eae500

Upload: natalia-goncalves

Post on 03-Jun-2018

216 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/12/2019 eae500

    1/11

    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.

    Airfoil/Wing Optimization

    Thomas A. ZangSystems Analysis and Concepts Directorate, NASA Langley Research Center, Hampton, VA, USA

    1 Introduction 1

    2 Optimization Formulation 2

    3 Shape Definition 4

    4 Mesh Generation 7

    5 Gradient Computation 7

    6 Optimization Using CFD 8

    7 Disclaimer 10

    Acknowledgments 10

    Notes 10

    References 10

    1 INTRODUCTION

    This chapter focuses on techniques for improving the aerody-

    namic characteristics of an aerospace vehicle through refine-

    ment of the vehicles shape. Methods for aerodynamic shape

    optimization have progressed through increasingly complex

    aerodynamic analysis tools roughly speaking, for linear

    aerodynamics (e.g., panel methods), transonic small distur-

    bance equations, and full potential equations in the 1970s; for

    linear aerodynamics with boundary-layer corrections in the

    1980s; for Euler equations from the 1980s through the mid-

    1990s; for NavierStokes equations on structured meshesin the 1990s; for Euler and NavierStokes equations on

    unstructured meshes from the late 1990s onward; and for

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

    sion in this chapter is confined to external flows, most of

    the discussion is equally applicable to internal flows, suchas those through aircraft engines. Readers interested in sum-

    maries of the important developments (and contributors) in

    this field should consult review articles such as those by

    Labrujere and Slooff (1993), Newmanet al.(1999), Reuther

    et al.(1999), and Mohammadi and Pironneau (2004), as well

    as the hundreds of references therein.

    Figure 1 illustrates the key processes and variables for

    aerodynamic shape optimization. (The gradient variables are

    shown in parentheses, as they are only relevant for gradient-

    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 descriptions of the surface. The surface description

    is used by the Mesh Generation process, which produces the

    computational mesh y used by the Aerodynamic Analysis

    component. The analysis output is the state variable q (and

    its gradientq). The Objective and Constraint Evaluationprocess supplies the objective functionf, the inequality and

    Objectiveand constraint

    evaluation

    Shapedefinition

    Aerodynamicanalysis

    Optimizationx

    y

    q

    f, gj, hk(f, gj,hk)

    Meshgeneration

    s

    (q)

    Figure 1. Key processes and variables for aerodynamic shapeoptimization.

    DOI: 10.1002/9780470686652.eae500

  • 8/12/2019 eae500

    2/11

    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.

    2 Aerospace System Optimization

    equality constraintsg andh (perhaps along with their gradi-

    ents) back to the Optimization process.

    Those components of the optimization process that are of

    special importance to aerodynamic shape optimization are the

    optimization formulation (objectives and constraints), shape

    definition, mesh generation, and gradient computation (for

    those methods that utilize gradient information). These top-

    ics are covered in that order. The initial expository material

    on the optimization formulation is given in the context of

    linear aerodynamics. Thereafter, the emphasis is on the im-

    portant considerations for aerodynamic shape optimization

    using the Euler or NavierStokes equations, that is, what are

    commonly called computational fluid dynamics (CFD) tools.

    The chapter concludes with some considerations for the use

    of CFD in aerodynamic shape optimization. The discussion

    is restricted to use of optimization for improving an existing

    design, that is, optimization that starts from an existing base-

    line design and is constrained to maintain the same topology

    for the shape.

    2 OPTIMIZATION FORMULATION

    The state variable q is a function of the spatial coordinate,

    which is denoted by the vector x. (The spatial variable has

    the unusual hat because in this volume the symbol x with-

    out the hat is reserved for the vector of design variables.)

    The state variable is a solution of the governing equations for

    an appropriate conceptual model for the flow (compressible

    or incompressible, viscous or inviscid, turbulent or laminar,

    nonlinear or linear). We write the governing equations gener-ically as

    r(q(x),x) = 0 (1)

    The optimization problems that are considered in this

    chapter have the form

    minx

    f(x; q(x; x),x) (2)

    subject to

    gj(x; q(x; x),x) 0, j= 1, . . . , m (3)

    hk(x; q(x;x),x) = 0, k= 1, . . . , p (4)

    xiL xi xiU , i = 1, . . . , n (5)

    where x is the vector of design variables, f is the objec-

    tive function,gjandhkare inequality and equality constraint

    functions, respectively, andxiL and xiU are upper and lower

    bounds on the i-th design variable. The design variables x

    have been added to the argument list of the state variable to

    emphasize its parametric dependence upon them. Likewise,

    we now write the state equation (1) as

    r(q(x;x),x;x)=

    0 (6)

    We necessarily havehk= rk for k= 1, . . . ,dimension ofr,as each component ofr yields a constraint. Because the state

    vector is defined implicitly, as in equation (6), the objective

    and constraint functions in equations (2)(3) have a depen-

    dence uponq andx. For aerodynamic shape optimization, the

    design variables characterize the shape of the airfoil, wing or

    aircraft.

