You are on page 1of 10

AIAA 98-0123

TWO-DIMENSIONAL AIRCRAFT HIGH LIFT SYSTEM DESIGN AND OPTIMIZATION


Eric Besnard,* Adeline Schmitz,* Erwan Boscher,+ Nicolas Garcia,+ and Tuncer Cebeci Aerospace Engineering Department California State University, Long Beach Abstract A two-dimensional aircraft high-lift system design and optimization method, which can be easily extended to three dimensions, is presented. The need for such a tool is assessed. The method uses a gradient based local optimizer. The aerodynamic performance is predicted using an Interactive Boundary Layer (IBL) approach. Methods to represent general multi-element airfoils by a set of design variables are described. The representation of airfoils by general shape functions as well as element positioning (deflection angle, gap, overlap) is considered. The accuracy of the IBL approach when Reynolds number, element gap, overlap and deflection are varied is investigated. The design/optimization approach is first validated for an inverse design by matching a pressure coefficient distribution. Next, the method is applied to multi-element airfoil lift to drag ratio and single airfoil maximum lift coefficient maximization. Results demonstrate the appropriateness of the approach for high lift system design and optimization. 1.0 Introduction The ultimate goal of an aircraft high lift system design team is to define the simplest configuration which, for prescribed constraints, will meet the take-off, climb, and landing requirements usually expressed in terms of maximum L/D and/or maximum CL.1 Most current methods for transport aircraft high lift system design rely on an extensive use of wind tunnel testing in conjunction with simple Computational Fluid Dynamics (CFD) analysis. The CFD tools usually consist of threedimensional inviscid methods to identify critical wing sections. Then, the high lift system design and optimization is performed with two*

dimensional data based on previous designs, simple viscous-inviscid methods, and wind tunnel testing for the critical sections of the wing. These approaches do not allow to vary the positions of the various elements, such as slats and flaps, in a systematic fashion. Also, structural deformations under the high wing loading and their effects on element positioning are difficult to take into account during the optimization process. Significant improvements can be reached by using advanced computational methods in conjunction with optimization methods to design better high lift systems at lower cost. This paper focuses on the aerodynamic optimization of two-dimensional configurations for high lift applications. The approach is general and can also be extended to three-dimensional configurations, including aeroelastic effects. A general optimization process is illustrated in Fig. 1. An initial set of s design variables, x = ( x i ) 1 i s , which might represent the configuration designed by experienced engineers, is supplied to the optimizer. Then, for this design, the objective function, f, is evaluated and the constraints, gi, are analyzed to check whether they are violated or not. If the optimum is not reached, these values are fed back to the optimizer that modifies the design vector x. The process is repeated until convergence. For the application to high lift aerodynamic optimization, the three main components of the numerical method are, (1) the representation of a configuration by a set of design variables, x, (2) the optimization method, and (3) the evaluation of the aerodynamic performance, i.e. f, for a given configuration. The constraints gi are analyzed at the stage appropriate for the problem considered.

Graduate Student, AIAA Student member Visiting Research Student, Ecole Polytechnique, France Professor and Chair, AIAA Fellow Copyright 1998 by the American Institute of Aeronautics and Astronautics Inc. All rights reserved.
+

1 American Institute of Aeronautics and Astronautics

AIAA 98-0123
Initial Design Variables

Define Configuration

Objective Function & Constraint Calculations

Optimizer

predictions is briefly described in Sect. 3 and is evaluated for high lift applications in Sect. 4. The design approach is validated in Sect. 5.1 for an inverse design problem. Finally, Sects. 5.2 and 5.3 present the applications of the high lift design and optimization method to the maximization of lift-todrag ratio of a three-element airfoil by varying gaps, overlaps, deflection angles, and the maximization of single airfoil maximum lift coefficient by varying the airfoil shape. The paper ends with a summary of the more important conclusions. 2.0 Design variables to represent airfoils The aircraft high lift system designer is usually given a wing designed for cruise conditions. The maximum chord of the slat and/or flap(s) is usually dictated by the size of the wing box determined for structural and fuel capacity considerations.1 Therefore, very little leverage exists on the shape of these elements but more on the spacing with respect to each other (gap and overlap). However, modifying the shape of an element may be useful and should not be ruled out in the optimization process. Defining general shapes is illustrated here by considering an entire single airfoil. Upper and lower surfaces of the airfoil are represented by y( x ) = y 0 ( x ) + x i f i ( x )
i =1 s

YES

Optimum ? NO New D. V.

