figure 4. Baseline and target pressure distributions (courtesy of
R.L. Campbell).
parameterization should compactly cover the design space
of interest while presenting no undue difficulties. In partic-
ular, parameterization-induced waviness and discontinuities
years for choosing target pressure distributions to achieve in (the slope and curvature of) the surface are undesirable.
desired performance measures. Methods for matching partic- These features are undesirable because they present manu-
ular flow-field characteristics rather than directly minimizing facturing and CFD analysis difficulties (since flow fields are
a global objective function such as drag are referred to as quite sensitive to such discontinuities). Sometimes, optimiz-
inverse design methods; the more conventional alternative is ing the shape of the entire surface is desired, but other times
called a direct optimization method. One could apply formal optimization is only applied to a portion of the surface. The
optimization to the inverse problem with the objective func- alternatives summarized below apply in both cases.
tion given by equation (8). However, there is a wide variety The usual custom for aerodynamic shape parameterization
of inverse design methods that “minimize” the objective in is to represent the surface as a baseline surface plus a pertur-
equation (8) much faster than direct optimization methods. bation expressed as an expansion in terms of basis functions.
An example of one such method is provided in Section 6 (see Consider the case of airfoils. The surface S is one dimen-
Labrujère and Slooff, 1993 for descriptions of many others). sional, and the basis functions bk (ξ) depend upon a single
In practical applications, numerous geometric constraints coordinate ξ. The expansion is given by
are needed for obtaining an acceptable result, for example,
requiring the external shape to accommodate internal struc- 
d

tures (see Figure 2), minimal radius of curvature of the lead- s(ξ; x) = xk bk (ξ) (9)
ing edge, and minimal angle of the trailing edge. Furthermore, k=1

aerodynamic shapes are required to operate over a range of


with the coefficients xk in the expansion serving as the de-
conditions, especially in Mach number. This is addressed
sign variables. The perturbation s(ξ; x) may be applied to
with a multi-objective optimization formulation (commonly
whatever functions are used for the surface definition, for
referred to as multipoint optimization in the aerodynamic
example, upper or lower airfoil surface, mean camber line,
optimization community) (see Drela, 1998 for an extensive
airfoil thickness. A variety of basis functions have been used,
discussion of constraints and objectives for airfoil optimiza-
some global and some local. Since the leading and trailing
tion).
edges of airfoils are fixed in most design problems, having
The NACA airfoil family used for the example above was
basis functions that vanish at both end points is desirable.
convenient for this demonstration because the parameteriza-
The sine functions are one option for a complete set of
tion guaranteed a simple airfoil shape. Furthermore, the de-
global basis functions. Drela (1998) recommends using them
sign variables corresponded directly with physical quantities
in the form
for which experienced aerodynamicists have considerable in-
tuition. However, the optimization problems of interest for 1
many decades now require consideration of a much broader bk (ξ) = sin(πkξ) (10)
k
range of shapes than those covered by the NACA airfoil fam-
ilies. In the following section, we cover the fundamentals of They form an orthogonal set for ξ ∈ [0, 1]. The sine basis
parameterizations that can represent a more general range of functions are illustrated in Figure 5a. Suitable combinations
shapes. of orthogonal polynomials can also be chosen to ensure that

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 US Government in the US and © 2010 John Wiley & Sons, Ltd in the rest of the world.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae500
10.1002/9780470686652.eae500, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/9780470686652.eae500 by Readcube (Labtiva Inc.), Wiley Online Library on [25/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Airfoil/Wing Optimization 5

1 1

0.5 0.5

bk bk
0 0

k =1 k=2
−0.5 k =2 −0.5 k=3
k =3 k=4
k =4 k=5
−1 −1
0 0.2 0.4 0.6 0.8 1 −1 −0.5 0 0.5 1
(a) x̂ (b) x̂

Figure 5. Examples of global basis functions: (a) sine; (b) Legendre.