    To illustrate the basic principles, we consider optimization

    of airfoils in the four-digit NACA series (see, e.g., Abbott and

    von Doenhoff, 1959, Chapter 6). These airfoils are described

    by the maximum camber (m), the distance of the location

    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.

    For a four-digit NACA airfoil, say, a NACA d1d2d3d4airfoil,

    the first digit d1 is m (in hundredths), the second digit is p

    (in tenths), and the last two digits are t(in hundredths). Thus,

    a NACA 2416 airfoil hasa maximum camberof 0.20c located

    atx= 0.10c, and a maximum thickness of 0.16c, wherec isthe airfoil chord. Figure 2b is indicative of internal structures

    that constrain the geometry; such important constraints are

    ignored in the present illustration.

    The conceptual model used for the airfoil analysis in this

    example is inviscid, irrotational, incompressible, constant-density flow. The velocity potential can be used as the

    (scalar) state variable. The mathematical model is given by

    the Laplace equation

    r= (7)

    where denotes the gradient operator (with respect to x).The Laplace equation is subject to a homogeneous Neumann

    boundary condition ( n = 0, wheren is unit vector nor-mal to the surface) on the airfoil surface, an homogeneous

    Dirichlet condition (= 0) at infinity, and a Kutta condition(upper and lower surface pressures match) at the trailing edgeThe computational model is based upon a panel (boundary-

    element) method1. This very simple model ignores many im-

    portant effects such as viscosity, compressibility, vorticity,

    and nonlinearity but provides a useful illustration of some

    basic concepts of aerodynamic shape optimization.

    For this example, the design variablesx= (m,p,t), withthese three NACA airfoil parameters treated as continuous

    variables; their upper and lower bounds are taken to be

    xL= (0.0, 0.1, 0.0) andxU= (0.1, 0.5, 0.2), respectively. In

    DOI: 10.1002/9780470686652.eae500

  • 8/12/2019 eae500

    3/11

    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.

    Airfoil/Wing Optimization 3

    Chord c(a) (b)

    Mean camberline Chord

    line

    p c

    m c

    t cLeading edge

    Trailingedge

    z

    x

    Fuel tank Stiffeners and ribs

    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 = 1. The keyperformance outputs are the (section) lift, drag, and pitching

    moment coefficients, cl, cd and cm, respectively. (The mo-

    ment reference center is the quarter-chord point.) Four rep-

    resentative cases are considered: (i) maximize lift (f= cl)with no constraints, (ii) maximize lift with inequality con-

    straints on the pitching moment (g1= cm 0.04) and drag(g2= cd 0.02), (iii) minimize drag (f= cd) with inequal-ity constraints on the pitching moment (g1= cm 0.04)and lift (g2= cm 0.30), and (iv) minimize lift/drag (f=cl/cd) with inequality constraints on the pitching moment(g1= cm 0.04) and lift (g2= cl 0.30. (There are noequality constraints in these examples.) The gradient-based

    optimizations employ Sequential Quadratic Programming,

    using finite differences to compute the gradients and the

    BroydenFletcherGoldfarbShanno (BFGS) method to ap-

    proximate the Hessian2. The starting point for the optimiza-

    tions isx0= (xL+ xU)/2.The optimal solutions to these four problems are given in

    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). Thecorresponding airfoils are

    shown in Figure 3. The key performance parameters are also

    included in the table. The impact of constraints is especially

    dramatic in the contrast between the results for the first two

    cases (maximization of lift). The conceptual model for this

    simple example has neglected several important physical ef-

    fects, most noticeably those of viscosity and compressibility.

    The resulting optimal designs are by no means representa-

    tive of airfoils that would be useful in practice.