Optimum Design
Fig. 1. Flowchart of the numerical optimization.

The optimization method must satisfy the objective of reaching the global optimum with a minimum amount of function calls, since these function calls tend to drive the cost of the optimization process. Advanced stochastic methods such as Simulated Annealing (SA),2 Genetic Algorithms (GA),3-4 and Monte Carlo (MC) and Quasi Monte Carlo (QMC) methods5 have been proposed for global optimization. For CFD applications, some of these methods compare favorably with gradient based approaches for low dimensionality problems.2 Most recent CFD optimization applications,6-7 however, use gradient based optimization methods because of the low number of function calls they require to obtain an improved design. In the present study, the method of Modified Feasible Directions (MFD) of DOT8 is used. One optimization iteration consists of first determining a Search Direction which defines how the design variables will be changed. The search direction depends on the gradients of the objective function and of the constraints, if any. In the present study, all gradients are calculated by finite difference. The second step, called One-Dimensional Search, is to determine how far to move in that direction. However, since DOT is a local optimizer, convergence to the global optimum is not guaranteed. The next section describes the representation of a configuration by a set of design variables where a distinction between shapes and element positioning is made. The IBL method used for aerodynamic performance

(1)

where x is the coordinate along the airfoil chord, y0 is a reference airfoil, e.g. a NACA 0012 airfoil, ( x i )1 i s are the design variables and ( f i )1 i s are the base functions. Several types of functions such as Hicks-Henne functions,9 Wagner functions, Legendre and Patched Polynomials,10 etc., can be used. Hicks-Henne functions are selected for the present application and are given by: f i ( x ) = x a (1 x )e bx ln ( 0.5 ) ln ( a ) f j ( x ) = sin x
b

(2)

where a and b control the center and thickness of the perturbation, and x is the normalized coordinate along the chord. They have the advantage of being space based functions, as opposed to frequency based functions (like Wagner functions), and thus allow for greater local control of the design. For the case of multi-element airfoils, element deflection angles, gaps and overlaps are used as design variables, as shown in Fig. 2.

2 American Institute of Aeronautics and Astronautics

AIAA 98-0123 4.0 Evaluation of the IBL method


r0 gap

: deflection
overlap (< 0)

Fig. 2. Definitions of gap, overlap, and deflection angle for a multi-element configuration.

It is clear, however, that the two could be combined to provide the most general and widest design space. 3.0 Interactive Boundary Layer (IBL) approach for performance predictions For a given configuration, the flowfield can be calculated by either solving the NavierStokes equations or employing an Interactive Boundary Layer (IBL) approach, which is based on the interactive solution of the inviscid and boundary layer equations. While the latter is not as general as the former, it offers a good compromise between the efficiency and the accuracy needed in a design environment and is therefore selected here. The IBL method has been used extensively for multi-element airfoil flowfield predictions and is described in greater detail in previous publications.11 Its ingredients are shown in Fig. 3. The inviscid flow field is computed by a panel method with compressibility corrections based on the Prandtl-Glauert formula. Once the external velocity distribution is known, the boundary layer equations are solved in an inverse mode using the Hilbert integral formulation to allow for the computation of separated flows.12 Transition is determined as part of the solution procedure, employing either the en-method or correlation formulas. The turbulent flow calculations employ a modified Cebeci-Smith eddy viscosity formulation validated for both accelerating and adverse pressure gradient flows.13 The displacement thickness and blowing velocity distributions are used to simulate the viscous effects in the inviscid method. The procedure is repeated until convergence.
Inviscid Method 2D Inverse B.L. Method Transition calculation

The calculation method of Sect. 2 was applied and validated for numerous single and multielement airfoils.11 It is evaluated here for the specific purpose of optimization. Maximum lift coefficients are calculated and compared with experimental data for single airfoils in Sect. 4.1. For multi-element airfoils, parametric studies are conducted and trend evaluations are conducted in Sect. 4.2. 4.1 Single Airfoils Fig. 4 shows a comparison between measurements,14 Navier-Stokes calculations,13 and results obtained with the present method for a NASA supercritical airfoil tested at a free stream Mach number M = 0.284 and chord Reynolds number RC = 4.02 106.
1.5

1.0
Cl

0.5

e xperiment Navier-Stokes present method

0.0 0.0

5.0

Fig. 4. Lift coefficient for a NASA supercritical airfoil.

