You are on page 1of 154

AIRFOIL OPTIMIZATION FOR MORPHING AIRCRAFT

A Thesis Submitted to the Faculty of Purdue University by Howoong Namgoong

In Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy

December 2005

ii

I dedicate this thesis to my father, Young Kyu Namgoong in heaven.

iii

ACKNOWLEDGMENTS

Thanks to God for being my guidance of the journey of life. It has been a privilege to be a student of Drs. William A. Crossley and Anastasios S. Lyrintzis. I was able to open my eyes toward the world of design optimization and morphing aircraft with a tremendous help from Dr. Crossley. I learned great knowledge about aerodynamics and received precious advice from Dr. Lyrintzis. I will cherish and miss the moments that we met together for five years. Special thanks to my committee members, Dr. Scott D. King, Dr. Marc H. Williams and Dr. Terrence A. Weisshaar for their invaluable comments and lectures. I also thank to my colleagues and staffs in Purdue AAE department. This work was partially supported by the Air Force Research Laboratory, contract F33615-00-C-3051, and by a Purdue Research Foundation grant. I would like to share this great moment with my lovely wife, Miran who completes my life, and my beautiful son, Young who gives me another reason for living. I will not forget the support from my three sisters, Ran, Eun and Yoon and my brothers in law. I also like to thank my father and mother in law for their support and prayer. Lastly, my deep appreciation goes to my mother, Mal Soon Park who showed me the meaning of true love.

iv

TABLE OF CONTENTS

Page LIST OF TABLES........................................................................................................... viii LIST OF FIGURES ........................................................................................................... ix ABSTRACT...................................................................................................................... xv CHAPTER 1 INTRODUCTION ....................................................................................... 1 1.1 Background .............................................................................................................. 1 1.1.1 Airfoil Design Methods .................................................................................... 1 1.2 Issues in Airfoil Design Optimization ..................................................................... 4 1.2.1 Objective and Constraint Functions.................................................................. 4 1.2.2 Design Variables............................................................................................... 5 1.2.3 Global Optimization.......................................................................................... 9 1.2.4 Flow Models ..................................................................................................... 9 1.3 Morphing Aircraft.................................................................................................. 11 1.3.1 Return to the First Flight................................................................................. 11 1.3.2 Definition of Morphing Aircraft ..................................................................... 12 1.3.3 Morphing Aircraft Related Programs ............................................................. 12 1.4 Motivation of Research.......................................................................................... 17 1.4.1 Global Optimization Issues in Airfoil Optimization....................................... 17 1.4.2 Morphing Airfoil Design ................................................................................ 18

v Page 1.5 Thesis Objectives ................................................................................................... 19 1.6 Thesis Organization ............................................................................................... 19 CHAPTER 2 TRANSONIC AIRFOIL OPTIMIZATION .............................................. 21 2.1 Design Variables.................................................................................................... 22 2.2 Base Airfoils .......................................................................................................... 28 2.3 Trimming ............................................................................................................... 31 2.4 Objective Function................................................................................................. 32 2.5 Accuracy of Function Evaluations......................................................................... 33 2.6 Computational Cost ............................................................................................... 38 2.6.1 Parallel Genetic Algorithm ............................................................................. 38 2.6.2 Gradient Based Optimization Method ............................................................ 40 2.7 Objectives and Fitness Function Formulation ....................................................... 40 2.7.1 Objectives ....................................................................................................... 40 2.7.2 Fitness Function Formulation ......................................................................... 41 2.8 Single-point Optimization Results Comparison .................................................... 43 2.8.1 GA Results ...................................................................................................... 43 2.8.2 GM Results ..................................................................................................... 53 2.9 Multi-point Optimization Results Comparison...................................................... 57 2.10 Lessons Learned from GA and GM Results ........................................................ 62 2.10.1 Multimodal Design Space............................................................................. 62 2.10.2 Computational Efficiency ............................................................................. 63 2.11 Summary .............................................................................................................. 64 CHAPTER 3 AIRFOIL OPTIMIZATION FOR MORPHING AIRCRAFT .................. 65 3.1 Problem Description .............................................................................................. 65

vi Page 3.2 Objective Function Formulation ............................................................................ 67 3.2.1 Aerodynamics Only Investigation (Single-objective Approach).................... 67 3.2.2 Energy Based Optimization (Multi-objective Approach)............................... 68 3.2.3 Multi-objective Optimization.......................................................................... 69 3.3 Design Variables.................................................................................................... 72 3.3.1 Optimization Algorithm.................................................................................. 73 3.3.2 Flow Solver..................................................................................................... 73 3.4 Optimization Results.............................................................................................. 74 3.4.1 Aerodynamic-only Optimization Results ....................................................... 74 3.4.2 Energy Based Optimization Results ............................................................... 76 3.4.3 Cp and Energy Comparison............................................................................. 82 3.5 Transonic Morphing Airfoil................................................................................... 88 3.5.1 Problem Definition.......................................................................................... 88 3.5.2 Objective Function.......................................................................................... 88 3.5.3 Flow Solver, Design method and Parameters ................................................. 89 3.5.4 Aerodynamics-only Result.............................................................................. 89 3.5.5 Energy-based Design Results ......................................................................... 91 3.6 Summary ................................................................................................................ 99 CHAPTER 4 ACTUATION ENERGY MODELING INCLUDING AERODYNAMIC WORK .................................................................................................... 100 4.1 Description of Concept ........................................................................................ 100 4.2 Formulation.......................................................................................................... 103 4.3 Sensorcraft Problem............................................................................................. 106 4.3.1 Problem Definition........................................................................................ 106 4.3.2 Stiffness Approximation ............................................................................... 107 4.4 Results and Comparison ...................................................................................... 109

vii Page 4.4.1 Effect of Aerodynamic Work Term.............................................................. 109 4.4.2 Effect of Stiffness Change ............................................................................ 122 4.5 Summary .............................................................................................................. 130 CHAPTER 5 CONCLUSIONS ..................................................................................... 131 5.1 Future Directions ................................................................................................. 132 5.2 Contributions........................................................................................................ 133 LIST OF REFERENCES................................................................................................ 134 VITA ............................................................................................................................... 139

viii

LIST OF TABLES

Table

Page

Table 2.1 Error norm comparison (Target airfoil: Whitcomb supercritical airfoil) ....... 24 Table 2.2 N-S solution comparison with experimental data........................................... 34 Table 2.3 Problem formulations ..................................................................................... 40 Table 2.4 Optimization methods..................................................................................... 42 Table 2.5 Single-point design result (GA)...................................................................... 44 Table 2.6 Single-point design result (GM) ..................................................................... 54 Table 2.7 Multi-point design result................................................................................. 57 Table 2.8 Comparisons of drag values and computational costs for GA and GM, (singlepoint runs) ....................................................................................................... 63 Table 3.1 Airfoil design conditions ................................................................................ 66 Table 3.2 Drag comparison of aerodynamics only design.............................................. 74 Table 3.3 Drag and Relative Strain Energy comparison (weighted sum method).......... 76 Table 3.4 Drag and Relative Strain Energy comparison (-constraint method) ............. 78 Table 3.5 Transonic Morphing UAV Mission Profile .................................................... 88 Table 3.6 -constraint target values for Cd constraints ................................................... 97 Table 4.1 Actuation energy formulation....................................................................... 110

ix

LIST OF FIGURES

Figure

Page

Figure 1.1 Discrete approach using airfoil coordinates ................................................... 6 Figure 1.2 Bezier approach using control points ............................................................. 7 Figure 1.3 Figure 1.4 Figure 1.5 The hierarchy of mathematical models ........................................................ 10 Front-view of Wright Flyer, illustrating wing warping (From Combs, Kill Devil Hill, Houghton Mifflin, Boston,1979.) .............................................. 11 MAW modifications to F-111 (From NASA TM-4606).............................. 13 Figure 1.6 Flight-determined drag polar comparison (From NASA TM-4606)............ 13 Figure 1.7 Smart Technologies ...................................................................................... 14 Figure 1.8 Morphing airplane (NASA).......................................................................... 15 Figure 1.9 Sliding skins concept (Image: NextGen)...................................................... 16 Figure 1.10 Folding wing concept (Image: Lockheed Martin)........................................ 16 Figure 2.1 Shape functions ............................................................................................ 25 Figure 2.2 Original versus reconstructed airfoils........................................................... 26 Figure 2.3 NACA0012 with each shape function applied ............................................. 27 Figure 2.4 Configuration of base airfoils....................................................................... 28 Figure 2.5 Design space available using the NACA0012 base airfoil........................... 29 Figure 2.6 Design space available using the RAE2822 base airfoil .............................. 30 Figure 2.7 Design space available using the Whitcomb base airfoil ............................. 30 Figure 2.8 Schematics of trimming process................................................................... 32 Figure 2.9 Grid using 79 by 29 points (RAE2822 airfoil)............................................. 35 Figure 2.10 Grid using 129 by 30 points (RAE2822 airfoil)........................................... 35 Figure 2.11 Grid using 217 by 61 points (RAE2822 airfoil)........................................... 36

x Figure Page

Figure 2.12 Pressure coefficient distribution of N-S prediction vs. published experiment (RAE 2822 airfoil, M=0.74, =3.19) ....................................................... 36 Figure 2.13 Lift coefficient convergence history............................................................. 37 Figure 2.14 Drag coefficient convergence history........................................................... 37 Figure 2.15 Speed-up of parallel GA, computational wall time vs. number of CPUs..... 39 Figure 2.16 Best fitness value convergence history for all three GA runs. ..................... 44 Figure 2.17 Best airfoil shapes (top) and pressure coefficient distributions (bottom) in selected generations of the GA using the NACA 0012 base airfoil. ........... 45 Figure 2.18 Best airfoil shapes (top) and pressure coefficient distributions (bottom) in selected generations of the GA using the RAE 2822 base airfoil................ 46 Figure 2.19 Best airfoil shapes (top) and pressure coefficient distributions (bottom) in selected generations of the GA using the Whitcomb supercritical base airfoil............................................................................................................ 47 Figure 2.20 Mach contour for the NACA0012 airfoil (M=0.74, Cl=0.605, = 12.64D )

(top) and Mach contour for the best airfoil shape from 80th generations of GA runs using the NACA0012 base airfoil (M=0.74, Cl=0.733)(bottom) .. 48 Figure 2.21 Mach contour for the RAE2822 airfoil (M=0.74, Cl=0.733)(top) and Mach contour for the best airfoil shape from 80th generations of GA runs using the RAE2822 base airfoil (M=0.74, Cl=0.733)(bottom).................................... 49 Figure 2.22 Mach contour for the Whitcomb airfoil (M=0.74, Cl=0.733)(top) and Mach contour for the best airfoil shape from 80th generations of GA runs using the Whitcomb base airfoil (M=0.74, Cl=0.733)(bottom)................................... 50 Figure 2.23 Best airfoil shapes from 80th generations of all three GA runs. ................... 52 Figure 2.24 Pressure coefficient distributions for best airfoils from 80th generations of GA runs........................................................................................................ 52 Figure 2.25 Convergence history for CONMIN with unmodified base airfoils as the starting shapes.............................................................................................. 54 Figure 2.26 Airfoil shape designs (top) and pressure coefficient distributions (bottom) generated during CONMIN search using RAE2822 base airfoil................. 55

xi Figure Page

Figure 2.27 Airfoil shape designs (top) and pressure coefficient distributions (bottom) generated during CONMIN search using Whitcomb base airfoil................ 56 Figure 2.28 Airfoil shapes for base, single-point, two-point design [GA (above) and CONMIN (below) results] ........................................................................... 58 Figure 2.29 Pressure coefficient distributions for two-point objective function results at
M=0.68 [GA (top) and CONMIN (bottom) results] .................................... 59

Figure 2.30 Pressure coefficient distributions for two-point objective function results at
M=0.74 [GA (top) and CONMIN (bottom) results] .................................... 60

Figure 2.31 Drag divergence diagram ............................................................................. 61 Figure 3.1 Notional high-altitude, long endurance aircraft concept .............................. 66 Figure 3.2 Internal linear spring model for strain energy .............................................. 69 Figure 3.3 Best airfoil shape (single-point optimization) .............................................. 75 Figure 3.4 Best airfoil shape (multi-point optimization) ............................................... 75 Figure 3.5 Weighted sum approach Pareto front ........................................................... 77 Figure 3.6

-Constraint approach Pareto front ............................................................. 78

Figure 3.7 Airfoil set found by multi-objective optimization including energy............ 79 Figure 3.8 Pareto set from N-Branch Tournament GA.................................................. 80 Figure 3.9 Airfoil shapes from energy based design (N-Branch tournament GA)........ 81 Figure 3.10 Convergence history of Pareto set from N-Branch Tournament GA. .......... 81 Figure 3.11 Cp distribution comparison (top), Airfoil comparison (bottom)................... 83 Figure 3.12 Cp distribution comparison (top), Airfoil comparison (bottom)................... 84 Figure 3.13 Cp distribution comparison (top), Airfoil comparison (bottom) .................. 85 Figure 3.14 Strain energy distributions (U12) .................................................................. 86 Figure 3.15 Strain energy distributions (U23) .................................................................. 87 Figure 3.16 Strain energy distributions (U13) .................................................................. 87 Figure 3.17 Best airfoil shape (Single-point optimization) ............................................. 90 Figure 3.18 Best airfoil shape (Multi-point optimization)............................................... 90 Figure 3.19 Pareto-front of transonic morphing wing case ............................................. 91 Figure 3.20 Pareto-front of transonic morphing wing case (Rescaled picture) ............... 92

xii Figure Page

Figure 3.21 Airfoil shapes from energy based design (N-Branch tournament GA) ........ 93 Figure 3.22 Airfoil shapes of Pareto front (Design condition 1) ..................................... 93 Figure 3.23 Airfoil shapes of Pareto front (Design condition 2) ..................................... 94 Figure 3.24 Cp distribution comparison (top), Airfoil shape comparison (bottom)........ 95 Figure 3.25 Cp distribution comparison (top), Airfoil shape comparison (bottom)........ 96 Figure 3.26 GA convergence history of -constraint method......................................... 98 Figure 3.27 Drag coefficient values................................................................................ 98 Figure 4.1 Figure 4.2 Figure 4.3 Figure 4.4 Figure 4.5 Aerodynamic force distributions on the airfoil surface............................. 101 The illustration of the aerodynamic work ................................................. 102 Simple spring airfoil structure model ........................................................ 104 Linear aerodynamic force variation........................................................... 105 Panel distribution on the airfoil ................................................................. 106 at state-1(Cl=0.85) and state-2 (Cl=1.52)................................................... 107 Figure 4.7 Aerodynamic force distribution at each control point................................ 108 Figure 4.8 Pareto front comparison (GA Generation 1000) ........................................ 112 Figure 4.9 Airfoil shapes [Aerodynamic load included case, drag objective (0.85)] .. 113 Figure 4.10 Airfoil shapes [Strain energy only case, drag objective (0.85)] ................. 113 Figure 4.11 Airfoil shapes [Aerodynamic load included case, drag objective (0.78)] .. 114 Figure 4.12 Airfoil shapes [Strain energy only case, drag objective (0.78)] ................. 114 Figure 4.13 Airfoil shapes [Aerodynamic load included case, drag objective (0.75)] .. 115 Figure 4.14 Airfoil shapes [Strain energy only case, drag objective (0.75)] ................. 115 Figure 4.15 Strain Energy Distribution [Drag objective (0.85)].................................... 116 Figure 4.16 Aerodynamic force distribution on the airfoil surface at both design conditions [Aerodynamic load included case, drag objective (0.85)] ....... 117 Figure 4.17 Magnified picture of the aerodynamic force distribution (Center area) [Aerodynamic load included case, drag objective (0.85)] ......................... 118 Figure 4.18 Magnified picture of the aerodynamic force distribution (Trailing edge area) [Aerodynamic load included case, drag objective (0.85)] ......................... 118

Figure 4.6 Normal component of aerodynamic force acting on the surface of the airfoil

xiii Figure Page

Figure 4.19 Cp distribution comparison [Drag objective (0.85)] .................................. 119 Figure 4.20 Cp distribution comparison [Drag objective (0.85)] .................................. 120 Figure 4.21 Cp distribution comparison [Drag objective (0.85)] .................................. 121 Figure 4.22 Pareto front for different stiffness .............................................................. 122 Figure 4.23 Airfoil shape comparison [Drag objective (0.85), Design Cl (0.85)] ......... 123 Figure 4.24 Airfoil shape comparison [Drag objective (0.85), Design Cl (0.18)] ......... 124 Figure 4.25 Airfoil shape comparison [Drag objective (0.85), Design Cl (1.52)] ......... 124 Figure 4.26 Airfoil shape comparison [Drag objective (0.75), Design Cl (0.85)] ......... 125 Figure 4.27 Airfoil shape comparison [Drag objective (0.75), Design Cl (1.18)] ......... 125 Figure 4.28 Airfoil shape comparison [Drag objective (0.75), Design Cl (1.52)] ......... 126 Figure 4.29 Airfoil shape comparison [Drag objective (0.85), Design Cl (1.52)] ......... 127 Figure 4.30 Airfoil shape comparison [Drag objective (0.85), Design Cl (1.52)] ......... 128 Figure 4.31 Airfoil shape comparison [Drag objective (0.87), Design Cl (0.85)] ......... 129 Figure 4.32 Airfoil shape comparison [Drag objective (0.87), Design Cl (1.52)] ......... 129

xiv

NOMENCLATURE

Cd
C di
Cl

Coefficient of drag Coefficient of drag at design condition i Coefficient of lift Coefficient of lift at design condition i Coefficient of pitching moment Coefficient of skin friction Airfoil chord length Free stream Mach number Airfoil coordinate Relative strain energy of airfoil 1 and airfoil 2 Relative strain energy of airfoil 2 and airfoil 3 Relative strain energy of airfoil 1 and airfoil 3 Weighting factor for multipoint design Design variables Error tolerance Reference drag coefficient

Cli

Cm

Cf

c
M

x
U 12

U 23 U 13

xv

ABSTRACT

Namgoong, Howoong, Ph.D., Purdue University, December, 2005. Airfoil Optimization for Morphing Aircraft. Major Professors: William A. Crossley and Anastasios S. Lyrintzis. Continuous variation of the aircraft wing shape to improve aerodynamic performance over a wide range of flight conditions is one of the objectives of morphing aircraft design efforts. This is being pursued because of the development of new materials and actuation systems that might allow this shape change. The main purpose of this research is to establish appropriate problem formulations and optimization strategies to design an airfoil for morphing aircraft that include the energy required for shape change. A morphing aircraft can deform its wing shape, so the aircraft wing has different optimum shapes as the flight condition changes. The actuation energy needed for moving the airfoil surface is modeled and used as another design objective. Several multi-objective approaches are applied to a low-speed, incompressible flow problem and to a problem involving low-speed and transonic flow. The resulting solutions provide the best tradeoff between low drag, high energy and higher drag, low energy sets of airfoil shapes. From this range of solutions, design decisions can be made about how much energy is needed to achieve a desired aerodynamic performance. Additionally, an approach to model aerodynamic work, which would be more realistic and may allow using pressure on the airfoil to assist a morphing shape change, was formulated and used as part of the energy objective. These results suggest that it may be possible to design a morphing airfoil that exploits the airflow to reduce actuator energy.

