You are on page 1of 11

ICAS 2002 CONGRESS

FROM ANALYSIS TO DESIGN OF HIGH-LIFT CONFIGURATIONS USING A


NEWTON–KRYLOV ALGORITHM

Marian Nemec and David W. Zingg


University of Toronto Institute for Aerospace Studies
4925 Dufferin Street, Toronto, ON
Canada, M3H 5T6
marian@oddjob.utias.utoronto.ca

Keywords: optimal shape design, airfoil, Navier–Stokes, high-lift, discrete gradient, adjoint, flow
sensitivities, Newton’s method, GMRES

Abstract the design of high-lift multi-element configura-


tions is an active area of research [5, 7, 2, 11, 1,
An efficient multi-block Newton–Krylov algo- 19, 13]. This challenging optimization problem
rithm using the compressible Navier–Stokes features complex geometry, with strict geometry
equations is presented for the analysis and de- constraints, and complex flow physics, such as
sign of high-lift airfoil configurations. The pre- regions of separated flow and confluent boundary
conditioned generalized minimum residual (GM- layers and wakes.
RES) method is applied to solve the discrete- A gradient-based algorithm for aerodynamic
adjoint equation, leading to a fast computation shape optimization problems can be divided into
of accurate objective function gradients. Further- four modules: 1) an optimizer, 2) a flow solver, 3)
more, the GMRES method is used in conjunc- a gradient computation algorithm, and 4) a grid-
tion with an inexact-Newton approach to obtain perturbation strategy. The accuracy of the opti-
fast solutions of the Navier–Stokes equations. mization ultimately depends on the modeling of
Optimization constraints are enforced through a the flowfield, and hence, the flow solver. Accu-
penalty formulation, and the resulting uncon- rate modeling of the complex flowfields encoun-
strained problem is solved via a quasi-Newton tered in high-lift problems requires the solution
method. Several design examples are provided of the compressible Navier–Stokes equations in
which demonstrate that this algorithm provides conjunction with a turbulence model. A detailed
an effective and practical tool for the design of review of flow solvers and results for the pre-
multi-element airfoil configurations. diction of high lift is provided by Rumsey and
Ying [24].
1 Introduction For a given number of objective function and
gradient evaluations, the efficiency of the opti-
The design of an efficient high-lift configuration mization procedure is dominated by the time re-
can significantly improve the aerodynamic per- quired to solve the flowfield equations and com-
formance of an aircraft, as well as provide weight pute the gradient. In particular, significant com-
savings and reductions in the complexity of the putational effort is required to obtain accurate
high-lift system [27]. Consequently, the applica- gradients. Generally, the cost of one gradient
tion of gradient-based optimization algorithms to evaluation is equivalent to between one and two

Copyright c 2002 by M. Nemec and D.W. Zingg. flowfield solutions [23, 18, 19, 11, 6, 12].
Published by the International Council of the Aeronautical In [16], we presented an accurate and efficient
Sciences, with permission.

173.1
MARIAN NEMEC AND DAVID W. ZINGG

algorithm for the calculation of the objective 0.3


Original Points
Fitted Points
function gradient via the discrete-adjoint method. 0.2
Control Points
The adjoint equation is solved using the pre- 0.1
5

Y/C
12
conditioned generalized minimum residual (GM- 0 1 4 FY
1
RES) Krylov subspace method [25]. Overall, the -0.1
8 FX
gradient is obtained in just one-fifth to one-half
-0.2
0 0.25 0.5 0.75 1 1.25
of the time required for a warm-started flow solu- X/C

tion. The performance of the algorithm is demon-


Fig. 1 B-spline curves and flap translation design
strated for several design examples, all based on
variables
single-element airfoils.
The objectives of this work are to extend and
apply the Newton–Krylov algorithm presented OVERLAP (-)

in [16] to the analysis and design of high-lift


multi-element configurations. The new algo-
rithm is based on an established flow solver TOR- GAP