10.0

15.0

Fig. 5 presents a comparison between measured15-16 and calculated maximum lift coefficients, cl,max, for a variety of single airfoils as a function of Reynolds number. Results show good agreement between calculations and experimental data, except for the FX 74-CL5-140 and FX 72-MS150B airfoils at RC = 106 where transition location may have been at the source of the discrepancy.
FX 72 MS 150A FX 72 MS 150B FX 74 CL5 140 NACA 0012 NACA23012

2.5 2.0 1.5 10 (a)


6

2.5 2.0 1.5 10 Rc


7

Cl,max
Blowing Velocity & Displacement thickness

10 (b)

10 Rc

Fig. 3. Interactive Boundary Layer approach.

Fig. 5. Cl,max for several airfoils at various Reynolds numbers, (a) measured and (b) calculated.

3 American Institute of Aeronautics and Astronautics

AIAA 98-0123 4.2.Multi-element airfoils For multi-element airfoils, the designer is often interested in positioning the elements to maximize or minimize a given objective function (e.g. maximize lift to drag ratio for a given lift coefficient). The ability of the calculation method to accurately predict changes in objective function value when gaps, overlaps and element deflections are varied is therefore critical. Also, being able to predict Reynolds number trends is important in order to correlate wind tunnel and flight test data. The present section evaluates the performance of the IBL approach for meeting these requirements. Similar studies were previously conducted using a compressible Navier-Stokes method.17 Baseline Wind tunnel measurements were performed on several three-element airfoil configurations at the NASA Langley Low Turbulence Pressure Tunnel (LTPT) at various Reynolds and Mach numbers.18-20 The baseline test case considered here has the slat and the flap deflected at 30o with gap/overlap riggings of 0.0295c/-0.025c and 0.0127c/0.0025c, respectively, where c is the airfoil chord with slat and flap retracted. The configuration is shown in Fig. 6. The test Reynolds number was 9 106 and the freestream Mach number was 0.2. Lift, drag and moment coefficients are presented in Fig. 7 and compared with experimental data. Several experimental data sets were available17 and are represented. It is important to note that the present results were obtained without using the experimental transition location, because, in general, the designer does not have that information. Instead, the transition predicted by the calculation method was used.
Cd

pressures near stall conditions and shows that agreement is also quite satisfactory.
5.0 4.0
MAIN ELT. TOT.

Cl

3.0 2.0 1.0


FLAP

measurements17 calculations SLAT

0.0
(a)

5.0

10.0

15.0

20.0

25.0

0.20 0.15 0.10 0.05 0.00


(b)

5.0

10.0

15.0

20.0

25.0

-0.30 -0.40 Cm -0.50 -0.60 -0.70 5.0 10.0 15.0 20.0 25.0

(c)

Fig. 7. Comparison of measured and calculated force coefficients for the baseline three-element airfoil, (a) lift, (b) drag, and (c) moment coefficients.
18.0

14.0

measurements calculations

10.0 -Cp

Fig. 6. Baseline three-element airfoil configuration.

6.0

The lift coefficients are slightly overpredicted, possibly due to modeling approximations used in the IBL method11 and/or due to experimental setup, such as wall effects. The calculated drag coefficient is larger than the measured one, probably for the same reasons. However, such large discrepancies between measured and calculated drag coefficients were not observed in test cases reported previously.11 Agreement with the measured moment coefficient is quite satisfactory. Fig. 8 shows a comparison between measured and calculated

2.0

-2.0 -0.2

0.0

0.2

0.4 x/c

0.6

0.8

1.0

1.2

Fig. 8. Comparison of measured and calculated pressure coefficient for the baseline configuration at = 21.34o.

All incremental values presented below are referenced to the present configuration and freestream flow conditions.

