multidisciplinary design optimization of a morphing

137
Multidisciplinary Design Optimization of a Morphing Wingtip Concept with Multiple Morphing Stages at Cruise by Michael Leahy A thesis submitted in conformity with the requirements for the degree of Master of Applied Science Mechanical and Industrial Engineering University of Toronto © Copyright by Michael Leahy 2013

Upload: others

Post on 21-Oct-2021

8 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Multidisciplinary Design Optimization of a Morphing

Multidisciplinary Design Optimization of a Morphing Wingtip

Concept with Multiple Morphing Stages at Cruise

by

Michael Leahy

A thesis submitted in conformity with the requirements

for the degree of Master of Applied Science

Mechanical and Industrial Engineering

University of Toronto

© Copyright by Michael Leahy 2013

Page 2: Multidisciplinary Design Optimization of a Morphing

ii

Multidisciplinary Design Optimization of a Morphing Wingtip

Concept with Multiple Morphing Stages at Cruise

Michael Leahy

Master of Applied Science

Mechanical and Industrial

University of Toronto

2013

Abstract

Morphing an aircraft wingtip can provide substantial performance improvement. Most civil

transport aircraft are optimized for range but for other flight conditions such as take-off and

climb they are used as constraints. These constraints could potentially reduce the performance of

an aircraft at cruise. By altering the shape of the wingtip, we can force the load distribution to

adapt to the required flight condition to improve performance. Using a Variable Geometry Truss

Mechanism (VGTM) concept to morph the wingtip of an aircraft with a Multidisciplinary Design

Optimization (MDO) framework, the current work will attempt to find an optimal wing and

wingtip shape to minimize fuel consumption for multiple morphing stages during cruise. This

optimization routine was conducted with a Particle Swarm Optimization (PSO) algorithm using

different fidelity tools to analyze the aerodynamic and structural disciplines.

Page 3: Multidisciplinary Design Optimization of a Morphing

iii

Acknowledgments

I would like to thank my supervisor Dr. Kamran Behdinan for his continuous help and guidance

throughout this work. I would also like to thank my parents and my friend Susan for their

encouragement.

I also appreciate the support from Dr. Fengfeng (Jeff) Xi from Ryerson University and the

financial support from Bombardier and the Natural Sciences and Engineering Research Council

of Canada is also greatly acknowledged.

Page 4: Multidisciplinary Design Optimization of a Morphing

iv

Table of Contents

Acknowledgments .......................................................................................................................... iii

Table of Contents ........................................................................................................................... iv

List of Tables ............................................................................................................................... viii

List of Figures ................................................................................................................................ xi

Nomenclature ................................................................................................................................ xv

List of Appendices ....................................................................................................................... xvi

Chapter 1 The Morphing Aircraft ................................................................................................... 1

1.1 Introduction ......................................................................................................................... 1

1.2 Defining a Morphing Aircraft ............................................................................................. 1

1.3 Current Morphing Aircraft Research .................................................................................. 2

1.3.1 Span Variation ........................................................................................................ 4

1.3.2 Sweep-Span ............................................................................................................. 6

1.3.3 Wing Twist .............................................................................................................. 8

1.4 Wingtip Design ................................................................................................................. 10

1.4.1 Induced Drag ......................................................................................................... 11

1.4.2 Current Morphing Wingtip Research .................................................................... 14

1.5 Thesis Motivation and Overview ...................................................................................... 15

Chapter 2 Variable Geometry Truss Mechanism .......................................................................... 17

2.1 Introduction ....................................................................................................................... 17

2.2 Kinematics ........................................................................................................................ 18

2.3 Modular Discretization ..................................................................................................... 20

2.4 VGTM Structural Layout .................................................................................................. 21

Chapter 3 Multidisciplinary Design Optimization Architectures and Problem Definition .......... 24

3.1 Introduction ....................................................................................................................... 24

Page 5: Multidisciplinary Design Optimization of a Morphing

v

3.1.1 Multi-Disciplinary Feasible .................................................................................. 25

3.1.2 Individual Discipline Feasible .............................................................................. 26

3.1.3 All-At-Once .......................................................................................................... 28

3.1.4 Additional MDO Architectures ............................................................................. 28

3.2 MDO Test Cases ............................................................................................................... 30

3.3 Objective Function ............................................................................................................ 32

3.4 Constraints ........................................................................................................................ 33

3.5 Design Variables ............................................................................................................... 35

Chapter 4 Particle Swarm Optimization Algorithm ...................................................................... 40

4.1 Introduction ....................................................................................................................... 40

4.2 Literature Review .............................................................................................................. 41

4.3 Social and Cognitive Parameter Variations ...................................................................... 42

4.3.1 Linear variation of acceleration constants ............................................................ 42

4.3.2 Step variation of acceleration constants ................................................................ 43

4.3.3 Sinusoidal variation of acceleration constants ...................................................... 44

4.4 Unconstrained Test Cases ................................................................................................. 45

4.5 Constrained Test Case Results .......................................................................................... 46

4.5.1 Problem Description [64] ...................................................................................... 47

4.5.2 Constrained Structural Optimization Results ........................................................ 49

Chapter 5 Discipline Computer Models ....................................................................................... 52

5.1 Introduction ....................................................................................................................... 52

5.2 Low Fidelity Models ......................................................................................................... 52

5.2.1 Aerodynamic Model ............................................................................................. 53

5.2.2 Structural Model ................................................................................................... 54

5.2.3 Aero-Structural Model .......................................................................................... 57

5.2.4 Fluid-Structure Interaction .................................................................................... 59

Page 6: Multidisciplinary Design Optimization of a Morphing

vi

5.3 High Fidelity Models ........................................................................................................ 61

5.3.1 Aerodynamic Model ............................................................................................. 61

5.3.2 Structure Model .................................................................................................... 63

5.3.3 Finite Element Mesh ............................................................................................. 65

5.3.4 VGTM Model ....................................................................................................... 66

5.3.5 Aero-Structural Analysis ...................................................................................... 67

Chapter 6 Multidisciplinary Design Optimization Results ........................................................... 69

6.1 Introduction ....................................................................................................................... 69

6.2 Low Fidelity Results ......................................................................................................... 69

6.2.1 Optimal Design for Two Morphing Stages ........................................................... 72

6.2.3 Optimal Design for Four Morphing Stages .......................................................... 76

6.2.4 Low Fidelity Summary ......................................................................................... 81

6.3 High Fidelity Results ........................................................................................................ 82

Chapter 7 Conclusion .................................................................................................................... 89

7.1 Thesis Conclusion ............................................................................................................. 89

7.2 Future Work ...................................................................................................................... 90

Bibliography ................................................................................................................................ 92

Appendix A – PSO Results ......................................................................................................... 100

A.1 Linear Test Case Results for Static and Dynamic .......................................................... 100

A.2 Step Test Case Results for Static and Dynamic .............................................................. 102

A.3 Sinusoidal Test Case Results for Static and Dynamic .................................................... 104

Appendix B – Beam Finite Element Model Validation .............................................................. 108

Appendix C – MDO Result Extras ............................................................................................. 111

C.1 Optimal Reference Wing Mass .................................................................................. 111

C.2 Two Morphing Stages .................................................................................................. 112

C.3 Four Morphing Stages ................................................................................................. 114

Page 7: Multidisciplinary Design Optimization of a Morphing

vii

C.4 High Fidelity Results .................................................................................................... 118

Page 8: Multidisciplinary Design Optimization of a Morphing

viii

List of Tables

Table 1.1 - Advantages and disadvantages of different morphing degrees of freedom.................. 2

Table 3.1 - Common MDO frameworks ....................................................................................... 28

Table 3.2 - Possible objective for each mission leg ...................................................................... 31

Table 3.3 - Design variable list for low fidelity models ............................................................... 35

Table 3.4 - Design variable list for high fidelity models .............................................................. 37

Table 4.1 - Unconstrained benchmark problems .......................................................................... 45

Table 4.2 - Load conditions for the 25-bar truss ........................................................................... 47

Table 4.3 – Test cases for 25-bar space truss ................................................................................ 49

Table 4.4 - Constrained test case results ....................................................................................... 50

Table 5.1 - Reference wing geometry ........................................................................................... 52

Table 5.2 - Comparing cross-sectional data from ANSYS to current FE model (X and Y Scaling

are set to 0.5) ................................................................................................................................. 56

Table 6.1 - Optimal design for wing structure of reference wing ................................................. 70

Table 6.2 - Performance results for the 2 Morphing stage case .................................................... 74

Table 6.3 - Performance results for 4 morphing stages ................................................................ 76

Page 9: Multidisciplinary Design Optimization of a Morphing

ix

Table 6.4 - Structural Optimization Results Comparison ............................................................. 82

Table 6.5 - First and second morphing stage performance results ............................................... 87

Table A.1 - Linear test case for unconstraint benchmark problems ........................................... 100

Table A.2 - Linear test case results for unconstraint benchmark problems ................................ 101

Table A.3 - Step test case for unconstraint benchmark problems ............................................... 102

Table A.4 - Step test case results for unconstraint benchmark problems ................................... 103

Table A.5 - Sinusoidal test case for unconstraint benchmark problems ..................................... 104

Table A.6 – Sinusoidal test case results for unconstraint benchmark problems ......................... 105

Table A.7 – Best overall social and cognitive distributions ....................................................... 107

Table B.1 - Material and geometric properties for problem 5.2 ................................................. 108

Table B.2 - Nodal displacement and nodal reaction comparison ............................................... 108

Table B.3 - Material and geometric properties for problem 5.8 ................................................. 109

Table B.4 - Nodal displacement comparisons ............................................................................ 110

Table C.1 - Optimal module design for 2 morphing stage case study ........................................ 112

Table C.2 – Optimal main wing and wingtip design for 2 morphing stage case study .............. 112

Table C.3 - Optimal module design for 4 morphing stage case study ........................................ 114

Table C.4 - Optimal main wing and wingtip design for 4 morphing stage case study ............... 115

Page 10: Multidisciplinary Design Optimization of a Morphing

x

Table C.5 - High fidelity first and second stage module results ................................................. 118

Table C.6 - High fidelity main wing and wingtip results ........................................................... 119

Page 11: Multidisciplinary Design Optimization of a Morphing

xi

List of Figures

Figure 1.1 - Airfoil section with pressure and shear forces .......................................................... 10

Figure 1.2 Escape of air around the wing tip [31] ........................................................................ 11

Figure 1.3 - Shed vortex sheet from trailing edge ........................................................................ 11

Figure 1.4 - Effect of downwash on the local flow over a local airfoil section of a finite wing [32]

....................................................................................................................................................... 12

Figure 1.5 - Different wingtip designs [31] .................................................................................. 13

Figure 1.6 - Boeing 767-400 with raked wingtip .......................................................................... 14

Figure 2.1 - VGTM like structure (left) and traditional truss structure (right) ............................. 17

Figure 2.2 - Base configuration of the modular VGTM morphing wing [45] .............................. 18

Figure 2.3 - Two coordinate systems arbitrarily orientated in 3D space ...................................... 19

Figure 2.4 - Discretized morphing wing used for the current work .............................................. 21

Figure 2.5 - Module structural layout ........................................................................................... 22

Figure 2.6 - Module degree of freedoms ...................................................................................... 23

Figure 3.1 - MDO structure for three disciplines .......................................................................... 24

Figure 3.2 - MDF architecture for aeroelastic analysis ................................................................. 25

Figure 3.3 - IDF architecture for aeroelastic analysis ................................................................... 27

Figure 3.4 - Aircraft mission showing mission legs ..................................................................... 31

Page 12: Multidisciplinary Design Optimization of a Morphing

xii

Figure 3.5 - Design space without constraint on objective (left) design space with constraint on

objective (right) ............................................................................................................................. 34

Figure 4.1 - Linear variation of particle acceleration constants .................................................... 43

Figure 4.2 - Step variation of particle acceleration constants ....................................................... 44

Figure 4.3 - Sinusoidal variation of particle acceleration constants ............................................. 45

Figure 4.4 - 25-bar space truss ...................................................................................................... 47

Figure 5.1 - Wingbox geometry confined within airfoil ............................................................... 55

Figure 5.2 - FE nodes (dots) with aerodynamic model ................................................................. 57

Figure 5.3 - Equivalent nodal loads for concentrated point loads ................................................ 58

Figure 5.4 - Deformed and undeformed wing shape .................................................................... 58

Figure 5.5 - Convergence plots for an elastic wing at constant angle of attack ........................... 60

Figure 5.6 - Convergence plots for an elastic wing while simultaneously being trimmed to

specified weight requirement ........................................................................................................ 61

Figure 5.7 - ICEM CFD unstructured surface mesh ..................................................................... 62

Figure 5.8 - Internal Structure layout ............................................................................................ 64

Figure 5.9 - Wing geometry without top skin panels .................................................................... 64

Figure 5.10 - FE mesh of wing ..................................................................................................... 66

Figure 5. 11 - FE mesh of VGTM structure .................................................................................. 67

Page 13: Multidisciplinary Design Optimization of a Morphing

xiii

Figure 6.1 - Mass convergence plot for reference wing ............................................................... 70

Figure 6.2 - Stress distribution for reference wing ....................................................................... 71

Figure 6.3 - Bending stiffness distribution for reference wing ..................................................... 71

Figure 6.4 - Fuel convergence for 2 morphing stage case ............................................................ 72

Figure 6.5 - Morphed wingtip for the first stage ........................................................................... 73

Figure 6.6 - Morphed wingtip for the second stage ...................................................................... 74

Figure 6.7 – Stress constraint distribution for 2 morphing stages ................................................ 75

Figure 6.8 - Fuel convergence for 4 morphing stages .................................................................. 76

Figure 6.9 - Aerodynamic performance of the 4 morphing stages ............................................... 78

Figure 6.10 - First morphing stage wingtip shape for the 4 morphing stage case ........................ 79

Figure 6.11 - Second morphing stage wingtip shape for the 4 morphing stage case .................... 79

Figure 6.12 - Third morphing stage wingtip shape for the 4 morphing stage case ....................... 80

Figure 6.13 - Fourth morphing stage wingtip shape for the 4 morphing stage case ..................... 80

Figure 6.14 - Stress Constraint distribution for 4 morphing stage case ........................................ 81

Figure 6.15 - Von-Mises stress contours for optimal reference wing structure ............................ 83

Figure 6.16 – High fidelity first morphing stage optimal wing shape .......................................... 84

Figure 6.17 – High fidelity first morphing stage optimal wingtip shape ...................................... 84

Figure 6.18 - High fidelity second morphing stage optimal wingtip shape .................................. 85

Figure 6.19 - Pressure contours of wingtip region for first and second stage .............................. 86

Figure 6.20 - Von-Mises stress contours for first morphing stage ............................................... 87

Page 14: Multidisciplinary Design Optimization of a Morphing

xiv

Figure 6.21 - Von-Mises stress contours for second morphing stage ........................................... 87

Figure B.1 - Problem 5.2 ............................................................................................................. 108

Figure B.2 - Problem 5.8 ............................................................................................................. 109

Figure C.1- Main wing and wing-module junction cross-sectional areas .................................. 111

Figure C.2 - Modules and wingtip cross-sectional area .............................................................. 111

Figure C.3 - First morphing stage for 2 morphing stage case ..................................................... 113

Figure C.4 - Second morphing stage for 2 morphing stage case ................................................ 114

Figure C.5 - First morphing stage for 4 morphing stage case ..................................................... 116

Figure C.6 – Second morphing stage for 4 morphing stage case ................................................ 117

Figure C.7 - Third morphing stage for 4 morphing stage case ................................................... 117

Figure C.8 - Fourth morphing stage for 4 morphing stage case ................................................. 118

Figure C.9 - Pressure distribution of first morphing stage wing showing shockwave at trailing

edge ............................................................................................................................................. 120

Figure C.10 - Pressure distribution of second morphing stage wing showing shockwave at

trailing edge ................................................................................................................................ 121

Figure C.11 - Comparing deformed to undeformed wing for the first morphing stage .............. 121

Figure C.12 - Comparing deformed and undeformed for second morphing stage ..................... 121

Page 15: Multidisciplinary Design Optimization of a Morphing

xv

Nomenclature

R Range ΔW Fuel Consumed

N Morphing Stage or Swarm Size σ Stress

W Weight σY Yield Stress

L Lift c1 Cognitive Acceleration Constant

D Drag c2 Social Acceleration Constant

V Free Stream Velocity A Cross-Sectional Area

CL Lift Coefficient M Mach Number or Bending

Moment

CD Drag Coefficient Ix Out-Of-Plane Second Area

Moment of Inertia

CL/CD Lift to Drag Ratio Iy In-Plane Second Area Moment

of Inertia

Ri,i-1 Rotation Matrix Γ Dihedral/Cant Angle

Λ Sweep Angle θ Twist Angle

b span t Airfoil thickness

U,M,L Upper, Middle, Lower Group

Page 16: Multidisciplinary Design Optimization of a Morphing

xvi

List of Appendices

Appendix A – PSO Results ………………………………………………………….100

Appendix B – Beam Finite Element Model Validation ……………………..……....108

Appendix C – MDO Result Extras ………………………………………………..…111

Page 17: Multidisciplinary Design Optimization of a Morphing

1

Chapter 1

The Morphing Aircraft

1.1 Introduction

Morphing vehicles are classified as structures exploiting actuation mechanisms to achieve

substantial wing area change resulting in an improved flight envelope. The concept behind

morphing vehicles stems from the desire to expand the mission capabilities of aircraft platforms.

Radical wing area changes enable a single aircraft to perform the roll of multiple aircraft. The

recent advances in new materials and structures are allowing engineers to design aircraft that can

meet specific environment requirements. Much of these technologies and applications allow

large changes in shape to maximize performance and efficiency of aircraft. The problems with

large shape changes for large aircraft are the actuation power requirements and structural stress

constraints. To address this problem, shape changing near the wingtip, where the loading is the

lowest is preferred. With rising fuel costs, aircraft designers are motivated to improve the fuel

efficiency of existing aircraft. By morphing the wingtip region we can redistribute the

aerodynamic loads thus improving vehicle performance. Using a Variable Geometry Truss

Mechanism (VGTM) concept the shape the wingtip, the question is what is the optimal shape

subjected to structural and performance constraints? Using a Particle Swarm Optimization (PSO)

algorithm with a Multidisciplinary Design Optimization (MDO) framework an optimal wing and

wingtip shape will be determined and compared to a conventional wing.

1.2 Defining a Morphing Aircraft

The evolution of birds has inspired humans to fly for centuries. The idea of “Morphing Aircraft"

is a direct result of this inspiration. According to [1] avian morphology permits a wide range of

wing configurations that can be used in a variety of flight conditions which are used for a

particular flight task. A bird will loiter with a high aspect ratio wing until it detects its prey, upon

detection it will retract its wings to reduce wing area and dive towards its prey. Morphing wings

have the potential to improve aircraft performance, however there is no agreed upon definition of

"Morphing Aircraft".

Page 18: Multidisciplinary Design Optimization of a Morphing

2

In the mid-1990s NASA initialed the "aircraft morphing" project. This project defined morphing

as "efficient multi-point adaptability". Efficient implied lighter and/or more energy efficient

conventional systems, multi-point implied the accommodation of different flight requirements,

and adaptability implied increased versatility [2]. This is a general definition that includes small

and large shape changes, and virtual shapes through micro flow control. The Defense Advanced

Research Projects Agency (DARPA) initiated the "Morphing Aircraft Structures" program in

2002. In summary the goal of the program was to develop aircraft that can change their shape

substantially, provide superior system capability, and uses designs that integrates innovative

technology [3]. This definition applies mostly to Unmanned Aerial Vehicles (UAV), since large

shape changes would be difficult for larger transport aircraft. In [4] the authors defined a

morphing wing as a bird like wing that has the ability to adapt to accommodate multiple flight

regimes or to obtain better flight performance. This definition implies seamless large shape

changes to achieve its goal. In [5], morphing aircraft were defined as a set of technologies that

increase a vehicles performance by manipulating certain characteristics to better match the

vehicle state to the environment and task at hand. This problem with this definition is it allows

existing technologies such as flaps and retractable landing gear to be considered morphing. The

NATO Research and Technology Organization, Applied Vehicle Technology Technical Team on

Morphing Vehicles developed the following definition: real-time adaptation to enable multi-

point optimized performance [2]. This definition points towards the desired capability for

morphing vehicles that is distinct from conventional vehicles.