the basis functions vanish at the end points. For example, one (Hicks and Henne, 1978) functions, which are defined by
can use
b(ξ) = {sin[πξ log(1/2)/ log(t1 ) ]}t2 (12)
1
bk (ξ) = √ (Lk−1 (ξ) − Lk+1 (ξ)) for k ≥ 1 (11)
2 (k + 1) with the domain now again normalized to ξ ∈ [0, 1]; t1 con-
trols the location of the peak and t2 its width. Figure 6a
for ξ ∈ [−1, 1], where Lk (ξ) is the Legendre polynomial of illustrates some of these functions. Although these functions
degree k. The set of basis functions given by equation (11) is lack the completeness property possessed by the sine func-
nearly orthogonal – the inner product of bk and bl vanishes tions and the orthogonal polynomials, they have proven very
except for l = k − 2, k, k + 2. The Legendre basis functions useful in the hands of skilled designers.
are illustrated in Figure 5b. These form a complete set of A more conventional choice of local basis functions are
basis functions. They are advantageous for approximating splines, which are piecewise polynomials. Figure 6b illus-
functions with a high degree of smoothness (more than, say, trates the case of cubic B-splines on uniformly distributed
four continuous derivatives), because they exhibit much faster knots (denoted by the dots on the x̂-axis); the curves are la-
convergence than the sine functions in such cases (see Canuto beled by the x̂ coordinate of the center of the spline. Unlike
et al., 2006, Chapter 2). the global basis functions shown in Figure 5, the underly-
A noteworthy set of local basis functions specifically ing B-spline basis functions effect only local changes in the
chosen for airfoil parameterization are the Hicks–Henne shape. On the other hand, they have only a finite number

1 1
c =0.3
c =0.5
0.75 0.75 c =0.7

bk bk
0.5 0.5

t 1 = 0.25, t 2 = 0.5
0.25 0.25
t 1 = 0.50, t 2 = 1.0

t 1 = 0.75, t 2 = 1.5

0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
(a) x̂ (b) x̂

Figure 6. Examples of local basis functions: (a) Hicks–Henne; (b) B-spline.

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 US Government in the US and © 2010 John Wiley & Sons, Ltd in the rest of the world.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae500
10.1002/9780470686652.eae500, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/9780470686652.eae500 by Readcube (Labtiva Inc.), Wiley Online Library on [25/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
6 Aerospace System Optimization

0.2 0.2
Baseline Baseline
Camber perturbation Thickness perturbation

0.1 0.1

ẑ ẑ
0 0

−0.1 −0.1

−0.2 −0.2
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
(a) x̂ (b) x̂

Figure 7. NURBS-based perturbations for (a) airfoil camber; (b) thickness.

of continuous derivatives, namely, p − 1 continuous deriva- systems. Figure 7 illustrates airfoil deformations based on
tives for splines of order p. Hence, if p > 2, then the shape NURBS expansions of the airfoil camber and thickness.
perturbation, along with its first and second derivatives, is Using the locations of the surface mesh points as design
continuous. This includes the cubic B-splines (p = 3) shown variables is yet another approach. The main challenge here
in Figure 6b. A generalization of B-splines, called nonuni- is picking appropriate geometric constraints and/or filtering
form rational B-splines (NURBS), has become a fairly widely procedures to ensure a smooth surface. See Li and Krist
used parameterization, particularly for complex shapes, such (2005) for one among many approaches to surface smooth-
as full aircraft configurations. NURBS represent functions ing.
(in this case surfaces) as a rational function, with both the For the parameterization of wings, the same approaches
numerator and denominator consisting of B-spline expan- apply, but, of course, there are now additional types of design
sions. Piegl and Tiller (1996) provide a comprehensive de- variables. Some representative wing parameters with a direct
scription of NURBS. See Samareh (2001) for a summary of aerodynamic interpretation are illustrated in Figure 8. The
their mathematical description and a short survey of vari- symbols indicate locations where the parameters are defined.
ous uses in aerodynamic shape optimization. An important The root chord, tip chord, semi-span, and leading edge
advantage of a NURBS representation for the shape is that sweep angle are planform variables; these are usually fixed
they are compatible with most computer-aided design (CAD) early in the design process. The airfoil section parameters


Leading edge
sweep angle

Camber and thickness
Root chord Twist and shear

Tip chord

Semi-span

Figure 8. Typical design variables for wings.