4 American Institute of Aeronautics and Astronautics

AIAA 98-0123 Effect of flap gap Data was also available for a configuration where the flap gap was increased to 0.015c. Fig. 9 shows the increments in total lift, flap lift, moment, and drag coefficients for various angles of attack. It is interesting to note that the differences between experimental data sets taken at the same wind tunnel conditions are often larger than the increments determined by the calculation method. Also, some measurements might be questionable such as the increase in drag coefficient measured around = 8o which reaches 30%, unless drastic changes, such as sudden increase in flow separation, may have resulted from the gap change during experiments.
0.10 0.05 Calculations Measurements
(a)

Effect of flap deflection The flap deflection was also increased to 35o, with a new flap rigging of 0.0095c/0.0c. Fig. 10 shows a comparison between measured and calculated total lift and flap lift coefficient increments. Results agree fairly well with experimental data, but it should be noted that the differences are much larger than in the previous section, thus suggesting that the method is at least able to predict trends reliably.
0.15 0.10 0.05

Cl

0.00 -0.05 -0.10 5 Measurements Calculations 10 15 20 25

Cl

0.00 -0.05

0.05 0.00

Cl
-0.10 5.0
(a)

-0.05 -0.10 5 10 15 20 25

10.0

15.0

20.0

25.0
(b)

0.05

Cl

0.00 -0.05 5.0

Fig. 10. Comparison of measured and calculated lift coefficient increments due to an increase in flap deflection and change of flap rigging; (a) total lift coefficient and (b) flap lift coefficient.

10.0

15.0

20.0

25.0

(b)

0.01 0.00 -0.01 -0.02

Effect of Reynolds number The baseline configuration was also tested at a chord Reynolds number of 5 106. Results are presented and compared with data in Fig. 11. Increments in total lift coefficient are well predicted. For drag and moment coefficients, even if variations with angle of attack are not replicated, the trends are captured. Both moment and drag coefficients increase. Conclusions Overall, comparisons between experimental data and calculations show that, for moderate variations of the parameters, the trends are satisfactorily captured even if their magnitude and variation with angle of attack might not be exactly matched. Also, for rather small variations in design parameters, calculated differences are of the same order of magnitude as the differences between two wind tunnel measurements, thus making it difficult to draw conclusions. Finally,

C m

10

15

20

25

(c)

0.015 0.010

Cd 0.005
0.000 -0.005
(d)

10

15

20

25

Fig. 9. Comparison of measured and calculated force coefficient increments due to an increase in flap gap; (a) total lift coefficient, (b) flap lift coefficient, (c) moment coefficient, and (d) drag coefficient.

5 American Institute of Aeronautics and Astronautics

AIAA 98-0123 differences in maximum lift coefficient are fairly well predicted (Figs. 9a, 10a, 11a). These observations suggest that, for the purpose of design/optimization where short turn around is critical, the use of an IBL approach compared to solving the Navier-Stokes equations can be quite sufficient, even without modeling merging shear layers and coves.
0.00 -0.05 -0.10

Cl

-0.15 -0.20 -0.25 5.0 Measurements Calculations 10.0 15.0 20.0 25.0

(a)

0.040 0.030

Cm 0.020
0.010 0.000
(b) 0.040 0.030

5.0

10.0

15.0

20.0

25.0

distribution of a S1223 airfoil15 at = 4o. The airfoil is represented by 22 design variables using the Hicks-Henne base functions defined in Sect. 2; the initial design corresponds to a NACA 0012 airfoil. The objective function to minimize corresponds to some integral of the pressure difference between the desired pressure distribution and the one calculated using the IBL approach of Sect. 3 for a given set of design variables. Fig. 12 shows the initial design point (NACA 0012 airfoil and its pressure distribution at = 4o), the objective point (S1223 airfoil and its pressure distribution for the same angle of attack), and the airfoil and pressure distribution obtained after convergence of the optimization process. The method converges in about 30 design iterations. Each iteration consists of calculating the gradient (i.e. search direction) which requires as many function evaluations as there are design variables (in this case 22) and then determining how far to move in that direction, usually requiring another 3 to 5 function evaluations. Both pressure distribution and airfoil ordinates match very well with the S1223 airfoil, thus showing that the method has been successfully implemented.
-3.0

Cd 0.020
0.010 0.000 5.0 10.0 15.0 20.0 25.0
-2.0

Initial (NACA 0012) S1223 Design

(c) Fig. 11. Comparison of measured and calculated force coefficient increments due to a reduction in Reynolds number; (a) total lift coefficient, (b) moment coefficient, and (c) drag coefficient.