CHAPTER 1 INTRODUCTION

An airfoil is the two dimensional section shape of the wing of an aircraft. The shape of the airfoil should be designed in the preliminary phase of aircraft development based on the aircraft performance requirements. Airfoil design is very important for aircraft, because the airfoil can determine, to a large extent, the aircrafts performance. This is especially true in the transonic regime, where aircraft speed is limited by drag divergence caused by shock wave induced separation. With the assistance of optimization algorithms and computational fluid dynamics (CFD), a new innovative airfoil shape that has minimum drag could be designed with less development cost and fewer man-hours. Firstly, this chapter reviews briefly the different airfoil design methods and discusses the optimization method issues in designing airfoil. Secondly, morphing aircraft is introduced to provide a framework for the research motivations.
1.1 Background

1.1.1 Airfoil Design Methods Methods of aerodynamic shape design have been developed over the past century of powered-aircraft history and can be categorized by four approaches 1. 1) Cut-and-try approach 2) Indirect method 3) Inverse design techniques 3) Optimization techniques

2 In the cut-and-try approach, the design engineer specifies geometry based on the flow field analysis result from wind-tunnel or flight tests. The Clark Y airfoil 2 (1920s) is a good example of this approach to airfoil design. The National Advisory Committee for Aeronautics (NACA)
3

systematized this approach by perturbing successful airfoil

geometries to generate a series of related airfoils. The NACA four-digit airfoils (1930s) are one of the most famous NACA series airfoils. Jacobs2, who was the leading experimentalist in NACA, designed completely new airfoil shapes with large regions of favorable pressure gradient, the laminar-flow airfoils. The most successful of the NACA laminar-flow airfoils was the six series. NACA six-series laminar-flow airfoils were used on almost all high-speed airplanes in the 1940s and 1950s and are still in use today. The invention of the jet engine led the high speed flight era, and the drag divergence of aircraft at high speed was a big problem in aircraft design. In 1960s, Whitcomb 4 of NASA Langley proposed an airfoil shape with supersonic flow over a major portion of the upper surface and subsonic drag rise well beyond the critical Mach number. This supercritical airfoil was designed on the basis of intuitive reasoning and substantiating experimentation. The cut-and-try design approach is time consuming and very expensive, and it also depends on the engineers expertise. Indirect methods are characterized by the fact the designer does not have control over the geometry or over aerodynamic quantities such as lift, pitching moment and pressure distribution directly. Rather than specifying such quantities, the designer manipulates a number of non-physical parameters. The hodograph and fictitious gas methods are in this category. One of the earliest indirect design procedures is the hodograph method by Bauer, et al 5. The hodograph method uses a variable transformation to the hodograph plane, which linearizes the partial differential equations for compressible potential flow. Solutions are constructed by superposition of fundamental solutions of the hodograph equations. Theoretically, it is possible to model shock waves, but it turns out to be difficult in practice. Hence, the hodograph method is limited to generating shock-free flows. Another limitation of the hodograph method is that the concept cannot be extended to three dimensions. This limitation has been overcome in the Sobieczkys fictitious gas concept 6 . When the solution to the fictitious gas flow problem is known, the correct

3 supersonic flow field inside the sonic surfaces is determined by solving an initial value problem with the initial data given on the sonic surface. The new correct flow inside the sonic umbrella defines a new stream surface that is tangent to the contour at the intersection of the sonic surface and original body. In this way, part of the body is modified, and it is considered as a shock-free redesign method. The drawback of this method is that the redesign problem does not always have a useful solution. This is associated with the character of the initial value problem to be solved in the supersonic part of the flow field. After the hodograph method was introduced, Barger and Brooks 7 published the inverse method. The inverse method alters the airfoil iteratively to produce a pressure distribution that matches a specified target distribution. The target pressures are used to determine velocities along the airfoil surface. These velocities are the boundary conditions for the flow field solution. After each iteration, the surface on which these boundary conditions are applied is changed in an attempt to reduce the normal component of velocity. When the normal component of velocity is zero, the surface is the proper airfoil shape. One of the first transonic inverse methods was developed by Tranen 8 (1974). Volpe and Menikc 9 (1985) formulated the first correctly posed inverse method for two-dimensional flow. Because the inverse shape design is based on the specified pressure distribution, it contains the dilemma of choosing the best surface target pressure for which a solution may not necessarily exist. Another drawback of the inverse method for aerodynamic shape design is that it is capable of creating only single pointdesigns. Therefore, the resulting shapes will have the desired aerodynamic characteristics only at the design conditions. It is also difficult to treat geometric constraints. In 1974, Hicks, et al. 10, developed one of the first optimization methods for airfoil design using Vanderplaatss optimization module CONMIN 11 , which is based on the feasible directions and the conjugate gradient method. Kennery 12 developed an airfoil design code in 1983, using a quasi-Newton optimization method, which is known to be more efficient than the conjugate gradient method.

4 As another optimization approach, control theory has been applied to systems of partial differential equations governing fluid flow 13 , 14 . This is essentially a gradient search optimization approach where the gradient information is obtained by formulating and solving a set of adjoint partial differential equations rather than evaluating the derivatives using finite differencing. The adjoint method has some difficulties, even though it can acquire a solution faster than a typical gradient-based optimization algorithm for which numerical derivatives are computed. One drawback is that it does not always allow for flow separation. The method also suffers from a tendency to converge to any of the numerous local minima, like most of the gradient search optimizers. An additional drawback is that it requires the derivation of an entirely new system of partial differential equations in terms of some non-physical adjoint variables and specification of their boundary conditions. Recently, stochastic optimization methods (e.g. genetic algorithms 15 and simulated annealing 16 ) are used for aerodynamic optimization. Since the 1980s, the Genetic Algorithm (GA) has been applied to many engineering optimization problems 17,18. Based on Darwins survival of the fittest concept, the GA performs optimization tasks by evolving a population of highly fit designs over many generations. The airfoil optimization method has also many issues, and more details are described in next section.
1.2 Issues in Airfoil Design Optimization

Optimization algorithms have changed the traditional aircraft design process; and, currently, many optimization techniques are applied during the airfoil design process. The major issues of airfoil design optimization are summarized in this section. 1.2.1 Objective and Constraint Functions A general form of the numerical optimization problem can be represented as Equation (1.1). The selection of the appropriate objective function, design variables and constraint functions are crucial to the success of the optimization methods. Conventionally, the airfoil optimization problem is a drag minimization problem with constraints of lift, moment, and/or thickness of the airfoil.

5 Minimize : Subject to :
F ( i )

(1.1) j=1,J i=1,K

g j ( i ) 0

iL i iU
where,

i: Design Variables

Airfoils designed for a single-point objective condition using the statement in Equation (1.1) may have poor performance in flow conditions other than the design condition. Some of the off-design performance problems of single-point optimization are presented in Ref.19. To overcome the problem of single-point designs, a multi-point optimization is suggested. Multiple objectives are combined into one objective using weighting factors as shown in Equation (1.2). Minimize :
Subject to : where,

m =1

Fm ( i )

(1.2)

g j ( i ) 0 j=1,J

m =weight factors

m =1

= 1.0 i=1,K

iL i iU

Constraints should be applied to the optimization problem to acquire a reasonable design result. In airfoil design optimization, lift is often constrained to be equal to a prescribed value. Thickness of the airfoil also can be another constraint. A thin airfoil is superior for aerodynamic performance. However, the thickness of the airfoil is confined by structural stiffness and space for fuel. Pitching moment of the airfoil is also candidate for constraints. An airfoil with a low pitching moment results in a wing that reduces the aircrafts trim drag. 1.2.2 Design Variables Design variables that affect and control the value of the objective function are needed in an optimization problem. A method to express the airfoil geometry accurately with few

6 variables is required in airfoil optimization. An airfoil is a smooth curve and there are many different approaches to approximate the curve 20.
Discrete approach

The discrete approach uses the coordinates of the geometry illustrated in Figure 1.1 as design variables. This approach is easy to implement, but the geometry changes are limited by the number of design variables. It is difficult to maintain a smooth geometry with a small number of design variables. Thus, the optimization solution may be impractical to achieve.

Figure 1.1 Discrete approach using airfoil coordinates


Polynomial and spline approach

Using polynomial and spline representations for shape parameterization can reduce the total number of design variables. A polynomial can describe a curve in a very compact form with a small set of design variables. As an example, a curve can be described as the polynomial shown in Equation (1.3) F (u ) = ci u i
i =0 n 1

(1.3)

Where n is the number of design variables, and u is the parameter coordinate along the curve. The ci is a set of coefficient vectors corresponding to coordinates and the components of these vectors can be used as design variables. The Bezier representation is another mathematical form for representing curves. F (u ) = pi Bi , p (u )
i =0 n 1

(1.4)

Where n is the number of control points (design variables), and the Bi , p (u ) are degree p Bernstein 21 polynomials. The pi are control points and they are typically used as design variables as shown in Figure 1.2.

Figure 1.2 Bezier approach using control points

The Bezier form is an effective and accurate representation for shape optimization of simple curves. However, complex curves require a high-degree Bezier form. As the degree of a Bezier curve increases, so does the round-off error. Also, it is very inefficient to compute a high-degree Bezier curve. A B-spline uses several low-degree Bezier segments to cover the entire curve resulting in one composite curve. A multisegmented B-spline curve can be described by F (u ) = pi N i , p (u )
i =0 n 1

(1.5)

Where pi are the B-spline control points, p is the degree, and N i , p (u ) is the i-th B-spline basis function of degree p. The drawback of the regular B-spline representation is its inability to represent implicit conic sections accurately. However, a special form of Bspline, nonuniform rational B-spline (NURBS) can represent quadric primitives as well as free-form geometry. A NURBS curve is defined as

R (u ) =

N
i =1 n i =1

i, p

(u )Wi pi (1.6) (u )Wi

i, p

Where the pi are the control points, Wi are the weights, and the N i , p are degree p B-spline basis functions. Despite recent progress, it is still difficult to parameterize complex shapes using polynomial and spline representations. This approach requires a large number of control points, and optimization is prone to creating irregular or wavy geometry with the spline representation.

Analytical approach (shape functions)

8 Hicks and Henne 22 introduced a compact formulation for parameterization of airfoils. The formulation is based on linearly adding shape functions to a baseline shape. The contribution of each parameter is determined by the value of the participating coefficient (which may be used as a design variable) associated with that function. The y-coordinates of an airfoil can be expressed by Equation (1.7). The optimum geometry and the efficiency of the design method may rely on the selection of the shape functions and the number of shape functions used to represent the airfoil.
y( x) = i f i ( x)

(1.7)

(i: Design Variables, fi: Shape Functions, x : x coordinates of curve) Mathematically, any continuous function or curve defined on a closed interval can be represented by an infinite series of normal mode functions that form a complete set of bases. However, an infinite number of series is not needed if the approximate curve represented by shape functions is within the limit of a required norm of error tolerance. Base airfoils can be added to this representation of an airfoil as shown in (1.8). Especially for gradient-based optimization methods, starting from a typical airfoil would be a reasonable strategy for searching the optimum airfoil. This would be achieved by setting the initial values of i = 0 . For a non-gradient-based method, such as the GA (Genetic Algorithm), including a base airfoil will increase the smoothness of airfoil shapes with the same upper and lower bounds on the shape function multiplier design variables.
y ( x ) = y ( x ) Base Airfoil + i f i ( x )

(1.8)

(i: Design Variables, fi: Shape Functions) This method is very effective for wing parameterization, but it is difficult to generalize it for a complex geometry and hard to find very different shapes (i.e. the design space is limited to the perturbations from the base shape available from the

i f i ( x) terms).

9 1.2.3 Global Optimization

Gradient-based optimization
In the early stages of the design process, the search for optimal airfoil shapes encompasses a broad range of possibilities. If the aerodynamic design space is smooth enough (e.g. continuous first derivatives), Gradient-based Methods (GM) usually have performance advantages over their global optimization counterparts. However, the aerodynamic performance of an airfoil is very sensitive to the surface geometry, and it is difficult to guarantee the convexness or unimodality of the objective functions used in airfoil optimization. Moreover, in transonic airfoil design, the drag objective function itself may be discontinuous due to shock waves 23. When a CFD solver is used as the function evaluator, the iterative nature of CFD can also lead to small discontinuities of the design space and false minima due to the solver convergence tolerance. One of the well-known concerns of using gradient-based optimization techniques is that they conclude their search at a point where some form of the Kuhn-Tucker conditions are satisfied; however, these conditions only describe a local optimum point. For airfoil design, this means that an airfoil shape found by a gradient-based optimizer is likely the locally optimal solution nearest to the initial airfoil shape.

Non gradient-based optimization


Recently, the Genetic Algorithm (GA) has emerged as a viable (although more costly) alternative for airfoil optimization24,
25, 26

because of its global search nature. A

GA has the ability to search highly multimodal, discontinuous design spaces. The GA also locates designs at, or near, the global optimum without requiring an initial design point. If numerous local optimum points exist in the airfoil design space, global optimization techniques will be an appropriate way to design airfoils despite the penalty of computational time increases. Furthermore, parallel and/or distributed computing can mitigate some of the computational expense associated with the GA. 1.2.4 Flow Models The optimization algorithm requires a function evaluator, and the accuracy of the function evaluator affects the reliability and validity of the optimum solution.

10 Computational Fluid Dynamics (CFD) algorithms have been developed for the past two decades and are considered an important tool for aerodynamic design 27 . With the development of CFD, the optimization method is combined with CFD and used as a design method. As CFD methods improve, flow fields around airfoils can be analyzed using more complex physics to improve accuracy. Figure 1.3 indicates a hierarchy of CFD methods at different levels of simplification. It should be noted that as the accuracy of flow physics increases, the computational cost also increases. Inviscid calculations with boundary layer coupling can predict lift and drag quite well for attached flows; however, it is not easy to find a converged solution between the outer inviscid region and the inside boundary layer with the onset of separation. Solving the full viscous equations (i.e. RANS: Reynolds Averaged Navier Stokes) is needed for complex separated flows. RANS requires turbulence modeling. Recently, with the increase of computing power; research is going on about Large-Eddy Simulation (LES) 28, Direct Numerical Simulation (DNS) 29, and Detached Eddy Simulation (DES) 30 which resolve part (or all for DNS) of the flow directly. However, due to their significant computational time these flow models are not yet used as an early design tool.

V. DES, LES, DNS (2000s and beyond) +Turbulence


Increasing Complexity More Accurate Flow Physics

IV. RANS (1990s) +Viscosity III. Euler (1980s) +Rotation II. Nonlinear Potential (1970s) +Nonlinearities I. Linear Potential (1960s) Inviscid, Irrotational Linear

Decreasing Computational Costs

Figure 1.3 The hierarchy of mathematical models

11 Several airfoil design methods are reviewed in this section. With the increase of computer power and the development of numerical methods, the optimization approach is widely used for designing airfoil and aircraft. Optimization would provide innovative design and a logical, systematic, decision process. However, optimization is still computationally expensive when using CFD for analysis.
1.3 Morphing Aircraft

Significant research has been carried out to understand how birds fly to gain inspiration for improved man-made aircraft 31. Birds can morph their wing shape to attain maneuverability and maximum performance in different flight conditions. Recent enhancement of materials and actuation technology has encouraged the concept of a Morphing Aircraft. Morphing aircraft and related programs are briefly introduced in this section to provide research motivations. 1.3.1 Return to the First Flight In 1903, Wright Flyer succeeded the first heavier-than-air powered flight in human history. One of the unique contributions of Wright brothers efforts was the design of successful flight control mechanism of all three axis of the airplane. Particularly, they used wing warping for lateral (roll) control illustrated in Figure 1.4. This unique design might have been inspired by a birds control system composed of bending and twisting of its wing.

Figure 1.4 Front-view of Wright Flyer, illustrating wing warping (From Combs, Kill

Devil Hill, Houghton Mifflin, Boston,1979.)


As the aircraft became bigger and faster, the aerodynamic efficiency of the warping has been sacrificed by the requirement of structural stiffness. Modern aircraft have a fixed, stiff wing with small hinged control surfaces and high-lift devices. These fixed-

12 geometry wings are often designed for one mission capability or are designed as a compromise among several capabilities. Also, discrete control surfaces deteriorate aerodynamic efficiency by adding leakage and protuberance drag. Recently developed smart material techniques can change the shape of the aircraft while maintaining stiffness. This advancement turns our eye back on the first flying machine and draws our attention to the design of a single aircraft that has multi-mission capability and aerodynamic efficiency through morphing of the wing. 1.3.2 Definition of Morphing Aircraft Usually, the word morphing means substantial shape change or transfiguration. In the context of NASAs research on future flight vehicles31, morphing is defined as efficient, multi-point adaptability. Efficiency implies mechanical simplicity and system weight reduction. Multi-point denotes accommodating diverse mission scenarios, and adaptability means extensive versatility and resilience. A Purdue research group defined the morphing aircraft as A multi-role aircraft that, through the use of morphing technologies (e.g. innovative actuators, effectors, mechanisms), can change its shape to perform each of several dissimilar mission roles as though the aircraft had been designed for each specific role 32. Current conventional aircraft also reduce the penalties of single shape design through the deflection of typical leading and trailing edge hinged control surfaces and high lift devices. However, these hinged and discrete systems are highly complicated mechanical devices with reliability problems and have gaps and external mechanisms that induce a drag increase. A morphing aircraft may use smooth, deformable leading and trailing edges, or possibly fully deformable airfoil sections instead of conventional discrete movable surfaces. 1.3.3 Morphing Aircraft Related Programs

Mission Adaptive Wing (MAW) program


In the early 1980s, the supercritical wing on the F-111 aircraft was replaced with a wing called a mission adaptive wing (MAW) 33 illustrated in Figure 1.5. NASA and the

13 US Air Force launched a joint program called Advanced Fighter Technology Integration (AFTI).

Figure 1.5 MAW modifications to F-111 (From NASA TM-4606)

The Mission adaptive wing is composed of numerous hinge points covered with glass reinforced plastic and mechanical actuators to alter the camber of the wing during flight. Figure 1.6 shows a set of trimmed flight test results illustrating the drag reduction capability through variable camber over the baseline aircraft results for the two different Mach numbers. The drag reduction changes from about 8 percent at the design cruise point (CL=0.4, M=0.70) to over 20 percent at an off-design condition (CL=0.8, M=0.70).

Figure 1.6 Flight-determined drag polar comparison (From NASA TM-4606)

14 Although the successful flight test of the MAW F-111 proved the benefits of variable camber, the wing was generally considered too heavy and complex for practical applications. The challenge was to find a way to easily bend a wing that is stiff and strong enough to sustain the high loads of an aircraft in flight, with motors small enough to fit inside its narrow space.

Smart Wing program


DARPA (Defense Advanced Research Projects Agency) initiated the Smart Wing program 34 in 1995 to incorporate the benefits of variable camber of MAW and variable wing twist of Active Aeroelatic Wing (AAW). The overall objective of the Smart Wing program was to develop smart technologies and demonstrate novel actuation systems to improve the aerodynamic and aeroelastic performance of military aircraft. Many researchers have investigated the use of fully integrated adaptive material actuator systems (so called "smart technologies") for performance enhancing shape control. The smart wing utilizes nickel-titanium (NiTi) shape-memory-alloy (SMA) to actuate trailingedge control surfaces and a SMA internal torque tube to provide hinge-less, smoothly contoured shape control (Figure 1.7) and variable spanwise wing twist. These devices offer a significant advantage over conventional wings because they have no flowdisturbing hinge lines.

SMA Wires

Figure 1.7 Smart Technologies

Active Aeroelastic Wing (AAW)


The Active Aeroelastic Wing (AAW) 35 flight research aircraft (F/A-18) completed its first flight test in early May 2003. The Active Aeroelastic Wing (AAW) concept is turning the wings aeroelastic flexibility into a net benefit through the use of multiple

15 leading and trailing edge control surfaces activated by a digital flight control system. AAW utilizes the energy of the air stream to achieve the desired wing twist with very little control surface motion. The wing then creates the needed control forces with less effort than an inflexible wing. At higher dynamic pressures, the AAW control surfaces are used as "tabs" that promote wing twist for added control force capability instead of trying to overcome control surface losses due to wing elastic twist. At these high dynamic pressures, large amounts of control power can be generated using the control surfaces to twist the wing. In the same design, the AAW control can also minimize drag at low wing strain conditions and minimize structural loads at high wing strain conditions.

NASAs Morphing Aircraft program


NASAs morphing program31, established in 2002, focuses on the areas of adaptive materials and structures, micro active flow control, and biologically inspired technologies. The project strategically incorporates both micro fluid dynamics and small/large-scale structure shape change. This project also addresses the intertwined functions of vehicle aerodynamics, structures and controls to seek new innovations that may only be possible at the intersection of disciplines. In this program, multi-disciplinary approaches of advanced adaptive technologies are a crucial part for leading development of future air and space vehicles beyond simple replacements of conventional technologies. Figure 1.8 is an artistic concept of a morphing airplane.

Figure 1.8 Morphing airplane (NASA)

16

DARPAs Morphing Aircraft structures program


Research is going on in DARPAs morphing aircraft structures project to create shape-changing, multi-mission aircraft using smart materials. The type of geometric adjustments that DARPA investigates include a 200% change in aspect ratio, 50% change in wing area, 50% change in wing twist, and a 20-degree change in wing sweep. As an example, sliding skins and folding wings could morph a plane from fast, attack configurations to as slower long distance shape. Figure 1.9 and Figure 1.10 illustrate these concepts 36.

Figure 1.9 Sliding skins concept (Image: NextGen)

Figure 1.10 Folding wing concept (Image: Lockheed Martin)

One of the attractions of using shape changes to control an aircraft inflight is that it could eliminate the need for traditional flight-control surfaces, a large source of radar reflections. Even more important is that morphing could greatly improve performance at two dissimilar flight conditions. Researchers in the past have tried to change wing shapes using more conventional, actuator-based approaches; however, DARPA considers them less efficient than a smart materials-based solution. The success of the project will depend on whether the wing designs are stiff enough to handle the aerodynamic forces and carry the aircraft loads. In addition, a shape-changing wing will weigh more than a traditional structure of the same size, thus the weight penalty should be minimized for success of the morphing program.

17
1.4 Motivation of Research

1.4.1 Global Optimization Issues in Airfoil Optimization Optimization algorithms have changed the aerodynamic shape design process and have provided a logical and systematic decision-making procedure. Optimization may provide a possibility to find innovative designs through parameter adjustments beyond human perception or intuition. A diverse array of optimization techniques have been applied for aerodynamic design problems. However, the use of local optimization tools (gradient-based methods; GM) may risk missing the best designs. Recently, the Genetic Algorithm (GA), which is known as global search algorithm, has emerged as viable alternative for airfoil optimization. If the aerodynamic design space is unimodal and convex, Gradient-based Methods (GM) usually have performance advantages over their global optimization counterparts. However, the aerodynamic performance of an airfoil is very sensitive to the surface geometry, and it is difficult to guarantee the convexness or unimodality of the objective and constraint functions used in airfoil optimization. Furthermore, in transonic airfoil design problem, the presence of a shock wave affects discontinuity and multimodality of the design space. Benefits of using global optimization methods have been addressed by several authors. Obayashi and Tsukahara 37 compared GA and GM and showed the necessity of global optimization in airfoil design using a simple problem with a few variables. Recently, Holst and Pulliam 38, compared GA and adjoint methods for transonic airfoil design focusing on the multi-objective problem and computational efficiency. Most of the previous research was focused on the methodologies and the advantages of using each optimization algorithm on designing airfoil. However, several issues of the actual airfoil optimization process still need to be addressed and investigated. A thorough comparison of various aspects of a GM and the GA is required to decide the appropriate optimization method for airfoil design.