    0 0.2 0.4 0.6 0.8 1

    0.2

    0.1

    0

    0.1

    0.2

    x

    z

    max cl

    max cl [g

    i:c

    mandc

    d]

    min cd [g

    i:c

    mandc

    l]

    min cl/ cd [gi:cmand cl]

    Figure 3. Optimal shapes for the NACA four-digit airfoil example.

    Another widely used objective function is the difference

    (in the least squares sense) between the surface pressure

    distribution and a target pressure distribution:

    f(x) =

    S

    || p(x;x) ptarget(x) ||2 (8)

    where the integral is taken over the airfoil or wing surface

    S of interest, p is the pressure in the flow produced by an

    airfoil described by the design variables x, andptarget is the

    target pressure. The basic concept is illustrated in Figure 4.

    The solid line is the pressure coefficient (Cp) for the present

    (baseline) airfoil. The dotted line is the desired (target) pres-

    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 0.3212 cl cmand cd 0.029 0.10 0.15 0.44 0.436 0.0020 0.0403 cd cm and cl 0.016 0.30 0.10 0.0010 0.300 0.0019 0.0364 cl/cd cm and cl 0.028 0.11 0.10 369 0.403 0.0011 0.040

    DOI: 10.1002/9780470686652.eae500

  • 8/12/2019 eae500

    4/11

    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.

    4 Aerospace System Optimization

    Figure 4. Baseline and target pressure distributions (courtesy ofR.L. Campbell).

    years for choosing target pressure distributions to achieve

    desired performance measures. Methods for matching partic-

    ular flow-field characteristics rather than directly minimizing

    a global objective function such as drag are referred to as

    inverse design methods; the more conventional alternative is

    called adirect optimization method. One could apply formal

    optimization to the inverse problem with the objective func-

    tion given by equation (8). However, there is a wide variety

    of inverse design methods that minimize the objective in

    equation (8) much faster than direct optimization methods.

    An example of one such method is provided in Section 6 (see

    Labrujere and Slooff, 1993 for descriptions of many others).

    In practical applications, numerous geometric constraints

    are needed for obtaining an acceptable result, for example,

    requiring the external shape to accommodate internal struc-tures (see Figure 2), minimal radius of curvature of the lead-

    ingedge, andminimal angle of thetrailingedge.Furthermore,

    aerodynamic shapes are required to operate over a range of

    conditions, especially in Mach number. This is addressed

    with a multi-objective optimization formulation (commonly

    referred to as multipoint optimization in the aerodynamic

    optimization community) (see Drela, 1998 for an extensive

    discussion of constraints and objectives for airfoil optimiza-

    tion).

    The NACA airfoil family used for the example above was

    convenient for this demonstration because the parameteriza-

    tion guaranteed a simple airfoil shape. Furthermore, the de-sign variables corresponded directly with physical quantities

    for which experienced aerodynamicists have considerable in-

    tuition. However, the optimization problems of interest for

    many decades now require consideration of a much broader

    range of shapes than those covered by the NACA airfoil fam-

    ilies. In the following section, we cover the fundamentals of

    parameterizations that can represent a more general range of

    shapes.

    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 manyoptimization 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

    parameterization should compactly cover the design space

    of interest while presenting no undue difficulties. In partic-

    ular, parameterization-induced waviness and discontinuities

    in (the slope and curvature of) the surface are undesirable.

    These features are undesirable because they present manu-

    facturing and CFD analysis difficulties (since flow fields are

    quite sensitive to such discontinuities). Sometimes, optimiz-

    ing the shape of the entire surface is desired, but other times

    optimization is only applied to a portion of the surface. The

    alternatives summarized below apply in both cases.

    The usual custom for aerodynamic shape parameterization

    is to represent the surface as a baseline surface plus a pertur-

    bation expressed as an expansion in terms of basis functions.

    Consider the case of airfoils. The surface S is one dimen-

    sional, and the basis functions bk() depend upon a single

    coordinate. The expansion is given by

    s(;x) =d

    k=1xkbk() (9)

    with the coefficients xk in the expansion serving as the de-

    sign variables. The perturbation s(;x) may be applied to

    whatever functions are used for the surface definition, for

    example, upper or lower airfoil surface, mean camber line,

    airfoil thickness. A variety of basis functions have been used,

    some global and some local. Since the leading and trailing

    edges of airfoils are fixed in most design problems, having

    basis functions that vanish at both end points is desirable.

