Você está na página 1de 234

Thesis presented to the Faculty of the Division of Graduate Studies of the

Technological Institute of Aeronautics in partial fulfillment of the

requirements for the Degree of Doctor in Science in the Course of

Aeronautical and Mechanical Engineering, Area of Aerodynamics,

Propulsion and Energy.

Roberto Gil Annes da Silva

A STUDY ON CORRECTION METHODS FOR

AEROELASTIC ANALYSIS IN TRANSONIC FLOW

...................................................................
Prof. Dr. João Luiz Filgueiras de Azevedo
Advisor

...........................................................
Dr. Olympio Achilles de Faria Mello
Advisor

......................................................
Prof. Dr. Homero Santiago Maciel
Head of the Division of Graduate Studies

Campo Montenegro
São José dos Campos, SP – Brasil
2004
Dados Internacionais de Catalogação-na-Publicação (CIP)
Divisão Biblioteca Central do ITA/CTA
Silva, Roberto Gil Annes da Silva
A Study on Correction Methods for Aeroelastic Analysis in Transonic Flow /
Roberto Gil Annes da Silva.
São José dos Campos, 2004.
233f.

Tese de doutorado– Curso de Engenharia Aeronáutica e Mecânica, e área de Aerodinâmica, Propulsão


e Energia – Instituto Tecnológico de Aeronáutica, 2004. Orientadores: Prof. Dr. João Luiz Filgueiras
de Azevedo e Prof. Dr. Olympio Achilles de Faria Mello.

1. Aeroelasticidade . 2. Aerodinâmica não estacionária. 3. Escoamento transônico. 4. Correção. 5.


Dinâmica dos fluidos computacional. 6. Sustentação aerodinâmica. 7 Mecânica dos fluidos. 8 Física. I.
Centro Técnico Aeroespacial. Instituto Tecnológico de Aeronáutica. Divisão de Engenharia Aeronáutica.
II.Título

REFERÊNCIA BIBLIOGRÁFICA

SILVA, Roberto Gil Annes da. A Study on Correction Methods for Aeroelastic Analysis
in Transonic Flow. 2004. 233f. Tese de doutorado – Instituto Tecnológico de Aeronáutica,
São José dos Campos.

CESSÃO DE DIREITOS

NOME DO AUTOR : Roberto Gil Annes da Silva


TÍTULO DO TRABALHO : A Study on Correction Methods for Aeroelastic Analysis in
Transonic Flow
TIPO DO TRABALHO/ANO : Tese / 2004

É concedida ao Instituto Tecnológico de Aeronáutica permissão para reproduzir cópias desta


tese e para emprestar ou vender cópias somente para propósitos acadêmicos e científicos. O
autor reserva outros direitos de publicação e nenhuma parte desta tese pode ser reproduzida
sem a autorização do autor.

______________________
Roberto Gil Annes da Silva
Rua Salto Grande, no 264, apto. 102, Perdizes, São Paulo,
SP – CEP: 01257-020, Brasil.
A STUDY ON CORRECTION METHODS FOR

AEROELASTIC ANALYSIS IN TRANSONIC FLOW

Roberto Gil Annes da Silva

Thesis Committee Composition:

Prof. Dr. Maher Nasr Bismark-Nasr .........................................................Chairperson - ITA

Prof. Dr. João Luiz Filgueiras de Azevedo .....................................................Advisor - ITA

Prof. Dr. Olympio Achilles de Faria Mello .....................................................Advisor - IAE

Prof. Dr. Donizeti de Andrade .......................................................................Member - ITA

Prof. Dr. Armando Miguel Awruch ..........................................................Member - UFRGS

Prof. Dr. Flávio Donizeti Marques ................................................................Member - USP

ITA
DEDICATION

To my wife Michelle and my children Julia and Pedro.


ACKNOWLEDGEMENTS

I would like to express my sincere gratitude to my advisors, Prof. Dr. João Luiz

Filgueiras de Azevedo and Prof. Dr. Olympio Achilles de Faria Mello, for their guidance and

advice during the course of this work. Their encouragement, thoughtfulness, and supervision

are deeply acknowledge

I would like to thank Prof. Dr. Danny D. Liu of Arizona State University and Mr. P.C.

Chen of ZONA Technology for their valuable guidance, constructive ideas, and generosity in

sharing their experiences. I have been very fortunate to receive invaluable lessons especially

during my stay in Arizona, for their valuable support and assistance. ZONA technology has

been very kind to let me use their facilities during the final stage of the present work

I wish to express my sincere thanks to Dr. Lei Tang, Dr. Luciano Amaury dos Santos,

and Dr. Maurício Guimarães Silva in helping me processing some of the results presented in

this work.

I would like to thank Dr. Dario Baldelli, Dr. Reggie Chang, Dr. Xiaowei Gao, Mr.

Darius Sarhaddi, Ms. Jennifer Sherr for their personal attention and help in providing me the

best conditions during my stay in Arizona.

I would like to express my gratitude to my colleagues Major Antonio Carlos Ponce

Alonso, and Major Guilherme Augusto Vargas Cesar, who provide me the conditions for the

development this work

And finally I would to express my profound gratitude to my parents, Arnaldo and

Irene and to my wife Michelle, who guided me with encouragement.


ABSTRACT

The work presents a study of correction techniques to compute unsteady transonic

pressure distributions and aeroelastic stability in this flow regime. The methodologies herein

investigated are based on corrections of pressure distributions by the weighting of the lifting

surface self-induced downwash, resulting from aeroelastic structural displacements or

prescribed motions. A number approaches were investigated.

An investigation into the linear/nonlinear behavior of unsteady transonic flows was

also conducted. It was concluded from such investigation that unsteady transonic flows

present a linear behavior with respect to small aeroelastic structural displacements around a

steady nonlinear mean flow. Such behavior is the basis for further development of downwash

correction methods.

The correction of pressure distributions through the weighting of the lifting surface

self-induced downwash is also known as downwash weighting method. This method has been

enhanced leading to a new downwash correction technique. The procedure may be divided in

two steps, where the first step is a nonlinear steady mean flow correction, with nonlinear

pressure differences considered as reference conditions to correct the self induced downwash.

The second step is the correction of the unsteady component of the downwash, where the

corresponding reference unsteady pressure differences are predicted by a linear aerodynamic

model, based on the potential flow equations.


This extended downwash correction method led to a rational formulation named as

“successive kernel expansion method” (SKEM). The unsteady pressures and aeroelastic

stability boundaries computations using such method led to good agreement with

experimental measurements. This procedure is a rapid form to compute the transonic flutter

speed boundaries, compared to computational aeroelasticity and experimental techniques.


RESUMO

O trabalho apresenta um estudo sobre técnicas de correção para calcular distribuições

de pressão não lineares e não estacionárias, assim como a estabilidade aeroelástica em

escoamentos transônicos. As metodologias investigadas neste trabalho são baseadas na

correção de distribuições de pressão através da ponderação das velocidades normais (ou

downwash) induzidas pelos deslocamentos estruturais de natureza aeroelástica, ou devido a

movimentos prescritos. Diferentes formas de aproximação foram investigadas.

Investigou-se o comportamento linear/não-linear de escoamentos transônicos não

estacionários. Conclui-se de tal estudo que os escoamentos transônicos não estacionários

apresentam um comportamento linear, com respeito às pequenas deformações estruturais de

natureza aeroelástica, ao redor de um escoamento estacionário não linear. Tais observações

relacionadas à este comportamento serão base para os desenvolvimentos adicionais dos

métodos de correção.

A correção das pressões através da ponderação do vetor de velocidades normais

induzidas é também conhecido como um método de ponderação do downwash. Este método

foi aperfeiçoado, levando a um novo procedimento de correção de downwash. Tal

procedimento pode ser divido em dois passos, sendo o primeiro baseado em uma correção do

escoamento estacionário não linear médio, onde diferenças de pressão não lineares serão

consideradas condições de referência para corrigir o downwash. O segundo passo consiste na

correção da parcela não estacionária do downwash, onde as pressões de referência são não
estacionárias e calculadas a partir de um modelo aerodinâmico linear baseado na equação do

escoamento potencial.

Esta extensão do método de correção do downwash resulta em uma formulação

racional, denominada como “método de expansão sucessiva do kernel” (successive kernel

expansion method - SKEM). Tal método apresentou bons resultados tanto com relação ao

cálculo de pressões não estacionárias quanto à estabilidade aeroelástica, em concordância com

resultados experimentais. Este procedimento é uma forma rápida de calcular as fronteiras de

estabilidade aeroelástica no regime transônico, quando comparado com métodos de

aeroelasticidade computacional ou técnicas experimentais.


CONTENTS

List of Figures …………………………………………………………..……………… IV

List of Tables ….…………………………………...…………………………….….… XII

List of .Symbols ………………………………………………………………………. XIII

Chapter 1 – Introduction …………….………………………………..……………...…... 1

1.1 – Background ...........…….………………………………………………………. 1

1.1.1 Aeroelastic Stability Behavior in Transonic Flow................................ 1

1.1.2 Approximate Solutions.......................................................................... 5

1.1.3 Force Matching Methods....................................................................... 7

1.1.4 Pressure Matching Methods................................................................ 10

1.1.5 Dau-Garner Type Methods................................................................... 13

1.1.6 Modal Aerodynamic Influence Coefficient Matrix (MAIC) Methods .15

1.2 – Analytical Review of Correction Methods.......................................................... 19

1.2.1 Application of Approximate Solutions for Transonic Flow

Computation......................................................................................... 19

1.2.2 Proposal of a Correction Technique..................................................... 24

1.3 – Objective of the Present Work.....………………………………………...…… 26

1.4 – Outline of the Thesis...............................................…………………………… 27

Chapter 2 – Unsteady Aerodynamic Modeling and Aeroelastic Analysis.……………... 29

2.1 – Integral Solution of the Linearized Potential Flow Equation……..…………... 29

2.2 – Discrete Element Kernel Function Methods …………………………….…… 35

2.3 – The Aerodynamic Influence Coefficients Approach.......................................... 37

2.4 – Aeroelastic Model............................................................................................... 40

2.4.1 – Aeroelastic Equations of Motion......................................................... 40

2.4.2 – Modal Approach ................................................................................. 43

I
2.4.3 – Simple Harmonic Motion Approach.................................................... 44

2.4.4 – Interconnection Between Aerodynamics and Structures..................... 45

2.5 – Flutter Solution Techniques................................................................................ 47

2.5.1 – Flutter Computation Based on the k-method........................................ 48

2.5.2 – The p-k method..................................................................................... 49

2.5.3 – The g-method........................................................................................ 51

2.5.4 – p-method for the Solution of a State Space Aeroelastic System.......... 53

Chapter 3 – Unsteady Transonic Flow Behavior……………………..………………..... 57

3.1 – Linear/Nonlinear Behavior Investigation...………………..…………….…….. 57

3.2 – The Navier Stokes Model....................…………..…………………….…….... 58

3.2.1 – Governing Equations........................................................................... 58

3.2.2 – Numerical Method............................................................................... 60

3.2.3 – Turbulence Model................................................................................ 66

3.2.4 – Boundary Conditions........................................................................... 67

3.3 – Transonic Flow Investigations of a Low Aspect Ratio Wing............................. 69

3.3.1 – Preliminary Considerations..................................................................69

3.3.2 – Validation of the Computational Procedure for Steady Flow

Computations....................................................................................... 70

3.3.3 - Validation of the Computational Procedure for Unsteady Flow

Computations....................................................................................... 72

3.3.4 – Investigation Methodology.................................................................. 76

3.3.5 – Analysis of the Linear/Nonlinear Behavior......................................... 77

Chapter 4 – Correction Methods for Approximate Nonlinear Aerodynamic Modeling.... 85

4.1 – Developments on the Pressure Matching Approach................…………….….. 85

II
4.1.1 - Downwash Weighting Method Based on Steady Reference

Conditions............................................................................................. 85

4.1.2 - Downwash Weighting Method Based on Unsteady Reference

Conditions............................................................................................. 89

4.2 – Extension of the Downwash Weighting Method Based on the Kernel Expansion

Hypothesis.……………………………............................................................. 92

4.2.1 – Motivation............................................................................................ 92

4.2.2 – A Modified Downwash Correction Method........................................ 95

Chapter 5 – Results and Discussion .......………...……………...……………….…….. 101

5.1 – Evaluation of Giesing´s Method for Transonic Flutter Computation................ 101

5.2 – Downwash Weighting Methods........................................................................ 107

5.2.1 – Investigation of the AGARD Wing Unsteady Pressure Distribution .108

5.2.2 – AGARD Wing Aeroelastic Stability Analysis................................... 134

5.3 – Evaluation Of the Successive Kernel Expansion Method................................. 143

5.3.1 – Preliminary Considerations................................................................ 143

5.3.2 – Validation of Unsteady Pressures Computations............................... 144

5.3.3 – Validation of Flutter Boundary Computations ................................. 165

Chapter 6 – Conclusions and Final Remarks……...……………...……………….…… 176

6.1 – Conclusions....................................................................................................... 176

6.2 –Contributions...................................................................................................... 181

6.3 – Recommendations for Future Work.................................................................. 182

Chapter 7 – References ….……………………………………….…………………...... 184

Appendix A – Giesing´s Force Matching Method.........................……….……………. 199

Appendix B – The Dau-Garner Method .…………...................……………..……..….. 202

III
LIST OF FIGURES

Figure 1.1 The transonic dip phenomenon. (taken from Tijdeman, 1977)................................ 3

Figure 3.1- Finite differences computational mesh surrounding the F-5-NLR wing and its

profile...................................................................................................................................... 71

Figure 3.2: Steady pressure distribution over the F-5 NLR wing at M∞ = 0.95..................... 71

Figure 3.3: Real and imaginary parts of the pressure coefficients over the F-5 NLR wing

(M=0.95, 35.2% spanwise station, frequency = 40 Hz)......................................................... 74

Figure 3.4: Real and imaginary parts of the pressure coefficients over the F-5 NLR wing

(M=0.95, 72.1% spanwise station, frequency = 40 Hz)......................................................... 74

Figure 3.5: Real and imaginary parts of the pressure coefficients over the F-5 NLR wing

(M=0.95, 97.7% spanwise station, frequency = 40 Hz)......................................................... 75

Figure 3.6: Differential pressure coefficient time history over F-5 wing (M=0.95, ∆α=1.0º;

k=0.1)..................................................................................................................................... 78

Figure 3.7: Frequency contents (amplitudes and phases) of differential pressure coefficient

over F-5 wing (M=0.95, ∆α=1.0º; k=0.1).............................................................................. 78

Figure 3.8: Effect of dynamic angle of attack on unsteady aerodynamic coefficients and shock

motion (amplitudes and phases; station 35.5%; k=0.2)......................................................... 79

Figure 3.9: Effect of dynamic angle of attack on unsteady aerodynamic coefficients and shock

motion (amplitudes and phases; station 72.1%; k=0.2)......................................................... 81

Figure 3.10: Effect of dynamic angle of attack on unsteady aerodynamic coefficients and

shock motion (amplitudes and phases; station 97.7%; k=0.2).............................................. 82

IV
Figure 3.11: Boundaries for linear behavior based on the CM criteria.................................... 83

Figure 3.12: Boundaries for linear behavior based on the shock displacement criteria.......... 83

Figure 5.1: Sketch of the AGARD 445.6 wing and the corresponding discrete element

aerodynamic model................................................................................................................ 102

Figure 5.2 : Comparison between the unsteady pressures obtained from the Giesing´s method,

ZONA 6 and CFL3D - AGARD 445.6 wing, weakened model #3, M∞= 0.96..................... 104

Figure 5.3 : Flutter boundaries (dynamic pressure and flutter frequency) for the AGARD wing

445.6, weakened model #3..................................................................................................... 105

Figure 5.4 : Sketch of the PAPA wing and the corresponding discrete element aerodynamic

model...................................................................................................................................... 105

Figure 5.5 : Flutter boundaries (dynamic pressure and flutter frequency) for the PAPA wing,

at a reference angle of attack of α = 1.0o............................................................................... 106

Figure 5.6 : Flutter boundaries (dynamic pressure and flutter frequency) for the PAPA wing,

at a reference angle of attack of α =-2.0o.............................................................................. 106

Figure 5.7: DLM paneling of the AGARD wing................................................................... 108

Figure 5.8 : AGARD wing CFD aerodynamic mesh............................................................. 109

Figure 5.9 : AGARD wing 445.6 NACA 65A004 profile.................................................... 110

Figure 5.10 : Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞=0.678, ∆α = 0.25o.......................................................................................................... 111

Figure 5.11: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞=0.678, ∆α = 0.5o............................................................................................................ 112

V
Figure 5.12: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞=0.678, ∆α = 1.0o............................................................................................................ 112

Figure 5.13: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞=0.678, ∆α = 1.5o. .......................................................................................................... 113

Figure 5.14: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞=0.678, ∆α = 2.0o............................................................................................................ 113

Figure 5.15: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞=0.678, ∆α = 2.5o............................................................................................................. 114

Figure 5.16: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞=0.678, ∆α = 3.0o............................................................................................................. 114

Figure 5.17: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞=0.901, ∆α = 0.25o........................................................................................................... 115

Figure 5.18: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞=0.901, ∆α = 0.5o............................................................................................................ 115

Figure 5.19: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞=0.901, ∆α = 1.0o............................................................................................................ 116

station
Figure 5.20: Unsteady pressure distributions for the wing at 30.8% of the span, M∞=

0.901, ∆α = 1.5o.................................................................................................................... 116

Figure 5.21: Unsteady pressure distributions for the wing station at 30.8% of the span, M∞=

0.901, ∆α = 2.0o.................................................................................................................... 117

Figure 5.22: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞=0.901, ∆α = 2.5o............................................................................................................ 117

VI
Figure 5.23: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞=0.901, ∆α = 3.0o........................................................................................................... 118

Figure 5.24: Unsteady pressure distributions for the wing station at 30.8% of the span, M∞=

0.96, ∆α = 0.25o................................................................................................................... 118

Figure 5.25: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞=0.96, ∆α = 0.5o ............................................................................................................. 119

Figure 5.26: Unsteady pressure distributions for the wing station at 30.8% of the span, M∞=

0.96, ∆α = 0.25o..................................................................................................................... 119

Figure 5.27: Unsteady pressure distributions for the wing station at 30.8% of the span, M∞=

0.96, ∆α = 1.5o....................................................................................................................... 120

Figure 5.28: Unsteady pressure distributions for the wing station at 30.8% of the span, M∞=

0.96, ∆α = 2.0o....................................................................................................................... 120

Figure 5.29: Unsteady pressure distributions for the wing station at 30.8% of the span, M∞=

0.96, ∆α = 2.5o....................................................................................................................... 121

Figure 5.30: Unsteady pressure distributions for the wing station at 30.8% of the span, M∞=

0.96, ∆α = 3.0o....................................................................................................................... 121

Figure 5.31 : Contour plot of instantaneous pressures, at spanwise station 30,8%, M∞=0.678,

α = 2.0o.....…………….......................................................................................................... 122

Figure 5.32 : Instantaneous pressure distributions, at spanwise station 30,8%, M∞=0.678,

∆α = 2.0o..........…….............................................................................................................. 123

Figure 5.33: Contour plot of instantaneous pressures, at spanwise station 30,8%, M∞=0.678,

α = 3.0o……………............................................................................................................... 123

VII
Figure 5.34: Instantaneous pressure distributions, at spanwise station 30,8%, M∞=0.678,

∆α = 3.0o..........…….............................................................................................................. 124

Figure 5.35: Contour plot of instantaneous pressures, at spanwise station 30,8%, M∞=0.901,

α = 2.0o……………............................................................................................................... 125

Figure 5.36: Instantaneous pressure distributions, at spanwise station 30,8%, M∞=0.901,

∆α = 2.0o................................................................................................................................ 126

Figure 5.37: Contour plot of instantaneous pressures, at spanwise station 30,8%, M∞=0.901,

α = 3.0o...................................................……………............................................................ 126

Figure 5.38: Instantaneous pressure distributions, at spanwise station 30,8%, M∞=0.901,

∆α = 3.0o...................................................................................................................……..... 127

Figure 5.39: Contour plot of instantaneous pressures, at spanwise station 30,8%, M∞=0.960,

α = 2.0o.............................................................................................................…………….. 128

Figure 5.40: Instantaneous pressure distributions, at spanwise station 30,8%, M∞=0.960,

∆α = 2.0o............................................................................................................................… 128

Figure 5.41: Contour plot of instantaneous pressures, at spanwise station 30,8%, M∞=0.960,

α = 3.0o...............................................................................................……………................ 129

Figure 5.42: Instantaneous pressure distributions, at spanwise station 30,8%, M∞=0.960,

∆α = 3.0o................................................................................................................................ 129

Figure 5.43: Instantaneous velocity vector fields over the leading edge of the AGARD wing

445.6....................................................................................................................................... 131

Figure 5.44. Non-dimensional phase angles of the pressure difference distributions for

spanwise station 30.8%, M∞=0.96. ....................................................................................... 132

VIII
Figure 5.45. Amplitudes of the pressure difference distributions for spanwise station 30.8 %,

Mach =0.96. .......................................................................................................................... 133

Figure 5.46: AGARD wing 445.6 results - comparison of the flutter computation results

between steady and unsteady downwash weighting methods............................................... 136

Figure 5.47: AGARD wing 445.6 results - comparison of the flutter computation results

between the unsteady downwash weighting methods and other methods............................. 137

Figure 5.48: Comparison between the flutter speed plots as a function of the amplitude of the

motion.................................................................................................................................... 140

Figure 5.49: Flutter speed behavior as a function of the amplitude of the motion................ 141

Figure 5.50: Flutter frequency behavior as a function of the amplitude of the motion......... 142

Figure 5.51 : Sketch of the F-5 wing and the corresponding discrete element aerodynamic

model mesh. .......................................................................................................................... 145

Figure 5.52 : Steady pressure distribution for the F-5 wing at M = 0.9 and α = 0°.............. 146

Figure 5.53 : Pressure coefficient ratio for the F-5 wing at M = 0.9 and ∆α = 0.5°............. 147

Figure 5.54 : Real part of the pressure distribution for the F-5 wing at M = 0.9 and reduced

frequency k = 0.275............................................................................................................... 148

Figure 5.55 : Imaginary part of the pressure distribution for the F-5 wing at M = 0.9 and

reduced frequency k = 0.275.................................................................................................. 149

Figure 5.56 : Steady pressure distribution for the F-5 wing at M∞ = 0.948 and α = 0°......... 151

Figure 5.57 : Pressure coefficient ratio for the F-5 wing at M∞ = 0.948 and ∆α = 0.5°........ 152

Figure 5.58 : Real part of the pressure distribution for the F-5 wing at M∞ = 0.948 and reduced

frequency k = 0.264............................................................................................................... 153

IX
Figure 5.59 : Imaginary part of the pressure distribution for the F-5 wing at M∞ = 0.948 and

reduced frequency k = 0.264.................................................................................................. 154

Figure 5.60 : Sketch of the LANN wing and the corresponding discrete element aerodynamic

model..................................................................................................................................... 155

Figure 5.61 : Steady pressure distribution for the LANN wing at M∞ =0.822 and α = 0.6°. 156

Figure 5.62 : Pressure coefficient ratio for the LANN wing at M∞ =0.822 and ∆α = 0.15°. 157

Figure 5.63 : Real part of the pressure distribution for the LANN wing at M∞ =0.822 and

reduced frequency k = 0.105.................................................................................................. 158

Figure 5.64 : Imaginary part of the pressure distribution for the LANN wing at M∞ =0.822 and

reduced frequency k = 0.105.................................................................................................. 159

Figure 5.65 : Sketch of the Lessing wing and the corresponding discrete element aerodynamic

model...................................................................................................................................... 160

Figure 5.66 : Steady pressure distribution for the Lessing wing at M∞ =0.9 and α = 0°....... 161

Figure 5.67 : Pressure coefficient ratio for the Lessing wing at M = 0.9 and ∆α = 0.2°....... 162

Figure 5.68 : Pressure distribution (amplitudes) for the Lessing wing at M∞ =0.9 and reduced

frequency k = 0.13................................................................................................................. 163

Figure 5.69 : Pressure distribution (phases) for the Lessing wing at M∞ =0.9 and reduced

frequency k = 0.13................................................................................................................. 164

Figure 5.70 - Flutter boundaries (dynamic pressure and flutter frequency) for the AGARD

wing 445.6, weakened model #3........................................................................................... 166

Figure 5.71 : Dynamic pressures and frequencies in the flutter boundaries of the PAPA wing,

at angle of attack, α = 1.0o.................................................................................................... 170

X
Figure 5.72 : Dynamic pressures and frequencies in the flutter boundaries of the PAPA wing,

at angle of attack, α = -2.0o................................................................................................... 171

Figure 5.73 : Sketch of the YXX wing and the corresponding discrete element aerodynamic

model...................................................................................................................................... 172

Figure 5.74 : Flutter dynamic pressure for the YXX wing, at a reference angle of attack

α=-1.0o................................................................................................................................... 174

Figure 5.75 : Flutter frequency for the YXX wing, at a reference angle of attack α=-1.0o.. 174

XI
LIST OF TABLES

Table 5.1: Flow conditions for AGARD wing 445.6 (weakened # 3 ) aeroelastic analysis.. 103

Table 5.2: Reduced frequencies for AGARD wing 445.6 aeroelastic analysis..................... 110

Table 5.3: Flutter speeds and frequencies for AGARD wing 445.6...................................... 135

Table 5.4: Flutter speeds and frequencies for AGARD wing 445.6...................................... 136

Table 5.5: Flutter speeds and frequencies for AGARD wing 445.6. until ∆α=1.5o............. 140

Table 5.6: Flutter speeds and frequencies for AGARD wing 445.6, until ∆α=3.0o............. 140

XII
LIST OF SYMBOLS

a Speed of the sound

A+ = 26.0 van Driest constant

A = ∂Eˆ ∂Qˆ Flux Jacobean matrix

A (U ∞ ) State space matrix of an aeroelastic system

A1 , A4 , A6 Auxiliary metric terms (Eq. 3.24)

[ AIC ] Aerodynamic influence coefficients matrix

b Reference length equals to the semi-chord

c = 2b Lifting surface streamwise section chord

CL Sectional lift coefficient

CM Sectional moment coefficient about the quarter-chord

cp (in Prandtl number) Specific heat at constants pressure

Cp Pressure coefficient

cr Root chord

d Distance from the wall

[ D] Kernel function matrix

[ D ]n , n = 1, m Constant coefficient matrices, obtained from the asymptotic series

expansion

Dij Element of the kernel function matrix

Dˆ en ,n +1 Explicit fourth order dissipation

e Total energy per unit volume

Eˆ , Fˆ , Gˆ Inviscid flux vectors

XIII
[F ]. Substantial derivative operator

g =γ k Damping factor

[G ] Spline matrix operator

gs . artificial structural damping

h Displacement mode shape vector

[H ] Aerodynamic transfer function matrix

[I ] Identity matrix

i Index corresponding to a receiving point in the discrete element

kernel function aerodynamic model

i Imaginary number ( i = − 1 )

j Index corresponding to a sending point in the discrete element

kernel function aerodynamic model

J Jacobian transformation between the curvilinear and cartesian

systems

[K ] Structural stiffness matrix

 K  = [ Φ ] [ K ][ Φ ] Structural modal stiffness matrix


T

k = ωb U ∞ Reduced frequency based on the free-stream flow speed

k (in Prandtl number) Thermal conductivity

kr = ω c a∞ = k ⋅ ( M ∞ c b ) Reduced frequency based on the speed of the sound

Kϕ Kernel of the integral relation in term of the velocity potential

Kψ Kernel of the integral relation in term of the acceleration potential

L Laplacian operator

XIV
{L} Total aerodynamic load vector

{La } Induced aerodynamic load vector

{Le } External aerodynamic load vector

m Mixing length

[M ] Structural mass matrix

 M  = [ Φ ] [ M ][ Φ ] Structural modal mass matrix


T

M∞ Free-stream Mach number

O ( ∆τ ) Order of error in finite difference

approximation

p Pressure

p (in flutter solution) = γ k + ik , aeroelastic eigenvalue

Pr = µ c p k Prandtl number

PrT = µT c p kT Turbulent Prandtl number

[Q ] Aerodynamic loads influence coefficients matrix

Q̂ vector of unknown flow properties

q dynamic pressure

{q} Generalized coordinates

Qap  Aerodynamic loads influence coefficients matrix approximated by

a Padé polynomial

Q j  Matrix coefficients of the Padé polynomial

{q } lag
Lag states of an aeroelastic system

XV
R = X 2 + β 2 (Y 2 + z 2 ) distances between the sending and receiving points in the discrete

element kernel function model

Rˆ , Sˆ , Tˆ Viscous flux vectors

R5, S5, T5 Auxiliary variables for vicous fluxes, given in, Eq. (3.7)

Re = ρ∞ a∞ c µ∞ Reynolds number based on free-stream speed of sound

Re1 = ρ∞U ∞ cr µ ∞ Reynolds number based on free-stream flow speed

Re( ), Im( ) Real and imaginary parts of a complex number

s = iω . Laplace variable

sb Nondimensional Laplace variable


s=
U∞

S Lifting surface area

[S ] Integration matrix

t time

Tξ , Tζ . Eigenvectors (transformation) matrices

{u } physical displacement vector of the structure

u,v,w cartesian components of velocity

Uc, , Vc , Wc Contravariant components of velocity

{w} Downwash vector

{w} Downwash vector divided by amplitude of the quasi-steady motion

amplitude ∆α .

W Wake surface area

[WT ] Weighting matrix operator

x Cartesian coordinate, aligned with the undisturbed flow speed

xs , ys ,zs Cartesian coordinates in the structural model coordinate system

XVI
y Cartesian coordinate, aligned with the spanwise direction

z Cartesian coordinate, normal to the lifting surface

α Angle of attack

β = 1− M ∞2 Prandtl- Glauert constant

βn Lag parameter

γ Ratio of specific heats

γ Damping decaying rate

δ Difference operator corresponding to standard second order central

difference

δ Difference operator corresponding to half-point central difference

{δα∆C p } Dimensionless pressure coefficient disturbance rate vector

{δ h} Vector of the virtual displacement at the aerodynamic control

points

{δ h } = ∆α
x
Vector of the quasi-steady disturbance in effective angle of attack

{δu } Vector of the virtual displacement at the structural nodes

∆α Amplitude of the dynamic angle of attack

{δ∆C } nl
p
Vector of the quasi-steady pressure coefficient differences

∆ξ , ∇ξ Forward and backward difference operators

∆ϕ Doublet singularity strength

∆τ Computational non-dimensional time step

εE Coefficient of explicit numerical dissipation

εI Coefficient of implicit numerical dissipation

XVII
ζ Coordinate in curvilinear system, aligned with the spanwise

direction

η Coordinate in curvilinear system, normal to the lifting surface,

κ von Kármán constant,

{λ} State space variables vector of an aeroelastic system

Λξ , Λζ contents the eigenvalues of the matrices A and C

µ Molecular viscosity

µT Eddy viscosity

ξ Coordinate in curvilinear system, tangent to the lifting surface

boundary, in the streamwise direction.

ξ ,η Planar position of doublet (discrete kernel function method)

ρ Density

σ Point source singularity strength

τ = a∞t c non-dimensional time

τ max Maximum shear stress in the boundary layer

τ wall Shear stress in the wall

τ xx ,τ xy ,τ xz ,τ yz ,τ yy ,τ zz Shear stresses

φ Perturbation velocity potential

φ1 , φ2 , φ3 Auxiliary variables in numerical dissipation

{φ} Vector mode shape (columns of the of the modal matrix)

[Φ ] Modal transformation matrix, composed by mode shape vectors

ϕ Frequency domain velocity potential

ϕD Unsteady doublet singularity potential

XVIII
ϕS Unsteady source singularity potential

ψ Acceleration potential

ω Oscillation frequency associated to the harmonic motion.

ω mean vorticity

Subscripts

A aerodynamic

ap approximate

e external

i,j,k indexes indicatign a point in the computational fluid dynamic grid

i imaginary

I, and F refer to an initial position and a final position respectively

LE leading edge

nlag number of lag states

R real

s structural

T turbulent

t Partial derivative with respect to time

TE Trailing edge

x Partial derivative with respect to x

y Partial derivative with respect to y

z Partial derivative with respect to z

∞ condition at free-stream

XIX
Superscripts

aero Aerodynamic

l lower

n, n + 1 time levels

nl nonlinear

str structural

u upper

XX
Chapter 1 – Introduction 1

1 - INTRODUCTION

1.1 Background

1.1.1 Aeroelastic Stability Behavior in Transonic Flow

The most common way in performing aeroelastic analyses in the aeronautical industry

is by the use of the structural-dynamic/unsteady aerodynamic commercial codes based on

linear unsteady aerodynamic modeling techniques combined with a structural dynamics

solver. Usually, linear aerodynamic modeling techniques are based on discrete element

solutions of the linear equations of the fluid flow. Discrete element methods, also named as

panel methods, are defined as numerical procedures which require the body to be discretized

into many small elements, also named as panels. Each discrete element contains a control

point where the boundary condition is imposed.

The fluid flow is represented by the linearized potential flow equation (Landahl,

1961), which has an integral solution in terms of the source and doublet singularity

distribution over the body surface (Chen et al. 1993). The resulting velocity potentials over

each of the elements are related to control point displacements through the boundary

conditions which requires the flow to be tangential to the body at all times. This condition is

represented by the substantial derivative of the body displacements leading to induced

velocities, normal to the body, when the tangential flow condition is satisfied.

These velocities are named as downwash, which depends on the pattern of the

displacements of the body undergoing an unsteady motion. Once the velocity potentials are

computed by the integral solution, the derivatives of those potentials with respect to the

direction normal to the body surface, lead to induced downwash, which are related to the
Chapter 1 – Introduction 2

downwash due to the body motion by the boundary condition. The remaining unknown is the

strength of the elementary singularity distribution along each of the elements.

However, the solution of the same linearized potential flow equation, in terms of the

acceleration potential instead of the velocity potential would greatly simplifies the

aerodynamic modeling. This is so because the acceleration potential, is the linearized

pressure, thus the body wake domain would not be discretized because there is no pressure

jump across the wake. Therefore, the solution of the aerodynamic model in terms of

elementary integral solution would be reduced to a pressure to downwash relationship, where

the assembly of these elementary integral solution gives a matrix whose coefficients represent

the aerodynamic influence of the discrete elements pulsating pressures to control point

displacements.

The resulting system of equations can be expressed as a linear operator applied to a

downwash vector resulted from the substantial derivatives of the body motion, in other words,

from the boundary conditions. The matrix which defines this linear operator is named as the

aerodynamic influence coefficient matrix (AIC) and its dimension is the same as the number

of panels. The body undergoing unsteady motions may be represented by the substantial

derivatives of control points displacements accordingly a mode shape pattern. Therefore, the

resulting downwash due to the body modal motion will be defined hereafter as downwash

mode shape. Summarizing, the mode shape displacements are related to known boundary

conditions, leading to a downwash vector. Furthermore, the AIC matrix is also known and is

obtained from the assembly of the integral solutions of the linearized potential flow equation.

Thus, the solution of the unsteady flow may be represented by the resulting pressure

distribution computed from the application of the AIC matrix operator to the downwash

vector.
Chapter 1 – Introduction 3

Examples of discrete element methods are the doublet lattice method – DLM (Albano

and Rodden, 1969), the unsteady subsonic vortex lattice method – UVLM (Soviero and

Bortolus, 1992), the constant pressure method –CPM (Appa, 1987), the harmonic gradient

method –HGM Chen and Liu, 1985), and others.

The application of these aforementioned methods are limited to purely subsonic or

supersonic flows, since the governing equations over which the method were developed are

based on a linearized unsteady potential flow hypothesis. However, the aeroelastic behavior

of an aircraft is typically critical in the transonic flight regime, where nonlinear phenomena

related to embedded moving shock waves and viscosity, play an important role in aeroelastic

stability. As discussed by Ashley (1980), the shock wave movement and strength profoundly

affects the flexure-torsion flutter mechanism. One consequence of this behavior is the so

called “transonic dip”, which is graphically represented in Figure (1.1).

1
linear theory
measurements
Dimensionless Flutter Speed

0.95

0.9

0.85

0.8

0.75
0.7 0.8 0.9 1 1.1 1.2
Mach Number

Figure 1.1 Flutter speed boundary as a function of the freestream Mach number, exposing the

transonic dip phenomenon.


Chapter 1 – Introduction 4

This phenomenon is characterized by a decrease of the slope of the flutter speed plot

as a function of the Mach number, when compared to the same plot obtained from a linear

aeroelastic analysis. Therefore, it is necessary to pay special attention to the flutter

phenomenon under these circumstances. This is especially significant as most modern aircraft

fly under transonic flow conditions.

One of the feasible alternatives for analyzing the aeroelastic stability in nonlinear flow

conditions is by the use of time accurate computational fluid dynamic (CFD) solutions of the

nonlinear fluid equations coupled with structural dynamic representation of the vehicle.

Another approach is obtained by wind tunnel testing of aeroelastic models, under tronsonic

flow conditions. However, wind tunnel testing for aeroelastic investigation regarding flutter

boundary computation is not usual because this class of experiments involves expensive

models and high operational costs. In most cases of experimental aeroelastic stability

investigation, the objective is to obtain accurate measurement data to validate computational

procedures applied to the solution of nonlinear equations. Examples of experiments on

unsteady aerodynamics regarding pressures measurements are reported by Lessing et al..

(1960), Tijdeman et al.. (1978), and Malone et al. (1983). Notable examples of experimental

flutter investigations were presented by Isogai (1983), Yates (1988), Farmer et al.. (1988),

and Hong et al. (2003). The other way to evaluate the transonic aeroelastic behavior, is from

flutter flight testing, which is the most hazardous and expensive option in terms of operational

costs. This approach may be used either for experimental flutter boundaries identification or

to verify the aeroelastic subcritical aerodynamic damping at specific flight envelope points to

validate aeroelastic numerical models.