1.3 Current Morphing Aircraft Research

There have been several approaches to morphing an aircraft wing structure but the objective of

the morphing process has been the same, which is to improve the aerodynamic performance by

morphing the wing span, sweep, and twist. Other morphing options include the dihedral/cant or

the chord length of the wing. Each of these morphing degrees of freedom has their own

particular influence on aerodynamic performance which is summarized in the table below.

Table 1.1 - Advantages and disadvantages of different morphing degrees of freedom

Morphing

Degree of

Freedom

Description Advantages Disadvantages

Page 19: Multidisciplinary Design Optimization of a Morphing

3

Span

Increased Range,

Endurance

Decreased fuel

consumption

Increased

bending

stress at

wing root

Increased

structural

weight

Slower roll

control

authority

Sweep

Reduced takeoff

length

Increased

maneuverability at

low speeds

Increased

weight

from

sweeping

mechanis

m

Aircraft

longitudina

l stability

will be

affected

Twist

Increased roll

control (reduces

the chances of

stall at the

wingtip)

Increase in

drag is a

possibility

Page 20: Multidisciplinary Design Optimization of a Morphing

4

Dihedral

Improved lateral

stability

Lower

maneuvera

bility

Chord/Airfoil

Potential increase

in lift

Potential

increase in

drag

This report is primarily concerned with the span, sweep, twist, and dihedral morphing degrees of

freedom. Chord (or airfoil morphing) is a subject of detailed research in academia but it is not the

main focus here. A literature review is presented in the proceeding sections.

1.3.1 Span Variation

Increasing the wing span will improve the aerodynamic efficiency by decreasing the induced

drag. This translates into increased range and endurance by improved fuel efficiency. However

aircraft with large wing spans have poor maneuverability (increases roll damping). This creates a

very low roll rate, which would be unattractive to high performance aircraft designers. The other

problem with an increased span is the increase in root bending moment which creates larger

stresses.

The concept of a variable span is not a new one. In 1931 Ivan Makonine developed a telescoping

span aircraft. An inner wing section was extended from the outer wing section using pneumatic

actuation to increase wing span and area. The span increased by 62% and area by 57% [5]. In [6]

they created a telescopic wing, which could increase wing span by 10 inches. They used a radio

controlled aircraft, and an increase of 19% in range was achieved compared to the baseline wing.

A change in static stability of about 5% was also determined. In [7] and [8], they developed a

morphing wing concept that used a pressurized telescopic spar to create large spanwise changes.

Page 21: Multidisciplinary Design Optimization of a Morphing

5

Their telescopic spar design consisted of three concentric circular tubes that can support the

aerodynamic loading. Their initial design in 2003 suffered from parasitic drag increase due to the

seams on the wing which resulted in a 25% reduction in lift to drag ratio compared to its fixed

wing counterpart. In 2007 they improved their design by adding a second telescopic spar. Their

design was able to achieve a 230% increase in aspect ratio while supporting aerodynamic loads.

The spar was tested under different loading conditions and different air pressure for the actuator.

The response of the wingtip in the longitudinal direction was measured during extension and

retraction. High pressurization was able to achieve the maximum span in the quickest time. Their

Wind tunnel test results were compared to theoretical results and a solid wing with the same

dimensions at 40%and 100% span extension. For each span length, the morphing wing

underperformed the solid wing. The authors mentioned that this was most likely caused by lower

skin stiffness and the seams on the wing. Their design was able to achieve lift to drag ratios as

high as 16. In [9] they investigated the aerodynamic and aeroelastic characteristics of a variable

span morphing wing applied to a long range cruise missile. A finite element model consisting of

four wingbox sections for the main wing and two wingbox sections for the telescoping wing was

used. Their results showed a reduction in spanwise lift and induced drag but increased wing

deflections with increasing span. The wing divergent natural frequencies were also determined

and showed that the natural frequencies would decrease with increasing span. In [10] they tested

a morphing telescopic wing in a channel studying its aerodynamic and roll effects for wing-in-

ground effects. The change in aspect ratio increased the lift to drag ratio compared to only the

ground and side wall effects. It was found that the extension of the wing tip was inefficient at

controlling the rolling moment; however the aspect ratio change was small. In [11] they used

asymmetric span extension to determine roll performance. They were able to achieve a roll

moment coefficient of 1.5. They extended their model in [12] to include the effects of morphing

into the aircraft dynamics model and it was shown that span extension was effective for roll

control. Span extension induces roll damping that is greater than aileron control. In [13] they

invested asymmetric span extension for roll control of a cruise missile. They showed that the

control of the missile can be larger compared to a conventional tail surface. They presented a

nonlinear control model to control the roll, angle of attack, and side slip. In [14] they developed a

morphing wing for a UAV with span extension. It was built and tested in a wind tunnel and their

design could increase the wing area by 100%. A silicone elastomer skin which had embedded

unidirectional carbon fibre to minimize poission's ratio effects and increase chord wise in-plane

Page 22: Multidisciplinary Design Optimization of a Morphing

6

stiffness. A morphing core substructure with zero poission ratio was also used. The honeycomb

substructure has low stiffness in the direction of actuation allowing for large displacements.

Wind tunnel tests at two different speeds at three different angles of attack were conducted to

determine the amount of out of plane skin deformation which was only 0.5mm.

Lockheed Martin developed a folding wing approach that incorporated flexible skins along the

fold line. Their concept has two main configurations: a loiter configuration where the wing span

is extended by unfolding the wing and a dash configuration where the wing is folded up against

the fuselage to reduce the wing span. Testing had shown that the model can with stand the

aerodynamic loads at each configuration [15] and [16].

1.3.2 Sweep-Span

For high speed aircraft, their wings are designed with a sweep angle to delay the drag rise at high

Mach numbers. However at low speeds maneuverability is poor and longer take-off distance is

required. Varying the sweep reduces the take-off and landing distance and improves low sweep

performance by decreasing the sweep angle. Changing the wing sweep also moves the centre of

gravity and aerodynamic centre, thus effecting longitudinal stability. The sweep angle can also

affect the lateral stability by creating an effective dihedral. Also, a variable sweep wing will

change its span length. The application of variable sweep wing has been in use for several

decades. The Northrop Grumman F-14 Tomcat has the ability to change its sweep angle from 20

to 68 degrees. The B-1 Lancer bomber can also change sweep for high speed cruise. Other swept

wing concepts have been designed and tested in the Soviet Union. One of the main issues for

these swept wing designs is the heavy weight penalty associated with the sweeping mechanism.

In [17] they performed an experimental analysis of a variable geometry aircraft that can change

both sweep and span. The goal of the wind tunnel test was to determine the influence of platform

variables on the models aerodynamic characteristics. As the wing sweep was increased the

aerodynamic centre shifted more towards the tail section of the model, however since the wing is

being swept back the centre of gravity is also shifting towards the tail section. The aircraft is

more stable since there is an increase in the static margin. At high lift coefficients the induced

drag is dominant over profile drag, a full span extension and full sweep showed better drag

performance at lift coefficients greater than 0.4 compared to an unmorphed baseline. However at

lower lift coefficients the baseline wing performed better. In [18] they presented a bird like

Page 23: Multidisciplinary Design Optimization of a Morphing

7

design with an inner and outer section that can increase its aspect ratio from 4.7 to 8.5. Using a

lightweight piezo-composite actuator their design could change camber as well. They were able

to achieve a significant increase in lift coefficient for an increasing aspect ratio and a reduction in

drag at different angles of attack.

NextGen Aeronautics developed a UAV referred to as the ‘batwing’, which can change its sweep

and span [19] and [20]. The wing can change its aspect ratio by 200%, it span by 40% and wing

area by 70%. Wing sweep and area are controlled through four bar linkages, revolute, and slider

mechanisms. Their morphing UAV successfully completed flight tests in 2006. Another larger

morphing UAV was built and tested in 2007. This design was capable of 40% change in wing

area, 177% change in aspect ratio, and 73% change in span.

In [21] they performed an optimization problem based on the static equilibrium and principle of

virtual work to determine the optimal location of an actuator in a scissor mechanism.

Optimizations were conducted for single and multiple cell mechanism. They used an Sequential

Quadratic Programming (SQP) algorithm for their optimization with the location of the actuator

as the design variable. The wing skin was modeled using linear spring elements. Optimizations

were conducted with different link thicknesses. Their results showed that a distributed actuation

system was more efficient with a flexible mechanism. This work was extended in [22]. A

nonlinear finite element model was used to determine the nodal displacements in a scissor like

mechanism in a two stage optimization routine. The first stage used a GA to determine the

existence of an actuator. The actuators were either active or in active in the cell using a discreet

variable. Constraints were imposed on the number of active actuators and the extension of the

mechanism was restricted. The second stage optimization used an SQP approach to determine

the position of the actuator within each cell. For both stages the objective was to maximize the

efficiency of the mechanism defined as the ration of output energy to input energy. The skin was

modeled by linear springs connecting opposite nodes. a high and low aerodynamic load case

were examined with skin representing isotropic and orthotropic stiffness. Their results showed

that the efficiency is greatly dependent on the skin stiffness and applied external loading. Also

the morphing efficiency under high loads were better than lower loads. In [23] they presented a

methodology that employed planar unit cells as the layout as the internal structure. The sizing of

this mechanism was based on simple beam theory to determine the bending moments assuming a

Page 24: Multidisciplinary Design Optimization of a Morphing

8

linearly distributed aerodynamic load. A prototype was designed and built using a cable

actuation mechanism. This mechanism can change the span and sweep angle. In [24] they

extended the NextGen concept by determining the optimal actuator orientation for rigid and

flexible structures.

1.3.3 Wing Twist

Creating a wash-in or wash-out of the wing can significantly alter the span wise loading of the

aerodynamic loads. This can improve aircraft performance and control authority. The Wright

brothers used a wing warping approach to vary the twist of their wing for roll control. Increased

aircraft speed forced designers to build stiffer wings to prevent wing divergence, aileron reversal,

and flutter. This came at with a significant weight penalty.

Taking advantage of wing flexibility to reduce structural weight was the goal of Rockwell

International Active Flexible Wing program in the 1980s. The objective was to use the control

forces to twist the wing structure to enhance roll performance. Other programs such as the Active

Aeroelastic Wing program funded by NASA and the USAF, and the Active Aeroelastic Aircraft

Structures in Europe explored other approaches to active aeroelastic concepts for roll control.

In [25] they proposed a Variable Stiffness Spar (VSS) concept to improve the aircraft roll

performance. The mechanism consists of a segmented spar having articulated joints at the

connections with wing ribs and an electric actuator capable of rotating the spar. With the spar in

the horizontal position there is very little bending stiffness compared to the vertical position.

Their concept allows the stiffness to be controlled as a function of flight conditions. An

optimization was performed at different transonic Mach numbers to maximize the roll rate

subjected to flutter, control surface hinge moment, and maximum deflection constraints. Two

VSS concepts were tested: for the first concept a single VSS was used and for the second, a

torsion-free concept where two stiff spars positioned close together carried most of the bending

moment with two VSS (one VSS near the leading edge and the other near the trailing edge). The

torsion-free concept allows for more control over the torsion stiffness. The single VSS was able

to achieve a 6-22% increase in roll rate and the second approach was able to achieve a 29-126%

improvement. In [26] they performed wind tunnel tests on a wing with multiple control surfaces

and VSS mechanism. The VSS was located at 60% of the chord extending to 58% of the wing

span. The effects of the VSS on modal frequencies, lift curve lope, and control surface

Page 25: Multidisciplinary Design Optimization of a Morphing

9

effectiveness were determined. Their predicted modal frequencies were lower than the measured

frequencies suggesting that the wind tunnel model was stiffer than the FE model. The horizontal

orientation of the VSS shows a decrease in lift curve slope compared to the vertical position

however the difference between them is small at subsonic speeds compared to transonic and

supersonic. Similarly the differences in control surface derivatives for the horizontal and vertical

position were only apparent at subsonic speeds. In [27] they used a Shape Memory Alloy (SMA)

spar approach for the enhancement of aircraft roll performance has been presented to replace the

VSS concept which was mechanical actuation system with a large weight penalty. It was

observed that the baseline aircraft has about 65 % of loss in the roll performance due to wing

flexibility at Mach number 0.95. With activation of SMA spar located at the leading edge, a 59

% roll rate improvement was achieved. And 61% roll rate improvement was achieved by using a

SMA spar located at the trailing edge.

In [28] they introduced the Flexspar configuration which employs an aerodynamic shell that can

pivot around a high strength main spar. A piezoelectric actuator is mounted inside the

aerodynamic shell with one end bonded to the spar and the other to the shell. As the piezoelectric

actuator is energized there is a change in pitch angle. Wind tunnel tests showed deflections of +/-

11o with collocated pitch axis, aerodynamic centre, and centre of gravity. With the quarter chord

moved ahead of the pitch axis the deflections are magnified to +/-16o. In [29] they built a series

of piezoelectric specimens to demonstrate the utility of the new post-buckled pre-compressed

(PBP) actuator scheme. These actuator elements were designed as actuators for grid fins for a

new class of subscale UAVs (vertical take-off and landing capable). An extensive series of tests

with varying levels of axial pre-compression showed that static deflections could be controllably

magnified approximately 4.5-fold over conventional levels, with excellent agreement between

theory and experiment. In [30] they presented a new class of flight control actuators integrated

into a flexible wing, allowing it to be deformed upon actuation. By using PBP actuators their aim

was to reduce weight, complexity and power consumption with respect to conventional actuators,

while increasing control bandwidth and control authority. By applying axial compression to

piezoelectric bimorph bender actuators, significantly higher deflections can be achieved than for

conventional piezoelectric bender actuators. It was shown during flight with a 1.4m span UAV

that wing morphing could produce 38% more roll control and 3.7 times greater control

derivatives than conventional approaches. Because PBP actuators are solid state and do not

Page 26: Multidisciplinary Design Optimization of a Morphing

10

employ any linkages, pushrods or gears, they operate very efficiently This efficiency lead to a

99.6% decrease in power consumption, 87% reduction in flight control system-related weight, an

order of magnitude increase in control actuation bandwidth, and an order of magnitude fewer

parts than conventional electromechanical servoactuators [30].

1.4 Wingtip Design

The intended purpose of all wingtip devices is to reduce the overall drag of the aircraft. Drag is

just the flight-direction component of the total aerodynamic force. This aerodynamic force can

be separated into a component parallel to the local surface (shear force) and a component

perpendicular to the surface (pressure force).

Figure 1.1 - Airfoil section with pressure and shear forces

When these two components are integrated over the entire wing surface, the resulting forces are

referred to as the "skin-friction" drag and the pressure drag. The skin-friction drag is a result of

viscous effects in the boundary layers on the airplane’s surfaces. The pressure drag is a result of

a more complicated combination of flow mechanisms, including viscous effects, shocks, and the

global effects of lift. Induced drag is a significant part of the drag due to the global effects of lift

and it will be discussed in more detail.

Page 27: Multidisciplinary Design Optimization of a Morphing

11

1.4.1 Induced Drag

When a wing generates aerodynamic lift, the air on the top surface has lower pressure relative to

the bottom surface. The air will ‘escape’ from below the wing and out around the tip to the top of

the wing in a circular fashion. This generates a tip vortex shown in Figure 1.2.

Figure 1.2 - Escape of air around the wing tip [31]

In actuality the vortices that feeds into the vortex cores generally comes from the entire span of

the trailing edge, not just from the wingtips. The shed vortices from the trailing edge roll up near

the outer edges of the wing creating two distinct vortex cores. This is show in Figure 1.3.

Figure 1.3 - Shed vortex sheet from trailing edge

Page 28: Multidisciplinary Design Optimization of a Morphing

12

The rotary motion of the air within the shed wing vortices reduces the effective angle of attack of

the air on the wing. This is best illustrated with the 2D airfoil in Figure 1.4. For a 2D airfoil the

lift and drag forces are taken relative to the global frame of reference. The airfoil is rotated with

a geometric angle of attack (α) relative to the direction of travel of the wing velocity (V). Since

the aircraft wing is finite there is an induced angle of attack (αi) caused by the downwash effect

(w) of the shed vortices. The air is deflected by the downwash which causes the total lift vector

to tilt back. The aft component of this lift vector is the induced drag (Di). The magnitude of the

induced drag is determined by the spanwise distribution of vortices shed downstream of the wing

trailing edge, which is related in turn to the spanwise lift distribution.

Figure 1.4 - Effect of downwash on the local flow over a local airfoil section of a finite wing

[32]

For high-lift low-speed conditions such as during takeoff or for high subsonic cruise speed, the

induced drag from the vortices can account for up to half of all drag [33]. To reduce the induced

drag effect several wingtip concepts have been implemented as shown in Figure 1.5. The first

studies on wingtip devices were conducted in 1897 by English engineer Frederick W.

Lanchester. His design called for endplates to be attached to the wingtip, and this design showed

a decrease in induced drag at high lift coefficient. However during cruise the airflow would

Page 29: Multidisciplinary Design Optimization of a Morphing

13

separate near the tip increasing the profile drag of the wing [33]. The most common wingtip

device is the Hoerner wingtip [34] due to its relative ease of construction, however it can only

achieve modest improvements in the lift to drag ratio. For commercial transports and business

jets, the winglet is the preferred wingtip device. First developed by Richard T. Whitcomb in the

1970’s, his design was specifically intended for lifting and subsonic Mach numbers. Wind tunnel

tests at Langley showed a reduction in induced drag of almost 20 percent and an increase in wing

lift-drag ratio of roughly 9 percent. This was twice the performance enhancement given by

simple tip extensions [35]. The problem with winglets is the increase in bending moment at the

wing mount attached to the fuselage. This increase in bending moment is attributed to the

increase in wing loading and added weight at the wingtip. The wings flutter speed can also be

adversely affected. Also due to the higher aerodynamic loading at the wingtip, roll performance

can be greatly affected.

Figure 1.5 - Different wingtip designs [31]

Another wingtip device used mainly by Boeing is the raked wingtip shown in Figure 1.6. They

are described as an integrated wingtip extension, as they are horizontal additions to the existing

wing, rather than the previously described vertical solutions. The main difference between a

traditional winglet and a raked wingtip is the amount of cant. They are another wingtip

technology to increase efficiency, to improve climb performance, and to shorten takeoff distance,

all by increasing the effective aspect ratio of the wing by disrupting the detrimental wingtip

Page 30: Multidisciplinary Design Optimization of a Morphing

14

vortices. Such technologies and retrofits promise between 6 and 12 percent increases in range

[36]. In testing by Boeing and NASA, raked wingtips have been shown to reduce drag by as

much as 5.5%, as opposed to improvements of 3.5% to 4.5% from conventional winglets [37].

Figure 1.6 - Boeing 767-400 with raked wingtip

The benefit of induced drag reduction is an increase in range, fuel savings, increased payload

capacity, and reduced takeoff length. However such improvements only occur at the design lift

coefficient that the winglet has been optimized for. Morphing technology presents the possibility

of actively reconfiguring the outer wing system geometry for the given flight condition to

provide maximum benefit throughout the entire aircraft operational envelope.

1.4.2 Current Morphing Wingtip Research

The concept of wing with a large area change is suitable for small aircraft since the aerodynamic

loading requirements is relatively small. Business jets and commercial transport aircraft

however, require very high aerodynamic loads to achieve flight and with these high loads come

large structural stresses and aeroelastic instabilities. To achieve large shape changes for large

aircraft would require significant actuator power and will increase the structural weight.

In [38], [39], [40] and [41] they investigated the use of variable cant wingtips for morphing

aircraft control. Their concept consisted of a pair of winglets with adjustable cant angle,

independently actuated and mounted at the tips of a baseline flying wing. Computations with a

Page 31: Multidisciplinary Design Optimization of a Morphing

15

vortex lattice model and wind-tunnel tests demonstrate the viability of the concept, with

individual and/or dual winglet deflection producing multi-axis coupled control moments.