Cp

-1.0

0.0

5.0 Applications to aircraft high lift system design/optimization The design/optimization approach is first validated for an inverse design by matching a pressure coefficient distribution. Then, the method is applied to multi-element airfoil lift-todrag ratio maximization by varying gaps, overlaps and deflection angles. Finally, the shape of a single airfoil is optimized to yield the highest maximum lift coefficient at a given Reynolds number. 5.1 Inverse design: pressure matching To validate the approach, the design optimization is applied to the matching of a prescribed pressure distribution. The pressure to be matched is the inviscid pressure

1.0 0.0

0.2

0.4

0.6

0.8

1.0

(a)
0.20 0.15 0.10

x/c

y/c

0.05 0.00 -0.05 -0.10 0.0

0.2

0.4

0.6

0.8

1.0

(b)

x/c

Fig. 12. Inverse design: matching of the S1223 airfoil inviscid pressure distribution at = 4o; (a) pressure distribution and (b) airfoil ordinates.

6 American Institute of Aeronautics and Astronautics

AIAA 98-0123 5.2 Lift-to-drag ratio optimization As previously mentioned, a designer might be interested in minimizing drag at a given lift coefficient, thus minimizing engine requirements or lowering noise limits at takeoff. This problem, equivalent to maximizing liftto-drag ratio (L/D) at fixed lift conditions, is discussed here. The baseline of Sect 4.2 is taken as the initial design. Angle of attack, slat and flap gaps, overlaps, and deflection angles are varied to maximize L/D at a fixed lift coefficient cl = 4.15, corresponding to = 16o for the baseline configuration. Fig. 13 shows the evolution of L/D with design iterations. The baseline has a L/D of 66.7 and the design configuration a L/D of 77.7, which yields a 16.5% improvement. In this case, each iteration requires approximately 12 function evaluations, including 7 for gradient calculations. Convergence is reached in 10 iterations. Each lift-to-drag ratio evaluation requires around 3-3.5 minutes on an IBM RISC 6000, which comes out to a total of about 6 hours for the optimization. However, it must be noted that the program has not been optimized to minimize CPU usage and the time required for a function evaluation could possibly be reduced to 1.5 minutes. This would reduce the optimization time by half.
80

drag coefficients and L/D ratio for each element. Also shown is the large improvement in L/D obtained for the main element, even if L/D was slightly reduced for the other elements.
Design variable slat deflection slat overlap slat gap flap deflection flap overlap flap gap Initial 30 -2.50 % 2.95 % 30 0.25 % 1.27 % 16 Design 30.017 -1.80 % 1.87 % 29.88 0.21% 1.45 % 15.38

Table 1. Initial and optimum design variables.


0.05

0.00 y/c -0.05

-0.10

Initial Design

(a)

-0.15 -0.10

-0.05

0.00 x/c

0.05

0.10

0.10 0.05 0.00


y/c

-0.05 -0.10

75

-0.15 -0.20 0.70

L/D
70

0.80

0.90

1.00

1.10

1.20

x/c (b) Fig. 14. Initial and design configurations, (a) slat and (b) flap positions.

65

10

Initial slat Lift Drag L/D main element flap Lift Drag L/D Lift Drag L/D TOTAL Lift Drag L/D 0.617 0.0101 61.11 3.139 0.0470 66.78 0.398 0.0050 79.19 4.153 0.0623 66.68

Design 0.642 0.0106 60.59 3.063 0.0367 83.45 0.416 0.0058 72.30 4.121 0.0531 77.68

iteration #

Fig. 13. Evolution of L/D with design iterations.

Table 1 gives the initial and the optimum designs. The parameter driving the optimization is the slat gap, which is reduced from 0.0295c to 0.0187c. During optimization, the constraint violation criterion was set to 1%, which explains why the final lift coefficient was slightly decreased from 4.15 to 4.12. A smaller margin could be used. Figs. 14 and 15 show the corresponding changes in configuration and pressure distributions, respectively. Fig. 15 shows that the load is transferred from the main element to the slat, and to a lesser extent, to the flap. Table 2 lists the changes in lift and

Table 2. Lift and drag changes between initial and final design for each element.