18 1.4.2 Morphing Airfoil Design Recently, a deformable wing concept was introduced as a future aircraft technology. A newly introduced morphing aircraft can change its wing freely like a bird. Most of the previous morphing aircraft related research was focused on device development for changing the shape of the aircraft. However, a system level approach on the overall performance of morphing aircraft including vehicle weight, size, cost has not been investigated as much. The concept of morphing aircraft has generated a new design goal for aerodynamic configuration design and optimization area. Most aircraft wings are designed as a single shape resulting in a compromise of performance at several design conditions. A single point design will have performance penalties at flight conditions other than the design condition. Generally, a multi-point problem formulation is used for airfoil or wing design. The multi-point optimization includes tradeoffs between the multiple design objectives. In the case of a morphing aircraft, the design objective is no longer a single aerodynamic objective as in the conventional aerodynamic shape optimization. It does not require aerodynamic tradeoffs between multiple design objectives. The morphing aircraft can adjust the wing shape to the optimal shape for any flight condition encountered by the aircraft. This adaptation will improve the aerodynamic performance of the aircraft. However, morphing aircraft requires an extra mechanism to change the wing shape. Therefore, to account for the total performance benefits of morphing aircraft, a new design strategy is needed to embrace the variation of the wing shape in the optimization process that addresses the effort needed to morph the shape of the airfoil. Another issue of morphing aircraft optimization is how to model the morphing cost. In previous research, Prock 39 et al suggested using a simple strain energy model for morphing aircraft design. This model is based on the assumption that the strain energy stored in the wing will be similar to the energy required for the actuator. Also, it is assumed that the actuator energy will be proportional to the morphing cost. This simple spring actuation model, however, neglected the effect of aerodynamic load. In real flight conditions, the effect of aerodynamic force cannot be neglected. If the stiffness of the wing is very high, the strain energy will be enough to account for the actuation energy.

19 However, if the stiffness decreases, which is required for morphing aircraft, the aerodynamic force is relatively more important and needs to be included in the actuation energy model.
1.5 Thesis Objectives

The first objective of this research is to investigate the issues associated with using a global search for transonic airfoil design by comparing the Genetic Algorithm (GA) with a Gradient-based optimization method (GM). The issues to be addressed in this transonic airfoil design study will include selecting design variables and their boundaries, trimming, computational cost, existence of local optima, CFD grid selection and choosing the base airfoil. The second objective is to develop a design process and strategy that incorporates morphing cost in the design of morphing airfoils, based on the method developed through the transonic airfoil optimization. The resulting airfoil set should provide the desired aerodynamic performance and also satisfy multiple missions with the smallest actuation energy (morphing cost). This will show the tradeoffs between improved performance and actuation energy. The third objective is to improve the accuracy of the morphing energy model by incorporating an aerodynamic load model. The suggested actuation model (morphing cost model) with an aerodynamic load term would be more realistic compared to the simple strain energy model.
1.6 Thesis Organization

In Chapter 2, most of the optimization issues discussed in section 1.2 are addressed for the transonic airfoil optimization problem and the development of the transonic airfoil optimization process is covered. Chapter 3 focuses on formulating a new objective function for morphing aircraft optimization. The airfoil optimization strategies developed in Chapter 2 are employed. Chapter 4 discusses the actuation energy modeling for the morphing aircraft. Aerodynamic work is modeled and added to the simple strain energy

20 model used in Chapter3. Finally, Chapter 5 summarizes the work with concluding remarks and future directions.

21

CHAPTER 2 TRANSONIC AIRFOIL OPTIMIZATION

To investigate the issues in airfoil optimization, drag minimization problem in the transonic regime is selected as a test case. Transonic airflows present some of the most complex challenges in aerodynamics because of the air behavior as it transitions from subsonic to supersonic speeds. The feasible direction method (CONMIN module) and Genetic Algorithm (GA) were selected as optimization algorithm. In transonic flow, the major source of the drag is the wave drag due to the shock wave. Thus, it is expected that the gradient-based method will stop at the first shock free conditions. However, the shock wave is not the only source of the drag in transonic flow. Many different shock-free airfoils creating a several local minima may exist. This can be tested by comparing the gradient based optimization results started from several different initial points, and this will be shown in this chapter. Because of the expected multimodal design space, a global optimization method seems more appropriate for the transonic airfoil optimization problem. An accurate flow solver, enough to resolve all the drag components, is needed to investigate the transonic airfoil optimization problem, which will increase the computational time dramatically for a global optimization technique. Computational cost is a primary issue for the GA to be considered as an alternative to a gradient-based method. Fortunately, the nature of the GA algorithm is well suited to parallel processing and could reduce GAs computational cost to an affordable level. The parallelization issue of the GA is also presented in this chapter.

22 The detailed process of designing the minimum drag airfoil using the GA and GM and the results comparisons between these two methods are presented to help to identify the issues.
2.1 Design Variables

An airfoil shape can be regarded as a curve and there are many ways to parameterize a curve. The intend is to represent the airfoil with a small number of variables with certain accuracy and maintaining reasonable shapes during the search. Samareh 40 surveyed shape parameterization techniques and compared the pros and cons of each method. However, there are no clear guidelines to select one most efficient parameterization method. The shape representation is especially important for global optimization. A large number of design variables can increase computational cost. Also, during a global search, a wide range of variable values may be investigated. If combinations of variables happen to represent an unrealistic airfoil, e.g. with upper and lower surfaces crossed, the global search method may likely encounter this combination. To choose an appropriate shape representation and set of design variables, a simple reconstruction test is performed. The reconstructed shape is acquired by the least-square solution of the difference between the original airfoil shape and the regenerated airfoil shape. Two parameterization methods are tested. One is the analytical approach 41 that uses shape functions added to a base airfoil shape and the other is the spline40 approach. For the analytical approach, the y-coordinate positions are represented by Equation (2.1). Three different shape functions (modified Hicks-Henne41, NACA normal mode 42 and Wagner functions 43) are selected for this test, and the NACA0012 airfoil is used as the base airfoil. As the design variables i vary from zero, the represented airfoil shape varies from the base airfoil shape.

23

y ( x, i ) = y ( x ) Base Airfoil + i f i ( x )
(i: Design Variables, fi: Shape Functions)

(2.1)

A Bezier curve 44, described in Equation (2.2), is tested as an example of a spline approach, where the design variables are multiplied by a polynomial function. There is no base airfoil in this case.

y ( x, i ) = i Bi , p ( x )
(i: Design Variables, Bi,p: degree P Bernstein polynomials )

(2.2)

All shape functions used in this thesis and Bernstein polynomials are shown in Figure 2.1. For the investigation of shape representations, the Whitcomb supercritical airfoil45 are selected as a target airfoils; the NACA0012 is used as the base airfoil (as mentioned above). Approximate sets of y-coordinates were regenerated by the solution design variables that minimize (yactual y(i)) at several ordinate(x) locations on the airfoil. The least square solution is shown in Equation (2.3).

i = ( AT A) 1 AT b
(b: y coordinate of target airfoil, A: matrix composed of x ordinate of each shape functions)

(2.3)

A total of 16 design variables (8 variables for each surface) are used in this test. The norm of the error was compared in Table 2.1, and the shape difference is compared in Figure 2.2. Table 2.1 shows that the analytical method with the modified Hicks-Henne functions has the best reconstruction capability for both the upper and lower surfaces. In this research, the analytical method with modified Hicks-Henne functions is chosen for representing airfoils. The design variables are multipliers that determine the magnitude of the shape function as it is added to the baseline shape. Figure 2.1 depicts the shape functions and Figure 2.3 shows their individual effects on a baseline NACA 0012 airfoil when one of the values is 0.015.

24

Table 2.1 Error norm comparison (Target airfoil: Whitcomb supercritical airfoil)
Modified Hicks-Henne functions Upper surface NACA Normal Mode functions 0.00088 Wagner Functions 0.00130 Bezier Curve 0.01680

y original y estimate
Lower surface

0.00200
2

y original y estimate

0.00330
2

0.00510

0.00440

0.01620

25
f k ( x ) = sin[ (1 x )e ( k ) ], k = 1,2 f k ( x) = sin 3 [x e ( k ) ], k = 3,4,5 f k ( x) = sin[x e ( k ) ], k = 6,7,8

f1 ( x) = x x
f 2 ( x) = x(1 x)

f k +1 ( x) = x k (1 x) , for (k=2,3,4,5)
f 7 ( x) = 3 x x f 8 ( x) = 4 x 3 x

where,
e(k ) = ln(0.5) ln(1 x k ) , k = 1,2

e(k ) = ln(0.5) ln( xk ) ,

k = 3,4,5,6,7,8

x k = 0.06, 0.13, 0.2, 0.4, 0.6, 0.8, 0.87, 0.94


1 0.25

y 0.5 y

0.5 x

0.5 x

(a) modified Hicks-Henne functions

(b) NACA normal mode functions

f1 ( x) =

+ sin( ) sin 2 2
sin(k ) sin((k 1) ) + for k>1 k

n i n i Fn ,i ( x) = i x (1 x)
where, n=10, , i= 0, n i = i!(n i )!

f k ( x) =

n!

where, = 2 sin 1 ( x )
0.5 1

y 0.5

-0.4 0

0.5 x

0.5

(c) Wagner functions

(d) Bernstein polynomials

Figure 2.1 Shape functions

26

0.06

0.06

0.04

0.04

0.02

Whitcomb Reconstructed (Hicks-Henne functions)

0.02

Whitcomb Reconstructed (NACA mode functions)

-0.02

-0.02

-0.04

-0.04

-0.06

0.2

0.4

0.6

0.8

-0.06

0.2

0.4

0.6

0.8

(a) Hicks-Henne functions

(b) NACA normal mode functions

0.06

0.06

0.04

0.04

0.02

Whitcomb Reconstructed (Wagner functions)

0.02

Whitcomb Reconstructed (Bezier functions)

Y
0 0.2 0.4 0.6 0.8 1

-0.02

-0.02

-0.04

-0.04

-0.06

-0.06

0.2

0.4

0.6

0.8

(c) Wagner functions

(d) Bezier curve

Figure 2.2 Original versus reconstructed airfoils

27

0.1 0.075 0.05 0.025 0

-0.025 -0.05 -0.075

0.25

0.5

0.75

Figure 2.3 NACA0012 with each shape function applied

28

2.2 Base Airfoils

Using the analytic shape representation also entails the additional issue of selecting a base airfoil. Three base airfoils are chosen from Ref. 46; these airfoils are the NACA 0012, the RAE 2822, and the Whitcomb supercritical airfoil. The NACA 0012 is a subsonic, symmetric airfoil, while the RAE 2822 and Whitcomb are cambered airfoils originally designed to reduce wave drag in transonic flight conditions. Figure 2.4 shows the difference of base airfoils. Transonic airfoils have smaller surface curvatures than the NACA0012 on the upper surface and have a concave lower surface at the rear of the airfoil.

0.15
NACA0012 RAE2822 WHITCOMB

0.1

0.05

-0.05

-0.1

Base airfoils comparison


0 0.25 0.5 X 0.75 1

Figure 2.4 Configuration of base airfoils As part of investigating issues for global optimization of airfoils, a global method (GA) and a gradient-based method (GM) will be compared. The role of the base airfoils is different for the GA and the GM. In the case of the GM, an initial design point is required; this initial shape can be changed by starting from a different airfoil. If all i are set to zero for the initial design in Equation (2.1), then GM will start from the base airfoil. However, the GA does not have an initial design point. Instead, the GA has randomly generated initial populations that are generated using shape functions applied to the base

29 airfoils. Therefore, through altering the base airfoils, the GA can have a different range of design space (possible airfoil shapes) with the same values for the upper and lower bounds on the design variables. Figures 2.5-2.7 show the available design space using different base airfoils with the modified Hicks-Henne shape functions. Because the shape function perturbation is added to the airfoil shapes as depicted in Equation (2.1), changing the base airfoil alters the design space even with the same bounds on each design variable. The maximum limit is with all i at their upper bound; the minimum limit is with all i at their lower bound. The search method, either global or gradient based, should be able to find any airfoil shape with surfaces between the two bounds. The Whitcomb base airfoil results in a notably different design space than the other two, especially in the aft portion of the airfoil. If one global minimum shape exists and, moreover, that point is within the design space region shared by all base airfoils used in the shape representation, the global optimum point can be expected to be found by a global optimization method regardless of the base airfoils. On the other hand, if the global minimum shape lies outside the bounds of the design space for one or more base airfoils, a different solution will be acquired for these airfoils.

0.15
NACA0012 Maximum Limit Minimum Limit

0.1

0.05

-0.05

-0.1

0.25

0.5 X

0.75

Figure 2.5 Design space available using the NACA0012 base airfoil

30

0.15
RAE2822 Maximum Limit Minimum Limit

0.1

0.05

-0.05

-0.1

0.25

0.5 X

0.75

Figure 2.6 Design space available using the RAE2822 base airfoil

0.15
Whitcomb Maximum Limit Minimum Limit

0.1

0.05

-0.05

-0.1

0.25

0.5 X

0.75

Figure 2.7 Design space available using the Whitcomb base airfoil

31
2.3 Trimming

In optimal airfoil shape design, the angle of attack also can be, and often is, used as one of the design variables, so that the solution contains the shape and angle of attack needed to minimize drag and meet the lift constraints. Instead of using the angle of attack as a design variable, trimming has been applied in this research. This process may reduce the complexity of the design space and help find a global optimum design, because the number of design variables is smaller. Also, it is possible that two different airfoil shapes at different angles of attack can have same lift coefficient. However, trimming requires more function evaluations, thereby increasing the computational cost for each function evaluation compared to using the angle of attack as the one of the variables. In this approach, the lift curve slope of an airfoil shape is estimated using two flow solutions, and then the angle of attack corresponding to the desired lift is predicted. Then, this angle of attack is used as an input for the flow solver to provide values of Cl , Cd and Cm for the constraint and objective functions (Figure 2.8). Thus, two extra evaluations are needed to find the angle of attack for the design lift coefficient. As a result of this process, the airfoil will meet the specified lift coefficient, if the shape has a constant lift curve slope. Occasionally, some airfoil shapes do not have a linear Cl versus trend, in that case more than two evaluations are needed to resolve the design Cl values. This could be an issue for the GA, especially in early generations, when several airfoil shapes do not have a linear lift slope (maybe due to early separation regions) or have a shape impossible to satisfy the design Cl by only changing the angle of attack. In this research, an explicit penalty type constraint is still used to handle these poor airfoils while limiting the trimming iterations to two extra flow evaluations.

32

Cl

2 [ rad 1 ]
Cldes

real slope

1 = 0 D

Figure 2.8 Schematics of trimming process

2.4 Objective Function

The simplest airfoil shape optimization problem is a single-point optimization. For this, the objective function is the drag coefficient at one specified Mach number (see Equation (2.4)); this is to be minimized while providing a specified value for lift coefficient. This formulation will be investigated with both the GA and a GM. Minimize: Subject to:
F ( i ) = C d1
C l = C l1 , imin i imax

(2.4)

Airfoils designed for a single flow condition may have poor performance at offdesign conditions. This is well presented in Ref. 47. To overcome these problems of single-point designs, a multi-point optimization approach is often used. Multiple objectives representing drag at several flow conditions can be combined into one objective using a weighted sum. For a two-point shape optimization, the problem formulation is shown below utilizing a weighting factor ( ) in the objective function.

33 Minimize: Subject to:


F ( i ) = * C d1 + (1 ) * C d 2
Cl = Cl1 , Cl = Cl2 , imin i imax

(2.5)

In this formulation, the intent is to minimize the drag coefficient at two different Mach numbers. The choice of the weighting factors affects the contribution of each term in the objective function. The weighting factors of used by Drela47 in his discussion of airfoil optimization appear to provide a rational objective function and are used herein. Equality constraints ensure the desired lift coefficient is obtained in both Mach number conditions. Also, this can be extended to problems with three or more Mach number conditions.
2.5 Accuracy of Function Evaluations

To consider the viscosity effects in transonic airfoil design, the TURNS (Transonic Unsteady Rotor Navier-Stokes) code 48 developed at NASA Ames is used in this research. The TURNS code was originally developed to research the unsteady flow around a rotor blade, however this code can be used for a 2-D airfoil. In the TURNS code the inviscid flux is calculated based on an upwind-biased flux-difference scheme originally suggested by Roe 49 . To acquire higher order accuracy the Van Leer MUSCLE (Monotone Upstream-centered scheme for the conservative laws approach) 50 is applied with flux limiters in order to be total variation diminishing (TVD). For the implicit operator the Lower-Upper-Symmetric Gauss Seidel (LU-SGS) scheme suggested by Jameson and Yoon 51, 52 is used. A large time step ( t = 50 , where the reference time scale, t = C / a ) is applied to acquire a steady state solution needed for optimization algorithm, as suggested in Ref. 48. Finding a proper grid size is another issue in aerodynamic shape design, especially for global optimization that requires many function evaluations. A fine grid will require a long time for solver convergence and a sparse grid will affect the reliability of the solutions. A parametric study for the grid was performed to determine a proper grid size.

34 The RAE2822 airfoil is selected for the grid test and the flight speed was M=0.74 with an angle of attack = 3.19 D . Figures 2.9 to 2.11 show the grid systems selected for the test. The computation results were compared with experimental data 53. A comparison with other grids in Figure 2.12 showed that a 12930 grid arrangement and resolution is a good compromise between accuracy and efficiency. 101 grid points are located on the surface of airfoil. Figure 2.12 compares the surface pressure distribution from TURNS with the experimental data for the RAE2822 airfoil at the design point. The convergence history (Figures 2.13-2.14) generally showed that after 2000 iterations, the difference of Cl and C d between subsequent iterations was less than 10 5 . From this result, an upper limit of 2000 flow solver iterations is set for each airfoil evaluation during the optimization. The maximum residual is reduced about four orders of magnitude during a typical solution. Each function evaluation required a wall clock time of approximately 220 sec for 2000 iterations on one processor of a Linux cluster machine that has AMD Athlon 1.2GHz CPUs. A typical GA run requires many function evaluations (i.e. 35,000 runs or more in this research), thus the choice of the grid and the iterations required has a significant impact on the total CPU time.

Table 2.2 N-S solution comparison with experimental data


Cl Experiment 79 by 25 Grid 129 by 30 Grid 217 by 61 Grid 0.73300 0.72871 0.73743 0.77010 Cd 0.01880 0.02600 0.02533 0.02584 Cm -0.08600 -0.08796 -0.08794 -0.09135 CPU time (2000 iterations) N/A 81 Sec. 168 Sec. 618 Sec.

35

6 0.5 4 2 0 -2 -4 -0.5 -6 0

-5

0 X

10

-0.5

0 X

0.5

Figure 2.9 Grid using 79 by 29 points (RAE2822 airfoil) Physical domain (left), magnified picture near the airfoil (right)

6 0.5 4 2 0 -2 -4 -0.5 -6 0

-5

0 X

10

-0.5

0 X

0.5

Figure 2.10 Grid using 129 by 30 points (RAE2822 airfoil) Physical domain (left), magnified picture near the airfoil (right)

36

6 0.5 4 2 0 -2 -4 -0.5 -6

-5

0 X

10

-0.5

0 X

0.5

Figure 2.11 Grid using 217 by 61 points (RAE2822 airfoil) Physical domain (left), magnified picture near the airfoil (right)

-1.5

-1

-0.5

Cp

0.5

Experiment Grid (79 by 25) Grid (129 by 30) Grid (217 by 60)

1.5

-0.25

0.25 x

0.5

0.75

Figure 2.12 Pressure coefficient distribution of N-S prediction vs. published experiment (RAE 2822 airfoil, M=0.74, =3.19)

37

0.9

0.8

0.7

Cl

0.6

0.5

0.4

0.3 1000 2000 3000 4000 5000

Iterations

Figure 2.13 Lift coefficient convergence history

10

-1

Cd
10
-2

10

-3

1000

2000

3000

4000

5000

Iterations

Figure 2.14 Drag coefficient convergence history

38

2.6 Computational Cost

2.6.1 Parallel Genetic Algorithm Since its first description, the GA has been applied to many engineering optimization problems 54,55. The GA has the ability to search highly multimodal, discontinuous design spaces. The GA can also locate designs at, or near, the global optimum without requiring a good initial design point. Because the transonic airfoil design space has local minima and discontinuities caused by shock waves, the GA is an appropriate search method for transonic optimization. However, using a GA for design optimization can be computationally expensive. To overcome the computational time problem, the GA is adapted to a coarsegrain parallel implementation. The GA algorithm is inherently parallelizable, because for each airfoil out of the total airfoil population (which is usually several hundreds), the objective function evaluation (i.e. the Navier-Stokes solver) can be done in parallel and independently of other airfoils. In this research, a manager-worker type parallelization is applied to convert a serial GA into a parallel program. The manager CPU generates the population for each generation and also distributes/gathers the design variables to/from worker CPUs. The worker CPUs are computing the function using the design variables of each individual from the manager CPU. Generally, the population size is much larger than the CPU numbers; the population is divided by the available CPUs. A Linux-Cluster machine was used for calculation following the basic approach of Ref. 56. To illustrate the speed-up of the parallel GA, a single flow condition problem was solved using TURNS for the N-S evaluation needed in the objective function. Figure 2.15 shows the total wall-time for the parallel GA versus the number of processors. The total computational time for the 24 Navier-Stokes evaluations in the Linux cluster machine (with nodes connected via a 100base-T Ethernet) decreases almost exponentially as the number of processors increases. The difference from the ideal speed-up is caused by the communication time.