Chapter 1 – Introduction 5

1.1.2 - Approximate Solutions

There have been several attempts to solve the transonic aeroelastic problem using

combined procedures which relate linear models to measured data for the correction of

unsteady linear aerodynamic models. A good review on correction techniques was presented

by Palacios et al. (2001), describing the most employed methods concerning transonic flutter

prediction via combined procedures. Such approaches are named here as mixed procedures.

The purpose of such procedures is to correct the linear aerodynamic models to take into

account nonlinear effects, unpredictable by the linearized potential-based equations of the

fluid flow.

The methodologies for the solution of the transonic aeroelastic problem based on

mixed procedures are referred also as semi-empirical corrections. These corrections can be

performed by the multiplication, addition or the whole replacement of the AIC matrix. This

approach is an adequate tool for engineering applications, because the methodology employed

is less expensive than direct use of CFD techniques. The correction techniques, which have

been applied to unsteady loading calculation for static or dynamic stability analysis, are here

classified in four major groups:

– Force matching methods - (Yates, 1966; Giesing et al., 1976; Zwaan, 1985; Pitt

and Goodman, 1987; Brink-Spalink and Bruns, 2001 );

– Pressure matching methods - (Rodden and Revell, 1962; Bergh, 1964; Bergh and

Zwaan, 1966; Luber and Schmid, 1982; Baker, 1997; Baker et al., 2000; Jadic et

al., 1999; Jadic et al., 2000);

– Dau-Garner type procedures – (Garner, 1977; Dau, 1992; SenGupta, 1996;

Schulze and Vogel, 1989; Yonemoto, 1985);


Chapter 1 – Introduction 6

– Modal aerodynamic influence coefficients (MAIC) matrix replacement – (Suciu et

al., 1990; Suciu, 2000; Suciu, 2003; Liu et al., 1988; Chen et al., 1996 ).

The main idea of the first procedure is to match reference nonlinear forces and

moments, which may be obtained from experiments or accurate CFD solutions of the

nonlinear flow governing equations. In this case, nonlinearities such as pressure jumps due to

shock waves and viscous effects are embedded in such reference quantities. The second form

to proceed with the correction is the matching of the pressures taken as reference conditions.

The same nonlinearities are present in the reference conditions, in this case pressure

distributions. The Dau-Garner class of methods are a procedures where the unsteadiness of

the resulting nonlinear corrected pressures are computed based on steady nonlinear

information with the aid of semi-empirical relations. The procedure which named this class of

methods will be further detailed in sub-section 1.1.5, and its formulation is presented in

Appendix B.

Finally, the MAIC procedures consist in generating a modal aerodynamic influence

coefficient matrix referred to measured or computed nonlinear pressures or loading due to

given modal displacements of the lifting surface. This new matrix is then substituted in the

aereolastic equations of motion, where the generalized unsteady aerodynamic forces are

related to the associated modal displacements of the lifting surface, or downwash mode

shapes. The aeroelastic analysis is performed taking into account transonic flow effects

embedded in such new modal aerodynamic influence coefficient matrix. In the next sub-

sections, brief descriptions of the aforementioned methods will be presented, so that it will be

clear how they compare to each other.


Chapter 1 – Introduction 7

1.1.3 - Force Matching Methods

One of the first approaches for developing correction techniques for matching

experimental forces was introduced by Yates (1966). His method is based on a modification

of the strip theory, which is based on Theodorsen’s (1934) typical section unsteady loading

solution. The objective is to take into account nonlinear effects, due to both viscous and

transonic phenomena. This is performed by the replacement of the lifting and moment

coefficient derivatives, obtained from Theodorsen’s (1934) method by the same quantities

determined from measurements. This method is named as the modified strip theory (Yates,

1966) and the published results were in good agreement with the experiments for the case of a

wing with high aspect ratio.

Giesing et al.(1976) developed a method to correct the subsonic discrete lifting surface

theory based on a reference experimental or nonlinear computed loading. Their method

included a procedure to obtain a pre- or post-multiplicative operator to correct the AIC matrix

by the use of weighting factors. The purpose of such operators is to match a single reference

steady spanwise loading of the lifting surface.

Each spanwise station corresponds to a set of panels aligned in the streamwise

direction. However, there are less known quantities, i.e., section loads, than the necessary

weighting factors since the number of the spanwise stations is smaller than the number of

panels. In order to circumvent this problem, the weighting factors are obtained using the least

squares method. With such method, the pressures associated to the reference nonlinear

loading are not restored after the multiplication of the AIC matrix by the weighting factors.

However, the corrected loading condition will be identical to the reference loading condition.

The motivation for the development of Giesing et al.procedure (1976) is the

requirement of a safe prediction of the control surface flutter, because the employment of
Chapter 1 – Introduction 8

potential theory based methods usually underestimates this type of flutter. The reason for this

underestimation is related to absence of representing nonlinear effects, such as, the boundary

layer thickening. This correction factor technique was successfully applied for this class of

problems in subsonic cases. In the transonic flow situation, there are important effects yet to

be considered, such as the control surface buzz and the transonic flutter on the control surface

itself. However, the authors of the cited reference did not present results regarding those

situations.

Zwaan (1985) presented an evaluation of methods to calculate unsteady transonic

loads for aircraft design and flutter analysis. He reviewed a set of correction methods based

on the matching of experimental or computed spanwise lift and moment distributions. All of

them are based on matching those spanwise loading quantities. The differences between the

evaluated methods are related to the correction of the three-dimensional effects, the use of

two-dimensional and three-dimensional reference data and the origin of the nonlinear data,

i.e., either from experiments or from nonlinear computations. Most of the reference nonlinear

loadings were based on wind tunnel measurements. However, Zwann also employed

nonlinear loading resulting from steady or unsteady finite difference solutions of the two-

dimensional transonic small disturbance equation. Zwaan concluded from the evaluation of

such procedures that they require further improvement. The aeroelastic stability behavior, in

some cases, did not agree with the test results in a quantitative sense. However, qualitatively,

all the methods predicted the flutter speed dip. All of them were clearly dependent on the

reference flowfield conditions. These conditions are effects such as reference angle of attack,

wind tunnel test section interference, and whether the computed reference nonlinear loading is

of three-dimensional or two-dimensional nature.

Pitt and Goodman (1987) presented a method for correcting a linear stripwise AIC

matrix, which relates the strip displacements to their loads. Their procedure is based on a
Chapter 1 – Introduction 9

post-multiplication of the stripwise AIC matrix by weighting factors. The weighting factors

are obtained from the inversion of this AIC matrix multiplied by the reference loading

condition. In turn, the reference loading is obtained from the solution of the transonic small

disturbance equations in a steady flow condition. Pitt and Goodman (1987) presented results

in which the transonic dip phenomenon was captured, and the computed flutter speeds were in

agreement with experiment. Some small differences between computed and experimental

results were attributed to viscous effects by the authors.

Brink-Spalink and Bruns (2001) presented an extension of the method of Giesing et

al.(1976). Instead of using a simple least square procedure to match desired loads, as

performed by Giesing et al.(1976), these authors introduced a weighted least squares

procedure, based on a linear quadratic optimization (Brink-Spalink and Bruns, 2001). The

objective of such optimization procedure is to minimize the least squares deviation from the

theoretical linear AIC matrix, while matching multiple loading conditions. Therefore, another

feature added in their procedure, when compared to Giesing’s et al. (1976) method, is the

possibility to include different sets of loads due to distinct displacement conditions. These

displacement conditions can range from a frozen elastic mode shape to a control surface rigid

rotation, for example. Brink-Spalink and Bruns’ (2001) procedure presents better chances to

guarantee the matching of the desired loading condition without severe distortion in the

pressure distributions. This is so, because the uncorrected AIC matrix is taken as reference for

the minimization condition. In other words, the goal of the minimization process is to

compute a weighting matrix which does not distort severely the AIC, preserving interference

effects as much as possible. However, the shock jump conditions for the case of transonic

flow situations cannot be restored, because the AIC matrices computed from a linear

aerodynamic model does not have the capability to predict nonlinearities. The resulting

weighting operator will be a fully populated matrix, instead of the diagonal operator obtained
Chapter 1 – Introduction 10

by Giesing et al.(1976) procedure. The authors did not present conclusive results with regard

to flutter computation. They only presented comparisons between the resulting corrected AIC

matrix and the original one.

1.1.4 - Pressure Matching Methods

Rodden and Revell (1962) presented a procedure based on the pre-multiplication of

the AIC matrix by a diagonal weighting operator. This operator is obtained from the ratio

between the experimental and theoretical steady pressures. In a similar approach, Bergh(1964)

and Bergh and Zwaan (1966) correct the AIC matrix by the use of a diagonal pre-multiplying

weighting operator. However, these authors extended Rodden and Revell´s procedure to

compute the weighting operator based on the ratio between measured and theoretical unsteady

pressures, which are related to a given downwash mode shape. Their main concern is to verify

if is possible to extent the use of a unique weighting matrix to restore unsteady pressures for

different mode shapes and reduced frequencies. Their conclusions are that this extension is

admissible for different mode shapes, however, around a small reduced frequency variation.

Luber and Schmid (1982) presented a method for the aeroelastic analysis in transonic

flow based on a procedure similar to the one presented by Rodden and Revell (1962).

However, Luber and Schmid (1982) employ the ratio between the nonlinear and the linear

quasi-steady pressure slopes as opposed to the ratio between the pressures, as performed by

Rodden and Revell (1962). The resulting ratio of the pressure slopes will compose a diagonal

weighting operator, which role is to match a nonlinear pressures distribution correcting the

linear pressures distribution. Luber and Schmid (1982) perform this correction in two forms,

as a multiplicative weighting operator or as an additive weighting operator. These weighting

operators are used to corrected the AIC matrix as a multiplying diagonal operator or as an
Chapter 1 – Introduction 11

incremental contribution to be added to the AIC matrix, respectively. Luber and Schmid

(1982) presented results regarding unsteady pressures and flutter computations. The computed

unsteady pressures results showed a wrong trend when compared to experimental data. The

flutter computation also did not present good results. The computed flutter speeds were

nonconservative when compared to experimental measurements.

Another recent approach for transonic flutter analysis was presented by Baker (1997),

and it was later extended for static aeroelastic analysis by Baker et al. (2000). Baker and his

co-workers perform an additive correction based on the increment of the linear AIC matrix by

a frequency dependent nonlinear term. This nonlinear term is obtained from a nonlinear

unsteady CFD computation of the lifting surface undergoing a impulse-type displacement.

The pressures are computed from a discrete Fourier transform of the aerodynamic response in

the time domain. The resulting frequency dependent unsteady pressures are used to obtain the

nonlinear term of the AIC matrix, at each of the reduced frequencies of interest. Each of the

frequency dependent nonlinear terms are obtained by the scaling of those pressures by the

chosen reference mode shape, resulting in an additive diagonal operator. This method requires

a refined time marching CFD solution to get accurate frequency response results. Since

Baker´s (1997) procedure employs an unsteady CFD solution, it is a more expensive approach

to correct the linear aerodynamics. Furthermore, Baker’s (1997) results present a fairly good

agreement with the experiments in the case of flutter computations for a standard aeroelastic

wing configuration.

Another extension of the Giesing et al. (1976) correction method was presented by

Jadic et al.. (1999, 2000), who developed a procedure which takes pressures instead of loads

as reference conditions. Furthermore, their procedure admits the employment of multiple

reference pressure conditions, such as the method of Brink-Spalink and Bruns (2001). Each of

these reference pressure conditions is associated to a reference mode shape. The main
Chapter 1 – Introduction 12

difference between Jadic’s et al. (1999, 2000) and Brink-Spalink and Bruns’ (2001) methods

is how the weighting operator is obtained.

The main concern of Jadic et al.. (1999, 2000) procedure is to avoid the use of

minimization techniques for the determination of the weighting factors. However, usually the

number of the given reference steady pressures is smaller than the order of the AIC matrix.

Therefore, the resulting pressure matrix, which contains the given pressures due to prescribed

mode shapes will be a rectangular matrix. At this step, the computation of the weighing

operator is performed by the inversion of the matrix of given pressures. However, since it is a

rectangular matrix, the computation of the inversion requires a minimization technique.

Jadic’s et al.. (1999, 2000) suggestion to avoid the minimization technique is based on the

computation of a set of linear steady pressures based on a linearly independent artificial mode

shape basis. Since these complementary pressure conditions are known, they will be included

in the given reference pressure matrix. Then, a modified square pressure matrix is

constructed, in which each of its columns are related to a given modal displacement. These

displacements, on the other hand, can be either associated to the reference mode shapes or to

the artificially generated ones. The resulting pressure matrix is inverted and multiplied by an

exclusively linear pressure matrix, constructed based on the same modal basis of the modified

pressure matrix. The resulting weighting operator will be a fully populated matrix to be

multiplied to the AIC matrix to match the reference pressure conditions.

As discussed by Jadic et al. (1999, 2000), their procedure presents a good strategy to

avoid the severe pressure distortion typical of the Giesing et al. (1976) method, because the

DLM interference effects are not strongly modified. Some subsonic flutter analysis results

were presented by the authors showing a good agreement between the computed data and the

experiment. However, further extension for solving transonic flutter problems (Jadic et al..

2001), based on steady nonlinear given pressure modes, did not present good results. The
Chapter 1 – Introduction 13

reason for the bad results are because transonic flow nonlinearities cannot be restored, since

the additional artificial mode shapes are generated from the DLM method.

A disadvantage of this method, either for static aeroelasticity or flutter calculations is

the need of additional CFD computations on a set of artificially deformed lifting surface mode

shapes, in order to generate sufficient nonlinear reference conditions to compute an effective

weighting matrix. On the other hand, only a single simulation is necessary to compute the

reference conditions for obtaining a diagonal weighting matrix, which would be sufficiently

effective to perform the correction.

1.1.5 - Dau-Garner Type Methods

This class of correction methods has been developed with the objective of computing

approximate nonlinear unsteady pressures in the frequency domain to be used for flutter

calculation in transonic flow. The computation of these nonlinear pressures is performed with

the aid of semi-empirical relations, taking as reference steady state nonlinear pressure

distributions.

The first approach in this class of procedures was introduced by Garner (1977), who

presented the theoretical background of such methods. The extension of the theory was

performed by Dau (1992), resulting in the well-known Dau-Garner method (SenGupta, 1996).

The objective of Dau (1992) is to develop an expedient procedure for transonic aeroelastic

stability analysis and unsteady loads calculation for industrial applications. The mathematical

formulation of the method is presented in Appendix B.

The basic hypothesis of the Dau-Garner method (Dau, 1992) is an empirical relation

introduced by Garner (1977), where it is assumed that the ratio between the unsteady and

steady velocity potential gradients remains the same whether the flow is transonic or purely
Chapter 1 – Introduction 14

subsonic. Once this hypothesis is assumed, it is possible to express the unsteady velocity

potential gradient as a function of the given steady state velocity potential gradient. The ratio

between the linear unsteady and steady velocity potential gradients can be computed from a

linear theory such as the doublet lattice method.

The additional condition to be considered is the relation between unsteady pressures

and the unsteady velocity potential, which may be derived from the Euler momentum

equations, in a small disturbances context. The complete derivation of such relation is

presented by Dau (1992). When the hypothesis due to Garner (1977) is considered, one

should note that the unsteady velocity potential gradient is a function of the steady state

velocity potential gradient. Likewise, the steady velocity potential gradient can be written as a

function of a steady state pressure distribution, as given by the pressure to velocity potential

relation, which was derived from the Euler momentum equations. Therefore, the unsteady

velocity potential distribution can be written as a function of steady pressure distribution

times the ratio between the unsteady and steady velocity potential gradients.

The introduction of the nonlinearities associated with transonic flow is performed by

the substitution of the steady state linear pressures in the resulting equation by the nonlinear

ones, in such way that both real and imaginary parts of the pressures are modified with respect

to a nonlinear steady state pressure distribution.

SenGupta (1996) and Schulze and Vogel (1989) presented results of the application of

this method for unsteady pressure computations. In some of the unsteady pressure

distributions, the phase components present a discrepant behavior, when compared with

experimental measurements. In most of the results obtained by SenGupta (1996), the

computed phases were different from the experimental ones by approximately about 50

degrees. This fact indicate a shortcoming in the pressure phase calculation. Schulze and Vogel

(1989) applied the Dau-Garner method for unsteady pressures calculation of the LANN wing.
Chapter 1 – Introduction 15

Again, when their solution was compared to a finite difference computation of the transonic

small disturbance equation and also to experimental data, the computed pressures presented

wrong trends with regard to the imaginary pressure components.

Another different approach to compute unsteady pressures in transonic flow was

presented by Yonemoto (1985). His procedure is divided in two steps: an amplitude

correction, such as the procedure presented by Luber and Schmid (1982), and a phase

correction based on a semi-empirical relationship. The phase correction relationship is the

same presented by Tijdeman (1977), and it was obtained from experimental observations.

Yonemoto’s procedure was applied for flutter computation of the YXX wing model

(Yonemoto, 1985). This wing has a supercritical profile and it was tested in a wind tunnel at

transonic flow conditions. The computed flutter boundaries agree qualitatively with the

experimental ones, including the capture of the transonic dip for the test case studied.

Yonemoto suggested that further refinement of this procedure was needed for better

quantitative results.

1.1.6 - Modal Aerodynamic Influence Coefficient Matrix (MAIC) Methods

The linear operator which relates pressures to the resulting downwash due to body

motions is the aerodynamic influence coefficient matrix (AIC). However, when these body

motions are written in terms of modal generalized coordinates instead of physical

displacements coordinates the matrix operator, which would satisfy the relation between the

resulting pressures modes to displacement mode shapes, will be the modal aerodynamic

influence coefficient matrix (MAIC). These pressure mode shapes are integrated along the

body area result in a generalized aerodynamic loading. The idea of correction procedures

based on the MAIC concept consists in its replacement of the modal aerodynamic influence
Chapter 1 – Introduction 16

coefficients matrix, computed from linear theory, by a externally computed one which takes

into account a nonlinear aerodynamic behavior.

The aerodynamic derivative factoring procedure from Suciu et al., (1990) is classified

as a MAIC correcting procedure. This method is a stripwise force matching procedure such as

presented in the methods of Pitt and Goodman’s (1987) and Zwaan’s (1985). However,

differently from these two methods, Suciu et al. replace the modal aerodynamic influence

coefficients matrix by an externally computed one.

The main objective of the work of Suciu et al. (1990) is the correction of control

surface and tab hinge moments of a wing based on steady state experimental measurements of

these quantities. Each of these loading quantities are related to a given mode shape pattern

associated to the lifting surface displacement. Suciu et al. assumed that, any lifting surface

displacement mode shape can be decomposed in a linear combination of the typical section

degrees of freedom (Theodorsen, 1934). Therefore, the vector form of the downwash is

transformed into a matrix form, where each of the columns is associated to one of the chosen

typical section degrees of freedom. After the multiplication of this new downwash matrix by

the AIC matrix it is possible to restore a pressure matrix due to the displacements of those

degrees of freedom. The result is that each of the lifting surface elements will include

contributions due the typical section degrees of freedom. By integrating all the pressures

related to elements which belong to a given strip, it is possible to quantify the strip loading,

that is, the lift, moment and controls (aileron and tab) hinge moments.

Correction factors are reduced from the relation between the decomposed linear

loading and the corresponding nonlinear reference data. This is performed from the ratio

between each of the stripwise loading derivative, due to a given typical section degree of

freedom, and the measured experimental loading derivatives. The following step is to use

these correction factors to multiply the decomposed linear loading, which are generated for
Chapter 1 – Introduction 17

each of the reduced frequencies set of the aeroelastic stability analysis to be performed. The

addition of the lift and moments derivatives will result in the lift and moments due to the

typical section degree of freedom displacements. The generalized aerodynamic loading

matrix, for each of those reduced frequencies are computed from the multiplication of the

decomposed mode shape matrix times the corrected stripwise loading. One should note that

there is one correction factor set for each of the mode shapes taken as reference.

The results concerning this method are restricted to subsonic aeroelastic stability

analysis, and present improved prediction of flutter speeds with respect to flutter results using

an uncorrected linear model. Suciu (2000) presented some comparisons between the results

from the above procedure and the Giesing et al.(1976) procedure.. Further generalization of

this procedure, to take into account the interference effects between two or more lifting

surfaces are presented in Suciu (2003). It should be noted that no transonic flutter analysis has

been presented using such method.

An interesting approach for the transonic unsteady aerodynamic loading for flutter

analysis in this regime was presented by Liu et al. (1988), named as Transonic Equivalent

Strip (TES) method. The procedure is divided in two consecutive correction steps, a stripwise

correction followed by a spanwise correction. In the first step, a chordwise mean flow

correction is performed based on steady pressures inputs from CFD codes or measured data.

An inverse airfoil design procedure is employed to generate a new profile of an equivalent

airfoil, for each of the spanwise stations of the lifting surface, being the reference conditions

for the design algorithm the given nonlinear steady pressure distribution. The computation of

unsteady pressures in the profile, now subjected to new boundary conditions, is based on a

time domain impulse response finite difference solution of the transonic small disturbance

equation. The unsteady pressures in the frequency domain are computed from the Fourier

transform of the time response of these pressures. The second step consists in a spanwise
Chapter 1 – Introduction 18

correction of these two-dimensional unsteady pressures to take into account the three-

dimensional effects. This correction is a phase adjustment based on the three-dimensional

linear wave propagation analogy, referred to a given downwash mode. This three-dimensional

wave analogy may be understood as the interference effect computed by the DLM.

The computation of the unsteady pressures distribution is performed through the

application of the spanwise phase correction to the two-dimensional computed pressures.

Since the downwash modes and the associated unsteady pressures modes are known, it is

possible to construct a modal aerodynamic influence coefficients matrix (MAIC) (Chen et al..

1996). Therefore, the generalized aerodynamic forces are computed and replaced in the

aeroelastic equations of motion in terms of modal coordinates, for further application in

dynamic stability analysis or unsteady pressures calculation.

The results regarding the application of this method for both unsteady pressures

computation or flutter calculation are in good agreement with the experimental data, in the

cases of isolated lifting surfaces. One disadvantage of this procedure resides in the

performance of the inverse airfoil design algorithm, when handling complex typical section

profiles such as a combination of wing-pylon-nacelles or external stores with the wing profile.

However, this method is not so expensive like three-dimensional time domain computational

aeroelasticity simulation, since it depends on two-dimensional finite difference solutions of

the transonic small disturbances equation.


Chapter 1 – Introduction 19

1.2 Analytical Review of Correction Methods

The correction techniques reviewed in the last section present advantages and

disadvantages regarding the application context. Most of them employ steady state reference

data, which may be loads or pressures. Some of them are based on the computation of

corrected unsteady pressures from semi-empirical relations, or with the aid of computational

fluid dynamics simulations. The objective of this section is to present an analytical review of

the most applied correction methodologies, with special attention to the ones developed for

transonic flow computation. The emphasis will be to identify their failures, in order to look

for guidelines which can help the development or the enhancement of correction procedures

for approximate modeling of the transonic unsteady flow.

1.2.1 Application of Approximate Solutions for Transonic Flow Computation.

The Giesing et al.(1976) procedure depends on a least squares technique. Thus, there

is no guarantee that the modified pressure distribution, resulting from the corrected AIC

matrix, will be the same as the experimental or the ones computed with nonlinear method. In

the presence of shock waves, this pressure distortion can modify severely the aeroelastic

behavior. Furthermore, the nature of the transonic dip depends on the shock wave

characteristics with regard to its positioning and strength. Nevertheless, at least the modified

pressure distribution will guarantee that the sectional lift and moment will be the same. But

this fact is not sufficient to yield a complete restoration of the nonlinear quasi-steady flow

shock behavior as it would not be the same as the reference condition regarding the shock

dynamics contribution. Care must be taken when using such methods to unsteady transonic

flow correction because the shock wave dynamics plays an important role on the transonic
Chapter 1 – Introduction 20

flutter. These considerations are valid for all force matching procedures, which correct the

discrete element solution AIC matrix.

In the case of the application of correction factors directly to the strip loads, the

nonlinear contribution is implicitly included in reference loading condition. This strategy is

done in Yates (1966) strip theory, and in some of the correction methods presented by Zwaan

(1985). Special attention need to be paid with regard to Yates’ (1966) method, since his

procedure is best suited for wing with high aspect ratio, because the three-dimensional effects

are not taken into account in unsteady loads computation.

The method of Brink-Spalink and Bruns (2000) is based on an optimization technique,

where the difference of the theoretical linear AIC matrix, and the one resulting from its

product by the weighting matrix is forced to be a minimum. This fact represents an

improvement when comparing with the method of Giesing et al.(1976). Brink-Spalink and

Bruns’ (2000) method presents better chances to guarantee the matching of the desired

loading condition without severe distortion the pressure distribution. However, is not possible

to guarantee that shock jump conditions present in transonic flow conditions can be restored

since the reference AIC matrix, taken from the DLM method does not have the capability to

model such transonic flow nonlinearity.

The best way to preserve the reference nonlinear flow conditions regarding the shock

structure is the direct matching of pressures by the use of a weighting operator. However, the

pressure weighting methods introduced by Rodden and Revell (1962), Bergh (1964), Bergh

and Zwaan (1966), and Luber and Schmid (1982) present problems with regard to the

evaluation of the pressure ratios necessary for the computation of the correction factors. If in

any case there is a zero or very low pressure value in the denominator, the correction will

present a tendency to blow up to very high correction factors values implying in a bad

conditioning of the resulting corrected AIC matrix.


Chapter 1 – Introduction 21

A multiple reference pressure modes condition procedure is presented by Jadic et al.

(1999,2000). Their method is based on the introduction of a fully populated pre-multiplying

correction matrix as a pressure weighting method. The use of reference pressures associated

multiple mode shapes is a form to avoid the problem of the scaling of the pressures by small

displacements, even though the weighting operator is not applied to the downwash. But the

problem regarding shock wave characteristics of the reference flow remains the same, since

the additional pressures due to the artificial mode shapes are obtained from a linear theory.

This fact is the reason for the method to fail in the evaluation of the transonic flutter, as

presented in the results of Jadic et al.. (2001).

The other way to reformulate the adjustment of the pressures is to correct the

downwash vector, defined as the substantial derivative of the displacements of the control

points. A work which explores a downwash correction methodology is the procedure

developed by Pitt and Goodman (1987). These authors presented in their work a procedure

based on the post-multiplication of a modified AIC matrix, which relate the strip loads to its

degrees of freedom. This procedure is a force matching method because the reference

conditions are strip loads. The choice of the correction of the downwash is based on

considerations highlighted by MacCain’s (1985). This author discussed in his work that the

post-multiplication weighting operators allows modifying both the real and imaginary parts of

the downwash, thus changing the pressures amplitude and phase angles.

The results presented by Pitt and Goodman (1987) are in agreement with experimental

measured flutter speeds. However, there is no indications in the work of Pitt and Goodman

(1987) with regard to the computation of the loading phases, which can confirm MacCains’s

(1985) observations. The discrepancies found with regard to the flutter speed computation

were attributed by the authors to viscous effects. The reference loading is obtained from the

numerical solution of the transonic small disturbance equation. Indeed, viscous effects alter
Chapter 1 – Introduction 22

the strength and location of shock waves over the wing surface, which in its turn, may also

have a significant effect on the flutter computations.

Another approach is the computation of the nonlinear unsteady pressures based on

semi-empirical hypothesis or relations. The Dau-Garner method (Dau, K, 1992) is a procedure

based on a rational formulation, whose nonlinear unsteady pressures are computed with the

aid of semi-empirical relations for unsteady transonic flows derived by Garner (1977). This

method may be applied, in an aeroelastic analysis context, by the inclusion of downwash

weighting matrices referred to these new computed unsteady nonlinear pressures. Each of

these new reference pressure conditions is related to a given reduced frequency. Therefore,

the corresponding frequency dependent AIC matrix is multiplied by a weighting operator,

which is dependent of the same frequency, to match those new computed reference

conditions.

This method presents a failure with regard to the pressures phase computation,

identified by SenGunpta (1996). The correction of the imaginary component is not effective

because for low reduced frequency value it is simply reduced to a phase shift of 180 degrees.

A mathematical explanation of this behavior is presented in Appendix B. This is probably the

main cause of the pressure phases differences detected by SenGupta (1996), in about 50o ,

when comparing with experimental results.

The correction methods based on the modal aerodynamic influence coefficient

(MAIC) matrix replacement are presented by Suciu et al., (1990) and Liu et al., (1988). It is

possible to understand the Dau-Garner method (Dau, 1992) as a MAIC approach when its

computed pressures are used to construct the generalized aerodynamic loads matrices. The

aerodynamic derivative factoring scheme presented by Suciu et al., (1990) is based on the

correction of the loading derivatives of a lifting surface divided in typical sections. The

reference conditions used to generate a corrected spanwise loading are: steady state lift and
Chapter 1 – Introduction 23

moment derivatives associated to each of the typical section degrees of freedom. Suciu et al.,

(1990) did not present results regarding transonic aeroelastic stability investigations, since

their main concern is the control surface flutter investigation in subsonic flow. However their

procedure allows the use of reference steady or unsteady transonic loading derivatives,

resulting in MAIC matrices, which can be defined for each of the reduced frequencies of

interest. It is an interesting approach which is similar of the work presented by Liu et al.

(1988), where the MAIC approach is also adopted (Chen et al. 2000). The difference between

these two approaches is that one uses pressures distributions to construct the MAIC matrix

(Liu et al. 1988), and the other uses the spanwise loading decomposed in the typical section

degrees of freedom (Suciu et al., 1990) .

The transonic equivalent strip method (Liu et al. 1983) is procedure based on spanwise

computations of two-dimensional unsteady pressures distribution along the chord length in the

frequency domain. This computation is performed by solving the two-dimensional form of the

transonic small disturbance equations for each of the strips of a lifting surface. The results

presented by Liu et al. (1983) show a good agreement of the computed unsteady pressures

and flutter speeds when compared to the experimental measurements. The efficiency of this

procedure is due to the fact that the unsteady pressure computation is performed based on the

solution of a nonlinear equation. However, transonic equivalent strip method and a possible

extension of Suciu’s et al. (1990) procedure for transonic flutter computation are dependent

on nonlinear unsteady finite difference solutions or measured unsteady pressures.


Chapter 1 – Introduction 24

1.2.2 – Proposal of a Correction Technique

From the analytical review of the correction methods, presented in section 1.2.1, one

should observe that the pressure correction based on the downwash weighting was not

explored by other authors. The combination of the matching of the pressures through the

downwash weighting presents good robustness and the preservation of the mean steady flow

nonlinear characteristics. These features indicate that this approach may be developed for the

prediction of the approximate unsteady transonic aerodynamic loading. Other advantages with

regard to the application of this approach are mentioned in what follows.

The matching of the pressures based on downwash correction consists in the

modification of the control points displacements vector to satisfy a given reference pressure

distribution. The advantage of this method when comparing with the pressures correction, is

that the weighting factors computation depends on a linear system equation solution instead of

a simple pressure ratio. For this reason, even though there are null displacements or pressures

given as reference conditions, the solution of the linear system of equation will not imply in

weighting factors blow up, as it happens in the pressures matching procedures at those

conditions.

Another advantage to be pointed out is related to the use of pressures instead of loads

as reference conditions. The use of loads to correct an AIC matrix which relates downwash to

pressures, may lead to distortions of the pressure distributions regarding shock wave

positioning and strength. This problem does not happen when one uses a properly defined

AIC matrix, which relates loads to the corresponding displacements in the typical section

degrees of freedom. However, the AIC matrices which results from the conventional

modeling based on discrete element solutions of the linearized potential flow equations relates

pressures to displacements ant control points. Therefore, it is natural to employ the pressure
Chapter 1 – Introduction 25

matching for the correction of the discrete element models. The nonlinear reference pressures

may be computed from steady state CFD solutions or from experimental measurements.

The computation of the weighting function is based on quasi-steady pressure slopes, as

performed by Luber and Schmid (1988). The choice of the pressure rates instead of their

absolute values is made in order to have displacement-independent weighting functions. The

following step is the setting of a prescribed quasi steady downwash mode shape. One

suggestion is the use of a pitch mode, for example. However, it is possible to employ other

pattern of displacements, such as a frozen structural mode shape. The resulting weighting

function will post-multiply the AIC matrix, that is, it will operate over the downwash vector.

The aeroelastic analysis is performed considering such weighting operator as a post-multiplier

of the AIC matrices, for all reduced frequencies to be explored in an aeroelastic analysis.

Generally, the pressure or load matching procedures, based on steady reference

conditions, fail in obtaining unsteady pressures at higher frequencies. The magnitude of these

pressures are well approximated, but the phases computation presents wrong trends. This is so

because the pressure amplitudes do not present a great variation with the reduced frequencies

as it happens with the pressure phases (Ashley, 1980). The reason for this fact is the absence

of unsteady transonic flow information in the reference conditions (Lui et al., 1983; Ashley,

1980).

The conclusion of this analytical review is that the downwash correction procedure

based on steady reference pressures may be a good option for the discrete element

aerodynamic approximate modeling for transonic flow. Another advantage of such procedure

lies in the fact that they are inexpensive in the computational point of view, since it depends

on steady state computed pressures. The other advantage is that the matching of the pressures

does not distort the nonlinear steady flow characteristics, and also is sufficiently robust in

terms of the computation of the correcting weighting functions. However, the procedure based
Chapter 1 – Introduction 26

on weighting functions computed from steady pressures does not present good results with

regard to the approximation of the imaginary counterpart of the nonlinear unsteady pressures.

Therefore, there is the need to improve such procedure, to adequately compute unsteady

transonic pressures.

1.3 Objectives of the Present Work

The objective of the present work is to study correction techniques applied to unsteady

transonic flow computation. Downwash weighting procedures are the chosen methods for the

investigation of correction procedures, based on their robustness and advantageous features

that were identified, such as cost and compatibility with the physics of the problem. They also

present the advantage in providing a simple way to modify the pressures obtained from the

linear theory by the replacement of externally computed or measured pressures distributions.

These procedures will be investigated using either steady or unsteady pressures. An extension

of this formulation will also be developed to circumvent the problem of the wrong pressure

phases computation, without the use of unsteady reference pressures data. The methodology

of the present work is summarized below:

- Understand the small disturbance nonlinear transonic flow characteristic with

regard to its behavior when subjected to small disturbance conditions;

- Evaluate the downwash correction techniques using either steady or unsteady

reference pressures, and also to understand its behavior based on the nature of those reference

conditions;

- Enhance the downwash correction procedure, to obtain the correct transonic

unsteady pressures based on steady reference conditions, independently of the reduced


Chapter 1 – Introduction 27

frequency of interest. This development should be performed accordingly to a rational

formulation based on the governing equations of the unsteady linear potential flow.

1.4 Outline of the Thesis.

The present chapter presented a review of correction techniques and the statement of

the objectives of the current investigation. Chapter 2 presents the theoretical background of

unsteady linear aerodynamic modeling based on discrete element solutions, and the

aeroelastic modeling for stability analysis. A linear/nonlinear behavior investigation based on

a finite difference solution technique of the Navier-Stokes equations for unsteady transonic

flow computation is presented in chapter 3. This investigation is performed for the better

understanding of the transonic unsteady flow in a small disturbance context. Some results

with regard to this investigation are also presented in the same chapter. Chapter 4 presents

the development of the downwash correction procedure applied to the discrete element

solutions of the linearized potential flow equation. The same chapter presents the

development of an extension of the downwash correction method, which will be named as the

kernel expansion procedure. This method is based on a rational formulation, which goal is to

search for a mathematically and physically supported correction procedure which bypass the

failure and disadvantages of the downwash correction methods, with regard to unsteady

transonic pressures computation.

The results of the downwash correction and the kernel expansion methods, as well as

the comments and discussion about them are presented in chapter 5. The overall discussion

and conclusions with regard to the studies and developments are presented in chapter 6. The

same chapter presents the contribution and suggestions for future works. The bibliography is
Chapter 1 – Introduction 28

presented in chapter 7, and the Giesing et al. (1976) and the Dau-Garner methods (Dau,

1992) are revisited in the A and B appendices respectively.


2 – UNSTEADY AERODYNAMIC MODELING AND AEROELASTIC ANALYSIS

2.1 Integral Solution of the Linearized Potential Flow Equation.

The aerodynamic modeling of unsteady linear potential flows is performed based on

methods such as the discrete kernel function approach. This section presents a brief review of

the discrete kernel function theory applied to subsonic unsteady aerodynamic modeling. The

basic governing equation over which the discrete element methods are developed is the small

disturbances linearized potential flow equation (Landahl, 1961).

φ ( x, y , z , t ) = β 2φxx + φ yy + φzz − 2 ( M ∞2 U ∞ )φxt − ( M ∞2 U ∞ )φtt = 0 (2.1)

∂2 ∂2 ∂2 ∂2 ∂2
where ≡ β2 + +
∂x 2 ∂ y 2 ∂z 2
− 2 ( M 2
∞ U ∞ ) ∂x∂t
− ( M 2
∞ U ∞ ) ∂t 2
is a linear operator

applied to the small disturbances potential.