7 American Institute of Aeronautics and Astronautics

AIAA 98-0123
-12.0

Initial Design
-8.0

Cp
-4.0

0.0 0.0 0.2 0.4 0.6 0.8 1.0

(a)
-10.0 -8.0 -6.0

x/c

Fig. 16 shows the FX 74-CL5-140 airfoil and its representation with 22 design variables (denoted initial). A small discrepancy at the leading edge of the airfoil can be seen, related to the type of base functions used for the representation. The resulting initial airfoil has a maximum lift coefficient of 2.40 at =12, compared to 2.36 found in Sect. 4.1 for the FX 74-CL5-140 airfoil. The slope of the FX 74-CL5-140 airfoil is very large near the leading edge and cannot be represented by the Hicks-Henne functions, which have their slope related to the height of their maximum. One would need an extra design variable to control this slope.
0.03 0.02

Cp

-4.0 -2.0 0.0 2.0 0.0

FX 74 CL5 140 Initial (22 D.V .) Design

y/c

0.01 0.00

0.2

0.4

0.6

0.8

1.0

(b)
-4.0 -3.0 -2.0

x/c
0.00 0.01 0.02 0.03 0.04

x/c
Fig. 16. Comparison of leading edges shapes between the FX 74 CL5 140, its representation with 22 HicksHenne functions, and the result of the optimization.

Cp

-1.0 0.0 1.0 2.0 0.0

0.2

0.4

0.6

0.8

1.0

(c)

x/c

Fig. 15. Pressure distributions of initial and design configurations on the (a) slat, (b) main element and (c) flap.

5.3 Maximum Lift Optimization The approach is also applied to the determination of an airfoil shape that maximizes the maximum lift coefficient at a Reynolds number of 2 million. Similar to the pressure coefficient matching problem previously discussed, the airfoil is represented by a set of 22 design variables using the HicksHenne functions. The initial design corresponds to the airfoil that leads to the highest cl,max in Sect. 4.1, the FX 74-CL5-140 airfoil. Since the reference airfoil, y0(x), is a NACA 0012 airfoil, a nonzero initial set of design variables was used to represent the FX 74-CL5-140 airfoil. This requirement could be, however, easily relaxed.

8 American Institute of Aeronautics and Astronautics

Fig. 17 presents the evolution of the objective function, cl,max, with the number of design iterations. Fig. 18 and 19 show the initial and design airfoil shape and pressure coefficient distribution at the angle of attack corresponding to maximum lift, respectively. The shape of the leading edge of the design airfoil is shown in greater detail in Fig. 16. The design configuration leads to a 10% increase in maximum lift coefficient (design cl,max = 2.64), principally due to a stronger flow acceleration on the upper surface of the airfoil. However, a better cl,max could certainly be obtained for the reasons exposed below. First, the flow velocity on the lower surface, though small, could be reduced by thinning the airfoil. The optimizer attempted to do so, and since no constraint on the airfoil thickness was enforced, the upper and lower surfaces eventually crossed near the trailing edge, leading to an early termination of the optimization process. This situation can be circumvented by setting a minimum thickness constraint along the chord. Also, a rapid flow acceleration near the leading edge followed by a rapid recompression is observed in Fig. 19. This acceleration/deceleration phase is probably undesirable for improving cl,max.

AIAA 98-0123 Its elimination would require inclusion of base functions which allow for modification of the airfoil slope at the leading edge.
2.70

2.60

Cl,max
2.50

2.40

iteration #
Fig. 17. Evolution of the airfoil maximum lift coefficient during optimization. 0.3 0.2 0.1

drastically reduced if the program was optimized for minimum CPU usage. In conclusion, the method is able to significantly improve the design. Some improvements, however, still must be incorporated. Special attention must be paid to the representation of the airfoil at the leading edge; extra functions can be added to the base of Hicks-Henne functions to control the slope of the profile near the leading edge. Also, the optimization should constrain the airfoil thickness to some minimum value. These two issues are currently being addressed and the method should eventually lead to a higher maximum lift coefficient. 6.0 Concluding remarks The paper described the necessary ingredients for aerodynamic high lift optimization. These comprise appropriate representation of a configuration by a set of design variables, a gradient based optimization method, and an efficient Interactive Boundary Layer (IBL) approach for aerodynamic performance predictions. Both general shapes and element positioning are considered as design variables. The IBL approach is evaluated for the specific purpose of high lift optimization. Lift, maximum lift, drag, and moment coefficients are satisfactorily predicted. The optimization method is successfully validated by considering an inverse design problem. Then, the approach is applied to the optimization of a multi-element airfoil configuration for maximum L/D, and to the design of an airfoil shape for maximum cl,max at RC = 2 106. For both test cases, large improvements in objective function are obtained thus demonstrating the potential for using the approach in a design environment. Future work will include the introduction of global search methods to expand the design space and to allow for discrete design variable values. References 1. N.B. Nield, An Overview of the 777 High Lift Aerodynamic Design, Proceedings of High Lift and Separation Control Conference, Royal Aeronautical Society, March 1995. 2. S. Aly, M. Ogot, and R. Pelz, Stochastic Approach to Optimal Aerodynamic Shape Design, J. Aircraft, Vol. 33, No. 5, pp. 956961, Sept. 1996. 3. J. Lee and P. Hajela, Parallel Genetic Algorithm Implementation in Multidisciplinary Rotor Blade Design, J.

y/c
0.0 -0.1 -0.2 0.0

Initial Design
0.2 0.4 0.6 0.8 1.0

x/c
Fig. 18. Initial and design configurations.