Comparisons between the experimental and computational results showed reasonable to good

agreement. However they concluded that a single pair of adjustable winglets cannot substitute for

all the conventional control surfaces at the same time if one wants a full control envelope.

In [42] they proposed the Morphing Winglet (MORPHLET) concept. The objective of the

MORPHLET project was to investigate the use of adaptive materials and structures technology

to dynamically tailor the external shape of the wingtip devices in order to improve multi-phase

mission performance, maneuverability and integrated economics. The twist, cant, and span of

four modules at the wing tip were adjusted to maximize the wings performance at three different

points in the aircrafts mission: start of cruise, start of final cruise and start of descent. A

significant increase in the specific air range was achieved at the cost of a heavier wing structure.

The aerodynamic analysis was performed using a low fidelity vortex lattice approach. In [43]

they performed a multidisciplinary design optimization to the problem of maximizing the

performance benefit through the retrofit of a morphing non-planar outer wing system to a

commercial narrowbody aircraft. Their work involved a two-level combined aero-structural-

control-performance optimization that requires the coupling of aerodynamic and structural

design modules. However their analysis was performed using low fidelity tools which are

sufficient for conceptual design only. They were able to achieve a substantial improvement in

specific air range of 4-6%.

1.5 Thesis Motivation and Overview

Researchers have been inspired by the morphing capabilities of insects and birds that improve

their performance at a variety of flight conditions. Using advanced adaptive structural design and

smart materials, imitating these features for aircraft performance improvement is becoming a

reality. Much of the work being done has been applied to UAVs for military applications. Areas

of particular interest for morphing technology are the wingtips of civil aircraft.

Modern commercial aircraft are optimized only for a narrow range of flight conditions. As a

consequence, off-design conditions results in poor performance. Most civil transport aircraft are

optimized for range but for other flight conditions such as take-off and climb they are used as

Page 32: Multidisciplinary Design Optimization of a Morphing

16

constraints. These constraints could potentially reduce the performance of an aircraft at cruise.

By morphing the wingtip to different shapes, we can alter the load distribution and adapt the

wing performance to the required flight condition.

Most of the work done with the current morphing design from [44], [45] and [46] have focused

largely on either aerodynamics or module kinematics or structural layout. The intent of the

current work is to analyze the combined aerodynamic and structural analyses in a

Multidisciplinary Optimization (MDO) routine. This MDO routine will use either low fidelity or

high fidelity computer models to determine a superior design compared to a baseline aircraft

wing. There are three main objectives:

1. Develop and test an optimization algorithm

2. Integrate aerodynamic and structural discipline codes.

3. Determine the optimal wingtip shape for different MDO architectures

This thesis is organized as follows: the second chapter will describe the VGTM concept to be

used as the morphing mechanism, the third chapter will describe the MDO architecture, the

fourth chapter will describe the PSO algorithm and it will test is performance against several

benchmark functions, the fifth chapter will describe the discipline models used for the low and

high fidelity models, the sixth chapter will present the results of the MDO analysis, and the

seventh chapter will be the conclusion.

Page 33: Multidisciplinary Design Optimization of a Morphing

17

Chapter 2

Variable Geometry Truss Mechanism

2.1 Introduction

The current work does not focus on the design of a morphing mechanism; however it is

important to have an understanding of how the mechanism works. The mechanism that has been

adopted for the current work has been developed at Ryerson University, [44] and [45]. Their

concept follows a Variable Geometry Truss Mechanism (VGTM), which are truss structures that

use active and passive prismatic actuators instead of rigid links to achieve kinematic motion.

This is shown in the figure below.

Figure 2.1 - VGTM like structure (left) and traditional truss structure (right)

VGTM structures are capable of large, rigid displacements and possess high stiffness to mass

ratios [45]. Unlike a traditional VGTM, any truss mechanism used for wing morphing is subject

to considerable size constraints. As wings are narrow at the wingtip, a VGTM would either have

very limited motion, or disrupt the integrity of the wing surface, both of which are undesirable

[44].

Page 34: Multidisciplinary Design Optimization of a Morphing

18

2.2 Kinematics

VGTMs can provide the morphing wingtip with kinematic capabilities and can also support

aerodynamic and structural loading as they are sufficiently rigid. The spar and stringer

arrangement of conventional wingtips is replaced in favor of VGTMs that kinematically connect

airfoil ribs. The truss architecture shown in Figure 2.2 below consists of seven active and passive

kinematic branches.

Figure 2.2 - Base configuration of the modular VGTM morphing wing [45]

The four active branches relate to a specific morphing degree of freedom; qΓ, qb, qΛ, and qα

correspond to the branch actuators that control module cant, span, sweep, and twist, respectively

[45]. That is, the VGTM modules are capable of individual or simultaneous cant, span, sweep,

and twist morphing. The modules also maintain the flexibility to adapt to changes in morphing

requirements through reconfiguration of the base module.

The position and orientation of the module platform relative to the global coordinate system can

be described by:

[ ] (2.1)

Page 35: Multidisciplinary Design Optimization of a Morphing

19

The location of the reference point of the ith

module platform, denoted by , relative to the

global coordinates system, is the summation of the ith module base and the position of the

reference point of the module platform relative to the module base . Before summation,

must be converted to the global reference using the rotation matrix [ ]. The two subscripts

represent the rotation between two coordinate systems attached to two adjacent bodies. For this

case ‘o’ represents the global coordinate system and ‘i’ is the ith module orientation. The

orientation of a reference point relative to the global orientation is the product of the rotation

matrices up to the current coordinate system ‘n’.

∏ ( )

(2.2)

For example if there were two coordinate systems (n=2), shown in Figure 2.3 below, we can

calculate the position of the module platform’s reference point using (2.1).

[ ][ ] (2.3)

Figure 2.3 - Two coordinate systems arbitrarily orientated in 3D space

The rotation matrix uses the desired orientation of the module platforms relative to the base

module as input. These orientations are the dihedral (Γ), twist (θ), and sweep (Λ):

Page 36: Multidisciplinary Design Optimization of a Morphing

20

( ) [

]

( ) [

]

( ) [

]

(2.4)

The product of these three matrices will give the final rotation matrix:

( ) ( ) ( ) (2.5)

2.3 Modular Discretization

The VGTM can be applied in a modular way to meet the kinematic requirements of the

morphing system. Modular discretization has been examined by [45]. Their objective was to

determine the number of morphing wing modules and the respective spacing required to emulate

a known wing shape and satisfy a corresponding flight requirement. The curvature and twist

distribution from the reference wing shape quarter chord line were extracted and used to

determine the spacing of the discretized wing modules. Using ANSYS © Fluent ™, each

modular morphing wing configuration was evaluated for its lift and drag performance. At a

constant angle of attack and incompressible flow, their results showed maximum lift to drag ratio

with seven modules. With additional modules, there was a decline in the lift to drag ratios caused

by adverse flow conditions in regions near the module tips due to the discontinuity between two

adjacent module surfaces [45]. This work was extended to find an optimal wingtip shape for

climb, cruise, and descent flight conditions by [46]. Using the number of modules and the sweep,

span, dihedral, and twist for each module, a Harmony Search algorithm was used to find an

optimal shape for each flight condition (cruise, climb, and descent). A figure of the discretized

wing is shown below used for this work. For the current work the wing will have three main

Page 37: Multidisciplinary Design Optimization of a Morphing

21

components: the main wing, the modules, and the wingtip. The main wing and wingtip are

inactive sections (i.e. no morphing) and modules which are active.

Figure 2.4 - Discretized morphing wing used for the current work

2.4 VGTM Structural Layout

In a conventional wing of a commercial or business aircraft, the combination of spars and skins

typically provides the necessary torsional stiffness, whereas the combination of spars and the

skin-stringers provide the necessary bending stiffness. By replacing the primary structural

elements with a series of linear actuators and passive members that provide only axial stiffness;

it will require a locking and unlocking at an appropriate sequence to achieve the dual purpose of

a structure and mechanism [47]. A simplified representation of the morphing mechanism is

shown below.

Page 38: Multidisciplinary Design Optimization of a Morphing

22

Figure 2.5 - Module structural layout

The passive members exist to provide structural stiffness and the active members provide the

necessary kinematic motion. Figure 2.6 below shows the mechanism shaping the platform airfoil

into different motions.

Page 39: Multidisciplinary Design Optimization of a Morphing

23

Figure 2.6 - Module degree of freedoms

There are three main design issues using conventional VGTMs as an internal mechanism for

morphing wings:

1. The primary structural elements are replaced by actuators which will add weight to the

structure

2. Lack of stiffness due to the geometric restrictions at the wingtip

3. Lack of fault tolerance

To address the second issue, actuators with larger diameters could provide the extra stiffness, but

this of course will add more weight. The third issue could be addressed by adding redundant

actuation but this will again increase weight and it will increase the complexity of the design.

Page 40: Multidisciplinary Design Optimization of a Morphing

24

Chapter 3

Multidisciplinary Design Optimization Architectures and

Problem Definition

3.1 Introduction

Multidisciplinary Design Optimization, MDO, "can be described as a methodology for design of

complex engineering systems that are governed by mutually interacting physical phenomena and

made up of distinct interacting subsystems" [48]. In other words, MDO utilizes optimization

techniques to solve problems that feature various disciplines as shown in Figure 3.1. These

discipline platforms are incorporated simultaneously by passing information between them and

an optimal solution is provided. Since the interaction between disciplines is taken into

consideration, the multidisciplinary solution is superior when compared to the individual

optimization of each discipline. The complexity of engineering problems has sparked a growing

interest in multidisciplinary optimization.

Figure 3.1 - MDO structure for three disciplines

Multidisciplinary Analysis (MDA) is a process by which equilibrium is established by

simultaneously solving state equations of disciplinary analysis. For example aerodynamics and

structural performance (stress, deflection …etc.) are both highly coupled. The deformation of the

Discipline 1

Discipline 2

Discipline 3

OPTIMIZER

Page 41: Multidisciplinary Design Optimization of a Morphing

25

wing structure is dependent on the aerodynamic loading which is in turn dependent on the wings

deformation. This makes the MDA an iterative process. Fixed-point iteration is the easiest to

implement but convergence is dependent on the initial guess at the beginning. Newton iteration

is another approach but it can be computationally expensive with the determination of the

sensitivity derivatives.

3.1.1 Multi-Disciplinary Feasible

Multi-Disciplinary Feasible (MDF) method is the most common way of solving an MDO

problem and it is the easiest to implement. The MDF method works to achieve multidisciplinary

feasibility at each iteration of the optimization. Multidisciplinary feasibility essentially means

that at each iteration all of the design variables satisfy each discipline and there are no conflicting

variables. The MDF method links a single optimizer with a MDA. The optimizer is responsible

for selecting both global and local design variables for the MDA. Achieving a multidisciplinary

feasible design essentially provides the coupling variables to the system. Once the feasibility is

achieved all design variables are known and the analysis calculates the values of the objective

function and the constraints for use in the optimizer. A figure of the MDF architecture is shown

below:

Figure 3.2 - MDF architecture for aeroelastic analysis

Page 42: Multidisciplinary Design Optimization of a Morphing

26

For the above architecture we are using an aeroelastic analysis of an aircraft wing as an example.

Some of the notations have been adopted from [49]. The purpose for MDO is to optimize an

objective function ‘f’ that is dependent on the global and local design variables (XD) which are

the input to the MDA. The objective function is also dependent on the aerodynamic analysis

A(XD) and structural analysis S(XD). The design variables are passed to the disciplines to execute

a MDA, the aerodynamic pressure forces from the wing are turned into structural loads by a

mapping process (MAS), and after the structural analysis the displacements are used to reshape

the aerodynamic grid using another mapping process (MSA). The first output of the MDA is the

convergence of the aerodynamic discipline (A) which is dependent on the design variables XD

and the coupling between the aerodynamic and structure (GAS) discipline which are also

dependent on the design variables and the structural performance. Likewise the second output is

the convergence of the structure (S) which is dependent on the structure and aerodynamic

coupling (GSA). Constraints are imposed on each discipline (CD) such as the limits on the design

variables and maximum allowable stress of the wing structure.

A primary advantage of the MDF architecture is that, at each iteration, multidisciplinary

feasibility is achieved by only treating global and local design variables as optimization

variables, and leaving the coupling variables to be solved by the MDA. Another advantage of

this method is that it is easy to implement and is suitable for small to medium sized engineering

problems. However there are several disadvantages that need to be mentioned. The main

disadvantage of the MDF method is that it is computationally expensive. This is because a MDA

must be performed for every iteration of the optimization and for large engineering problem this

can add a significant amount of computational time. Finding a feasible point for the MDA might

be difficult as well. Depending on the solution method for the MDA a feasible point might take

several iterations or it may not converge at all.

3.1.2 Individual Discipline Feasible

The Individual Discipline Feasible (IDF) method is an alternative to the possibly

computationally expensive MDF method because it avoids a complete MDA upon each iteration.

Instead of achieving multidisciplinary feasible design at every iteration of the optimization, IDF

enforces individual discipline feasibility for each discipline and iteration. A discipline is

considered to be feasible when the equations the discipline code is intended to satisfy are solved

Page 43: Multidisciplinary Design Optimization of a Morphing

27

[49]. For example when the CFD model converges to find the pressures given an input shape, the

aerodynamic discipline is considered to be feasible. Likewise for the FE model, when the stress

and displacements are determined the structural discipline is considered to be feasible. The IDF

method only achieves a complete multidisciplinary feasible design when the solution has

converged.

The architecture links a single optimizer and a decoupled disciplinary analysis for each

discipline. The specific analysis variables that have been promoted are those that represent

communication, or coupling, between analysis disciplines via interdisciplinary mappings. This

changes the MDF architecture by forcing the introductions of surrogate design variables that will

allow the coupling variables to be explicitly optimized by the optimizer [49]. The IDF

architecture is shown below.

Figure 3.3 - IDF architecture for aeroelastic analysis

The extra constraints must be included in order to drive the optimization to a feasible design

upon convergence. The optimizer selects the surrogate design variables (XMAS and XMSA) and the

global/local variables (XD). These are then introduced to each discipline for analysis where the

actual values of the coupling variables can be calculated. These actual coupling variables are the

additional equality constraints (GAS for aerodynamics and GSA for structure). XMAS is the

Page 44: Multidisciplinary Design Optimization of a Morphing

28

surrogate design variable for the aero-structural coupling and XMSA is the surrogate design

variable for the structural-aero coupling. These surrogate variables could be the coefficients of a

spline curve or a spline surface. For example the surrogate design variables for the structural

discipline could use spline coefficients to create the aerodynamic loading for the FE model.

The advantage of this method is that a complete MDA is avoided upon each iteration which will

improve computational time. However, this is not always the case since satisfying the coupling

variables equality constraint can be difficult and there is an increase in the problem dimension.

3.1.3 All-At-Once

Another interesting MDO architecture is the All-At-Once (AAO) method. It is the exact opposite

extreme of the MDF method. For the AAO, we do not seek to obtain feasibility for the analysis

problem. In other words discipline feasibility, multidiscipline feasibility, and equations within

the discipline itself are not satisfying feasibility. For example an uncoverged CFD or nonlinear

FE model are considered infeasible. This approach attempts to reduce the computational time by

avoiding feasibility when the design is far from the optimum. The problem with this approach is

that it introduces several constraints and increases the dimension of the problem. The AAO

approach will not be used in the current work.

3.1.4 Additional MDO Architectures

Table 3.1 below list several common MDO frameworks with a brief description of the technique

with some of the advantages and disadvantages associated with each approach. Each method has

been summarized from [50].

Table 3.1 - Common MDO frameworks

Framework Description Advantages Disadvantages

Multidisciplinary feasible

(MDF)

This approach couples an

optimization code to the

MDA process.

Excellent for problems

with strong coupling. State

equations are satisfied

during the optimization.

Computationally

expensive. Disciplinary

code integration can be

difficult.

Individual discipline

feasible (IDF)

Coupling variables are

treated as optimization

Only one cycle of the

disciplinary analysis code

Increase in problem

dimensionality. Coupling

Page 45: Multidisciplinary Design Optimization of a Morphing

29

variables and auxiliary

variables are used to

satisfy individual

disciplines

is required. variables may not satisfy

the physics of the problem.

Disciplinary analysis

optimization (DAO)

Each discipline chooses its

own system level auxiliary

variables in disciplinary

analysis. Convergence is

ensured by placing

equality constraints on the

system variables.

Ensures disciplinary

constraints are satisfied at

each iteration of the

optimization cycle.

Increased problem

dimensionality. Several

equality constraints must

be satisfied.

All-at-once (AAO) The state equations and

optimization problem are

solved simultaneously.

State equations are

considered equality

constraints on the system

level optimization.

Eliminate iterative

disciplinary analysis.

Reduced computation cost.

Difficult to satisfy state

equations’ equality

constraints. Difficult to

obtain a global optimum.

Collaborative

Optimization (CO)

The optimization problem

is solved at two levels:

system level and discipline

level. Each discipline

selects its own set of

design variables and then

each discipline-specific

optimization minimizes

the difference between

discipline-chosen design

variables and design

variables provided by the

system-level optimizer

Independent execution of

disciplines allows for

concurrent processing of

disciplinary optimization.

Subsystem optimization

could have a reduced

dimensionality. Each

discipline has its own

autonomy to select its own

optimization process.

System-level constraints

can have discontinuous

derivatives making this

difficult to solve with

gradient optimizers. Large

number of coupling

variables can lead to

excessive computational

costs and convergence

difficulties.

Concurrent subspace

optimization (CSSO)

The system-level is treated

as another discipline and

each discipline has its own

optimizer. This method

Subsystem and system-

level are solved

independently from each

other reducing

Uses approximation

models for subsystem and

system-level optimization.

Approximation models are

Page 46: Multidisciplinary Design Optimization of a Morphing

30

uses decomposition if the

design variables and

approximations so as to

ensure multidisciplinary

fidelity in each step of the

optimization problem.

computational time. accurate within small

regions surrounding the

point of interest.

Bilevel integrated system

synthesis (BLISS)

Formulates the

disciplinary objective

function from the given

system-level objective

function such that

minimization of the

disciplinary objective

function will result in the

reduction of the system-

level objective function

value.

System-level optimization

process guides subsystem-

level optimization to

improve overall system

design. Reduction of the

optimization problem to a

subsystem-level

optimization process

reduces problem

dimensionality

Individual disciplines do

not have autonomy to

select their own objective

function. Difficult to solve

using gradient optimizers

Disciplinary interaction

variable elimination

(DIVE)

Reduce optimization

complexity by elimination

certain variables. Each

discipline constructs a

metamodel.

Variable elimination

reduces model dimension.

Construction of

metamodels can be

computationally

expensive.

Which method to use will be highly dependent on the problem to be solved, problems with high

coupling between disciplines and separation of these disciplines proves to be difficult, the MDF

approach will be the most appropriate but more computationally expensive.

3.2 MDO Test Cases

The overall objective of the VGTM concept is to find an optimal wing shape for each mission

leg of an aircraft’s flight. A simple mission is presented in Figure 3.4. Each circle represents a

shape change in the wing. The hollow circle represents the start of a mission leg and a filled

circle represents a shape change during its respective mission leg. More specific details on each

mission leg will be given later. A morphing wing should change its shape during each mission

leg since the weight of the aircraft is decreasing due to fuel being consumed. Conventional wings

are optimized for beginning of cruise but at subsequent intervals during the mission legs the

Page 47: Multidisciplinary Design Optimization of a Morphing

31

wing will be suboptimal. The work done in [51] has shown that a continuous shape changing of a

morphing wing concept during each mission leg will provide the best results. Using a multi-level

optimization routine, they broke down each mission leg into different segments and optimized

the wings performance to reduce the fuel consumption. They determined that an aircraft whose

wing area is shrinking, they referred to this as “cruise-shrink”, will give optimal results.