Under the assumption of harmonic motion, the potential φ is related to a frequency

domain potential ϕ by:

φ ( x, y, z, t ) = ϕ ( x, y, z, k ) eikt , (2.2)

and Eq. (2.1) is written in the frequency domain as:

ϕ ( x, y , z , k ) = β 2ϕ xx + ϕ yy + ϕ zz − 2ikM ∞2 ϕ x + k 2 M ∞2 ϕ = 0 . (2.3)

where is a linear operator applied to the small disturbance potential in frequency domain,

(x,y,z) are the cartesian coordinates, M ∞ the free-stream Mach number, β 2 = 1− M ∞2 ,

k = ωb U ∞ is the nondimensional reduced frequency referred to a free-stream flow speed

U ∞ and a reference length b, and ω is the oscillation frequency associated to the harmonic

motion.
Chapter 2 – Unsteady Aerodynamic Modeling and Aeroelastic Analysis 30

The development of discrete element kernel function methods is based on integral

solutions of Eq. (2.1) for a configuration of interest. The integral solution is obtained by the

application of Green’s theorem to Eq. (2.1) (Chen et al., 1993), leading to:

ϕ ( x, y , z , k ) = −∫∫ ϕS dS + ∫∫ ϕD d ( S + W ) , (2.4)
S S +W

in terms of unsteady source and doublet singularity distributions ϕS and ϕD , respectively,

over a surface S and its associated wake surfaces W. The harmonic oscillating potential, for

source singularity distribution with superficial density σ , is given by:

 ikM ∞ ( M ∞ R− X ) 
σ (ξ , η , 0, k )  e β 2 
ϕS =   , (2.5)
4π  R 
 

which is a function of geometrical parameters, reduced frequency and Mach number. The

coordinates (ξ , η , 0) define the position of an equivalent point source of strength σ dS

resulting from the density integration over the element surface S; and ( x, y, z ) are the

corresponding coordinates of a receiving point in the field to be influenced by the singularity.

The distances between these two distinct points are given by R = X 2 + β 2 (Y 2 + z 2 ) , where

X = x − ξ and Y = y − η . The corresponding doublet potential is given by the derivative of

the source potential with respect to the normal direction :

 ikM ∞ ( M ∞ R− X ) 
∆ϕ (ξ , η , 0, k ) ∂  e β 2 
ϕD =   . (2.6)
4π ∂n  R 
 

One should note that there is a common term inside the parenthesis of Eq. (2.5) and Eq. (2.6).

This term is named the kernel of the integral relation, denoted as Kϕ and given by:

ikM ∞
( M ∞ R− X )
β2
e
ϕ0 = = Kϕ (2.7)
R
Chapter 2 – Unsteady Aerodynamic Modeling and Aeroelastic Analysis 31

The kernel is the free space Green’s function of Eq. (2.3). It represents the solution of the Eq.

(2.3) over an unbounded three-dimensional domain with a point unsteady source of unitary

intensity, concentrated in the point (ξ , η , 0) (Chen et al., 1993). The notation for the doublet

singularity as ∆ϕ is used because the doublet strength can be understood as velocity

potential jump across the surface over which the doublets are distributed. Rewriting Eq. (2.4)

in terms of the kernel function one obtains:

1 1 ∂K
ϕ ( x, y , z , ik ) = − ∫∫ σ (ξ , η , 0, ik )Kϕ dS + ∫∫ ∆ϕ (ξ , η , 0, ik ) ϕ d ( S + W ) (2.8)
4π S 4π S +W ∂n

The change of the notation of the variable argument k to ik is introduced to reinforce the

simple harmonic motion nature associated to the unsteady perturbation potential.

The objective of discrete kernel function methods is to solve the problem of unsteady

aerodynamic modeling of general aircraft configurations, by subdividing its surface, as well

as the associated wakes, in discrete elements. It is assumed that each of these elements contain

elementary solutions of the governing equations. The main concern of this brief review is the

unsteady aerodynamic modeling of planar lifting surfaces, which was chosen for simplicity.

This is the reason why the out of plane coordinate of the source point is set to zero in Eq.

(2.8).

In the case of linear unsteady aerodynamics of thin lifting surfaces, such as a wing, the

thickness effects are of second order. Therefore, it is assumed that only a doublet singularity

distributed on the lifting surface plane may be used to represent the wing. The source

singularity is usually employed to model thickness effects whereas doublet singularity is used

to generate lift. This fact reduces Eq. (2.8) to a simpler form given by:

1 ∂K
ϕ ( x, y , z , ik ) = ∫∫ ∆ϕ (ξ , η , 0, ik ) ϕ d ( S + W ) . (2.9)
4π (S +W ) ∂n
Chapter 2 – Unsteady Aerodynamic Modeling and Aeroelastic Analysis 32

The integral relation presented in Eq. (2.9) represents the doublet singularity distribution over

the lifting surface and its wake (S+W). The discretization of wing-like components and their

associated wake surfaces imply that discrete element aerodynamic modeling is required not

only for the lifting surface, but also for the wake surfaces. This fact greatly increases the

computational effort, because the wake needs to be discretized from the lifting surface trailing

edge downstream to a position sufficiently far from the wing, to simulate far field conditions.

Therefore, the acceleration potential, which is directly proportional to a perturbation pressure

field, is introduced in order to avoid the need of wake modeling, since there is no pressure

jump across wake surfaces. Consequently, the integration limits of Eq. (2.9) are restricted to

the lifting surface domain S.

The linearized potential flow equation problem can be formulated either based on

velocity potential or acceleration potential approaches. The acceleration potential ψ is

defined as the linearized unsteady pressure derived from the linearized unsteady Bernoulli

equation (Bisplinghoff et al., 1951; Liu, 1988), and is given by :

∂∆ϕ ∆C p (ik )
∆ψ ( x, y, z , ik ) = + ik ∆ϕ = − . . (2.10)
∂x 2

Furthermore, quantities such as the velocity potential and its derivative with respect to

the flow direction “x”, and the proper acceleration potential satisfy (∆ϕ) = 0 , (∆ϕ ) = 0 x

and ψ = 0 , respectively, since is a linear operator. Eq. (2.10) is an ordinary differential

equation. Therefore, the velocity potential is obtained by the integration of the acceleration

potential as:
x

ϕ ( x, y, z, ik ) = e ikx
∫e
ikx0
ψ ( x0 , y, z , ik ) dx0 , (2.11)
−∞

where xo is an auxiliary variable. The transformations between the unsteady doublet

singularity and the pressures are given by Eq. (2.10) and (2.11). The choice of the lower limit
Chapter 2 – Unsteady Aerodynamic Modeling and Aeroelastic Analysis 33

of the integral is to satisfy the condition that ϕ vanishes as x → −∞ , that is, ahead of the

lifting surface. When the transformation from the velocity to the acceleration potential is

performed, by the substitution of Eq. (2.11) in Eq. (2.9), the velocity potential is given as a

function of the perturbation pressure field by :

1
ϕ ( x, y , z , ik ) = − ∆C p (ξ , η , 0, ik )K ψ ( X , Y , 0, ik ) dS .
8π ∫∫
(2.12)
S

The remaining function inside the integrand of Eq. (2.12) to be rewritten in terms of

the acceleration potential is a new kernel function K ψ of the integral relation given by Eq.

(2.12). This new kernel function is obtained as a function of the acceleration potential from

the relationship between the harmonically oscillating velocity potential and the acceleration

potential as:
X
∂Kϕ
K ψ ( X , Y , 0) = e−ikX ∫e
ikx0
( x0 , Y , 0) dx0 (2.13)
−∞
∂n

and the final expression for the new kernel function is given by:

 X  ikM ∞  
  ( M ∞ R− x0 )
 
∂  −ikX β2


e
K ψ ( X , Y , 0, ik ) = e eikx0   dx0  (2.14)
 
 x0 + β (Y + z )  
∂n  2 2 2 2

 −∞

The objective at the present step is to establish a relationship between the boundary

conditions and the solution for the velocity potential given by Eq. (2.12), which is a function

of the acceleration potential, or in other words, the linearized pressures. The downwash may

be understood either as the normal component of the induced flow due to a doublet potential

field or as the induced velocities due to displacements of the lifting surface, when the flow

tangency condition over the lifting surface is satisfied. This tangential flow condition in a

small disturbances context, suggests representing the downwash as the substantial derivatives

of the body displacements. On the other hand, the downwash may also be represented as the
Chapter 2 – Unsteady Aerodynamic Modeling and Aeroelastic Analysis 34

derivative of the potential with respect to the normal direction, satisfying the following

relation:

∂h
ϕn ( x, y , 0, ik ) = w ( x, y , 0, ik ) = ( x, y, 0) + ikh ( x, y , 0) , (2.15)
∂x

where the right hand side of Eq. (2.15) results from the application of a substantial derivative

applied to a displacement mode shape vector h, in the reduced frequency domain context.

This is a Neumann boundary value problem, since the velocities are related to potential

derivatives with respect to the normal direction n.

Since a planar lifting surface has been assumed, the normal direction n is aligned with

the “z” axis. Thus, the gradient of the velocity potential field ϕ with respect to the “z”

coordinate, which is normal to the flow direction, will be the downwash. The evaluation of

Eq. (2.12) for the doublet potential over the lifting surface, that is, when z → 0 , leads to the

downwash rewritten as:

 ∂  1 
ϕz ( x, y, 0, ik )z =0 = lim  − ∫∫ ∆C p (ξ , η , 0, ik )K ψ ( X , Y , 0, ik ) dS  (2.16)
z → 0  ∂z
  8π S 

Replacing Eq. (2.14) in Eq. (2.16) yields:

    ikM ∞     
    ( M ∞ R− x0 )
   
∂ 1 ∂  X
e β2
   
ϕz ( x, y, 0, ik )z =0 = lim  − ∫∫ ∆C p (ξ , η , 0, ik ) e−ikX ∫ eikx0   dx0  dS  
z → 0  ∂z
  8π S ∂z   x02 + β 2 (Y 2 + z 2 )    
   −∞
    


(2.17)

Since pressure distributions over lifting surfaces are not functions of the normal coordinate

“z”, the final expression for the downwash is given by:

   
  2   ikM ∞
 
X
 ( M ∞ R− x0 )
∂  β2

∫∫ ∫
1 e
ϕz ( x, y, 0, ik )z=0 = − ∆C p (ξ , η , 0, ik )lim  2 e−ikX ikx0 
e  
 dx0  dS

8π  z→0 ∂ z 
  x + β (Y + z )  
2 2 2 2 
 
S 


  −∞  0   

(2.18)
Chapter 2 – Unsteady Aerodynamic Modeling and Aeroelastic Analysis 35

This solution is also known as the Küssner relation between the acceleration potential

(linearized pressure) and the downwash between two distinct points, where ∆C p (ξ , η , 0, ik )

is the pressure coefficient difference associated to doublet strength (Küssner, 1940).

2.2 – Discrete Element Kernel Function Methods

The formulation presented in section 2.1 is the basic development of the discrete

kernel function methods. The evaluation of the kernel function is a difficult task due to the

complexity of the solution of the resulting integral relations. There are several ways to solve

the integral inside the kernel given by Eq (2.14), considering a number of approaches. These

approaches have been developed to simplify the solution of the kernel function and

correspond to the well-known methods available in the literature.

Sulaeman (2001) presented an evaluation of discrete element kernel function methods

for linear unsteady potential flow modeling. There are several methods that have been

developed and implemented, including the ones due to Landahl and Stark (1968), Albano and

Rodden (1969), Jordan (1978), and Giesing et al. (1971), Ueda and Dowell (1982) Ichikawa

et al. (1985), Rodden et al.(1997,1998), and van Zyl (1998,1999). Sulaeman (2001) reviewed

the formulations of the doublet lattice method of Rodden et al. (1997,1998), the doublet point

method of Ueda and Dowell (1982) and the hybrid doublet lattice/doublet point method of

Eversman and Pitt (1991). He performed an evaluation of these formulations and also

presented expedient approaches to solve the kernel function integrals.

The main differences between these discrete kernel function approaches concern how

the assumed singularity is distributed over an element. The integration of the associated

kernel over each of these elements is also performed in different ways for each method. The

doublet lattice method, for example, assumes that elementary doublets are distributed over the
Chapter 2 – Unsteady Aerodynamic Modeling and Aeroelastic Analysis 36

¼ chord line of each of the panels. An additional point known as the control point needs to be

considered, and is the location over which the downwash induced by the pulsating unsteady

pressures will be computed. This location is the receiving point which is over the intersection

point between the ¾ chord line and the panel midspan line. Therefore, all elements are

influenced by the line of doublets over each of the ¼ chords of each of the panels, yet at the

same time, the downwash is computed in each of the control points.

The doublet point method of Ueda and Dowell (1982) assumes a doublet singularity

concentrated over the intersection point of the ¼ chord line of each panel and its midspan line,

and the receiving point at the same position as defined for the doublet lattice method.

Eversman and Pitt (1991) developed a combined approach between the doublet lattice and the

doublet point methods, named as the doublet hybrid method. The objective of this procedure

is to extend the doublet point method to the nonplanar case. These authors concluded that the

best approach is to utilize the nonplanar doublet point method only for sending and receiving

pairs which are distant. If the sending and receiving points are close, they suggest the use of

the nonplanar doublet lattice method. This combined technique is the basis of their doublet

hybrid method.

In subsonic aeroelastic analysis, the most widely-used unsteady aerodynamic method

is the doublet lattice method of Rodden and Johnson (1994), for its ready applicability to

complex nonplanar configurations. In that method, the integral relation given by Eq. (2.18) is

solved after the application of a variable transformation based in Richardson-Landahl’s

transform (Rodden et al, 1970), which greatly simplifies the evaluation of the integral by a

exponential substitution approach, proposed by Laschka (1963), also used by Giesing et al.,

(1971). Further refinement, with regard to kernel function integration along the element area,

was presented by Rodden et al. (1997,1998). Another discrete element kernel function method

approach was presented by Chen et al. (1993). These authors perform the unsteady subsonic
Chapter 2 – Unsteady Aerodynamic Modeling and Aeroelastic Analysis 37

aerodynamics modeling for bodies and wings with external stores, using a method named

ZONA 6.

One should note that both source and doublet distributions are enclosed in the integral

solutions of the linearized potential flow given by Eq. (2.4). However, the evaluation of the

kernel function needs to be performed distinctly for the case of a wing-like lifting surface or a

body element. This is performed assuming either a surface doublet or a source distribution,

over the lifting surface panel and body panel respectively.

In all the aforementioned methods, it is assumed that the doublet strength is constant

over a line, a surface or concentrated in a point. The way one quantifies the pressures for the

downwash computation is not relevant, since all methods result in algebraic systems of

equations. However, a pressure distribution along each element area is conveniently

considered. This feature greatly simplifies obtaining a linear system of equations that

represents the unsteady aerodynamic model.

2.3 – The Aerodynamic Influence Coefficients Approach

The basic concept of discrete element methods is the subdivision of the lifting surface

(or body) domain into discrete elements, named also as panels. The configuration of interest

discretized into panels is known as a panel model. Each panel contains a control point where

the boundary condition is imposed. The integral equation is approximated by the summation

of elementary integrals associated with each panel.

These elementary integrals at each panel, as well as the aerodynamic interference of

one panel onto others leads to a linear algebraic system of equations relating the pressure

coefficient differences to downwash. The algebraic form of the resulting system of equations

does not depend on the discrete element kernel function method. The assembly of the
Chapter 2 – Unsteady Aerodynamic Modeling and Aeroelastic Analysis 38

elementary integral solutions results in a matrix which elements represent the aerodynamic

influence of the panels into the control points. This matrix is named as the aerodynamic

influence coefficients matrix, which relates the structural deformations to the aerodynamic

forces. This linear relationship between the panels and the control points is justified by the

elementary solutions superposition principle (Bismarck-Nasr, 1993).

The approach used here is based on assuming constant pressure distribution along each

of the elements which discretizes the lifting surfaces and bodies (Chen et al., 1993). In other

words, it is assumed that the pressures associated to the singularity strengths are uniformly

distributed over each of the lifting surface elements Chen et al., 1993). Since these pressures

are assumed constant along the element surfaces, Eq. (2.18) is rewritten as:

∆C pj  ∂  
ϕzi ( x, y, 0, ik )z=0 = −

(ξ , η , 0, ik )
∫∫ Sj
lim

 ( K ψ ) dS
 z→0  ∂z  

(2.19)

The integral relationship (2.19) between the downwash at a receiving point “i”, and

the pressures at a sending panel “j’ can be written as a system of equations as:

ϕ zi = wi = Dij ∆C p j , (2.20)

where each of the matrix elements Dij is the result of the integration of the kernel function

over the given “jth” lifting surface element geometry (Chen, 1993):

ξj ηj
1   ∂  
Dij = − ∫
8π ξ ∫ lim  ∂z K ( x − ξ , y − η , 0, M
z →0
ψ i n i n ∞ , k ) d ξn d ηn ,

(2.21)
j−1 η j−1

In the doublet lattice method this integration is performed as a line integral, since the

doublet distribution is over a line along the ¼ chord length. In the ZONA 6 method, the

integration of the kernel function is performed along the panel area. This feature turns the

ZONA 6 method of higher order than the doublet lattice method.

The inverse of the matrix operator D multiplied by the downwash vector yields the

pressure distribution. In other words, the solution of the system of equations gives the doublet
Chapter 2 – Unsteady Aerodynamic Modeling and Aeroelastic Analysis 39

strength at each panel referred to a given downwash which is related to a displacement mode

shape. The resulting inverse matrix is named as the aerodynamic influence coefficients matrix

AIC, which is a function of the reduced frequency, and is related to the pressure distribution

by:

{∆C p (ik )} =  AIC (ik ) {w(ik )} . (2.22)

One should recall that a simple harmonic motion is assumed, hence the dependence on

ik. The coefficients of this matrix may be interpreted as rates of pressure variation due to a

given displacement amplitude input associated to the boundary conditions. Then, the

determination of the pressure coefficient vector in Eq. (2.22) is performed from the known

downwash, which is related to the amplitude of the pitch and plunge motion at each element.

The substantial derivative of a given displacement mode is composed of a derivative of the

normal direction displacement with respect to the main flow direction plus the associated

velocity scaled by the undisturbed flow speed which, in a small disturbance sense, represents

an angle of attack. Therefore, from the boundary conditions for those small perturbations, the

relationship between the normalwash and a solid boundary displacement is rewritten as:

∂h ( x , y , 0)
{w ( x, y, 0, ik )} = + ikh ( x, y , 0) =  F (ik ) {h ( x, y , 0)} (2.23)
∂x

The substantioal dierivative applied to a given modal displacement vector {h}, which

appears in Eq. (2.23) is denoted by thematrix operator  F (ik ) . The resulting aerodynamic

loading vector, {La (ik )} may be expressed as the multiplication of the pressures by an

integration matrix, [S], which is constructed from the panel elements geometrical

characteristics. The resulting final expression for the unsteady loading over the lifting surface

is given by:

{La (ik )} = q∞ [ S ]  AIC (ik )  F (ik ) {h} (2.24)


Chapter 2 – Unsteady Aerodynamic Modeling and Aeroelastic Analysis 40

where the matrix  F (ik ) is given by:

∂ ⋅ 
 F (ik ) (⋅) =  ( ) + ik (⋅) (2.25)
   ∂x 
 

The subsonic discrete kernel function approach will be further employed as the

unsteady aerodynamic theory for computation of unsteady pressures and loads for aeroelastic

response and stability analysis, to be discussed in the next section.

2.4 – Aeroelastic Model

The aeroelastic response of a given airframe is a result of the mutual interaction of

inertial, elastic and aerodynamic forces induced by static or dynamic deformations of the

structure and other disturbances, such as gusts, atmospheric turbulence, control surface

actuation, etc.

2.4.1 – Aeroelastic Equations of Motion

The equations of motion of a discrete aeroelastic system can be represented as an

equilibrium relation between the structural and aerodynamic forces as:

[ M ]{u ( xs , ys , zs , t )} + [ K ]{u ( xs , ys , zs , t )} = {L ( u ( xs , ys , zs , t ) , u ( xs , ys , zs , t ) )} , (2.26)

where {u ( xs , ys ,zs ,t )} is the physical displacement vector, which is associated to the

displacements of the structure in a structural coordinate system ( xs , ys , zs ) . This vector is a

function of the geometric coordinates of the system and it will be denoted hereafter as {u (t )}

for simplicity. The matrices [M] and [K] are the mass and stiffness matrices respectively,
Chapter 2 – Unsteady Aerodynamic Modeling and Aeroelastic Analysis 41

{ }
which represents the structural dynamic properties of the structure, and L ( u ( t ) , u ( t ) ) is the

aerodynamic load vector.

The matrices [M] and [K] are usually obtained from a finite element model (Bismarck-

Nasr, 1993) of the airframe. Such model is based on the discretization of the structural domain

in elements connected by nodes. At these nodes, the loading is applied and the resulting

displacements due to the structural deformation are computed. The aerodynamic loading

present in Eq. (2.26) is applied at the nodes of the structure, and the displacement vector

{u ( x, y,z,t )} is a function of the nodal displacements in the cartesian coordinates ( x, y,z ) .

The aerodynamic load vector is a function of the structural displacements and it can be

{ }
split in two parts, the induced aerodynamic forces La ( u ( t ) , u ( t ) ) and the external forces

{Le ( t )} . The external forces are usually due to gust loads, atmospheric turbulence, control
surface forces or store ejection loads. They are considered in most cases for the aeroelastic

dynamic response investigation. The induced aerodynamic load vector may be obtained from

any unsteady aerodynamic theory. These loads depends on the structural deformations

{u (t )} as an aerodynamic feedback system. Considering the dependence of the induced

aerodynamic loading with the structural displacements, Eq. (2.26) is rewritten as :

[ M ]{u ( t )} + [ K ]{u ( t )} − {La ( u ( t ) , u ( t ) )} = {Le ( t )} (2.27)

The investigation of the dynamic stability of this aeroelastic system is known as the

flutter problem. This problem consists in a investigation of the stability behavior withe respect

to parameters, such as dynamic pressure, density or flight speed. Usually the induced

{ }
aerodynamic load vector La ( u ( t ) , u ( t ) ) is a nonlinear function of the displacements of the

structure, specially in transonic flow situations. Therefore, in such cases, the flutter analysis

should be performed by a time marching procedure solving the homogeneous form of Eq.
Chapter 2 – Unsteady Aerodynamic Modeling and Aeroelastic Analysis 42

(2.27), based on known initial conditions at t = 0. The stability boundary is obtained by

examining the decay or growth of the structural response with respect to any chosen

parameter variation.

Time marching simulations require the solution of the nonlinear aerodynamic

equations of the flow. These solutions are based on computational fluid dynamic procedures,

where the nonlinear governing equations are solved by numerical methods, such as the finite

difference, finite volume or finite elements approaches. Since the solution of the nonlinear

governing equations is performed in the time domain, in most cases this approach to solve

flutter problems is computationally expensive. However, it is the most accurate way in

obtaining the stability behavior of an aeroelastic system, because of the high fidelity

aerodynamic modeling.

The mathematical structural dynamic model, which is the left hand side of Eq. (2.26)

is in most cases linear, since in aeroelastic phenomena the amplitudes of deformations are

small (Bisplinghoff et al., 1996). When the aerodynamic response may be assumed linear

with respect to the structural deformations, the system of equations (2.27) is linear. The flutter

boundary computation should be performed by solving the complex eigenvalue problem of

this system, for a given parameter variation.

The solution of the eigenvalue problem requires that the homogeneous aeroelastic

system of equations :

[ M ]{u ( t )} + [ K ]{u ( t )} − {La ( u ( t ) , u ( t ) )} = 0 , (2.28)

be written into the Laplace domain. The transformation of the induced aerodynamic loads to

the Laplace domain is performed by a convolution integral as :


 U 
La ( u ( t ) , u ( t ) ) = q∞  H  ∞ ( t − τ )   {u (τ )} dτ
  b  (2.29)
0
Chapter 2 – Unsteady Aerodynamic Modeling and Aeroelastic Analysis 43

where [ H ] represents the aerodynamic transfer function matrix, q∞ is the dynamic pressure, b

is a reference length equal to half the reference chord, and U ∞ is the undisturbed flow speed.

The aerodynamic loading in the Laplace domain is written as :

{L ( u ( s ) )} = q
a ∞
H
 ( ) {u(s)}
sb
U∞
(2.30)

with  H
 ( ) as the Laplace domain counterpart of  H ( ) , and sb U
sb
U∞
U ∞t
b ∞ being the non-

dimensional Laplace variable. The resulting eigenvalue problem is obtained by transforming

Eq. (2.28) to the Laplace domain as:

 2 
 s [ M ] + [ K ] − q∞  H ( )  {u(s)} = 0
sb
U∞
(2.31)

2.4.2 – Modal Approach

In some cases, there is a large number of degrees of freedom associated with the

discretization of the structure, which implies in large mass and stiffness matrices. The most

adequate approach to be adopted is the transformation of the structural dynamic system model

to a modal space. This transformation is performed by the eigensolution of the homogeneous

form of Eq. (2.26) in the Laplace domain. The resulting eigenvalues are known as the natural

frequencies, and the eigenvectors are the associated mode shape vectors. Therefore, in a

modal representation, the structural displacements of the structure are written as a linear

combination of the generalized coordinates {q ( t )} and the physical displacements of the

structure {u ( t )} as:

{u ( t )} = [Φ ]{q ( t )} , (2.32)

where [ Φ ] is the modal transformation matrix, composed by mode shape vectors, each of
Chapter 2 – Unsteady Aerodynamic Modeling and Aeroelastic Analysis 44

them associated to a given natural frequency of the structure.

The critical flutter modes usually result from the coupling of low order structural

modes, leading to a mode coalescense. The low order modes are also associated to the lower

natural frequencies of the structure. Therefore, the structural deformation of the airframe in

the flutter mode can be represented by the superposition of the lower order structural modes.

For this reason it is possible to reduce the size of the matrix system of equations by employing

the modal approach. This is performed by the choice of the lower order modes to represent the

structural dynamic behavior of the structure. The resulting aeroelastic system of equations

written in the modal space is given by:

 2    
 s  M  +  K  − q∞ Q ( )  {q(s)} = 0
sb
U∞
(2.33)

where :

 M  = [ Φ ] [ M ][ Φ ]  K  = [ Φ ] [ K ][ Φ ] ,
T T
and (2.34)

are the generalized mass and stiffness matrices respectively, and

Q
 ( ) = [Φ]  H ( ) [Φ] ,
sb
U∞
T sb
U∞
(2.35)

is the generalized aerodynamic forces matrix, with {q ( s )} being the generalized coordinates

vector in the Laplace domain.

2.4.3 Simple Harmonic Motion Approach

The computation of the aerodynamic transfer functions in the Laplace domain is a

complicated task. For this reason, unsteady aerodynamic methods are often formulated in the

frequency domain assuming simple harmonic motion for the body and a simple harmonic

aerodynamic model in the frequency domain. The aerodynamic influence coefficients


Chapter 2 – Unsteady Aerodynamic Modeling and Aeroelastic Analysis 45

approach may be understood as an aerodynamic transfer function, however, it is computed in

a simple harmonic reduced frequency domain rather than in the Laplace domain.

Previously in this chapter, the discrete kernel function approach was presented, where

the aerodynamic transfer function was obtained in the form of the aerodynamic influence

coefficients matrix, based on the simple harmonic hypothesis. Its role is to relate the structural

deformations to the discrete element aerodynamic pressures. The induced aerodynamic loads

in the frequency domain are obtained after the multiplication of the pressure coefficients by

the dynamic pressures and by an integration matrix, resulting in Eq. (2.24).

The induced aerodynamic loading is obtained by a convolution integral which

transforms the aerodynamic loading from the time domain to the Laplace domain. Since the

AIC matrix is only available in the simple harmonic reduced frequency domain, the imaginary

counterpart of aeroelastic system of equations, given by Eq. (2.33), is obtained replacing s by

iω. leading to:

 −ω 2  M  +  K  − q∞ Q ( ik )   {q (iω )} = 0 (2.36)


 

with {u ( iω )} = [ Φ ]{q ( iω )} .

The assumption of simple harmonic motion is valid for a stability analysis of the

aeroelastic system, since damping at this condition is zero. Moreover, this hypothesis is

mathematically consistent with the assumption of the simple harmonic motion introduced by

the unsteady aerodynamic formulation.

2.4.4 – Interconnection Between Aerodynamics and Structures

It should be pointed out that the aerodynamic discretization usually does not match the

structural dynamic model discretization. The aerodynamic degrees of freedom are the

displacements at the panel control points. On the other hand, the structural degrees of freedom
Chapter 2 – Unsteady Aerodynamic Modeling and Aeroelastic Analysis 46

are the modal displacements associated with the finite element model. Furthermore, each of

the models can be defined based on a particular coordinate system, which need not to be

coincident. Therefore, the exchange of displacement information between the structural and

the aerodynamic degrees of freedom requires an interpolation procedure, which may be

regarded as an operator [G] applied to the following transformations:

{h ( x, y,0)} = [G ]{u ( xs , ys ,zs )} and {φis } = [G ]{φia } (2.37)

where {u ( xs , ys , zs )} is the finite element model nodal displacements vector and {h ( x, y, 0)} ,

is the displacement vector at the aerodynamic control points. [G] is a general matrix

{ }
transformation operator, φis is the mode shape vector of the modal matrix [ Φ s ] , referred to

the structural degrees of freedom, as well as the vector φia { } and the modal matrix [Φ ] are a

referred to the aerodynamic degrees of freedom.

The aerodynamic loading must be also transferred from the aerodynamic to the

structural grid points. For this case, the loading transfer is performed based on the virtual

work principle, which implies that the distributed loading applied to the aerodynamic and

structural reference points must be equivalent. In other words, this equivalence relationship is

given by:

{δ h ( x, y,0)}
T
{Laero
a ( x, y, 0)} = {δu ( xs , ys ,z s )} { La ( xs , ys ,z s )}
T str (2.38)

T T
where, {δ h ( x, y, 0)} and {δu ( xs , ys ,zs )} are the virtual displacements, and the aerodynamic

loading at the structural grid points and at the aerodynamic control points are denoted as

{Lstr
a ( xs , ys ,zs )} and, {Laero
a ( x, y,0)} , respectively. The resulting relationship between the loads

in these two reference systems is given by:

{L str
a ( xs , ys , zs )} = [G ]T {Laero
a ( x, y,0)} (2.39)
Chapter 2 – Unsteady Aerodynamic Modeling and Aeroelastic Analysis 47

Replacing Eq. (2.25) in Eq. (2.39) and using Eq. (2.37) leads to:

{L str
a ( xs , ys ,zs ,ik )} = q∞ [G ]T [ S ]  AIC (ik )  F (ik ) [G ]{u ( xs , ys ,zs )} . (2.40)

The generalized aerodynamic forces matrix results from the transformation:

Q (ik ) = [Φ a ]T [G ]T [ S ]  AIC (ik )  F (ik ) [G ][Φ a ] (2.41)


    

The most adequate interpolation operator for the connection between the structural

grid points and aerodynamic control points is based on spline function approximations. The

resulting operator G is named here as the spline matrix, which coefficients are obtained from

the interpolation functions. A description of the spline matrix interpolation methods is

presented in Zona Technology (2003), and Rodden and Johnson (1994). Further information

about the procedure for obtaining the coefficients of the spline matrix operator is presented by

Harder and Desmarais (1972), and the mathematical procedure is presented by Silva (1994),

Zona Technology (2003), and Rodden and Johnson (1994).

2.5 – Flutter Solution Techniques

The stability of an aeroelastic system is evaluated by using flutter solution techniques.

These methods are based on the solution of the eigenvalue problem with respect to a given

parameter variation. Some examples of flutter solutions techniques are the well known “k”

method due to Theodorsen (1935), also presented by Rodden and Johnson, (1994), and

Rodden and Berlinger, (1982); the “pk” method - Irwin and Guyett (1965) and Hassig (1971);

and the “g” method (Chen, 1999). The latter will be applied for the flutter computations in the

present work.

A brief review of such flutter solution techniques will be presented. Further

information about the solution algorithm, with regard to the k and pk-methods are presented
Chapter 2 – Unsteady Aerodynamic Modeling and Aeroelastic Analysis 48

by Rodden and Johnson (1994). The detailed procedure of the g method is presented in the

works of Chen (1999) and Zona Technology (2003).

2.5.1 – Flutter Computation Based on the k-method

The flutter solution of an aeroelastic system is defined as the stability analysis of the

linear system of equations (2.36), which results from the assumption of a purely harmonic

motion, assuming s = iω . In order to evaluate the stability behavior with respect to a given

parameter variation, it is introduced an artificial damping, since the eigensolution of Eq.

(2.36), in its original form, results in imaginary eigenvalues. Therefore, Eq. (2.36), is

rewritten as:

 −ω 2  M  + (1 + ig s )  K  − q∞ Q ( ik )   {q (iω )} = 0 , (2.42)


 

where the new term ig s is an added artificial structural damping, originally introduced by

Theodorsen (1935).

This approach is known as the k-method flutter solution. It is mathematically

consistent with the assumption of simple harmonic motion introduced in the unsteady

aerodynamic formulation. The damping is only an artifice used to seek out the flutter stability

margin, and cannot be interpreted as having physical significance as a measure of the decay

rate of the aeroelastic system response. However, the method is capable of providing the

correct prediction of the flutter boundary, which, in other words, is the flutter speed at zero

damping.
Chapter 2 – Unsteady Aerodynamic Modeling and Aeroelastic Analysis 49

2.5.2 – The pk-method

The damping values at flow conditions below the flutter speed can serve as guidelines

for conducting wind tunnel tests or flight flutter tests. For this reason, it is necessary to

consider a different approach for the flutter solution with regard to the computation of the true

damping of the aeroelastic system. Recalling Eq. (2.33) one should note that it can be

rewritten as a function of a non-dimensional Laplace parameter as:

 b  2 
  p 2
 M  + 
    ∞K  − q  Q ( p )   {q ( p )} = 0
 (2.43)
 U ∞  

where p = γ k + ik = sb U ∞ , with γ as the transient decay rate coefficient. This is the so-

called p-method equation. Its solution can provide the true damping of the aeroelastic system.

The remaining problem is the availability of the aerodynamic transfer function, which

is readily available for simple harmonic motion only. One way to circumvent this problem is a

method originally proposed by Irwin and Guyett (1965), known as the p-k flutter solution

technique. The basic assumption of this approach is the replacement of the simple harmonic

motion aerodynamic transfer function in the p-method equation (2.33). The resulting

aeroelastic system to be solved is given by:

 b  2 
  p  M  +  K  − q∞ Q ( ik )   {q ( p )} = 0
2
. (2.44)
 U ∞  

The p-k method equation is mathematically inconsistent since the eigenvalue p is

complex, (corresponding to damped oscillatory motion), while Q ( ik )  is obtained based on

undamped harmonic motion (hence the purely harmonic imaginary argument). Nevertheless,

it is believed that this method is a good approximation in finding the damped solution The

rationale for the p-k method is that for oscillatory motion with slowly increasing or decreasing

amplitude the generalized aerodynamic force based on constant amplitude harmonic motion is
Chapter 2 – Unsteady Aerodynamic Modeling and Aeroelastic Analysis 50

a good approximation. A detailed description, with regard to the superiority of the p-k method

in comparison with the k-method, is presented by Hassig (1972).

Rodden et al. (1979) presented a modification of the original p-k method. Their

formulation shows the effectiveness of the p-k method in comparison with the k method. This

modified method is based on the introduction of an aerodynamic damping matrix into Eq.

(2.44) resulting in:

 U ∞  2 2 
  p  M  +  K  − q∞ QI ( ik ) k  p − q∞ QR ( ik )   {q ( p )} = 0 (2.45)
 b  

where the complex aerodynamic influence coefficients matrix Q ( ik )  is written as:

Q ( ik )  = QR ( ik )  + i QI ( ik ) (2.46)

with QR ( ik )  and QI ( ik )  as the real and imaginary parts of Q ( ik )  , respectively.

Substituting the non-dimensional Laplace variable, which multiplies the imaginary part of

Q ( ik )  in (2.45), yields:

 U ∞  2 2 
  p  M  +  K  − q∞ QI ( ik ) k  γ k − q∞ Q ( ik )   {q ( p )} = 0 (2.47)
 b  

One should note that Eq. (2.47) presents an extra term, when comparing with Eq. (2.44). This

term is the added damping matrix.

The eigensolution of the system of equations (2.47) is performed by varying the

parameter q∞ , or the associate flow speed U ∞ or the density ρ , for each of the modes of

interest. The resulting eigenvalues are associated to each of the chosen modes, and their

imaginary parts are compared with the reduced frequency by an iterative procedure. This

procedure requires the repeated interpolation of the matrices Q ( ik )  from a given reduced

frequency set, during the process of eigenvalue extraction. As soon as the imaginary part of p
Chapter 2 – Unsteady Aerodynamic Modeling and Aeroelastic Analysis 51

matches the associated reduced frequency value used to compute Q ( ik )  , the associated

frequency and damping of the eigenvalue are computed and stored as a flutter solution point.

2.5.3 – The g-method

Another recent approach for the flutter solution of the aeroelastic system is presented

by Chen (2000), named as the g-method. It is a damping perturbation method, where it is

suggested that a first order term, in terms of the derivative of the aerodynamic influence

coefficients matrix with respect to the damping g, could be included in the flutter equation in

the Laplace domain. This method is rigorously derived from the Laplace domain

aerodynamics, where it is assumed the aerodynamic transfer function matrix Q ( p )  can be

defined for the whole Laplace domain. This assumption is based on the analytic continuation

of the aerodynamic transfer function matrix for decayed motion, since it is analytic for

divergent or constant amplitude motion. Therefore, assuming that Q ( p )  is analytic, it can

be expanded along the imaginary axis, around small values of the damping g as:

 ∂ Q ( p )  
Q ( p )  Q ( ik )  + g    (2.48)
 ∂ g 
  g =0

The derivative or the aerodynamic transfer function Q ( p )  with respect to the

damping is not available from the unsteady aerodynamic theory. Since the p-domain

aerodynamic transfer function is analytic, it must satisfy the Cauchy-Riemann relations such

that the derivative of Q ( p )  with respect to the damping can be written as the derivative of

Q ( p )  with respect to the complex reduced frequency ik, leading to:

Q ( p )  Q ( ik )  + g Q ' ( ik )  (2.49)


Chapter 2 – Unsteady Aerodynamic Modeling and Aeroelastic Analysis 52

where g = γ k . Substituting Eq. (2.49 in (2.43) yields the g method equation :

 U ∞  2 2 
 p  M  +  K  − q∞ Q ( ik )  γ k − q∞ Q ( ik )   {q ( p )} = 0
'
 (2.50)
 b  

One should note that the difference between the pk method flutter equation (2.47) and

the g method equation lies in the damping matrices, Q ' ( ik )  and QI ( ik ) k  , respectively.