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 US Government in the US and © 2010 John Wiley & Sons, Ltd in the rest of the world.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae500
10.1002/9780470686652.eae500, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/9780470686652.eae500 by Readcube (Labtiva Inc.), Wiley Online Library on [25/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Airfoil/Wing Optimization 7

(camber and thickness) as well as the wing twist and shear CFD volume mesh for a wing–fuselage configuration. The
are often the focus of wing shape design. The twist angle at coordinates of these volume mesh-points are denoted by y in
a given airfoil section is the difference between the airfoil Figure 1.
section incident angle at the root and the incident angle of For aerodynamic shape optimization, changes to the sur-
that airfoil section. Similarly, the shear (dihedral) is the dif- face of the airfoil resulting from design variable changes
ference between the airfoil leading edge ẑ coordinate for the require adjustments to the volume mesh as well, and these
root and the ẑ coordinate for the particular airfoil section. adjustments need to occur automatically. Nowadays, com-
putational meshes suitable for even viscous CFD analysis
can be generated automatically from the surface shape defi-
4 MESH GENERATION nition for airfoils and wings, but this state has proven to be
very elusive for complex aircraft configurations subjected to
Unlike panel methods, for which a surface mesh is sufficient, Navier–Stokes analyses – some type of user intervention is
CFD methods require a volume mesh for numerical solution typically needed if the mesh generation must be performed
of the state equation. For compressible flow, the state variable ab initio. Thompson, Soni and Weatherill (1998) provide an
q is given by extensive description of methods for CFD mesh generation.
Moreover, as discussed in the next section, analytically based
q = (ρ, ρu, ρE)T (13) gradients are highly desirable, and these are not generally
available from ab initio mesh generation packages. Hence,
where ρ is the density, u the velocity, and the (specific) total many aerodynamic shape optimization processes use special
energy E = e + 21 u · u , where e denotes the (specific) inter- processes that lend themselves to analytically based gradi-
nal energy. For steady, inviscid flow described by the Euler ents for obtaining the volume mesh as a perturbation upon
equations, we have the mesh associated with the baseline shape. Samareh (2001)
contains an overview of the various approaches to volume
r(q(x̂), x̂) = ∇ˆ · F (14) mesh generation.

where the flux F = (ρu, ρuuT , ρuE + pu)T . The Navier–


Stokes equations have the same form as equation (14) but
with additional terms in the flux function. Figure 9 illustrates 5 GRADIENT COMPUTATION
the surface and symmetry plane portions of an unstructured
For gradient-based optimization methods, the derivatives of
the objective and constraints with respect to the design vari-
ables (terms in parentheses in Figure 1) are needed. For non-
linear CFD methods, computation of the gradients3 using
finite differences has several disadvantages: (i) computation
of each gradient requires another full CFD solution since
the equations are nonlinear; (ii) extensive trial and error is
necessary to choose the appropriate step size for each de-
sign variable; and (iii) for some difficult CFD problems, the
analysis code may be simply unable to converge to the level
needed for accurate finite-difference gradients. Computation
of these derivatives through quasi-analytical means is pre-
ferred. (The adjective quasi-analytical is used because the
equations for the gradients are derived analytically but solved
numerically.) In order to distinguish the state variable and the
state equations, which are functions, from their discrete rep-
resentations, which are vectors, we use the symbols q̃ and
r̃, respectively, for the latter; similarly ỹ is the vector of the
mesh-point coordinates. For a problem with N mesh-points,
Figure 9. Surface mesh for wing optimization in the presence of the length of q̃ and r̃ is 5N since q has five components at
the fuselage. Reproduced with permission from Nielsen and Park each point, and the length of ỹ is 3N. (Boundary conditions
(2006) c AIAA. may slightly alter these lengths.)