    The sine functions are one option for a complete set of

    global basis functions. Drela (1998) recommends using them

    in the form

    bk() =1

    ksin(k) (10)

    They form an orthogonal set for [0, 1]. The sine basisfunctions are illustrated in Figure 5a. Suitable combinations

    of orthogonal polynomials can also be chosen to ensure that

    DOI: 10.1002/9780470686652.eae500

  • 8/12/2019 eae500

    5/11

    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.

    Airfoil/Wing Optimization 5

    0 0.2 0.4 0.6 0.8 11

    0.5

    0

    0.5

    1

    x x

    bk

    (a)

    k=1

    k=2

    k=3

    k=4

    1 0.5 0 0.5 11

    0.5

    0

    0.5

    1

    bk

    (b)

    k=2

    k=3

    k=4

    k=5

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

    the basis functions vanish at the end points. For example, one

    can use

    bk() =1

    2

    (k + 1) (Lk1() Lk+1()) for k 1 (11)

    for [1, 1], where Lk() is the Legendre polynomial ofdegreek. The set of basis functions given by equation (11) is

    nearly orthogonal the inner product ofbk andbl vanishes

    except forl= k 2, k , k + 2. The Legendre basis functionsare illustrated in Figure 5b. These form a complete set of

    basis functions. They are advantageous for approximating

    functions with a high degree of smoothness (more than, say,

    fourcontinuous derivatives), because theyexhibitmuch fasterconvergencethan thesine functionsin such cases (see Canuto

    et al., 2006, Chapter 2).

    A noteworthy set of local basis functions specifically

    chosen for airfoil parameterization are the HicksHenne

    (Hicks and Henne, 1978) functions, which are defined by

    b() = {sin[log(1/2)/ log(t1)]}t2 (12)

    with the domain now again normalized to [0, 1];t1 con-trols the location of the peak and t2 its width. Figure 6a

    illustrates some of these functions. Although these functions

    lack the completeness property possessed by the sine func-

    tions and the orthogonal polynomials, they have proven very

    useful in the hands of skilled designers.

    A more conventional choice of local basis functions are

    splines, which are piecewise polynomials. Figure 6b illus-

    trates the case of cubic B-splines on uniformly distributed

    knots (denoted by the dots on the x-axis); the curves are la-beled by the xcoordinate of the center of the spline. Unlike

    the global basis functions shown in Figure 5, the underly-

    ing B-spline basis functions effect only local changes in the

    shape. On the other hand, they have only a finite number

    0 0.2 0.4 0.6 0.8 10

    0.25

    0.5

    0.75

    1

    x x

    bk

    (a) (b)

    t1=0.25, t2=0.5

    t1=0.50, t2=1.0

    t1=0.75, t2=1.5

    0 0.2 0.4 0.6 0.8 10

    0.25

    0.5

    0.75

    1

    bk

    c=0.3

    c=0.5

    c=0.7

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

    DOI: 10.1002/9780470686652.eae500

  • 8/12/2019 eae500

    6/11

    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.

    6 Aerospace System Optimization

    0 0.2 0.4 0.6 0.8 10.2

    0.1

    0

    0.1

    0.2

    x x

    z z

    (a) (b)

    Baseline

    Camber perturbation

    0 0.2 0.4 0.6 0.8 10.2

    0.1

    0

    0.1

    0.2

    Baseline

    Thickness perturbation

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

    of continuous derivatives, namely,p 1 continuous deriva-tives for splines of order p. Hence, ifp >2, then the shape

    perturbation, along with its first and second derivatives, is

    continuous. This includes the cubic B-splines (p = 3) shownin Figure 6b. A generalization of B-splines, called nonuni-

    form rational B-splines(NURBS),has become a fairly widely

    used parameterization, particularly for complex shapes, such

    as full aircraft configurations. NURBS represent functions

    (in this case surfaces) as a rational function, with both the

    numerator and denominator consisting of B-spline expan-

    sions. Piegl and Tiller (1996) provide a comprehensive de-

    scription of NURBS. See Samareh (2001) for a summary of

    their mathematical description and a short survey of vari-

    ous uses in aerodynamic shape optimization. An important

    advantage of a NURBS representation for the shape is thatthey are compatible with most computer-aided design (CAD)

    systems. Figure 7 illustrates airfoil deformations based on

    NURBS expansions of the airfoil camber and thickness.

    Using the locations of the surface mesh points as design

    variables is yet another approach. The main challenge here

    is picking appropriate geometric constraints and/or filtering

    procedures to ensure a smooth surface. See Li and Krist

    (2005) for one among many approaches to surface smooth-

    ing.