At zero reduced frequency, both methods will provide the same flutter boundary. In fact the

damping matrix in the pk-method is a special case of the damping matrix in the g method.

This fact can be shown by expanding Q ( ik )  by a Taylor’s expansion as:

1
Q ( ik )  = Q ( 0 )  + ik Q ' ( 0 )  + ( ik ) Q '' ( 0 )  + …
2
(2.51)
2

Since the derivative terms at ik = 0 are real, the aerodynamic transfer function for

simple harmonic motion can be split into a real and imaginary part as follows:

 1   1 
Q ( ik )  =  Q ( 0 )  − k 2 Q '' ( 0 )  + …  + i  k Q ' ( 0 )  − k 3 Q ''' ( 0 )  + …  ,(2.52)
 2   6 

dividing the imaginary part of Eq. (2.52) by k leads to the damping matrix introduced in the

pk method:

1
QI ( ik ) k  = Q ' ( 0 )  − k 2 Q ''' ( 0 )  + … (2.53)
6

Differentiating the Taylor expansion series with respect to ik gives the damping matrix

introduced by the g method as:

Q ' ( ik )  = Q ' ( 0 )  + ik Q '' ( 0 ) + … (2.54)

A comparison between Eqs. (2.53) and (2.54) indicates that QI ( ik ) k  , and

Q ' ( ik )  are equal if Q ( ik )  is a linear function of k, at k = 0 . However, the aerodynamic

transfer function matrix for simple harmonic motions is a highly nonlinear function of the

reduced frequency. The frequency dependent terms resulting form the expansion of the pk-
Chapter 2 – Unsteady Aerodynamic Modeling and Aeroelastic Analysis 53

method aerodynamic damping matrix may introduce errors in its approximation. The

consequence of this fact is the production of unrealistic roots, during the extraction of the

aeroelastic system eigenvalues. The aforementioned comments indicate the advantage of the

g-method when comparing it with the pk-method.

2.5.4 – p-method for the Solution of a State Space Aeroelastic System.

The flutter solutions techniques presented in subsections 2.5.1 to 2.5.3 are based on an

unsteady aerodynamic model which is valid for simple harmonic motion of the body. This

approach is valid for flutter stability boundary computation. However, at conditions other than

the flutter boundary, the body is undergoing aeroelastic motion which is not purely harmonic.

Therefore, rigorously speaking, the aerodynamic damping computed with those methods does

not have a physical significance. The p-k and g methods are good solutions for computing an

estimate of the damping for arbitrary motion, but still are approximate techniques.

In order to seek out a solution which can provide the true damping of the aeroeastic

system, the p-method is based on unsteady aerodynamics approximated by rational functions

in terms of the Laplace variable s, that is, for arbitrary motions. Looking at Eq. (2.33), one

should observe that the aerodynamic influence coefficients matrix is a function of the

nondimensional Laplace variable p = sb U ∞ . The p-method consists in approximating a set

of aerodynamic influence coefficient matrices, computed for simple harmonic motion, by a

representative function of the Laplace variable s. This function is known as a Padé polynomial

(Vepa 1977), and has zero to second order terms in the Laplace variable s, as well as a series

of poles, that is:

s [Qn + 2 ]
2
 b  2 b 
nlag
Qap ( s )  = [Q0 ] + [Q1 ] s   [ 2] 
+ Q s  ∑
+ (2.55)
 U∞  ( )
 U ∞  n =1 s + b∞ β n
U
Chapter 2 – Unsteady Aerodynamic Modeling and Aeroelastic Analysis 54

It should be recalled that sb U ∞ = ik on the imaginary axis. The βn are called lag

parameters since they are physically related to the inherent lag between motion and

aerodynamic loading in unsteady phenomena, and nlag is the number if these terms. The

values of these parameters should be obtained either by optimization procedures or by user's

experience. The coefficients Q j  , j = 1, , 2 + nlag of the Padé polynomial are obtained using

the least-square method in order to fit a set of tabulated matrices Q ( km )  . This

approximation is performed for each of the coefficients of the aerodynamic influence

coefficient matrices, independently of the other matrix elements. A more detailed description

of such procedure may be found in the work of Silva (1994).

One should note that the rational function approximation was performed based on

simple harmonic motion unsteady aerodynamics. However, the resulting approximate

aerodynamic influence coefficient matrix is valid for unsteady arbitrary motion. This fact is

justified by the analytical extension principle associated to the rational function, which now is

defined for the entire complex plane. Therefore, the resulting approximation leads to an

unsteady aerodynamic model for arbitrary motions. Once the rational approximation has been

obtained (Eq. 2.55), one can rewrite Eq. (2.33) as :

 s 2  M  +  K   {q ( s )} = q∞ Qap ( s )  {q ( s )} (2.56)
 

The most adequate approach to compute the aeroelastic stability of the system, where

the aerodynamic influence coefficient matrix is approximated by Padé polynomials, is by

transforming the system of equations from the frequency domain to the time domain. Thus, an

inverse Laplace transform should be performed leading to Eq. (2.56) written in the time

domain as :
Chapter 2 – Unsteady Aerodynamic Modeling and Aeroelastic Analysis 55

 b 
 M  {q (t )} +  K  {q (t )} = q∞ [Q0 ]{q (t )} + q∞ [Q1 ]{q (t )}  +
 U∞ 
 b 
2
nlag (2.57)
+ q∞ [Q2 ]{q (t )}   + q∞ ∑ [Qn + 2 ]{qlag (t )}n
 U∞  n =1

The new generalized coordinates qlag (t ) are the so-called aerodynamically augmented

or lag states, and they appear because of the rational form of the poles series. They may be

obtained by applying the convolution theorem to the product of functions:

 s [Q ] 
 n+2
 {q ( s )} , (2.58)
 ( )
 s + b βn 
U∞

where the inverse Laplace transforms are known:

 [Q ] 
L −1  s {q ( s )} = {q ( t )}  = [Qn + 2 ] e −(U ∞ b ) β nt
and L −1  n+2
. (2.59)
 ( )
 s + b βn 
U∞

Therefore,
t

{q ( t )} = ∫ {q (τ )}e (
lag n
− U ∞ b ) β n ( t −τ )
dτ . (2.60)
0

The differentiation of Eq. (2.60) gives a set of ordinary differential equations for the lag

states, which can be included in the set of equations of the aeroelastic system:

{q ( t )} = {q ( t )} − Ub
lag n
β n {qlag ( t )}n , (2.61)

and a vector of state space variables of the aereolastic system may be written as :

T
{λ ( t )} = {q ( t )} {q ( t )} {qlag ( t )}1 …{qlag ( t )}n  ,
T T T T
(2.62)
 lag 

The resulting aeroelastic system of equations is represented by an unique matrix with

only the freestream flow speed as argument:

{λ ( t )} = A (U ) {λ ( t )} ,
∞ (2.63)

where the state space matrix  A (U ∞ )  is the state matrix of the state space aeroelastic
Chapter 2 – Unsteady Aerodynamic Modeling and Aeroelastic Analysis 56

representation of the system:

 − M −1 B − M −1 K − M −1Q3 − M −1Qnlag + 2 
 
 I 0 0 0 

A (U ∞ ) =  I 0 − ( )β I
U∞
b 3 0 
 , (2.64)
 
 
 I

0 0 ( )
− b β nlag + 2 I 
U∞

with,

2
 b   K  =  K  − q∞ [Q0 ]
 M  =  M  − q∞   [Q2 ] ,
 U∞  . (2.65)
 b 
 B  =  B  − q∞   [Q1 ] , Qn  = q∞ Qn 
 lag   lag 
U
 ∞

The stability investigation of such system of equations should be performed through

its eigensolutions. The resulting eigenvalues lead to aeroelastic frequency and aerodynamic

damping, which in this method has physical significance. Silva (1994) pointed out that the

differences between the computed aerodynamic damping based on the p-k method and p-

method are small. However, the expediency of state space aeroelastic modeling lies in the fact

that it is the most adequate formulation for modeling aeroservoelasticity problems.

Another fact to be pointed out is that the g-method formulation is based on an

expansion of the aerodynamic influence coefficient matrix as a first order series, with respect

to its derivative in relation to the damping g. Furthermore, the resulting approximation for the

aerodynamic influence coefficient matrix for arbitrary motion is valid based on the same

analytic continuation principle. This allows for a comparison between the p and g methods,

where p-method employs a rational function to compute the unsteady aerodynamics for

arbitrary motion, while for the g-method the unsteady aerodynamics is computed from the

approximation of the aerodynamic influence coefficients matrix by a first order series

expansion in terms of the derivative with respect to g.


3 –UNSTEADY TRANSONIC FLOW BEHAVIOR

3.1 - Linear/Nonlinear Behavior Investigation

Approximate flutter analyses in transonic flow are typically based on linear solutions

of an aeroelastic system of equations, with the inclusion of the nonlinear transonic effects, by

applying corrections to the linear aerodynamic model. However, such analyses implicitly

require that transonic aerodynamic loads, taken as reference conditions, be “locally” linear

with respect to structural deformation of a give lifting surface.

In the work of Dowell et al. (1983), some studies in the linear behavior of the

nonlinear unsteady transonic flow were performed in order to identify the range of

parameters, such as phase angle and shock strength, related to reduced frequency and angle of

attack, over which the linear behavior occurs. The objective in this case is to understand the

limits of linear behavior for aeroelastic applications, where most analyses apply linear

equations and the motions are small. Some conclusions arose from the results presented by

Dowell et al. (1983), such as, the linearity of shock wave motion with respect to a small angle

of attack change, which is the case of aeroelastic applications, and the importance of an

accurate mean flow calculation to take into account the correct shock wave position and

strength.

This chapter presents an extension of the linear/nonlinear behavior investigation

presented by Dowell et al. (1983). The nonlinear aerodynamic model over which their

investigation was performed is based on finite difference solutions of the transonic small

disturbances equations. The analysis test cases were restricted to computational simulations of

two-dimensional profiles oscillating in harmonic motion. In the present investigation,

unsteady transonic flow computations are performed based on solutions of the Navier-Stokes
Chapter 3 – Unsteady Transonic Flow Behavior 58

equations, using the finite difference procedure to be presented in the next section. The results

with regard to time domain simulations of the Navier-Stokes equations are used to verify the

“locally” linear behavior of aerodynamic loads with respect to the dynamic angle of attack.

With the present numerical technique, viscous and three-dimensional effects are taken into

account and shown to be significant.

3.2 – The Navier Stokes Nonlinear Model

The nonlinear aerodynamic computations were based in the finite difference solutions

of the Navies Stokes equations in order to have steady and unsteady pressures distribution

over the lifting surfaces. The Navier-Stokes solver used for unsteady flow computations

(Mello, 1994) is a modified version of a code developed by Sankar and Kwon (1990). It is a

finite difference solution of the vector form of the full Reynolds-averaged, three-dimensional

Navier-Stokes equations based on an arbitrary curvilinear coordinate system The time

marching consists in a two-point backward difference at the new time level n + 1 . All spatial

derivatives are approximated by standard second-order central differences (Mello, 1994). The

mathematical formulation and the description of the numerical method are here reproduced

from the work of Mello (1994).

3.2.1 – Governing Equations

The vector form of the full Reynolds-averaged, three-dimensional Navier-Stokes

equations based on an arbitrary curvilinear coordinate system can be written in non-

dimensional form as:

Qˆτ + Eˆξ + Fˆη + Gˆ ς =


Re
(
1 ˆ
Rξ + Sˆη + Tˆς ) (3.1)
Chapter 3 – Unsteady Transonic Flow Behavior 59

where Q̂ is the vector of unknown flow properties; Ê , F̂ , Ĝ are the inviscid flux vectors;

R̂ , Ŝ , Tˆ are the viscous flux vectors and Re = ρ∞ a∞ c µ∞ is the Reynolds number based on the

free-stream speed of sound a∞ , density ρ∞ , viscosity µ ∞ and reference chord c . The

resulting non-dimensional flux vectors are given by,

 ρ   ρU c   ρVc 
 ρu   ρ uU + ξ p   ρ uV + η p 
1   1   1  
c x c x

Qˆ =  ρ v  ; ˆ
E =  ρ vU c + ξ y p  ; ˆ
F =  ρ vVc + η y p 
J  J J
ρw ρ wU c + ξ z p  ρ wVc + η z p 
     
 e  ( e + p )U c − ξt p  ( e + p ) Vc − ηt p 

 ρWc   0 
 ρ uW + ζ p  ξ τ + ξ τ + ξ τ 
1 
z xz 
1  
c x x xx y xy
ˆ ˆ 
G =  ρ vWc + ζ y p  ; R = ξ xτ xy + ξ yτ yy + ξ zτ xz 
J J
ρ wWc + ζ z p  ξ τ + ξ yτ yz + ξ zτ zz 
   x xz 
( e + p ) Wc − ζ t p   ξ x R5 + ξ y S5 + ξ zT5 

 0   0 
η τ + η τ + η τ  ζ τ + ζ τ + ζ τ 
1 
z xz 
1 
x xx y xy x xx y xy z xz 
ˆ  ˆ 
S = η xτ xy + η yτ yy + η zτ xz  ; T = ζ xτ xy + ζ yτ yy + ζ zτ xz  (3.2)
J J
η τ + η yτ yz + η zτ zz  ζ τ + ζ yτ yz + ζ zτ zz 
 x xz   x xz 
 η x R5 + η y S5 + η zT5   ζ x R5 + ζ y S5 + ζ zT5 

with J as a jacobian transformation between the curvilinear and cartesian systems written as:

J =  yξ ( xζ zη − xη zζ ) + yη ( xξ zζ − xζ zξ ) + yζ ( xη zξ − xξ zη )  .
−1
(3.3)

Uc, Vc and Wc are the contravariant velocities defined as :

U c = ξt + ξ x u + ξ y v + ξ z w
Vc = ηt + η x u + η y v + η z w (3.4)
Wc = ζ t + ζ x u + ζ y v + ζ z w

and the pressure p is related to the total energy and kinetic energy per unit volume by:

 1 
 2
(
p = (γ + 1) e − ρ u 2 + v 2 + w2  .

) (3.5)
Chapter 3 – Unsteady Transonic Flow Behavior 60

The shear stresses are :

4 2 
τ xx = µ 
3
( uξ ξ x + uηη x + uζ ζ x ) − ( vξ ξ y + vηη y + vζ ζ y + wξ ξ z + wηη z + wζ ζ z ) 
3 
τ xy = µ ( uξ ξ y + uηη y + uζ ζ y ) + ( vξ ξ x + vηη x + vζ ζ x ) 
τ xz = µ ( uξ ξ z + uηη z + uζ ζ z ) + ( wξ ξ x + wηη x + wζ ζ x ) 
4 2 
τ yy = µ 
3
( vξ ξ y + vηη y + vζ ζ y ) − ( uξ ξ x + uηη x + uζ ζ x + wξ ξ z + wηη z + wζ ζ z ) 
3 
τ yz = µ ( vξ ξ z + vηη z + vζ ζ z ) + ( wξ ξ y + wηη y + wζ ζ y ) 
4 2  (3.6)
τ zz = µ  ( wξ ξ z + wηη z + wζ ζ z ) − ( uξ ξ x + uηη x + uζ ζ x + vξ ξ y + vηη y + vζ ζ y ) 
3 3 

with the R5, S5, and T5 variables defined as:

µ  ∂a 2 ∂a 2 ∂a 2 
R5 = uτ xx + vτ xy + wτ xz + ξ
 x + η + ζ 
Pr (γ + 1)  ∂ξ ∂η ∂ζ 
x x

µ  ∂a 2 ∂a 2 ∂a 2 
S5 = uτ xy + vτ yy + wτ yz + ξ y +ηy +ζ y 
Pr (γ + 1)  ∂ξ ∂η ∂ζ  (3.7)
µ  ∂a ∂a
2
∂a  2 2
T5 = uτ xz + vτ yz + wτ zz + ξz + ηz +ζ z 
Pr (γ + 1)  ∂ξ ∂η ∂ζ 

where Pr = µ c p k is the Prandtl number and a is the speed of the sound.

In turbulent flows, the molecular viscosity µ appearing in Eqs. (3.6) and (3.7) is

replaced by µ + µT , and the quantity µ Pr in Eq. (3.7) is replaced by µ Pr + µT PrT , where

µT is an eddy viscosity and PrT is the turbulent Prandtl number.

3.2.2 – Numerical Method

The time derivative, Q̂τ , of Eqn. (3.1) is approximated using two-point backward difference

at the new time level n + 1 ,

Qˆ n +1 − Qˆ n ∆Qˆ n +1
Qˆτ = + O ( ∆τ ) = + O ( ∆τ ) . (3.8)
∆τ ∆τ
Chapter 3 – Unsteady Transonic Flow Behavior 61

All spatial derivatives are approximated by standard second-order central differences and are

represented by the difference operators δ,

Eˆ i +1, j ,k − Eˆ i −1, j ,k
( Eˆ )ξ
i , j ,k
( )
= δ ξ Eˆ
i , j ,k
(
+ O ∆ξ 2 = ) 2∆ξ
+ O ( ∆ξ 2 ) . (3.9)

The streamwise and normal derivatives, Êξ and Ĝζ , are evaluated implicitly at the new

time level n + 1 . The spanwise derivative, F̂η , is evaluated explicitly at the old time level n

but uses the n + 1 values as soon as they become available.

n +1 n n +1 n +1 n +1 n +1 n +1
Qˆ i , j , k − Qˆ i , j ,k
n
Eˆ − Eˆ i −1, j , k Fˆ i , j +1, k − Fˆ i , j −1, k Gˆ i , j , k +1 − Gˆ i , j ,k −1
+ i +1, j ,k + + (3.10)
∆τ 2 2 2

This semi-explicit treatment of the spanwise derivative enables the scheme to solve

implicitly for ∆Qˆ n +1 at all points at a given spanwise station at a time. To eliminate any

dependency the solution may have on the sweeping direction, the solver reverses the direction

of spanwise sweeping with every sweep.

n +1 n n +1 n +1 n +1 n +1 n +1
Qˆ i , j , k − Qˆ i , j ,k
n
Eˆ − Eˆ i −1, j , k Fˆ i , j +1, k − Fˆ i , j −1, k Gˆ i , j , k +1 − Gˆ i , j ,k −1
+ i +1, j ,k + + (3.11)
∆τ 2 2 2

The viscous terms R̂ξ , Ŝη and Tˆζ are evaluated explicitly, using half-point central

differences denoted here by the difference operator δ , so that the computational stencil for

the stress terms uses only three nodes in each of the three directions. Explicit treatment of the

stress terms still permits the use of large time steps since the Reynolds numbers of interest

here are fairly large. With the above described time and space discretizations, Eqn. (3.1)

becomes:

(
∆Qˆ n +1 + ∆τ δ ξ Eˆ n +1 + δη Fˆ n ,n +1 + δ ς Gˆ n +1 = ) 1
Re
(
δ ξ Rˆ n ,n+1 + δη Sˆ n ,n +1 + δ ς Tˆ n ,n+1 ) (3.12)

Note that all viscous terms include η-derivatives, for which known values at the new time

level “n+1” are used, hence the notation δ ξ Rˆ n ,n +1 .


Chapter 3 – Unsteady Transonic Flow Behavior 62

Application of Eq. (3.12) to the grid points leads to a system of non-linear, block

penta-diagonal matrix equations for the unknown ∆Qˆ n +1 = Qˆ n +1 − Qˆ n . The convection fluxes

Ê , F̂ , and Ĝ are non-linear functions of the vector of unknown flow properties Q̂ .

Therefore, Eq. (3.12) is then linearized using the Jacobean matrices A = ∂Eˆ ∂Qˆ and

C = ∂Gˆ ∂Qˆ in order to write it as a linear system of equations. Since the spanwise derivative

term is obtained explicitly, it is not necessary to compute the Jacobean matrix B = ∂Fˆ ∂Qˆ .

The Jacobean matrices A and C are presented in the work of Mello (1994). The linearization

of each term of the Eq. (3.12) is obtained as follows:

n
 ∂Eˆ  ˆ n +1 ˆ n
Eˆ n +1
= Eˆ + 
n
ˆ ( 2
) ( )
 Q − Q + O ∆τ = Eˆ + A ∆Qˆ + O ∆τ
n n n +1 2
( )
 ∂Q 
n +1
n ∆ Qˆ (3.13)
A

n
 ∂Gˆ  ˆ n +1 ˆ n
Gˆ n +1
= Eˆ + n
ˆ ( 2
) ( )
 Q − Q + O ∆τ = Gˆ + C ∆Qˆ + O ∆τ
n n n +1 2
( )
 ∂Q  n +1
n ∆ Qˆ (3.14).
C

Replacing the linearized flux vectors in the left hand side of Eq. (3.12), and setting all the

semi-explicit terms in the right hand side yields:

 (  )
 I + ∆τ δ ξ An + δ ζ C n  ∆Qˆ n +1 = RHS ˆ n ,n +1

∆τ
(
= −∆τ δ ξ Eˆ n + δη Fˆ n ,n +1 + δ ζ Gˆ n + )
Re
(
δ ξ Rˆ n ,n +1 + δη Sˆ n ,n +1 + δ ζ Tˆ n ,n +1 ) (3.15)

The resulting system of linear equation has a block pentadiagonal matrix structure,

which is considerably expensive to solve. A solution to circumvent the computational cost is

the use of an approximate factorization algorithm. The chosen approximate factorization is

presented by Beam and Warming (1978), where the system of equation reduces from a block

pentadiagonal to a block tridiagonal structure:

 (  )
 I + ∆τ δ ξ An + δ ζ C n  ∆Qˆ n +1 =  I + ∆τδ ξ An   I + ∆τδ ζ C n  ∆Qˆ n +1 + O ∆τ 3
   ( ) (3.16)
Chapter 3 – Unsteady Transonic Flow Behavior 63

which, allows the system of equations (3.15) to be written as:

 I + ∆τδ ξ An   I + ∆τδ ζ C n  ∆Qˆ n +1 = RHS


ˆ n ,n +1 (3.17)
  

Note that there is no loss of temporal accuracy, because the error incurred due to the

( )
approximate factorization is of order O ∆τ 3 . The system of equations (3.17) may now be

solved in two steps, each involving only a block-tridiagonal system of equations:

 I + ∆τδ ξ An  ∆Qˆ *n +1 = RHSˆ n ,n +1


 
 I + ∆τδ ζ C n  ∆Qˆ n +1 = ∆Qˆ *n +1
  (3.18)

However, the computational effort may be reduced by employing a diagonalization

algorithm, which reduce the block tridiagonal matrices to single diagonal ones. The procedure

to be employed is the Pulliam and Chausse (1981) algorithm. Since the Jacobean matrices A

and C have a complete set of eigenvalues and the corresponding eigenvectors, the following

similarity transformation may be used to diagonalize A and C :

A = Tξ Λξ Tξ−1 ;and C = Tζ Λζ Tζ−1 (3.19)

where the diagonal matrices Λξ and Λζ contents the eigenvalues of the matrices A and C

respectively, as far as the associated eigenvectors (transformation) matrices Tξ and Tζ . These

matrices are presented in the work of Mello (1994). Applying the similarity transformations

(2.22) to Eq. (2.20) and using the identities [ I ] = Tk Tk−1 yields:

(
 T T −1
) ( ) (
+ ∆τδ ξ Tξ Λξ Tξ−1   Tζ Tζ−1 ) ( )
+ ∆τδ ζ Tζ Λζ Tζ−1  ∆Qˆ n +1 = RHS
n n n n
ˆ n ,n +1 (3.20)
 ξ ξ    .

The fundamental simplification of the diagonal algorithm is obtained by moving Tξ

and Tζ outside the difference operators δ ξ and δ ζ , respectively:

( )
Tξn [ I ] + ∆τδ ξ Λξn  N n [ I ] + ∆τδ ς Λςn  Tτ−1 n
∆Qˆ n +1 = RHS
ˆ n ,n +1 (3.21)

where
Chapter 3 – Unsteady Transonic Flow Behavior 64

ˆ n , n +1 = ∆τ δ Eˆ n + δ Fˆ n ,n +1 + δ Gˆ n + ∆τ δ Rˆ n ,n +1 + δ Sˆ n , n +1 + δ Tˆ n ,n +1
RHS ( ) ( ) , (3.21)
ξ η ς ξ η ς
Re

and N = Tξ−1Tζ . This simplification introduces and error because Tξ and Tζ are functions of

(ξ ,η , ζ ) and cannot be arbitrarily brought in or out the derivatives. It is however believed that

the errors introduced are of order O ∆τ 2 ( ) and have the effect of making the scheme first

order accurate in time, the same order as the previous approximations.

The solution of Eqn. (3.20) involves two block-tridiagonal systems where the blocks

are diagonal matrices. The solution of Eq. (3.21) is obtained through the following steps:

( ) RHS
[ I ] + ∆τδ ξ Λξn  ∆Qˆ *n +1 = Tξ−1
n
ˆ n , n +1

[ I ] + ∆τδ ζ Λζn  ∆Qˆ **n +1 = ( N ) ∆Qˆ


n
−1 * n +1
(3.22)
∆Qˆ n +1 = T −1∆Qˆ **n +1
ζ .

The use of standard central differences to approximate the spatial derivatives can give

rise to the growth of high frequency errors in the numerical solution with time. To control this

growth, a set of 2nd/4th difference nonlinear, spectral radius based, explicit artificial

dissipation terms are added to the discretized equations. A second order implicit dissipation is

used to help the overall numerical stability of the scheme (Pulliam and Steger, 1986). The

systems of equations in (3.22) are modified as follows:

[ I ] + ∆τδ ξ Λξn − ∆τε I ∇ξ φ1∆ξ J  ∆Qˆ *n +1 = Tξ−1 ( ) ( RHS )


n
ˆ n , n +1
− Dˆ en ,n +1
(3.23)
[ I ] + ∆τδ ζ Λζ − ∆τε I ∇ζ φ3∆ζ J  ∆Qˆ **n +1 = ( N ) ∆Qˆ
n
n −1 * n +1
,

where ε I is the coefficient of implicit numerical dissipation, ∇ξ and ∆ξ denote backward

and forward difference operators, respectively, that is., ∇ξ ( i )i , j ,k = ( i )i , j ,k − ( i )i −1, j ,k  ∆ξ and


 

∆ξ ( i )i , j ,k = ( i )i +1, j ,k − ( i )i , j ,k  ∆ξ . The quantities φ1 and φ3 are defined as:


 
Chapter 3 – Unsteady Transonic Flow Behavior 65

 
V +a A4 Wc + a A6 
(φ1 )i , j ,k =  J −1
( )
 i + 2 , j ,k

U c + a A1  1 1 + max  c
U + a
, 

A1 U c + a A1 i + 1 , j ,k 

  c
 2  , (3.24)
 
(φ3 )i , j ,k =  J −1 ( 1 + max  U c + a A1 Vc + a A4 
Wc + a A6  )
 i , j ,k + 12  W + a
, 
A6 Wc + a A6 i , j ,k + 1 

  c 2 

with A1 = ξ x2 + ξ y2 + ξ z2 , A4 = η x2 + η y2 + η z2 and A6 = ζ x2 + ζ y2 + ζ z2 as auxiliary metric terms.

Dˆ en ,n +1 is the explicit fourth order dissipation, given by:

Dˆ en ,n +1 = ∆τε E ∇ξ φ1 ( ∆ξ ∇ξ ∆ξ ) + φ2 ( ∇η ∆η ) + ∇ζ φ3 ( ∆ζ ∇ζ ∆ζ )  JQˆ n


2
, (3.25)
 

where ε E is the coefficient of explicit numerical dissipation and

 
 U c + a A1 Wc + a A6 
( φ 2 )i , j , k
1
(
= J −1 Vc + a A4
2
) 1
i , j + ,k

1 + max  V + a A , V + a A 

 . (3.26)
2   c 4 c 4 i , j + 1 , k 
 2 

For points adjacent to the computational boundaries, second-order explicit dissipation is used

instead of Eq. (3.25).

The above described fourth order dissipation may lead to oscillations near shocks. To

avoid this problem, a switching function based on the second normalized streamwise

derivative of pressure :

pi +1 − 2 pi + pi −1
, (3.27)
pi +1 − 2 pi + pi −1

is used to replace the fourth order dissipation with second order dissipation near shocks

(Jameson et al., 1981).


Chapter 3 – Unsteady Transonic Flow Behavior 66

3.2.3 Turbulence Model

A slightly modified version of the Baldwin-Lomax algebraic turbulence model

(Baldwin and Lomax, 1978), is used, where the maximum shear stress is used instead of the

wall shear stress because in the vicinity of separation points, the shear stress values approach

zero at the wall. In this model, two layers are considered; in the inner layer, µT is given by:

µTinnerlayer = ρ 2
m ω (3.28)

where ω is the mean vorticity, given by:

2 2 2
 ∂w ∂v   ∂u ∂w   ∂v ∂u 
ω =  −  + −  + −  (3.29)
 ∂y ∂z   ∂z ∂x   ∂x ∂y 

and m is the mixing length, given by:

m =κd (1− e − d + A+
) (3.30)

where κ = 0.41 is the von Kármán constant, d is the distance from the wall, A+ = 26.0 is the

van Driest constant, and

ρτ max
d+ = d (3.31)
µ∞

The modification with respect to the original Baldwin-Lomax model is apparent in Eq.

(3.31), where τ max is used instead of τ wall . In the outer layer, µT is given by:

( µT )outerlayer = K c ρ c1Fw Fk (3.32)

where K c = 0.0168 is Clauser's constant, c1 = 1.6 is an empirical constant, Fw is given by:

 2
d maxU dif 
Fw = min  d max Fmax , 0.25  (3.33)
 Fmax
 

where
Chapter 3 – Unsteady Transonic Flow Behavior 67

 ω
Fmax = max  m  , (3.34)
 κ 

U dif = max ( )
u 2 + v 2 + w2 − min ( u 2 + v 2 + w2 ) , (3.35)

and d max is the distance from the wall where Fmax occurs. Also in Eq. (3.32) Fκ is given by:

−1
  0.3d  
6

Fκ = 1 + 5.5    (3.36)
  d max  

The switch between inner and outer zones occurs at the distance d c , defined as the smallest

distance from the wall for which ( µT )innerlayer = ( µT )outerlayer , i.e., the values from Eqs. (3.28)

and (3.32) are the same.

It should be noted that this change to the Baldwin-Lomax model allows the method to

treat mild separation, but it is not clear to what extent the results would be valid for massive

separation. Since the scope of the present investigation is the computation of unsteady flow

behavior of oscillating wings at moderate amplitudes of angle of attack, this model is

adequate and will be applied to the problems at hand (Mello and Azevedo, 2000).

3.2.4 Boundary Conditions

The boundary conditions must be specified along the solid surface, at the wing root

and the far field boundaries. The solid surface corresponds to the plane k=1. The unknown

conservative variables ( Q̂ ) are obtained for values between k=2 and k=(kmax-1). At the end

of each iteration, the new conservative variables values are computed at the solid surface, that

is, for k=1 where the densities and pressures are computed assuming a adiabatic wall
Chapter 3 – Unsteady Transonic Flow Behavior 68

condition, in other words the normal derivatives ∂p ∂n = ∂ρ ∂n = 0 . For the case of near

orthogonal grids this conditions are satisfied as :

4 ρi , j ,2 − ρi , j ,3 4 pi , j ,2 − pi , j ,3
ρi , j ,1 = and pi , j ,1 = . (3.37)
3 3

The velocities at the surface are computed from the non-slip condition as, ui , j ,1 = ( xτ )i , j ,1 ,

vi , j ,1 = ( yτ )i , j ,1 and wi , j ,1 = ( zτ )i , j ,1 .

The wing root corresponds to the plane j = 1. The unknown conservable variables

vector includes values from j = 2 to j = (jmax-1). Their values at k = 1, that is, at the root, are

not updated; when computing the residual at the j = 2. The fluxes at j = 1 are computed with

the symmetry condition that the contravariant velocity normal to the boundary vanishes, i.e.,

V = 0. The pressure values at j = 1 are computed using zeroth order extrapolation, so that

pi ,1,k = pi ,2,k .

The far field boundaries are located outboard of the wing tip, downstream beyond the

trailing edge. The downstream (i = 1 and i = imax) and outboard (j = jmax) boundaries are

treated in the same way. The velocity normal to the boundary is computed. Then, the

boundary conditions are imposed depending on whether it is an inflow or outflow and

whether it is subsonic or supersonic:

•Supersonic outflow: all variables are extrapolated from the interior of the domain;

•Subsonic outflow: the pressure is fixed to be the free-stream value and the other

variables are extrapolated;

•Subsonic inflow: the density is extrapolated from the interior of the domain and the

other variables are fixed from the free-stream;

•Supersonic inflow: all variables are fixed to be the free-stream values.


Chapter 3 – Unsteady Transonic Flow Behavior 69

3.3 – Transonic Flow Investigations of a Low Aspect Ratio Wing

3.3.1 - Preliminary Considerations

The finite difference solution of the Navier Stokes equations has been applied to the

simulation of the F-5 wing, since it is a low aspect ratio wing, where the three-dimensional

effects are more significant. The numerical method used in the present work has been

previously validated (Mello, 1994) against experimental data (Tijdemann et al., 1978) for the

case of a model of the F-5 wing at several Mach numbers up to 0.95, in steady state

conditions, and oscillating at frequencies of 20 and 40 Hz. This wing was tested in wind

tunnel, under transonic flow conditions, oscillating in rigid pitch around an axis perpendicular

to the root midchord, around a steady state angle α o = 0.0 0 , and in the steady state situation.

In the work of Mello (1994) the computational simulations were performed employing

a hybrid Navier-Stokes/full potential method. This method is based on the subdivision of the

computational domain in two zones. In the inner zone, where the viscous effects are more

relevant, the Navier Stokes equations are solved, at the same time that in the outer zone, the

solution is based on the full-potential equations. Further descriptions about this numerical

approach are found in the work of Mello (1994). Even though of the hybrid method presents

good results and a lower computational cost, in the present investigations, the Navier-Stokes

solution has been employed because, at the present time, its computational cost is not

prohibitive, even for unsteady flow computations.


Chapter 3 – Unsteady Transonic Flow Behavior 70

3.3.2 - Validation of the Computational Procedure for Steady Flow Computations

In a first step, a steady state simulation of the F-5 wing is performed for the validation

of the Navier-Stokes method. The chosen undisturbed flow conditions to perform the desired

validation are: M ∞ = 0.95 , and Reynolds number, here defined as Re1 = ρ∞U ∞ cr µ ∞ is 11

million, scaled by the wing root chord cr = 0.6396 meters. This wing was tested in the NLR

laboratories, under these same transonic flow conditions, at steady state angle of attack

α o = 0.0 0 . This undisturbed flow speed is chosen because, at this situation, the strongest shock

waves occur, that is, where the transonic effects are more pronounced. The numerical method

used here does not include any prediction of transition location, which may be arbitrarily set.

For the conditions under consideration the flow is considered completely turbulent.

The computational mesh here employed was an algebraic generated “C” type

topology, with 141 points in the ξ direction, tangent to the lifting surface boundary, where

121 points are over the lifting surface. In the η direction, normal to the wing surface, there are

41 points between the solid surface and the limit of the computational mesh. And finally, in

the ζ direction there are 22 points aligned with the spanwise direction, where 17 points are

over the lifting surface, and the remaining are between the wing tip and the computational

domain limit. The computational mesh, which surrounds the lifting surface, as well as its

profile are presented in Fig. (3.1).


Chapter 3 – Unsteady Transonic Flow Behavior 71

0.2
upper surface
lower surface
0.1

t/c 0.0
0 0.2 0.4 0.6 0.8 1
-0.1

-0.2
x/c

Figure 3.1- Finite difference computational mesh surrounding the F-5-NLR wing and

its profile.

For consistency, the same boundary conditions were employed in the validation,

namely no-slip at the solid surface, approximate non-reflective conditions at the far field and

no-slip wall boundary conditions at the wing root, since this validation is based on

experimental measurements for a wind tunnel model. A comparison between the computed

steady pressures and corresponding experimental measurement are presented in Fig. (3.2) for

three spanwise stations.

-0.8 -0.8 -0.8


-0.6 -0.6 -0.6
-0.4 -0.4 -0.4
-0.2 -0.2 -0.2
0 0 0
C C C (calc) C C (calc)
P P 0.2 P,L P 0.2
0.2 C
P,L
(calc) P,L
C (calc) C (calc)
C (calc) P,U P,U
0.4 P,U 0.4 C (exp) 0.4 C (exp)
C (exp) P,L P,L
P,L 0.6 C (exp) 0.6 C (exp)
0.6 C (exp) P,U P,U
P,U
0.8 0.8 0.8
1 1 1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x/c x/c x/c

35,2 % 72,1 % 97,7 %

Figure 3.2: Steady pressure distribution over the F-5 NLR wing at M∞ = 0.95.
Chapter 3 – Unsteady Transonic Flow Behavior 72

The dominating feature to be noted is the formation of a shock wave in the upper and

lower surfaces. The matching of the pressures in both surfaces are well predicted, except at

station 97,7 %, where the suction peak is overestimated. Nevertheless, the results are in good

agreement with the experimental data, qualifying the numerical method for the unsteady

computation of this wing at the same undisturbed flow conditions.