NADO [15, 9, 14]. The validation and perfor-


mance of the new algorithm are demonstrated
for lift-enhancement design examples based on
a two-element take-off configuration.
Fig. 2 Definition of gap and overlap distances
2 Problem Formulation
where CD and CL represent the target drag and lift
The aerodynamic shape optimization problem
coefficients, respectively. The weights ωD and
consists of determining values of design variables
ωL are user specified constants. This objective
X , such that the objective function J is minimized
can be used for both lift-enhancement and lift-
min J (X ; Q) (1) constrained drag minimization problems.
X The design variables are based on a B-spline
subject to constraint equations C j , parameterization [3, 16] of the airfoil. An exam-
ple is shown in Fig. 1, where a B-spline curve
C j (X ; Q)  0 j = 1; : : : ; Nc (2) is fitted over the upper surface of the main ele-
ment, and also the upper surface of the flap for the
where the vector Q denotes the conservative flow- NLR 7301 configuration [28]. The vertical coor-
field variables and Nc denotes the number of con- dinates of the B-spline control points are used as
straint equations. The flowfield variables are design variables. Depending on the problem of
forced to satisfy the governing flowfield equa- interest, additional design variables may include
tions, F , within a feasible region of the design the angle of attack, and the horizontal and verti-
space Ω, cal translation associated with each high-lift ele-
ment in multi-element configurations, labeled as
F (X ; Q) = 0 8X 2Ω (3) Fx and Fy in Fig. 1. The horizontal and vertical
translation design variables control the gap and
which implicitly defines Q = f (X ). overlap distances in the slot region of the airfoil,
The objective function is given by as defined in Fig. 2.
8  2  2 The constraint equations, Eq. 2, represent air-
if CD > CD
>
< ωL 1 + ωD
CL CD
CL 1 
CD foil thickness constraints that are used to ensure
J =  2
>
: ωL 1 CL feasible designs. The constraints are given by
CL otherwise
(4) h (z j ) h(z j )  0 (5)

173.2
From Analysis to Design of High-Lift Configurations Using a Newton–Krylov Algorithm

where h (z j ) represents the minimum allowable construct a local cubic interpolant. Note that
thickness at location z j expressed as a fraction of the optimization problem is based on the discrete
the airfoil’s chord. For multi-element configura- form of the flowfield equations. Using the dis-
tions, it is also necessary to constrain the gap and crete approach, we expect the gradient to vanish
overlap distances. These constraints are required at the optimum solution. In the following sec-
in order to prevent collisions among the elements tions, we present the formulation for the penal-
and to ensure a reasonable computational grid. ized objective function, as well as the algorithms
The governing flow equations are the com- used for the flowfield evaluation (objective func-
pressible two-dimensional thin-layer Navier– tion value), the gradient evaluation, and the grid-
Stokes equations in generalized coordinates, perturbation strategy.
∂Eb(Qb; X ) ∂F
b(Qb; X ) b b
1 ∂S(Q; X ) 3.1 Objective with Constraints
+ = Re (6)
∂ξ ∂η ∂η
A penalty method is used to combine the objec-
where Q b = J 1 Q = J 1 [ρ; ρu; ρv; e℄T is the vec- tive function with the constraint equations
tor of conservative dependent state variables, ξ
and η are the streamwise and normal general- Nc

ized coordinates, respectively, and J is the Jaco-


J = JO + ωT ∑ Cj (7)
j=1
bian of the coordinate transformation from Carte-
sian coordinates. Vectors Eb and Fb represent the where JO refers to Eq. 4. The constraint equa-
convective flux vectors, the viscous flux vector is tions represent thickness, gap, and overlap con-
b and Re denotes the Reynolds number.
given by S, straints, which are cast using a quadratic penalty
Sutherland’s law is used to determine the laminar term. For example, the thickness constraint,
viscosity. The equations are in non-dimensional based on Eq. 5, is given by
form. For further details, see [22]. The turbulent   2
viscosity is modeled with the Spalart–Allmaras 1 h(z j )=h (z j )
if h(z j ) < h (z j )
Cj =
turbulence model [26]. All cases considered in 0 otherwise
this study are assumed to be fully turbulent, and (8)
therefore, the laminar-turbulent trip terms are not where ωT is a user specified constant. A similar
used. quadratic term is used to enforce the lower and
upper bounds for the gap and overlap distances.
3 Numerical Method
3.2 Flowfield Evaluation
The aerodynamic shape optimization problem
The spatial discretization of the flowfield equa-
defined by Eqs. 1–3 is cast as an unconstrained
tions, Eq. 6, is the same as that used in TOR-
problem. This is accomplished by lifting the side
NADO [15] (see also ARC2D [22]) for structured
constraints, Eq. 2, into the objective function J
multi-block H-topology grids. The discretization
using a penalty method. Furthermore, the con-
consists of second-order centered-difference op-
straint imposed by the flowfield equations, Eq. 3,
erators with second- and fourth-difference scalar
is satisfied at every point within the feasible de-
artificial dissipation. The Spalart–Allmaras tur-
sign space, and consequently these equations do
bulence model is discretized as described in [26,
not explicitly appear in the formulation of the op-
9]. Overall, the spatial discretization leads to a
timization problem.
nonlinear system of equations
The unconstrained problem is solved using
the BFGS quasi-Newton method in conjunction b) = 0
R(X ; Q (9)
with a backtracking line search [20, 16]. At
each step of the line search, the objective func- Although R is written as a function of the design
tion value and gradient are required in order to variables, we emphasize that during a flowfield