    For the parameterization of wings, the same approaches

    apply, but, of course, there are now additional types of design

    variables. Some representative wing parameters with a direct

    aerodynamic interpretation are illustrated in Figure 8. The

    symbols indicate locations where the parameters are defined.

    The root chord, tip chord, semi-span, and leading edge

    sweep angle are planform variables; these are usually fixedearly in the design process. The airfoil section parameters

    Camber and thickness

    Twist and shear

    Leading edgesweep angle

    Root chord

    Tip chord

    Semi-span

    y

    x

    Figure 8. Typical design variables for wings.

    DOI: 10.1002/9780470686652.eae500

  • 8/12/2019 eae500

    7/11

    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.

    Airfoil/Wing Optimization 7

    (camber and thickness) as well as the wing twist and shear

    are often the focus of wing shape design. The twist angle at

    a given airfoil section is the difference between the airfoil

    section incident angle at the root and the incident angle of

    that airfoil section. Similarly, the shear (dihedral) is the dif-

    ference between the airfoil leading edge zcoordinate for the

    root and the zcoordinate for the particular airfoil section.

    4 MESH GENERATION

    Unlike panel methods, for which a surface mesh is sufficient,

    CFD methods require a volume mesh for numerical solution

    of the state equation. For compressible flow, the state variable

    qis given by

    q= (, u, E)T (13)

    where is the density, uthe velocity, and the (specific) total

    energyE= e + 12u u, whereedenotes the (specific) inter-

    nal energy. For steady, inviscid flow described by the Euler

    equations, we have

    r(q(x),x) = F (14)

    where the flux F= (u, uuT, uE + pu)T. The NavierStokes equations have the same form as equation (14) but

    with additional terms in the flux function. Figure 9 illustrates

    the surface and symmetry plane portions of an unstructured

    Figure 9. Surface mesh for wing optimization in the presence ofthe fuselage. Reproduced with permission from Nielsen and Park(2006) c AIAA.

    CFD volume mesh for a wingfuselage configuration. The

    coordinates of these volume mesh-points are denoted by yin

    Figure 1.

    For aerodynamic shape optimization, changes to the sur-

    face of the airfoil resulting from design variable changes

    require adjustments to the volume mesh as well, and these

    adjustments need to occur automatically. Nowadays, com-

    putational meshes suitable for even viscous CFD analysis

    can be generated automatically from the surface shape defi-

    nition for airfoils and wings, but this state has proven to be

    very elusive for complex aircraft configurations subjected to

    NavierStokes analyses some type of user intervention is

    typically needed if the mesh generation must be performed

    ab initio. Thompson, Soni and Weatherill (1998) provide an

    extensive description of methods for CFD mesh generation.

    Moreover, as discussed in the next section, analytically based

    gradients are highly desirable, and these are not generally

    available from ab initio mesh generation packages. Hence,

    many aerodynamic shape optimization processes use specialprocesses that lend themselves to analytically based gradi-

    ents for obtaining the volume mesh as a perturbation upon

    the mesh associated with the baseline shape. Samareh (2001)

    contains an overview of the various approaches to volume

    mesh generation.

    5 GRADIENT COMPUTATION

    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 solvednumerically.) 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 q and

    r, respectively, for the latter; similarly y is the vector of the

    mesh-point coordinates. For a problem withNmesh-points,

    the length ofq and r is 5Nsince q has five components at

    each point, and the length ofyis 3N. (Boundary conditions

    may slightly alter these lengths.)

    DOI: 10.1002/9780470686652.eae500

  • 8/12/2019 eae500

    8/11

    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.

    8 Aerospace System Optimization

    Recall that a change in a design variable is propagated

    through the shape definition and mesh generation processes

    (Figure 1), with the computational mesh y depending upon

    x. In particular, the gradient of, say, the objective function

    with respect to the particular design variable xjis given by

    df

    dxj= f

    xj+

    5Nk=1

    f

    qk

    qk

    xj

    +

    3Nl=1

    f

    yl

    yl

    xj

    (15)

    Evaluation of this requires determination of the three explicit

    partial derivatives off plus the qk/xj andyl/xjterms.

    The first term on the right-hand side is usually straightfor-

    ward to compute. (For the airfoil example in Section 2, all

    these derivatives vanish, as the design variables do not appear

    explicitly in the expressions for the lift, drag, and pitching

    moment.) Likewise, an analytical expression can typically

    be derived straightforwardly and then evaluated numerically

    forf/qk and f/yl. Theyl/xjterms represent the influ-ence of the design variables upon the computational mesh.