3.3.3 – Validation of the Computational Procedure for Unsteady Flow

Computations

The time domain unsteady flow simulations were performed from a steady state

converged solution. The unsteady flow simulation initiates imposing a harmonic pitching

motion to the wing, around the root midchord axis. Under these pitching oscillation, the lifting

surface deforms aeroelastically, since the experimentally tested wing presents flexibility.

Therefore, the lifting surface will harmonically move at a given vibration mode shape. The

test condition to be simulated is presented by Tijdeman et al. (1978), for Mach number 0.95,

and the harmonic motion frequency is 40 Hz.

The computation of the first harmonic components of the pressures is performed by

Fourier transformations applied to pressures time histories, given by:


τ1 + 2π / kr
k
( )
Re C p (kr ) =
π ∫ C p (τ ) sin ( krτ ) dτ , and (3.38)
τ 1

τ1 + 2π / kr
k
( )
Im C p ( kr ) =
π ∫ C p (τ ) cos ( krτ ) dτ , (3.39)
τ1

where kr is the reduced frequency of a harmonic motion, and τ is the non-dimensional time,

as defined in the non-dimensionalization of the Navier-Stokes equations. Here, the reduced

frequency is defined as kr = ω c a∞ , where ω is the circular frequency c , is a reference chord,


Chapter 3 – Unsteady Transonic Flow Behavior 73

and a∞ is the speed of the sound, in conformity with the non-dimensional time defined by

τ = a∞t c . Since the time domain simulation of the Navier-Stokes equations is based on a

discrete time marching algorithm, the Fourier integrals, presented in relations (3.38) and

(3.39), are approximated by discrete Fourier transforms as follows:

kr M ∞ ∆τ m1 + mT

( )
Re C p ( kr ) =
2π∆α
∑C pm sin ( kr m∆τ ) , (3.40)
m = m1

kr M ∞ ∆τ m1 + mT

( )
Im C p ( kr ) =
2π∆α
∑ C p m cos ( kr m∆τ ) . (3.41)
m = m1

where ∆τ is the computational non-dimensional time step over which the time marching

simulation was performed. Therefore, the complex first harmonic components of the pressure

coefficient differences C p ( kr ) are obtained from the real and imaginary part of the first

harmonic components, given by Eqs (3.40), and (3.41) as:

{C ( ik )} = ( Re {C ( ik )}) + i ( Im {C ( ik )}) .
p r p r p r
(3.42)

The time simulation period is twice the time expended in one complete harmonic

motion cycle, in order to avoid undesired transient effects due to the transition from a steady

state converged solution to the unsteady one. Hence, unsteady pressures to be acquired will be

related to the second simulation cycle. This needs to be done in order to avoid spurious effects

related to undesired frequency contents, when the transformation from the time to the

frequency domain via Fourier analysis of the acquired pressured coefficients time history is

performed. In Figs. (3.3) through (3.5) the first harmonic components of the pressure

coefficients at given spanwise stations are presented for the aforementioned test conditions.

The time domain unsteady flow simulations were performed for two time steps, in order to

evaluate its effect on the unsteady flow solution.


Chapter 3 – Unsteady Transonic Flow Behavior 74

20 4
Re C (calc, ∆t=.005)
P,i,L
Re C (calc, ∆t=.005) 2
15 Re C
P,i,U
(exp)
P,i,L
Re C (exp)
Re C
P,i,U
(calc, ∆t=.002)
0
10 P,i,L
ReCpi Re C (calc, ∆t=.002)
P,i,L -2
ImCpi
5 Im C (calc, ∆t=.005)
P,i,L
Im C (calc, ∆t=.005)
-4 P,i,U
Im C (exp)
P,i,L
0 Im C (exp)
-6 P,i,U
Im C (calc, ∆t=.002)
P,i,L
-5 Im C (calc, ∆t=.002)
P,i,U
-8

-10 -10
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x/c x/c

Figure 3.3: Real and imaginary parts of the pressure coefficients over the F-5 NLR

wing (M∞ =0.95, 35.2% spanwise station, frequency = 40 Hz).

30 3
25 Re C
P,i,L
(calc, ∆t=.005)
Re C (calc, ∆t=.005) 0
P,i,U
20 Re C
P,i,L
(exp)
Re C (exp) -3
P,i,U
15 Re C (calc, ∆t=.002)
P,i,L
ReCpi Re C (calc, ∆t=.002) Im C
P,i,L
(calc, ∆t=.005)
10 P,i,U ImCpi -6 Im C (calc, ∆t=.005)
P,i,U
Im C (exp)
P,i,L
5 Im C (exp)
-9 P,i,U
Im C (calc, ∆t=.002)
P,i,L
0 Im C (calc, ∆t=.002)
P,i,U
-12
-5

-10 -15
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x/c x/c

Figure 3.4: Real and imaginary parts of the pressure coefficients over the F-5 NLR

wing (M∞ =0.95, 72.1% spanwise station, frequency = 40 Hz).


Chapter 3 – Unsteady Transonic Flow Behavior 75

3
Re C (calc, ∆t=.005)
40 P,i,L
Re C
P,i,U
(calc, ∆t=.005) 0
Re C (exp)
P,i,L -3
30 Re C (exp)
P,i,U
Re C (calc, ∆t=.002)
Re C
P,i,L
(calc, ∆t=.002)
-6
P,i,U
20 -9 Im C (calc, ∆t=.005)
ImCpi Im C
P,i,L
(calc, ∆t=.005)
ReCpi -12 P,i,U
Im C (exp)
10 P,i,L
Im C (exp)
-15 P,i,U
Im C (calc, ∆t=.002)
P,i,L
0 -18 Im C (calc, ∆t=.002)
P,i,U
-21
-10
-24
0 0.2 0.4 0.6 0.8 1
x/c 0 0.2 0.4 0.6 0.8 1
x/c

Figure 3.5: Real and imaginary parts of the pressure coefficients over the F-5 NLR

wing (M∞ =0.95, 97.7% spanwise station, frequency = 40 Hz).

A good agreement in unsteady pressure coefficients was observed over all wing

stations, although a few discrepancies were noted at spanwise station 97.7 %, with respect to

difficulties in obtaining a smooth grid around the it within reasonably economical grid sizes.

The computations with a smaller time step show some improvement in the real part of the

pressures, mainly with regard to the shock capturing, as shown in Fig. (3.4). However, further

reduction of the time step does not guarantee a good shock capturing, since this capability

mainly depends on the discretization of the domain around the shock position. Overall, the

computed unsteady pressures correlate well with experimental measurements, validating the

numerical method to be used for the linear/nonlinear behavior investigation to be further

developed.
Chapter 3 – Unsteady Transonic Flow Behavior 76

3.3.4 –Investigation Methodology

The linear/nonlinear behavior investigation is performed based on the same test case

employed to validate the numerical procedure. The undisturbed flow conditions, as well as the

mean angle of attack position, are the same which the validation of the numerical procedure

was based on. However, the time domain simulation of the F-5 wing is performed based on a

rigid body oscillation in pitch, that is, without considering the lifting surface flexibility,

around an axis perpendicular to the root midchord. The rigid body pitching amplitudes is also

denoted here as “dynamic angles of attack”, were 0.125o to 1.5o , varying in equally spaced

steps of 0.125o. The simulations were conducted for reduced frequencies of 0.1, 0.15, 0.2,

0.25 and 0.3.

The objective is to evaluate the behavior of the pressures and the associated loading,

as a function of the displacements of the wing and the reduced frequency. The aim was to

establish the limit for linear behavior with dynamic angle of attack for the range of reduced

frequencies tested, and at each spanwise station, so that three-dimensional effects could be

identified. From the application of the discrete Fourier transformations to the pressure time

histories, given by Eqs. (3.40) and (3.41), the same transformations are applied to the lift

coefficient CL and moment coefficient CM about the quarter-chord. These loading derivatives

are computed by the integration of the pressures along the non-dimensional chord length as:

(τ ) − C pupper (τ ) d 
1 x
CL (τ ) =
π ∫
C lower
p  (3.43)
0 c

(τ ) − C plower (τ ) 
2 x 1  x
CM (τ ) =
π ∫
C upper
p − d   (3.44)
0  c 4  c

The shock positioning along the wing section chord is determined by using the

criterion of maximum pressure coefficient gradient along the chord length. However, the grid

resolution in around the shock positions is coarse, thus a polynomial interpolation is needed in
Chapter 3 – Unsteady Transonic Flow Behavior 77

order to find the maximum gradient position. A polynomial interpolation of third degree was

chosen, in order to yield a linear expression for the second derivative with respect to the non-

dimensional chordwise coordinate, from which the maximum gradient position can be found.

The points employed in the interpolation are chosen to be nearby the shock location, taking

into account the imposed mesh displacements required by the unsteady computations. After

the shock location time history is determined, the same discrete Fourier transformations are

applied leading to first harmonic counterparts of the shock displacement as a function of the

reduced frequency.

3.3.5 Analysis of the Linear/Nonlinear Behavior

In the present investigation, it was found that obtaining frequency content for the

aerodynamic coefficients Cp, CL and CM gave more consistent results than the ones for the

shock position, which depends on how good the polynomial interpolation is. Inadequate

polynomials would easily result in spurious shock positions. Three representative spanwise

stations at 35.5%, 72.1% and 97.7% were chosen for the present investigation, in order to

evaluate the three-dimensional effects. In the following, some results concerning the linearity

studies will be presented.

Figure (3.6) presents a sample time history for the pressure coefficient differential at

the three chosen spanwise stations. The associated corresponding frequency contents, for the

three spanwise stations under consideration for a dynamic angle of attack ∆α=1.0º and

reduced frequency k=0.1 are presented tin Fig (3.7).


Chapter 3 – Unsteady Transonic Flow Behavior 78

0.4
35.5%
0.3 72.1%
97.7%

0.2

∆Cp
0.1

0
0.0 20.0 40.0 60.0 80.0

-0.1

-0.2 τ (non-dimensional)

Figure 3.6 : Pressure difference coefficient time history over F-5 wing (M∞ =0.95,

∆α=1.0º; k=0.1).

0.30 300 35.5%


35.5% 72.1%
0.25 250
Phase angles (deg)

72.1% 97.7%
Amplitudes

0.20 97.7% 200

0.15 150

0.10 100

0.05 50
0.00 0
0.1 0.2 0.3 0.4 0.1 0.2 0.3 0.4
k k

Figure 3.7: Frequency contents (amplitudes and phases) of differential pressure

coefficient over F-5 wing (M∞ = 0.95, ∆α=1.0º; k=0.1).

It may be observed that the pressure coefficient time response is not in a single

harmonic of the movement frequency, especially near the wing tip. This indicates a non-

linear, hysteresis-like behavior. Also, three-dimensional effects are clearly present, as a one-

harmonic sinusoidal behavior is evident at the inboard station.. Even though the first harmonic
Chapter 3 – Unsteady Transonic Flow Behavior 79

is dominant for all stations, the ones near the tip have significant components in the second

and third harmonics. Also, these higher harmonics are out of phase with respect to the wing

motion.

It should be noted that these higher harmonic effects are not included in linear

methods such as the doublet-lattice method and correction methods based on modification of

the aerodynamic influence coefficient matrices would not be readily adaptable to incorporate

higher-harmonic corrections. However, there are situations where a second mode is close to a

multiple of a first mode. In such case, the presence of a second harmonic in the lower mode

aerodynamic response may facilitate the exchange of energy between the modes and be a

contributing factor to coalescence

At reduced frequency 0.2, the first harmonic of the response in lift and moment

coefficients, upper and lower surface shock motion were computed for the various dynamic

angles of attack. The results for station 35.5% span are shown in Fig. (3.8) for the amplitudes

and phase angles, of aerodynamic coefficients and shock motion. Linear trend lines are placed

in the amplitude plot to illustrate deviations from linear behavior. Those are not drawn in the

phase plot, because during the analysis of the phase lag variation a basic linear behavior could

not be established and thus a deviation from linearity could not be characterized.

0.35
CL
0.30
CM x10 30
CL
0.25 DXsu x10
Amplitudes

CM
Phase (deg)

DXsl x10
0.20 20 DXsu
DXsl /10
0.15

0.10 10
0.05

0.00 0
0 0.5 1 1.5 2 0 0.5 1 1.5 2
∆α (deg) ∆α (deg)

Figure 3.8: Effect of dynamic angle of attack on unsteady aerodynamic coefficients

and shock motion (amplitudes and phases; station 35.5%; k=0.2).


Chapter 3 – Unsteady Transonic Flow Behavior 80

In order to identify the limit for each case, a linear equation was estimated so that it

would start at the origin and pass through the first and second points, which corresponded to a

dynamic angles of attack of 0.125 and 0.250 degrees. A linear equation was then reduced for

these first values, at which the linear behavior would be expected. Then this linear

relationship is used to extrapolate to the following points and a percent deviation of the actual

unsteady coefficient or shock displacement to the extrapolated value was computed. The first

point where the deviation exceeded 5% and the point before that were used in an interpolation

to find the dynamic angle of attack where the linear limit was.

Two criteria were used to establish the linear boundary: 5% deviation in pitching

moment amplitude and 5% deviation in shock motion amplitude. For the latter, the deviation

is in the upper surface case only. The lower surface shock is weaker than the upper surface

one, because the assimetry of the wing profile (see Fig. (3.1) ). The shock movement then

presents a significant dispersion probably promoted by the turbulent boundary layer.

Using the above criteria, the linear limits for the test cases corresponding to Fig. (3.8),

that is, reduced frequency 0.2 and station 35.5%, were found to be ∆α=0.61º for the pitching

moment criterion and ∆α=0.58º for the shock displacement criterion. The results for the same

reduced frequency of 0.2 at station 72.1% span are shown in Fig. (3.9) for the amplitudes and

phase angles of aerodynamic coefficients and shock motion. Linear trend lines are placed in

the amplitude plot to illustrate deviations from linear behavior. It is apparent from Fig. (3.9)

that the deviation from linear behavior occur at smaller dynamic angles. Indeed, using the

above described procedure to compute the linear limits yields linear limits of ∆α=0.34º for the

pitching moment criterion and ∆α=0.50º for the shock displacement criterion.
Chapter 3 – Unsteady Transonic Flow Behavior 81

0.25
50
0.20 CL 40
CL
Amplitudes

Phase (deg)
0.15 CM 30 CM
x10
DXsu
DXsu
0.10 20
x10 DXsl /10
DXsl
0.05 x10
10

0
0.00
0 0.5 1 1.5 2
0 0.5 1 1.5 2 ∆α (deg)
∆α (deg)

Figure 3.9: Effect of dynamic angle of attack on unsteady aerodynamic coefficients

and shock motion (amplitudes and phases; station 72.1%; k=0.2).

The results for the same reduced frequency of 0.2 at station 97.7% span are shown in

Fig. (3.10) for the amplitudes and phase angles of aerodynamic coefficients and shock

motion. Again, linear trend lines are placed in the amplitude plot to illustrate deviations from

linear behavior. The deviation from linear behavior seem to occur even earlier now in terms

of dynamic angles. For this case, the computed linear limits are ∆α=0.32º for the pitching

moment criterion and ∆α=0.27º for the shock displacement criterion. Even though the results

at the outboard station should be regarded with caution, the overall presented results show that

the linear boundary is very dependent on the spanwise location, outboard being more severe.

Three-dimensional effects are therefore significant and using two-dimensional linear

boundaries in a general problem would seem to be questionable. Another feature to be noted

is that the shock strength increases with the proximity to the wing tip. Then, the linear

boundary could be affected by the upper surface shock strength.


Chapter 3 – Unsteady Transonic Flow Behavior 82

0.25 30
CL CL
CM
0.20 CM 20 DXsu

Phase (deg)
Amplitudes

x10 DXsl /10


0.15 DXsu
10
DXsl
0.10
x10
0
0.05

0.00 -10
0 0.5 1 1.5 2 0 0.5 1 1.5 2
∆α (deg) ∆α (deg)

Figure 3.10: Effect of dynamic angle of attack on unsteady aerodynamic coefficients

and shock motion (amplitudes and phases; station 97.7%; k=0.2).

This linear/nonlinear behavior investigation presented some indications that unsteady

transonic flow is strongly dependent of the amplitude of the motion. However, in aeroelastic

stability analysis the local mean angle of attack variations are small, since the linear unsteady

aerodynamic models are developed considering small disturbances of the boundary

conditions. Then, the associated displacements are usually less than the aforementioned linear

limit values, indicating that, the unsteady character of the transonic flow can be considered as

a contribution predicted by a small disturbance linear aerodynamic model, to be added to a

transonic flow steady pressures distribution. Otherwise, as soon as the amplitude of the

aeroelastic deformations grows up, other non-linear phenomena can appear, and the suggested

linear aerodynamic model representation need to be further investigated.

Simulations for the several reduced frequencies with the corresponding determination

of linear limits allow for a comparison of these limits at different spanwise stations using both

linearity criteria. This is illustrated in Fig. (3.11), where the linear boundaries are plotted

against reduced frequency for the three spanwise stations (35.5%; 72.1% and 97.7%) for the

pitching moment and Fig. (3.12) for the upper surface shock displacement criteria.
Chapter 3 – Unsteady Transonic Flow Behavior 83

0.30

0.25

k Non-linear
Linear
0.20

35.5% span
0.15 72.1% span
97.7% span

0.10
0.0 0.2 0.4 0.6 0.8 1.0
∆α (deg)

Figure 3.11: Boundaries for linear behavior based on the CM criteria.

0.30

0.25

k Non-linear
0.20

Linear 35.5% span


0.15 72.1% span
97.7% span

0.10
0.0 0.2 0.4 0.6 0.8 1.0
∆α (deg)

Figure 3.12: Boundaries for linear behavior based on the shock displacement criteria.
Chapter 3 – Unsteady Transonic Flow Behavior 84

The results shown in Figs. (3.10) and (3.11) indicate that the shock displacement

criterion is more conservative than the pitching moment criterion, as found by Dowell et al.

(1983). In addition, the linear boundaries based on both pitching moment and shock

displacement criteria clearly depend on the spanwise station. For the first criterion, at the

inner station the behavior is similar to the two–dimensional, inviscid case as presented by

Dowell et al (1983). For outer spanwise stations, the shock is stronger and three-dimensional

effects are more pronounced. Consequently, the linear behavior is affected, as may be noted in

Fig. (3.10). A somewhat surprising result is that the shock movement-based linear limit does

not seem to increase with reduced frequency for the inner stations, and even decreases with

reduced frequency for the outer station

The linear limits calculated using the moment coefficient criterion were found to

depend significantly on the spanwise station. This indicates that simply using two-

dimensional linear limits, as presented by Dowell et al. (1983), would not be appropriate

when dealing with correction methods. Notwithstanding the dependence of the linear limits on

reduced frequency and spanwise station, it is clear that some degree of linear behavior may be

assumed allowing the approximation of an unsteady transonic flow by means of small

disturbance linear governing equations.


4 - CORRECTION METHODS FOR APPROXIMATE NONLINEAR

AERODYNAMIC MODELING

4.1 Developments on the Pressure Matching Approach

The main purpose of the correction methods to be developed here is the computation

of unsteady transonic flows for dynamic aeroelastic response and aeroelastic stability

analysis. Among the correction methods reviewed in the first chapter, the chosen procedure to

be developed, and further extended, is the pressure matching based on the downwash

weighting approach. The main idea of such procedure is the modification of the downwash

vector inside the pressure to downwash linear relationship. The transonic flow reference

conditions can be based on, either computational fluid dynamics (CFD) solutions of the

nonlinear fluid dynamic governing equations, or experimental data.

4.1.1 – Downwash Weighting Method Based on Steady Reference Conditions

The basic equation over which this procedure is developed, is the algebraic pressure to

downwash relationship, derived from the application of discrete element kernel function

methods to model the linearized potential flow equation. This relation is rewritten in a

simplified form as a function only of the reduced frequency as:

{∆C p (ik )} =  AIC (ik ) {w(ik )} . (4.1)

The downwash vector is related to the lifting surface displacements by boundary conditions

defined in a small disturbances context, and may be regarded as an effective dynamic angle of

attack at each of the lifting surface elements. In the frequency domain, these boundary

conditions are rewritten here as:


Chapter 4 – Correction Methods for Approximate Nonlinear Aerodynamic Modeling 86


{w ( x, y, 0, ik )} = {h ( x, y, 0)} + ik {h ( x, y, 0)} , (4.2)
∂x

recalling that {h ( x, y, 0)} is an out of plane lifting surface displacement mode shape, and

{w ( x, y, 0, ik )} is the resulting downwash with respect to the modal motion. In steady state

conditions, this downwash is reduced to the derivative of the associated mode shape

displacements with respect to the flow direction, “x” as:


{w ( x, y, 0)} = {h ( x, y, 0)} , (4.3)
∂x

leading to a steady state pressure to downwash relationship given by:

{∆Cp (ik = 0)} = [ AIC (ik = 0)]{hx } . (4.4)

The downwash correction method to be developed is based on the matching of

externally generated nonlinear quasi-steady pressure differences. The quasi-steady

approximation is defined here as the variation between two states of any given property at a

infinitely large time step. The computation of the nonlinear quasi-steady pressure differences

is performed based on a quasi-steady modal displacement, where the pressure coefficient

differences are computed at these two distinct mode shape deformation states, independently

of any time variation. For the present investigations, the nonlinear pressure differences are

obtained from CFD solutions of the Navier-Stokes equations by the use of the numerical

method presented in chapter 3. The chosen displacement of the lifting surface is a rigid body

pitching rotation mode, where the associated quasi-steady displacement is the same as a

variation in angle of attack given by ∆α = α F − α I . The indexes “I”, and “F” refer to an initial

and a final mode shape displacement state, respectively.

The nonlinear pressure coefficient differences are here denoted by the superscript “nl”.

They are computed from the solutions of the Navier-Stokes equations, as the difference of the

pressures between lower {C lp } and upper {C up } surfaces of the lifting surface as:
Chapter 4 – Correction Methods for Approximate Nonlinear Aerodynamic Modeling 87

{∆C } = {C } − {C }
nl
p
l
p
u
p
. (4.5)

The resulting quasi-steady pressure coefficient difference contributions are defined as the

subtraction between these quantities at a final and a initial position of the lifting surface

displacement as:

{δ∆C } = {∆C } − {∆C }


nl
p
nl
p F
nl
p I
, (4.6)

which may be regarded as pressure coefficients disturbances due to the quasi steady motion.

The resulting pressure disturbance can be related to the lifting surface quasi-steady motion, in

the same way as the absolute pressure coefficients differences are related to the lifting surface

displacements, accordingly with Eq. (4.4), as:

{δ∆C p (ik = 0)} = [ AIC (ik = 0)]{αF − αI } (4.7)

However, the pressure disturbance to downwash relationship, given by Eq. (4.7), is a

result from a linear unsteady aerodynamic model. Thus, the downwash needs to be weighted

by the introduction of a multiplying matrix operator, here denoted as [WT (ik = 0)] . Its role is

to match the nonlinear quasi-steady reference conditions. The argument of the weighting

operator, denoted as ik = 0 , indicates that, its role is the matching of quasi-steady pressure

differences. Therefore, the resulting equation for the quasi-steady downwash correction is

given by:

{δ∆C pnl (ik = 0)} = [ AIC (ik = 0)][WT (ik = 0)]{αF − αI } . (4.8)

In order to have a weighting function, which is independent of the amplitude of the

quasi-steady motion, a non-dimensional pressure coefficient disturbance rate is defined,

obtained from the division of the quasi-steady pressure disturbances by the corresponding

displacements as:

  
 ∂ {δ∆C nl }   {∆C p }F − {∆C p }I  
 nl nl
 p  
  =  = {δα ∆C pnl (ik = 0)} . (4.9)
 ∂α 
 (ik =0)  (αF − αI ) 

 (ik =0)

Chapter 4 – Correction Methods for Approximate Nonlinear Aerodynamic Modeling 88

These pressure disturbance rates are used as new reference conditions for the

weighting operator computation. Likewise, the downwash vector needs to be divided by the

amplitude of the displacement in order to have Eq. (4.8) mathematically consistent. The

downwash vector is the same as the variation in angle of attack ∆α = α F − α I . Thus, the

remaining term in the right hand side of Eq. (4.8) is the desired operator resulting from the

solution of the linear system of equations given by:

[ AIC (ik = 0)]


−1
{δα∆C pnl (ik = 0)} = [WT (ik = 0)]{1} . (4.10)

One should note that the left-hand-side of Eq. (4.10) may be regarded as the ratio of a

modified disturbance downwash mode {δ h } ,


nl
x by the quasi–steady mode

{δ h } = ∆α = α
x F − α I . This modified disturbance downwash vector, multiplied by the

aerodynamic influence coefficients matrix generates the nonlinear pressure coefficients

difference rates. The resulting downwash weighting coefficients result from the ratio between

each of the elements of both downwash vectors as:

(hxnl )i
(WT )ii = , (4.11)
(hx )i
which may compose the elements of a matrix operator [WT (ik = 0)] . This matrix is

diagonal in order to satisfy the dimensions of the matrix multiplication inside Eq. (4.10).

The pressure differences associated to the reference conditions can be obtained from

different sources, such as CFD computations or experimental measurements. In both cases,

the data acquisition points usually do not agree with the panels control points. Hence, an

interpolation is necessary for the determination of those quasi-steady pressure differences at

the desired locations, in agreement with the discrete element method aerodynamic mesh.

These interpolations can be since surface splines (Zona Technology, 2003) to linear

chordwise interpolations (Silva et al. 1999).


Chapter 4 – Correction Methods for Approximate Nonlinear Aerodynamic Modeling 89

The subsequent step is the modification of the unsteady aerodynamic loading (Eq.

2.25) by the introduction of the weighting operator, leading to a approximate nonlinear

aerodynamic loading given by:

{Lnl
a (ik )} = q∞ [ S ]  AIC (ik ) [WT (ik = 0)]{w (ik )} (4.12)

In steady state conditions, the nonlinear aerodynamic loading is completely restored

by the application of the weighting operator. Looking at Eq. (4.12), one should note that, even

though the weighting matrix is defined based on quasi-steady conditions, it is assumed that

the weighting of the downwash will be performed for all the set of reduced frequencies to be

considered in the solution of the aeroelastic stability problem.

4.1.2 – Downwash Weighting Method Based on Unsteady Reference Conditions

Another capability of downwash correction procedures to be investigated is the use of

computed or experimental unsteady pressure differences. In such cases, the reference

conditions are based on frequency-dependent pressures, computed from the time domain

aerodynamic response. Pressure time histories, for example, may result from the lifting

surface displacements due to a prescribed motion. These displacements are arbitrary, as they

may be, for example, an impulse-type or an oscillating harmonic motions. For the present

downwash correction investigation, the unsteady pressure differences computation was

performed based on a time domain CFD solution of the Navier-Stokes equations, using the

finite differences approach presented in Chapter 3.

The chosen displacement may be any unsteady motion associated to a displacement

mode shape of the lifting surface. In the present investigations, it is chosen a rigid body

harmonic pitching rotation around the root midchord axis of the lifting surface, at a preset

reduced frequency kr . This motion leads to an unsteady downwash, which also may be
Chapter 4 – Correction Methods for Approximate Nonlinear Aerodynamic Modeling 90

regarded as an unsteady perturbation in angle of attack, with amplitude equals to the

amplitude of the pitching motion ∆α.

Since the linear aerodynamic model to be corrected is defined in the frequency

domain, it shall be necessary to transform the reference time domain pressure differences to

the frequency domain. The most adequate approach is the first harmonic components

computation of the pressure differences using a Fourier transform algorithm applied to their

time histories. The Fourier transformations used herein to obtain the frequency domain

components of the pressure differences are given by:


τ1 + 2π / kr
k
(
Re ∆C (ikr ) =
nl
p ) π ∫ ∆C p (τ ) sin ( krτ ) dτ , and (4.13)
τ1

τ1 + 2π / kr
k
(
Im ∆C nl
p ( ikr ) ) =
π ∫ ∆C p (τ ) cos ( krτ ) dτ , (4.14)
τ1

where kr is given by kr = ω c a∞ , τ = a∞t c is the non-dimensional time, as defined in the

non-dimensionalization of the Navier-Stokes equations. Since the time domain simulation of

the Navier-Stokes equations is based on a discrete time marching algorithm, the Fourier

integrals, presented in relations (4.13) and (4.14), shall be approximated by discrete Fourier

transforms as follows:

k M ∆τ m1 + mT

(
Re ∆C nl
p ( ikr ) ) = r ∞
2π∆α
∑ ∆C p m sin ( kr m∆τ ) , (4.15)
m = m1

kr M ∞ ∆τ m1 + mT

(
Im ∆C pnl ( ikr ) = ) 2π∆α
∑ ∆C p m cos ( kr m∆τ ) . (4.16)
m = m1

where ∆τ , is the computational non-dimensional time step over which the time marching

simulation was performed. Therefore, the complex first harmonic components of the pressure

coefficient differences ∆C pnl ( ik ) are obtained from the real and imaginary part of the first

harmonic components, given by Eqs (4.15), and (4.16) as:


Chapter 4 – Correction Methods for Approximate Nonlinear Aerodynamic Modeling 91

{∆C ( ik )} = ( Re {∆C ( ik )}) + i ( Im {∆C ( ik )})


nl
p r
nl
p r
nl
p r
. (4.17)

{ }
Since the unsteady pressure differences ∆C pnl ( ikr ) are computed, it is possible to determine

correction factors, which satisfy a system of equations analogue to Eq. (4.10). Special care

needs to be taken regarding definition of the reduced frequency for the Navier-Stokes

simulations. The reduced frequency differs from that defined for the discrete element kernel

function method by a factor of kr = k ⋅ ( M ∞c b) , since k = ωb U ∞ . Thus, in order to have

unsteady pressures computed by the Navier-Stokes simulation and kernel function methods at

the same reduced frequency, the aforementioned relation needs to be satisfied. Another

feature to be noted is that the unsteady pressure coefficients are divided by the amplitude of

the motion ∆α , as indicated in Eqs (4.15) and (4.16).

The system of equations (4.8) is now rewritten as function of a frequency dependent

weighting operator given by:

{∆C ( ik )} =  AIC ( ik ) WT ( ik ) {w ( ik )}


nl
p r r r r , (4.18)

where {w ( ik )} ,
r is the downwash vector divided by amplitude of the motion ∆α . The

diagonal weighting matrix coefficients are obtained from the ratio between a modified

unsteady downwash, computed by:

{wnl (ikr )} =  AIC (ikr ) {∆C pnl (ikr )}


−1
, (4.19)

and the known downwash associated to the prescribed lifting surface motion, leading to:

(wnl (ikr ))i


(WT (ikr ))ii = . (4.20)
(w (ikr ))i
The nonlinear unsteady aerodynamic loading computation is performed by introducing in Eq.

(2.25), the correcting weighting operator, which multiplies the downwash vector, leading to

an approximate aerodynamic loading given by:


Chapter 4 – Correction Methods for Approximate Nonlinear Aerodynamic Modeling 92

{Lnl
a (ikr )} = q∞ [ S ]  AIC (ikr ) [WT (ikr )]{w (ikr )} (4.21)

The same comments regarding the steady loading calculation are valid for the

unsteady case. The choice of the reduced frequency over which the nonlinear unsteady

pressure differences are referred is arbitrary, however a good suggestion is the flutter reduced

frequency predicted from the purely linear stability analysis of the aeroelastic model. This

assumption allows the best correction of the unsteady aerodynamic loading at the critical

reduced frequency, representing an improvement on the prediction of transonic effects near

the flutter boundary. However, for different reduced frequency values, there is no guarantee

that the computation of the unsteady transonic loading has physical significance. This is so

because the weighting function is computed based on a given reduced frequency value, which

shall be different from the one in this different unsteady flow condition.

4.2 – Extension of the Downwash Weighting Method Based on the Kernel

Expansion Hypothesis

4.2.1 – Motivation

An alternative for the enhancement of the nonlinear unsteady pressure based

procedure is the generation of weighting operators computed from reference pressure

differences at different reduced frequency values. These pressure differences in the frequency

domain may be computed by two possible approaches, as the Fourier transform of an impulse

or a harmonic time domain aerodynamic responses. These resulting pressure differences are

either computed for frequency spectrum, in the case of the impulse type aerodynamic

response, or at a single frequency value for the case of the harmonic aerodynamic response.
Chapter 4 – Correction Methods for Approximate Nonlinear Aerodynamic Modeling 93

The computation of the pressure differences based on impulse type aerodynamic

response requires accurate time domain CFD simulation. This approach may be regarded as

an extension of the previous unsteady pressure based downwash correction method, where the

harmonic nature of the aerodynamic response restricts the correction to single frequency

dependent reference pressure differences. Therefore, the resulting weighting function will be

consistent for the correction of the aerodynamic model computed for the same reduced

frequency associated to the reference data.

The frequency contents of the impulse aerodynamic response is theoretically infinite,

however, its time domain computational simulation restricts the frequency response to a finite

range of values (Ballhaus and Goorjian, 1978). The suggested approach is to compute a set of

weighting functions, based on unsteady pressure differences at each of the reduced frequency

identified by the Fourier transformation applied to the time domain aerodynamic impulse

response. Therefore, the correction of the linear aerodynamic model is performed based on the

same algorithm employed for the correction based on unsteady pressure differences at a single

reduced frequency. The difference is that, now each weighting function is applied to correct

the corresponding pressure to downwash relation resulting from the linear unsteady

aerodynamic modeling for the same reduced frequency.

The disadvantage of approaches based on unsteady reference pressures is the

computational cost, since it depends on time domain CFD computation to simulate the

nonlinear unsteady aerodynamic loading. The costs are prohibitive, when the aerodynamic

configuration and governing equations are more complex.

Another possibility to minimize the computational cost is to perform the downwash

correction based on nonlinear steady pressure differences. However, in this case the

procedure, as it is defined, is not capable to take into account properly unsteady transonic

effects, since they are not embedded in the steady state solutions of the nonlinear flow.
Chapter 4 – Correction Methods for Approximate Nonlinear Aerodynamic Modeling 94

Furthermore, the use of a single frequency dependent condition for the correction of the

aerodynamic model for different reduced frequency values is inconsistent. This fact can be

better understood considering the splitting of the aerodynamic influence coefficients matrix,

denoted as [AIC], in real and imaginary parts as [ AICR ] and [ AICI ] , respectively. The

computation of the weighting function is based on pressure differences computed at a reduced

frequency kr , where for the steady state, this value is zero. Then, the resulting pressure to

downwash relationship, is rewritten as:

{∆C nl
p (ik ) } = [ AIC R (ik ) + AICI (ik ) ][WT (ikr ) ]{hx + ikh} =
= [ AICR (ik ) ][WT (ik r ) ]{hx } + ik [ AICR (ik ) ][WT (ik r ) ]{h} + (4.22)
+ [ AICI (ik ) ][WT (ikr ) ]{hx } + ik [ AICI (ik ) ][WT (ikr ) ]{h}

One should note that none of the parcels of the relation above present a physical meaning,

since they are arbitrary, unless for the case when the reduced frequency associated to the

reference unsteady pressures is the same of the pressure to downwash equation, that is,

kr = k . Thus, a correct pressure distribution recovery is possible only by the following

relationships:

{∆C p (ik = 0)} = [ AIC (ik = 0)][WT (ik = 0)]{h }


nl
x
, (4.23)
{∆C p (ik )} = [ AIC (ik )][WT (ik )]{h + ik h}
nl
r r r x r

where it is considered the corresponding frequency dependent aerodynamic influence

coefficients matrices relating the downwash to the steady and unsteady nonlinear pressure

differences, respectively.

The above discussion suggests that the imaginary part of the pressure differences

cannot be obtained from the given steady pressure differences. A suggested approach is to

obtain the unsteady counterpart of the reference pressures, for example based on the subsonic

equation for the linearized unsteady potential flow. That is, it shall be necessary to search a

way to correct the linear unsteady aerodynamic model rewriting the downwash correction in
Chapter 4 – Correction Methods for Approximate Nonlinear Aerodynamic Modeling 95

such way that, the nonlinear steady component contributes to the correction of nonlinear

unsteady pressure differences.

4.2.2 A Modified Downwash Correction Method

Linear unsteady aerodynamic theories, applied to the aeroelastic modeling and

analysis, are developed considering small disturbances around mean angle of attack variations

of the lifting surface. In chapter 3, a linear/nonlinear behavior investigation was presented,

indicating that unsteady transonic flow behavior is strongly dependent on the amplitude of the

motion. However, aeroelastic deformations are usually smaller than the computed linear

limits, in terms of amplitudes of the motion, which were identified in chapter 3. Based on this

observation, it is inferred that, for aeroelastic analysis in a small disturbance context,

unsteadiness of transonic flows present a linear behavior with regard to aerodynamic

derivatives and shock dynamics, when the amplitude of lifting surfaces undergoing unsteady

motion are below the linear limits.

In other words, the unsteadiness of transonic flow may be understood as the

superposition of a linear contribution to a mean steady state nonlinear pressure distribution. A

linear unsteady pressure small disturbance contribution, for example, can be computed as a

contribution predicted by a small disturbance linear aerodynamic model. The same

conclusions were indicated by Dowell et al. (1983), where they pointed out that the unsteady

pressure disturbances behavior, is linear with respect to the aeroelastic deformations around

the mean steady state pressure distribution, at sub-transonic flow conditions.

From these observations, it is possible to assume the hypothesis that an unsteady

contribution to a steady nonlinear mean flow, in small disturbance transonic flows, can be

represented by a subsonic linear model. Thus, a correction procedure based on steady mean
Chapter 4 – Correction Methods for Approximate Nonlinear Aerodynamic Modeling 96

flow pressure distribution as reference condition, added by an unsteady contribution predicted

by the linearized potential flow equation, has a great chance to present good results regarding

unsteady aerodynamic loading prediction for aeroelastic stability analysis.