Take-Off

Cruise Loiter

Descent

Land

Climb

Figure 3.4 - Aircraft mission showing mission legs

Table 3.2 below shows a list of possible objectives for each mission leg.

Table 3.2 - Possible objective for each mission leg

Mission Leg Possible Objectives

Take-Off Minimize take-off distance

Minimize noise

Climb Maximize rate of climb

Maximize climb angle

Cruise Maximize range

Minimize fuel consumption

Loiter Maximize endurance

Descent Maximize sink rate

Land Minimize landing distance

Page 48: Multidisciplinary Design Optimization of a Morphing

32

Minimize stall speed

3.3 Objective Function

Attempting to perform an optimization for all of the mission legs simultaneously would require

significant computational resources. Commercial and business jet aircraft will devote most of

their flight time to cruise. Therefore it will be selected as the dominant mission leg that should be

optimized. For commercial aircraft, the fuel totals amount 38% of operations and maintenance

costs [31, p. 570]. With rising fuel costs, aircraft designers have been motivated towards

improving aircraft fuel efficiency. Since business jets or commercial aircraft will consume most

of its fuel during cruise, minimizing the fuel consumed for this mission leg was selected to be the

objective. Maximizing range are other objectives that can be considered, however the amount of

fuel consumed will need to be known aprioi, and since aircraft are typically designed with a

range requirement with fuel efficiency in mind, minimizing fuel would be a better option.

The objective of the low and higher fidelity test cases is to minimize the fuel consumption for a

known thrust specific fuel consumption (ct) and range (R) at a given flight speed (V). The range

for an aircraft is given below [31]:

(

) (3.1)

L is the lift, D is the drag, Wi is the fuel at the end of cruise and Wi-1 is the fuel at the beginning.

Since we are interested in minimizing fuel, the equation can be rewritten as:

( ) (3.2)

From the equation above, if we maximize the velocity and lift, and minimize the specific fuel

consumption, range and drag, we will minimize fuel consumption. This equation is

contradictory: high speed means high drag, high lift means high drag, and high drag means high

thrust (higher specific fuel consumption). However, for the optimization problems below the

specific fuel consumption, Mach number, and range will be held constant at 2e-4 /s, 0.85, and

6000 km respectively. With these constants, the only variables that will influence the objective

are the lift and drag.

Page 49: Multidisciplinary Design Optimization of a Morphing

33

With multiple morphing transitions during cruise, there is now a multi-objective problem. The

objective will be to minimize the fuel consumed for each morphing stage at cruise.

(3.3)

The multi-objective problem will follow a weighted sum approach. With multiple morphing

states at cruise, a weight of 1/N has been implemented (with N being the number of morphing

states). Including this weighting to the fuel objective will give an equal weight of 1. An uneven

weight will introduce bias to the objective with the large weighting. Also each objective will

need to be normalized with respect to a reference value to prevent bias.

3.4 Constraints

There are two intended purposes of constraints: the first is to provide limitations to the

optimization algorithm to make it more efficient and the second is to satisfy a design

requirement. For this MDO routine upper and lower limits have been specified for all design

variables. This constraint forces the PSO algorithm to search in a region where the optimal

design will most likely be found. Constraining the bending stress to be below the yield stress

with a factor of safety is a design requirement to ensure a safe wing structure. The material

properties for the wing are Aluminum 2014-T6 (Young’s Modulus is 73.1 GPa, Yield Stress ( )

414 GPa, and density 2790 kg/m3).A factor of safety (FS) of 1.5 was used.

(3.4)

A constraint has been imposed on the fuel consumption objective. This was done for two

reasons: the first is a design requirement because we want an improved performance compared to

a baseline wing and the second reason was to improve algorithm convergence. This is illustrated

in the figure below.

Page 50: Multidisciplinary Design Optimization of a Morphing

34

Figure 3.5 - Design space without constraint on objective (left) design space with constraint

on objective (right)

The figure shows an objective (F) with global and local minima with only one design variable

(V). The left figure is without an objective constraint. This constraint absence increases the size

of the feasible design space for the algorithm to search and it may end up getting stuck in the

local minima. The algorithm may find the global optimum eventually, but it will likely require

several more iterations. On the other hand, if we introduce a constraint on the objective we can

encourage exploration closer to the global optimum. This figure is meant for illustrative purposes

only.

(3.5)

The objective above is directly influenced by the aerodynamic discipline only. Therefore to keep

the mass of the structure within a reasonable limit a constraint will also be imposed on the

structural mass. This constraint will confine the mass of the wing to be below the mass of the

reference wing. However, to give some leeway to the optimizer the reference wing mass will be

increased by 10%.

(3.6)

Page 51: Multidisciplinary Design Optimization of a Morphing

35

In order to achieve discipline feasibility for aircraft performance, steady level flight conditions

must be satisfied (i.e. lift must equal weight).

|

| (3.7)

The aircraft needs to be trimmed using the angle of attack to find this equality constraint. This

constraint is necessary to find an equilibrium state to satisfy the aircraft performance discipline.

Depending on the method (IDF or MDF) used satisfying this constraint can be difficult, so a

tolerance (tol) of 0.01 was used. This means that the lift must be within 1% of the weight

estimation. For the IDF approach the angle of attack is a variable used for trimming. However

for the MDF a different approach is used. Since the lift for an aircraft at low angles of attack

(less than 8o) is relatively linear, an interpolation/extrapolation scheme can be used to find the

trim state. Using this approach, a trimmed state can be found in only three function calls of the

aerodynamic model as opposed to a fixed-point iterative scheme which could take several more

function calls. However there is no guarantee that achieving a trimmed state using this approach

is possible, so the trimmed constraint above is still important.

3.5 Design Variables

The number of design variables for the high and low fidelity models are dependent on the

number of morphing transitions (N) during cruise. If there is only one morphing state at the

beginning of cruise, there are a total of 31 variables. If the number of morphing states were to

increase to two, the number of variables becomes 46.

Table 3.3 – Design variable list for low fidelity models

Morphing Stage Wing

Component Discipline Design Variable

Upper

Limit

Lower

Limit

Initial Stage Main Wing Aerodynamic

Span (m) 6 9

Sweep (o) 40 20

Taper 0.3 0.1

Page 52: Multidisciplinary Design Optimization of a Morphing

36

Chord (m) 5 3

Main

Structure

Leading Edge

Offset (%chord)

0.3 0.1

Trailing Edge Offset

(%chord)

0.9 0.6

X Scale (Root and

Tip)

0.99 0.7

Y Scale (Root and

Tip)

0.99 0.7

Wingtip

Aerodynamic

Taper 1 0.3

Sweep (o) 20 -20

Span (m) 1 0.5

Twist (o) 5 -5

Dihedral (o) 20 -5

Structure

X-Scale 0.99 0.7

Y-Scale 0.99 0.7

VGTM (3 Mods)

Aerodynamic

Span (m) 0.5 0.25

Twist (o) 5 -5

Dihedral (o) 20 -5

Sweep (o) 20 -20

Taper 1 0.7

Structure

X Scale 0.99 0.7

Y Scale 0.99 0.7

Page 53: Multidisciplinary Design Optimization of a Morphing

37

Flight Condition Performance Angle of Attack (o) 3 -3

Additional

Stages

VGTM (3 Mods) Aerodynamic

Span (m) 0.5 0.25

Twist (o) 5 -5

Dihedral (o) 20 -5

Sweep (o) 20 -20

Flight Condition Performance Angle of Attack (o) 3 -3

Total Number 16+N*15

The variables for the higher fidelity case are slightly different from the low fidelity case and are

listed in the table below. From the previous low fidelity case all the variables were continuous.

For the high fidelity case two discrete variables were included. The number of stringers and ribs

are considered to be discrete. To implement the discrete variables in the PSO algorithm a simple

rounding of the number being passed to the discipline code takes place.

Table 3.4 - Design variable list for high fidelity models

Morphing

Stage

Wing

Component Discipline Design Variable

Upper

Limit Lower Limit

Initial Stage Main Wing

Aerodynamic

Span (m) 9 7

Sweep (o) 40 20

Twist (o) 8 2

Taper 0.3 0.1

Chord (m) 5 3

Main

Structure

Ribs 24 16

Stringers 8 3

Skin Thickness (m) 0.025 0.015

Page 54: Multidisciplinary Design Optimization of a Morphing

38

Rib Thickness (m) 0.01 0.001

Spar Thickness (m) 0.01 0.001

Stringer Radius (m) 0.01 0.001

Spar Location (m) 0.9 0.7

Structural Taper 0.6 0.2

Wingtip

Aerodynamic

Span (m) 1 0.25

Taper 1 0.7

Structure Spar Thickness (m) 0.008 0.001

VGTM (3 Mods)

Aerodynamic

Sweep (o) 15 -15

Span (m) 0.75 0.25

Twist (o) 5 -5

Dihedral (o) 20 0

Structure

Passive Member Radius (m) 0.01 0.001

Active Member Radius (m) 0.01 0.001

Skin Thickness (m) 0.008 0.001

Rib Thickness (m) 0.008 0.001

Flight Condition Performance Angle of Attack (o) 6 0

Additional

Stages VGTM (3 Mods) Aerodynamic

Sweep (o) 15 -15

Span (m) 0.75 0.25

Twist (o) 5 -5

Dihedral (o) 20 0

Page 55: Multidisciplinary Design Optimization of a Morphing

39

Flight Condition Performance Angle of Attack (o) 6 0

Total Number 23+15*N

Page 56: Multidisciplinary Design Optimization of a Morphing

40

Chapter 4

Particle Swarm Optimization Algorithm

4.1 Introduction

Particle Swarm Optimization (PSO) was developed by [52] as a stochastic optimization

algorithm based on social simulation models. The PSO algorithm uses a swarm of particles to

travel stochastically in the search space. The velocity vector is used to update the particles

current position through

(4.1)

The velocity update vector is given by (4.2).

(

) (

) (4.2)

The velocity of the ith

particle with the jth

dimension is updated based on the information from

previous time step t and an inertia weight constant given by w from [53] [54] [55]. The inertia

weight was introduced to help concentrate the swarm near the promising regions. Values greater

than 1 encourage the swarm to explore the design space which is desirable for the early stages of

optimization. Smaller values reduce the velocity step to exploit the best solutions. and are

uniformly distributed random numbers between 0 and 1. is the best position of the particle that

it has ever visited is held in memory. In other words it is the best position or combination of

design variables that yielded the best objective. In addition to the personal best position there is a

global best . This represents the best position out of all of the particles at the current iteration.

The best positions are communicated to the entire population with the intent to encourage the

particles to search the more promising regions in the search space. and are the cognitive and

social learning factors respectively (accelerations constants) supplied by the user. The values

associated with the two learning factors influence how much confidence the particle has in itself

(cognitive) or in the swarm (social). If is less than then the particles will be attracted to the

global best solution and this global best position may not be the objectives global minimum

solution, and a premature convergence is possible. If is greater than the swarm will be more

Page 57: Multidisciplinary Design Optimization of a Morphing

41

biased towards itself and convergence may take several iterations. Also the two constants control

the magnitude of the velocity vector step size. After the position has been updated using (4.1) it

is necessary to find the new personal best and global best. If the personal best is smaller than the

previous personal best then the personal best is updated to the current value. Otherwise the

personal best is not updated. Each value of the particles new personal best is compared to the

previous global best. If the new particle’s personal best is smaller than the current global a new

global is created.

4.2 Literature Review

Several enhancements have been used to improve the basic algorithm. Using a constant value of

inertia weight showed improvement to the PSO algorithm, however linearly varying the inertia

weight from 0.9 to 0.4 has shown substantial improvement in performance compared to a

constant inertia weight [54]. A dynamic approach to changing the inertia weight was used in [56]

and [57]. The advantage of a dynamic reduction approach is the fewer function evaluations

necessary for a satisfactory solution. This has been shown in [58] with a 10 and 25 bar truss

example. The Time Varying Acceleration Coefficient (TVAC) proposed from [59], for the early

stage of the optimization having a larger initial value for the cognitive over the social allows the

particles to have more confidence in its exploration. For the later stages of optimization, to

facilitate exploitation having a larger social factor would be more beneficial to obtain the global

optimum. A mutation operator was also introduced to improve swarm diversity.

A slight modification to the standard PSO algorithm has been introduced. The social and

cognitive acceleration constants will not vary with time but will vary with the particle elitism.

The intent of this approach is to try and jostle a particle free of a local optimum. A similar

approach has been used in [60], where they used the particles local and global best values to

determine their acceleration constants. c1 and c2 were assumed to be the same. They were able to

successfully reduce the average number of iterations and the average fitness for several

benchmark problems. Their results were improved further by incorporating mutation, crossover,

and root mean square variant to tune theirs results. In [61] they applied a self-adaptive inertia

weight (SAIW) and social acceleration coefficient (SAIC) PSO. The SAIC was tested against

high dimensionality Rastrigin, Sphere, and Griewank functions. The results were excellent for

Page 58: Multidisciplinary Design Optimization of a Morphing

42

the Griewank and Sphere test cases but were poor for the Rastrigin; the authors incorporated a

mutation operator from [62] to their algorithm to maintain diversity.

4.3 Social and Cognitive Parameter Variations

For this section, the three different social and cognitive parameter variations will be described.

The three variations will follow a linear, step, or sinusoidal trend.

4.3.1 Linear variation of acceleration constants

The cognitive and social parameters will follow a linear distribution from the best to worst

particle. Equations 4.3 and 4.4 are the cognitive and social acceleration constants respectively.

c1best and c1worst are the cognitive parameter for the best and worst particle with c2best and c2worst

are the social parameters for the best and worst particle. N is the number of particles in the

swarm.

( )( )

( ) (4.3)

( )( )

( ) (4.4)

For the above equations the ith

particle at the current iteration is given a rank given by j based on

its eliteness. The best partcle is identified with an index of 1 and a worst particle has a rank of N

which is the swarm size. A sample distribution is given in Figure 4.1 with c1best = c2worst = 1 and

c1worst = c2best = 3.

Page 59: Multidisciplinary Design Optimization of a Morphing

43

Figure 4.1 - Linear variation of particle acceleration constants

A large value for c1best over c2best promotes exploration and may require many function

evaluations to converge, however with greater social attraction exploitation of the best position is

preferred but may force a premature convergence. Values of c1worst which are greater than c2worst

encourage exploration of the particles such as the second half of the ranked swarm in Figure 4.1.

4.3.2 Step variation of acceleration constants

In addition to the linear variation described previously, a step variation will also be examined.

The step variation separates the swarm into three distinct groups: upper, middle, and lower

groups. Figure 4.2 below is an example of the step distribution. For this example the upper group

(U) has a larger cognitive learning capability compared to its social forcing a larger step towards

its own personal best, the middle group (M) is an average of the social and cognitive parameters

(this is just the SPSO algorithm), the lower group (L) has a higher social learning than cognitive

forcing the particles to take a larger step towards the more promising regions found in the search

space.

Page 60: Multidisciplinary Design Optimization of a Morphing

44

Figure 4.2 - Step variation of particle acceleration constants

4.3.3 Sinusoidal variation of acceleration constants

Another variation being considered is a sinusoidal variation of the particles acceleration

constants. Figure 4.3 below shows an example distribution determined from equations 4.5 and

4.6. ω is the frequency, is the phase, and j is the particle rank index.

(

) (4.5)

(

) (4.6)

Page 61: Multidisciplinary Design Optimization of a Morphing

45

Figure 4.3 - Sinusoidal variation of particle acceleration constants

For the three different social and cognitive variations described, the parameters will be

determined statically and dynamically. For the static case, the social and cognitive parameters

will be determined by the initial swarm and will remain constant throughout the optimization

process. However for the dynamic case, the social and cognitive parameters will be revaluated at

each iteration.

4.4 Unconstrained Test Cases

Four benchmark functions listed in Table 4.1 will be evaluated for the current work. The

Rastrigin, Griewank, and Rosenbrock problems are multimodal equations with several local

minima thus making them a good test for the algorithms exploration capability. The Sphere

problem is unimodal used to test the exploitation capabilities of the algorithm. Each function has

a global optimum of 0.

Table 4.1 - Unconstrained benchmark problems

Function Limits Vmax

Rastrigin ( ) ∑ ( ( ) )

10

Page 62: Multidisciplinary Design Optimization of a Morphing

46

Rosenbrock ( ) ∑ (

) ( ) 30

Sphere ( ) ∑

50

Griewank ( ) ∑

∏ (

√ )

600

Each function will be used to compare the standard PSO algorithm (SPSO) and the constriction

based PSO (CPSO) [63] to the proposed methods. Each benchmark problem will have 100 trial

runs, a dimension of 30, and a swarm size of 30. The criteria for comparison include: the mean

and standard deviation after 3000 iterations have been completed. All test cases will have a time

varying inertia weight (TVIW) from 0.9 to 0.2 (with the exception of the CPSO). The TVIW

uses the specified number of iteration to change the inertia weight in a linear fashion. A large

inertia weight is beneficial for exploration at the early stage of optimization but a lower value is

best to exploit the more promising regions in the design space. The initial population is randomly

generated within the specified bounds on the design variables. These limits on the design

variables confine the algorithm to a specified search space to improve algorithm efficiency. Also

each particle receives an initial randomly generated velocity confined to a maximum velocity.

The maximum velocity will limit the step size of the particle. If the velocity is too large the

algorithm will not converge. The results for each benchmark problems are presented in appendix

A. From the results it was concluded that a sinusoidal variation demonstrated the best

performance

4.5 Constrained Test Case Results

The previous section dealt with only unconstrained problems. Most engineering design

optimization problems require at least some constraints to be imposed. To handle constraints for

the PSO algorithm, an exterior quadratic penalty function approach will be used.

∑ { }

(4.7)

Page 63: Multidisciplinary Design Optimization of a Morphing

47

This equation states that the original objective is penalized with a penalty factor P set to 1e7

and the summation of all the violated constraints . A constraint is considered to be violated if

its value is greater than zero. The penalized objective is sent to the optimizer. In this section

we will use a straight forward structural optimization problem to analyze the modified PSO

algorithm’s performance. The problem description is presented next.

4.5.1 Problem Description [64]

The 25-bar space truss shown in the figure below is required to support two load conditions

given in Table 4.2. Constraints have been imposed on member stresses as well as Euler buckling.

A minimum and maximum allowable area is specified for each member. The allowable stresses

for all members are specified as σmax (40e3 psi) in both tension and compression. The Young’s

Modulus is 107 psi and the density (ρ) is 0.1 Ib/in

3.

Figure 4.4 - 25-bar space truss

Table 4.2 - Load conditions for the 25-bar truss

Joint

Page 64: Multidisciplinary Design Optimization of a Morphing

48

Load 1 2 3 6

Fx (1) 0 0 0 0

Fy (1) 20e3 -20e3 0 0

Fz (1) -5e3 -5e3 0 0

Fx (2) 1e3 0 500 500

Fy (2) 10e3 10e3 0 0

Fz (2) -5e3 -5e3 0 0

The members are assumed to be tubular with a nominal diameter/thickness ratio of 100, so that

the buckling stress in member i becomes

(4.8)

Where Ai and li are the cross-sectional and length, respectively, of member i. The cross-sectional

areas linked as:

A1, A2=A3=A4=A5, A6=A7=A8=A9,

A10=A11, A12=A13, A14=A15=A16=A17,

A18=A19=A20=A21, A22=A23=A24=A25

(4.9)

The objective of the problem is to minimize the weight of the 25-bar space truss

(4.10)

With the following constraints:

Page 65: Multidisciplinary Design Optimization of a Morphing

49

| |

(4.11)

Where σi,j is the stress induced in member i under load condition j. σi,j must also be less than the

critical buckling stress pi.

4.5.2 Constrained Structural Optimization Results

The optimal mass is 233.07 Ib with the design vector X=[0.1, 0.8023, 0.748, 0.1, 0.124, 0.571,

0.978, 0.802]. The test cases that are being compared are in Table 4.3.