    The surface definition techniques illustrated for airfoils in

    Figures 57, as well as some mesh perturbation techniques,

    lend themselves to analytical evaluation of these terms.

    Computation of theqk/xjterms is more involved. Con-

    sider the state equation given by equation (6). Its solution

    q, considered as a function of the design variables x, satis-

    fies dr/dx= 0. Hence, implicit differentiation of the stateequation yields the following expression:

    5N

    k=1

    ri

    qk

    qk

    xj =

    ri

    xj

    3N

    l=1

    ri

    yl

    yl

    xj (16)

    Equation (16) is linearin thedesired qk/xj term. Thesize of

    the linear system is 5N 5N. CFD computations for wingstypically use O(106) mesh-points, and complex configura-

    tions can easily take upwards of 108 mesh-points. Hence, the

    linear systemis extremely large in three dimensions so large

    that direct solution methods are impractical. An efficient it-

    erative solution method is surveyed in Newmanet al.(1999)

    and described in detail in Koriviet al.(1994).

    For most aerodynamic shape optimization problems, the

    number of design variables far exceeds the number of ob-

    jectives and constraints. Hence, the adjoint approach for ob-taining the gradients is much more efficient than the direct

    approach described above. Details may be found, for exam-

    ple, in Newmanet al.(1999) and Reutheret al.(1999). Both

    continuous and discrete adjoints have been used in aero-

    dynamics. (The phrase continuous adjoint refers to one in

    which the continuous adjoint equations are first derived from

    the governing partial differential equations (PDEs) and then

    discretized; a discrete adjoint is one in which the adjoint

    equations are derived from the algebraic equations resulting

    from the discretization of the PDEs.) See studies by Newman

    et al. (1999) and Reuther et al. (1999) for extended lists of

    references to the early experiences of various groups with

    both approaches. Some of the subtle issues that must be ad-

    dressed for consistent aerodynamic gradients are the treat-

    ment of boundary conditions and mesh gradients (see Giles

    et al., 2003; Nielsen and Park, 2006 for detailed recommen-

    dations). An approach to obtaining second-order derivatives

    by combining direct and adjoint techniques is described by

    Shermanet al.(1996).

    Yet another alternative to computing the gradients is the

    use of complex variables, which we illustrate on a simple

    real-valued scalar functionf. In this method, the function is

    evaluated at the complex point x + i, where i denotes theimaginary unit, for small. A Taylor expansion yields

    f(x + i) = f(x) + (i)df

    dx + O(2

    ) (17)

    Thus, for small , the real part of f(x + i) is a good ap-proximation tof(x), and the imaginary part divided by is

    a good approximation to the gradient df/dx. Note that this

    approach does not involve the subtraction of nearly equal

    quantities, as occurs for gradients computed via finite dif-

    ferences. Hence, round-off errors are not a concern for the

    complex variable approach. Provided that the source code is

    available, implementation of this approach can require little

    more than changing the type of variables from real to com-

    plex. A prototypical use of this approach for aerodynamic

    shape optimization is given by Nielsen and Kleb (2006).

    6 OPTIMIZATION USING CFD

    Over the past several decades, a broad assortment of direct

    optimization and inverse design methods has been applied to

    aerodynamic shape optimization using CFD. The principal

    distinctions between the approaches are (i) use of global per-

    formance measures such as lift, drag, and pitching moment in

    the objectives and constraints versus inverse objectives such

    as matching a prescribed pressure distribution, (ii) gradient-

    based methods versus methods using only function values,and (iii) having the optimizer invoke the aerodynamic anal-

    ysis tool itself versus having it rely upon surrogate models

    constructed off-line.

    Most of the review articles cited in Section 1 concentrate

    on direct optimization of global performance measures using

    gradient-based methods that directly call CFD analysis tools.

    Several difficulties with this approach can arise. For exam-

    ple, the analysis code may produce solutions that result in

    DOI: 10.1002/9780470686652.eae500

  • 8/12/2019 eae500

    9/11

    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.

    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.

    Nevertheless, many groups have used gradient-based

    optimization of global performance measures without re-

    course to surrogate models to produce impressive CFD-based

    optimization results for multi-element airfoils, wings, wing

    fuselage configurations, and even more complex shapes.