The extension of the standard downwash correction method is performed considering

the hypothesis mentioned above. The procedure is divided in two steps, the first being a

nonlinear steady mean flow correction as is performed for the correction of the steady

downwash, when nonlinear pressure differences are considered as reference conditions. The

second step is the correction of the unsteady downwash, where the unsteady counterpart of the

reference pressures to be added to the steady nonlinear reference pressures will compose new

reference pressure differences. These unsteady pressure contributions are predicted by a linear

unsteady flow aerodynamic model. Recalling that the downwash is related to the pressure

differences by an integral equation, which is written in a general form as:

1
w ( x, y, 0, ik ) = − ∆Cp (ξ , η ) K ψ ( x − ξ , y − η , 0, ik , M ) d ξ d η ,
8π ∫∫
(4.24)
A

it may also be written in an algebraic form as:

{w ( ik )} = [ D(ik )]{∆C ( ik )}
p . (4.25)

The kernel function integral relations between the sending and receiving points are

represented by the matrix operator [ D(ik ) ] , leading to a resulting equation which relates the

downwash to the pressure difference.

The matrix [ D(ik ) ] , hereafter named as the kernel function matrix, is computed for

subsonic flows and is a function of the reduced frequency. Therefore, it is assumed, as a

second hypothesis that, it is possible to expand the subsonic kernel function matrix as an

asymptotic series around small reduced frequencies. This kernel is a function of the reduced

frequency, and the domain of dependency is continuos and analytic at this circumstances.

Consequently, (Miles, 1950) the function may be expanded in an asymptotic series around
Chapter 4 – Correction Methods for Approximate Nonlinear Aerodynamic Modeling 97

reduced frequency values smaller than 1.0. Therefore it is possible to express all dependent

variables in the equations in terms of the reduced frequency as the expansion parameter (Liu,

1977; Hui, 1969). The resulting expanded kernel matrix is given by:

[ D(ik )] ≅ ( ik ) [ D ]0 + ( ik ) [ D ]1 + ( ik ) [ D ]2 + + ( ik ) [ D ]n
0 1 2 n

n , (4.26)
= [ D ]0 + ∑ ( ik ) [ D ]m
m

m =1

where k < 1.0 , and [ D ]n , n = 1, m , are constant coefficient matrices, obtained from the

asymptotic series expansion.

The transonic flutter reduced frequency is normally less than 1.0. This assumption is

justified based on Ashley (1980) and Tijdemann (1977) observations. Both authors pointed

out that flutter instability usually results from the coalescence of low frequency aeroelastic

modes, such as the coupling of a first structural bending and a first structural torsion mode.

Since under those conditions, the flutter flow speed is high and the associated frequency is

low, the resulting flutter reduced frequency is low.

The first term of the expansion series (4.26) is the steady state kernel function matrix

([ D ] = [ D (ik = 0)]) , since this term is the unique matrix which is independent of the reduced
0

frequency. The purpose of the kernel expansion is to allow the introduction of nonlinear

steady state reference conditions in place of the original steady state kernel function. This may

be performed by the replacement of the original kernel by a non linear steady state kernel

matrix in Eq. (4.26), which was computed from the downwash weighting based on the

nonlinear steady pressure differences using the following equation:

{∆C pnl (ik = 0)} = [ AIC (ik = 0)][WT (ik = 0)]{hx } (4.27)

One should observe that the downwash correction may be understood either as a left

multiplication of the downwash, or a right multiplication of aerodynamic influence

coefficients matrix. The right multiplication results in a aerodynamic influence coefficients


Chapter 4 – Correction Methods for Approximate Nonlinear Aerodynamic Modeling 98

matrix, in such way that its multiplication by the downwash results in steady nonlinear

pressure difference distribution. Thus, the downwash for a known steady displacement of the

lifting surface is given by:

−1
{hx } = {w(ik = 0)} =  AIC nl (ik = 0) {∆C pnl (ik = 0)}
(4.28)
=  D nl (ik = 0) {∆C pnl (ik = 0)}

−1
where  AIC nl (ik = 0) is a modified steady state aerodynamic influence coefficients
 

matrix, corrected by nonlinear pressure differences distribution. The inverse of this matrix is

denoted as  D nl (ik = 0) corresponding to a modified kernel function matrix, computed from

the weighting procedure based on the nonlinear steady pressure differences distribution.

Therefore, the resulting kernel expansion modified by the replacement of the steady

nonlinear kernel is given by:

 D nl (ik )  ≅  D nl ( ik = 0 )  + ( ik ) [ D ]1 + ( ik ) [ D ]2 + + ( ik ) [ D ]n
1 2 n

n . (4.29)
=  D nl ( ik = 0 )  + ∑ ( ik ) [ D ]m
m

m =1

The remaining term of the expansion is a function of the reduced frequency and may be

written as:

∑ ( ik ) [ D ] ≅ [ D(ik )] − [ D(ik = 0)]


m
m
(4.30)
m =1

After replacing Eq. (4.30) into the right hand side of Eq. (4.29), a modified kernel function

matrix, which takes into account the mean flow nonlinear contribution, is given by:

 D nl (ik )  =  D(ik ) − D(ik = 0) + D nl (ik = 0)  (4.31)

The inversion of  D nl (ik )  results in a modified aerodynamic influence coefficients matrix

 AIC nl (ik )  . Then, the nonlinear unsteady pressure difference is given by:

{∆C ( ik )} =  AIC
nl
p
nl
(ik )  {w(ik )} (4.32)
Chapter 4 – Correction Methods for Approximate Nonlinear Aerodynamic Modeling 99

Once the nonlinear unsteady pressures are established, it is possible to obtain a

frequency dependent correction matrix satisfying the above relation, written in downwash

correction form as:

{∆C ( ik )} = [ AIC (ik )] WT ( ik ) {w(ik )}


nl
p (4.33)

for a given reduced frequency of interest. Then the downwash correcting matrix is obtained

from:

 D (ik ) {∆C pnl (ik )} = WT (ik ) {w(ik )} . (4.34)


   

One should note that that the left hand side represents the modified downwash related to the

unsteady non-linear pressure differences. Therefore, the weighting function will be a diagonal

matrix whose elements are obtained from the relation between both computed downwash :

WT (ik ) =
{wnl (ik )}i
  ii . (4.35)
{w(ik )}i

WT (ik ) is a complex matrix and carries the information about the nonlinear unsteady
 

pressure difference contribution. It can be used to right multiply the aerodynamic influence

coefficients matrices for the unsteady aerodynamic loading computation as:

{L nl
a (ik )} = q∞ [ S ]  AIC (ik ) [WT (ik )]{w (ik )} (4.36)

This extension of the downwash weighting method, based on the steady kernel matrix

replacement inside the series expansion of the kernel function will be denominated hereafter

as the successive kernel expansion method (SKEM). This procedure can be applied to any

discrete element kernel function method. However, considering the hypothesis

aforementioned in the text, i.e., its application is restricted to the computation of nonlinear

flow in sub-transonic situations. The extension of such procedure for the full transonic flow

range, that is, for M ∞ ≥ 1.0 , should be regarded as a future development to be performed.
Chapter 4 – Correction Methods for Approximate Nonlinear Aerodynamic Modeling 100

The choice of the denomination of the proposed procedure as successive refers to its

applicability for aeroelastic stability analysis. In such situation, the method is applied in a

successive manner to correct each of the linear aerodynamic influence coefficients defined for

a reduced frequency domain set, defined for the computation of the flutter boundaries as a

function of a given parameter of interest. The investigation to be performed about the

efficiency of such method will be presented in the results chapter both for unsteady pressures

computation and flutter calculations.


5 – RESULTS AND DISCUSSION

The main concern of this Chapter is to present results from the application of

correction methods which were presented in Chapter 4. However, the first results to be

presented herein are obtained from the force matching method of Giesing et al. (1976). The

objective of this first investigation is to analyze the performance of such method, when

handling transonic aeroelastic stability analysis. The description of Giesing’s method is

presented in Appendix A.

Downwash correction methods based on steady and unsteady data (Sub-sect. 4.1.1 and

4.1.2, respectively) will be applied for aeroelastic stability analysis, in transonic flow. For

better understanding of the results with regard to these correction approaches, the unsteady

flow behavior over the prospective test case is investigated, in order to correlate unsteady

pressure distributions with the flutter boundaries calculated by such procedures.

Furthermore, the successive kernel expansion method (SKEM), which development is

presented in the Sect. 4.2 of Chapter 4, will be validated for unsteady pressure computations

and stability aeroelastic analysis. Comparisons with experimental data will be performed to

show the performance of such method with regard to the prediction of unsteady transonic

flow and flutter computation.

5.1 - Evaluation of Giesing´s Method for Transonic Flutter Computation

This section presents an evaluation of Giesing’s method (Giesing et al., 1976) with

respect to its performance in transonic aeroelastic stability analysis. This is a force matching

method, where the computation of the weighting matrix is based on strip loads. Each strip is

composed by a set of panels along the streamwise direction. As the reference loads are taken
Chapter 5 – Results and Discussion 102

for each strip, while correction factors are to be calculated for each panel, it is clear that there

are fewer available data (reference loads) than unknowns (correction factors). Therefore the

correction factors need to be computed by a minimization technique, which is described in

Appendix A, where the Giesing’s method is briefly descripted.

Giesing´s method was tested for two standard aeroelastic configurations, the AGARD

wing 445.6 weakened model #3 (Yates, 1988) and the PAPA wing (Pitch And Plunge

Apparatus, Farmer and Rivera., 1988). The first test case under investigation, is well known

as a standard aeroelastic configuration. It has an aspect ratio of 4 and a NACA 65A004 airfoil

section.

Figure 5.1 : Sketch of the AGARD 445.6 wing and the corresponding discrete element

aerodynamic model.

The AGARD 445.6 wing is modeled by ZONA 6 method, implemented in the ZAERO

software system (Zona Technology, 2003), as an isolated wing. The flow conditions which

were used for analysis belong to a subset of those presented in the work of Yates (1988),

which consists in an experimental investigation of the aeroelastic behavior of this wing under

transonic flow conditions. The model under investigation is described in Yates (1988),
Chapter 5 – Results and Discussion 103

regarding its structural and geometrical characteristics. The Mach numbers and air densities

for the cases considered here are presented in Table 5.1.

Table 5.1: Flow conditions for AGARD wing 445.6 (weakened # 3) aeroelastic

analysis.

Mach Reynolds Density (kg/m3) Density (slug/ft3)


0.678 1,410×106 0.2082 0,000404
0.901 0,911×106 0.0995 0,000193
0.960 0,627×106 0.0634 0,000123

The discrete element aerodynamic model, based on the ZONA 6 method, is comprised

of 220 panels, uniformly distributed along the chord length. The flutter computations were

performed using the g-method. Since the objective is to identify flutter boundaries, the use of

the k, pk or g methods is indifferent, because for that purpose, they give the same results. The

advantage of the g-method, in comparison with the others, is a better prediction of the sub-

critical aeroelastic response, since the predicted damping is more realistic.

The stripwise lifting surface loading was computed based on the integration of

nonlinear steady pressure distributions, computed by the CFL3D Navier-Stokes

computational fluid dynamic solver (Thomas et al., 1990, and Rumsey et al., 1997), The

nonlinear flow simulations were performed at two distinct angles of attack of 0.0o and 0.5o.

Quasi-steady pressure disturbances are computed and scaled by the amplitude of the motion,

turning them independent of the amplitude of the disturbances, if one considers that the flow

behaves linearly in this range. Once these loads are obtained, the procedure allows the

computation of a force matching weighting matrix which restores exactly the nonlinear

reference loading condition. However, the pressures are not restored since the force matching

is a least squares based procedure, that is, there are less unknowns, which are the correction

factors, than equations, since the weighting matrix needs to have the same dimension of the
Chapter 5 – Results and Discussion 104

AIC matrix. Then, this weighting matrix does not guarantee a correct pressure distribution, as

it only assures the load matching. In Figure (5.2) the complex pressure distribution, after the

multiplication of the weighting matrix from Giesing´s method, is compared with the nonlinear

pressures obtained by the Navier-Stokes solution, and with the ones obtained from ZONA 6

linear solution.

14.0 0.5

12.0 Giesing's Method 0


10.0 Navier-Stokes 0 0.2 0.4 0.6 0.8 1
-0.5

Imag (∆ Cp)
Real ( ∆ Cp)

ZONA 6 (linear)
8.0
-1
6.0
-1.5 Giesing's Method
4.0

-2
Navier-Stokes
2.0
ZONA 6 (linear)
0.0 -2.5
0 0.2 0.4 0.6 0.8 1
X/C X/C

Figure 5.2 : Comparison between the unsteady pressures obtained from the Giesing´s method,

ZONA 6 and CFL3D - AGARD 445.6 wing, weakened model #3, M∞ =0.96.

The results from the flutter solution using Giesing´s correction procedure are

presented in Figure (5.3). In the same figure are presented results for the downwash weighting

method (DWM) based on the same reference flow conditions used for Giesing’s method, i.e.

for the same quasi-steady pressure disturbances. One should observe that the flutter stability

margin is over-predicted in terms of the flutter dynamic pressure. This is an indication that

there is a failure in such procedure when handling transonic effects, which may be justified by

the anomalous shock wave dynamics behavior observed in Fig. (5.2).


Chapter 5 – Results and Discussion 105

140.0 24.0

130.0 AGARD 445.6 ∆α = 0.5 AGARD 445.6 ∆α = 0.5


22.0
120.0

Frequency (Hz)
Dyn. Press.

20.0
110.0

100.0 18.0

90.0 DWM
Giesing
16.0 DWM
80.0 Giesing
ZONA 6 14.0
70.0 ZONA 6
experiment
experiment
60.0 12.0
0.6 0.7 0.8 0.9 1 0.6 0.7 0.8 0.9 1
Mach Mach

Figure 5.3 : Flutter boundaries (dynamic pressure and flutter frequency) for the AGARD wing

445.6, weakened model #3.

The other test case is the flutter computation of the PAPA wing (Farmer and Rivera,

1988) at two distinct reference mean angles of attack. It has a cambered supercritical airfoil

with maximum thickness of 12% and a rectangular planform with chord of 16 inches (0.4064

m) and semispan of 32 inches (0.8128 m). The structural support provides two rigid body

modes: a plunge mode (3.43 Hz) and a pitch mode (5.44 Hz). The ZONA 6 aerodynamic

model is comprised of 200 panels, equally distributed along the chord length (Fig. 5.4).

Figure 5.4 : Sketch of the PAPA wing and the corresponding discrete element aerodynamic

model.
Chapter 5 – Results and Discussion 106

The flutter of this wing results from the coalescence of these two modes, depending on

the mean angle of attack of the wing, and from the undisturbed flow conditions. The flutter

boundaries, for a few Mach numbers near the transonic flow regime are presented in Figures

(5.5) and (5.6).

180.0 4.7
papa +1 deg papa +1 deg
4.65
175.0
4.6
170.0
Dyn. Press.

DWM 4.55

Freq (Hz)
Giesing
165.0 ZONA 6 4.5
experimental
4.45
160.0
DWM
4.4
Giesing
155.0
4.35 ZONA 6
experimental
150.0 4.3
0.6 0.65 0.7 0.75 0.8 0.85 0.6 0.65 0.7 0.75 0.8 0.85
Mach Mach

Figure 5.5 : Flutter boundaries (dynamic pressure and flutter frequency) for the PAPA wing at

a reference angle of attack of α = 1.0o.

230.0 4.65
DWM DWM
Giesing 4.6 Giesing
210.0 ZONA 6 ZONA 6
4.55
experimental
Dyn. Press.

experimental
Freq (Hz)

4.5
190.0
papa -2 deg 4.45

4.4
170.0
4.35 papa -2 deg

150.0 4.3
0.65 0.7 0.75 0.8 0.85 0.65 0.7 0.75 0.8 0.85
Mach Mach

Figure 5.6 : Flutter boundaries (dynamic pressure and flutter frequency) for the PAPA wing,

at a reference angle of attack of α =-2.0o.


Chapter 5 – Results and Discussion 107

These results are in better agreement with the experimental ones when comparing the

dynamic pressure in the case of the PAPA wing at α=1.0o . One should note that the results of

the steady downwash weighting method and from Giesing´s method are nearly coincident for

both the dynamic pressure and flutter frequency. However the flutter frequencies still remain

underestimated, when compared to the experimental measurements. At higher Mach numbers,

Giesing’s method results are worse in the case of the PAPA wing at α=-2.0o, when compared

to the results at the same Mach number and α = 1.0o. Such behavior is expected, because

transonic effects at higher angle of attack amplitudes are stronger than at lower amplitudes,

and this method does not guarantee that shock positioning and strength be restored, thus

giving worse results for flow conditions with stronger transonic effects. Certainly this method

will fail on the prediction of transonic flutter speeds, turning it inadequate for such

computations.

5.2 –Downwash Weighting Methods

The results to be presented in this section are obtained using downwash correction

methods, based on steady and unsteady pressures as the nonlinear reference conditions. The

test case to be investigated is the AGARD wing 445.6 weakened model #3 (Yates, 1988), in

subsonic to transonic flow conditions. As a first step, the unsteady pressure distribution over

the AGARD 445.6 wing are investigated in order to understand the fluid dynamic behavior

and its relation with the results with regard to the flutter computation for this wing.
Chapter 5 – Results and Discussion 108

5.2.1 – Investigation of the AGARD Wing Unsteady Pressure Distributions.

The test case under investigation is the AGARD wing 445.6, weakened model # 3

standard aeroelastic configuration. It has an aspect ratio of 4.0 and a NACA 65A004 airfoil

section. This wing is modeled by the doublet lattice method, implemented in

MSC/NASTRANTM software system, as an isolated wing in different flow conditions, where

the discrete element model is comprised of 240 panels (Fig. 5.7).

Figure 5.7: DLM paneling of the AGARD 445.6 wing.

The test conditions belong to a subset of those presented in the work of Yates (1988),

which consists in an experimental investigation of the aeroelastic behavior of this wing under

transonic flow conditions. The model under investigation is described in Yates (1988),

regarding its structural and geometrical characteristics. The Mach numbers and air densities

for the cases considered here are presented in Table (5.1).


Chapter 5 – Results and Discussion 109

The computation of the nonlinear unsteady pressures is performed from the finite

differences solution of the Navier-Stokes equations, using the method described in Chapter

3. The computational mesh surrounding the wing is an algebraic generated “C” type topology,

with 141 points in the ξ direction, tangent to the lifting surface boundary, where 121 points

are over the lifting surface solid surface. In the η direction, normal to the wing surface, there

are 41 points between the solid surface and the limit of the computational mesh. And finally,

in the ζ direction there are 25 points aligned with the spanwise direction, where 17 points are

over the lifting surface, and the remaining are between the wing tip and the computational

domain limit. The computational mesh which surrounds the wing, as well as its profile at

spanwise station 30.8 % are presented in Fig. (5.8) and (5.9) respectively.

1.1

0.9

0.8

0.7

0.6

0.5

0.4

0.3

0.2

0.1

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1 1.2

Figure 5.8 : AGARD wing CFD aerodynamic mesh.


Chapter 5 – Results and Discussion 110

The investigation of the unsteady transonic flow behavior around the AGARD wing

445.6 is here performed examining pressure distributions at spanwise station 30.8 % Fig (5.9).

The main concern of this investigation is to verify the influence of the amplitude of the

motion on pressure distributions. Thus, this station was chosen because the flow is less

subjected to three-dimensional effects as indicated by Silva et al. (2001).

Three reduced frequencies are chosen from the test conditions presented by Yates

(1988). These frequencies are presented in Table 5.2, where, k denotes the reduced frequency

k = ωb U ∞ , while kr = ω c a∞ is another way of representing the reduced frequency, used in

the Navier-Stokes code. For both linear and nonlinear computations, c= 2b = 1.0 (m).

Table 5.2: Reduced frequencies for AGARD wing 445.6 aeroelastic analysis.

Mach kr k
0,678 0.48839 0.36016
0,901 0.34076 0.18910
0,960 0.28252 0.14714

0.4

0.3

0.2

0.1

-0.1

-0.2

-0.3

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7

Figure 5.9 : AGARD wing 445.6 NACA 65A004 profile, spanwise station 30.8 %.
Chapter 5 – Results and Discussion 111

Pressure distributions are computed from the doublet lattice method and the unsteady

Navier-Stokes simulation for the reduced frequencies and corresponding Mach numbers,

presented in Table (5.2). In the case of the nonlinear solution, they were obtained for different

amplitudes of oscillation, leading to pressure coefficient differences properly scaled by the

magnitude of the disturbance. The computations were performed for different amplitudes in

order to quantify the influence of the amplitude of the motion on the pressure ratios, and to

compare these to the ones computed from the linear model.

The computed unsteady pressure distributions along the chord at spanwise station 30,8

%, are presented in Figs. (5.10) through (5.30). Both real and imaginary parts of these

pressures are presented at a given Mach number (0.678, 0.901 and 0.960), the corresponding

reduced frequency k given in Table (5.2), and amplitude of oscillation. The chosen amplitudes

were between 0.25o to 3.0o in angle of attack rotation around the root midchord axis.

0.8 0.02
0.7 Doublet Lattice 0.01
0
0.6 Navier-Stokes
-0.01 0 0.2 0.4 0.6 0.8 1
Real (∆ Cp)

0.5
Imag (∆ Cp)

-0.02
0.4
-0.03
0.3 -0.04
0.2 -0.05 Doublet Lattice
0.1 -0.06 Navier-Stokes
0 -0.07
0 0.2 0.4 0.6 0.8 1 -0.08
-0.09
x/c x/c

Figure 5.10 : Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞=0.678, ∆α = 0.25o.
Chapter 5 – Results and Discussion 112

1.6 0.04
1.4 0.02
Doublet Lattice 0
1.2
-0.02 0 0.2 0.4 0.6 0.8 1
Real (∆ Cp)

1 Navier-Stokes

Imag (∆ Cp)
-0.04
0.8 -0.06
0.6 -0.08
Doublet Lattice
0.4 -0.1
-0.12 Navier-Stokes
0.2
-0.14
0
-0.16
0 0.2 0.4 0.6 0.8 1
-0.18
x/c x/c

Figure 5.11: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞=0.678, ∆α = 0.5o

3 0.1
Doublet Lattice 0.05
2.5
Navier-Stokes 0
Real (∆ Cp)

2
-0.05 0 0.2 0.4 0.6 0.8 1
Imag (∆ Cp)

1.5 -0.1

1 -0.15
-0.2
0.5 Doublet Lattice
-0.25
0
Navier-Stokes
-0.3
0 0.2 0.4 0.6 0.8 1
-0.35
x/c x/c

Figure 5.12: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞=0.678, ∆α = 1.0o
Chapter 5 – Results and Discussion 113

4.5 0.2
4 Doublet Lattice
0.1
3.5 Navier-Stokes
0
Real (∆ Cp)

3
0 0.2 0.4 0.6 0.8 1

Imag (∆ Cp)
2.5 -0.1
2
-0.2
1.5
1 -0.3
0.5 -0.4
0 Doublet Lattice
-0.5
0 0.2 0.4 0.6 0.8 1 Navier-Stokes
-0.6
x/c x/c

Figure 5.13: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞=0.678, ∆α = 1.5o.

6 0.2

5 0.1
Doublet Lattice
0
Real (∆ Cp)

4
Navier-Stokes 0 0.2 0.4 0.6 0.8 1
Imag (∆ Cp)

-0.1
3
-0.2
2
-0.3 Doublet Lattice
1
-0.4 Navier-Stokes
0 -0.5
0 0.2 0.4 0.6 0.8 1
-0.6
x/c x/c

Figure 5.14: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞=0.678, ∆α = 2.0o
Chapter 5 – Results and Discussion 114

8 0.2
7 Doublet Lattice 0.1
6 Navier-Stokes 0
Real (∆ Cp)

5 -0.1 0 0.2 0.4 0.6 0.8 1

Imag(∆ Cp)
4 -0.2
-0.3
3
-0.4 Doublet Lattice
2
-0.5 Navier-Stokes
1
-0.6
0
-0.7
0 0.2 0.4 0.6 0.8 1
-0.8
x/c x/c

Figure 5.15: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞=0.678, ∆α = 2.5o

9 0.4
8
Doublet Lattice 0.2
7
Real (∆ Cp)

6 Navier-Stokes 0
Imag (∆ Cp)

5 0 0.2 0.4 0.6 0.8 1


-0.2
4
3 -0.4
2 Doublet Lattice
1 -0.6
Navier-Stokes
0 -0.8
0 0.2 0.4 0.6 0.8 1
-1
x/c x/c

Figure 5.16: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞=0.678, ∆α = 3.0o
Chapter 5 – Results and Discussion 115

0.8 0.02
0.7 Doublet Lattice
0
0.6 Navier-Stokes 0 0.2 0.4 0.6 0.8 1
Real (∆ Cp)

-0.02
0.5

Imag (∆ Cp)
0.4 -0.04
0.3
-0.06
0.2
-0.08
0.1
Doublet Lattice
0 -0.1
0 0.2 0.4 0.6 0.8 1 Navier-Stokes
-0.12
x/c x/c

Figure 5.17: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞=0.901, ∆α = 0.25o

1.6 0.05
1.4 Doublet Lattice
0
1.2 Navier-Stokes
0 0.2 0.4 0.6 0.8 1
Real (∆ Cp)

1 -0.05
Imag (∆ Cp)

0.8
-0.1
0.6
0.4 -0.15
0.2
-0.2 Doublet Lattice
0
0 0.2 0.4 0.6 0.8 1 Navier-Stokes
-0.25
x/c x/c

Figure 5.18: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞=0.901, ∆α = 0.5o
Chapter 5 – Results and Discussion 116

3 0.1
0.05
2.5
Doublet Lattice 0
-0.05 0 0.2 0.4 0.6 0.8 1
Real (∆ Cp)

2 Navier-Stokes

Imag (∆ Cp)
-0.1
1.5 -0.15
1 -0.2
Doublet Lattice
-0.25
0.5
Navier-Stokes
-0.3
-0.35
0
-0.4
0 0.2 0.4 0.6 0.8 1
-0.45
x/c x/c

Figure 5.19: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞=0.901, ∆α = 1.0o

4.5 0.2
4 0.1
3.5 0
Doublet Lattice
Real (∆ Cp)

3
Navier-Stokes -0.1 0 0.2 0.4 0.6 0.8 1
Imag (∆ Cp)

2.5
-0.2
2
-0.3
1.5
1 -0.4 Doublet Lattice
0.5 -0.5 Navier-Stokes
0 -0.6
0 0.2 0.4 0.6 0.8 1 -0.7
x/c x/c

Figure 5.20: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞= 0.901, ∆α = 1.5o


Chapter 5 – Results and Discussion 117

6 0.2

5
Doublet Lattice
0
Navier-Stokes 0 0.2 0.4 0.6 0.8 1
Real (∆ Cp)

4
-0.2

Imag (∆ Cp)
3
-0.4
2
-0.6 Doublet Lattice
1 Navier-Stokes
-0.8
0
0 0.2 0.4 0.6 0.8 1 -1
x/c x/c

Figure 5.21: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞= 0.901, ∆α = 2.0o

8 0.2
7 Doublet Lattice 0
6 Navier-Stokes 0 0.2 0.4 0.6 0.8 1
Real (∆ Cp)

-0.2
5
Imag(∆ Cp)

4 -0.4
3
-0.6
2 Doublet Lattice
1 -0.8 Navier-Stokes
0 -1
0 0.2 0.4 0.6 0.8 1
-1.2
x/c x/c

Figure 5.22: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞=0.901, ∆α = 2.5o
Chapter 5 – Results and Discussion 118

9 0.4
8 Doublet Lattice 0.2
7 0
Navier-Stokes
Real (∆ Cp)

6
-0.2 0 0.2 0.4 0.6 0.8 1

Imag (∆ Cp)
5
-0.4
4
3 -0.6

2 -0.8 Doublet Lattice


1 -1 Navier-Stokes
0 -1.2
0 0.2 0.4 0.6 0.8 1
-1.4
x/c x/c

Figure 5.23: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞=0.901, ∆α = 3.0o

0.8 0.02
0.7 Doublet Lattice
0
0.6 Navier-Stokes 0 0.2 0.4 0.6 0.8 1
Real (∆ Cp)

0.5 -0.02
Imag (∆ Cp)

0.4
-0.04
0.3
Doublet Lattice
0.2 -0.06
Navier-Stokes
0.1
-0.08
0
0 0.2 0.4 0.6 0.8 1 -0.1

x/c x/c

Figure 5.24: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞= 0.96, ∆α = 0.25o


Chapter 5 – Results and Discussion 119

1.6 0.05
1.4 Doublet Lattice
0
1.2 Navier-Stokes
0 0.2 0.4 0.6 0.8 1
Real (∆ Cp)

Imag (∆ Cp)
-0.05
0.8
0.6
-0.1
0.4
Doublet Lattice
0.2 -0.15 Navier-Stokes
0
0 0.2 0.4 0.6 0.8 1 -0.2
x/c x/c

Figure 5.25: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞=0.96, ∆α = 0.5o

3 0.1

Doublet Lattice 0.05


2.5
0
Navier-Stokes
Real (∆ Cp)

2 -0.05 0 0.2 0.4 0.6 0.8 1


Imag (∆ Cp)

-0.1
1.5
-0.15
1 -0.2

0.5 -0.25
-0.3 Doublet Lattice
0 -0.35
0 0.2 0.4 0.6 0.8 1
Navier-Stokes
-0.4
x/c x/c

Figure 5.26: Unsteady pressure distributions for the wing station at 30.8% of the span, M∞=

0.96, ∆α = 0.25o
Chapter 5 – Results and Discussion 120

4.5 0.1
4
Doublet Lattice 0
3.5
0 0.2 0.4 0.6 0.8 1
Navier-Stokes -0.1
Real (∆ Cp)

Imag (∆ Cp)
2.5 -0.2
2
1.5 -0.3

1 -0.4 Doublet Lattice


0.5 Navier-Stokes
-0.5
0
0 0.2 0.4 0.6 0.8 1
-0.6
x/c x/c

Figure 5.27: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞= 0.96, ∆α = 1.5o

6 0.2
0.1
5 Doublet Lattice
0
Navier-Stokes
Real (∆ Cp)

4 -0.1 0 0.2 0.4 0.6 0.8 1


Imag (∆ Cp)

-0.2
3
-0.3
2 -0.4

1 -0.5 Doublet Lattice


-0.6 Navier-Stokes
0 -0.7
0 0.2 0.4 0.6 0.8 1
-0.8
x/c x/c

Figure 5.28: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞= 0.96, ∆α = 2.0o


Chapter 5 – Results and Discussion 121

8 0.2
7 Doublet Lattice 0
6 Navier-Stokes 0 0.2 0.4 0.6 0.8 1
Real (∆ Cp)

5 -0.2

Imag (∆ Cp)
4
-0.4
3
2 -0.6 Doublet Lattice
1 Navier-Stokes
-0.8
0
0 0.2 0.4 0.6 0.8 1
-1
x/c x/c

Figure 5.29: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞= 0.96, ∆α = 2.5o

9 0.2
8
0
7 Doublet Lattice
0 0.2 0.4 0.6 0.8 1
Navier-Stokes -0.2
Real (∆ Cp)

6
Imag (∆ Cp)

5 -0.4
4
-0.6 Doublet Lattice
3
2 -0.8 Navier-Stokes
1
-1
0
0 0.2 0.4 0.6 0.8 1 -1.2

x/c x/c

Figure 5.30: Unsteady pressure distributions for the wing station at 30.8% of the span,

M∞= 0.96, ∆α = 3.0o.


Chapter 5 – Results and Discussion 122

The results presented in Figs. (5.10) through (5.16) are the complex pressure

distributions for a subsonic (M∞ = 0.678) flow condition. One should note that there is a good

agreement between the results from the Navier-Stokes and the doublet lattice computations,

up to ∆α = 1.5o. For these low amplitudes of motion, there is a distinct leading edge suction

peak. However, examining the results presented in Figs (5.14) through (5.16), it is possible to

note that there is a deviation near the leading edge, for both real and imaginary parts of the

pressures. In order to further investigate this behavior, a contour plot of the pressures at the

maximum angle α = 2.0o is shown in Fig (5.31), while instantaneous pressures at several

angle of attack are plotted in Fig. (5.32) for ∆α = 2.0o. Similar plots for the motion amplitude

∆α= 3.0o and maximum angle α = 3.0o are shown in Figs. (5.34) and (5.33).

Figure 5.31 : Contour plot of instantaneous pressures at spanwise station 30,8%, M∞=0.678,
α = 2.0o.
Chapter 5 – Results and Discussion 123

1.80
α =2.0
1.60 α =1.8
α =1.6
1.40
α =1.4
1.20 α =1.2
1.00
α =1.0
α =0.8
Cp

0.80
α =0.6
0.60 α =0.4
α =0.2
0.40

0.20

0.00
0.0 0.2 0.4 0.6 0.8 1.0
-0.20 x/c

Figure 5.32 : Instantaneous pressure distributions at spanwise station 30,8%, M∞=0.678,

∆α = 2.0o.

Figure 5.33: Contour plot of instantaneous pressures at spanwise station 30,8%, M∞=0.678,
α = 3.0o.
Chapter 5 – Results and Discussion 124

1.80
α =3.0
1.60 α =2.7
α =2.4
1.40
α =2.1
1.20 α =1.8
1.00
α =1.5
α =1.2
Cp

0.80
α =0.9
0.60 α =0.6
α =0.3
0.40

0.20

0.00
0.0 0.2 0.4 0.6 0.8 1.0
-0.20 x/c

Figure 5.34: Instantaneous pressure fields and distributions at spanwise station 30,8%,
M∞=0.678, ∆α = 3.0o.

For both angles of attack, one should observe the appearance of a secondary suction

peak which moves towards the trailing edge with the increase in motion amplitude. The

displacement of this secondary pressure suction peak is associated to the aerodynamic lag of

the flow around the finite thickness profile, computed by the Navier-Stokes solution. The

surface pressure distribution plots in Figs (5.32) and (5.34), indicate that the pressure

distribution is quite influenced by the motion amplitude. For example, for the same

instantaneous angle of attack of 1.8o, it is possible to note the dependency of the aerodynamic

lag on the motion amplitude. For ∆α = 2.00 (Fig. 5.32), the position of the secondary pressure

peak is around x c ≅ 0.1 , while for ∆α = 3.00 (Fig. 5.34), the position of the secondary

pressure peak is x c ≅ 0.2 . Since these secondary pressure peaks are moving with time, and

its displacements increase with the amplitude of the motion, then the deviation from the linear
Chapter 5 – Results and Discussion 125

complex pressure distributions are larger for the higher angle of attack amplitudes (Figs 5.14

through 5.16).

The downstream displacement of the secondary pressure peak is more pronounced in

the subsonic Mach number because, under this condition, the aerodynamic lag effects are

more relevant. However, examining instantaneous pressure plots (Fig 5.36 and 5.38) and the

pressure contours at the maximum angle of attack (Fig 5.36 and 5.38) for M∞=0.901, it is

possible to note a behavior similar to the subsonic case,i.e., M∞=0.678, but with less intensity,

as expected.

Figure 5.35: Contour plot of instantaneous pressures at spanwise station 30,8%, M∞=0.901,

α = 2.0o.
Chapter 5 – Results and Discussion 126

1.80
α =2.0
1.60 α =1.8
α =1.6
1.40
α =1.4
1.20 α =1.2
1.00
α =1.0
α =0.8
Cp

0.80
α =0.6
0.60 α =0.4
α =0.2
0.40

0.20

0.00
0.0 0.2 0.4 0.6 0.8 1.0
-0.20 x/c

Figure 5.36: Instantaneous pressure distributions at spanwise station 30,8%, M∞=0.901,

∆α = 2.0o.

Figure 5.37: Contour plot of instantaneous pressures at spanwise station 30,8%, M∞=0.901,

α = 3.0o.
Chapter 5 – Results and Discussion 127

1.80
α =3.0
1.60 α =2.7
α =2.4
1.40
α =2.1
1.20 α =1.8
1.00
α =1.5
α =1.2
Cp

0.80
α =0.9
0.60 α =0.6
α =0.3
0.40

0.20

0.00
0.0 0.2 0.4 0.6 0.8 1.0
-0.20 x/c

Figure 5.38: Instantaneous pressure distributions at spanwise station 30,8%, M∞=0.901,

∆α = 3.0o.

The differences that have been observed between the doublet lattice method and the

Navier-Stokes results at M∞=0.96 (Figs. 5.24 through 5.30) are associated to a shock wave

formation. Contour plots of the instantaneous pressures at maximum angle of attack situations

are presented in Figs. (5.39) and (5.41) for this Mach number. In latter situation, there is a

slight trend for appearance of a secondary peak (Figs. 5.42), because the aerodynamic lag is

suppressed by the high undisturbed flow speed energy.


Chapter 5 – Results and Discussion 128

Figure 5.39: Contour plot of instantaneous pressures at spanwise station 30,8%, M∞=0.960,

α = 2.0o.

1.60
α =2.0
1.40 α =1.8
α =1.6
1.20
α =1.4
1.00
α =1.2
α =1.0
0.80 α =0.8
Cp

α =0.6
0.60
α =0.4
0.40 α =0.2

0.20

0.00
0.0 0.2 0.4 0.6 0.8 1.0

-0.20 x/c

Figure 5.40: Instantaneous pressure distributions at spanwise station 30,8%, M∞=0.960,

∆α = 2.0o.
Chapter 5 – Results and Discussion 129

Figure 5.41: Contour plot of instantaneous pressures at spanwise station 30,8%, M∞=0.960,

α = 3.0o.