39

3000 2500 2000 Computational Wall-time (Sec.) 1500 1000

Ideal Speed-Up Parallel GA

500

5 10 Number of CPUs

15

20 25 30

Figure 2.15 Speed-up of parallel GA, computational wall time vs. number of CPUs.

40 2.6.2 Gradient Based Optimization Method If the objective function and constraints provide a unimodal convex domain and are also differentiable, a gradient-based optimization method can find the global optimum solution much faster than the GA. However, it is very hard to prove convexness and ensure continuity of functions for general engineering design problems. The transonic airfoil design problem is also difficult to characterize as convex and unimodal, because the presence of a shock wave affects the discontinuity and multimodality of the design space. For this effort, the method of feasible directions 57, as provided by the CONMIN11 subroutines, is used as the gradient-based optimizer. In this research, optimization runs started from different initial airfoil shape designs are used to check the consistency of optimum points found by the GM. This can give some indication of multiple local minima appearing in the design space.
2.7 Objectives and Fitness Function Formulation

2.7.1 Objectives For the single-point optimization problem, the objective is to minimize the drag when the free stream velocity is M=0.74 while producing a lift coefficient Cl=0.733. A summary of this single-point optimization problem is presented in Table 2.3 Table 2.3 Problem formulations Single Point Minimize
F ( i ) = C d1

Multi Point Minimize (where, M 1 =0.74)


F ( i ) = * C d1 + (1 ) * C d 2

Subject to
C l = C l1 = 0.733 ,
0.015 i 0.015

(where, M 1 =0.68, M 2 =0.74,=1/3) Subject to


Cl = Cl1design = Cl2 design = 0.733 ,

0.015 i 0.015

41 A single-point design usually has poor performance in off-design conditions. To overcome this, a multi-point design is suggested and applied herein. A two-point design case was set-up to investigate the effects of the objective function selection. The two design Mach numbers are M=0.68 and M=0.74 as suggested by Drela47. The design lift coefficient is 0.733 for both Mach numbers. Thus, the drag coefficient at the second design point (M=0.68) is also included in the objective function Equation (2.4) with the weighting factor of =1/3 in Equation (2.5). To satisfy the lift coefficient condition, the trimming used in single-point optimization was applied for each design condition separately, which results in two different angles of attack one for each Mach number. The computational time of the multi-point design increased to twice that of the single-point design. 2.7.2 Fitness Function Formulation For the Genetic Algorithm, constraint violations are added to the objective function using an exterior penalty. Even though trimming is already applied, a lift coefficient penalty term is added to the objective function to ensure that the lift coefficient constraint is satisfied. This is because some airfoil shapes, particularly those in the initial, randomly generated population, do not have a linear lift curve slope. For convenience, the fitness function is multiplied by 100 to increase the value, because the order of magnitude of the drag coefficient is very small. The equality Cl constraint has been changed to an inequality constraint to maintain the error smaller than the tolerance value (=0.002, see Table 2.4). The input values for GA are described in Table 2.4. The GA begins with a set of solutions, represented by chromosomes, called the population. Solutions from one population are taken and selected based on their fitness to form a new population. The more suitable individuals, represented by binary code, receive more chances to reproduce. More details of the GA can be found in Ref. 54. The population size and mutation rate were selected using empirically derived guidelines 58 for GA using tournament selection and uniform crossover. Seven bits represent each of the 16 shape function multipliers, (8 for upper surface and 8 for lower surface) for a total chromosome length of 112 bits.

42 Elitism, which copies the best design from the current generation into the next generation, is used in the tournament selection here. Each run was halted after 80 generations with no other stopping criteria. The gradient-based method CONMIN uses the method of feasible directions 59 to perform its search through the design space. In the case of the GM, the airfoil trimming strategy is applied, but the constraint on lift coefficient is set separately to handle shapes without linear Cl vs. behavior. For comparison to the GA, an inequality-type constraint was used as in Table 2.4. Because the CONMIN uses finite differencing for the gradient information, a quadratic constraint form is needed to be continuously differentiable. The TURNS code is also used as the independent flow solver.

Table 2.4 Optimization methods GA Maximize


F ( x i ) = 100 * {C d1 + r * [ MAX (0, g 1 )] 2 }

GM Minimize
F ( x i ) = 100 * {C d1 }

Where
i
min

g1 =

Cl Cl1

Subject to
i
min

g1 =

(Cl Cl1 ) 2

1 0 ,

i imax , r = 0.0001

i imax

(8)

Base airfoils

Starting airfoils

NACA0012 RAE2822 Whitcomb Super Critical Airfoil

NACA0012 RAE2822 Whitcomb Super Critical Airfoil

GA parameters Population size: 448 Resolution: 7 bit Design variables: 16 Total chromosome length bits: 112 Variable limits: 0.015 i 0.015 Mutation probability: 0.0022

43

2.8 Single-point Optimization Results Comparison

2.8.1 GA Results Drag coefficients of the best airfoils after 80 generations are presented in Figure 2.16 for runs conducted using each of the three base airfoils. The Cd values of these airfoils are compared to the Cd of the base airfoil. The drag data for NACA0012 base airfoil in Table 2.5 is not available, because the NACA0012 airfoil can not obtain the design Cl =0.733 at the speed of M=0.74 due to a strong shock wave. Figure 2.16 presents the fitness history of the GA. Here, the fitness is maximized, which is corresponding to reducing the drag objective. This trend shows a monotonic behavior, because elitism has been applied for the GA. In Figure 2.16, although the best fitness values in the initial generations are different, the best fitness values are converging to about the same value as the generations increase. Using the shape function approach to represent changes in the airfoil shape requires a base airfoil. However, the initial generation of the GA is generated randomly, so the GAs search does not begin with the base airfoil. The best airfoil shape encountered in selected generations during a parallel GA run is shown in Figures 2.17-2.19 along with the base airfoil sections for each of three runs. For all three runs, the airfoils decrease in overall and adjust camber over subsequent generations. The pressure distribution plots in Figures 2.17-2.19 show that the best airfoils of early generations have small shocks on the upper surface. However, after generation 80, the airfoils maintain the specified design lift coefficient without shockwaves. Mach contours drawn in Figures 2.20-2.22 also show the reduction of the shock wave after the optimization. Figure 2.20 (top) shows the result of trimming using the NACA0012 airfoil. A strong shock near the leading edge induced large separation. This picture explains the reason why the NACA0012 cannot satisfy the design Cl .

44

Table 2.5 Single-point design result (GA)


Base airfoil NACA0012 RAE2822 Whitcomb Method GA GA GA Design Mach 0.74 0.74 0.74 Design Lift Coeff. 0.733 0.733 0.733 Drag(Cd) (Base Airfoil) N/A 0.02238 0.02189 Best airfoil (Generation-80) 0.01498 0.01454 0.01462 Drag reduction from base (%) N/A -35.0 -33.2

-1.4 -1.5 -1.6 -1.7 Fitness -1.8 -1.9 -2 -2.1 -2.2 -2.3
Base Airfoil (NACA0012) Base Airfoil (RAE2822) Base Airfoil (Whitcomb)

10

20

30

40 50 Generation

60

70

80

Figure 2.16 Best fitness value convergence history for all three GA runs.

45

0.3

0.2

0.1 Y 0

-0.1

NACA0012 Generation 1 Generation 10 Generation 30 Generation 40 Generation 80

0.25

0.5 X

0.75

-1.5

-1

-0.5

Cp

0.5
Generation 1 Generation 10 Generation 30 Generation 40 Generation 80

M=0.74 Cl=0.733 Base Airfoil [NACA0012]

1.5

0.25

0.5 X

0.75

Figure 2.17 Best airfoil shapes (top) and pressure coefficient distributions (bottom) in selected generations of the GA using the NACA 0012 base airfoil.

46

0.3

0.2

0.1 Y 0

-0.1

RAE2822 Generation 1 Generation 30 Generation 40 Generation 80

0.25

0.5 X

0.75

-1.5

-1

-0.5

Cp

0.5
RAE2822 Generation 1 Generation 30 Generation 40 Generation 80

M=0.74 Cl=0.733 Base Airfoil [RAE2822]

1.5

0.25

0.5 X

0.75

Figure 2.18 Best airfoil shapes (top) and pressure coefficient distributions (bottom) in selected generations of the GA using the RAE 2822 base airfoil.

47

0.3

0.2

0.1 Y 0

-0.1

Whitcomb Generation 1 Generation 10 Generation 30 Generation 40 Generation 80

0.25

0.5 X

0.75

-1.5

-1

-0.5

Cp

0.5
Whitcomb Generation 1 Generation 10 Generation 30 Generation 40 Generation 80

M=0.74 Cl=0.733 Base Airfoil [Whitcomb]

1.5

0.25

0.5 X

0.75

Figure 2.19 Best airfoil shapes (top) and pressure coefficient distributions (bottom) in selected generations of the GA using the Whitcomb supercritical base airfoil.

48

0.95 0.9 0.9 0.85 0.7 1 1 0.15 0.1 0.2

0.85 0.55 0.8

0.9 0.8 1

0.9 0.8

1.1

0.65

0.75 0.65

0.6

Figure 2.20 Mach contour for the NACA0012 airfoil (M=0.74, Cl=0.605, = 12.64D ) (top) and Mach contour for the best airfoil shape from 80th generations of GA runs using the NACA0012 base airfoil (M=0.74, Cl=0.733)(bottom)

49

0.9 0.8 1 0.85

0.95

1.2

0.6

0.7

0.8

0.7

0.8

0.9 1

0.9

0.85 0.7 1.1

0.65

0.75 0.75

0.6

0.65

Figure 2.21 Mach contour for the RAE2822 airfoil (M=0.74, Cl=0.733)(top) and Mach contour for the best airfoil shape from 80th generations of GA runs using the RAE2822 base airfoil (M=0.74, Cl=0.733)(bottom).

50

0.9 0.8

0.9

0.85

1.25 0.6

0.8 0.7

0.65

0.8

0.9 0.85

0.7 1

0.6 0.65 0.75 0.65

Figure 2.22 Mach contour for the Whitcomb airfoil (M=0.74, Cl=0.733)(top) and Mach contour for the best airfoil shape from 80th generations of GA runs using the Whitcomb base airfoil (M=0.74, Cl=0.733)(bottom).

51 The final best airfoil shapes from the three runs are compared in Figure 2.23. The upper surfaces of the airfoils are similar, but there are some differences in the lower surfaces. Because the wave drag is dominant in the transonic region, the upper surface shape is more important for reducing wave drag than the lower surface for a lifting airfoil. So, it would be expected that low wave drag airfoils might have similar upper surface shapes. In Figure 2.24 the pressure coefficients of the best airfoils after 80 generations are compared. The upper surface pressure contours exhibit some variety. The peak Cp values are about the same for all three shapes; however, the locations of the peak Cp values are different. All three best airfoils of the GA results approach the lowest limit of the lower surface position. This shows that a thinner airfoil reduces the drag. Finally, very similar best airfoils were expected regardless of the base airfoils used for GA. However, the global optimum point appears to be outside of the limits of geometry constraints (especially on lower surface of the airfoil), so the optimum point is dependent on the surface limits, which vary with the base airfoils. This suggests the necessity of defining more constraints such as limits on the moment coefficient or thickness to find a unique optimum configuration. It may be possible that two airfoils could have similar Cl and Cd, but very different Cm. The leading edge shape of the Whitcomb base airfoil is steeper than other base airfoils, so the surface limits made it difficult to have gradual acceleration near the leading edge when using the Whitcomb shape as the base airfoil. None of these pressure distributions indicate a strong shock; hence, the low predicted values of wave drag.

52

0.25

Generation 80, Base Airfoil (NACA0012) Generation 80, Base Airfoil (RAE2822) Generation 80, Base Airfoil (Whitcomb)

-0.25

0.25

0.5 X

0.75

Figure 2.23 Best airfoil shapes from 80th generations of all three GA runs.

-1.5

-1

-0.5

Cp

0.5

Generation 80, Base Airfoil (NACA0012) Generation 80, Base Airfoil (RAE2822) Generation 80, Base Airfoil (Whitcomb)

1.5

0.25

0.5 X

0.75

Figure 2.24 Pressure coefficient distributions for best airfoils from 80th generations of GA runs.

53 2.8.2 GM Results The initial design for the search is the base airfoil. Figure 2.25 shows the convergence history of the CONMIN program using the RAE2822 and the Whitcomb airfoil as the starting point. In the CONMIN subroutine, the objective and constraint function values are normalized and adjusted for calculations. In the formulation of the GA fitness function, it is necessary to find good values for the lift coefficient penalty multiplier. Similarly, scaling of variables is needed for the GM and to find the proper scale for the numerical gradient. To determine these, trial and error is required. When using the NACA0012 as the starting airfoil, CONMIN could not find any feasible directions because of the strong shock wave associated with the NACA0012 at the design conditions. As an outcome, no optimized airfoils could be found. This is noted as N/A in Table 2.6. Figure 2.26 (top) shows the change of the airfoil shape and the change of pressure coefficient during the iterations with the RAE2822 as the starting airfoil. Only small changes to the shape are made during the search. The airfoil is modified to reduce the shock of the airfoils. The final airfoil surface pressure in Figure 2.26 (bottom) shows a nearly isentropic compression on the upper surface. Figure 2.27 also shows that the final airfoil shape does not change much when using the Whitcomb airfoil as the base shape. The final airfoil starting from the Whitcomb airfoil also does not have strong shock wave, but the negative peak value of upper surface pressure is higher than the final airfoil from RAE2822. A different converged solution was acquired from two different initial airfoils by the use of the gradient-based search method. This supports the notion that the transonic airfoil design space has many local optimum design points, and the GM simply finds the local optimum nearest to the initial airfoil shape.

54

Table 2.6 Single-point design result (GM)


Base airfoil NACA0012 RAE2822 Whitcomb Method GM GM GM Design Mach 0.74 0.74 0.74 Design Lift (Cl) 0.733 0.733 0.733 Drag(Cd) Base Airfoil N/A 0.02238 0.02189 Drag(Cd) Converged Airfoil N/A 0.01546 0.01620 Drag reduction from base (%) N/A -30.0 -25.9

1.4 1.5 1.6 1.7 Fitness 1.8 1.9 2 2.1 2.2 2.3 Starting Airfoil (RAE2822) Starting Airfoil (Whitcomb) 0 1 2 3 4 5 Iterations 6 7 8 9

Figure 2.25 Convergence history for CONMIN with unmodified base airfoils as the starting shapes.

55

0.3

0.2

0.1 Y 0 -0.1

RAE2822 Iteration 1 Iteration 3 Final Iteration

0.25

0.5 X

0.75

-1.5

-1

-0.5

Cp

0.5

RAE2822 Iteration 1 Iteration 3 Final Iteration

M=0.74 Cl=0.733 Starting Airfoil [RAE2822]

1.5

0.25

0.5 X

0.75

Figure 2.26 Airfoil shape designs (top) and pressure coefficient distributions (bottom) generated during CONMIN search using RAE2822 base airfoil.

56

0.3

0.2

0.1 Y 0 -0.1
Whitcomb Iteration 1 Iteration 3 Final Iteration

0.25

0.5 X

0.75

-1.5

-1

-0.5

Cp

0.5

Whitcomb Iteration 1 Iteration 3 Final Iteration

M=0.74 Cl=0.733 Starting Airfoil [RAE2822]

1.5

0.25

0.5 X

0.75

Figure 2.27 Airfoil shape designs (top) and pressure coefficient distributions (bottom) generated during CONMIN search using Whitcomb base airfoil.

57
2.9 Multi-point Optimization Results Comparison

A comparison between the GA and a GM was made for the multi-point optimization. The RAE2822 airfoil is used as the base/starting airfoil for both the GA and the GM. Table 2.7 shows the results of the two-point optimization using the GA and CONMIN. Shapes for the base, single-point, and two-point airfoils are compared in Figure 2.28 using an exaggerated y-axis scale. For the GA, the multi-point solution showed a smaller leading edge radius than the single-point solution. In the case of the GM, all three shapes are very similar to the initial design, again, suggesting the existence of local minima. Actually, the GM result shows an increased drag at M=0.68 in order to reduce drag at M=0.74. The pressure coefficient distributions for these two airfoils appear in Figure 2.29 and Figure 2.30. At M=0.68, it appears that a shock exists for the base airfoil and the GM results, but not for the GA results. The GA results show a weaker shock wave and have lower drag at both design Mach numbers than the GM results. The drag divergence diagram is plotted in Figure 2.31 to illustrate the effect of the two-point design formulation. For both the single-point and multi-point formulation, the GA and GM generated shapes all have a drag divergence Mach number near M=0.75. The results from the GM, for single- and multi-point formulations, have higher drag than those of the GA results. At lower Mach numbers, the GM results have higher drag than the original RAE2822, which suggests the poor off design performance of these designs. In contrast, the GA shapes have a much lower drag across the range of Mach numbers, indicating that the global search results may be less sensitive to the design conditions. Table 2.7 Multi-point design result
Base airfoil Method GA RAE2822 GM Design Mach 0.68 0.74 0.68 0.74 Design Lift (Cl) 0.733 0.733 0.733 0.733 Drag(Cd) Base Airfoil 0.01513 0.02238 0.01513 0.02238 Drag (Cd) Optimized 0.01373 0.01444 0.01590 0.01544 Drag reduction from base (%) -9.2 -35.4 5.0 -31.0

58

0.1

RAE2822 Single-Point (GA, Generation 80) Multi-Point (GA, Generation 80)

-0.1

0.25

0.5 X

0.75

0.1

RAE2822 Single-Point (GM, Final Iteration) Multi-Point (GM, Final Iteration)

-0.1

0.25

0.5 X

0.75

Figure 2.28 Airfoil shapes for base, single-point, two-point design [GA (above) and CONMIN (below) results]

59

-2

M=0.68 Base Airfoil [RAE2822]

-1.5

-1

-0.5 Cp 0 0.5 1

RAE2822 Single-Point (GA, Generation 80) Multi-Point (GA, Generation 80)

1.5

0.25

0.5 X

0.75

-2

M=0.68 Base Airfoil [RAE2822]

-1.5

-1

-0.5 Cp 0 0.5 1

RAE2822 Single-Point (GM, Final Iteration) Multi-Point (GM, Final Iteration)

1.5

0.25

0.5 X

0.75

Figure 2.29 Pressure coefficient distributions for two-point objective function results at M=0.68 [GA (top) and CONMIN (bottom) results]

60

-2

M=0.74 Base Airfoil [RAE2822]

-1.5

-1

-0.5 Cp 0 0.5 1

RAE2822 Single-Point (GA, Generation 80) Multi-Point (GA, Generation 80)

1.5

0.25

0.5 X

0.75

-2

M=0.74 Base Airfoil [RAE2822]

-1.5

-1

-0.5 Cp 0 0.5 1

RAE2822 Single-Point (GM, Final Iteration) Multi-Point (GM, Final Iteration)

1.5

0.25

0.5 X

0.75

Figure 2.30 Pressure coefficient distributions for two-point objective function results at M=0.74 [GA (top) and CONMIN (bottom) results]

61

0.03
rae2822 GA (Single-Point) GA (Multi-Point) GM (Single-Point) GM (Multi-Point)

Cd

0.02

0.01

0.66

0.68

0.7 M

0.72

0.74

0.76

Figure 2.31 Drag divergence diagram

62

2.10 Lessons Learned from GA and GM Results

2.10.1 Multimodal Design Space The drag coefficients of the single-point GM results, denoted in Table 2.6 are quite different from each other. The final shape of the airfoils found using the GM is not very different from the starting airfoil. Thus, when the starting airfoil is different, the GM found very different local optimum solutions. This suggests that the transonic airfoil design space is not unimodal. Because the design space appears multimodal, numerous runs started from different initial design points may be needed to locate the best obtainable shape. In addition, the GM could not use the NACA0012 as an initial design point, because it is not in the feasible region. However, the GA could find a near-optimum solution even with the NACA0012 base airfoil, because the GA does not need an initial point. The best GA solution shows a smaller drag than the best GM solution. Compared to the GM solution, the GA solution has a much smaller thickness to chord ratio. This suggests that the GA solution has a smaller profile drag. Both the GA and the GM solutions do not have a strong shock wave, so the problem evolves to reducing the profile drag rather than the wave drag after achieving a shock free profile. In the case of the GM, the iteration process stopped at a shock free solution near to the starting airfoil. The GM could not modify the lower surface to reduce the profile drag by thinning the airfoil. However, the GA could obtain solutions that are quite similar in shape, shockless, and thinner than the original profile even though totally different base airfoils were used. The GA results also showed an unfamiliar lower surface leading edge shape, which has a negative surface pressure, that seems to have a thrust component. Another noticeable difference is that the trailing edges of the GA results are very thin and cambered. This very thin trailing edge results in a low profile drag. However, this may cause structural and manufacturing problems.