173.3
MARIAN NEMEC AND DAVID W. ZINGG

solution the design variables, and consequently and storage requirements of the incomplete fac-
the computational grid, are constants. torization. This approach is similar to the ‘di-
Eq. 9 is solved in a fully-coupled manner, agonal shift’ strategy suggested by Chow and
where convergence to steady state is achieved Saad [4]. The present preconditioning matrix is a
using the preconditioned GMRES algorithm in compromise between a preconditioner based on a
conjunction with an inexact-Newton strategy [21, first-order upwind discretization of the flowfield
16]. The main components include matrix- equations and a preconditioner based on the ac-
free GMRES(40) and a block-fill incomplete tual second-order discretization. This novel ‘in-
LU (BFILU) preconditioner. The matrix-vector termediate’ preconditioner is significantly more
products required at each GMRES iteration are effective than either of these more commonly
formed with first-order finite-differences. Right used approaches.
preconditioning is used, and the preconditioner is
based on an approximate-flow-Jacobian matrix. 3.3 Gradient Evaluation
The level of fill for most cases is 2 [BFILU(2)],
The gradient, G , of the objective function
but difficult multi-element cases may require
J [X ; Q(X )℄ is given by
BFILU(4). The approximate-factorization algo-
rithm [22, 15] is used to reduce the initial resid- dJ ∂J ∂J dQ
ual by three orders of magnitude in order to avoid G= = + (11)
dX ∂X ∂Q dX
Newton startup problems.
The approximate-flow-Jacobian matrix used where we reduce the vector of design variables,
for the preconditioner is identical to the flow- X , to a scalar in order to clearly distinguish be-
Jacobian matrix, ∂R=∂Q, except for the treat- tween partial and total derivatives. For problems
ment of the artificial-dissipation coefficients [16]. with multiple design variables, it may be helpful
Hence, the preconditioner contains the contribu- to note that G and ∂J =∂X are [1  ND ℄ row vec-
tions from all components of the residual vec- tors, ∂J =∂Q is a [1  NF ℄ row vector, and dQ=dX
tor, namely inviscid and viscous fluxes, bound- is a [NF  ND ℄ matrix, where ND represents the
ary conditions, block interfaces, and the tur- number of design variables and NF represents the
bulence model. The artificial-dissipation coef- number of flowfield variables. We assume that
ficients, which include the spectral radius and the implicit function Q(X ) is sufficiently smooth
the pressure switch, are assumed to be constant even in the presence of flow discontinuities such
with respect to the flowfield variables. Fur- as shock waves. See [10, 8] for further details.
thermore, the preconditioning matrix is formed The difficulty in Eq. 11 is the evaluation of
with only second-difference dissipation, but the the term dQ=dX , referred to as the flow sensitivi-
second-difference coefficient is combined with ties. Evaluation of the partial derivatives, ∂J =∂X
the fourth-difference coefficient as follows, and ∂J =∂Q, is relatively straightforward and is
(2) (2) (4)
described at the end of this section. The flow sen-
dl = dr + φ dr (10) sitivities are obtained by differentiating the flow-
where the subscript r denotes the contribution field equations, Eq. 9, with respect to the design
from the right-hand side, and the subscript l de- variables, which yields the following large sys-
notes the resulting left-hand side value used in tem of linear equations
forming the preconditioner. This modification ∂R dQ ∂R
does not affect the steady-state solution. Fast = (12)
∂Q dX ∂X
convergence is obtained with the value of φ set
to 6.0, which has been determined through nu- The direct, or flow-sensitivity, method results
merical experiments. from solving Eq. 12 for the flow sensitivities
Eq. 10 improves the diagonal dominance of dQ=dX and using these values in Eq. 11 to ob-
the preconditioning matrix and reduces the work tain the gradient.