    Figure 10, from Nielsen and Park (2006), shows the objec-

    tive function convergence for unconstrained maximization of

    the lift-to-drag ratio of the vehicle illustrated in Figure 9 attransonic conditions using a NavierStokes CFD code. The

    optimum was achieved in a few dozen function evaluations.

    The surface definition for the wing utilized NURBS rep-

    resentations of the camber, thickness, twist, and shear (see

    Figure 8) perturbations from the baseline shape. (The fuse-

    lage was taken as fixed.) Gradients were computed with a

    quasi-analytical adjoint method.

    One particular approach to inverse design that has

    seen considerable use in industry applications is based on

    Function evaluation

    L/Dratio

    0 5 10 15 20 25 30 35

    L /D = +32%

    Figure 10. Lift/drag convergence from a wing optimization.Reproduced with permission from Nielsen and Park (2006)c AIAA.

    Figure 11. Design improvement regimes for an aircraft (courtesyof J. Hooker and A. Agelastos (Lockheed-Martin) and W. Milholenand R. Campbell (NASA)).

    modifying the surface curvature in proportion to the desired

    change in surface pressure. The surface geometry is modified

    at the same time that the flow solver is converging; hence, this

    method costs little more than a single CFD analysis. Details

    can be found in the study by Campbell (1992). Figure 11,

    taken from joint Lockheed-Air Force-NASA work, illustrates

    the various regions of the aircraft to which this inverse de-

    sign method was applied. (PAI refers to propulsion-airframe

    integration.) Collectively, these improvements in the local

    aircraft shape enabled the design to meet its performance

    goals.

    The most common non-gradient-based optimization

    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 asthe accompanying disadvantage of taking significantly more

    computational time for convergence than gradient-based

    methods.

    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-

    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 thedesign space that cause difficulties for theanalysis

    code, since the surrogate model enables the optimization toproceed 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.

    For the relatively expensive Euler and NavierStokes mod-

    els, use of surrogates entails the usual compromise between

    accuracy (relative to the CFD model) and speed. All types of

    surrogates design of experiments, kriging, neural networks,

    radial basis functions have been fruitfully employed on this

    DOI: 10.1002/9780470686652.eae500

  • 8/12/2019 eae500

    10/11

    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.

    10 Aerospace System Optimization

    application. These types of surrogates all suffer from the curse

    of dimensionality, limiting the number of design variables in

    practice toO(10).

    Of course, one would like optimization using a surrogate

    to converge to the optimum corresponding to that for the

    underlying analysis code. Several optimization methods that

    judiciously mix calls to the surrogate with calls to the actual

    code have been proven mathematically to converge to the

    true optimum. See Bookeret al.(1999) for discussion of one

    method in a general setting, and see Alexandrovet al.(2001)

    for another method with an aerodynamic shape optimization

    application; in practice, the latter seems to reduce the overall

    cost by a factor of 35 compared with always invoking the

    CFD code. When the surrogate is just a lower fidelity model,

    for example, using an Euler code instead of a NavierStokes

    code or a coarse-grid solution in place of a fine-grid solution,

    then the number of design variables is not limited as it is for

    generic surrogates.

    In addition to the considerations mentioned above, oneneeds to weigh the computational costs of the various meth-

    ods, at least for CFD. (Linear aerodynamics methods are

    so fast on desktop computers that CPU time is rarely an

    issue, regardless of the choice of optimization method.)

    Roughly speaking, in terms of the cost of a single CFD anal-

    ysis, an inverse design method takes O(1) analysis time,

    gradient-based optimization using the adjoint formulation

    takesO(10) analysis time, and non-gradient-based methods

    (including surrogate ones) take O(100 10 000) analysistime. All these methods have their niches. Non-gradient-

    based methods and surrogate models enable exploration of

    large regions of the design space. Gradient-based methodsfacilitate refinement of a local optimum, and inverse design

    methods help fine-tune performance in local regions of the

    surface.

    7 DISCLAIMER

    This chapter is declared a work of the U.S. Government and

    is not subject to copyright protection in the United States.

    ACKNOWLEDGMENTS

    The author gratefully acknowledges numerous discussions

    with hiscolleagues at theNASALangley Research Center for

    alltheir contributions to hisunderstanding of this subject. The

    numerical examples in Section 3 utilized publicly available

    codes of L.N. Sankar. J.A. Samareh supplied the code used

    for Figure 7.