63 2.10.2 Computational Efficiency Table 2.8 compares the drag results and computational time from each optimization method for the single point optimization. The total computational time is the product of the number of function evaluations and time needed for one function evaluation using the Navier-Stokes method with the trimming approach. Because the GA is always run for 80 generations with the same population size for this effort, all the GA times are the same. In the case of the RAE2822 base airfoil, the GA result has much smaller drag coefficient than the GM result, suggesting a much better airfoil design. However, a serial GA needs about 230 times more computational time than the GM. When a parallel GA with 40CPUs is used, the GA requires only about 6 times longer than the GM. Even though the GM method used (CONMIN) in this research is not as fast as the adjoint method for this airfoil optimization problem, it is reasonable to say that the GAs penalty of a higher computational cost can be alleviated using parallelization. The benefit of using the GA is a better (lower drag) transonic airfoil shape.

Table 2.8 Comparisons of drag values and computational costs for GA and GM, (single-point runs)
Number of Base airfoil Method GA NACA 0012 GM GA RAE2822 GM GA Whitcomb GM 0.01620 264 0.01547 0.01462 459 107,520 N/A 0.01454 N/A 107,520 Drag (Cd) 0.01498 function evaluation 107,520 Computational time (Wall clock time) ~ 6 days and 19 hours (with 40 parallel processors) N/A ~ 6 days and 19 hours (with 40 parallel processors) ~1 days and 4 hours ~ 6 days and 19 hours (with 40 parallel processors) ~16 hours

64

2.11 Summary

A manager-worker type parallel Genetic Algorithm has been developed and applied for the transonic airfoil design problem. The parameterization of the airfoil shape is one important issue that was investigated. A simple reconstruction test showed that the analytical method using shape functions can have good accuracy while keeping reasonable shapes. However, the selection of the base airfoil is important for this method for both the GM and the GA. Trimming is an another issue for optimization especially for the GA. Trimming can simplify the design space, however, adding an external penalty type constraint is still necessary to limit the computational cost increase. Both the GA and the GM found shock-free airfoils. However, the GA results showed smaller total drag than the GM results. This suggests the GA results have a smaller profile drag than the GM results and the GA was not affected by the shock-free local minima in transonic flow. Thus, by using the GA new shapes for the lower surfaces were found. The GM stopped at different shock-free conditions when changing the starting airfoil. This suggests that the design space of transonic airfoil has probably numerous local optimum points. The GA found airfoils with similar upper surface shapes. However, the lower surfaces were different, because they are limited by the boundaries of the design space. This implies that the global optimum point is outside of the boundaries for the lower surfaces. This result suggests that it is an issue for using the GA to define a design space boundary that includes the global optimum design. Parallelization is a key issue in using the GA method in transonic airfoil optimization. Without using parallel computing, the GA requires unrealistic CPU time when combined with a high fidelity numerical solver such as a Navier-stokes solver. The nature of the population-based search with small communication requirements makes the GA suitable for an efficient parallelization and reduces the significant computation time to an affordable level.

65

CHAPTER 3 AIRFOIL OPTIMIZATION FOR MORPHING AIRCRAFT

By varying the wings airfoil shape in flight, a morphing aircraft could meet a single point aerodynamic objective at each point in its flight envelope. However, morphing aircraft will require an actuation system to morph their shapes. One of the objectives of this research is to include a measure of the energy needed for morphing as part of the design objectives. This research is under the assumption that the energy needed to change airfoil shape will be proportional to the actuation system weight. A simple strain energy model has been introduced in Ref. 39 as an objective to consider the shape variation of an aircraft wing. Prock39, et al. suggested a singleobjective optimization approach, that is setting the strain energy as a single objective while addressing the aerodynamic performance via constraints. In this research, a multi-objective optimization strategy is proposed instead of a single objective approach. The relative strain energy is used as another design objective along with a combined drag coefficient objective. Three different multi-objective optimization methods are applied to this problem. Guided by the conclusions of Chapter 2, the parallel GA is selected as an optimization algorithm for the airfoil optimization of a morphing aircraft.
3.1 Problem Description

Some recent studies 60 by the US Air Force Research Laboratory (AFRL) have focused upon a high-altitude, long-endurance aircraft platform. A notional representation of this concept appears in Figure 3.1. This aircrafts design mission includes a 40+ hour loiter segment, during which the aircraft will experience a significant weight reduction as it consumes most of its fuel. If the aircraft is intended to loiter at a constant altitude and

66 constant airspeed, a fixed geometry wing would not be operating at its most efficient conditions throughout the mission. However, if the aircraft utilized a wing with morphing airfoil sections, it would be possible to change airfoil shape throughout the mission in order to improve the endurance performance of the aircraft.

Figure 3.1 Notional high-altitude, long endurance aircraft concept Based upon systems studies from AFRL, the required lift coefficients are known at various times during the long loiter segment. To begin the energy objective investigations, the flight conditions at three points in time provide the airfoil shape design conditions. These are summarized in Table 3.1. Near the start of the loiter segment before much of the fuel is consumed, the aircrafts weight is high, and the required design lift coefficient is also high. Because of the desire for constant altitude, constant velocity loiter, the Reynolds number and Mach number for all three conditions are the same, and the lift coefficient should be reduced over time due to aircraft weight loss from fuel consumption. The low Mach numbers should not require aerodynamic analysis that includes wave drag.

Table 3.1 Airfoil design conditions


Condition 1 Condition 2 Condition 3

Design lift coefficient Mach number Reynolds number

1.52 0.6 1.5106

1.18 0.6 1.5106

0.85 0.6 1.5106

67
3.2 Objective Function Formulation

A set of optimization runs are conducted first using only aerodynamic concerns. Aerodynamics-only optimization will be presented in this research as single-point and multi-point optimization. In the aerodynamics-only design, the objective is the minimization of the aerodynamic drag with the constraint of keeping the design lift. The single-point optimization is finding three different optimum shapes for each of the three flight conditions. These three different optimum shapes are expected to have the smallest drag at each corresponding flight condition. Therefore, these three airfoil shapes are ideal shapes from an aerodynamic point of view but have no consideration of energy required for shape change. The multi-point optimization finds only one airfoil shape. Aerodynamically this single airfoil has a compromised performance for each flight condition, but this is an ideal airfoil from the relative strain energy point of view because it requires zero strain energy. In the energy-based optimization, strain energy will be added as another objective, which makes the optimization problem multi-objective. 3.2.1 Aerodynamics Only Investigation (Single-objective Approach) 3.2.1.1 Single-point Optimization With no consideration of the energy needed to change the morphing airfoils shape, the airfoil would be able to adjust so that its performance at any given flight condition would match the result of a single-point optimization at the flight condition. Three different single-point optimizations are performed for each design condition described in Table 3.1 employing the same procedure showed in Chapter 2. (1) Minimize: C d Subject to Cl1 = 1.52 , 1 solver tolerance (2) Minimize: C d Subject to Cl 2 = 1.18 , 2 solver tolerance (3) Minimize: C d Subject to Cl3 = 0.85 , 3 solver tolerance 3.2.1.2 Multi-point Optimization The multi-point approach uses the weighted sum of drag coefficients as an objective function. A single airfoil shape is acquired by this approach that requires no strain energy. (3.1)

68 However, this aerodynamics tradeoff results in higher drag at each flight condition compared to the single point optimization results. Minimize:
Subject to: 1 1 1 C d1 + C d 2 + C d3 3 3 3 Cl1 = 1.52 , 1 solver tolerance Cl 2 = 1.18 , 2 solver tolerance Cl3 = 0.85 , 3 solver tolerance 3.2.2 Energy Based Optimization (Multi-objective Approach) The strain energy term is proposed as another objective in energy-based design for morphing airfoils. This research tested two traditional multi-objective methods (i.e. weighted-sum, -constraint) and one evolutionary multi-objective method (i.e. N-branch tournament GA 61) to incorporate the strain energy into the aerodynamic objective. 3.2.2.1 Strain Energy Model If a mechanism to provide shape changes exists, the energy to actuate this mechanism should be used to compute the energy objective function. Without such a mechanism, a simple model of the strain energy associated with changing the airfoil shapes is used in this thesis. There are several ways to model the strain energy needed to change the airfoil shape. All of these ways make use of the basic idea that the strain energy in a structure is proportional to the square of the change in length of the structure. Here, a simple strain energy model has been considered. This model is described by equation (3.3) using the internal linear spring model concept suggested by Prock, et al39.
n n 1 1 EA 2 2 U = k i Li = Li i =1 2 i =1 2 Li

(3.2)

(3.3)

In this equation, U is the strain energy; ki, the spring constant, EA the spring axial stiffness, and Li the spring deformation. With no real actuation system envisioned as yet, the spring model strain energy objective does not need to include the EA terms. This model assumes that springs connect the upper and lower airfoil surfaces; as the airfoil

69 morphs, the springs deform, which corresponds to an amount of strain energy. Figure 3.2 presents a simple illustration of this model. These springs are connected at points corresponding to the ends of panels used by the aerodynamic analysis.

( L / c)i

Internal springs connecting upper and lower surface

(L / c) i
Springs contract (or expand) to meet new airfoil shape

Figure 3.2 Internal linear spring model for strain energy

3.2.3 Multi-objective Optimization When an optimization problem involves more than one objective function, the task of finding the optimum is known as multi-objective optimization. For a multi-objective optimization, the objective function is the vector-valued function as shown in Equation (3.4).
G G G G G f ( x ) = [ f1 ( x ), f 2 ( x ),..., f k ( x )]T

(3.4)

In problems with more than one objectives, there is no single optimum solution. There exist a number of solutions that are all considered optimal because tradeoffs between conflicting objectives are important. Without any further information, no solution from the set of optimal solutions can be said to be better than any other.

G The decision variables x * F is Pareto optimal if there does not exist another G G G G G x F such that f i ( x ) f i ( x * ) for all i = 1,..., k and f j ( x ) < f j ( x * ) for at least one j. This

G definition says that x * is Pareto optimal if there exists no feasible vector of decision G variables x F that would decrease some criterion without causing a simultaneous
increase in at least one other criterion. This concept almost always gives not a single

70 solution, but rather a set of solutions called the Pareto optimal set. The vectors G x * corresponding to the solutions included in the Pareto optimal set are called

nondominated. The plot of the objective functions whose nondominated vectors are in the
Pareto optimal set is called the Pareto front (or Pareto frontier). Several approaches have been proposed to find the Pareto optimal set. In this research, a weighted sum method and - constraint methods are used as a first step to explore the Pareto optimal set. The weighted sum method combines a set of objectives into a single objective by premultiplying each objective with a user-supplied weight and adding the weighted objectives. The weight of an objective is usually chosen in proportion to the objectives relative importance in the problem. This method is the simplest approach and is easy to implement. However, linear combinations of weights do not work when the Pareto front is concave, regardless of the weight used. To alleviate the difficulties faced by the weighed sum approach in solving problems having nonconvex objective spaces, the - constraint method is used. This method keeps one of the objectives and restricts the rest of the objectives within user-specified values. The advantage of this method is that it can be used for any arbitrary problem with either convex or nonconvex objective space. However, a vector of values determining the location of the Pareto-optimal solution is needed. 3.2.3.1 Weighted Sum Approach In the weighted sum approach, the strain energy term is added to the aerodynamic term with a weight factor as denoted in Equation (3.5). The number of all the relative strain energies (i.e. the change in strain energy when moving from any design shape to
N another shape) between the N different airfoils is . In the case of three different 2

design conditions, there exist three different relative energies ( U12 , U 23 , U13 ). The definition of the strain energy is the same as shown in Equation (3.3). Instead of using the sum of the three relative energies, the maximum value of the relative energies is used as

71 energy objective. This is because the actuator needs to be big enough for the largest relative change of shape. Minimize: Subject to

1 f1 + 2 f 2
Cl1 = 1.52 , 1 solver tolerance Cl 2 = 1.18 , 2 solver tolerance Cl3 = 0.85 , 3 solver tolerance
(3.5)

Where,

f1 =

[max(U 12 ,U 23 ,U 13 )] , f
C1

1 3

1 c d1 + 1 3 cd 2 + 3 cd3

C2

C1, C2 = normalization factors

3.2.3.2

-Constraint Approach

Because the weighted sum approach described in Equation (3.5) does not constrain drag coefficient values, there is no guarantee that the multi-objective approach will find an aerodynamically better shape than the multi-point solution. To address this problem, constraints on the drag coefficient are enforced via penalty function in this approach. Minimize: Subject to:
f1

f 2 i (i=1,2,.m) Cl1 = 1.52 , 1 solver tolerance Cl 2 = 1.18 , 2 solver tolerance Cl3 = 0.85 , 3 solver tolerance

(3.6)

Where,

f1

[max(U12 ,U 23 ,U 13 )] , f =
C1

1 3

1 c d1 + 1 3 cd2 + 3 cd3

C2

C1, C2 = normalization factors

i =reference drag coefficient

72 3.2.3.3 N-Branch Tournament Selection Recently, many different versions of the Genetic Algorithm (GA) have been used for multi-objective optimization. An appropriately modified genetic algorithm approach can generate a large number of designs that represent the Pareto set for a multi-objective problem with similar computational effort required to solve a single objective problem with a genetic algorithm. In this research, the N-Branch Tournament GA61 is used as a evolutionary multiobjective genetic algorithm. N-branch tournament differs from non-dominance ranking approaches such as Multi-Objective Genetic Algorithm (MOGA) 62 because it uses the selection operator to perform multi-objective design rather than rely upon the formulation of a single fitness function. In N-branch tournament selection, designs compete once on a fitness value associated with each objective. Equation(3.7) is the problem statement for this optimization.
[max(U 12 ,U 23 ,U 13 )] G C1 f = 1 1 1 3 cd1 + 3 c d 2 + 3 c d3 C2

Minimize:

(3.7)

Subject to:

Cl1 = 1.52 , 1 solver tolerance Cl 2 = 1.18 , 2 solver tolerance Cl3 = 0.85 , 3 solver tolerance

Where,
3.3 Design Variables

C1, C2 = normalization factors

The design variables are used to describe the airfoil shape as shown in Equation (3.8). For the single-point and multi-point aerodynamics-only design, 16 design variables are needed as explained in Chapter 2. In the case of energy-based design, a total of 48 design variables (16 for each of the three flight conditions) are needed to describe shapes for each design condition. The NACA0012 airfoil is selected as a base airfoil for all shapes. The same modified Hicks-Henne shape functions and design variable bounds presented

73 in Chapter 2 are used here. Thus, the same limits on upper and lower surface shapes are provided.
y ( x ) k = y ( x ) k (Base Airfoil) + ki f ki ( x )

(3.8)

( ki : Design Variables, f ki : Shape Functions) 3.3.1 Optimization Algorithm The parallel genetic algorithm is used as an optimization algorithm as described previously in Chapter 2. However, for the multiobjective design with energy as an objective, the population size and other GA parameters are adjusted as follows (based on the criteria of Reference 58), because the number of variables is increased to 48.

Number of design variables = 48 Bits per design variable Chromosome length Population size Mutation probability Design variable limits = 7 bit = 336 = 1344 = 0.00037 = 0.015 i 0.015

3.3.2 Flow Solver In this stage of the research, a quick turn-around computational time is desired for the flow evaluation. Thus, the well-known XFOIL 63 code is selected as a function evaluator. XFOIL uses a linear-vorticity panel method with a Karman-Tsien compressibility correction for inviscid calculations. To model viscous layer influence, source distributions are superimposed on the airfoil and wake. The boundary layer and the transition equations are solved simultaneously with the inviscid flow field by a global Newton method. At high angles of attack, XFOIL has difficulties obtaining a converged solution as stall is approached. To prevent this problem, non converged solutions are given a high penalty by setting the fitness function arbitrarily large via a penalty function and are eventually eliminated during the design process. The solver tolerance value that used for this penalty calculation is 0.002.

74
3.4 Optimization Results

3.4.1 Aerodynamic-only Optimization Results Three single-point optimizations were performed using Equation 3.1, one for each of the three flight conditions, as described in Table 3.1. The drag coefficients of the designed airfoils are compared in Table 3.2. A multi-point design was also done using equation 3.2, and the results also appear in Table 3.2. For these designs, relative strain energy values are computed even though energy was not used in the problem. Because there is no convergence criterion for the GA, in Chapter 2, a fixed number of generations is used as a stopping criterion. However, in this chapter, 95% BSA(Bit-String Affinity) stopping criteria 64 is applied for all aerodynamic-only designs. Normally, obtaining 95% BSA for this problem requires about 300 GA generations. The single-point design results show lower drag than the multi-point design results, as expected. This is because the single-point designs have only one objective which is to minimize drag at one flight condition, but the multi-point design solution is a compromise solution over all three flight conditions. The multi-point solution is actually a minimum energy solution, because the result of the multi-point design is a single airfoil. Thus, the relative strain energies in Table 3.2 are all zero for the multi-point design. The relative energies for the single-point solutions shown in Table 3.2 are calculated following the energy model described in Equation (3.3). As a result, Table 3.2 indicates the range of tradeoff available for a morphing airfoil. The three single-point shapes are the best possible aerodynamic solution, while the multi point shapes is the best possible energy solution. Table 3.2 Drag comparison of aerodynamics only design
Cd1 single-point multi-point 0.005024 0.007356 Cd2 0.007015 0.008178 Cd3 0.010249 0.010649 U12 0.00704 0 U23 0.00189 0 U13 0.01241 0

The three airfoils of the single-point optimization are compared in Figure 3.3. These airfoils have different shapes; the camber of the airfoil is changed to meet the varying lift coefficient constraints. Figure 3.4 shows the airfoil from the multi-point optimization.

75

0.25 0.2 0.15 0.1 0.05 0 -0.05 -0.1 -0.15 -0.2 -0.25 0 0.25 0.5 X
Design Cl=0.85 Design Cl=1.18 Design Cl=1.52

0.75

Figure 3.3 Best airfoil shape (single-point optimization)

0.25 0.2 0.15 0.1 0.05 0 -0.05 -0.1 -0.15 -0.2


Multi-Point Optimization

-0.25

0.25

0.5 X

0.75

Figure 3.4 Best airfoil shape (multi-point optimization)

76 3.4.2 Energy Based Optimization Results Three different multi-objective optimization methods (weighted sum, -constraint and N branch tournament GA) were applied to acquire the Pareto front, and the results are compared. 3.4.2.1 Weighted Sum Approach Result The weighted sum approach was performed by changing the weighting factors ( 1 , 2 ) in Equation (3.5) and conducting a GA run for each combination 1 and 2 . The results for these runs are shown in Table 3.3 along with values use for 1 and 2 . The single-point and the multi-point results are also included for comparison. The two objectives in Equation (3.5) are plotted in Figure 3.5 (where C1 =0.01,
C 2 =0.01). Most of the drag objective values are higher than the multi-point designs. This

suggests that when the weighting factor 1 is higher, the weighted sum approach described in Equation (3.5) does not sufficiently reduce the drag coefficient objective values.

Table 3.3 Drag and Relative Strain Energy comparison (weighted sum method)
1
Multi-point case1 case2 case3 case4 case5 case6 case7 case8 Single-point 0.9 0.8 0.7 0.5 0.3 0.2 0.1 0.05 0.1 0.2 0.3 0.5 0.7 0.8 0.9 0.95

f1
0.000 0.010 0.011 0.015 0.021 0.037 0.069 0.151 0.320 1.240

f2
0.864 1.121 0.987 0.971 0.918 0.893 0.862 0.839 0.840 0.735

Cd1
0.00735 0.00903 0.00832 0.00818 0.00780 0.00753 0.00737 0.00731 0.00675 0.00502

Cd2
0.00817 0.01121 0.00992 0.00996 0.00881 0.00876 0.00837 0.00828 0.00853 0.00701

Cd3
0.01064 0.01457 0.01241 0.01200 0.01191 0.01144 0.01104 0.01047 0.01080 0.01025

U12
0 0.00009 0.00002 0.00013 0.00010 0.00024 0.00016 0.00151 0.00282 0.00704

U23
0 0.00013 0.00014 0.00018 0.00018 0.00043 0.00082 0.00187 0.00397 0.00189

U13
0 0.00012 0.00013 0.00018 0.00026 0.00046 0.00086 0.00164 0.00168 0.01241

77

1.2

1.1 f2 (Drag Objective)

1.0

energy based multi-point single-point

0.9

0.8

0.7 -0.2

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

f 1 (Energy Objective)

Figure 3.5 Weighted sum approach Pareto front

3.4.2.2

-constraint Approach Results

The results of previous section showed that weighted sum approach is not effective in studying the tradeoff between aerodynamic-only design and energy-related design, because the weighted sum approach cannot easily constrain drag. To overcome this problem, the -constraint approach was tried. Here, the drag objective is constrained by values found between the single-point and the multi-point design in this approach. Thus, the i values in Equation (3.6) are defined by the values within the limit of multi-point and single-point Cd value. Seven test cases were presented in Table 3.4. For all cases, the drag coefficients are lower than the multi-point, while keeping the strain energies lower than single-point design. Figure 3.6 shows the Pareto set found from the -constraint approach. The singlepoint and multi-point results of the aerodynamic-only designs are also plotted for comparison. In Figure 3.6 the x-axis values represents the energy objective f1 and the yaxis values are the drag objective f2 shown in Equation (3.6) (where, C1 =0.01, C2 =0.01).