173.4
From Analysis to Design of High-Lift Configurations Using a Newton–Krylov Algorithm

In order to formulate the discrete adjoint sensitivities is given by


method, substitute Eq. 12 into Eq. 11 and define
the following intermediate problem ∂R R(X + hei ; Q) R (X hei ; Q)
= (16)
∂X i 2h
  1
∂J ∂R where
ψ T
= (13)
∂Q ∂Q 
h = max ε jXij; 1  10 8
(17)
where ψ is a [NF  1℄ column vector. Post-
multiplication of both sides by ∂R=∂Q and apply- and i = 1; : : : ; ND . The ith unit vector is denoted
ing the transpose operator results in the following by ei , and a typical value of ε is 1  10 5 . It
linear system of equations is important to realize that Eq. 16 involves two
evaluations of only the residual vector per design
∂R T ∂J T variable and not two flowfield solutions. Note
ψ= (14)
∂Q ∂Q that the evaluation of residual sensitivities in-
cludes the evaluation of grid sensitivities, since
This is known as the adjoint equation, and the the design variables do not explicitly appear in
vector ψ represents the adjoint variables. The ex- the residual equations except for the angle of at-
pression for the gradient becomes tack design variable, see [16] for further details.

dJ ∂J ∂R 3.4 Grid-Perturbation Strategy


= ψT (15)
dX ∂X ∂X
As the shape and position of an airfoil evolve
The GMRES strategy from the flow solver, during the optimization process, the location of
discussed in Subsection 3.2, is adopted to solve the grid nodes is adjusted from the baseline con-
both the adjoint and flow-sensitivity equations. figuration to conform to the new configuration.
Fast adjoint and flow-sensitivity solutions are ob- In [16], we use an algebraic grid-perturbation
tained with BFILU(6), GMRES(85), and φ = 6:0. strategy that preserves the distance to the outer
For the flow-sensitivity equation, we use matrix- boundary and relocates the grid nodes in the
free GMRES. In addition to memory savings, normal direction proportional to the distance
the matrix-free approach is easier to implement, from the airfoil boundary. When the optimiza-
since an accurate differentiation of cumbersome tion problem involves the horizontal and vertical
functions in the residual equations, such as the translation of a slat or a flap, the use of this strat-
absolute value and min = max functions, is ‘auto- egy can result in significantly skewed grid cells
matically’ provided. Due to the transpose on the near the boundary.
left-hand-side of Eq. 14, the matrix-free approach In order to improve the quality of the mod-
is not possible for the adjoint equation. The flow- ified grid, we present a new grid-perturbation
Jacobian matrix is stored explicitly, where we in- strategy given by
clude the contribution from the spectral radius,
but we treat the pressure switch associated with ∆y
ynew [1 + cos (πSk )℄
old
k = yk + (18)
the artificial dissipation scheme as a constant. 2
The remaining terms in Eqs. 11 and 15, where ∆y represents the airfoil shape change. Sk
namely the objective function sensitivities is the normalized arclength distance given by
∂J =∂X and ∂J =∂Q, as well as the residual
sensitivity ∂R=∂X , are evaluated using centered 1 k
Lg i∑
differences. The use of centered differences for Sk = Li k = 2; : : : ; kmax 1 (19)
=2
the evaluation of the partial derivative terms is
not computationally expensive. For example, where S1 = 0, Li is the length of a segment be-
the centered-difference formula for the residual tween nodes k and k 1, and Lg is the grid-line

173.5
MARIAN NEMEC AND DAVID W. ZINGG

length from the body to the outer boundary. An 0


evaluation of the new grid-perturbation strategy
-2 A-F
is presented in [17], where we demonstrate that N-K
the new strategy provides superior estimates of -4
aerodynamic performance.

log( R )


-6

4 Validation -8

-10
Before presenting design examples for high-lift
configurations, we carefully validate the perfor- -12
mance of the flow solver and the gradient com-
-14
putation algorithm. All grids for this and the fol-
0 200 400 600 800
lowing sections consist of approximately 31,000
nodes. The off-wall spacing is 2  10 6 c, the
Time (s)