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 US Government in the US and © 2010 John Wiley & Sons, Ltd in the rest of the world.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae500
10.1002/9780470686652.eae500, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/9780470686652.eae500 by Readcube (Labtiva Inc.), Wiley Online Library on [25/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
8 Aerospace System Optimization

Recall that a change in a design variable is propagated equations are derived from the algebraic equations resulting
through the shape definition and mesh generation processes from the discretization of the PDEs.) See studies by Newman
(Figure 1), with the computational mesh y depending upon et al. (1999) and Reuther et al. (1999) for extended lists of
x. In particular, the gradient of, say, the objective function references to the early experiences of various groups with
with respect to the particular design variable xj is given by both approaches. Some of the subtle issues that must be ad-
dressed for consistent aerodynamic gradients are the treat-
 5N   3N 
  ment of boundary conditions and mesh gradients (see Giles
df ∂f ∂f ∂q̃k ∂f ∂ỹl
= + + (15) et al., 2003; Nielsen and Park, 2006 for detailed recommen-
dxj ∂xj k=1
∂q̃k ∂xj l=1
∂ỹl ∂xj
dations). An approach to obtaining second-order derivatives
by combining direct and adjoint techniques is described by
Evaluation of this requires determination of the three explicit
Sherman et al. (1996).
partial derivatives of f plus the ∂q̃k /∂xj and ∂ỹl /∂xj terms.
Yet another alternative to computing the gradients is the
The first term on the right-hand side is usually straightfor-
use of complex variables, which we illustrate on a simple
ward to compute. (For the airfoil example in Section 2, all
real-valued scalar function f . In this method, the function is
these derivatives vanish, as the design variables do not appear
evaluated at the complex point x + i, where i denotes the
explicitly in the expressions for the lift, drag, and pitching
imaginary unit, for small . A Taylor expansion yields
moment.) Likewise, an analytical expression can typically
be derived straightforwardly and then evaluated numerically df
for ∂f/∂q̃k and ∂f/∂ỹl . The ∂ỹl /∂xj terms represent the influ- f (x + i) = f (x) + (i) + O(2 ) (17)
dx
ence of the design variables upon the computational mesh.
The surface definition techniques illustrated for airfoils in Thus, for small , the real part of f (x + i) is a good ap-
Figures 5–7, as well as some mesh perturbation techniques, proximation to f (x), and the imaginary part divided by  is
lend themselves to analytical evaluation of these terms. a good approximation to the gradient df/dx. Note that this
Computation of the ∂q̃k /∂xj terms is more involved. Con- approach does not involve the subtraction of nearly equal
sider the state equation given by equation (6). Its solution quantities, as occurs for gradients computed via finite dif-
q, considered as a function of the design variables x, satis- ferences. Hence, round-off errors are not a concern for the
fies dr/dx = 0. Hence, implicit differentiation of the state complex variable approach. Provided that the source code is
equation yields the following expression: available, implementation of this approach can require little
more than changing the type of variables from real to com-
5N 
    3N  
∂r̃i ∂q̃k ∂r̃i ∂r̃i ∂ỹl plex. A prototypical use of this approach for aerodynamic
=− − (16)
k=1
∂q̃k ∂xj ∂xj l=1
∂ỹl ∂xj shape optimization is given by Nielsen and Kleb (2006).