78 Figure 3.6 shows that a tradeoff does exist between the multi-point design, which requires no strain energy, and the single point design, which requires a high strain energy for the lowest drag.

Table 3.4 Drag and Relative Strain Energy comparison (-constraint method)

Multi-Point case1 case2 case3 case4 case5 case6 case7 Single-Point


0.856 0.840 0.824 0.807 0.791 0.775 0.759

f1
0.000 0.036 0.095 0.147 0.173 0.225 0.261 0.461 1.241

f2
0.864 0.845 0.831 0.814 0.799 0.782 0.765 0.751 0.736

Cd1
0.00736 0.00705 0.00677 0.00646 0.00619 0.00588 0.00560 0.00532 0.00502

Cd2
0.00818 0.00799 0.00789 0.00771 0.00759 0.00743 0.00725 0.00715 0.00702

Cd3
0.01065 0.01058 0.01054 0.01050 0.01045 0.01040 0.01034 0.01030 0.01025

U12
0.00000 0.00013 0.00054 0.00022 0.00072 0.00072 0.00091 0.00226 0.00704

U23
0.00000 0.00036 0.00092 0.00119 0.00173 0.00220 0.00221 0.00226 0.00189

U13
0.00000 0.00036 0.00095 0.00147 0.00172 0.00225 0.00261 0.00461 0.01241

0.90

0.85 f 2 (D rag O b jec tive)

energy based 0.80 multi-point single-point

0.75

0.70 -0.2

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

f 1 (Energy Objective)

Figure 3.6 -Constraint approach Pareto front

79 The airfoil set in Figure 3.7 is one of the tradeoff solutions (f1=0.26, f2=0.765) selected from the Pareto-set in Figure 3.6. This airfoil set has a better aerodynamic performance than the multi-point design and also has a smaller strain energy than the set of single-point designs. The morphing airfoil in Figure 3.7 has a drag objective of 0.765, which is lower than the muli-point drag objective of 0.864. Also, the morphing airfoil in Figure 3.7 has an energy objective of 0.260, which is also lower than the set of singlepoint shapes of 1.240. This airfoil set shows little difference in the lower surface to reduce energy, but has more variation in the upper surface to reduce drag and maintain required Cl values.

0.25 0.2 0.15 0.1 0.05 0 -0.05 -0.1 -0.15 -0.2 -0.25 0 0.25 0.5 X
Design Cl=0.85 Design Cl=1.18 Design Cl=1.52

0.75

Figure 3.7 Airfoil set found by multi-objective optimization including energy

80

3.4.2.3 N-Branch Tournament Genetic Algorithm Figure 3.8 shows the Pareto set found from one run of the N-branch tournament GA. With less effort, many more tradeoff solutions were acquired from the N-branch tournament GA as compared to -constrained method. One run of the N-branch for 1000 generations found 163 Pareto optimal designs; this required 1,344,000 function evaluations. In contrast, the -constraint approach found only 7 Pareto optimal designs and required a total of 2,822,400 function evaluations. Figure 3.9 is one of the tradeoff solutions selected from the Pareto set shown in Figure 3.8. The morphing airfoil in Figure 3.9 has a drag objective of 0.765 and energy objective 0.2. The morphing airfoil shapes in Figure 3.9 and Figure 3.7 seem similar, but the energy objective of the airfoil shapes in Figure 3.9 is slightly less. The airfoil shapes in Figure 3.9 are much closer to each other than the single-point generated airfoil shapes in Figure 3.3, illustrating why the strain energy of the morphing airfoil in Figure 3.9 is smaller.
0.90

0.85 f2 (Drag Objective)

N-Branch 0.80 multi-point single-point

0.75

0.70 -0.2

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

f 1 (Energy Objective)

Figure 3.8 Pareto set from N-Branch Tournament GA.

81

0.25 0.2 0.15 0.1 0.05 0 -0.05 -0.1 -0.15 -0.2 -0.25 0 0.25 0.5 X
Design Cl=0.85 Design Cl=1.18 Design Cl=1.52

0.75

Figure 3.9 Airfoil shapes from energy based design (N-Branch tournament GA). Figure 3.10 presents the evolution of the Pareto-front as the generation number of the
N-branch GA increases. After generation, 800 the Pareto front does not change much.

From this Figure, a solution is assumed to be converged after 1,000 iterations.


0.95

0.90 N-branch [Generation 400] f2 (Drag Objective) N-branch [Generation 500] 0.85 N-branch [Generation 600] N-branch [Generation 700] N-branch [Generation 800] N-branch [Generation 900] 0.80 N-branch [Generation 1000] multi-point single-point 0.75

0.70 -0.20

0.30

0.80

1.30

f 1 (Energy Objective)

Figure 3.10 Convergence history of Pareto set from N-Branch Tournament GA.

82 3.4.3 Cp and Energy Comparison The C p distribution and airfoil shapes are compared in Figures 3.10-3.12. The selected airfoils for multi-objective methods are the same airfoils shown in Figures 3.7 and 3.9. Each of these figures compares the Cp and shape at one of the three design flight conditions. It is obvious from the figures that the shapes generated using energy as an objective compromise between energy and drag when compared with the single-point shapes and the multi-point design. One noticeable trend from the airfoil comparison is that, as the design Cl increases, the difference of the shapes decreases. This might be due to the fact that as the design Cl increases, it is more difficult to find airfoils that satisfy the high Cl with small drag. The XFOIL solver has some problems in finding converged solutions at high angles of attack that may include separated flow, so feasible designs may not exist for these high Cl at lower Cd values.

83

-3 -2.5 -2 -1.5 -1 single-point multi-point -constraint N-branch

Cp
-0.5 0 0.5 1 1.5
Design Cl=0.85 M=0.6 Re=1.5E6

0.25

0.5

0.75

0.25 0.2 0.15 0.1 0.05 single-point multi-point -constraint N-branch

-0.05 -0.1 -0.15 -0.2 -0.25 0 0.25 0.5


Design Cl=0.85 M=0.6 Re=1.5E6

0.75

Figure 3.11 Cp distribution comparison (top), Airfoil comparison (bottom) (Design Cl =0.85)

84

-3 -2.5 -2 -1.5 -1 single-point multi-point -constraint N-branch

Cp
-0.5 0 0.5 1 1.5
Design Cl=1.18 M=0.6 Re=1.5E6

0.25

0.5

0.75

0.25 0.2 0.15 0.1 0.05 single-point multi-point -constraint N-branch

-0.05 -0.1 -0.15 -0.2 -0.25 0 0.25 0.5


Design Cl=1.18 M=0.6 Re=1.5E6

0.75

Figure 3.12 Cp distribution comparison (top), Airfoil comparison (bottom) (Design Cl =1.18)

85

-3 -2.5 -2 -1.5 -1 single-point multi-point -constraint N-branch

Cp
-0.5 0 0.5 1 1.5
Design Cl=1.52 M=0.6 Re=1.5E6

0.25

0.5

0.75

0.25 0.2 0.15 0.1 0.05 single-point multi-point -constraint N-branch

-0.05 -0.1 -0.15 -0.2 -0.25 0 0.25 0.5


Design Cl=1.52 M=0.6 Re=1.5E6

0.75

Figure 3.13 Cp distribution comparison (top), Airfoil comparison (bottom) (Design Cl =1.52)

86 To see the strain energy reductions more clearly, the strain energy (see Equation 3.3) distribution along the airfoil is plotted using a NACA0012 shape for reference. For example, if the control point associated with a location on the lower surface trailing edge has a large displacement between two airfoil shapes, a large bar is placed at the corresponding control point on the NACA 0012 airfoil section. In this manner, a visual representation can be made to show which section of the airfoil is associated with the highest strain energy. The single point and energy-based designs are compared. Figures 3.13 through 3.15 show clearly that the strain energy (U23,U13) in the multi-objective designs has been reduced significantly from that of the single point shapes.

x 10 7 6 5
Energy

singlepoint constraint Nbranch

4 3 2

1 0.8

1 0.6 0 0.1 0.4 0 0.1 0 y 0.2 0.1 0 0.1 0 X

Figure 3.14 Strain energy distributions (U12)

87

x 10 7 6

singlepoint 5
Energy

constraint Nbranch

4 3 2

1 0.8

1 0.6 0 0.1 0.4 0 0.1 0 y 0.2 0.1 0 0.1 0 X

Figure 3.15 Strain energy distributions (U23)

x 10 7 6

singlepoint 5 constraint
Energy

4 Nbranch 3 2 0.8 1 0.6 1

0 0.1

0.4 0 0.1 0 y 0.2 0.1 0 0.1 0 X

Figure 3.16 Strain energy distributions (U13)

88
3.5 Transonic Morphing Airfoil

As another application of the morphing airfoil design strategy described in the previous sections, a transonic morphing airfoil design is performed. The main difference from the sensorcraft problem is the flight speed regime of the morphing aircraft. For this application, the speed regime varies from subsonic to transonic, and the altitude changes from low to high. This application requires a Navier-Stokes code for flow evaluation to capture the physics of the shock wave in the transonic regime. The importance of the energy-based optimization is increased here, because a greater shape change is expected for the optimal shapes compared to the low speed sensorcraft problem. 3.5.1 Problem Definition The flight conditions considered in this research reflect a notional transonic morphing UAV(Unmanned Aerial Vehicle). This UAV has a multi mission capability which includes features of a high altitude reconnaissance UAV and a combat UAV.
Transonic Morphing UAV mission statement

(1) High altitude reconnaissance capability (2) High speed dash/cruise from point to point The selected design conditions are shown in Table 3.5. The loiter mission is defined based on the Global Hawk UAV 65. In addition, this problem assumes that the aircraft can reduce its wing area for the dash segment. Table 3.5 Transonic Morphing UAV Mission Profile
Mach Altitude (ft) Dash Loiter 0.8 0.6 5000 60,000 0.20 1.00 1.98E+07 1.77E+06 25% 100% CL Re Wing Area

3.5.2 Objective Function To reduce the computational time while keeping the aspects of morphing airfoil strategy, only two design conditions were selected for this study. The selected mission

89 segments are the dash and loiter missions (see Table 3.5). These missions are expected to require large shape change for optimal aerodynamic design. The objective functions are described in Equation (3.9). Only one relative strain energy is needed for the two design conditions.
U 12 G f = 1 1 C d1 + C d 2 2 2 Cl1 = 0.2 , M=0.8, Re=1.98E7 Cl 2 = 1.0 , M=0.6, Re=1.77E6 3.5.3 Flow Solver, Design method and Parameters The TURNS47 code is the flow solver for this problem. The same grid points described in Chapter 2 are also applied in this problem. However, 1500 iterations are used for each solution here instead of 2000 iterations, to reduce the computational cost without paying a great penalty for accuracy. The same parallel GA described in Chapter 2 is used for this problem and the number of design variables is 16 for each airfoil shape (for a total of 32 variables). The RAE2822 airfoil is used as base airfoil. The resolution of the GA is set to 5 bits per design variable for this problem. The GA parameters for population size and mutation rate are selected following the guidelines of Ref.58. 3.5.4 Aerodynamics-only Result As before aerodynamics-only optimization were conducted to identify the extremes of the tradeoff between the drag objective and the energy objective. For this transonic morphing airfoil problem, the best solutions after 100 GA generations are used as converged solutions. For a single point optimization, this requires 76,800 function evaluations and 67 hours of computational wall-time with 30 CPUs. The single point optimization results are shown in Figure 3.16. As expected, the aerodynamically best airfoil shapes show a large difference in configuration compared to the sensor craft problem. Figure 3.17 presents the results for the multi-point design case.

minimize :
Subject to:

(3.9)

90

0.25 0.2 0.15 0.1 0.05 Design condition 1 Design condition 2

-0.05 -0.1 -0.15


Single Point Optimization

-0.2 -0.25 0 0.25

Design condition 1 : M=0.6, Cl=1.0, Re=7.9E6 Design condition 2 : M=0.8, Cl=0.2, Re=2.0E7

0.5

0.75

Figure 3.17 Best airfoil shape (Single-point optimization)

0.25 0.2 0.15 0.1 0.05

-0.05 -0.1 -0.15


Multi Point Optimization

-0.2 -0.25 0 0.25

Design condition 1 : M=0.6, Cl=1.0, Re=7.9E6 Design condition 2 : M=0.8, Cl=0.2, Re=2.0E7

0.5

0.75

Figure 3.18 Best airfoil shape (Multi-point optimization)

91 3.5.5 Energy-based Design Results For this problem, only the N-branch tournament GA is used to calculate the Pareto-set. Because computational time is high for this problem, the classical multi-objective methods (weighted sum or -constraint) are not affordable with the available computing power. The Pareto-set found by the N-branch tournament GA is shown in Figure 3.18.

1.7

1.6

1.5 f2 (Drag Objective)

1.4

Single-point Multi-point N-branch

1.3

1.2

1.1

1 -0.5 0 0.5 1 1.5 2 2.5 3 f 1 (Energy Objective)

Figure 3.19 Pareto-front of transonic morphing wing case

Figure 3.20 is a rescaled picture of Figure 3.19 to check the convergence of the Nbranch GA. It can be concluded that the Pareto set after 300 generations is quite close to a converged set because the difference of the Pareto set between the generation 230 and generation 300 is very small.

92

1.5

Single-point 1.4 f2 (Drag Objective) Multi-point N-branch GA (Generation 300) N-branch GA (Generation 230) N-branch GA (Generation 124) N-branch GA (Generation 100)

1.3

1.2

1.1 0 0.1 0.2 0.3 0.4 0.5 0.6 f 1 (Energy Objective)

Figure 3.20 Pareto-front of transonic morphing wing case (Rescaled picture)

One Pareto front solution point (f1=0.199, f2=1.202 from Figure 3.19) is selected and the airfoils are drawn in Figure 3.21. The shapes are very similar to each other, compared with the shapes found by the single point aerodynamic-only design (Figure 3.16). The shape difference of the solution airfoil sets is relatively greater than the subsonic airfoil solution sets of the sensorcraft application. This indicates that the energy-based design is more important for the transonic application than the subsonic case and suggests that this approach can identify shapes that will save actuation energy. All the airfoil shapes of the Pareto front are drawn in Figures 3.22 and 3.23. The best aerodynamic shape that has the lowest f2 value is drawn as a solid line. Figures 3.22 and 3.23 present the airfoil thickness variations when the design point of the Pareto front changes. There might be a tradeoff between the thickness of the airfoil and aerodynamic performance, because the best aerodynamic shapes are thinner than other airfoil shapes.

93

0.25 0.2 0.15 0.1 0.05 Design condition 1 Design condition 2

-0.05 -0.1 -0.15


Energy Based Optimization

-0.2 -0.25 0 0.25

Design condition 1 : M=0.6, Cl=1.0, Re=7.9E6 Design condition 2 : M=0.8, Cl=0.2, Re=2.0E7

0.5

0.75

Figure 3.21 Airfoil shapes from energy based design (N-Branch tournament GA)

0.1 Base airfoil Best aerodynamics Pareto front

0.08

0.06

0.04

0.02

0.02

0.04

0.06

0.08

0.1

0.1

0.2

0.3

0.4

0.5 X

0.6

0.7

0.8

0.9

Figure 3.22 Airfoil shapes of Pareto front (Design condition 1)

94

0.1 Base airfoil Best aerodynamics Pareto front

0.08

0.06

0.04

0.02

0.02

0.04

0.06

0.08

0.1

0.1

0.2

0.3

0.4

0.5 X

0.6

0.7

0.8

0.9

Figure 3.23 Airfoil shapes of Pareto front (Design condition 2)

Figures 3.24-3.25 compare the Cp distribution for shapes in Figure 3.21, the singlepoint set of shapes and the multi-point shape. The single-point solutions have smaller shocks near the leading edge of the airfoil compared to the other two solutions, which results in smaller wave drag. It is noticeable that the thickness of the single point design airfoils is much smaller than the other two airfoils. This agrees with the results of the transonic airfoil optimization study, shown in Chapter 2, that the profile drag is small when the airfoil thickness reduces. The thickness of the N-branch solution airfoil is similar to the multi-point design; however, the aerodynamic performance is better than the multi-point design. This is because the N-branch solution is a compromised solution between the aerodynamic performance and the shape variation. The thicker airfoils requires less shape change, but have higher drag. The N-branch solution also shows that changing the camber of a thin airfoil requires more strain energy compared to a thicker airfoil with the same camber variation.

95

-2

Single-point Multi-point N-branch

-1.5

-1

Cp

-0.5

0.5

1 0 0.2

Design Conditgion : M=0.6, Cl=1.0, Re=7.9E6

0.4

0.6

0.8

0.25 0.2 0.15 0.1 0.05 single-point multi-point N-branch

-0.05 -0.1 -0.15 -0.2


Design condition : M=0.6, Cl=1.0, Re=7.9E6

-0.25

0.25

0.5

0.75

Figure 3.24 Cp distribution comparison (top), Airfoil shape comparison (bottom) (Design condition: M=0.6, Cl=1.0, Re=7.9E6)

96

-2

Single-point Multi-point N-branch

-1.5

-1

Cp

-0.5

0.5

1 0 0.2

Design Conditgion : M=0.8, Cl=0.2, Re=2.0E7

0.4

0.6

0.8

0.25 0.2 0.15 0.1 0.05 single-point multi-point N-branch

-0.05 -0.1 -0.15 -0.2


Design condition : M=0.8, Cl=0.2, Re=2.0E7

-0.25

0.25

0.5

0.75

Figure 3.25 Cp distribution comparison (top), Airfoil shape comparison (bottom) (Design condition: M=0.8, Cl=0.2, Re=2.0E7)

97 Figure 3.19 shows that the converged Pareto set does not have low drag, high energy solutions when drag is lower than 1.19 and energy is higher than 0.03. To investigate this Pareto set further, the -constraint method is applied to attempt to find a low drag, high energy solution for the above region. The objective function for this test is Minimize : Subject to :
U 12

(3.10)

Cd1 0.018349 , Cl1 = 0.2 , M=0.8, Re=1.98E7 C d 2 0.004297 , Cl 2 = 1.0 , M=0.6, Re=1.77E6

The target drag coefficients are defined for a point in the above region based on the values of single- and multi-point designs as described in Table 3.6. Table 3.6 -constraint target values for Cd constraints
Design Condition1 single Target ( i ) multi 0.018164 0.018349 0.019639 Design Condition 2 0.003651 0.004297 0.008813 Drag Objective 1.090835 1.132313 1.422659

Figure 3.26 shows that the best fitness function values do not change after about 40 GA generations. From this graph, it is concluded that the genetic algorithm reached a converged solution. However, the converged solution is much larger than the expected value in Figure 3.19. Figure 3.27 plots the drag coefficient of the best solutions for each GA generations. The drag coefficient values at design condition 1 satisfy the target drag coefficient (see Table 3.6); however, the drag coefficient values at design condition 2 greatly exceed the target drag coefficient. This suggests that these results are infeasible given the traditional -constraint definition. This infeasible solution results in large objective value by the addition of penalties. This finding is consistent with the absence of solutions in the low drag, high energy region (as shown in Figure 3.19). This may indicate that the transonic morphing airfoil is more sensitive to the geometry change. The geometric resolution of the transonic morphing case may need to

98 be increased (e.g. finer resolution between discretized design variable values) to find feasible solutions with low drag and low energy. This increase of geometric resolution will require a larger computational effort, because the chromosome length and population size would also need to increase significantly.

-200

-250
Fitness function value

-300

-350

-400

-450 1 11 21 31 41 51 61 71 81 91
Generations

Figure 3.26 GA convergence history of -constraint method

0.03

0.025

Drag coefficient

0.02 Cd 1 (M=0.6,Cl=1.0,Re=1.7E6) Cd 2 (M=0.8,Cl=0.2,Re=2.0E7)

0.015

0.01

0.005

0 1 11 21 31 41 51 61 71 81 91

Generation

Figure 3.27 Drag coefficient values

99
3.6 Summary

The multi-objective optimization approach is applied for the morphing airfoil design problem instead of the single-objective approach. The two objectives are minimization of drag and morphing energy. Three different multi-objective optimization methods have been chosen and the results are compared. To calculate the energy objective for multi-point design conditions, the maximum value is used instead of a weighted summation of the relative strain energy values. This is because the actuator needs to be big enough to overcome the maximum actuation load. The sensorcraft problem showed that minimizing the drag objective is more difficult compared to minimizing the strain energy objective. Because of this, the weighted sum approach is not appropriate for this problem, but the -constraint method and N-branch tournament method do work for the sensorcraft problem. The N-branch tournament method result showed that this method is very efficient in finding tradeoff solutions compare to the -constraint method, because it could acquire more Pareto-front solutions with one single run of GA. As a second application of the morphing airfoil design strategy, transonic morphing airfoils are designed. For the transonic example, a difference of the optimal shape for the design conditions is greater than the subsonic case. This increases the importance of the energy-based design strategy for transonic morphing airfoil design. There were no low drag and low energy designs found by N-branch GA for transonic morphing airfoil and this appears to be the result of low geometric resolution of the possible airfoil sets.