distance to the outer boundary is 12c, the spac- Fig. 3 Flow-solver performance
ing at the H-topology grid stagnation points is
2  10 5 c, and the trailing-edge clustering is
2  10 3 c. The reported CPU times are obtained started flow solves typically converge in 2/3 of
on a 667 MHz Alpha 21264 processor (SPECfp the original flow solve time.
2000 rating of 562 peak).
4.2 Gradient Accuracy
4.1 Flow-Solver Performance
Finite-difference gradients provide a benchmark
A fast solution of the flowfield equations is a that is used to establish the accuracy of the gra-
critical component of an effective design algo- dient computation using the flow-sensitivity and
rithm, since an evaluation of the objective func- adjoint methods. A subsonic lift-enhancement
tion is required at each iteration of the optimizer. problem for the NLR 7301 configuration is con-
The performance of the flow solver is examined sidered. During the computation of the finite-
for the NLR 7301 configuration at M∞ = 0:25, difference gradient, the flowfield solution is con-
α = 8Æ and Re = 2:51  106 . Fig. 3 shows that verged 14 orders of magnitude. The flow-
the Newton–Krylov flow solver (denoted as NK) sensitivity and adjoint equations are converged 8
is approximately two to three times faster than the orders of magnitude.
approximate-factorization flow solver (denoted The freestream conditions are M∞ = 0:25,
as AF). For many cases, this speed-up can be α = 4Æ , and Re = 2:51  106 . We compute the
even larger. Initially, the convergence rate of gradient of the objective function, Eq. 4, with re-
both flow solvers is identical, since approximate- spect to control point 5 on the main airfoil (de-
factorization is used as a startup procedure for the noted as 5M), control point 4 on the flap (denoted
Newton–Krylov flow solver. as 4F), and the horizontal and vertical flap dis-
One of the main difficulties associated with placements (denoted as Fx and Fy , respectively),
Newton’s method is the startup procedure. The see Fig. 1. The target drag coefficient, CD , is set
Newton–Krylov flow solver is particularly well equal to the initial drag coefficient, while the tar-
suited for the design problem since once we ob- get lift coefficient, CL , is set equal to 2:2, which
tain the solution for the initial airfoil shape, we represents a 2.5% increase from the initial value.
warm-start the remaining flow solves. If the step- The values of ωL and ωD in Eq. 4 are both set
sizes during the line-search procedure are suffi- to 1.0 and there are no side constraints. Ta-
ciently small, the startup procedure using approx- ble 1 shows that there is an excellent agreement
imate factorization is not necessary. The warm- between the finite-difference, adjoint, and flow-

173.6
From Analysis to Design of High-Lift Configurations Using a Newton–Krylov Algorithm

0 S-MF D.V. 1
Table 1 Gradient accuracy S-MF D.V. 2
-1 S-MF D.V. 3
Design Finite Adjoint S-MFa S-MF D.V. 4
Adjoint
Variable Difference (% Diff.) (% Diff.)b
b -2

log( Residual )


5M -0.01228 0.02 -0.07 -3
4F -0.08533 -0.19 -0.07
-4
Fx -0.02591 0.06 -0.02
Fy -0.03363 -0.05 -0.05 -5

a matrix-free flow-sensitivity -6
b % Diff = (G GFD )=GFD  100
-7

-8
10 20 30 40 50 60 70 80
GMRES Iterations
sensitivity gradients.
Fig. 4 GMRES convergence (S-MF denotes
4.3 Gradient Computation Efficiency matrix-free flow-sensitivity, D.V. denotes design
variable)
Fig. 4 shows the convergence of the ad-
joint and flow-sensitivity equations for the lift-
enhancement problem introduced in the previous only three orders of magnitude in order to ob-
subsection. The residuals of the flow-sensitivity tain gradients of sufficient accuracy [23, 18, 12].
equation are shown for each design variable. This level of convergence is achieved in approx-
Fig. 4 highlights the influence of different right- imately 45 seconds, as shown in Fig. 5. For
hand-side vectors on the convergence of GM- the flowfield equations, we typically reduce the
RES(85). Note that the initial guess for the ad- residual by ten orders of magnitude in order to
joint and flow-sensitivity solution vectors is zero. prevent stalling of the line searches once the so-
The left-hand side of the flow-sensitivity equa- lution is close to the optimum. This conver-
tion, Eq. 12, is the flow-Jacobian matrix. This gence level is achieved in 245 seconds, and con-
matrix remains the same for each design variable. sequently, the gradient is obtained in less than
For the adjoint equation, Eq. 14, the left-hand one-fifth of the flow solve time.
side is the transpose of the flow-Jacobian ma-
trix. The transpose operator does not change the
5 Design Examples
eigenvalues of the flow Jacobian. Although the
flow-sensitivity equation converges faster than
5.1 Flap Position Optimization
the adjoint equation, the flow-sensitivity equation
must be solved for each design variable, resulting The goal of this design example is to determine
in a much longer gradient computation time. the optimal gap and overlap distances for the
Fig. 5 compares the convergence history of NLR 7301 configuration, such that the result-
the adjoint and flowfield equations with respect ing configuration achieves a higher lift coefficient
to CPU time. The time for the formation of the while maintaining the same (or lower) drag coef-
preconditioning matrices is included in Fig. 5.1 ficient. The freestream conditions are specified
It is necessary to converge the adjoint equation in Subsection 4.2. The initial values of CL and
CD are 2.145 and 0.04720, respectively. The ob-
1 In Fig. 5, the ‘flat step’ in the convergence of the jective function is given by Eq. 4, where we set
CL = 2:180 and CD equal to the initial drag coeffi-
flow solver after a three order-of-magnitude decrease in the
residual indicates the formation time of the preconditioner.
For the adjoint equation, this time is indicated at the start cient. The weights ωL and ωD are set to 1.0. The
of the convergence history. design variables are the horizontal and vertical