1.80
α =3.0
1.60 α =2.7
α =2.4
1.40
α =2.1
1.20 α =1.8
1.00
α =1.5
α =1.2
Cp

0.80
α =0.9
0.60 α =0.6
α =0.3
0.40

0.20

0.00
0.0 0.2 0.4 0.6 0.8 1.0
-0.20 x/c

Figure 5.42: Instantaneous pressure distributions at spanwise station 30,8%, M∞=0.960,

∆α = 3.0o.
Chapter 5 – Results and Discussion 130

One should observe in Figs (5.40 and 5.42) that there is a displacement of the shock

wave towards the trailing edge, from x c ≅ 0.50 to x c ≅ 0.57 for the angles of attack

α = 2.0o and α = 3.0o respectively. The transonic effects are more pronounced for this Mach

number, and are especially evident in the imaginary part of the pressures, leading to strong

differences between the linear and nonlinear predicted phases. As the pressure phases are

more sensitive to the shock movement than the pressure amplitudes, the main differences

when comparing the potential and the Navier-Stokes results will be in the imaginary part of

the pressure. At M∞=0.901, the transonic effects are weaker than the aforementioned

situation, as one should observe in Figs. (5.17) through (5.23). This fact can be explained

examining the airfoil profile maximum thickness. The AGARD wing has a 65A004 airfoil

with a maximum thickness ratio (t/c) equal to 4.0 %. Hence, the shock wave formation will

only occur at higher Mach numbers.

Another feature to be noted is the influence of the amplitude of the motion on the

shock wave position. In Figs (5.27) through (5.30), one may notice that the shock appears to

move aft as the amplitude is increased. The reason for this behavior is that, in steady flow

conditions, the shock wave position moves towards the trailing edge as the angle of attack is

increased, as shown by Silva et al., (2002 A). This behavior is also noted for unsteady flows,

as indicated by Dowell et al. (1983). Recalling that the pressures are scaled by the amplitude

of the motion, one should note that, for smaller angles of attack, the shock position is about

the same. This is so because int the low angle of attack range the transonic flow behaves

linearly, with respect to angle of attack, as concluded in Chapter 3.

Another aspect to be highlighted is that the formation of the moving secondary suction

peak is apparently not associated to flow separations. This is illustrated by the instantaneous

velocity vector plots presented in Figs (5.43) for Mach numbers 0.678 to 0.960 and maximum

angles α = 2.00 and α = 3.00 .


Chapter 5 – Results and Discussion 131

M∞ = 0.678, α =2.0o M∞ = 0.678, α =3.0o

M∞ = 0.901, α =2.0o M∞ = 0.901, α =3.0o

M∞ = 0.960, α =2.0o M∞ = 0.960, α =3.0o

Figure 5.43: Instantaneous velocity vector fields over the leading edge of the AGARD 445.6

wing.
Chapter 5 – Results and Discussion 132

Examining the velocity vector plots in Fig (5.43), one should observe that for

∆α = 2.00 and ∆α = 3.00 , there is no evidence of flow separation. However, the mesh around

the leading edge is not sufficiently refined to capture mild separation. Neither it was

guaranteed that the maximum displacement positions are the ones where the dynamic angle of

attack derivative is null, because the time domain Navier-Stokes simulation is performed in

discrete time steps. On the other hand, since the identified unsteady flow behavior regards

aerodynamic lag associated to the wing motion, the aforementioned modeling restrictions do

no impact on the conclusions of the present investigation.

The behavior of the phase angles of the scaled complex pressure difference along the

chord for M∞=0.96 is summarized in Fig (5.44), where Θ is the non-dimensional phase angle

resulting from the scaling of the phase angle θ and the amplitude ∆α , as Θ = θ ∆α .

50

40 0.25 deg
0.5 deg
30 1.0 deg
1.5 deg
2.0 deg
Θ (deg)

20
2.5 deg
10 3.0 deg

0
0 0.2 0.4 0.6 0.8 1
-10

-20
x/c
Figure 5.44: Non-dimensional phase angles of the pressure difference distributions for

spanwise station 30.8%, M∞=0.96.


Chapter 5 – Results and Discussion 133

As may be observed in Fig. (5.44), for this specific case (i.e. M∞=0.96, 30,8%

spanwise station) pressure phases remain around –8o up to x c ≅ 0.55 . Aft of that location,

the pressures phase advances as the trailing edge is approached. The corresponding pressure

difference amplitudes, scaled by the amplitude of the motion are summarized in Fig. (5.45). It

may be noticed that the scaled pressure difference amplitude changes only slightly as the

angle of attack amplitude is increased.

2.5
0.25 deg
0.5 deg
2 1.0 deg
1.5 deg
|∆Cp|/Da

1.5 2.0 deg


2.5 deg
3.0 deg
1

0.5

0
0 0.2 0.4 0.6 0.8 1
x/c

Figure 5.45. Amplitudes of the pressure difference distributions for spanwise station 30.8%,

M∞=0.96.

The pressure distribution behavior summarized in the results of this section suggest

that the amplitude of the motion plays an important role in the computation of the complex

unsteady pressure distributions. The amplitude of the pressures gives the lift, while the

pressure phases may be understood as a form to quantify the lag between the center of

pressure displacement and the lifting surface motion. Since the center of pressure position is
Chapter 5 – Results and Discussion 134

governed by the shock displacement, as indicated by Ashley (1980) and Tijdeman (1977), the

moment coefficient derivatives will be more sensitive to the nonlinearities of the flow over

the lifting surface. Therefore, from these observations it is suggested that the influence of the

nonlinearities will promote significant changes in the computation of the aeroelastic stability,

since flutter depends on the relative position of the center of pressure with respect to the

elastic axis position. In the next sub-section an investigation of the aeroelastic stability

behavior will be performed to evaluate the influence of the amplitude of the angle of attack on

the computation of flutter speeds based on nonlinear unsteady pressures.

5.2.2 AGARD 445.6 Wing Aeroelastic Stability Analysis

The computation of the aeroelastic stability of the AGARD wing is performed in order

to evaluate the efficiency of the downwash correction method, using either nonlinear steady

or unsteady pressure distribution as reference conditions. The unsteady aerodynamic

modeling is based on the doublet lattice method, implemented in the MSC/NASTRANTM

software system. The chosen flutter solution technique is the k-method, which is

mathematically consistent for the computation of the flutter boundary. The weighting

operators are computed to correct the pressure to downwash relation, resulting from the

modeling of the AGARD wing using the doublet lattice method.

The nonlinear unsteady pressures were obtained under quasi-steady motion and

unsteady motions to provide the corresponding reference conditions for the steady and

unsteady downwash weighting methods, respectively. In the latter case, the pressures were

computed under harmonic motion oscillations of the wing with amplitude ∆α = 2.00 , where

the nonlinear contribution due to unsteady transonic effects are more relevant. Otherwise, in

the former case, the chosen amplitude of the quasi steady motion is ∆α = 0.50 because the
Chapter 5 – Results and Discussion 135

nonlinear contribution comes from the steady mean transonic flow. Indeed, this amplitude is

sufficient because it will be computed quasi-steady pressure rates, which would be

independent of the amplitude of the motion. Furthermore, it was identified by Silva et al.

(2002 A) that for small angles of attack (up to 2.0o ) the flow behaves linearly with the

variation in angle of attack with regard to the shock wave displacement in steady flow

conditions. However, for larger angles of attack it is possible to have shock induced

separations, which are not desirable for the computation of the quasi-steady reference

pressure, because this phenomenon is out of the scope of the present investigation. The main

concern is to have a nonlinear steady mean flow distribution with regard to the shock

positioning and strength.

The weighting operators are introduced in the aeroelastic analysis as correction

factors, yielding the computed flutter speeds shown in Tables (5.3) and (5.4). Table (5.3)

presents a comparison between the transonic flutter computation based on the correction using

steady and unsteady nonlinear pressures and experimental results (Yates, 1988). Table (5.4)

includes comparisons with some well known aeroelastic analysis codes (Chen et al, 2000)

with unsteady downwash correction procedure results.

Table 5.3: Flutter speeds and frequencies for AGARD wing 445.6.

Experimental Linear Steady Correction Unsteady Correction


M∞ VF [m/s] ωF [Hz] VF [m/s] ωF [Hz] VF [m/s] ωF [Hz] VF [m/s] ωF [Hz]
0.678 231,37 17,98 239,89 20,18 213,82 21,25 239,30 23,76
0.901 296,69 16,09 299,30 16,38 275,65 17,35 284,76 17,12
0.960 309,01 13,89 329,18 14,57 315,97 15,65 298,63 15,41
Chapter 5 – Results and Discussion 136

Table 5.4: Flutter speeds and frequencies for AGARD wing 445.6.

Unsteady Correction CAPTSD (nonlinear) ZTAIC


M∞ (Bennet et al.,1989) (Chen et al., 2000)
VF [m/s] ωF [Hz] VF [m/s] ωF [Hz] VF [m/s] ωF [Hz]
0.678 239,30 23,76 234,09 19,2 231,95 19,30
0.901 284,76 17,12 290,17 15,8 294,19 16,38
0.95 N/A N/A 291,39 12,8 287,91 13,46
0.96 298,63 15,41 N/A N/A N/A N/A

The ZTAIC method is a modal aerodynamic influence coefficiet matrix correction

method, based on the transonic equivalent strip method developed by Liu et al. (1983) and

further extended by (Chen et al.,2000). The CAP-TSD method is a time domain finite

difference solution of the transonic small disturbance equation (Bennet et al.,1989) coupled

with a finite element structural dynamic model. The results shown in Tables (5.3) and (5.4)

are also represented in graphical form in Figs. (5.46) and (5.47), respectively, in terms of the

flutter speed index (FSI) as a function of the Mach number.

0.5 24 Experimental
0.48 Linear
0.46 22
Flutter Frequency (Hz)

Unsteady
0.44 C ti
20
0.42
FSI

0.4 18
0.38
Unst. Correction 16
0.36
Experimental
0.34
Linear 14
0.32 Steady correction
0.3 12
0.6 0.7 0.8 0.9 1 0.6 0.7 0.8 0.9 1

Mach Mach

Figure 5.46: AGARD wing 445.6 results - comparison of the flutter computation results

between steady and unsteady downwash weighting methods.


Chapter 5 – Results and Discussion 137

0.5 24 Experimental
0.48 Unsteady

Flutter Frequency (Hz)


0.46 22 Correction
ZTAIC
0.44
20
0.42
FSI

0.4 18
0.38 Unst. Correction
Experimental 16
0.36
0.34 ZTAIC
14
CAP-TSD
0.32
0.3 12
0.6 0.7 0.8 0.9 1 0.6 0.7 0.8 0.9 1
Mach Mach

Figure 5.47: AGARD wing 445.6 results - comparison of the flutter computation results

between the unsteady downwash weighting methods and other methods.

The flutter speed index is a nondimensional flutter speed defined as

FSI = V f ( b ωα µ ) ,
s where V f is the dimensional flutter speed, bs is the root semi-chord

length, and ωα is a reference frequency. The term inside the square root is the mass ratio

defined as µ = m ρϑ , where m is the mass of the wing, ρ is the flow density if the test

medium, ϑ is the volume of a conical frustum having streamwise root chord as a lower base

diameter, streamwise tip as the upper base diameter, and the wing span as height. The flutter

speed index computation is performed for each Mach number, depending on the freestream

flow properties and geometrical properties of the wing, as FSI M ∞ =0.678 = V f 1662.40 ,

FSI M ∞ =0.901 = V f 2405.18 , and FSI M ∞ =0.96 = V f 3012.81 .

One can observe in Fig. (5.46) that the flutter speed indexes computed from the

correction based on unsteady pressures indicates the presence of the transonic dip. This

phenomenon is characterized by a decrease of the slope of the flutter speed plot as a function

of the Mach number, when comparing with the linearly predicted one. In the same Figure are

also shown results from the correction based on steady pressures. One may notice that there is
Chapter 5 – Results and Discussion 138

a good agreement when comparing the dip slope between the unsteady pressures based

correction procedure and the experimental data. The steady pressures based correction

underestimates some of the flutter speeds, as well as the dip slope. The reason for these

discrepancies is related to the absence of a nonlinear unsteady pressures contribution, because

the reference pressures, over which the correction factors are computed, are from steady

nature.

One feature to be noted in the same results (Fig. 5.46) is that in the subsonic Mach

number case (M∞=0.678), the correction based on unsteady pressures did not change

significantly the linear prediction. This fact can be understood by resorting to Figs. (5.14),

(5.21) and (5.28), where one should observe the differences in the pressure phases from the

linear and nonlinear calculations, for each Mach number and for the same amplitude of the

motion ∆α = 2.00 . Considering those differences between the linear and nonlinear pressure

distributions, the resulting correction factors for the transonic flow conditions will properly

introduce the necessary changes in the downwash vector to take into account the nonlinear

behavior. However, at the subsonic flow condition the differences between the amplitude of

the linear and the nonlinear computed pressures are very small and not sufficient to introduce

changes in the computation of the corrected flutter speed. The same behavior is observed in

the frequency plots shown in Fig. (5.46). Note that at M∞=0.678 the flutter frequency is not

subjected to significant changes, when comparing it with the uncorrected results. On the other

hand, in the case of the computed flutter frequencies at M∞=0.901, and M∞=0.96, the

frequencies are subjected to important changes resulting in corrected values which are in

better agreement with the experimental ones.

The results presented by Chen et al. (2000), as indicted in Fig. (5.47), show that the

transonic dip phenomenon is well characterized in the solution with the ZTAIC method, and

the computed flutter speed also presents good agreement with the experimental data. The
Chapter 5 – Results and Discussion 139

ZTAIC procedure is based in a more comprehensive theory of AIC matrix correction. This

method employs two-dimensional finite differences solutions for the computation of three-

dimensional nonlinear unsteady pressures distributions, for a set of different downwash

modes, by the use of the transonic equivalent strip method (Liu et al., 1983). If one considers

that the downwash correction method is based on unsteady pressures computed for a single

pitch mode, it is clear that one should expect the ZTAIC method to yield a better correlation

with the experimental data. In Fig. (5.47), the CAP-TSD code calculations (Bennet et al.,

1989) also yields good results, since its formulation is based on a finite difference nonlinear

solution of the three dimensional form of the transonic small disturbance equations.

Furthermore, the ZTAIC and CAP-TSD methods represent adequately the severity in the

flutter dip phenomenon, which in this case is a desirable feature in transonic flutter

prediction. A disadvantage regarding such procedures is the dependency on two and three-

dimensional unsteady finite difference solutions of the nonlinear equations, respectively,

increasing the computational cost in comparison with the downwash correction method.

The next step is to perform a sensitivity analysis with regard to the variation of the

dynamic amplitude input, which is used to generate the nonlinear unsteady pressure

distribution, taken as reference conditions for the computation of the correction factors. The

objective is to understand the sensitivity of the computed aeroelastic system stability margins

with respect to the amplitude of the motion, which may be related to the linear/nonlinear

behavior. Tables (5.5) and (5.6) present the computed flutter speeds, based on the unsteady

downwash correction method, using the resulting pressures with respect to a set of

displacement amplitudes.
Chapter 5 – Results and Discussion 140

Table 5.5: Flutter speeds and frequencies for AGARD wing 445.6. until ∆α=1.5o.

∆α=0.25o ∆α=0.5o ∆α=1.0o ∆α=1.5o


M∞ VF [ft/s] ωF [Hz] VF [ft/s] ωF [Hz] VF [ft/s] ωF [Hz] VF [ft/s] ωF [Hz]
0.678 776.86 23.49 777.62 23.52 772.97 23.39 774.67 23.44
0.901 954.32 17.26 945.92 17.96 934.21 16.71 929.16 16.93
0.960 1046.1 15.36 1039.9 15.19 1020.5 15.47 999.35 15.26

Table 5.6: Flutter speeds and frequencies for AGARD wing 445.6, until ∆α=3.0o.

uncorrected ∆α=2.0o ∆α=2.5o ∆α=3.0o experimental


M∞ VF [ft/s] ωF [Hz] VF [ft/s] ωF [Hz] VF [ft/s] ωF [Hz] VF [ft/s] ωF [Hz] VF [ft/s] ωF [Hz]
0.678 787.03 17.54 785.09 23.76 818.58 20.97 861.69 28.93 759.1 17.98
0.901 981.95 15.28 934.24 17.12 983.37 17.01 1086.22 18.03 973.4 16.09
0.960 1080.0 14.35 979.72 15.41 1021.1 15.77 1167.15 17.17 1013.8 13.89

Tables (5.5) and (5.6) include comparisons with the results computed from the

uncorrected aeroelastic model and experimental measurements by Yates (1988). One should

observe that the flutter speeds present significant variation with the nature of the unsteady

pressure data used to compute the correction factors. These results are graphically represented

in Figs. (5.48), as the variation of the non-dimensional flutter speed (FSI) with the freestream

Mach number.

0.5 0.55

0.5
0.45

0.45
FSI

0.4
FSI

Experimental ∆α = 1.5
0.4 ∆α = 2.0
Linear
∆α = 0.25 ∆α = 3.0
0.35
0.35 Experimental
∆α = 0.5
∆α = 2.5
∆α = 1.0
Linear
0.3
0.3
0.6 0.7 0.8 0.9 1
Mach 0.6 0.7 0.8 0.9 1
Mach

Figure 5.48: Comparison between the flutter speed plots as a function of the amplitude of the

motion.
Chapter 5 – Results and Discussion 141

In Fig. (5.48), it is possible to notice that the behavior of the non-dimensional flutter

speed with the Mach number presents a certain proportionality in the slopes of the curves up

to ∆α=1.0o. As the motion amplitude approaches to this value, the predicted flutter speed for

M∞=0.96 approaches the corresponding experimental value. However, the transonic dip,

which is characterized as the slope of the non-dimensional flutter speed curve as a function of

the Mach number, is less pronounced than the experimental one. Also in Fig (5.48), another

set of computed flutter speeds is presented for greater amplitudes of the dynamic angle of

attack. The best results in approaching the transonic dip slope is when one considers a

dynamic angle amplitude of ∆α=2.0o, at the same time as the flutter speeds are slightly

underestimated. Above this value, an interesting result should be noted. The computed flutter

speeds at 2.5o are nearly coincident with the experimental ones in the transonic Mach number

range. Otherwise, it is possible to note an increase of the flutter speeds computed at 3.0o. In

Figs. (5.49) and (5.50) the behavior of the aeroelastic stability of tha AGARD 445.6 wing is

summarized graphically representing the flutter speeds and frequencies as functions of the

amplitude of the motion.

1200
Mach=0.678
Mach=0.901
1100
Mach=0.96

1000
Vf (ft/s)

900

800

700
0.00 0.25 0.50 0.75 1.00 1.25 1.50 1.75 2.00 2.25 2.50 2.75 3.00
∆α

Figure 5.49: Flutter speed behavior as a function of the amplitude of the motion.
Chapter 5 – Results and Discussion 142

35.0
Mach=0.678
30.0 Mach=0.901
Mach=0.96
25.0
Frequency (Hz)

20.0

15.0

10.0

5.0

0.0
0.00 0.25 0.50 0.75 1.00 1.25 1.50 1.75 2.00 2.25 2.50 2.75 3.00
∆α

Figure 5.50: Flutter frequency behavior as a function of the amplitude of the motion.

It should be noted that up to ∆α= 2.0o, the flutter speed decreases, and also the

transonic dip is more evident, reinforcing the fact that the transonic dip phenomenon is mainly

governed by the nonlinear unsteady flow conditions. However, as soon as the angle of attack

increases beyond 2.0o, it should be noted that the flutter speed increases with the amplitude of

the motion. On possible reason for this behavior may be related to the differences in the

complex pressures which may be observed in the pressure plots for the amplitudes 2.5o and

3.0o, given in Figs (5.15), (5.16), (5.22), (5.23), (5.29), and (5.30). Looking at these plots, one

may notice that there are significant differences between the linear and the nonlinear

pressures, with regard to the prediction of the leading edge suction peak. These differences

are resulting from the increase aerodynamic lag associated to the displacement of a secondary

pressure peak present due to the thickness of the profile, as it was identified in Figs (5.31)

through (5.42). Thus, the center of pressure position is changed in such way that it modifies

the moment on the wing. Consequently, its flutter characteristics will be a function of the

amplitudes of the motion, as one should observe in Fig (5.49) and (5.50), confirming that the
Chapter 5 – Results and Discussion 143

aerodynamic lag effect is playing and important role in the stability of the aeroelastic system.

Furthermore, it is possible to note that, up to the amplitude of 2.0o, the transonic dip is more

evident as the angle of attack increases, indicating the nonlinearity of such phenomenon, with

regard to the aerodynamic lag due to the thickness effect.

The present investigation indicates that computed flutter speeds using unsteady

downwash correction method depend on amplitudes of the motion.. In the linear/non-linear

investigations presented in Chapter 3, the transonic flow linear behavior limit could be

established for disturbances in angle of attack bellow 0.35o, assuming the moment coefficient

criterion. One should recall that moments play an important role in the flutter phenomenon.

Therefore, the variation of the flutter speeds and frequencies as a function of amplitudes of

the motion results from the nonlinear flow behavior, regarding thickness effects.

5.3 - Evaluation of the Successive Kernel Expansion Method (SKEM)

5.3.1 – Preliminary Considerations

The scope of this section is to present the results regarding application of the extension

of the downwash weighting method, which had its development was presented in Chapter 4.

This procedure is based on a nonlinear steady kernel matrix replacement inside the series

expansion of the kernel function. The procedure was denominated as the successive kernel

expansion method (SKEM). The discrete element kernel function method to be corrected is

presented in the work of Chen et al. (1993) and in Zona Technology (2003), named as ZONA

6 method, implemented in the ZAERO software system. This code was selected because it

allows comprehensive modifications in terms of programming, with regard to the

implementation of the correction procedure to be here investigated.


Chapter 5 – Results and Discussion 144

The validation of the successive kernel expansion method consists basically in three

test cases regarding unsteady pressure calculation, and three flutter cases. A broad variety of

wings and profiles were considered for the present studies, ranging from thin profiles, as the

F-5 wing and the AGARD 445.6 wing, to thick supercritical profiles, like the LANN, YXX

and PAPA wings, and also parabolic arc profiles in the Lessing wing case. The wing

planforms were also different for several cases, unswept (PAPA wing, Lessing wing) or swept

(AGARD 445.6, YXX, LANN wings), and with low aspect ratio (F-5, PAPA and Lessing

wings) or high aspect ratio (LANN and YXX wings). Complex configurations, as wing store

configurations and complete aircraft models were not considered in the present investigation,

since the purpose is to evaluate, as a first step, the efficiency of this new correction method,

based on wing alone standard aeroelastic configurations.

5.3.2 - Validation of Unsteady Pressure Computations

The validation to be here performed consists in the computation of unsteady pressure

distributions using the successive kernel expansion method in order to confront it with

experimental data. The nonlinear steady pressures distributions were computed by the CFL3D

Navier-Stokes computational fluid dynamic solver (Thomas et al., 1990, and Rumsey et al.,

1997), whose results for each test configuration were taken from the work of Chen et al.

(2003). The test cases under investigation are:

A) - F-5 wing pitching about 50% root chord at M∞ = 0.9 and k = 0.275;

B) - F-5 wing pitching about 50% root chord at M∞ = 0.948 and k = 0.264;

C) - LANN wing in pitch mode about 62% root chord at M∞ = 0.822 and k = 0.105;

D) - Lessing wing in first bending mode at M∞ = 0.9 and k = 0.13.


Chapter 5 – Results and Discussion 145

A) F-5 Wing Pitching About 50% Root Chord at M∞ = 0.9 and k = 0.275 :

Figure 5.51 - Sketch of the F-5 wing and the corresponding discrete element aerodynamic

model mesh.

The ZONA 6 aerodynamic model is comprised of 200 panels, uniformly distributed

along the chord length, over the wing surface (Fig. 5.51).The reference conditions for the

computation of the nonlinear steady kernel are based on steady state nonlinear pressure

distributions, which were computed using the CFL3D Navier-Stokes solution. At M∞ = 0.9

and α = 0° the pressure coefficients differences correlates very well with the test data

(Tijdeman et al., 1978), as one should note in Fig (5.52). Quasi-steady pressure disturbances

ratios are computed in order to obtain the downwash weighting operator, following the steady

downwash correction procedure, presented in Chapter 4. Therefore, the ratios δ∆C pnl need to

be computed by ( ∆C p (α = 0.5 ) − ∆C p (α = 0.0 ) ) ( ∆α = 0.5 ) ,


nl o nl o o
leading to the nonlinear

kernel function matrix  D nl (ik = 0)  , obtained from the post-multiplication of the AIC matrix
Chapter 5 – Results and Discussion 146

by the weighting operator. For the present test case investigation, the represented quasi-steady

motion is a rigid body pitch displacement around the 50% chord location.

-0.8 -0.8
y/2b=51.5% y/2b=87.5%
-0.4 -0.4
Steady Cp

Steady Cp
0 0

0.4 Experiment (Upper) 0.4 Experiment (Upper)


Experiment (Lower) Experiment (Lower)
0.8 CFL3D (Upper) CFL3D (Upper)
0.8
CFL3D (Lower) CFL3D (Lower)
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1

X/C
X/C
-0.8
y/2b=97.7%
-0.4
Steady Cp

0.4 Experiment (Upper)


Experiment (Lower)
0.8
CFL3D (Upper)
CFL3D (Lower)
0 0.2 0.4 0.6 0.8 1
X/C

Figure 5.52 : Steady pressure distribution for the F-5 wing at M∞ = 0.9 and α = 0°.

The quasi steady pressure ratios are represented in Fig. (5.53), where, it is possible to

observe that the pressure jump with respect to the shock wave is preserved. The subsequent

step is to compute the unsteady pressures for a known reduced frequency, using the

successive kernel expansion procedure.


Chapter 5 – Results and Discussion 147

Y/2b=51.5 % Y/2b=87.5 %
40 25

35
20
∆ Cp/∆α (steady)

30

∆ Cp/∆α (steady)
25 15

20
10
15
10 5
5
0
0
0.0 0.2 0.4 0.6 0.8 1.0
0.0 0.2 0.4 0.6 0.8 1.0
-5 -5
X/C X/C

60
Y/2b=97.7 %

50
∆ Cp/∆α (steady)

40

30

20

10

0
0.0 0.2 0.4 0.6 0.8 1.0
-10

X/C

Figure 5.53 : Pressure coefficient ratio for the F-5 wing at M∞ = 0.9 and ∆α = 0.5°.

In Fig. (5.54) and (5.55) the real and imaginary parts of the pressures respectively, are

compared with the steady downwash correction method (DWM), as well as the experimental

measurements. Fig (5.53) shows a good agreement between the results.


Chapter 5 – Results and Discussion 148

35 30
Experiment y/2b=51.5% Experiment
30 25 y/2b=87.5%
SKEM SKEM
25 20 DWM
DWM

Re (∆ Cp)
20 15
Re (∆ Cp)

15 10

10 5

5 0
0 0.2 0.4 0.6 0.8 1
0 -5
0 0.2 0.4 0.6 0.8 1
-5 -10

X/C X/C

60
Experiment
50 SKEM
DWM
40
y/2b=97.7%
Re (∆ Cp)

30

20

10

0
0 0.2 0.4 0.6 0.8 1
-10

X/C

Figure 5.54 : Real part of the pressure distribution for the F-5 wing at M∞ = 0.9 and reduced

frequency k = 0.275.
Chapter 5 – Results and Discussion 149

15 10
10
5
5

Im (∆ Cp)
0
Im (∆ Cp)

0
0 0.2 0.4 0.6 0.8 1
-5 0 0.5 1
-5 y/2b=87.5%
-10
Experiment Experiment
-15 -10
SKEM SKEM
-20 y/2b=51.5% DWM
DWM
-15
-25
X/C X/C

15

10

5
Im (∆ Cp)

0
0 0.2 0.4 0.6 0.8 1
-5
y/2b=97.7%
-10
Experiment
-15 SKEM
DWM
-20

X/C

Figure 5.55 : Imaginary part of the pressure distribution for the F-5 wing at M∞ = 0.9 and

reduced frequency k = 0.275.

However, the results presented in Fig. (5.55), for the imaginary part show that, the

steady pressures-based downwash correction leads to an erroneous shock jump behavior,

when comparing with the experimental measurements, whereas the results from the

successive kernel expansion method agree with the experimental measurements.


Chapter 5 – Results and Discussion 150

In the station nearest the tip the discrepancies between the computed and experimental

pressures are more pronounced. This fact is related to the quality of the nonlinearly computed

steady pressures from the Navier Stokes simulation. Near the tip, in most cases the mesh

presents some problems in representing the tip geometry and this fact can introduce some

discrepancies in computing the flow in this region. Therefore, problems regarding the

computation of pressure distributions lead to incorrect unsteady pressures distributions

predicted by the successive kernel expansion method, because this procedure depends on

those steady pressures distributions.

B) F-5 Wing Pitching About 50% Root Chord at M = 0.948 and k = 0.264

The next results to be presented are analogous with the previous ones, but for different

Mach number and reduced frequency. In this case, the Mach number is 0.948, and the reduced

frequency is 0.264. The same comparison was performed against experimental data (Tijdeman

et al., 1978), for a steady state zero angle of attack situation, in order to evaluate the quality of

the CFL3D solution, as presented in Fig. (5.56).

The quasi-steady pressure ratios are plotted in Fig. (5.57). The shock jump is well

characterized, however, at spanwise station 64.1%, there is a second peak near the leading

edge, caused by a distortion on the prediction of the leading edge suction peak at angle of

attack of 0.5o.
Chapter 5 – Results and Discussion 151

-0.8 -0.8
y/2b=35.2% y/2b=64.1%
-0.4 -0.4

Steady Cp
Steady Cp

0 0

0.4 Experiment (Upper) 0.4 Experiment (Upper)


Experiment (Lower) Experiment (Lower)
CFL3D (Upper) CFL3D (Upper)
0.8 0.8
CFL3D (Lower) CFL3D(Lower)

0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1

X/C X/C

-0.8
y/2b=81.7%
-0.4
Steady Cp

0.4 Experiment (Upper)


Experiment (Lower)
0.8
CFL3D (Upper)
CFL3D (Lower)
0 0.2 0.4 0.6 0.8 1
X/C

Figure 5.56 : Steady pressure distribution for the F-5 wing at M∞ = 0.948 and α = 0°.
Chapter 5 – Results and Discussion 152

Y/2b=35.2 % Y/2b=64.1 %
10 25
9
∆ Cp/∆α (steady)

8 20

∆ Cp/∆α (steady)
7
6 15
5
4 10
3
2 5
1
0 0
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0

X/C X/C

Y/2b=81.7 %
35

30
∆ Cp/∆α (steady)

25

20

15

10

0
0.0 0.2 0.4 0.6 0.8 1.0

X/C

Figure 5.57 : Pressure coefficient ratio for the F-5 wing at M∞ = 0.948 and ∆α = 0.5°

In this higher Mach number, one should notice that the results regarding successive

kernel expansion method are also in good agreement with the experimental data (Fig. 5.58),

even though with the presence of strong shock waves. The shock wave jump behavior is well

represented for both real and imaginary part of the pressure coefficients, as one should
Chapter 5 – Results and Discussion 153

observe in Figs. (5.58) and (5.59). However, near the leading edge at the spanwise station

64.1%, a discrepant behavior of the computed complex pressures is noted, since the same

distortion is embedded in the reference quasi-steady pressures.

25 20
Experiment
15
20 SKEM

15
DWM 10

Re (∆ Cp)
Re ( ∆ Cp)

5
y/2b=35.2%
10
0
5 0 0.2 0.4 0.6 0.8 1
-5 Experiment
0
SKEM y/2b=64.1%
-10
0 0.2 0.4 0.6 0.8 1 DWM
-5 -15

X/C X/C

25
Experiment
20 SKEM
DWM
Re (∆ Cp)

15
y/2b=81.7%

10

0
0 0.2 0.4 0.6 0.8 1

X/C

Figure 5.58 : Real part of the pressure distribution for the F-5 wing at M∞ = 0.948 and reduced

frequency k = 0.264.
Chapter 5 – Results and Discussion 154

6 6

4 y/2b=35.2% 4 y/2b=64.1%
2 2

0 0

Im (∆ Cp)
Im (∆ Cp)

0 0.2 0.4 0.6 0.8 1


-2 -2 0 0.5 1

-4 -4 Experiment
-6 Experiment -6 SKEM
SKEM -8
DWM
-8
DWM
-10 -10

X/C X/C

4 y/2b=81.7%
2

0
0 0.2 0.4 0.6 0.8 1
Im (∆ Cp)

-2

-4

-6

-8
Experiment
-10
SKEM
-12
DWM
-14

X/C

Figure 5.59 : Imaginary part of the pressure distribution for the F-5 wing at M∞ = 0.948 and

reduced frequency k = 0.264.

The main cause for the differences between the pressures computed using the

successive kernel expansion method, near the leading edge of spanwise station 64.1 % (Figs.

5.58 and 5.59), is related to problems regarding the nonlinear quasi-steady pressures

computation at this position. Therefore, from the results presented above, it is concluded that

an accurate unsteady pressure prediction depends strongly on a correct quasi-steady mean

flow pressure distribution.


Chapter 5 – Results and Discussion 155

C)- LANN Wing in Pitch About 62% Root Chord at M∞ = 0.822 and k = 0.105.

Figure 5.60 : Sketch of the LANN wing and the corresponding discrete element aerodynamic

model.

The LANN wing is a supercritical wing with aspect ratio of 7.92 and 25o of sweep

angle along ¼ chord axis. It consists in a good test case for evaluation of the present

procedure, since it is a supercritical profile configuration. The wing was discretized in 280

panels in order to perform unsteady pressures computation based on the ZONA 6 linear

aerodynamic model. The same steps were repeated with regard to the nonlinear pressure

distribution computation. A steady CFL3D Navier-Stokes solution at M∞ = 0.822 and α = 0.6°

was simulated (Fig. 5.61) and the pressures correlate very well with test data (Malone et al.,

1960). The unsteady pressure distributions presented by Malone et al., (1960) correspond to

the wing undergoing a rigid body pitch oscillation around the 62% of the root chord location.

The quasi-steady pressure coefficient ratio δ∆C pnl ∆α was computed for a smaller angle of

attack difference, by ( ∆C p (α = 0.75 ) − ∆C p (α = 0.6 ) ) ( ∆α = 0.15 ) ,


nl o nl o o
since this profile is

supercritical, and smaller angle of attack changes can strongly modify the transonic flow

behavior. The quasi-steady pressure differences ratios are plotted in Fig. (5.62), where it is
Chapter 5 – Results and Discussion 156

possible to note that, at spanwise station 82.5%, there are again two peaks, since this station is

near from the tip, and the same aforementioned meshing problem arose.

Cp at Y/L=0.475 Cp at Y/L=0.65
1.2 1.2

1 1

0.8 0.8

0.6 0.6

Cp (steady)
Cp (steady)

0.4 0.4

0.2 0.2

0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
-0.2 -0.2
Experiment (upper) Experiment (upper)
-0.4 -0.4
Experiment (lower) Experiment (lower)
-0.6 -0.6
CFL3D CFL3D
-0.8 -0.8
X/C
x/c

Cp at Y/L = 0.825
-1.2
y/2b=82.5%
-0.8
Steady Cp

-0.4

0.4
Experiment (Upper)
Experiment (Lower)
0.8 CFL3D (Upper)
CFL3D (Lower)
0 0.2 0.4 0.6 0.8 1

X/C

Figure 5.61: Steady pressure distribution for the LANN wing at M∞ =0.822 and α = 0.6°.
Chapter 5 – Results and Discussion 157

Y/L=0.475 Y/L=0.65
80 50

70
40
60

∆ Cp/∆α (steady)
∆ Cp/∆α (steady)

50 30

40
20
30

20 10

10
0
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
-10 -10
X/C X/C

Y/L=0.825
50

40
∆ Cp/∆α (steady)

30

20

10

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
-10

-20

X/C

Figure 5.62 : Pressure coefficient ratio for the LANN wing at M∞ =0.822 and ∆α = 0.15°.

In Figs. (5.63) and (5.64) one should observe a good agreement with test data is

obtained at the spanwise stations y/2b = 47.5% and 65% by the successive kernel expansion

method, while the steady downwash correction method gives incorrect shock jump in the

imaginary pressure component (Fig. 5.64). The oscillating pressure behavior near the shock

jump at y/2b = 82.5% computed by the successive kernel expansion method is due to
Chapter 5 – Results and Discussion 158

oscillations δ∆C pnl ∆α computed by CFL3D simulation. However, the shock positioning and

strength ate well predicted for the three spanwise stations.

80 50
Experiment Experiment
70
SKEM 40 SKEM
60 DWM DWM
50 30
y/2b=47.5% y/2b=65%
Re (∆ Cp)

Re (∆ Cp)
40
20
30

20 10

10
0
0
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
-10 -10
X/C X/C

50
Experiment
40
SKEM
30 DWM
20
Re (∆ Cp)

10

0
0 0.2 0.4 0.6 0.8 1
-10

-20
y/2b=82.5%
-30

X/C

Figure 5.63 : Real part of the pressure distribution for the LANN wing at M∞ =0.822 and

reduced frequency k = 0.105.


Chapter 5 – Results and Discussion 159

10 10
5
5
0
0
0 0.2 0.4 0.6 0.8 1
-5 0 0.5 1

Im (∆ Cp)
Im (∆ C p)

-5
-10
-10
-15 y/2b=47.5% y/2b=65%
-15 Experiment
-20 Experiment
-20 SKEM
-25 SKEM
DWM DWM
-30 -25
X/C X/C

10

8 y/2b=82.5%
6

2
Im (∆ Cp)

0
0 0.2 0.4 0.6 0.8 1
-2

-4
Experiment
-6
SKEM
-8
DWM
-10
X/C

Figure 5.64 : Imaginary part of the pressure distribution for the LANN wing at M∞=0.822 and

reduced frequency k = 0.105.