100

CHAPTER 4 ACTUATION ENERGY MODELING INCLUDING AERODYNAMIC WORK

4.1 Description of Concept

The morphing airfoil design strategy described in the previous chapter does not account for the effect of the external air load on the airfoil shape. If the structure of the wing is very stiff so that deformation caused by the external aerodynamic load can be ignored, then this assumption is quite appropriate. However, for a morphing aircraft, the aerodynamic load should not be ignored, because the shape change requires a flexible wing. The pressure distribution acting on an airfoil can be represented by a series of forces acting at specific control points on the airfoil. The direction and magnitude of these forces depend on both the flight condition and the local shape of the airfoil. Figure 4.1 displays two airfoils, each designed to minimize drag at a different lift coefficient. The aerodynamic forces acting on the airfoil are presented as vectors, and the two plots show that if the airfoil shape and flight conditions change, the aerodynamic force distribution also changes. Because the goal of the morphing airfoil is to change the shape of the airfoil, there is a possibility to acquire some assist from the aerodynamic force, or the possibility that the actuation effort needs to increase to overcome the aerodynamic force and the structural stiffness. Figure 4.2 illustrates how the variation in airfoil shape change impacts the direction and magnitude of the aerodynamic forces. This, in turn, can affect the required effort from the morphing actuators. In Figure 4.2, two exaggerated airfoil shapes are shown; the black dashed shape represents the initial shape and the blue solid shape represents the

101 next desired shape. Two points of interest are shown, P1 and P2. At P1, are of the control point describing the airfoil shape is to move from position a to position b. The direction G of this shape change is indicated by the vector ri . In the original shape, the aerodynamic force acting at this point is indicated by the dashed arrow applied at location a; this aerodynamic force has a component in the direction of the shape change (indicated by the red vector applied at a). This would suggest that the aerodynamic force could assist with the airfoil shape change. In contrast, the control point at P2 needs to move from location a to location b, which G means the move vector, ri , is opposite in direction to the component of the aerodynamic force at this point. This would indicate a possible resistance to the airfoil shape change, so the actuator would need to overcome the stiffness of the airfoil structure as well as the aerodynamic load.

0.6 0.5 0.4 0.3 0.2 0.1

0.6 0.5 0.4 0.3 0.2 0.1

0 -0.1 -0.2 -0.3 -0.4 -0.5 -0.2 -0.1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 X 1 1.1

Y
0 -0.1 -0.2 -0.3 -0.4 -0.5 -0.2 -0.1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 X 1 1.1

(a) Design Cl=0.85

(b) Design Cl=1.52

Figure 4.1 Aerodynamic force distributions on the airfoil surface

102

P2
b

P1
a

ri
a

ri

Figure 4.2 The illustration of the aerodynamic work

103

4.2 Formulation

To include the aerodynamic force in morphing, a simple spring model is employed to represent the structure of the airfoil as shown in Figure 4.3. Deforming the airfoil shape is modeled by a deformation of the linear spring. At each state in Figure 4.3, the force equilibrium is maintained. For clarity, the spring at only one control point will be presented. When the aerodynamic load is not included, the work done by the actuator to move the airfoil from State-1 to State-2 in Figure 4.3(a) can be described by Equation (4.1). Because the force and deformation have a linear relation for a linear elastic spring, the actuation work required to change the shape is given by Equation (4.2).
U 1 ( strain energy) + W12 ( actuator ) = U 2 (strain energy)
2 G 2 G 1 W12 ( actuator ) = U 2 U 1 = f a dx = kxdx = k L2 2 1 1

(4.1) (4.2)

On the other hand, when the aerodynamic force also acts on the airfoil structure as in Figure 4.3 (b), the aerodynamic work term needs to be included in the energy/work equation.
U 1 ( strain energy) + (W12 ( actuator ) + W12 ( aerodynamic ) ) = U 2 (strain energy)

(4.3)

Thus, the corresponding actuation work becomes W12 ( actuator ) = (U 2 U 1 ) W12 ( aerodynamic ) (4.4)

G If we assume that the aerodynamic force, f a , acting on the spring varies linearly from

State-1 to State-2 as the airfoil changes from shape 1 to shape 2 (see Figure 4.4), then the work done by the aerodynamic force can be given by Equation (4.5). The integral of the aerodynamic force is replaced by the average aerodynamic force.
G G G G G G 2 f + f G G G G G f1 + f 2 G 1 2 r dx = r L = 1 ( f1 r + f 2 r )L = f a dx = 2 2 2 1 1
2

W12( aerodynamic )

(4.5)

104

( L / c)i

Internal springs connecting upper and lower surface

(L / c) i
Springs contract (or expand) to meet new airfoil shape

G f1

G fa

1
G fs
L

G G f2 fa

G fs
k k k k

G fs

State-1

State-2

(a) no aerodynamic load


G f a = actuator force G f b = internal force in spring G f1 = aerodynamic force in shape 1 G f 2 = aerodynamic force in shape 2

State-1 State-2 (b) with aerodynamic load

k = stiffness of spring (EA)

1 = deformation of structure under aerodynamic load in shape 1 L= change of airfoil shape


Figure 4.3 Simple spring airfoil structure model

105

fa

f1

f2

L1

L2

Xspring

Figure 4.4 Linear aerodynamic force variation

Substituting Equation (4.2) and Equation (4.5) into Equation (4.4), provides following Equation (4.6) that is the actuation work at one single control point. W12 ( actuator ) = 1 1 G G G G kL2 ( f1 r + f 2 r ) L 2 2 (4.6)

In Equation 4.6, the aerodynamic term is subtracted from the strain energy term to calculate the actuation work term. This means that the aerodynamic work term should be increased to reduce the actuation energy.

106
4.3 Sensorcraft Problem

To demonstrate the aerodynamic work concept, the same design conditions of the sensorcraft problem in Chapter 3 are used here. 4.3.1 Problem Definition Because the actual airfoil structure and morphing mechanism are replace with linear springs, the spring constant k in Equation (4.6) is unknown and needs to be assumed. This assumption governs the stiffness of the airfoil. As a starting point, it is assumed that the spring is deformed about 0.001c (1/1000 of the chord length) by the aerodynamic force at State-1. This would represent a very minimal deformation of the airfoil under the aerodynamic load.

G G f1 r1 = k1
where, 1 = 0.001c = 0.001 , G G k c = 1000 * f1 r1

(4.7)

To define the reference value for k, two airfoil shapes designed for two different flight conditions are selected (condition 1 and condition 3 in Table 3.1). From the flow solver (XFOIL), Cp and Cf values can be obtained at each control point (See Figure 4.5). The aerodynamic force at each control point can be calculated by the following Equation.

G G G f1 = q[(C p1 * ds1 )n1 + (C f1 * ds1 )t1 ] G Where, t : unit tangential vector of the panel G n : unit normal vector of the panel
Control Point

(4.8)

Panel length: ds

G n

K t

Figure 4.5 Panel distribution on the airfoil

107 Figure 4.6 shows the normal component of the aerodynamic force calculated from Cp acting on each control point of the two airfoils. The tangential component of the aerodynamic force in Equation 4.8 is comparatively very small, so it is not drawn in Figure 4.6.

0.6 0.5 0.4 0.3 0.2 0.1 Y 0 -0.1 -0.2 -0.3 -0.4 -0.5 -0.2 -0.1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 X 1 1.1

G f2
K f1

K G f2 r
G G Gr f1 r

Figure 4.6 Normal component of aerodynamic force acting on the surface of the airfoil at state-1(Cl=0.85) and state-2 (Cl=1.52) 4.3.2 Stiffness Approximation Figure 4.7 shows the aerodynamic force terms calculated using Equation 4.8, at each control point of the selected airfoils from Figure 4.6. A total of 140 control points are

108 presented, and they are counted in counter clockwise direction. Thus, 0 is for the upper trailing edge control point, 70 for the leading edge and 140 for the lower surface trailing edge and so on.

G G G G From the values of f1 r q and f 2 r q in Figure 4.7, the value 0.03, which is near G G the largest value at any control point, is selected for f1 r q to assume the stiffness
value. Because the actuator needs to be big enough to overcome the largest value, the largest possible value is selected as a typical value.

G G f1 r1 = 0.03 q
Aerodynamic force distributioin 0.04 f1e/q f2e/q (f1e+f2e)/2q 0.03

(4.9)

0.02

0.01

control point counting


-0.01

-0.02

20

40

60 80 control point

100

120

140

Figure 4.7 Aerodynamic force distribution at each control point

From Equations 4.7 and 4.9, the stiffness can be assumed as Equation (4.10). This is based on the assumption that every spring is displa2ced 0.001c by the aerodynamic forces associated with the state-1 airfoil shape. k=30 q (4.10)

109 Then, the substitution of Equation (4.10) into Equation (4.6) results in the following
G G G G ( f1 r + f 2 r ) q 2 = 30L L 2 q

W12 ( actuator )

(4.11)

This is the work the actuator must do to move the airfoil control point from shape 1 to shape 2. If this is negative, this indicates the airloads could move the airfoil shape. In the sensorcraft problem, where the speed and altitude are same for both design conditions, Equation (4.11) can be rearranged as follows:

K G K G W12 ( actuator ) = Const. 30L2 ( f1 r / q + f 2 r / q )L

(4.12)

Where,

G G G f1 / q = [(C p1 * ds1 )n1 + (C f1 * ds1 )t1 ] G G G f 2 / q = [(C p2 * ds2 )n2 + (C f 2 * ds2 )t 2 ]

Because the value of q is not changed by the geometry variation, but by the flight condition, q can be considered as constant value in Equation 4.12. For the total actuator work needed to morph the airfoil shape, Equation 4.12 is used for each spring modeled in the airfoil, and these terms are added together.

4.4 Results and Comparison

4.4.1 Effect of Aerodynamic Work Term The updated actuator work expression, Equation (4.12), is applied to the sensorcraft problem. Objectives are to minimize the actuation energy and the drag at both flight conditions when the design lift changes from high to low. For the sensorcraft problem, the weight of the aircraft reduces as it consumes fuel. Thus, to maintain the same altitude, the lift should be reduced. Three airfoils are designed for three different flight conditions as shown in Table 3.1. The objective functions are shown in Equation (4.13). For

110 simplicity, U13 in Equation (2.7) is not included in this test. The airfoil will not change from 1 to 3 because the lift coefficient is always decreasing. Max(U 12 ,U 23 ) G f = 1 1 1 100( C d1 + C d 2 + C d 3 ) 3 3 3 Cl1 = 1.52 , 1 solver tolerance Cl 2 = 1.18 , 2 solver tolerance Cl3 = 0.85 , 3 solver tolerance To investigate the effect of the aerodynamic work term in the actuation energy formulation, two different N-branch GA runs are performed using two sets of different actuation energy formulations shown in Table 4.1. Table 4.1 Actuation energy formulation Aerodynamic load included case Strain energy only case

Minimize: Subject to:

(4.13)

K G K G n 2 U12 = 100 (30L12 i ( f1i ri / q + f 2 i ri / q)L12 i ) i =1 K G K G n 2 U 23 = 100 (30L23i ( f 2 i ri / q + f3i ri / q)L23i ) i =1

n 2 U 12 = 100 (30L12i ) i =1 n 2 U 23 = 100 (30L23i ) i =1

1000 generations (which is considered to be converged), are calculated for each Nbranch GA run. Figure 4.8 compares the Pareto-set of aerodynamic load included versus strain energy only case. Figure 4.8 shows that the objective values are not very different for the two sets of results. One of the possible reasons is that it is very difficult to find airfoils that satisfy the high design lift constraint with a low drag coefficient. The Paretoset shows that the aerodynamic force is acting adversely in most of the region. For this sensorcraft example, the total actuation energy is larger than the strain energy alone for designs with drag objectives below about 0.83. The reason why the very low energy solutions appear to gain assistance from the aerodynamic load at the expense of drag

111 performance would be that there are many possible designs which have small relative energy compared to possible designs with small drag. Also, in Figure 4.8, it is noticeable that the multi-point solution has a lower drag and a lower energy than several of the GA generated designs. Figures 4.9-4.14 compare the airfoil shapes for the aerodynamic work included and strain-energy-only cases designs near three different drag objective values (0.85, 0.78, and 0.75) in the Pareto-set were chosen for this comparison. Including the aerodynamic work term does not greatly affect resulting aerodynamic shapes. However, there are some differences, especially at drag objective 0.85, near the trailing edges. To see the difference more precisely, the actuation energy distribution of the airfoils associated with the drag objective of 0.85 is plotted in Figure 4.15. Figure 4.15 shows that U12 is larger than U13 and that the actuation energy is somewhat different near the trailing edge. It also shows some difference near the center of the upper surface. The aerodynamic force acting on the surface of the airfoil is drawn in Figures 4.16 through 4.18. Figure 4.17 shows that the aerodynamic force can reduce the actuation energy near the center of the upper surface. Figure 4.18 also indicates that the aerodynamic force is very small near the trailing edge and results in small total actuation energy. In Figures 4.19 - 4.21, the Cp distributions for each design point are compared. The Cp distribution shows that the difference between the aerodynamic load included case and the strain energy only case reduces as the design Cl increases. This shows that it is more difficult to find airfoils with small drag for high design Cl condition.

112

0.95

0.9

f 2 (Drag Objective)

0.85

Aerodynamic load included Strain energy only Multi-Point

0.8

Single-Point

0.75

0.7 0 20 40 f 1 (Energy Objective) 60 80

Figure 4.8 Pareto front comparison (GA Generation 1000)

113

0.25 0.2 0.15 0.1 0.05 Design Cl=0.85 (f1=0.85) Desing Cl=1.18 Design Cl=1.52

-0.05 -0.1 -0.15 -0.2


Aerodynamic Load Included

-0.25

0.25

0.5

0.75

Figure 4.9 Airfoil shapes [Aerodynamic load included case, drag objective (0.85)]

0.25 0.2 0.15 0.1 0.05 Design Cl=0.85 (f1=0.85) Design Cl=1.18 Design Cl=1.52

-0.05 -0.1 -0.15 -0.2


Strain Energy Only

-0.25

0.25

0.5

0.75

Figure 4.10 Airfoil shapes [Strain energy only case, drag objective (0.85)]

114

0.25 0.2 0.15 0.1 0.05 Design Cl=0.85 (F1=0.78) Design Cl=1.18 Design Cl=1.52

-0.05 -0.1 -0.15 -0.2


Aerodynamic Load Included

-0.25

0.25

0.5

0.75

Figure 4.11 Airfoil shapes [Aerodynamic load included case, drag objective (0.78)]

0.25 0.2 0.15 0.1 0.05 Design Cl=0.85 (f1=0.78) Design Cl=1.18 Design Cl=1.52

-0.05 -0.1 -0.15 -0.2


Strain Energy Only

-0.25

0.25

0.5

0.75

Figure 4.12 Airfoil shapes [Strain energy only case, drag objective (0.78)]

115

0.25 0.2 0.15 0.1 0.05 Design Cl=0.85 (f1=0.75) Design Cl=1.18 Design Cl=1.52

-0.05 -0.1 -0.15 -0.2


Aerodynamic Load Included

-0.25

0.25

0.5

0.75

Figure 4.13 Airfoil shapes [Aerodynamic load included case, drag objective (0.75)]

0.25 0.2 0.15 0.1 0.05 Design Cl=0.85 (f1=0.75) Design Cl=1.18 Design Cl=1.52

-0.05 -0.1 -0.15 -0.2


Strain Energy Only

-0.25

0.25

0.5

0.75

Figure 4.14 Airfoil shapes [Strain energy only case, drag objective (0.75)]

116

Strain Energy Only

0.5 Aerodynamic Load Included 0.4


Actuation Energy

0.3

0.2 1 0.1 0.6 0 0.1 0 0.1 0 0.1 y 0 0.2 X 0.4 0.8

U12 comparison

Strain Energy Only

Aerodynamic Load Included 0.5

0.4
Actuation Energy

0.3

0.2 1 0.1 0.6 0 0.1 0 0.1 0 0.1 y 0 0.2 X 0.4 0.8

U23 comparison Figure 4.15 Strain Energy Distribution [Drag objective (0.85)]

117

0.6

0.4

0.2

Y
0 -0.2 -0.4
Blue line: Design Cl=1.18 Red line: Design Cl=0.85

-0.2

0.2

0.4

0.6

0.8

X
Figure 4.16 Aerodynamic force distribution on the airfoil surface at both design conditions [Aerodynamic load included case, drag objective (0.85)]

118

0.2

Y
0

Blue line: Design Cl=1.18 Red line: Design Cl=0.85

0.4

0.6

0.8

Figure 4.17 Magnified picture of the aerodynamic force distribution (Center area) [Aerodynamic load included case, drag objective (0.85)]

Y
0

Blue line: Design Cl=1.18 Red line: Design Cl=0.85

0.8

Figure 4.18 Magnified picture of the aerodynamic force distribution (Trailing edge area) [Aerodynamic load included case, drag objective (0.85)]

119

0.25 0.2 0.15 0.1 0.05 Strain Energy Only Aerodynamic Load Included

-0.05 -0.1 -0.15 -0.2 -0.25 0.0 0.3 0.5


Design Cl=0.85 Re=1.5E6 M=0.6

0.8

1.0

-3 -2.5 -2 -1.5 Strain Energy Only Aerodynamic Load Included

-1 -0.5 0 0.5 1 0.0 0.3 0.5


Design Cl=0.85 Re=1.5E6 M=0.6

0.8

1.0

Figure 4.19 Cp distribution comparison [Drag objective (0.85)]

120

0.25 0.2 0.15 0.1 0.05 Strain Energy Only Aerodynamic Load Included

-0.05 -0.1 -0.15 -0.2 -0.25 0.0 0.3 0.5


Design Cl=1.18 Re=1.5E6 M=0.6

0.8

1.0

-3 -2.5 -2 -1.5 Strain Energy Only Aerodynamic Load Included

-1 -0.5 0 0.5 1 0.0 0.3 0.5


Design Cl=1.18 Re=1.5E6 M=0.6

0.8

1.0

Figure 4.20 Cp distribution comparison [Drag objective (0.85)]

121

0.25 0.2 0.15 0.1 0.05 Strain Energy Only Aerodynamic Load Included

-0.05 -0.1 -0.15 -0.2 -0.25 0.0 0.3 0.5


Design Cl=1.52 Re=1.5E6 M=0.6

0.8

1.0

-3 -2.5 -2 -1.5 Strain Energy Only Aerodynamic Load Included

-1 -0.5 0 0.5 1 0.0 0.3 0.5


Design Cl=1.52 Re=1.5E6 M=0.6

0.8

1.0

Figure 4.21 Cp distribution comparison [Drag objective (0.85)]

122 4.4.2 Effect of Stiffness Change Finally, the effect of changing the stiffness of the airfoil on the energy based design is investigated. To do this, three different stiffnesses (k=10q, k=30q, k=50q) are applied and three N-branch GA runs are performed to obtain a Pareto front for each of the three stiffness values. Changing the stiffness means to change the relative importance of the aerodynamic force term in the objective function described in Equation 4.11. For example, low stiffness of the airfoil means it is flexible or easy to deform. Thus a flexible airfoil requires small strain energy to deform, but the aerodynamic load term does not affected by stiffness change. Figure 4.22 shows that the Pareto set variation when the stiffness of the airfoil changes. Three different stiffnesses (k=10q, k=30q, k=50q) are applied in three different runs.

0.95

0.9

Aerodynamic load included (k=30q) Aerodynamic load included (k=10q)

f2 (Drag Objective)

0.85

Aerodynamic load included (k=50q) Strain energy only (k=30q)

0.8 Multi-point Single-point

0.75

0.7 0 20 40 f 1 (Energy Objective) 60 80

Figure 4.22 Pareto front for different stiffness

123 The Pareto front of the total actuation energy shifts as the stiffness changes in Figure 4.22. However, it is not obvious whether the stiffness change affects the designed airfoil shapes or not. To see the differences between the designed airfoil shapes, the designs near the drag objective 0.85 and 0.75 are selected and examined. Figures 4.23-4.25 compare the airfoil shapes for the drag objective value near 0.85 and Figures 4.26-4.28 show the airfoil shapes for the drag objective near 0.75. Figures 4.23-4.28 indicates that the designed airfoil shapes of the Pareto front are not affected much from the selected range of stiffness variation. Only small differences of trailing edge shape are seen in Figure 4.23, when the design lift coefficient is low (Cl =0.85) and the drag objective is high (f2=0.85).