173.7
MARIAN NEMEC AND DAVID W. ZINGG

Adjoint
-2 Flow Solve
Table 2 Gap-overlap optimization summary

Design CL CD Ga Ob
log( Residual )

-4
NLR 7301 2.145 0.04720 2.40 -5.31
-6 #1 2.165 0.04687 1.99 -4.28
#2 2.173 0.04677 1.95 -3.30
-8 Final 2.175 0.04675 2.02 -2.68
Target 2.180  0.0472
-10 a Gap (%c)
b Overlap (%c)
-12
0 50 100 150 200 250 300
CPU Time (s)

Fig. 5 Comparison of adjoint and flow solve con-


vergence times

0.05 NLR 7301


displacements of the trailing edge of the flap, as Design #1
Design #2
indicated in Fig. 1. The gap and overlap limits are 0.025 Final Design

set to 0:5%c and 1:0%c, respectively, based


Y/C

0
on the initial configuration. The weight, ωT , as-
sociated with the gap and overlap constraints is -0.025
set to 0:05. The gradient is computed using the
-0.05
adjoint method.
0.85 0.9 0.95 1
Table 2 and Fig. 6 summarize the results. X/C
Within a few flowfield and gradient evaluations,
the flap reaches the maximum allowable over- Fig. 6 Flap position summary
lap distance of approximately 4:3%c, at which
point the overlap penalty function becomes ac-
tive. The optimization converges to the design
#1 configuration, shown in Fig. 6. A new grid
is generated for this configuration and the corre- -5 -2.5

sponding values of CL and CD are given in Ta- -2


-4
ble 2. The optimization is restarted from the new NLR 7301
Final Design -1.5
grid with the same objective function. This pro- -3
cedure is continued until convergence to the final -1
Cp

Cp

-2
design is obtained (see Fig. 6), where the gap and -0.5
overlap constraints are no longer active. Note that -1
0
the drag objective is satisfied for all the designs.
Consequently, the optimization is purely attempt- 0 0.5

ing to maximize the lift coefficient. Overall, a 1 1


0 0.2 0.4 0.6 0.8 0 0.1 0.2 0.3
1:4% increase in the value of the lift coefficient is X/C X/C
obtained. This is achieved by an increased load-
ing on the main element as well as the flap, as Fig. 7 Cp distribution for main element and flap
shown in Fig. 7.
Example convergence histories for the design

173.8
From Analysis to Design of High-Lift Configurations Using a Newton–Krylov Algorithm

0 3
Design #2
Final Design 2.9
-1
2.8
-2
2.7
G24-O53
log( Gradient )

-3 2.6 G29-O05

Gap (%c)
2.5
-4
2.4
-5 2.3

2.2
-6
2.1
-7 1
2 2

-8 1.9
5 10 15 20 -5 -4 -3 -2 -1 0
Flow Solves and Gradient Evaluations Overlap (%c)

Fig. 8 Convergence histories for gap-overlap op- Fig. 9 Convergence to optimal gap-overlap dis-
timization tances from two distinct initial conditions

#2 and final configurations are shown in Fig. 8.