Table 4.3 – Test cases for 25-bar space truss

Test Case

1 Standard C1=2, C2=2 8 Sinusoidal ω=2π and φ=π/2

2 Linear

C1best=C2worst=3

C1worst=C2best=1

9 Sinusoidal ω=4π and φ=π/2

3 Linear

C1best=C2worst=1

C1worst=C2best=3

10 Sinusoidal ω=6π and φ=π/2

4 Step

0≤U≤20 C1=3 C2=1

11 Sinusoidal ω=8π and φ=π/2 20≤M≤60 C1=2 C2=2

60≤L≤100 C1=1 C2=3

5 Step

0≤U≤20 C1=1 C2=3

12 Sinusoidal ω=2π and φ=0 20≤M≤60 C1=2 C2=2

60≤L≤100 C1=3 C2=1

6 Step 0≤U≤40 C1=3 C2=1 13 Sinusoidal ω=4π and φ=0

Page 66: Multidisciplinary Design Optimization of a Morphing

50

40≤M≤80 C1=2 C2=2

80≤L≤100 C1=1 C2=3

7 Step

0≤U≤40 C1=1 C2=3

14 Sinusoidal ω=6π and φ=0 40≤M≤80 C1=2 C2=2

80≤L≤100 C1=3 C2=1

15 Sinusoidal ω=8π and φ=0

Each test case will be run 20 times to find the mean and standard deviation, have a swarm size of

20, a maximum iteration count of 400, and will use the TVIW. The performance metric used to

compare the test cases are the mean number of iterations/fitness required to achieve the target

weight of 233.07 Ib and the iteration/fitness standard deviation. The results are presented in

Table 4.4.

Table 4.4 - Constrained test case results

Test

Case

Mean

Fitness

STD

Fitness

Mean

Iteration

STD

Iteration

Test

Case

Mean

Fitness

STD

Fitness

Mean

Iteration

STD

Iteration

1 239.4 13.12 323.8 81.14 8 233 0.07275 278.3 26.22

2 238.4 12.8 310.4 70.38 9 233 0.05197 272.5 11.65

3 249.3 23.9 358.5 67.21 10 251.4 30.5 362.9 60.43

4 233.7 2.033 284.1 48.38 11 240.9 14.97 313.2 62.1

5 258.7 46.01 317.3 65.66 12 243.3 13.92 366.5 60.94

6 253.3 55.21 336.1 68.68 13 247 27.37 316.8 72.48

7 238.2 11.22 344.5 62.45 14 246.1 23.91 323.1 82.06

15 246.8 33.34 348.2 67.14

Page 67: Multidisciplinary Design Optimization of a Morphing

51

From the results above, test cases 8 and 9 achieved the best performance. The results were

consistent, since the average fitness was exactly the target weight of 233.07 Ib, with a very low

standard deviation. Test case 9 had the lowest number of iterations and the lowest standard

deviation. Therefor the 9th

test case will be used for the MDO routine.

Page 68: Multidisciplinary Design Optimization of a Morphing

52

Chapter 5

Aerodynamic and Structure Computer Models

5.1 Introduction

Selecting an appropriate discipline tool for the MDO process cannot be taken lightly. The

discipline tool needs to be accurate enough to model the physics of the problem. Complicated

models such as a non-linear finite element analysis can be more accurate for certain problems

compared to a linear model but the computational requirements will be a lot higher. A balance

between computational efficiency and model accuracy is required. This chapter will provide a

description of the low and high fidelity models being used for the current work.

5.2 Low Fidelity Models

Low fidelity models are typically used in the very early conceptual design phase. They are

simple models based on approximations and other simplifying assumptions. These models are

selected for their computational efficiency over their accuracy. This becomes very important for

optimization problems with a large number of function calls. This section will describe two low

fidelity approaches under consideration. A mid-sized business jet wing has been selected for

testing. Its geometric description has been listed below.

Table 5.1 - Reference wing geometry

Bombardier Challenger 300

Maximum Take-Off

Gross Weight (MTGW)

18000 kg Wing Reference Area (S) 40 m2

Cruise Mach Number 0.85 Wing Root Chord (Cr) 4 m

Cruise Altitude 10 km Taper Ratio (λ) 0.2

Wing Span (b) 18 m

Page 69: Multidisciplinary Design Optimization of a Morphing

53

5.2.1 Aerodynamic Model

The aerodynamic tool used for the current work is the well-known Tornado MATLAB code

which has been developed by Dr. Tomas Melin [65]. The Tornado code is a vortex lattice

method (VLM), programmed to be used in conceptual aircraft design. The aircraft geometry in

Tornado is fully three dimensional with a flexible, free-stream following wake. The Tornado

solver solves for forces and moments, from which the aerodynamic coefficients are computed.

Although Tornado is considered a low fidelity tool it has been selected for the current work for

its computational efficiency. The VLM is fast, which is a requirement of design studies that

entail a large number of function evaluations, such as the one here. Tornado is used for the

calculation of relevant aerodynamic coefficients and lift distribution over the wing. This lift

distribution is necessary for the structural model.

Tornado is a low fidelity model because it is based on potential flow theory which requires

several simplifying assumptions: incompressible, inviscid, irrotational, and airfoil thickness is

neglected. For a real aircraft at transonic flight these simplifying assumptions are completely

incorrect. At transonic speeds the air is compressible meaning its density will change with

pressure significantly. This rise in density will increase drag and lift. Failing to consider

compressibility effects for structural design may result in structural failure. Also aircraft

performance can be severely over estimated. An inviscid flow model ignores the drag effects

associated with the boundary layer on the wing surface. Ignoring the viscous effects will give an

optimistic drag prediction. An irrotational model assumes the flow remains attached to the wing

at all times. At high angles of attack the air will separate from the surface creating an increase in

drag. The airfoil thickness is strongly influenced by the Mach number. A thick airfoil will create

very large velocity gradients making such a design suitable for low speed flight. However at

transonic speeds, to avoid or reduce the presence of shockwaves, a thinner airfoil is preferred.

Tornado has the option to use correction factors and approximate models to address some of the

simplifying assumptions. The Prandlt-Glauret compressibility correction factor can be applied to

the aerodynamic coefficients to estimate the effects of transonic flight. This compressibility

correction is performed by adjusting the flows incompressible coefficients using:

Page 70: Multidisciplinary Design Optimization of a Morphing

54

√ (5.1)

The compressibility correction adjusts the coefficient (C) using the Mach number to obtain the

new compressibility coefficient (Cc). Using correction factors Tornado can estimate the zero-lift

drag of the wing. This drag is often referred to as the profile drag, which is determined when the

lift of the aircraft is zero. Using the coefficient of drag correction factor and the wing wetted area

the zero-lift drag coefficient for a wing section can be calculated.

In [66] Tornado was verified against wind tunnel data from an elliptical wing. At Mach numbers

of 0.3, 0.4, and 0.5, the lift slopes from the experimental and Tornado models were compared

and the errors were shown to be 0.9%, 0.2%, and -2% respectively. Slightly larger errors were

determined when comparing drag polars due to the absence of viscous drag effects. For the three

different Mach numbers there errors were -4%, -5%, and -4% respectively. Tornado has also

been used by the following papers [43], [46], [51], and [67]for their work.

5.2.2 Structural Model

The structural model of the wing is a 3D cantilever beam using linear finite elements. This 3D

beam model has 6 degrees of freedom at each node. The stiffness matrix can be found in several

textbooks on finite elements, but this was taken from [68]. The stiffness matrix in its current

form is fine for a beam that is straight, however the wing has a sweep angle and the wingtip

modules will have three different orientations.

The cross-section of the wing is shown in Figure 5.1 below. It is a simple beam that has a

polynomial shape confined inside an airfoil. This cross-section is just an approximation of a real

wingbox. A real wingbox can have several design variables and require more sophisticated

computer models, however it was selected for the following reasons; it requires only two design

variables, and its bending and torsion stiffness are dependent on the size of the wing. To create

the thickness of the wingbox a simple 2D scaling approach is used. The interior shape has its

coordinates defined by scaling the x and y coordinates of the exterior shape.

Page 71: Multidisciplinary Design Optimization of a Morphing

55

Figure 5.1 - Wingbox geometry confined within airfoil

The area moment of inertia and polar moment of inertia are defined respectively as:

∑(

)

( )

∑(

)

( )

(5.2)

The above equations are for a solid polygon shape only; however the cross-section of the wing

being considered is hollow. There are two large polynomials that can be used to define the cross-

section; an exterior and an interior. To find the overall area moment of inertia a simple

subtraction of the two polygons is possible:

(5.3)

Page 72: Multidisciplinary Design Optimization of a Morphing

56

Finding the polar moment of inertia is a bit more involved. We need to find the radial location ‘r’

of the centroid of each polygon relative to the origin of the cross-section and ‘A’ is the area. In

this case the polygons refer to the shapes defining the thickness of the cross-section. To find the

centroid and area of the polygons we use:

∑( )

∑( )

( )

∑( )

( )

(5.4)

To validate the above model, a sample section was created in ANSYS© to determine the

necessary cross-sectional properties. The results are compared in table Table 5.2.

Table 5.2 - Comparing cross-sectional data from ANSYS to current FE model (X and Y

Scaling are set to 0.5)

Structural Property ANSYS Current model

2.9714e-05 3.0576e-005

0.00171 0.0018

0.0405 0.0406

1.0904e-04 1.4093e-004

-0.043704 -0.0375

-0.0045851 -0.0039

Page 73: Multidisciplinary Design Optimization of a Morphing

57

To validate the FE solver example problems from [68] were tested. Two examples were chosen:

the first example was used to validate the frame setup with a distributed load in 2D and the

second example was chosen to assess its ability to model in 3D. The example problems are in

appendix B. The solver was able to replicate the results from [68].

5.2.3 Aero-Structural Model

To calculate the structural displacements and bending stress, the aerodynamic loads from

Tornado needs to be mapped to the finite element model. Instead of using equations 2.1 and 2.2

from chapter 2, the structural nodes will be determined from the Tornado model. This will ensure

model consistency between the two disciplines. The location of the beam centroid will be aligned

with the axis, which is the midpoint between the leading and trailing edge offset.

Figure 5.2 - FE nodes (dots) with aerodynamic model

To map the aerodynamic forces to the finite element model, the technique of equivalent nodal

forces have been used [68]. A beam can support distributed or concentrated nodal loading. In

order to account for these loading conditions we need to find an equivalent nodal load. From

Tornado the aerodynamic forces are concentrated nodal loads located between the FE nodes.

Page 74: Multidisciplinary Design Optimization of a Morphing

58

Using the equivalent nodal loads for a beam shown in Figure 5.3, it is possible to determine the

structural displacement and stresses.

Figure 5.3 - Equivalent nodal loads for concentrated point loads

To find the stress in the beam, we need to know the reaction forces and moments. Once the

reactions are known using the following equation it is possible to determine the bending stress:

(5.5)

The stress is dependent on the reaction moment M, the second area moment of inertia Ix and the

height of the wing section t. A wing deflection compared to its initial state is shown in the figure

below.

Figure 5.4 - Deformed and undeformed wing shape

Page 75: Multidisciplinary Design Optimization of a Morphing

59

5.2.4 Fluid-Structure Interaction

The previous section was describing a simple aero-structure problem. It assumes that the wing

deformation has no effect on the aerodynamic loading, which is an incorrect assumption. The

aerodynamic loading of the wing is dependent in the shape of the wing which is dependent on

the aerodynamic loading. The aerodynamic performance is now highly coupled with the

structural performance. Older aircraft were designed assuming a rigid wing for aerodynamic load

prediction. A rigid aircraft wing will produce more lift compared to a flexible wing; however this

higher lift would force designers to build a stiffer wing structure, which would increase the

weight of the aircraft. Figure 5.5 and Figure 5.6 will demonstrate this. For the initial state of the

aircraft, the wing was trimmed to the MTGW assuming a rigid wing. This will give a decent

starting angle of attack for convergence compared to a random starting angle of attack.

How the coupled Tornado-FE model works is very straight forward. Tornado requires four inputs

to generate a wing section: span, sweep, dihedral, and twist. The solution to the FE model

outputs the nodal displacements and rotations. The rotations of the nodes are used to transform

the Tornado model’s degree of freedoms (sweep, dihedral, and twist) into the new deformed

shape to determine the aerodynamic loading. It is assumed however, that due to the low

aerodynamic forces acting along the span, the length of the beam will remain constant.

Through several test cases, it was determined that the convergence can be unstable and it could

take several iterations. To improve convergence, numerical damping was applied to the beam

rotations before being used by Tornado. In the equations below, D is the damping set to 0.7 and

k is the iteration count.

( )

( )

( )

(5.6)

The next two figures show the lift to drag ratio and wingtip deflection convergence. It is clearly

shown that the static deflection for the rigid wing is far greater compared to the elastic wing due

to its higher loading.

Page 76: Multidisciplinary Design Optimization of a Morphing

60

Figure 5.5 - Convergence plots for an elastic wing at constant angle of attack

The convergence is slightly unstable at the beginning but it stabilizes quickly. There is a slight

decline in the lift to drag ratio from the initial rigid wing. For this example the rigid wing lift to

drag ratio was 25.2 and after convergence of the elastic wing the lift to drag ratio dropped to

22.8. There is a 9.5% drop in performance. The tip deflection dropped from 0.53 m to 0.29 m.

The above example was for a wing at constant angle of attack. Since the aircraft lift has been

altered due to its elastic deformation, the aircraft is no longer trimmed. Figure 5.6 below is an

example of the same wing being trimmed to the required weight.

Page 77: Multidisciplinary Design Optimization of a Morphing

61

Figure 5.6 - Convergence plots for an elastic wing while simultaneously being trimmed to

specified weight requirement

For this case, convergence required twice as many iterations. However the trimmed lift to drag

ratio is 23.8 which is only a 5.5% drop in performance. A more flexible wing will require more

iterations to find a static deflection.

5.3 High Fidelity Models

High fidelity models are typically used during the more detailed design phase. These models are

based on fewer assumptions and can model more complicated systems. However the

computational requirements can be extensive, such as CPU time and memory. This section will

describe two high fidelity approaches under consideration.

5.3.1 Aerodynamic Model

The aerodynamic model will use ANSYS CFX, which is Computational Fluid Dynamics (CFD)

software. CFX requires three steps: the first step is the Preprocessing, the second step is the

Solver, and the third step is the Postprocessor. The first step is where the problem boundary

Page 78: Multidisciplinary Design Optimization of a Morphing

62

conditions and the solver options are set for the solver. Since we are interested in transonic flight

conditions, air compressibility is an important factor in determining the aerodynamic loading.

The compressibility effect option was enabled under the solver options tab. Also temperature

damping was enabled since experience has shown that convergence of the energy equation can

be unstable especially if there are shockwaves present. Viscous effects are another important

factor to be considered for aerodynamic performance. However this creates two problems:

execution time and mesh quality. For an MDO solution, execution time of your discipline codes

is very important; to accurately predict the boundary layer of the wing, mesh resolution near the

surface needs to be very fine which would increase the number of nodes and in the boundary

layer there are large velocity gradients which could make a solution difficult to convergence. It

was decided to use an inviscid solution with free-slip boundary conditions on all surfaces for the

current work. The purpose of a wingtip device is to alter the pressure distribution of the main

wing. An inviscid model will be able to determine the pressure distribution and flow separation.

The meshing program used was ICEM CFD. The meshing approach for the current work uses an

unstructured mesh. This mesh type was selected because it is very easy to generate compared to

a structured mesh. A more complicated geometry (such as a randomly generated one from the

PSO algorithm) would make it difficult to create a structured mesh of good quality.

Figure 5.7 - ICEM CFD unstructured surface mesh

Page 79: Multidisciplinary Design Optimization of a Morphing

63

5.3.2 Structure Model

The finite element model that was used to calculate the structural response of the wing was

created in ANSYS. ANSYS has the capability for automatic model generation which is

necessary for the MDO process. This section details how the automatic generation of the model

is done and what types of elements are used to solve the problem.

The structural finite element model was created to contain the most basic elements of the

airframe structure. The main structural elements that will be used as variables are: skin thickness,

rib thickness, stringer areas, and spar thickness. In addition to these variables: the chordwise

location of the leading and trailing edge spars, the number of spars, the number of stringers, and

a structural taper are also considered to be variables.

The model was created by a series of keypoints, lines, and areas. The first step was to create the

ribs. Using equations 2.1 and 2.2 from chapter 2 and the airfoil coordinates it was possible to

generate the ribs. Using the spacing of the ribs given by and their orientations (Γ, θ, and Λ), it

is possible to create the airfoil keypoints in 3D space. The spacing of the ribs is dependent on the

span of the main wing and the number of ribs. Before we can mesh the wing we must first create

the areas from the keypoints. To do this we need to create the lines that will represent the edges

of the areas. The number of areas used for the rib will depend on the number of stringers.

The ribs have been orientated perpendicular to the leading edge spar. This will help improve

torsional stiffness compared to ribs that are parallel to the airflow. Also this perpendicular

orientation will make the manufacturing process easier. The same procedure for generating the

ribs was used to create the wing spars.

Page 80: Multidisciplinary Design Optimization of a Morphing

64

Figure 5.8 - Internal Structure layout

The stringers are relatively straight forward to create since they can be modeled only by lines.

The wing skin follows again the same procedure that was done to generate the rib areas. The

completed geometry is given below.

Figure 5.9 - Wing geometry without top skin panels

Rib

Stringer

Trailing Edge Spar Leading Edge Spar

Page 81: Multidisciplinary Design Optimization of a Morphing

65

5.3.3 Finite Element Mesh

The structures that make up the shape of the wing are curved areas and straight lines. In order to

accurately model the 3D wing properties appropriate 3D finite elements need to be chosen. Shell

elements were chosen for the curved areas because they can take into account the warping of the

element. They can also take into effect bending, shear, and torsional effects of the wing as well

as in-plane or normal loading. These shells were used to model the ribs, skin, and spars. Two

elements from ANSYS were considered: shell 281 and shell 181. Shell elements decouple the

deformation on the surface and the deformation in the normal direction, allowing for a simple

and efficient simulation of a thin structure. Both elements have 6 degrees of freedom for each

node, they can model thin to moderately thick structures, and they are suitable for linear and non-

linear applications. Shell 281 has 8 nodes which would allow for better modeling of curved

surfaces; however it would increase the total degree of freedom of the system which would

increase computational costs. For this reason shell 181 was selected. The stress for this element

can be calculated at the mid-plane, top surface, or bottom surface. All stresses to be exported are

from the mid-plane. The stringers are straight lines which can be modeled by beam elements.

Since shell 181 does not have any mid-side nodes, a beam element with only two nodes is

required, therefore beam 188 was chosen. The element is based on Timoshenko beam theory

which includes shear-deformation effects. Stringers assist with supporting the bending loads and

they provide stiffness to the skin panels. The stringers are modeled as a solid circular shaft. A

figure of the meshed model is shown below.

Page 82: Multidisciplinary Design Optimization of a Morphing

66

Figure 5.10 - FE mesh of wing

Each of the structural elements in the model has been structurally tapered to help decrease the

overall wing mass and to help give an even stress distribution from the wing root to wingtip. The

structural taper was implemented in a piece-wise linear fashion with each jump in structural

thickness occurring at each rib.

5.3.4 VGTM Model

To create the VGTM model, the same procedure for the main wing was implemented. Four

additional ribs were created. Three were for the morphing platform for each module and the last

rib was for the final wingtip. For the morphing modules to achieve kinematic motion the spars

and stringers were replaced with 4 passive and 4 active links. This was shown in Figure 2.5 in

chapter 2. To allow for kinematic motion a series of revolute and spherical joints will need to be

modeled, however this would require a non-linear solution which would substantially increase

computational time. Such modeling can be quite complicated and is beyond the scope of the

current work. Instead only static conditions will be considered. Working under the assumption

that some sort of a flexible or sliding skin panels will be used to allow the modules to change

shape; for the second stage the skin thickness will be reduced by a certain percentage. This

percentage will be dependent on the span of each module. For example, if the first module were

to extend by 10%, the thickness of the skin panels for this module for this stage will be reduced

by 10%. This change in thickness will lower the modules stiffness and make it difficult for the

Page 83: Multidisciplinary Design Optimization of a Morphing

67

skin to maintain a suitable aerodynamic shape. The FE model for the VGTM structure without

the skin is shown below.