    NOTES

    1. The underlying computations are performed with a

    MatlabTM

    code of L.N. Sankar, found on-line at

    http://www.ae.gatech.edu/people/lsankar/AE3903/.

    2. Performed with the MatlabTM routinefmincon.

    3. These gradients are often calledsensitivity derivatives.

    REFERENCES

    Abbott, I.A. andvon DoenhoffA.E. (1959) Theory of Wing Sections,Dover Publications, New York.

    Alexandrov, N.M., Lewis, R.M., Gumbert, C.R., Green, L.L. and

    Newman, P.A. (2001) Approximation and model management inaerodynamic optimization with variable-fidelity models. J. Air-

    craft,38(6), 10931101.Booker,A.J., Dennism, J.E.,Jr.,Frank, P.D.,Serafini, D.B.,Torczon,

    V. and Trosset, M.W. (1999) A rigorous framework for optimiza-tion of expensive functions by surrogates. Struct. Optim.,17(1),113.

    Campbell, R.L. (1992) An approach to constrained aerodynamicdesign with application to airfoils. NASA-TP-3260.

    Canuto, C., Hussaini, M.Y., Quarteroni, A., and Zang, T.A. (2006)

    Spectral Methods, Fundamentals in Single Domains, Springer,

    New York.

    Drela, M. (1998) Pros and cons of airfoil optimization, in Frontiersof Computational Fluid Dynamics (eds D.A. Caughey, and M.M.Hafez), World Scientific, Hackensack, pp. 363381.

    Giles,M.B., Duta, M.C.,Muller, J.-D.and Pierce, N.A.(2003) Algo-rithm developments for discrete adjoint methods.AIAA J.,41(2),198205.

    Hicks, R.M. and Henne, P.A. (1978) Wing design by numericaloptimization.J. Aircraft,15(7), 407412.

    Holst, T. (2005) Genetic algorithms applied to multi-objective

    aerospace shape optimization. J. Aero. Comput. Info. Comm.,2,217235.

    Korivi, V.M., Taylor, A.C. Ill, Newman, P.A., Hou, G.J.-W. andJones, H.E. (1994) An approximate factored incremental strategyfor calculating consistent discrete CFD sensitivity derivatives.J.Comput. Phys.,113(2), 336346.

    Labrujere, Th.E. and Slooff, J.W. (1993) Computational methodsfor the aerodynamic design of aircraft components. Ann. Rev.Fluid Mech.,25, 183214.

    Li, W. and Krist,S. (2005) Spline-based airfoil curvature smoothingand its applications.J. Aircraft,42(4), 10651074.

    Mohammadi, B. and Pironneau, O. (2004) Shape optimization influid mechanics.Ann. Rev. Fluid Mech.,36, 255279.

    Newman, J.C. III, Taylor, A.C. III, Barnwell, R.W., Newman,P.A. and Hou, G.J.-W. (1999) Overview of sensitivity analysisand shape optimization for complex aerodynamic configurations.AIAA J.,36(1), 8796

    DOI: 10.1002/9780470686652.eae500

  • 8/12/2019 eae500

    11/11

    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

    Airfoil/Wing Optimization 11

    Nielsen, E.J. and Kleb, W.L. (2006) Efficient construction ofdiscrete adjoint operators on unstructured grids using complexvariables.AIAA J.,44(4), 827836.

    Nielsen, E.J. and Park, M.A. (2006) Using an adjoint approach toeliminate mesh sensitivities in computational design. AIAA J.,44(5), 948953.

    Piegl, L. and Tiller, W. (1996) The NURBS Book, Springer,New York.

    Reuther, J.J., Jameson, A., Alonso, J.J., Rimlinger, M.J. andSaunders, D. (1999) Constrained multipoint aerodynamic shapeoptimizationusing an adjoint formulationand parallel computers.Parts 1 and 2.AIAA J.,36(1), 5174.

    Samareh, J.A. (2001) Survey of shape parameterization techniquesfor high-fidelity multidisciplinary shape optimization. AIAA J.,39(5), 877884.

    Sherman, L.L., Taylor, A.C. III, Green, L.L., Newman, P.A., Hou,G.J.-W. and Korivi, V.M. (1996) First- and second-order aerody-namic sensitivity derivatives via automatic differentiation with

    incremental iterative methods. J. Comput. Phys., 129(2), 307331.

    Thompson, J.R., Soni, B.K. and Weatherill, N.P. (eds) (1998)

    Handbook of Grid Generation, CRC Press, Boca Raton.