The oscillations in the L2 norm of the gradient
for design #2 are due to the presence of the gap
and overlap constraints. The norm of the gradient
is reduced by several orders of magnitude, which
indicates that the optimization converged to a lo-
cal minimum.
Given that the target value of the lift coeffi-
Fig. 10 Flap shape and position design variables
cient is not achieved at the final design configura-
tion (see Table 2), it is somewhat surprising that
further design improvements cannot be realized 5.2 Flap Shape and Position Optimization
by further extending the effective chord of the
configuration. The convergence of the gradient The flap position optimization example presented
in Fig. 8 indicates that a local optimum has been in Subsection 5.1 is expanded to include flap
found, but a global optimum is not guaranteed. In shape changes. As shown in Fig. 10, a B-spline
order to verify the uniqueness of the optimal so- curve is fitted over a portion of the upper surface
lution, the optimization is restarted from a differ- of the flap, such that the cruise (flap-stowed) con-
ent initial condition. The flap is re-positioned to a figuration is not affected by the shape modifica-
gap of 2:9%c and an overlap of 0:5%c, i.e., the tions. The design variables consist of the four
leading edge of the flap is almost aligned with the shaded control points, as well as the horizontal
trailing edge of the main element. Fig. 9 shows and vertical flap displacements.
that the optimization converges to the same opti- The objective function and all the optimiza-
mum solution. The data labeled ‘G24-O53’ show tion parameters remain unchanged from Sub-
the convergence to the optimum solution from the section 5.1 except for the target lift coefficient,
original configuration, with designs #1 and #2 in- which is increased to 2.2. Table 3 and Fig. 11
dicated, while the data labeled ‘G29-O05’ show summarize the results. The optimization is
the convergence to the same optimum solution started from the optimal gap and overlap values
from the new initial conditions. obtained previously, which is denoted as the ini-
tial configuration in Table 3 and Fig. 11. For the

173.9
MARIAN NEMEC AND DAVID W. ZINGG

-2.5
Table 3 Flap optimization summary -5

-2
Design CL CD Ga Ob -4 NLR 7301
Final Design
NLR 7301 2.145 0.04720 2.40 -5.31 -3
-1.5

Initial 2.175 0.04675 2.02 -2.68 -1

Cp

Cp
Final 2.2 0.04723 2.06 -2.40 -2

Target 2.2  0.0472 -1


-0.5

0
a Gap (%c)
b 0
Overlap (%c) 0.5

1 1
0 0.2 0.4 0.6 0.8 0 0.1 0.2 0.3
X/C X/C
0.05
NLR 7301
Initial Fig. 12 Cp distribution for main element and flap
0.025 6 D. V.
Y/C

0
mately one-fifth of the flow solve time when us-
-0.025 ing the adjoint method. Overall, the results indi-
-0.05
cate that the new algorithm provides an efficient
0.9 0.95 1 1.05 and robust tool for complex aerodynamic design
X/C
problems.
Fig. 11 Flap shape and position summary (D.V.
denotes design variable) Acknowledgments

This research was supported by the Natural


final design, the gap distance remains approxi-
Sciences and Engineering Research Council of
mately constant, but the thickness of the flap in-
Canada, as well as Bombardier Aerospace and an
creases considerably near the leading edge, as
OGS grant from the Government of Ontario.
shown in Fig. 11. The optimization converges
in 25 flowfield and gradient evaluations. A new
grid is generated for the final configuration, and References
the corresponding values of the lift and drag co-
[1] Alexandrov NM, Nielsen EJ, Lewis RM, and
efficients are provided in Table 3. Overall, the
Anderson WK. First-order model manage-
final design achieves a 2.5% increase in the lift
ment with variable-fidelity physics applied to
coefficient value while almost no drag penalty
multi-element airfoil optimization. AIAA Paper
is incurred when compared with the original 2000–4886, September 2000.
NLR 7301 configuration. The pressure distribu-
[2] Anderson WK and Bonhaus DL. Airfoil design
tions for the main element and flap are shown in on unstructured grids for turbulent flows. AIAA
Fig. 12. Journal, Vol. 37, No 2, pp 185–191, 1999.
[3] Anderson WK and Venkatakrishnan V. Aerody-
6 Conclusions namic design optimization on unstructured grids
with a continuous adjoint formulation. Comput-
A Newton–Krylov algorithm for the analysis and ers & Fluids, Vol. 28, pp 443–480, 1999.
design of multi-element airfoil configurations has [4] Chow E and Saad Y. Experimental study of ILU
been presented. The accuracy of the objective preconditioners for indefinite matrices. Jour-
function gradient, based on the discrete-adjoint nal of Computational and Applied Mathematics,
and flow-sensitivity methods, is excellent. Fur- Vol. 86, pp 387–414, 1997.
thermore, the gradient is obtained in approxi- [5] Drela M. Design and optimization method for