Figure 5.11 - FE mesh of VGTM structure

5.3.5 Aero-Structural Analysis

The aero-structural analysis is required to determine the structural stresses from the CFD

pressure forces. For this analysis the CFD and FE model are solved independently and are

integrated in MATLAB. After the CFX solver converges the forces associated with each node

are exported during the post-processing stage. These pressure forces will be mapped to the FE

model.

In order to map the CFD mesh to the FE mesh, the wing was split into its upper and lower

sections. This makes the process less prone to errors since it guarantees that the upper or lower

surface CFD pressure forces are mapped to the FE nodes on the upper or lower surface. The

mapping process used here is simple algorithm that picks a CFD grid point and tries to find the

nearest FE node:

Page 84: Multidisciplinary Design Optimization of a Morphing

68

‖ (5.7)

Where is the location of the CFD grid point and

is the location of the FE nodes. is

the distance between the CFD grid point and the FE node. The algorithm finds the nearest

structural node to apply the necessary force. Since the CFD grid is more refined compared to the

FE mesh, some structural nodes will receive forces from multiple CFD grids points. A figure

showing the mapped pressure forces to the FE model is shown below.

Figure 5.12 - Mapped CFD pressure forces to FE mesh

For the current work only a rigid aeroelastic analysis was conducted. It is possible to execute a

fluid-structure interaction analysis using a similar approach as described earlier however the

computational expense would be prohibitive.

Page 85: Multidisciplinary Design Optimization of a Morphing

69

Chapter 6

Low and High Fidelity Multidisciplinary Design

Optimization Results

6.1 Introduction

This chapter will present the results for the high and low fidelity test cases and these results will

be compared to a reference wing. The wing’s geometry has been listed in Table 5.1. For the low

fidelity model optimization routine two test cases have been executed; the cruise stage has been

segmented into 2 morphing stages and 4 morphing stages.

6.2 Low Fidelity Results

Before we can execute the MDO routine for the low fidelity case we need to find an optimal

structural mass of the reference wing. This step is necessary since this mass will be used as a

constraint in the MDO analysis. This constraint was described in chapter 3 section 2.2. The

aircraft loads were determined at an altitude of 10km at Mach 0.85. The wing was trimmed to a

MTGW of 18000kg. A load factor of 6 was applied to the pressure forces on the wing and a

factor of safety of 1.5 was applied to the structural stress constraint. The convergence of the

optimization routine is in Figure 6.1, with a final optimal mass of 385.7 kg.

Page 86: Multidisciplinary Design Optimization of a Morphing

70

Figure 6.1 - Mass convergence plot for reference wing

The convergence of the mass objective is really quick up to 100 iterations. After this point the

change in the objective is very small since the optimization algorithm is refining the design. The

optimal structural design for the reference wing is listed in Table 6.1. The leading and trailing

edge offset are active with the lower bound and the scaling in the X direction (for all the

variables) is active with the upper bound.

Table 6.1 - Optimal design for wing structure of reference wing

Leading Edge Offset 0.1 Module Scale X 0.99

Trailing Edge Offset 0.6 Module Scale Y 0.979

Wing Root Scale X 0.99 Wingtip Scale X 0.99

Wing Root Scale Y 0.962 Wingtip Scale Y 0.973

Wing Tip Scale X 0.99

Wing Tip Scale Y 0.951

The stress constriant distribution is shown below. If the constraint is close to zero, this indicates

that the stress is active with the yeild limit of the material. This active constraint occurs at the

Page 87: Multidisciplinary Design Optimization of a Morphing

71

wing root (high reaction load) and the wing-module junction (low bending stiffness). The

bending stiffness distribution is shown in Figure 6.3.

Figure 6.2 - Stress distribution for reference wing

Figure 6.3 - Bending stiffness distribution for reference wing

Page 88: Multidisciplinary Design Optimization of a Morphing

72

Because of the wing taper and scaling of the wingbox, the bending stiffness decreases in a non-

linear fashion. This gives a design with decent stiffness at the wing root and a very low stiffness

for the modules and wingtip. Since we are minimizing the mass, it is not surprising that the

highest stress concentration occurs at the wing root and at the main wing tip. The figures for the

cross-sectional areas of the wing root, modules, and wingtip are in appendix C.

6.2.1 Optimal Design for Two Morphing Stages

The MDO routine has been split into two optimizations. The first 400 iterations will use an IDF

approach (the wing will not be trimmed at each iteration) and for an additional 400 iterations an

MDF approach (the wing will be trimmed) will be used. For the IDF approach the structure and

aerodynamic models will only need to be called twice (for each morphing stage) for each particle

(16000 function calls for each discipline). For the MDF approach however, to trim the aircraft,

three calls to the aerodynamic model is required (48000 aerodynamic calls and 16000 structure

calls). The reason for this combined IDF-MDF approach is very simple: better efficiency. The

IDF approach is faster since it does not waste time trimming a wing design that is far from the

optimum. The MDF is more appropriate for the refinement stage during the optimization. The

convergence plot is shown below.

Figure 6.4 - Fuel convergence for 2 morphing stage case

Page 89: Multidisciplinary Design Optimization of a Morphing

73

During the IDF run, it required 73 iterations to find a feasible solution. Up to this point the

algorithm needed to find a design that would satisfy the trim constraint. There is a slight increase

in the objective at the start of the MDF routine. For the IDF the trim constraint was a loose

constraint and for the MDF the constraint was a lot tighter. This would change the load

distribution and affect the wing’s performance. The final wingtip geometry for the first and

second stage is in Figure 6.5 and Figure 6.6 respectively.

Figure 6.5 - Morphed wingtip for the first stage

For the first stage the wingtip follows a traditional winglet design. The modules are swept back

slightly with a large dihedral.

Page 90: Multidisciplinary Design Optimization of a Morphing

74

Figure 6.6 - Morphed wingtip for the second stage

The second stage design is similar to the first, with the only significant difference being the

amount of dihedral. Figure C.3 and Figure C.4 show the first and second morphing stage wing

shapes in 3 different orientations. The performance for the optimal and reference wing designs is

in Table 6.2.

Table 6.2 - Performance results for the 2 Morphing stage case

Stage 1 Stage 2

Reference Optimal Reference Optimal

ΔWf 1.58e4 1.31e4 1.45e4 1.19e4

Mass 387.5 422.7 - -

CL 0.34 0.482 0.31 0.444

CD 0.0135 0.0158 0.0124 0.0143

CL/CD 25.2 30.6 25 31.1

αtrim 0.0023 0.4 -0.2516 0.102

Page 91: Multidisciplinary Design Optimization of a Morphing

75

The total amount of fuel consumed for the optimal design is 2.502e4 N. This accounts for 14.2%

of the MTGW. The total amount of fuel consumed for the reference wing is 3.02e4 N which is

17.1% of the MTGW. The ratio of the optimal fuel consumed to the reference wing is 0.83,

which means a 17% fuel savings. The lift to drag ratio shows us that for the reference wing, the

ratio is declining slightly however for the optimal design the ratio is increasing. The reference

wing did not have a wingtip device to improve efficiency. If the reference wing had a wingtip

device and if we assume the lift to drag ratio of the optimal wing’s first stage is constant for the

6000 km range, the fuel consumption is 2.54e4 N. Comparing this to the optimal design for the

two morphing transitions, there is only a 1.2% reduction in fuel. The stress distribution is plotted

next in Figure 6.7.

Figure 6.7 – Stress constraint distribution for 2 morphing stages

The stress along the wing follows the same general trend as the reference wing, with the stress

peaking at the wing root and wingtip. Since the mass of the wing is not being minimized, the

constraints are not active at the root and the tip. The final mass was almost active with the

reference mass constraint. The 10% allowance on the mass constraint was more than sufficient to

confine the wing mass. A tighter constraint could have been implemented.

Page 92: Multidisciplinary Design Optimization of a Morphing

76

6.2.3 Optimal Design for Four Morphing Stages

To improve performance of the aircraft, additional morphing stages are required. The number of

morphing stages has increased to 4, which increases the problem dimension to 76. Just like

before the MDO routine has been split into two, with an IDF approach to start and an MDF to

finish. Each approach was executed for 400 iterations. Since the number of morphing stages is 4,

this doubles the number function calls compared to the previous test case (32000 for IDF, 96000

aerodynamic and 32000 structure for MDF). The convergence plot is presented in Figure 6.8.

Figure 6.8 - Fuel convergence for 4 morphing stages

For the IDF portion of the optimization, convergence is very gradual, however when the MDF

approach is used, the convergence is much more rapid up to 550 iterations. After this point the

convergence slows and only minor improvements in the objective are obtained.

Table 6.3 - Performance results for 4 morphing stages

Stage 1 Stage 2 Stage 3 Stage 4

Reference Optimal Reference Optimal Reference Optimal Reference Optimal

ΔWf 8071 6746 7725 6464 7410 6128 7120 5837

Mass 385.7 424.26 - - - - - -

Page 93: Multidisciplinary Design Optimization of a Morphing

77

CL 0.339 0.541 0.324 0.512 0.309 0.501 0.295 0.483

CD 0.0135 0.0179 0.0129 0.0169 0.0124 0.0162 0.0119 0.0155

CL/CD 25.2 30.3 25.1 30.4 25 30.8 24.8 31.2

αtrim 0.0023 0.76 -0.1276 0.653 -0.252 0.325 -0.371 0.266

The total fuel consumed for the entire cruise stage is 2.52e4 N, which amounts to approximately

14% of the MTGW. The total fuel consumed for the reference wing is 3.03e4 N, which amounts

to 17% of the MTGW. The ratio of the optimal to the reference fuel consumed is 0.832, which is

a fuel savings of 16.8%. This is slightly less than the fuel savings for the 2 morphing stage case.

Due to the increased number of design variables and the four objectives, it is possible that the

allowed number of iterations was insufficient for this case. More iteration would have provided a

better result. Again for this case, the design was compared to reference wing without a winglet.

If we assume that the lift to drag ratio for the first optimal stage is constant for the 6000 km

range, the fuel savings for a 4 morphing stage design is only 0.8%.

Comparing the figures in Figure 6.9, shows why there is a substantial improvement in fuel

consumption of the optimal wing compared to the reference wing.

Page 94: Multidisciplinary Design Optimization of a Morphing

78

Figure 6.9 - Aerodynamic performance of the 4 morphing stages

There is a substantial increase in the lift to drag ratios at the start of cruise for the optimal wing.

The optimal wing has a lift to drag ratio that increases during cruise, but the reference wing’s lift

to drag ratio declines slightly. Another interesting observation is that both the lift and drag

coefficients are declining during cruise in an almost linear trend. This decline in the coefficients

is due to the lower lift required to maintain a trimmed state. The slope of the lift coefficient is

approximately the same; however for the drag coefficient the rate of change in drag is slightly

higher for the optimal design. The shapes of the optimal wingtips are presented in Figure 6.10 to

Figure 6.13.

Page 95: Multidisciplinary Design Optimization of a Morphing

79

Figure 6.10 - First morphing stage wingtip shape for the 4 morphing stage case

For the first morphing stage above the wingtip, the modules have a greater sweep than the main

wing (similar to a raked wingtip) and have an increased dihedral (similar to a typical winglet).

Figure 6.11 - Second morphing stage wingtip shape for the 4 morphing stage case

Page 96: Multidisciplinary Design Optimization of a Morphing

80

The second morphing stage is very similar to the first stage but with a much larger dihedral.

Figure 6.12 - Third morphing stage wingtip shape for the 4 morphing stage case

The third stage seems to have taken to a very different design possibility. The modules have been

swept forward with a much smaller dihedral compared to the second stage.

Figure 6.13 - Fourth morphing stage wingtip shape for the 4 morphing stage case

Page 97: Multidisciplinary Design Optimization of a Morphing

81

The fourth stage has a similar design to the third, however the first two modules have been swept

back slightly but the third module has been swept forward. Also the span extension of the

morphing modules is much larger than that of the third stage. The stress constraint distribution

shows an active constraint at the main wing root and the junction point between the modules and

main wing. Also the mass constraint is active with the reference wing mass constraint.

Figure 6.14 - Stress Constraint distribution for 4 morphing stage case

6.2.4 Low Fidelity Summary

The previous results for the low fidelity cases do suffer from the aerodynamic and structural

dynamic model inaccuracies. For example for the two cases the sweep angle for the main wing

was 20o. This result should not be surprising since, Tornado cannot account for the wave drag

which occurs at transonic speeds. A lower swept wing would produce more lift at a lower angle

of attack. In hindsight it probably would have been better to keep the wing sweep constant. This

could also explain the swept forward shape for the last two stages for the 4 morphing stage case.

For both cases the root chord length is almost 3 m with a taper ratio of 0.3. The small chord

length greatly reduces the wing area so a much larger angle of attack is to trim the aircraft is

necessary. This will increase the lift induced drag of the aircraft but since the lift and drag do not

scale proportionately to each other this increase in drag is not that influential on the overall

objective. The mass of the wing is sensitive to only two constraints: stress and reference wing

Page 98: Multidisciplinary Design Optimization of a Morphing

82

mass. Because of the stochastic nature of the PSO algorithm, any design variable combination

that does not violate these constraints is considered to be acceptable. This is observed with the

stress constraint on the modules and wingtip for both cases. These stresses are very far from the

yield limit. Comparing this to the optimal reference wing stress, the stress of the modules and

wingtip are much higher. This implies that the mass of the optimal design of the 2 and 4

morphing stage case could be reduced even further.

6.3 High Fidelity Results

Just as before with the low fidelity model, an optimal structural of a conventional wing needs to

be determined. The aircraft loads were determined at an altitude of 10km at Mach 0.85. The

wing was trimmed to a MTGW of 18000kg. A load factor of 6 was applied to the pressure forces

on the wing and a factor of safety of 1.5 was applied to the structural stress constraint. Two

structural optimization test cases were executed to determine the effect of spar location. The

design variables were: skin, rib, and spar thickness; structural taper; the number of ribs and

stringers; and the radius of the stringer cross-section. The final results are presented in the table

below:

Table 6.4 - Structural Optimization Results Comparison

Design Variable Spar Location as a Variable Spar Location as a Constant

Mass (Element Volume) 0.57 0.67

Skin Thickness (m) 0.0161 0.019

Rib Thickness (m) 0.005 0.005

Spar Thickness (m) 0.005 0.005

Stringer Radius (m) 0.0001 0.005

Structural Taper 0.4 0.4

Number of Ribs 20 20

Number of Stringers 4 5

Page 99: Multidisciplinary Design Optimization of a Morphing

83

Spar Location %chord 0.8368, 0.25 0.8, 0.15

Comparing the results above it is obvious that having the spar location as a variable can

significantly reduce the mass of the structure. The skin thickness is higher for the fixed spar

location test because the skin will need to contribute more to resist the bending stress. The spar

located at the 0.25c location has a greater height due to the greater thickness of the spar which

increases its bending stiffness. The rib and spar thickness were the same for both cases, and the

stringer radius should have been the same but the lower limit on the variable was different. The

number of stringers for the second case was higher to address the bending stiffness. The number

of ribs was the same for both cases and inspection of the stress contours for the two cases

showed that the maximum stress occurs at the third rib near the trailing edge. It is important to

note that the spar, rib, stringer, and structural taper were active with the lower limit set on the

variables.

Figure 6.15 - Von-Mises stress contours for optimal reference wing structure

For the high fidelity case presented here the MDO routine was terminated early due to the

excessive computational time. After approximately three months of execution only 100 iterations

were completed. A combined IDF and MDF was used just as before; 50 iterations for the IDF

and 50 for the MDF. Even with this shortened optimization we can still learn from the results.

Figure 6.16 below shows the optimal wing shape and Figure 6.17 shows the optimal wingtip

shape for the first morphing stage. The shape of the wingtip has two obvious similarities: there is

Page 100: Multidisciplinary Design Optimization of a Morphing

84

an increase in the wingtip cant (just like a typical winglet design) and an increased wing sweep

angle (just like a raked wingtip).

Figure 6.16 – High fidelity first morphing stage optimal wing shape

Figure 6.17 – High fidelity first morphing stage optimal wingtip shape

Page 101: Multidisciplinary Design Optimization of a Morphing

85

Figure 6.18 shows the optimal wingtip shape for the second morphing stage. The design is very

similar to the first stage. The only notable main difference is the smaller cant angles.

Figure 6.18 - High fidelity second morphing stage optimal wingtip shape

Inspection of both morphing stage designs’ pressure contours near the wingtip has shown low

pressure on the lower surface of the wingtip shown in Figure 6.19. This implies that the wingtip

region has a negative angle of attack. This would reduce the aerodynamic loading at the wingtip,

which would decrease the stress concentration at the wing-module junction.

Page 102: Multidisciplinary Design Optimization of a Morphing

86

Figure 6.19 - Pressure contours of wingtip region for first (left) and second stage (right)

The Von-Mises stress contours are shown Figure 6.20 and Figure 6.21 for the first and second

stage respectively. The maximum stress occurs at the second rib, for the first stage, and there are

some stress concentrations at the junction between the main wing and the first module. The

second stage maximum stress occurs at the junction between the first and second modules. This

is due to the lower cant angle at the wingtip. The forces on the wing are normal to the wing’s

surface. A larger cant angle will give much larger forces in the x and z directions compared to a

lower cant angle where the x and z forces will be small. This smaller cant will produce higher

forces in the y direction which will result in higher stresses for the modules.

Page 103: Multidisciplinary Design Optimization of a Morphing

87

Figure 6.20 - Von-Mises stress contours for first morphing stage

Figure 6.21 - Von-Mises stress contours for second morphing stage

The results for the first and second stage results are in Table 6.5. The total fuel consumed for the

optimal design is 6.5e4 N compared to 6.7e4 N for the reference wing. This is only a 3%

reduction in fuel.

Table 6.5 - First and second morphing stage performance results

Stage 1 Stage 2

Page 104: Multidisciplinary Design Optimization of a Morphing

88

Reference Optimal Reference Optimal

ΔWf 3.6e4 3.5e4 3.1e4 2.98e4

Mass 0.57 0.52 - -

CL/CD 10.86 12.3 10.08 10.1

The final design for the high fidelity model can be found in appendix C. Comparing the results to

the low fidelity two morphing stage case: the sweep angle for the high fidelity is much larger

(35o vs. 20

o) due to the CFD’s sensitivity to transonic conditions, the root chord is larger (3.9 m

vs. 3 m) but with a smaller taper ratio (0.19 vs. 0.3), and the span is much larger as well (7.85 m

vs. 6.85 m). From the low fidelity case it was demonstrated that the lift to drag ratio should

increase for during cruise for better performance, however this is clearly not the case for the high

fidelity. There is definitely room for improvement with more iteration. The pressure contours are

shown in Figure C.9 and Figure C.10 in the appendix for both the first and second stage. There is

a discontinuity in the contour near the trailing edge suggesting that there is a shockwave present.

Page 105: Multidisciplinary Design Optimization of a Morphing

89

Chapter 7

Conclusion

Morphing an aircraft wingtip can provide a substantial performance improvement. Most civil

transport aircraft and business jets are optimized for one specific mission leg but for other flight

conditions they are suboptimal. By altering the shape of the wingtip, we can force the load

distribution to adapt to the required flight condition to improve performance. The concept

investigated in this thesis relies on the use of variable geometry truss mechanisms (VGTMs) to

permit wingtip morphing. The wing has been divided into two inactive components (the main

wing and wingtip) and three morphing modules, composed of VGTMs, which permit individual

or simultaneous cant, span, sweep, and twist morphing modes. The scope of the thesis work was

to prove the concepts viability to reduce fuel consumption with multiple morphing stages at

cruise. This was carried out through establishing and demonstrating an MDO design framework.