0.25 0.2 0.15 0.1 0.05 k=10q (Cl=0.85) k=50q (Cl=0.85) k=30q (Cl=0.85)

-0.05 -0.1 -0.15 -0.2 -0.25 0 0.25 0.5 0.75 1

Figure 4.23 Airfoil shape comparison [Drag objective (0.85), Design Cl (0.85)]

124

0.25 0.2 0.15 0.1 0.05 k=10q (Cl=1.18) k=30q (Cl=1.18) k=50q (Cl=1.18)

-0.05 -0.1 -0.15 -0.2 -0.25 0 0.25 0.5 0.75 1

Figure 4.24 Airfoil shape comparison [Drag objective (0.85), Design Cl (0.18)]

0.25 0.2 0.15 0.1 0.05 k=10q (Cl=1.52) k=30q (Cl=1.52) k=50q (Cl=1.52)

-0.05 -0.1 -0.15 -0.2 -0.25 0 0.25 0.5 0.75 1

Figure 4.25 Airfoil shape comparison [Drag objective (0.85), Design Cl (1.52)]

125

0.25 0.2 0.15 0.1 0.05 k=10q (Cl=0.85) k=30q (Cl=0.85) k=50q (Cl=0.85)

-0.05 -0.1 -0.15 -0.2 -0.25 0 0.25 0.5 0.75 1

Figure 4.26 Airfoil shape comparison [Drag objective (0.75), Design Cl (0.85)]

0.25 0.2 0.15 0.1 0.05 k=10q (Cl=1.18) k=30q (Cl=1.18) k=50q (Cl=1.18)

-0.05 -0.1 -0.15 -0.2 -0.25 0 0.25 0.5 0.75 1

Figure 4.27 Airfoil shape comparison [Drag objective (0.75), Design Cl (1.18)]

126

0.25 0.2 0.15 0.1 0.05 k=10q (Cl=1.52) k=30q (Cl=1.52) k=50q (Cl=1.52)

-0.05 -0.1 -0.15 -0.2 -0.25 0 0.25 0.5 0.75 1

Figure 4.28 Airfoil shape comparison [Drag objective (0.75), Design Cl (1.52)]

For further investigation of the effect of changing stiffness on the designed airfoil shape, the sensor craft problem is simplified to have only two design conditions as presented in Equation 4.14 below. Also, at this time, a smaller stiffness value k=1q is applied instead of k=10q.
Max(U 12 ) G f = 1 1 100( C d1 + C d 2 ) 2 2 Cl1 = 1.52 , 1 solver tolerance Cl2 = 0.85 , 2 solver tolerance Where, K K n 2 U 12 = 100 (kLi ( f1i / q + f 2i / q)Li ) i =1 n 2 U 12 = 100 (kLi ) i =1 k=1, 30, 50 or (4.14)

Minimize:
Subject to:

127 Figure 4.29 shows the Pareto front solutions of runs using the different stiffness values. In this figure, the values of the strain energy term and the values of the total actuation energy are drawn together. The difference between these two values presents the values of the aerodynamic work term. In the case of k=50q and k=30q, the Pareto front shape shows the same trend between the total actuation energy values and strain energy terms. This means that about this k values, the strain energy is the more dominant factor compared to the aerodynamic work term.

1.1

1.05

1 f2 (Drag Objective) k=q (total actuation energy) k=q (strain energy) k=30q (total actuation energy) k=30q (strain energy) k=50q (total actuation energy) k=50q (strain energy)

0.95

0.9

0.85

0.8

0.75 -10 0 10 20 30 40 50 60 70 f 1 (Energy Objective)

Figure 4.29 Airfoil shape comparison [Drag objective (0.85), Design Cl (1.52)]

However, in the case of k=1q, where the airfoil is more flexible, the aerodynamic work term becomes important. From the previous assumption in section 4.3.2, k=1q means that every spring is displaced 0.03c by the aerodynamic forces shown in Equation 4.9. Figure 4.30 is a rescaled picture of Figure 4.29 to present the Pareto set when k=1q. Figure 4.30 shows that the total actuation energy trend is different from the strain energy

128 trend for designs on the Pareto front. Especially at the small energy and high drag region, the aerodynamic work term is a large portion of total work compared to the designs from the high energy, low drag region. Figures 4.31 and 4.32 compare the airfoil shapes near the drag objective value f1=0.87. Figure 4.31 shows large changes of the designed airfoil shapes when the stiffness value k is equal to 1q. This indicates that as the stiffnesses become smaller, the total actuation energy also decreases and the relative importance of the aerodynamic force term increases. For an actual configuration, k could be chosen based on wing structural properties.

1.3 1.25 1.2 1.15 f2 (Drag Objective) 1.1 1.05 1 0.95 0.9 0.85 0.8 0.75 -1 0 1 f 1 (Energy Objective) 2 3 k=q (total actuation energy) k=q (strain energy)

Figure 4.30 Airfoil shape comparison [Drag objective (0.85), Design Cl (1.52)]

129

0.25 0.2 0.15 0.1 0.05 k=1q (Cl=0.85) k=30q (Cl=0.85) k=50q (Cl=0.85)

-0.05 -0.1 -0.15 -0.2 -0.25 0 0.25 0.5 0.75 1

Figure 4.31 Airfoil shape comparison [Drag objective (0.87), Design Cl (0.85)]

0.25 0.2 0.15 0.1 0.05 k=1q (Cl=1.52) k=30q (Cl=1.52) k=50q (Cl=1.52)

-0.05 -0.1 -0.15 -0.2 -0.25 0 0.25 0.5 0.75 1

Figure 4.32 Airfoil shape comparison [Drag objective (0.87), Design Cl (1.52)]

130
4.5 Summary

Aerodynamic work is modeled and included with the simple strain energy model to provide a more realistic morphing energy model. The enhanced morphing energy model that can account for the effect of airloads is applied to the sensorcraft problem. At high stiffness (k=30q), in most cases, the aerodynamic work acts adversely to change the airfoils. This means, the designs including aerodynamic work require more actuation energy compared to the strain energy only design. Only small actuation energy with large drag designs could get some assist from the aerodynamic work. By varying the stiffness of the airfoils, it is shown that the relative importance of the aerodynamic work increased by reducing the stiffness of the morphing airfoil if the morphing airfoil has small stiffness, it is more important to include the aerodynamic work term in actuation model.

131

CHAPTER 5 CONCLUSIONS

During the research, a parallel genetic algorithm based airfoil optimization strategy using shape functions is developed. Transonic airfoil design issues are investigated by comparing the GA(Genetic Algorithm) and a GM (Gradient based Method); this work employs a Navier-Stokes solver for the transonic drag prediction. Through the research of transonic airfoil design, it is shown that the GA is more robust in searching global optimal solution than the GM and the GA combined with the Navier-Stokes code is affordable for an airfoil design with the help of parallel computing. There has been great interest in developing technologies that may enable a morphing aircraft. Such an aircraft can change shape in flight, which would make it possible to adjust the wing to the best possible shape for any flight condition encountered by the aircraft. However, there is an actuation energy/cost associated with making these shape changes. The concept of a morphing aircraft presents a tradeoff between the aerodynamic performance and the energy or effort needed to morph the aircraft. The development of optimization strategies incorporating morphing cost is essential during the morphing aircraft design process to increase the benefits of using morphing technology. Several multi-objective optimization strategies to design an airfoil set for morphing aircraft are applied to a low-speed, incompressible flow problem and to a problem involving low-speed and transonic flow. In the efforts, the relative strain energy needed to change from on airfoil shape to another is presented as a design objective along with the drag objective. From this multi-objective optimization strategies, it is identified the best tradeoff designs of low actuation energy and low drag that lie between the multipoint shape and the set of single-point shapes. Through these optimization processes,

132 engineers can approach a solution that maximizes the benefits of the morphing technology, while minimizing the actuation cost. A simple strain energy model to present morphing cost has been used in previous work. A new aerodynamic work term is formulated and included into the actuation energy model in this research. Including the aerodynamic work in morphing airfoil optimization represents, using pressure on the airfoil to assist a morphing shape change. In the case of the sensorcraft application, some airfoil sets with small strain energy could get some assistance from the aerodynamic work. However, for most of the airfoil sets satisfying the constraints, the aerodynamic force increases the total actuation energy. This research also demonstrated that it is very important to include aerodynamic work term for actuation energy calculation when the morphing aircraft wing structure is very flexible.
5.1 Future Directions

In Chapter 3, a simple spring model has been adapted to represent the morphing cost. The spring model is adequate to show the concept and explain the energy-based design strategy for morphing aircraft. The strain energy, formulated in the simple spring model, depends only on the amount of deformation of the airfoil. In reality, the stiffness distribution of the airfoil or wing will be different depending on the design of the structure. If a detailed actuation method and structure design is provided, more realistic actuation energy can be calculated and will generate more accurate designs. However, the design strategy presented in this thesis can still work for these detailed designs. Finally, the morphing airfoil problem can be extended to the supersonic flow regime. This extension will require a dramatic change of the airfoil shape and the morphing cost will be increased correspondingly. Because the larger change in shape might provide a broader tradeoff between aerodynamic and actuation energy, the morphing airfoil optimization strategy suggested in this thesis might contribute more to the supersonic morphing airfoil design.

133
5.2 Contributions

A system-level design optimization strategy for a morphing airfoil incorporating minimization of the morphing actuation energy and maximization of the aerodynamic performance was previously not available. Most of the previous works related to a morphing aircraft were focused on the development of a light weight actuator device development. The design strategy presented in this thesis can provide the right reasons and necessities for morphing aircraft development. Also, the strategy will provide the directions and generate target shapes for the detail subsystem development and structure design. Also, a new aerodynamic work formulation on the morphing airfoil is demonstrated in this thesis. This aerodynamic work term can be added to the strain energy term to compute the total energy required for changing the shape of a morphing airfoil. The aerodynamic work term provides a more realistic actuation energy model especially for a flexible morphing aircraft. With this formulation, utilizing aerodynamic work suggests that it is possible to design a morphing airfoil that exploits the airflow to reduce the input actuation energy.

134

LIST OF REFERENCES

[1] Malone, J. B. et al., Airfoil Design Method Using the Navier-Stokes Equations, Journal of Aircraft, Vol.28, No3, 1989, pp.216-224. [2] Anderson, J. D., The Airplane: A History of Its Technology, AIAA, Virgina, 2002. [3] I.H.Abbott and A.E.V.Doenhoff, Theory of Wing Sections, Dover, New York, 1959. [4] Harris, C. D., NASA Supercritical Airfoils, NASA TP 2969, 1990. [5] Bauer, F., Garabedian, P., and Kron, D., A Theory of Supercritical Wing Sections, With Computer Programs and Examples, Volume 66 of Lecture Notes in Economics and Mathematical Systems, M.Beckmann and H.P. Kunzi, eds., Springer-Verlag, 1972. [6] Sobieczky, H., New Design Concepts for High Speed Air Transport, Springer-Verlag, New York, 1997. [7] Barger, R. L., and Brooks, C.W., A Streamline Curvature Method for Design of Supercritical and Subcritical Airfoils, NASA TN D-7770, 1974. [8] Tranen, T. L., A Rapid Computer Aided Transonic Airfoil Design Method, AIAA Paper 74-501, 1974. [ 9 ] Volpe, G., and Melnik, R. E., A Method for Designing Closed Airfoils for Arbitrary Supercritical Speed Distributions, AIAA Paper 85-5023, 1985. [10] Hicks, R. M., Murman, E. M., and Vaderplaats, G. N., An Assessment of Airfoil Design by Numerical Optimization, NASA TM X-3092, 1974. [11] Vanderplaats, G. N., CONMIN-A FORTRAN Program for constrained Function Minimization, Users Manual. NASA TM X-62282 [12] Kennelly, R. A., Improved Method for Transonic Airfoil Design-by-Optimization, AIAA Paper 83-1864, 1983.

135

[13] Jameson, A., Aerodynamic Design Via Control Theory, Inst. For Computer Application in Science and Engineering, Rept. 88-64, NASA Langly, Hampton, VA., Nov.1988. [14] Abergel. F., and Temam, R., On Some Control Problems in fluid Mechanics, Theoretical and Computational Fluid Dynamics, Vol. 1, 1990, pp. 303-325. [15] Holland, J. H., Adaptation in Natural and Artificial Systems, Univ. of Michigan Press, Ann Arbor, MI, 1974. [16] Kirkpatrick C.D. et al.,Optimisation by simulated annealing, Science, 220, 1983, pp 671680. [17] Goldberg, D. E., Genetic Algorithms in search optimization and machine learning, AddisonWesley, MA, 1989. [18] Hajela, P., Genetic Search - An Approach to the Nonconvex Optimization Problem, AIAA Journal, Vol. 28, No.7, Jul. 1990, pp. 1205-1210. [19] Drela, M., Pros & Cons of airfoil optimization, Frontiers of Computational Fluid Dynamics-1998,(D. A. Caughey and M. M. Hafez, eds.), World Scientific Publishers, 1998, pp.363-381. [20] Samareh, J. A., A Survey of Shape Parameterization Techniques, NASA CP 209136, 1999. [21] Salomon, D., Computer Graphic and Geometric Modeling, Springer-Verlag, New York, 1999. [22] Hicks, R. M., Henne, P. A., Wing Design by Numerical Optimization, AIAA Journal of Aircraft, Vol.15, No.7, 1978, pp 407-412. [23] Obayashi, S., and Tsukahara, T., Comparison of Optimization Algorithms for Aerodynamic Shape Design, AIAA Journal, Vol.35, No.8., Aug. 1997, pp. 1413-1415. [24] Holst L. T., and Pulliam H. T., Aerodynamic Shape Optimization Using Real-NumberEncoded Genetic Algorithm, AIAA Paper 2001-2473, 2001. [25] Quagliarella, D. and Della C. A., Genetic Algorithms Applied to the Aerodynamic Design of Transonic Airfoils, AIAA Paper 94-1896-CP, 1994. [26] Oyama, A. Multidisciplinary Optimization of Transonic Wing Design Based on Evolutionary Algorithms Coupled with CFD Solver, ECCOMAS, 2000.

136

[27] Jameson A., Essential Elements of Computational Algorithms for Aerodynamic Analysis and Design, NASA CR 97-206268 [28] Rogallo, R. and Moin, P.,Numerical Simulation of Turbulent Flows, Annual Review of Fluid Mechanics, Vol. 16, 1984, pp. 99-137. [29] Moin, Parviz and Mahesh, Krishnan, Direct Numerical Simulation: A Tool in Turbulence Research,Annual Reviw of Fluid Mechanics, Vol. 30, 1998, pp.539-578. [30] Sparlart, P. R., Jou, W. H., Strelets, M., and Allmaras, S. R.,comments on the Feasibility of LES for Wings and on a Hybrid RANS/LES Approach, in Advances in DNS/LES, ed. C. Liu and Z. Liu, Greyden Press, 1997, pp. 137-149. [31] McGowan, A.R. et al., Recent Results from NASAs Morphing Project, SPIE Paper No.4698-11,9th International Symposium on Smart Structure and Materials, 2002, SanDiego, California [32] Weisshaar,T.,Aeroelastic Tailoring for Energy Efficient Morphing Aircraft-Finding the Right Stuff, ICASE Morphing Lecture Series, 2001. [33] Renken, J. H., Mission Adaptive Wing Camber Control Systems for Transport Aircraft, AIAA Paper 85-5006, 1985. [34] Kudva, J. N., Martin, C. A., Scherer, L. B., Jardine, A. P., McGowan, A. R., Lake, R. C., Sendecky, G. and Sanders, B., Overview of the DARPA/AFRL/NASA Smart Wing Program, Proceedings of SPIE. Vol.3674, pp.230-236. [35] Pendleton, E., Bessette, D., Kehoe, M., and Perry, B., A Flight Research Program for Active Aeroelastic Wing Technology, AIAA Paper 96-1574, 1996. [36] Marks, P.,The next 100 years of flight, NewScientist.com news service, Dec. 2003. [37] Obayashi, S., and Tsukahara, T., Comparison of Optimization Algorithms for Aerodynamic Shape Design, AIAA Journal, Vol.35, No.8., Aug. 1997, pp. 1413-1415. [38] Holst L. T., and Pulliam H. T., Aerodynamic Shape Optimization Using Real-NumberEncoded Genetic Algorithm, AIAA Paper 2001-2473, 2001. [39] Prock, B. C., Weisshaar T. A., Crossley, W.A., Morphing airfoil shape change optimization with minimum actuator energy as an objective, 9th AIAA/ISSMO Symposium on Multidisciplinary Analysis and Optimization, Atlanta, Georgia, 2002.

137

[40] Smareh, J. A., A Survey of Shape Parameterization Techniques, NASA CP 209136, 1999. [41] Hager, J. O., Eyi, S., Lee, K. D., Two-point transonic airfoil design using optimization for improved off-design performace, AIAA Journal of Aircraft, Vol.31, No5, 1994, pp. 11431147. [42] Chang, I, et al.. Geometric Analysis of Wing Sections, NASA TM 110346, 1995. [43] Ramamoorthy, P., and Padmavathi, K., Airfoil Design by Optimization, Journal of Aircraft, Vol. 14, pp. 219-221, 1977. [44] Salomon, D., Computer Graphic and Geometric Modeling, Springer-Verlag, New York, 1999. [45] Harris, D., C., NASA Supercritical Airfoils, NASA TP 2969, 1990. [46] Selig, M., UIUC Airfoil Data Site, [http://amber.aae.uiuc.edu/~m-selig/ads.html], accessed Feb. 2002. [47] Drela, M., Pros & Cons of airfoil optimization, Frontiers of Computational Fluid Dynamics-1998,(D. A. Caughey and M. M. Hafez, eds.), World Scientific Publishers, 1998, pp.363-381. [48] Srinivasan, G. R., Baeder, J. D., Flowfield of a Lifting Rotor in Hover: A Navier-Stokes Simulation, AIAA Journal, Vol.30, No.10, 1992, pp.2371-2378. [49] Roe, P. L., Approximate Riemann solvers, parameter vectors and difference schemes, Journal of Computational Physics, 43, 1981, pp. 357-72. [50] Anderson, W. K., Thomas, J. L., and van Leer, B., A Comparison of Finite Volume Flux Vector Splitting for the Euler Equations, AIAA Paper 85-0122, Jan. 1985. [51] Jameson, A., and Yoon, S., Lower-Upper Implicit Schemes with Multiple Grids for the Euler Equaitons, AIAA Journal, Vol. 25, No. 7, 1987, pp.929-935. [52] Yoon,S.,and Jameson, A.,An LU-SSOR Scheme for the Euler and Navier-Stokes Equations, AIAA Paper 87-0600, Jan.1987. [53] Experimental Data base for Computer Program Assessment, AGARD advisory report, No.138, 1979.

138

[54] Goldberg, D. E., Genetic Algorithms in search optimization and machine learning, AddisonWesley, MA, 1989. [55] Hajela, P., Genetic Search - An Approach to the Nonconvex Optimization Problem, AIAA Journal, Vol. 28, No.7, Jul. 1990, pp. 1205-1210. [56] Jones, B. R., Crossley, W. A., Lyrintzis, A. S., Aerodynamic and Aeroacoustic Optimization of Airfoils via a Parallel Genetic Algorithm, AIAA Journal of Aircraft, Vol. 37, No. 5, Nov.-Dec. 2000, pp. 1088-1096. [57] Zoutendijk, G., Methods of Feasible Directions, Elsevier, Amsterdam, 1960. [58] Williams, E. A., Crossley, W. A., Empirically-derived population size and mutation rate guidelines for a genetic algorithm with uniform crossover, WSC2:2nd Online World Conference on Soft Computing in Engineering Design and Manufacturing, June 1997. [59] Zoutendijk, G., Methods of Feasible Directions, Elsevier, Amsterdam, 1960. [60] Reich, W. G., Bowman, C. J., Sanders, B., Large-Area Aerodynamic Control for HighAltitude Long-Endurance Sensor Platforms, Journal of Aircraft, Vol. 42, No.1, Jan 2005. [61] Martin, E. T., Hassan, R. A., Crossley, W. A., Comparison of the N-Tournament Genetic Algorithm and the Multi-Objective Genetic Algorithm, AIAA Journal, Vol.42, No.6, June 2004. [62] Fonseca, C., and Fleming, P., Genetic Algorithms for Multiobjective Optimization: Formulation, Discussion and Generalization, Proceedings of the Fifth International Conference on Genetic Algorithms, edited by S. Forrest, Morgan Kaufmann, San Mateo, CA, 1993, pp.416-423. [63] Drela, M., XFOIL: An Analysis and Design System for Low Reynolds Number Airfoils, Conference on Low Reynolds Number Airfoil Aerodynamics, University of Notre Dame, June 1989. [64] Crossley, W. A. , Nankani, K., Raymer, D. P., Comparison of Bit-String Affinity and Consecutive Generation Stopping Criteria for Genetic Algorithms, AIAA Paper 2004-449, 2004. [65] Global Hawk First Flight, Aviation Week & Space Technology, March 9, 1998.

139

VITA

03/1989 - 02/1993

Bachelor of Science Department of Aeronautical and Mechanical Engineering Hankuk Aviation University, Seoul, Korea

03/1993 - 02/1995

Master of Science Department of Aerospace Engineering Korea Advanced Institute of Science and Technology (KAIST), Taejon, Korea

03/1995 - 07/2000

Research Engineer Aircraft Division Korea Aerospace Research Institute, Taejon, Korea

08/2000 - 12/2005

Doctor of Philosophy School of Aeronautics and Astronautics Purdue University, West Lafayette, IN

You might also like