Equation (16) is linear in the desired ∂q̃k /∂xj term. The size of
the linear system is 5N × 5N. CFD computations for wings 6 OPTIMIZATION USING CFD
typically use O(106 ) mesh-points, and complex configura-
tions can easily take upwards of 108 mesh-points. Hence, the Over the past several decades, a broad assortment of direct
linear system is extremely large in three dimensions – so large optimization and inverse design methods has been applied to
that direct solution methods are impractical. An efficient it- aerodynamic shape optimization using CFD. The principal
erative solution method is surveyed in Newman et al. (1999) distinctions between the approaches are (i) use of global per-
and described in detail in Korivi et al. (1994). formance measures such as lift, drag, and pitching moment in
For most aerodynamic shape optimization problems, the the objectives and constraints versus inverse objectives such
number of design variables far exceeds the number of ob- as matching a prescribed pressure distribution, (ii) gradient-
jectives and constraints. Hence, the adjoint approach for ob- based methods versus methods using only function values,
taining the gradients is much more efficient than the direct and (iii) having the optimizer invoke the aerodynamic anal-
approach described above. Details may be found, for exam- ysis tool itself versus having it rely upon surrogate models
ple, in Newman et al. (1999) and Reuther et al. (1999). Both constructed off-line.
continuous and discrete adjoints have been used in aero- Most of the review articles cited in Section 1 concentrate
dynamics. (The phrase continuous adjoint refers to one in on direct optimization of global performance measures using
which the continuous adjoint equations are first derived from gradient-based methods that directly call CFD analysis tools.
the governing partial differential equations (PDEs) and then Several difficulties with this approach can arise. For exam-
discretized; a discrete adjoint is one in which the adjoint ple, the analysis code may produce solutions that result in

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 US Government in the US and © 2010 John Wiley & Sons, Ltd in the rest of the world.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae500
10.1002/9780470686652.eae500, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/9780470686652.eae500 by Readcube (Labtiva Inc.), Wiley Online Library on [25/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Airfoil/Wing Optimization 9

objective functions and constraints that are contaminated by


numerical noise. Such noise gives the appearance of many
local extrema and can render gradient-based methods un-
workable. Another potential difficulty is that there may be
portions of the desired design space for which the mesh gen-
eration or the analysis process simply fails. At an even more
fundamental level, the problem formulation (objective and
constraints) itself can omit important considerations with the
result that the optimization produces a design that is unreal-
istic from the perspectives of other disciplines. See the cited
review articles for extensive discussion of these issues. Figure 11. Design improvement regimes for an aircraft (courtesy
Nevertheless, many groups have used gradient-based of J. Hooker and A. Agelastos (Lockheed-Martin) and W. Milholen
optimization of global performance measures without re- and R. Campbell (NASA)).
course to surrogate models to produce impressive CFD-based
optimization results for multi-element airfoils, wings, wing–
fuselage configurations, and even more complex shapes. modifying the surface curvature in proportion to the desired
Figure 10, from Nielsen and Park (2006), shows the objec- change in surface pressure. The surface geometry is modified
tive function convergence for unconstrained maximization of at the same time that the flow solver is converging; hence, this
the lift-to-drag ratio of the vehicle illustrated in Figure 9 at method costs little more than a single CFD analysis. Details
transonic conditions using a Navier–Stokes CFD code. The can be found in the study by Campbell (1992). Figure 11,
optimum was achieved in a few dozen function evaluations. taken from joint Lockheed-Air Force-NASA work, illustrates
The surface definition for the wing utilized NURBS rep- the various regions of the aircraft to which this inverse de-
resentations of the camber, thickness, twist, and shear (see sign method was applied. (PAI refers to propulsion-airframe
Figure 8) perturbations from the baseline shape. (The fuse- integration.) Collectively, these improvements in the local
lage was taken as fixed.) Gradients were computed with a aircraft shape enabled the design to meet its performance
quasi-analytical adjoint method. goals.
One particular approach to inverse design that has The most common non-gradient-based optimization
seen considerable use in industry applications is based on methods that have been applied to aerodynamic shape op-
timization are genetic algorithms (see Holst, 2005 for an ex-
ample). These have exhibited the customary advantage of
greater likelihood of finding the global optimum, as well as
the accompanying disadvantage of taking significantly more
computational time for convergence than gradient-based
methods.
∆L / D = +32% In aerodynamics, as in other disciplines, surrogate models
are extensively used in optimization. In some cases, these are
used to deal with noisy analysis results. Linear aerodynam-
L /D ratio

ics methods are particularly prone to producing noisy results.