7.1 Thesis Conclusion

An aircraft with a morphing wingtip will never maximize its performance if the morphing

wingtip is not utilized to its full capability. Morphing the wingtip to a predefined shape only at

the start of a mission leg (cruise, loiter, climb, etc…), will not fully utilize the capability of a

morphing wingtip. Instead we might as well use a conventional winglet design because by using

a predefined shape we are now using a design that might be optimal for the beginning of the

mission leg but will become eventually suboptimal near the end. An aircraft will only see the

benefit of a morphing wingtip if the wingtip can change shape at different intervals during the

mission leg. For the current work the fuel consumption for cruise was minimized with two and

four morphing stages for a low fidelity model, and a two stage high fidelity model. A combined

effort using an IDF to start the optimization and an MDF to finish the optimization was

implemented. The purpose of this combined approach was to make the optimization process as

efficient as possible. An IDF approach is beneficial for starting the optimization since it requires

fewer discipline function calls and an MDF is more appropriate for the refinement stage of the

optimization routine. The fuel savings for the two and four morphing stage low fidelity cases

were 17% and 16.8% respectively. The four stage case should have produced better results since

there were more optimization points for cruise, but the defined number of iterations was

Page 106: Multidisciplinary Design Optimization of a Morphing

90

insufficient. The lower fidelity case was unable to predict the more complicated flow

phenomenon (wave drag, flow separation, viscous, etc…) and designed a wing with a small area

with a very low sweep angle. The higher fidelity model, using CFD, is a much better model for

the aerodynamic discipline but requires a lot more time to execute. The higher fidelity

optimization routine was terminated after 100 iterations and the design was able to achieve a 3%

reduction in fuel. This could definitely be improved if more iteration had been allowed. Stress

concentrations occurred near the wing root and wing-module junction. The MDO framework

was able to determine a design that could out perform a conventional wing; however there are a

few areas in which to improve which will be discussed next.

7.2 Future Work

The morphing concept and design framework established within this thesis has only considered a

few aspects of morphing wingtip design. Consequently, there is a lot of room to improve and

expand on the current work. Some of the possible ways in which this can be done are discussed

below.

The framework itself can be improved upon in several areas. The shortcomings of the linear

aerodynamic model used in the framework have been discussed. Consequently, one way to

expand on the current work would be to couple a more sophisticated aerodynamic tool to the

optimization routine. This was done using a CFD model however in an effort to reduce

computational time viscous effects were ignored and the convergence tolerance was not very

tight. Using a more sophisticated CFD model; viscous effects, shockwave-boundary layer

interaction, and the drag which accumulates due to the interacting junction flows between

adjacent modules can be determined. The low fidelity tool is, however, justified at the moment

due to its quick calculation of aerodynamic coefficients and wing loading. This is a requirement

of design studies that entail a large number of function evaluations such as the one here. A

combination of a low fidelity model and a high fidelity model could also be used. For example a

low fidelity model would be most beneficial during the early optimization stage and a more

advanced model for the refinement stage.

There are a few additional considerations that should be included into the current optimization

framework. For example, the number of morphing stages and the length of each could also be

considered as variables, the dynamic effects of the wingtip changing shape during cruise could

Page 107: Multidisciplinary Design Optimization of a Morphing

91

be used as a constraint, maintaining a stable aircraft is very important but it may require

additional variables for the tail section, and finally activation energy as an objective or constraint

for the VGTM could also be included.

In developing the discipline codes little regard was given toward skin. The skin material must

permit morphing, while maintaining some level of structural integrity and aerodynamic

consistency. For the low fidelity case the structural model completely ignored the skin, however

for the higher fidelity model, a simple attempt to include a morphing skin was employed where

the skin’s thickness would shrink with a module’s span extension. This would give the modules a

lower stiffness when extended. However this approach was only for a rigid aeroelastic analysis

of the wing. A more complete fluid-structure interaction would be a more appropriate. Since the

skin’s thickness is reduced, the ability for the skin panel to maintain aerodynamic consistently

would be a challenge. Also dynamic effect such as panel buffeting would also need to be

considered. Of course trying to meet these requirements in an optimization routine would take a

lot of time to execute.

As mentioned previously, an aircraft will not see the complete benefit of a morphing wingtip if it

does not change its shape during each mission leg. Multiple morphing stages would be required

to achieve its full potential. However there is no guarantee that the flight conditions will be the

same for an aircraft each time it is in operation. Presetting a morphing wingtip shape from wind

tunnel or CFD results will not necessarily provide optimum performance. It may be possible to

design an optimization algorithm to be included with the flight control system of an aircraft to

actively reconfigure the wing at different stages in flight. This of course will create some

difficulties for the optimizer since measurement data from the aircraft instruments could be

noisy.

These are just some of the areas in which to improve the current work. Morphing aircraft

research is a vast subject of interest in academia and in the industry. There are a lot of different

research areas to improve aircraft performance.

Page 108: Multidisciplinary Design Optimization of a Morphing

92

Bibliography

[1] M. Abdulrahim and R. Lind, "Flight Testing and Response Characteristics of a Variable

Gull Wing Morphing Aircraft," in Proceedings of AIAA Guidance, Navigation, and Control,

Conference and Exhibit, Providence, RI, 2004.

[2] A.-M. R. McGowan, D. D. Vicroy, R. C. Busan and A. S. Hahn, "Perspectives on Highly

Adaptive or Morphing Aircraft".NATO Research and Technology Organization.

[3] J. Wilson, Morphing UAVs Change the Shape of Warfare, vol. 42, Aerospace America,

2004, pp. 28-29.

[4] J. Bowman, B. Sanders and T. Weisshaar, "Evaluating the Impact of Morphing

Technologies an Aircraft Performance," in Proceedings of 43rd

AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference,

Denver, CO, 2002.

[5] T. Weisshaar, "Morphing Aircraft Technology - New Shapes for Aircraft Design," RTO-

MP-AVT-141, Neuilly-sur-Seine, France, 2006.

[6] L. Arrison, K. Birocco, C. Gaylord, B. Herndon, K. Manion and M. Metheny, "2002-2003

AE/ME Morphing Wing Design," Spring Semester Final Report, Virginia Tech Aerospace

Engineering Senior Design Project, 2003.

[7] J. Blondeau, J. Richeson and D. Pines, "Design, Development and Testing of a Morphing

Aspect Ratio Wing Using an Inflatable Telescopic Spar," in Proceedings of 44th

AIAA/ASME/ASCE/AHS/ASC Structures, Structutal Dynamics and Materials Conference,

Norfolk, VA, 2003.

[8] J. Blondeau and D. Pines, "Design and Testing of a Pneumatic Telescopic Wing for

Unmanned Aerial Vehicles," Journal of Aircraft, vol. 44, pp. 1088-1099, 2007.

Page 109: Multidisciplinary Design Optimization of a Morphing

93

[9] J. Bae, T. Seigler and D. Inman, "Aerodynamic and Static Aeroelastic Characteristics of a

Variable-Span Morphing Wing," Jornal of Aircraft, vol. 42, pp. 528-534, 2005.

[10] C. Han, K. Ruy and L. S., "Experimental Study of a Telescopic Wing Inside Channel,"

Engineering Note, Jornal of Aircraft, vol. 44, pp. 1029-1030, 2007.

[11] J. Henry, J. Blondeau and D. Pines, "Stability Analysis for UAVs with a Variable Aspect

Ratio Wing," in Proceedings of 46th AIAA/ASME/ASCE/AHS/ASC Structures, Structural

Dynamics, and Materials Conference, Austin, TX, 2005.

[12] J. Henry and D. Pines, "A Mathematical Model for Roll Dynamics by Use of Morphing-

Span Wing," in In Proceedings of 48th AIAA/ASME/ASCE/AHS/ASC Structures, Structural

Dynamics, and Materials Conference, Honolulu, HI, 2007.

[13] T. Seigler, J. Bae and D. Inman, "Flight Control of a Variable Span Cruise Missle," in In

Poceedings of 2004 ASME International Mechanical Engineering Congress and Exposition,

Anaheim, CA, 2004.

[14] R. D. Vocke III, C. S. Kothera, B. K. Woods and N. M. Wereley, "Development and

Testing of a Span-Extending Morphing Wing," Journal of Intelligent Material Systems and

Structures, 2011.

[15] M. Love, P. Zink, R. Stroud, D. Bye, S. Rizk and D. White, "Demonstration of Morphing

Technology through Ground and Wind Tunnel Tests," in In Proceedings of 48th

AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference,

Honolulu, HI, 2007.

[16] T. Ivanco, R. Scott, M. Love, S. Zink and T. Weisshaar, "Vakidation of the Lockheed Matin

Morphing Concept with Wind Tunnel Testing," in In Proceedings of 48th

AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference,

Honolulu, HI, 2007.

[17] D. Neal, M. Good, C. Johnston, H. Robertshaw, W. Mason and D. Inman, "Designing and

Page 110: Multidisciplinary Design Optimization of a Morphing

94

Wind-Tunnel Analysis of a Fully Adaptive Aircraft Configuration," in In Proceedings of

45th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials

Conference, Palm Springs, CA, 2004.

[18] Y. Heryanwan, H. Park, N. Goo, K. Yoon and Y. Byun, "Design and Deomonstration of a

Small Expandable Morphing Wing," in In Proceedings of SPIE 12th Annual International

Symposium on Smart Structures and Materials, San Diego, CA, 2005.

[19] G. Anderson and D. Cowan, "Aeroelastic Modeling Analysis and Testing of a Morphing

Wing Structure," in In Proceedings of 48th AIAA/ASME/ASCE/AHS/ASC Structures,

Structural Dynamics, and Materials Conference, Honolulu, HI, 2007.

[20] J. Flanagan, R. Strutzenberg, R. Meyers and J. Rodrian, "Development and Flight Testing of

a Morphing Aircraft, the NextGen MFX-1," in In Proceedings of 48th

AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference,

Honolulu, HI, 2007.

[21] J. Joo, B. Sanders, T. Johnson and M. Frecker, "Optimal Actuator Location within a

Morphing Wing Scissor Mechanism Configuration," in In Proceedings of SPIE Smart

Structures and Materials 2006: Modeling, Signal Processing, and Control, San Diego, CA,

2006.

[22] T. Johnson, M. Frecker, M. Abdalla, Z. Gurdal and D. Lindear, "Nonlinear Analysis and

Optimizatiom od Diamond Cell Morphing Wings," Journal of Intelligent Material Systems

and Structures, pp. 815-24, 2008.

[23] S. Bharti, M. Frecker, G. Lesieutre and J. Browne, "Tendon Actuated Cellular Mechanisms

for Morphing Aircraft Wing," in In Proceedings of SPIE Modeling, Signal Processing, and

Control for Smart Structures, San Diego, CA, 2007.

[24] B. Grady, "Multi-Objective Optimization of a Three Cell Morphing Wing Substructure,"

University of Dayton, Dayton, OH, 2010.

[25] P. Chen, D. Sarhaddi, R. Jha, D. Liu, K. Griffin and R. Yurkovich, "Variable Stiffness Spar

Page 111: Multidisciplinary Design Optimization of a Morphing

95

Approach for Aircraft Maneuver Enhancement Using ASTROS," Jounral of Aircraft, vol.

37, no. 5, 2000.

[26] J. Florance, J. Heeg, S. C.V. and P. Lively, "Variable Stiffness Spar Wind-Tunnel Modal

Development and Testing," in In Proceedings of 45th AIAA/ASME/ASCE/AHS/ASC

Structures, Structural Dynamics and Materials Conference, Palm Springs, CA, 2004.

[27] C. Nam, A. Chattopadhyay and Y. Kim, "Application of Shape Memory Alloy (SMA) Spars

for Aircraft Maneuver Enhancement," in In Proceedings of SPIE Smart Structures and

Materials 2002: Smart Structures and Integrated Systems, San Diego, CA, 2002.

[28] R. Barret, "Active Aeroelastic Tailoring of an Adaptive Flexspar Stabilator," Smart

Materials and Structures, vol. 5, pp. 723-730, 1996.

[29] R. Barret, R. McMurty, R. Vos, P. Tiso and R. De Breuker, "Post-buckled precompressed

piezoelectric flight control actuator design, development, and demonstration," Smart

Materials and Strucures, vol. 15, pp. 1323-1331, 2006.

[30] R. Vos, R. Barret and R. T. P. De Breuker, "Post-buckled precompressed elements: a new

clas of control actuators for morphing wing UAVs," Smart Materials and Structures, vol.

16, pp. 919-926, 2007.

[31] D. P. Raymer, Aircraft Design: A Conceptual Approach, 4th ed., Reston, Virgina: American

Institute of Aeronautics and Astronautics, 2006.

[32] J. D. Anderson, Fundamentals of Aerodynamics, 3rd, Ed., New York, New York: McGraw-

Hill, 2001.

[33] C. J. Bargsten and M. T. Gibson, "NASA Innovation in Aeronautics: Select Technologies

That Have Shapped Modern Aviation," NASA Technical Memorandum, 2011.

[34] S. Hoerner, "Fluid Dynamic Drag," Published by Author, 1957.

[35] R. T. Whitcomb, "A Design Approach and Selected Wind-Tunnel Results at High Subsonic

Page 112: Multidisciplinary Design Optimization of a Morphing

96

Speeds for Wing-Tip Mounted Winglets," NASA TN D-8260, Washington DC, 1976.

[36] J. F. Halpert, D. H. Prescott, T. R. Yechout and M. Amdt, "Aerodynamic Optimization and

Evaluation of KC-135R Winglets, Raked Wingtips, and a Wingspan Extension," in 48th

AIAA Aerospace Sciences Meeting Including the New Horizons Forum and Aerospace

Exposition, Orlando, FL, 2010.

[37] Boeing, "AERO Magazine," [Online]. Available:

http://www.boeing.com/commercial/aeromagazine/aero_17/wingtip_devices.html.

[Accessed 18 05 2013].

[38] P. Bourdin, A. Gatto and M. Friswell, "Potential of Articulated Split Wingtips for Morphing

Based Control of Flying Wing," in Proceedings of 25th AIAA Applied Aerodynamics

Conference, Miami, Florida, 2007.

[39] P. Bourdin, A. Gatto and M. Friswell, "Aircraft Control via Variable Cant-Angle Winglets,"

Journal of Aircraft, vol. 45, pp. 414-423, 2008.

[40] P. Bourdin, A. Gatto and M. Friswell, "Performing Coordinated Turns With Articulated

Wing Tips as a Multi-axis Control Effectors," Aeronautical Journal, vol. 114, pp. 35-47,

2010.

[41] A. Gatto, P. Bourdin and M. Friswell, "Experimental Investigation into Articulated Winglets

Effects on Flying Wing Surface Pressure Aerodynamics," Journal of Aircraft, vol. 47, pp.

1811-1815, 2010.

[42] N. Ursache, T. Melin, A. Isikveren and M. Friswell, "Morphing Winglets for Aircraft

Multiphase Improvement," in Proceedings of 7th AIAA Aviation Technology, Integration

and Operations Conference, Belfast, Northern Ireland, 2007.

[43] D. Smith, A. Isikveren, R. Ajaj and M. Friswell, "Multidisciplinary Design Optimization of

an Active Nonplannar Polymorphing Wing," in Proceedings of 27th International Congress

of the Aeronautical Sciences, Nice, France, 2010.

Page 113: Multidisciplinary Design Optimization of a Morphing

97

[44] A. D. Finistauri and F. Xi, "Type Synthesis and Kinematics of a Modular Variable

Geometry Truss Mechanism for Aircraft Morphing," in Proceedings of the ASME/IFToMM

INternational COnference on Reconfigurable Mechanisms and Robots, Farifield, NJ, 2009.

[45] A. D. Finistauri, F. Xi and P. Walsh, "Discretization Method for the Development of a

Modular Morphing Wing," Journal of Aircraft, vol. 49, no. 1, 2012.

[46] J. Morley, "Shape Optimization and Modular Discretization for the Development of a

Morphing Wingtip," University of Toronto, Toronto, 2012.

[47] A. Moosavin, F. Xi and S. Hasemi, "Design and Motion Control of a Fully Variable

Morphing Wings".

[48] J. Sobieszczanski-Sobieski, "Multidisciplinary Design Optimization: an Emerging New

Engineering Discipline," in Advances in Structural Optimization, Boston, Kluwer Academic

Publishers, 1995, p. 483.

[49] E. J. Cramer, J. E. Denis Jr., P. D. Frank, R. M. Lewis and G. R. Shubin, "Problem

Formulation for Multidisciplinary Optimization," SIAM Journal on Optimization, no. 4, pp.

754-776, November 1994.

[50] P. Breitkopf and R. F. Coelho, Eds., Multidisciplinary Design Optimization in

Computational Mechanics, London nad Hoboken: ISTE Ltd and JohnWile & Sons Inc,

2010.

[51] D. M. Skillen and W. A. Crossley, "Conceptual Morphing Aircraft Sizing Using a Multi-

Level Optimization Strategy," in 46th AIAA Aerospace Eciences Meeting and Exhibit, Reno,

Nevada, 2008.

[52] J. Kennedy and R. C. Eberhart, "Particle Swarm Optimization," in Proceedings of the 1995

IEEE International Conference on Neural Networks, Perth, Australia, 1995.

[53] R. C. Eberhart and Y. Shi, Comparison Between genetic algorithms and particle swarm

Page 114: Multidisciplinary Design Optimization of a Morphing

98

optimization, W. Potor, N. Saravanan, D. Waagen and A. Eiben, Eds., Berlin: Springer,

1998, pp. 611-616.

[54] Y. Shi and R. C. Eberhart, "A modified particle swarm optimizer," Anchorage, 1998a.

[55] Y. Shi and R. C. Eberhart, "Parameter selection in particle swarm optimization," in Lecture

Notes in Computer Science, London, Springer, 1998b.

[56] P. Fourie and A. Groenwold, "Particle Swarms in Size and Shaper Optimization," in Proc.

Workshop on Multidisciplinary Design Optimzation, Pretoria, South Africa, 2000.

[57] P. Fourie and A. Groenwold, "The Particle Swarm Optimization Algorithm in Size and

Shape Optimization," Structural and Multidisciplinary Optimization, vol. 23, no. 4, pp. 259-

267, May 2002.

[58] J. Schutte and A. Groenwold, "Optimial Sizing Design of Truss Structures Using the

Particle Swarm Optimzation," in Proceedings of the 9th AIAA/ISSMO Symposium on

Multidisciplinary Analysis and Optimization, 2002.

[59] A. Ratnaweera, S. Halgamuge and H. Watson, "Self Organizing Heirarchical Particle

Swarm Optimizer With Time Varying Acceleration Coefficients," IEEE Transactions on

Evolutionary Computation, vol. 8, no. 3, pp. 240-255, 2004.

[60] M. Senthil Arumugam, R. Rao and Arthi Chandramohan, "A New and Improved Version of

Particle Swarm Optimization Algorithm with Global-Local Best Parameters," vol. 16, pp.

331-357, 2008.

[61] A. Ratnaweera, S. Halgamuge and H. Watson, "Particle Swam Optimization with Self-

Adaptive Acceleration Coefficients," in Proceeding of the 1st Internation Conference on

Fuzzy Systems and Knowledge Discovery (FSKD 2002), 2003.

[62] E. Peer, F. Van Den Bergh and A. Engelbrecht, "Using Neighborhoods with the Guaranteed

Convergence PSO," in Proceeding of the IEEE Swarm Intelligence Symposium 2003 (SIS

2003), Indianapolis, Indiana, 2003.

Page 115: Multidisciplinary Design Optimization of a Morphing

99

[63] M. Clerc and J. Kennedy, "The particle swrm-explosion, stability, and convergence in

multidimensional complex space," IEEE Transactions on Evolutionary Computation, vol. 6,

no. 1, 2002.

[64] S. S. Rao, Engineering Optimization: Theory and Practice, 4th ed., Hoboken, New Jersey:

John Wiley & Sons,Inc., 2009.

[65] T. Melin, "Tornado, a vortex lattice MATLAB implementation for Linear Aerodynamic

Wing applications," Royal Institute of Technology, 2000.

[66] T. Melin, A. T. Isikveren and M. I. Friswell, "Induced-Drag Compressibility Correction for

Three Dimensional Vortex Lattice Methods," AIAA Journal of Aircraft Engineering Notes,

vol. 47, no. 4, 2010.