173.10
From Analysis to Design of High-Lift Configurations Using a Newton–Krylov Algorithm

multi-element airfoils. AIAA Paper 93–0969, gorithm for complex aerodynamic design. Proc
1993. 10th Annual Conference of the CFD Society of
[6] Elliott JK and Peraire J. Aerodynamic opti- Canada, Windsor, ON, June 2002.
mization on unstructured meshes with viscous [18] Nielsen EJ. Aerodynamic Design Sensitivi-
effects. AIAA Paper 97–1849, June 1997. ties on an Unstructured Mesh Using the Navier
[7] Eyi S, Lee KD, Rogers SE, and Kwak D. High- Stokes Equations and a Discrete Adjoint Formu-
lift design optimization using Navier–Stokes lation. PhD thesis, Virginia Polytechnic Insti-
equations. Journal of Aircraft, Vol. 33, No 3, tute and State University, 1998.
pp 499–504, 1996. [19] Nielsen EJ and Anderson WK. Recent improve-
[8] Giles MB and Pierce NA. An introduction to the ments in aerodynamic design optimization on
adjoint approach to design. Flow, Turbulence unstructured meshes. AIAA J., Vol. 40, No 6,
and Combustion, Vol. 65, No 3/4, pp 393–415, pp 1155–1163, 2002.
2000. [20] Nocedal J and Wright SJ. Numerical Optimiza-
[9] Godin P, Zingg DW, and Nelson TE. High-lift tion. Springer–Verlag Inc., New York, 1999.
aerodynamic computations with one- and two- [21] Pueyo A and Zingg DW. Efficient Newton–
equation turbulence models. AIAA J., Vol. 35, Krylov solver for aerodynamic computations.
No 2, pp 237–243, 1997. AIAA J., Vol. 36, No 11, pp 1991–1997, 1998.
[10] Gunzburger MD. Introduction into mathemati- [22] Pulliam TH. Efficient solution methods for the
cal aspects of flow control and optimization. In- Navier-Stokes equations. Technical report, Lec-
verse Design and Optimization Methods, Brus- ture Notes for the von Kármán Inst. for Fluid
sels, Belgium, April 1997. von Kármán Institute Dynamics Lecture Series, Brussels, Belgium,
For Fluid Dynamics. January 1986.
[11] Kim CS, Kim C, and Rho OH. Sensitivity anal- [23] Reuther JJ, Jameson A, Alonso JJ, Rimlinger
ysis for the Navier–Stokes equations with two- MJ, and Saunders D. Constrained multipoint
equation turbulence models. AIAA J., Vol. 39, aerodynamic shape optimization using an ad-
No 5, pp 838–845, 2001. joint formulation and parallel computers, part 2.
[12] Kim H-J, Sasaki D, Obayashi S, and Naka- Journal of Aircraft, Vol. 36, No 1, pp 61–74,
hashi K. Aerodynamic optimization of super- 1999.
sonic transport wing using unstructured adjoint [24] Rumsey CL and Ying SX. Prediction of high
method. AIAA Journal, Vol. 39, No 6, pp 1011– lift: Review of present CFD capability. Progress
1020, 2001. in Aerospace Sciences, Vol. 38, No 2, pp 145–
[13] Kim S, Alonso JJ, and Jameson A. Design opti- 180, 2002.
mization of high–lift configurations using a vis- [25] Saad Y and Schultz MH. GMRES: A gener-
cous continuous adjoint method. AIAA Paper alized minimal residual algorithm for solving
2002–0844, January 2002. nonsymmetric linear systems. SIAM J. Sci. &
[14] Nelson TE, Godin P, De Rango S, and Zingg Stat. Comput., Vol. 7, No 3, pp 856–869, 1986.
DW. Flow computations for a three-element air- [26] Spalart PR and Allmaras SR. A one-equation
foil system. Canadian Aeronautics and Space turbulence model for aerodynamic flows. AIAA
Journal, Vol. 45, No 2, pp 132–139, 1999. Paper 92–0439, January 1992.
[15] Nelson TE, Zingg DW, and Johnston GW. Com- [27] Van Dam CP. The aerodynamic design of
pressible Navier–Stokes computations of mul- multi-element high-lift systems for transport
tielement airfoil flows using multiblock grids. airplanes. Progress in Aerospace Sciences,
AIAA J., Vol. 32, No 3, pp 506–511, 1994. Vol. 38, No 2, pp 101–144, 2002.
[16] Nemec M and Zingg DW. Newton–Krylov algo- [28] Van Den Berg B. Boundary layer measurements
rithm for aerodynamic design using the Navier– on a two-dimensional wing with flap. NLR TR
Stokes equations. AIAA J., Vol. 40, No 6, 79009 U, January 1979.
pp 1146–1154, 2002.
[17] Nemec M and Zingg DW. A Newton–Krylov al-

173.11

You might also like