(Some surrogate models such as quadratic or cubic response
surfaces smooth the noise, but some other surrogate models
do not.) In other cases, surrogates are used to deal with por-
tions of the design space that cause difficulties for the analysis
code, since the surrogate model enables the optimization to
proceed even in the face of occasional analysis failures. Yet
another use is to permit efficient searches for global optima
using, say, pattern-search or genetic algorithm techniques.
0 5 10 15 20 25 30 35 For the relatively expensive Euler and Navier–Stokes mod-
Function evaluation els, use of surrogates entails the usual compromise between
Figure 10. Lift/drag convergence from a wing optimization. accuracy (relative to the CFD model) and speed. All types of
Reproduced with permission from Nielsen and Park (2006) surrogates – design of experiments, kriging, neural networks,

c AIAA. radial basis functions – have been fruitfully employed on this

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 US Government in the US and © 2010 John Wiley & Sons, Ltd in the rest of the world.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae500
10.1002/9780470686652.eae500, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/9780470686652.eae500 by Readcube (Labtiva Inc.), Wiley Online Library on [25/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
10 Aerospace System Optimization

application. These types of surrogates all suffer from the curse NOTES
of dimensionality, limiting the number of design variables in
practice to O(10). 1. The underlying computations are performed with a
Of course, one would like optimization using a surrogate TM
Matlab code of L.N. Sankar, found on-line at
to converge to the optimum corresponding to that for the http://www.ae.gatech.edu/people/lsankar/AE3903/.
underlying analysis code. Several optimization methods that
2. Performed with the MatlabTM routine fmincon.
judiciously mix calls to the surrogate with calls to the actual
code have been proven mathematically to converge to the 3. These gradients are often called sensitivity derivatives.
true optimum. See Booker et al. (1999) for discussion of one
method in a general setting, and see Alexandrov et al. (2001)
for another method with an aerodynamic shape optimization
application; in practice, the latter seems to reduce the overall REFERENCES
cost by a factor of 3–5 compared with always invoking the
CFD code. When the surrogate is just a lower fidelity model, Abbott, I.A. and von Doenhoff A.E. (1959) Theory of Wing Sections,
for example, using an Euler code instead of a Navier–Stokes Dover Publications, New York.
code or a coarse-grid solution in place of a fine-grid solution, Alexandrov, N.M., Lewis, R.M., Gumbert, C.R., Green, L.L. and
then the number of design variables is not limited as it is for Newman, P.A. (2001) Approximation and model management in
generic surrogates. aerodynamic optimization with variable-fidelity models. J. Air-
In addition to the considerations mentioned above, one craft, 38(6), 1093–1101.
needs to weigh the computational costs of the various meth- Booker, A.J., Dennism, J.E., Jr., Frank, P.D., Serafini, D.B., Torczon,
ods, at least for CFD. (Linear aerodynamics methods are V. and Trosset, M.W. (1999) A rigorous framework for optimiza-
tion of expensive functions by surrogates. Struct. Optim., 17(1),
so fast on desktop computers that CPU time is rarely an 1–13.
issue, regardless of the choice of optimization method.)
Campbell, R.L. (1992) An approach to constrained aerodynamic
Roughly speaking, in terms of the cost of a single CFD anal- design with application to airfoils. NASA-TP-3260.
ysis, an inverse design method takes O(1) analysis time, Canuto, C., Hussaini, M.Y., Quarteroni, A., and Zang, T.A. (2006)
gradient-based optimization using the adjoint formulation Spectral Methods, Fundamentals in Single Domains, Springer,
takes O(10) analysis time, and non-gradient-based methods New York.
(including surrogate ones) take O(100 − 10 000) analysis Drela, M. (1998) Pros and cons of airfoil optimization, in Frontiers
time. All these methods have their niches. Non-gradient- of Computational Fluid Dynamics (eds D.A. Caughey, and M.M.
based methods and surrogate models enable exploration of Hafez), World Scientific, Hackensack, pp. 363–381.
large regions of the design space. Gradient-based methods Giles, M.B., Duta, M.C., Müller, J.-D. and Pierce, N.A. (2003) Algo-
facilitate refinement of a local optimum, and inverse design rithm developments for discrete adjoint methods. AIAA J., 41(2),
198–205.
methods help fine-tune performance in local regions of the
Hicks, R.M. and Henne, P.A. (1978) Wing design by numerical
surface.
optimization. J. Aircraft, 15(7), 407–412.
Holst, T. (2005) Genetic algorithms applied to multi-objective
aerospace shape optimization. J. Aero. Comput. Info. Comm., 2,
7 DISCLAIMER 217–235.
Korivi, V.M., Taylor, A.C. Ill, Newman, P.A., Hou, G.J.-W. and
This chapter is declared a work of the U.S. Government and Jones, H.E. (1994) An approximate factored incremental strategy
for calculating consistent discrete CFD sensitivity derivatives. J.
is not subject to copyright protection in the United States. Comput. Phys., 113(2), 336–346.
Labrujère, Th.E. and Slooff, J.W. (1993) Computational methods
for the aerodynamic design of aircraft components. Ann. Rev.
ACKNOWLEDGMENTS Fluid Mech., 25, 183–214.
Li, W. and Krist, S. (2005) Spline-based airfoil curvature smoothing
The author gratefully acknowledges numerous discussions and its applications. J. Aircraft, 42(4), 1065–1074.
with his colleagues at the NASA Langley Research Center for Mohammadi, B. and Pironneau, O. (2004) Shape optimization in
fluid mechanics. Ann. Rev. Fluid Mech., 36, 255–279.
all their contributions to his understanding of this subject. The
numerical examples in Section 3 utilized publicly available Newman, J.C. III, Taylor, A.C. III, Barnwell, R.W., Newman,
P.A. and Hou, G.J.-W. (1999) Overview of sensitivity analysis
codes of L.N. Sankar. J.A. Samareh supplied the code used and shape optimization for complex aerodynamic configurations.
for Figure 7. AIAA J., 36(1), 87–96