-6.0

Initial Design
-4.0

Cp

-2.0

0.0

2.0 0.0

0.2

0.4

0.6

0.8

1.0

x/c
Fig. 19. Initial and design pressure distributions at the angle of attack corresponding to the maximum lift coefficient.

For this application, each maximum lift coefficient calculation requires about 8 minutes of CPU time on an IBM RISC 6000. With 22 design variables, about 16 hours of total CPU time is needed for this optimization. As previously mentioned, this time could be

9 American Institute of Aeronautics and Astronautics

AIAA 98-0123 Aircraft, Vol. 33, No. 5, pp. 962-969, Sept. 1996. B. Mantel, J. Priaux and B. Stoufflet, Review of the EUROPT project AERO0026, Optimum Design Methods for Aerodynamics, AGARD Report No. 803, 1994. H. Niederreiter, Random Number Generation and Quasi-Monte Carlo Methods, CBMS-NSF Regional Conference Series in Applied Mathematics, Society for Industrial and Applied Mathematics, 1992. J. Reuther and A. Jameson, A Comparison of Design Variables for Control Theory Based Airfoil Optimization, Proceedings of the 6th International CFD Symposium, Lake Tahoe, NV, Sept. 1996. S. Eyi, K.D. Lee, S.E. Rogers, and D. Kwak, High-Lift Design Optimization Using Navier-Stokes Equations, J. of Aircraft, Vol. 33, No. 3, May 1996. Vanderplaats, Miura & Associates, Inc., "DOT Users Manual, Version 4.10," VMA Engineering, 1994. R. Hicks and P. Henne, Wing Design by Numerical Optimization, J. Aircraft, Vol. 15, No. 7, pp. 407-413, July 1978. Eyi S., Performance Evaluation and Improvements of CFD-based Aerodynamic Design Optimization, Ph.D. Dissertation, University of Illinois, Urbana Champaign, 1995. T. Cebeci, E. Besnard and H.H. Chen, Calculation of Multi-element Airfoil Flows, Including Flap Wells, AIAA Paper No. 960056, Jan. 1996. T. Cebeci, An Engineering Approach to the Calculation of Aerodynamic Flows, to be published, 1997. T. Cebeci and K.C. Chang, An Improved Cebeci-Smith Turbulence Model for Boundary-Layer and Navier-Stokes Methods, Proceedings of the 20th ICAS/AIAA Aircraft System Conference, Sept. 1996. E. Omar, T. Zierten and A. Mahal, TwoDimensional Wind Tunnel Tests of a NASA Supercritical Airfoil with Various High Lift Systems, NASA CR-2215, 1977. M.S. Selig and J.J. Guglielmo, High-Lift Low Reynolds Number Airfoil Design, J. Aircraft, Vol. 34, No. 1, Jan. 1997. I.H. Abbott and A.E. Von Doenhoff, Theory of Wing Sections, Dover, 1959. F.T.Lynch, R.C. Potter, and F.W. Spaid, "Requirements for Effective High Lift CFD," Proceedings of the 20th ICAS/AIAA Aircraft System Conference, Sept. 1996. 18. W.O. Valarezo, "High Lift Testing at High Reynolds Numbers," AIAA Paper 92-3986, July 1992. 19. V.D. Chin, D.W. Peter, F.W. Spaid, and R.J. McGhee, "Flowfield Measurements About a Multi-element Airfoil at High Reynolds Numbers," AIAA Paper 93-3137, July 1993. 20. F.W. Spaid and F.T. Lynch, "High Reynolds Number, Multi-element Airfoil Flowfield Measurements," AIAA Paper 96-0682, Jan. 1996

4.

5.

6.

7.

8. 9. 10.

11.

12. 13.

14.

15. 16. 17.

10 American Institute of Aeronautics and Astronautics

You might also like