[67] M. D. Skillen and W. A. Crossley, "Modeling and Optimization for Morphing Wing

Concept Generation," NASA/CR-2007-214860, 2007.

[68] D. L. Logan, A First Course in the Finite Element Method, 4th ed., Toronto, Ontario:

Thomson, 2007.

Page 116: Multidisciplinary Design Optimization of a Morphing

100

Appendix A – PSO Results

A.1 Linear Test Case Results for Static and Dynamic

This section will present the results for the linear social and cognitive parameter distribution.

Table A.1 below lists the parameters that have been used. The best solutions for both the static

and dynamic approach are in bold.

Table A.1 - Linear test case for unconstraint benchmark problems

Test Case Parameters

SPSO TVIW, c1=c2=2

CPSO w=0.72984, c1=c2=1.49618

PSO-L1 TVIW, c1best = c2worst = 1 and c1worst = c2best = 3

PSO-L2 TVIW, c1best = c2worst = 3 and c1worst = c2best = 1

PSO-L3 TVIW, c1best = c2worst = 0 and c1worst = c2best = 2

PSO-L4 TVIW, c1best = c2worst = 2 and c1worst = c2best = 0

The PSO-L1 test case showed the best improvement for the Rastrigin function compared to the

SPSO and CPSO. The dynamic approach showed a decrease in performance compared to the

static (with PSO-L4 as the exception). PSO-L3 and PSO-L4 showed poor performance compared

to the SPSO and showed a small improvement compared to the CPSO. For the Rosenbrock

function, the PSO-L2 for both the static and dynamic approaches showed a substantial

improvement compared to the SPSO and CPSO. The dynamic approach was unable to improve

the static case. PSO-L3 and PSO-L4 showed very poor performance for both static and dynamic.

For the Sphere function, PSO-L1 and PSO-L2 performed well for the static approach, but

achieved poor performance for the dynamic. The dynamic approach most likely encouraged too

much exploration and allowed the algorithm to converge too slowly. PSO-L3 and PSO-L4 again

had very poor performance. For the Griewank function, the static approach PSO-L2 showed the

Page 117: Multidisciplinary Design Optimization of a Morphing

101

best performance for the static approach and PSO-L1 had the best performance for the dynamic

approach.

Table A.2 - Linear test case results for unconstraint benchmark problems

Static Dynamic

Function Test Case Mean

Fitness

STD

Fitness

Mean

Fitness

STD

Fitness

Rastrigin

SPSO 56.90118 24.33535 55.06076 24.94498

CPSO 308.4377 47.85343 301.9652 47.02567

PSO-L1 40.19204 16.80198 41.30647 15.76697

PSO-L2 42.05146 18.95228 49.71422 16.6171

PSO-L3 252.2899 38.12768 281.7518 40.84652

PSO-L4 241.5799 41.00807 228.4859 39.55752

Rosenbrock

SPSO 5934.871 41186.66 7399.068 27373.19

CPSO 14327244 5527449 14687113 6658570

PSO-L1 179.4968 513.4332 314.4676 590.1053

PSO-L2 92.00987 110.5895 109.1248 308.7556

PSO-L3 1545192 1045339 3451285 2634375

PSO-L4 1297824 987565.8 2560736 1430502

Sphere

SPSO 62.20384 214.8213 132.9173 306.1166

CPSO 3414.737 841.5883 3306.931 825.3003

PSO-L1 1.07E-14 7.78E-14 0.000502 0.001621

PSO-L2 1.31E-15 1.01E-14 0.470508 3.020789

PSO-L3 1046.318 384.7626 1437.601 478.9931

Page 118: Multidisciplinary Design Optimization of a Morphing

102

PSO-L4 959.9526 392.4152 1369.438 409.9016

Griewank

SPSO 6.212958 15.61322 5.437311 16.05851

CPSO 122.8694 28.17922 127.2405 30.91161

PSO-L1 0.05742 0.058148 0.033803 0.037524

PSO-L2 0.050278 0.048272 0.199606 0.267707

PSO-L3 36.37878 15.66795 54.6751 18.26063

PSO-L4 31.34928 12.91106 48.86456 15.62182

A.2 Step Test Case Results for Static and Dynamic

This section will present the results for the step distribution cases. The parameters are listed in

Table A.3. U represents the upper group, M is the middle group, and L is the low group. With

U=0.2 means that the top 20% of the swarm belongs to the upper group, with M=0.6 means that

the particles that lie between 20% and 60% are part of the middle group, and L=1 means the

particles that lie between 100% and 60% are part of the low group.

Table A.3 - Step test case for unconstraint benchmark problems

Test Case Parameters

SPSO TVIW, c1=c2=2

CPSO w=0.72984, c1=c2=1.49618

PSO-S1 TVIW, U=0.2 c1=3 c2=1 and M=0.6 c1=2 c2=2 and L=1 c1=1 c2=3

PSO-S2 TVIW, U=0.2 c1=1 c2=3 and M=0.6 c1=2 c2=2 and L=1 c1=3 c2=1

PSO-S3 TVIW, U=0.4 c1=3 c2=1 and M=0.8 c1=2 c2=2 and L=1 c1=1 c2=3

PSO-S4 TVIW, U=0.4 c1=1 c2=3 and M=0.8 c1=2 c2=2 and L=1 c1=3 c2=1

For the Rastrigin function, the PSO-S4 performed the best for the static case and the PSO-S1

performed the best for the dynamic case. With the exception of PSO-S4, the dynamic approach

Page 119: Multidisciplinary Design Optimization of a Morphing

103

for the other test cases yielded better results. The static test cases for the Rosenbrock function,

PSO-S1 to PSO-S3 performed poorly compared to their respective dynamic test cases. The

dynamic technique introduced more exploration to the optimization process which allowed for a

better solution. For the Sphere function, the static approach performs better for all test cases

except for PSO-S3. For the Griewank Function, the PSO-S1 performed the best for both the

static and dynamic techniques. For all the test cases PSO-S1 to PSO-S4, the dynamic technique

performed much better than the static. Comparing the results, PSO-S1 is the best option for

multi-modal functions. The smaller upper group with greater cognitive learning proved to be

more explorative than the other test cases.

Table A.4 - Step test case results for unconstraint benchmark problems

Static Dynamic

Function Test Case Mean

Fitness

STD

Fitness

Mean

Fitness

STD

Fitness

Rastrigin SPSO 56.90118 24.33535 55.06076 24.94498

CPSO 308.4377 47.85343 301.9652 47.02567

PSO-S1 41.7291 16.57188 38.12451 19.76411

PSO-S2 41.45962 16.3556 41.1666 13.6602

PSO-S3 43.06124 21.12839 42.63005 21.90669

PSO-S4 36.51926 12.15523 42.84084 29.20735

Rosenbrock SPSO 5934.871 41186.66 7399.068 27373.19

CPSO 14327244 5527449 14687113 6658570

PSO-S1 1970.379 12653.4 74.551 83.64479

PSO-S2 126.4138 311.8671 87.002 92.801

PSO-S3 199.2434 589.2289 91.0184 100.6304

PSO-S4 1035.156 8993.523 3721.739 17717.38

Page 120: Multidisciplinary Design Optimization of a Morphing

104

Sphere SPSO 62.20384 214.8213 132.9173 306.1166

CPSO 3414.737 841.5883 3306.931 825.3003

PSO-S1 5.15E-15 1.92E-14 5.92E-12 2.78E-11

PSO-S2 3.15E-11 3.12E-10 8.99E-08 8.99E-07

PSO-S3 8.95E-13 5.28E-12 9.53E-15 7.59E-14

PSO-S4 1.62E-14 7.71E-14 1.33E-11 5.84E-11

Griewank SPSO 6.212958 15.61322 5.437311 16.05851

CPSO 122.8694 28.17922 127.2405 30.91161

PSO-S1 0.052928 0.048512 0.029862 0.036609

PSO-S2 0.063826 0.059922 0.046133 0.045377

PSO-S3 0.07038 0.052941 0.047013 0.049769

PSO-S4 0.056182 0.040877 0.038422 0.035625

A.3 Sinusoidal Test Case Results for Static and Dynamic

This section will present the results for the sinusoidal distribution cases. The parameters are

listed in Table A.5.

Table A.5 - Sinusoidal test case for unconstraint benchmark problems

Test Case Parameters

SPSO TVIW, c1=c2=2

CPSO w=0.72984, c1=c2=1.49618

PSO-Sin1 TVIW, ω=2π and φ=π/2

PSO-Sin2 TVIW, ω=4π and φ=π/2

Page 121: Multidisciplinary Design Optimization of a Morphing

105

PSO-Sin3 TVIW, ω=6π and φ=π/2

PSO-Sin4 TVIW, ω=8π and φ=π/2

PSO-Sin5 TVIW, ω=2π and φ=0

PSO-Sin6 TVIW, ω=2π and φ=π

PSO-Sin3 provided the best solution for both the static and dynamic techniques for the Rastrigin

function; however their results are very similar. For each test case the static approach yielded

lower mean fitness values compare to the dynamic, but there is slight discernible difference in

performance for PSO-Sin1 to PSO-Sin4. PSO-Sin3 and PSO-Sin5for the Rosenbrock function;

the static and dynamic approach respectively have very similar results and are the best

performers. PSO-Sin1 for the dynamic approach had the best performance for the sphere

function. For all test cases of the Griewank function, except PS0-Sin5, the dynamic approach

was able to improve the mean fitness with PSO-Sin6 being the best performer.

Table A.6 – Sinusoidal test case results for unconstraint benchmark problems

Static Dynamic

Function Test Case Mean

Fitness

STD

Fitness

Mean

Fitness

STD

Fitness

Rastrigin SPSO 56.90118 24.33535 55.06076 24.94498

CPSO 308.4377 47.85343 301.9652 47.02567

PSO-Sin1 40.26834 15.29541 40.7902 17.28688

PSO-Sin2 37.88888 17.27999 38.4533 19.66409

PSO-Sin3 35.76533 14.66962 35.92581 13.06116

PSO-Sin4 37.68982 15.62986 38.84474 17.90841

PSO-Sin5 39.45006 17.18446 42.39678 14.60561

Page 122: Multidisciplinary Design Optimization of a Morphing

106

PSO-Sin6 36.92318 13.62176 41.57725 15.47511

Rosenbrock SPSO 5934.871 41186.66 7399.068 27373.19

CPSO 14327244 5527449 14687113 6658570

PSO-Sin1 122.27 141.3482 124.19 315.0992

PSO-Sin2 86.95107 108.5391 164.8982 514.0281

PSO-Sin3 80.87965 111.1392 127.6445 317.5367

PSO-Sin4 1017.51 8995.2 96.64395 123.9145

PSO-Sin5 109.3905 314.6066 80.86337 113.5386

PSO-Sin6 94.72582 109.1958 2000.426 12648.13

Sphere SPSO 62.20384 214.8213 132.9173 306.1166

CPSO 3414.737 841.5883 3306.931 825.3003

PSO-Sin1 2.26E-14 1.02E-13 3.67E-18 1.85E-17

PSO-Sin2 1.81E-14 7.28E-14 2.17E-15 7.94E-15

PSO-Sin3 2.54E-14 9.62E-14 7.83E-14 3.74E-13

PSO-Sin4 3.57E-14 1.52E-13 1.76E-13 9.17E-13

PSO-Sin5 8.21E-15 2.87E-14 1.12E-13 1.11E-12

PSO-Sin6 1.85E-14 8.04E-14 9.72E-06 2.94E-05

Griewank SPSO 6.212958 15.61322 5.437311 16.05851

CPSO 122.8694 28.17922 127.2405 30.91161

PSO-Sin1 0.055881 0.045171 0.051112 0.053198

PSO-Sin2 0.045247 0.041617 0.034147 0.044257

PSO-Sin3 0.059125 0.051338 0.032451 0.032623

Page 123: Multidisciplinary Design Optimization of a Morphing

107

PSO-Sin4 0.051446 0.045137 0.031386 0.029491

PSO-Sin5 0.047068 0.04089 0.06631 0.059792

PSO-Sin6 0.044868 0.045726 0.029877 0.035616

In summary, Table A.7 presents the results for the best social and cognitive distributions. For

each function the dynamic approach provided the best solutions. The sinusoidal and step

distributions achieved the best performance for all the benchmark functions.

Table A.7 – Best overall social and cognitive distributions

Function Static Dynamic

Rastrigin PSO-Sin3 PSO-Sin3

Rosenbrock PSO-Sin3 PSO-S1

Sphere PSO-L2 PSO-Sin1

Griewank PSO-Sin6 PSO-S1/PSO-Sin6

Page 124: Multidisciplinary Design Optimization of a Morphing

108

Appendix B – Beam Finite Element Model Validation

Figure B.1 - Problem 5.2

Table B.1 - Material and geometric properties for problem 5.2

A 100 in2

Iz 1000 in4

E 30,000 ksi

Table B.2 - Nodal displacement and nodal reaction comparison

[68] FE Model

dx 0.0033 in 0.00329 in

dy -0.0097 in -0.00974 in

rotz -0.0033 rad -0.00329 rad

F1x 26.6 kip 26.8 kip

F1y -2.268kip -2.261 kip

Page 125: Multidisciplinary Design Optimization of a Morphing

109

M1y -389 k-in -381.5 k-in

F3x 20.63 kip 20.6 kip

F3y 17.42 kip 17.4 kip

M3y 767.4 k-in 769.4 k-in

Figure B.2 - Problem 5.8

Table B.3 - Material and geometric properties for problem 5.8

A 10 in2

Iz and Iy 100 in4

J 50 in4

E 30,000 ksi

G 10,000 ksi

Page 126: Multidisciplinary Design Optimization of a Morphing

110

Table B.4 - Nodal displacement comparisons

[68] FE Model

dx 7.098e-5 in 7.09e-5 in

dy -0.014 in -0.014 in

dz -2.35e-5 in -2.35e-5 in

rotx -4e-3 rad -4e-3 rad

roty 1.78e-5 rad 1.78e-5 rad

rotz -1.03e-4 rad -1.03e-4 rad

Page 127: Multidisciplinary Design Optimization of a Morphing

111

Appendix C – MDO Result Extras

C.1 Optimal Reference Wing Mass

Figure C.1- Main wing and wing-module junction cross-sectional areas

Figure C.2 - Modules and wingtip cross-sectional area

Page 128: Multidisciplinary Design Optimization of a Morphing

112

C.2 Two Morphing Stages

Table C.1 - Optimal module design for 2 morphing stage case study

Design Variable Module 1 Module 2 Module 3

Stage 1

Sweep 9.6 -19.8 18.9

Dihedral 10.5 13.3 19.3

Twist -1.8 -1.28 3.1

Span 0.37 0.25 0.4

X-scale 0.865 0.865 0.865

Y-scale 0.82 0.82 0.82

Taper 0.7 0.917 0.7

Stage 2

Sweep -0.2 7.4 0.1

Dihedral 0 0 8.5

Twist -0.72 -1 2.45

Span 0.25 0.5 0.4

X-scale 0.865 0.865 0.865

Y-scale 0.82 0.82 0.82

Taper 0.7 0.917 0.7

Table C.2 – Optimal main wing and wingtip design for 2 morphing stage case study

Design Variable Main Wing Wingtip

Sweep 20 -3.6

Page 129: Multidisciplinary Design Optimization of a Morphing

113

Dihedral 0 1.7

Twist 0 -5

Span 6.85 0.723

X-scale 0.99 0.99

Y-scale 0.9 0.7

Taper 0.3 0.82

Chord 3.08 0.3443

Leading Edge 0.13 0.13

Trailing Edge 0.6 0.6

Figure C.3 - First morphing stage for 2 morphing stage case

Page 130: Multidisciplinary Design Optimization of a Morphing

114

Figure C.4 - Second morphing stage for 2 morphing stage case

C.3 Four Morphing Stages

Table C.3 - Optimal module design for 4 morphing stage case study

Design Variable Module 1 Module 2 Module 3

Stage 1

Sweep 1.85 20 -13.4

Dihedral 0 0 12.2

Twist -1.1 1.66 -2.8

Span 0.3 0.5 0.31

X-scale 0.99 0.99 0.99

Y-scale 0.882 0.882 0.882

Taper 0.89 1 0.78

Stage 2

Sweep -0.24 -0.81 20

Dihedral 0 20 18.5

Page 131: Multidisciplinary Design Optimization of a Morphing

115

Twist -3.3 5 -5

Span 0.5 0.4 0.5

X-scale 0.99 0.99 0.99

Y-scale 0.882 0.882 0.882

Taper 0.89 1 0.78

Stage 3

Sweep -20 20 -20

Dihedral 0 7 2.7

Twist 2.5 -1.7 -1.34

Span 0.36 0.2 0.31

X-scale 0.99 0.99 0.99

Y-scale 0.882 0.882 0.882

Taper 0.89 1 0.78

Stage 4

Sweep 18.5 -20 -20

Dihedral 8.5 0.96 0

Twist 1.14 -4.1 3.3

Span 0.34 0.39 0.25

X-scale 0.99 0.99 0.99

Y-scale 0.882 0.882 0.882

Taper 0.89 1 0.78

Table C.4 - Optimal main wing and wingtip design for 4 morphing stage case study

Main Wing Wingtip

Page 132: Multidisciplinary Design Optimization of a Morphing

116

Sweep 20 -11

Dihedral 0 0

Twist 0 2.5

Span 6 1

X-scale 0.99 0.75

Y-scale 0.932 0.813

Taper 0.3 0.61

Chord 3 0.38

Leading Edge 0.1 0.1

Trailing Edge 0.74 0.74

Figure C.5 - First morphing stage for 4 morphing stage case

Page 133: Multidisciplinary Design Optimization of a Morphing

117

Figure C.6 – Second morphing stage for 4 morphing stage case

Figure C.7 - Third morphing stage for 4 morphing stage case

Page 134: Multidisciplinary Design Optimization of a Morphing

118

Figure C.8 - Fourth morphing stage for 4 morphing stage case

C.4 High Fidelity Results

Table C.5 - High fidelity first and second stage module results

Design Variable Module 1 Module 2 Module 3

Stage 1

Sweep 13.04 -0.6 7.28

Dihedral 7.4 5.4 6.55

Twist 1.07 2.06 2.35

Span 0.43 0.325 0.464

Rib Thickness 0.006 0.006 0.006

Skin Thickness 0.0061 0.0061 0.0061

Taper 0.814 0.863 1

Passive Radius 0.0031 0.0031 0.0031

Active Radius 0.0038 0.0038 0.0038

Stage 2 Sweep -0.05 -4.7 6.85

Page 135: Multidisciplinary Design Optimization of a Morphing

119

Dihedral 1.9 4.95 3.24

Twist 2.67 0.67 -0.1

Span 0.29 0.35 0.49

Rib Thickness 0.006 0.006 0.006

Skin Thickness 0.0061 0.0061 0.0061

Taper 0.814 0.863 1

Passive Radius 0.0031 0.0031 0.0031

Active Radius 0.0038 0.0038 0.0038

Table C.6 - High fidelity main wing and wingtip results

Design Variable Main Wing Wingtip

Sweep 34.8 -3.6

Dihedral 0 1.7

Twist 6.2 -5

Span 7.85 0.757

Taper 0.19 0.82

Chord 3.9 0.3443

Ribs 19 0

Stringers 6 0

Leading Edge 0.25 0.25

Trailing Edge 0.72 0.721

Skin Thickness 0.0166 0.006

Page 136: Multidisciplinary Design Optimization of a Morphing

120

Rib Thickness 0.0052 0.0061

Spar Thickness 0.0086 0.0022

Stringer Radius 0.0049 0

Structural Taper 0.38 1

Figure C.9 - Pressure distribution of first morphing stage wing showing shockwave at

trailing edge

Page 137: Multidisciplinary Design Optimization of a Morphing

121

Figure C.10 - Pressure distribution of second morphing stage wing showing shockwave at

trailing edge

Figure C.11 - Comparing deformed to undeformed wing for the first morphing stage

Figure C.12 - Comparing deformed and undeformed for second morphing stage