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 US Government in the US and © 2010 John Wiley & Sons, Ltd in the rest of the world.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae500
10.1002/9780470686652.eae500, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/9780470686652.eae500 by Readcube (Labtiva Inc.), Wiley Online Library on [25/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Airfoil/Wing Optimization 11

Nielsen, E.J. and Kleb, W.L. (2006) Efficient construction of Samareh, J.A. (2001) Survey of shape parameterization techniques
discrete adjoint operators on unstructured grids using complex for high-fidelity multidisciplinary shape optimization. AIAA J.,
variables. AIAA J., 44(4), 827–836. 39(5), 877–884.
Nielsen, E.J. and Park, M.A. (2006) Using an adjoint approach to Sherman, L.L., Taylor, A.C. III, Green, L.L., Newman, P.A., Hou,
eliminate mesh sensitivities in computational design. AIAA J., G.J.-W. and Korivi, V.M. (1996) First- and second-order aerody-
44(5), 948–953. namic sensitivity derivatives via automatic differentiation with
Piegl, L. and Tiller, W. (1996) The NURBS Book, Springer, incremental iterative methods. J. Comput. Phys., 129(2), 307–
New York. 331.
Reuther, J.J., Jameson, A., Alonso, J.J., Rimlinger, M.J. and Thompson, J.R., Soni, B.K. and Weatherill, N.P. (eds) (1998)
Saunders, D. (1999) Constrained multipoint aerodynamic shape Handbook of Grid Generation, CRC Press, Boca Raton.
optimization using an adjoint formulation and parallel computers.
Parts 1 and 2. AIAA J., 36(1), 51–74.

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 US Government in the US and © 2010 John Wiley & Sons, Ltd in the rest of the world.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae500

You might also like