Chapter 5 – Results and Discussion 160

D) Lessing Wing in First Bending Mode at M = 0.9 and k = 0.13

Figure 5.65 : Sketch of the Lessing wing and the corresponding discrete element aerodynamic

model.

The Lessing wing has a rectangular planform with an aspect ratio equal to 3.0, and a

5% parabolic arc profile section. Fig. (5.65) presents the sketch of the wing as well as the

linear aerodynamic model mesh employed in the linear unsteady pressures computation via

the ZONA 6 aerodynamic theory, comprised by 220 panels, equally distributed along the

chord length. Unsteady pressure measurements were performed for this test wing, considering

a bending mode displacement of the lifting surface (Lessing et al., 1960). The objective of

this test case is to demonstrate that the successive kernel expansion method can give accurate

unsteady pressure of a bending mode even if δ∆C pnl ∆α is computed based on a quasi steady

pitch mode at k = 0. Steady pressures distribution were computed by the CFL3D Navier

Stokes code, and the pressures distribution for M∞ =0.9 and α = 0° are presented in Fig.

(5.66), showing a good agreement with experimental data.

The computation of the quasi steady pressure ratios δ∆C pnl ∆α was performed as

( ∆C p (α = 0.2 ) − ∆C p (α = 0.0 ) ) ( ∆α = 0.2 ) , and the resulting distributions are presented in


nl o nl o o
Chapter 5 – Results and Discussion 161

Fig. (5.67). One should note that in the station near the wing tip there is an oscillation of the

computed quasi-steady pressures ratios due to the same reason mentioned in the previous test

cases, regarding the CFD aerodynamic meshing problems.

-0.4
-0.5
y/2b=50% -0.3
-0.4

-0.3 -0.2

Steady Cp
-0.2 -0.1
Steady Cp

-0.1 0 y/2b=70%
0
0.1
0.1 Experiment (Upper)
Experiment (Upper) 0.2
Experiment (Lower)
0.2 Experiment (Lower)
0.3 CFL3D (Upper)
0.3 CFL3D (Upper)
CFL3D (Lower) CFL3D (Lower)
0.4
0.4
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1

X/C X/C

-0.4

-0.3

-0.2
Steady Cp

-0.1

0
y/2b=90%
0.1
Experiment (Upper)
0.2
Experiment (Lower)
0.3 CFL3D (Upper)
CFL3D (Lower)
0.4
0 0.2 0.4 0.6 0.8 1

X/C

Figure 5.66 : Steady pressure distribution for the Lessing wing at M∞ =0.9 and α = 0°.
Chapter 5 – Results and Discussion 162

Y/2b=50 % Y/2b=70 %
20 18
16

15 14
∆ Cp/∆α (steady)

∆ Cp/∆α (steady)
12
10
10
8
6
5
4
2
0
0
0.0 0.2 0.4 0.6 0.8 1.0
-2 0.0 0.2 0.4 0.6 0.8 1.0

-5 -4

X/C X/C

Y/2b=90 %
14

12
∆ Cp/∆α (steady)

10

0
0.0 0.2 0.4 0.6 0.8 1.0
-2

X/C

Figure 5.67 : Pressure coefficient ratio for the Lessing wing at M = 0.9 and ∆α = 0.2°.

In the midspan station (50%), there is also a slight oscillation immediately after the

shock wave position, towards to the leading edge. This oscillation is reflected in the

computation of the unsteady pressures amplitudes for the bending displacement of the wing,

as one should observe in the corresponding plot in Fig. (5.68).


Chapter 5 – Results and Discussion 163

14 16
Experiment (Test 1) Experiment (Test 1)
14
12 Experiment (Test 2) Experiment (Test 2)
ZONA6 (Linear) 12 ZONA6 (Linear)
10 SKEM SKEM

Magnitude
Magnitude

10
8
y/2b=50 % 8 y/2b= 70%
6
6

4 4

2 2

0
0
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
X/C X/C

16

Experiment (Test 1)
14
Experiment (Test 2)
12
ZONA6 (Linear)
Magnitude

10 SKEM
8
y/2b=90%
6

0
0 0.2 0.4 0.6 0.8 1

X/C

Figure 5.68 : Pressure distribution amplitudes for the Lessing wing at M∞ =0.9 and reduced

frequency k = 0.13.
Chapter 5 – Results and Discussion 164

300 300
Experiment (Test 1) Experiment (Test 1)
250 Experiment (Test 2) 250 Experiment (Test 2)
ZONA6 (Linear) ZONA6 (Linear)
Phase Angle (deg)

Phase Angle (deg)


200 SKEM 200 SKEM

150 150

100 100

50 50
y/2b=50% y/2b=70%
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
X/C X/C

300
Experiment (Test 1)
250 Experiment (Test 2)
Phase Angle (deg)

ZONA6 (Linear)
200
SKEM

150

100

50
y/2b=90%
0
0 0.2 0.4 0.6 0.8 1
X/C

Figure 5.69 : Pressure distribution phases for the Lessing wing at M∞ =0.9 and reduced

frequency k = 0.13.

However, an excellent agreement with test data is obtained by the successive kernel

expansion method, as presented in the results summarized in the plots of Figs. (5.68), and

(5.69), whereas the linear method such as ZONA 6 fails to predict the unsteady shock effects,

as expected. These results are encouraging, since it is possible to note that this method,

employing a quasi steady pitch motion, has the capability of correcting pressures with respect

to a bending type motion.


Chapter 5 – Results and Discussion 165

In all of the test cases it was observed that the successive kernel expansion method

was capable to restore the nonlinear unsteady pressure distribution and the shock wave

dynamics. Therefore, from these results, it is possible to conclude that the unsteady character

of the small disturbance transonic flow can be computed as a contribution predicted by a

small disturbance linear aerodynamic model.

However, it is clear that the successive kernel expansion method performance is

strongly dependent on the quality of the computed steady mean flow pressure distribution.

Any distortion in the computation of the pressures difference coefficient ratios will be

reflected in the computation of the approximate unsteady aerodynamic loading.

5.3.3 - Validation of the Successive Kernel Expansion Method for Flutter

Boundaries Computation

Flutter calculations were performed taking three standard aeroelastic wing

configurations, for which the aeroelastic stability was evaluated in wind tunnel tests. The

following test cases were selected:

A) - AGARD 445.6 wing - weakened model # 3 (Yates, 1988);

B) - PAPA wing (Pitch And Plunge Apparatus, Farmer and Rivera., 1988);

C) - YXX wing (Isogai, 1983).

The successive kernel expansion method (SKEM) is adopted for unsteady pressures

computation in order to provide the unsteady aerodynamic loading for flutter computations.

The nonlinear pressure distributions were computed for these wings undergoing quasi-steady

motion, as it was performed for the cases of transonic unsteady pressures prediction,

investigated in the last section.


Chapter 5 – Results and Discussion 166

A) Flutter Boundary of the AGARD 445.6 Weakened Wing

The investigation of the AGARD wing 445.6 aeroelastic stability is revisited, with the

objective to evaluate the performance of the successive kernel expansion procedure for the

transonic flutter prediction.. The reference nonlinear steady pressure distribution was

computed by the numerical method for solving the Navier-Stokes solver presented in

Chapter 3 at the angles of attack 0.0o and 0.5o , in order to compute the reference nonlinear

quasi-steady pressures disturbances. The discrete element aerodynamic model, based on the

ZONA 6 method, is comprised of 220 panels, uniformly distributed along the chord length

(Fig 5.1). The flutter computations were performed using the g-method. Flutter boundaries at

M∞ = 0.678, 0.901 and 0.95 are computed using ZONA6 (linear method), the steady

downwash weighting method (DWM) and the successive kernel expansion method, and the

results are presented in Fig. (5.70).

0.5 24.0
0.48 DWM
0.46 22.0 ZONA 6
Experiment
0.44
Frequency (Hz)

20.0 SKEM
0.42
FSI

0.4 18.0
0.38 DWM
0.36 ZONA 6 16.0
Experiment
0.34 SKEM
14.0
0.32
0.3 12.0
0.6 0.7 0.8 0.9 1 0.6 0.7 0.8 0.9 1
Mach Mach

Figure 5.70 : Flutter boundaries (dynamic pressure and flutter frequency) for the AGARD

wing 445.6, weakened model #3.


Chapter 5 – Results and Discussion 167

Among these three methods, the results obtained by the latter procedure present the

best correlation with the test data, when observing the capability in capturing the transonic

dip, and in decreasing the flutter speed, approaching the experimental results. Note that the

flutter speeds predicted by the steady downwash correction method applied to the ZONA 6

aerodynamic model lead to flutter speeds higher than the ones computed by the same

correction procedure using the doublet lattice method, for which the results were presented in

Fig. (5.46).

Previous computations indicated that large correction factors, resulting from the

correction of the pressures near the leading edge, lead to distortions of the computed unsteady

pressures. These large correction factors, which are computed by the steady downwash

correction, are the reason for the large differences between the computed flutter speed by the

steady downwash correction procedure with respect to the other results, already presented in

Fig (5.46). Those large correction factors would tend to increase the unsteady aerodynamic

loading in such way that the flutter speeds are underestimated, because there is an

overestimation of the magnitude of the unsteady loading.

A strategy to circumvent the problem is to suppress the correction of the pressures at

panels near the leading edge. This approach is consistent, because the objective of the present

investigation is to correct an unsteady aerodynamic loading to take into account transonic

effects, instead of potential effects such as the differences in the computation of the leading

edge suction peak. Furthermore, the main cause for these differences is related to a mesh

interpolation problem. One should observe that the control point is at the centroid of the

panel. Near the leading edge, the pressure distribution along the chord, presents higher

gradients when comparing to the remaining portion of the airfoil section. Thus, the

interpolated pressures at the control point are lower in magnitude than the pressures between

the control point and the leading edge of the wing, predicted at the mesh points of the finite
Chapter 5 – Results and Discussion 168

difference Navier Stokes numerical model. Since the purpose of the correction method is to

take into account the transonic flow behavior in the sense of the contribution of the shock

wave positioning and strength, the suppression of correction effects of potential nature does

not impair the procedure.

In the steady downwash correction the pressure phases are not corrected, resulting in

more conservative flutter dynamic pressures, when comparing with strictly linear results. In

addition, one should note that the results of this method, for transonic Mach numbers, have

the same behavior of the uncorrected linear ones, regarding the transonic dip curve slope. This

is an indication that the transonic dip behavior is closely related to the contribution of the

imaginary part of the pressures.

In the case of the successive kernel expansion procedure, the computed correction

factors result from the matching of nonlinear unsteady pressures, composed by nonlinear

steady mean flow pressures and a linear unsteady pressures contributions. Therefore, the

correction factor will take into account in an approximate form the unsteady transonic flow

behavior, since its computation is referred to those nonlinear unsteady pressures. In summary,

the dip slope, predicted by the successive kernel expansion method is increased, because the

correction procedure takes into account the influence of the real and imaginary part of the

transonic unsteady pressures. Thus, this improvement in the reference conditions leads to

better results in approaching the experimental measurements

The plots of the flutter frequency as a function of the Mach number, (Fig. 5.70)

indicate that the successive kernel expansion method overpredicts the flutter frequencies. The

reason for this distortion is that a larger change in the flutter frequency is expected since the

successive kernel expansion procedure introduces an important contribution to the imaginary

part of the complex eigenvalue problem involved in the flutter solution. Furthermore, the

correction factors are computed based on pressure disturbances ratios with respect to a given
Chapter 5 – Results and Discussion 169

pitch amplitude of the motion. However, the flutter mode is not a single pitch displacements,

as it consists in a flexure-torsion mode shape displacement, in the case of this wing under

investigation. Therefore, there is no guarantee that the computed correction factors will

properly restore the pressures distribution at the flutter mode, even considering that these

correction factors are computed independently of the motion nature.

It should be observed that the aerodynamic interference effects are computed by the

discrete kernel function method, which in the present test case is the ZONA 6. Interference

effects obtained from a pitch downwash mode are different from the interference effects due

to a flexure-torsion mode, even if the reduced frequency is the same. This fact suggest the use

of a “frozen” flutter mode or any other elastic mode, instead of a single pitch mode, for the

computation of the nonlinear quasi-steady pressures conditions to obtain the correction

factors. This approach allows to take into account the nonlinear interference effects, which are

embedded in the reference quasi-steady pressure distribution.

These observations are also valid for the explanation of the results with regard to the

flutter speeds. The differences of the computed flutter speeds using the successive kernel

expansion method with respect to the experimental results (Fig. 5.70) are justified by the same

reason, that is, the absence of a correction which takes into account the interference effects

may lead to differences between the computed and the experimental flutter speeds.
Chapter 5 – Results and Discussion 170

B) Flutter Boundary of PAPA Wing at α= 1°

The next test case is the investigation of the aeroelastic stability margins of the PAPA

wing. This wing is a good test case to evaluate the capability in correcting the unsteady

aerodynamics based on a supercritical flow pressure distribution, and for a low aspect ratio

wing. The ZONA 6 aerodynamic model is comprised of 200 panels, equally distributed along

the chord length. The same modeling features, employed in the AGARD wing 445.6 case are

considered, such as the suppression of correction factors for the leading edge panels. The

flutter solution technique employed is the g-method. The results of the flutter computation of

the PAPA wing, at two distinct angles of attack, are presented in Figs. (5.71) and (5.72).

200 6
Experiment α = +1 (deg) Experiment α = +1 (deg)
Flutter Frequency (Hz)

190 Zona 6 Zona 6


Dyn. Pressure (psf)

DWM 5.5 DWM


180 SKEM SKEM
170
5
160

150
4.5
140

130 4
0.3 0.4 0.5 0.6 0.7 0.8 0.9 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Mach Mach

Figure 5.71 : Dynamic pressures and frequencies in the flutter boundaries of the PAPA wing

at angle of attack, α = 1.0o


Chapter 5 – Results and Discussion 171

200 6
Experiment Experiment
190 Zona 6 Zona 6

Flutter Frequency (Hz)


DWM DWM
Dyn. Pressure (psf)

SKEM 5.5 SKEM


180

170
5
160

150
4.5
140
α = -2 (deg) α = -2 (deg)
130 4
0.3 0.4 0.5 0.6 0.7 0.8 0.9 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Mach Mach

Figure 5.72 : Dynamic pressures and frequencies in the flutter boundaries of the PAPA wing

at angle of attack, α = -2.0o.

Another feature to be highlighted, with regard to the successive kernel expansion

method, is the capability to predict the flutter of the wing at different initial angles of attack,

instead of the linear theory, which can only predict the flutter margins for a zero angle of

attack. This is so, because the steady mean flow can be computed at a given steady state angle

of attack, and the nonlinear unsteady pressures, used to compute the correction factors, will be

obtained from the contribution of the nonlinear steady mean flow pressures added to unsteady

linear pressures. Therefore, the angle of attack contribution is included in the mean flow

conditions, because unsteady linear pressures, predicted by the linear unsteady aerodynamic

model is independent of the angle of attack.

One should observe that the downwash correction method largely underpredicts the

flutter dynamic pressure while the flutter results of the successive kernel expansion method

correlate well with test data, both for the subsonic and the transonic Mach numbers. This is so

because the unsteady components of the nonlinear pressures, introduced by the correction

procedure, play an important role in the flutter computation. However, the flutter frequencies
Chapter 5 – Results and Discussion 172

are underestimated, as one should observe in Figs. (5.71) and (5.72). The reason for those

differences are the same presented in the case of the AGARD wing 445.6. In the present test

case, the flutter results from the coalescence of the pitch an plunge mode, while the nonlinear

mean flow reference pressures are obtained for a quasi steady rigid body pitch mode.

C) - Flutter Boundary of the YXX Wing

The YXX wing is an elastic swept wing with supercritical profile, which was tested in

wind tunnel under transonic flow conditions. It has an aspect ratio of 10.0, a 17o quarter chord

sweep angle, and supercritical airfoil tapered from 16% of thickness at the root section to 12%

at the tip section. Its planform presents a crank in the trailing edge, and demands a more

detailed linear aerodynamic model. The wing was modeled by the ZONA 6 method, where

the lifting surface was subdivided in 360 panels, equally spaced along the chordwise

direction.

Figure 5.73 : Sketch of the YXX wing and the corresponding discrete element aerodynamic

model.
Chapter 5 – Results and Discussion 173

The flutter boundaries were measured by Yonemoto (1984), and also computed by

Isogai (1983). This last author identified that the flutter boundaries are extremely sensitive to

the location of the center of the pressure; a shift in 2% of the elastic axis causes a 70% of

change in the flutter dynamic pressure, at M∞ =0.7. Furthermore, this wing presents a large

aeroelastic static deformation, even at low angle of attack. Therefore, such static aeroelastic

deformation may alter significantly the position of the center of the pressure, indicating the

flutter boundary depends on the static aeroelastic behavior of the structure.

Chen et al. (2003) present an overset filed-panel method for unsteady transonic

aerodynamic computations. The YXX wing test case was also investigated, with regard to its

aeroelastic stability. The results of the ZTRAN method show that the variation of the static

angle of attack promotes a strong impact on the wing flutter boundary at transonic flow

conditions. The discrepancies of the ZTRAN method results and the experimental data are

believed to be due to the exclusion of the static aeroelastic deformation in the CFL3D

aerodynamic mesh, since it is considered in the Navier-Stokes computations that the wing is

rigid. In addition, Chen et al. (2003) concluded in their work that the reference angle of attack

condition, which presents the best correlation with experimental data is α = -1.0o. This is so,

because among the angles of attack of 2.0o, 0.0o, and –1.0o the steady background flow at α=-

1.0o , computed by the CFL3D should be the closest one to that of a static aeroelastic

deformed wing.

The quasi-steady pressure distribution to be used in the successive kernel expansion

method, and the steady downwash weighting method, was computed with the CFL3D Navier

Stokes code, for two angles of attack : -0.8o and –1.0o. The selection of the reference angle of

attack as –1.0o is made in order to compare the results of the present investigation with the

ones from the ZTRAN method. The flutter results are presented in Figs. (5.74), and (5.75).
Chapter 5 – Results and Discussion 174

10
Dyn. Pressure (KPa)

6
Experiment (AOA=0 deg)
Experiment (AOA=2 deg)
ZONA6 (Linear)
4 SKEM (AOA=-1 deg)
ZTRAN (AOA=-1 deg)
DWM (AOA=-1 deg)

2
0.6 0.65 0.7 0.75 0.8 0.85 0.9
Mach

Figure 5.74 : Flutter dynamic pressure for the YXX wing, at angle of attack α = -1.0o.

250

200
Flutter Frequency (Hz)

150

Experiment (AOA=0 deg)


Experiment (AOA=2 deg)
100 ZONA6 (Linear)
SKEM (AOA=-1 deg)
ZTRAN (AOA=-1 deg)
DWM (AOA=-1 deg)
50
0.6 0.65 0.7 0.75 0.8 0.85 0.9
Mach

Figure 5.75 : Flutter frequency for the YXX wing, at angle of attack α=-1.0o.
Chapter 5 – Results and Discussion 175

One should notice that these results presents a large overprediction of the transonic dip

phenomenon, with respect to the experiment, for both methods under investigation (DWM

and SKEM). In addition these results are also different from the ones computed by the

ZTRAN method, at α = -1.0o. However, the transonic dip behavior is represented in these

results. The overall quality of this solution is questionable, since the pressure ratios due to the

quasi-steady motion were based on a variation in angle of attack amplitude, which play an

essential role for supercritical wing profiles. At angles of attack of –0.8o and –1.0o, one should

observe that differences between the shock wave structure in the steady nonlinear pressure

field are quite significant. Probably the expected linear behavior concerning the pressures

under quasi steady motion is failing, and the changes in the flow are unrealistic.
6 – CONCLUSIONS AND FINAL REMARKS

6.1 Conclusions

A study on correction methods for aeroelastic stability analysis in transonic flow has

been developed. This investigation can be outlined in three major steps. First, a

linear/nonlinear behavior investigation of unsteady transonic flows has been conducted, in

order to allow a better understanding of the linear limits. Such knowledge, in turn, allows

correction methods to be developed and investigated. Second, downwash correction methods

for transonic flutter computation were evaluated, including a sensitivity analysis of the

computed transonic flutter speeds as a function of unsteady reference pressures resulting from

different amplitudes of the lifting surface motion. Finally, an extension of downwash

correction methods was developed, leading to a new formulation named as “successive kernel

expansion method”, for which the theoretical background and hypotheses are supported by the

subsonic discrete kernel function method and observations on the transonic flow behavior,

respectively. The conclusions that can be drawn from the present investigation are outlined in

the forthcoming paragraphs.

Simulations of unsteady transonic flow over the NLR F-5 wing model were performed

in order to study the linearity of unsteady loads and shock displacement with respect to

dynamic angle of attack amplitude and frequency. A linear behavior was quantified in terms

of the variation of lift and moment coefficient derivatives and shock displacement as a

function of the amplitude of the motion of the NLR F-5 wing.

The aerodynamic response and shock displacement were found to include higher

harmonics, which are usually not considered in correction procedures applied to linear

aeroelastic methods. In flutter analysis, there are situations in which a second mode is close to
Chapter 6 – Conclusions and Final Remarks 177

a multiple of a first mode. In such case, the presence of a second harmonic in the lower mode

aerodynamic response may facilitate the exchange of energy between the modes and be a

contributing factor to coalescence. For this reason, the present results seem to indicate the

need for considering such harmonics in future correction methods.

Linear boundaries were computed for a few reduced frequencies and spanwise stations

along the NLR F-5 wing, using two criteria, based on moment coefficient and shock

displacement. These boundaries were found to be more conservative for the shock

displacement criterion. The linear limits calculated using the moment coefficient criterion

were found to depend significantly on the spanwise station. This indicates that simply using

two-dimensional linear limits would not be appropriate when dealing with correction

methods.

The linear/nonlinear behavior investigation indicated that the unsteady transonic flow

is strongly dependent on the amplitude of the disturbances it is subjected. Notwithstanding the

dependence of the linear limits on reduced frequency and spanwise station, it is clear that

some degree of linear behavior may be assumed for aeroelastic methods that employ corrected

coefficients around mean flow reference conditions. However, this assumption of linearity

should be used with caution and would not apply to problems such as aeroelastic response or

limit cycle oscillations, unless the amplitude of deformation is kept below previously

identified linear limits that take into account reduced frequency and spanwise station. In a

context of small disturbances associated to lifting surface motion, unsteady transonic flow

behaves linearly around a nonlinear mean flow. The identified linear limits were higher than

the small amplitudes associated to aeroelastic dynamic deformations of the structure, which

ensures the feasibility of applying correction methods in this context.

The evaluation of the downwash weighting method (DWM) confirmed that the

matching of the pressures, based on the correction of the linear discrete element kernel
Chapter 6 – Conclusions and Final Remarks 178

function method (ZONA 6, doublet lattice method), gives better results than force matching

methods for approximate transonic flow computations. This is due to the fact that the steady

mean flow nonlinear pressure distribution is preserved, whereas the force matching methods

do not present this capability. One should observe that in force matching procedures, there are

fewer unknowns (weighting factors) than the order of the AIC matrix. Thus, the use of the

least squares method is required to compute the correction factors. The resulting corrected

pressures computed by such procedure present a wrong trend regarding the shock wave

dynamics. Consequently, flutter boundary computations using such procedures do not present

good results and thus are not adequate for the approximate modeling of unsteady transonic

flow for aeroelastic applications.

Downwash weighting methods, using either steady or unsteady pressures as reference

conditions were investigated. The transonic dip was captured with application of both

correction approaches. However, the decrease of the flutter speed curve slope was more

evident and closer to experiments when the unsteady downwash correction method was

employed. The flutter speed curve slope obtained by the steady downwash weighting method

was about the same as the one obtained by the uncorrected aeroelastic stability analysis. It

may be concluded that the unsteady flow contribution, embedded in the unsteady reference

pressures, plays an important role in the transonic flutter phenomenon. However, the unsteady

downwash weighting method requires unsteady CFD computations, increasing its cost with

respect to the steady downwash weighting procedure. Furthermore, in a strict sense, it should

be applied only at the specific reduced frequency for which the unsteady pressures were

obtained. Also, the unsteady downwash weighting method depends on the nature of the

unsteady reference pressures, as indicated by the sensitivity analysis to be commented in what

follows.
Chapter 6 – Conclusions and Final Remarks 179

A sensitivity analysis was performed in order to investigate the role of the unsteady

reference pressure disturbances, which are referred to a given amplitude of the lifting surface

motion on the prediction of the transonic flutter by the use of the downwash correction

method. This investigation indicated that the flutter boundaries computed by unsteady

downwash correction methods are significantly dependent on the unsteady pressures taken as

reference conditions.

The above is consistent with the observations regarding the linear/non-linear

investigations, where it was found that the linear limit associated to the lifting moment is

more sensitive to the amplitude of the motion. The flutter instability is mainly governed by

the lifting moment, which is associated to the streamwise displacement of the center of

pressures. Investigation of the local two dimensional flow over the AGARD 445.6 wing

indicated that there is a formation of a secondary suction peak which displacement may alter

the position of the center of pressures. The appearance of this secondary suction peak results

from the Navier-Stokes computation of the flow around the leading edge of a profile with

finite thickness. Therefore, aerodynamic lag appears, which is dependent on the amplitude of

the motion at a given reduced frequency. These observations explain the dependence of the

computed flutter speed on these amplitudes. The overall results confirm that the nonlinear

effects associated to the profile thickness play an important role in flutter prediction by

correction methods based on nonlinear unsteady pressure distributions.

The development of the successive kernel expansion method arose from the

aforementioned observations regarding the unsteady transonic flow behavior. The advantages

of such procedure is the dependency on a nonlinear steady state pressure distribution, instead

of nonlinear unsteady transonic pressures computation or measurement. This method was

validated for unsteady pressures calculation and flutter computations. The results showed that

the quality of the unsteady transonic pressure distributions computed by the successive kernel
Chapter 6 – Conclusions and Final Remarks 180

expansion methods depends on the quality of the steady pressures taken as reference

conditions. The shock wave dynamics was restored when compared with experimental

measurements, independently of shock strength. Some discrepancies were observed in the

unsteady pressures computation near the lifting surface tip. These are due to the CFD method

failing to accurately predict the viscous flow field there. Care must be taken in order to have a

good CFD solution because its quality has an important role in the computation of the ratio

between the steady pressure and the quasi steady motion amplitude.

The flutter computation using the successive kernel expansion method also led to good

results when predicting the transonic dip. However, some slight discrepancies were found,

mainly related to the flutter frequency prediction. The reason for these discrepancies is related

to the nature of the steady reference condition, regarding the absence of correction of

interference effects at the flutter mode, which may justify the over prediction of those flutter

frequencies. In general, the transonic dip, as well as the computed flutter speeds and

frequencies are in good agreement with experimental measurements. This procedure is a rapid

form to compute the transonic flutter speed boundaries.

The downwash weighing method based on unsteady reference pressures requires time

accurate CFD simulation. For that method, The unsteady flow Navier-Stokes solution,

expended approximately 8.2 CPU hours for the simulation of two cycles of pitching motion of

the AGARD 445.6 wing, in a Silicon GraphicsTM Octane II workstation, with two R12000

RISC architecture processors. On the other hand, the successive kernel expansion method

employs a steady nonlinear pressure distribution. The time expended for a steady flow

simulation of the same wing is approximately 1.3 hours elapsed time. The additional

computational effort expended by the ZONA 6 method modified by the successive kernel

expansion method is less than five minutes of elapsed time, on a computer with a single 2.8

GHz PentiumTM IV processor.


Chapter 6 – Conclusions and Final Remarks 181

In summary, there is a restriction in the unsteady downwash correction method, for it

is not suitable for computing unsteady transonic flow for reduced frequencies other than the

one corresponding to the reference unsteady pressures. The successive kernel expansion

method does not have the same restriction, for it has the capability of computing unsteady

pressures based on steady state pressures, for any reduced frequency of interest. Thus, if one

were to conduct computations for a number of reduced frequencies, the time required by the

unsteady downwash correction method would be the number of reduced frequencies of

interest multiplied by the expended time for the unsteady Navier-Stokes CFD solution, while

the time require by the successive kernel expansion method would essentially not depend on

the number of reduced frequencies.

6.2 Contributions

The first contribution of the present work is the linear/nonlinear investigation in terms

of Navier-Stokes solutions. The improvement when comparing with the existing literature is

the use of three-dimensional Navier-Stokes solutions instead of two-dimensional transonic

small disturbances solutions, for the identification of the linear limits.

Another contribution is a comprehensive study of the downwash correction techniques

extending their application to the matching of pressures instead of strip loads, as it was

originally proposed by Pitt and Goodman (1987). Furthermore, the capability to correct the

pressures to downwash linear relation using either steady or unsteady reference nonlinear

pressure conditions was introduced, as was the idea to correct the downwash using unsteady

pressures computed for the flutter reduced frequency predicted by the uncorrected linear

aeroelastic model, in order to evaluate the stability of the aeroelastic system.

An expedient way to compute unsteady transonic flow was developed, named in this
Chapter 6 – Conclusions and Final Remarks 182

work as the successive kernel extension method (SKEM). The application of such procedure

can be performed either for aeroelastic response or stability analysis, without the need of

unsteady CFD computations or measurements. The expediency of such method lies in the idea

of employing nonlinear steady state mean flow conditions, which are inexpensive to compute,

when compared with the unsteady ones, which were applied in the downwash correction

procedure. This feature makes this method suitable for application in industrial developments,

due to its computational efficiency.

6.3 – Recommendations for Future Work

The successive kernel expansion method may be further extended to cover full

transonic ranges including Mach numbers M∞ ≥ 1.0. The employment of different quasi-

steady patterns of motion to generate the steady non-linear reference field is another approach

to be investigated. In order to improve the performance of the procedure, a next step in the

continuation of the present effort might be to use a correction matrix which is constructed

from a set of steady downwash modes. However, such approach increases the computational

cost of the aeroelastic analysis since more steady CFD simulations would need to be

performed in order to create the correction factors. As the transonic dip phenomenon is

predicted, and the computed flutter speeds are conservative, the important aspect which

remains to be analyzed, hence, is whether the additional costs of such an approach are

worthwhile. Otherwise, the computational cost might be increased so much that it would be

near that of full computational aeroelasticity.

Additional investigation on the unsteady downwash correction method needs to be

performed in order to explain nonlinear behavior regarding the variation of the flutter speed

with the amplitude of the motion. It is suggested further investigations on viscous effects,
Chapter 6 – Conclusions and Final Remarks 183

increasing the aerodynamic mesh refinement and confronting the results against Euler

computations.

Another approach which might be investigated is the use of a set of nonlinear unsteady

pressures computed from the transformation to the frequency domain of a impulse-type

aerodynamic time response. Such approach would circumvent the disadvantage in having a

single set of unsteady pressures for the correction of the pressure to downwash linear relation,

for different reduced frequency values.

The successive kernel expansion method results are encouraging, suggesting its further

extension. Complex configurations and high frequency tests cases were not performed. It is

suggested that the method be evaluated under these conditions in order to extend its

applicability .
7 – REFERENCES

1. Albano, E., and Rodden, W. P., “A Doublet Lattice Method for Calculating

Lift Distributions on Oscillating Surfaces in Subsonic Flows”, AIAA Journal, vol. 7, No. 2,

1969 pp 279-285.

2. Appa, K., “Constant Pressure Panel Method for Supersonic Unsteady Airload

Analysis”, Journal of Aircraft, vol. 24, No. 10, 1987, pp 696-702.

3. Ashley, H., “Role of Shocks in the ‘Sub-Transonic’ Flutter Phenomenon”,

Journal of Aircraft, Vol. 17, No. 3, March, 1980, pp 187-195.

4. Baker, M. L., "CFD Based Corrections for Linear Aerodynamic Methods".

AGARD R822, October 1997

5. Baker, M. L., Yuan, K., and Goggin, P.J., “Calculation of Corrections to Linear

Aerodynamic Methods for Static and Dynamic Analysis and Design”, Proc. of the 39th AIAA/

ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference and

Exhibit, Paper AIAA-98-2072, Long Beach, California, April 20-23, 1998.

6. Baldwin, B., and Lomax, H., “Thin Layer Approximation and Algebraic Model

for Separated Turbulent Flows”, AIAA-Paper-78-257, AIAA 16th Aerospace Sciences

Meeting, Huntsville Alabama, January, 1978.

7. Ballhaus, W.F., and Goorjian, P.M., “Computation of Unsteady Transonic

Flows by the Indicial Method,” AIAA Journal, Vol. 16, No. 2, Feb. 1978, pp. 117-124.

8. Beam, R. M., and Warming, R. F., “An Implicit Factored Scheme for the

Compressible Navier-Stokes Equations”, AIAA Journal, Vol. 16, April 1978, pp. 393-402.
Chapter 7 – References 185

9. Bennet, R. M., and Edwards, J. W., “An Overview of Recent Developments in

Computational Aeroelasticity”, AIAA Paper 98-2421, 29th Fluid Dynamics Conference, June

15-18, Albuquerque, NM, 1998.

10. Bennet, R. M., Batina, J. T., and Cunningham, H. J., “Wing Flutter

Calculations with the CAP-TSD Unsteady Transonic Small-Disturbance Program”, Journal of

Aircraft, Vol. 26, No, 9, 1989, pp. 876-882.

11. Bergh, H., and Zwaan, R. J., “A Method for Estimating Unsteady Pressure

Distributions for Arbitrary Vibration Modes from Theory and form Measured Distributions

for One Single Mode”, NLR-TR F.250,1966.

12. Bertin, J.J. and Smith, M.L., Aerodynamics for Engineers, Prentice-Hall, New

Jersey, 1979, USA.

13. Bismark-Nasr, M. N., “Finite Elements in Applied Mechanics”, São Paulo,

Abete Gráfica, v.1 p. 591, 1993.

14. Bisplinghoff, R., L., Ashley, H., and Halfman, R. L., Aeroelasticity, Addison-

Wesley, 1955, Reprint Dover Publications, Inc., Mineola, N.Y., 1996.

15. Bisplinghoff, R., L., and Ashley, H., Principles of Aeroelasticity. John Wiley &

Sons, N.Y., 1962.

16. Brink-Spalink, J., and Bruns, J. M., “Correction Of Unsteady Aerodynamic

Influence Coefficients Using Experimental or CFD Data”, AIAA-2000-1498, 41st

AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and Materials Conference

and Exhibit, Atlanta, Georgia, 3-6 April, 2000.


Chapter 7 – References 186

17. Chen, P. C., Gao, X. W., and Tang, L., “An Overset Field Panel Method for

Unsteady Transonic Aerodynamic Influence Coefficients Matrix Generation”, AIAA Paper-

2004-1512, In: Proceedings of the 45th AIAA/ASME/ASCE/AHS/ASC Structures, Structural

Dynamics & Materials Conference 12th AIAA/ASME/AHS Adaptive Structures Conference,

6th AIAA Non-Deterministic Approaches Forum, 5th AIAA Gossamer Spacecraft Forum,

Palm Springs, California, USA, 19 - 22 April, 2004.

18. Chen, P.C., and Liu, D.D., “Unsteady Supersonic Computation of Arbitrary

Wing-Body Configurations Including External Stores,” Journal of Aircraft, Vol. 27, No. 2,

Feb. 1990,pp. 108-116.

19. Chen, P.C., “A Damping Perturbation Method for Flutter Solution: The g-

Method,” International Forum on Aeroelasticity and Structural Dynamics, Hampton, VA, Jun.

22- 25, 1999.

20. Chen, P.C., Lee, H.W., and Liu, D.D., “Unsteady Subsonic Aerodynamics for

Bodies and Wings with External Stores Including Wake Effect”, Journal of Aircraft, Vol. 30,

No. 5, Sep.-Oct. 1993, pp. 618-628.

21. Chen, P.C., and Liu, D.D., “A Harmonic Gradient Method for Unsteady

Supersonic Flow Calculations”, Journal of Aircraft, Vol. 22, No. 5, Sep.-Oct. 1985, pp. 371-

379.

22. Chen. P. C., Sarhaddi, D., and Liu, D. D.,2000, “Transonic Aerodynamic

Influence Coefficient Approach for Aeroelastic and MDO Applications”, Journal of Aircraft,

Vol. 37, No. 1, pp. 85-94.


Chapter 7 – References 187

23. Dau, K., “A Semi-Empirical Method for Calculating Pressures on Oscillating

Wings in Unsteady Transonic Flow”, Report No. DA-EF24-B08/92, Deutche Airbus,

October, 1992.

24. Dowell, E. H. , Bland, S. R., and Williams, M. H., “Linear/Nonlinear Behavior

in Unsteady Transonic Aerodynamics”, AIAA Journal, Vol. 21, No. 1, January, 1983, pp. 38-

46.

25. Eversman, W., and Pitt, D.M., “Hybrid Doublet Lattice/Doublet Point Method

for Lifting Surfaces in Subsonic Flow, “ Journal of Aircraft, Vol. 28, No. 9, Sept. 1991, pp.

572 – 578.

26. Farmer, M. G., and Rivera, J. A., “Experimental Flutter results of a Cambered

Supercritical Wing on a Pitch and Plunge Apparatus”, presented at Aerospace Flutter and

Dynamics Council Meeting, NASA Langley Research Center, May, 1988.

27. Fung, K.Y., and Chung, A., “Computations of Unsteady Transonic

Aerodynamics Using Prescribed Steady Pressures,” Journal of Aircraft, Vol. 20, no. 12,

December 1983, pp. 1058-1061.

28. Garner, H. C., "A practical framework for evaluation of oscillating

aerodynamic loadings on wings in supercritical flows", RAE Technical Memorandum

Structures 900, United Kingdom, 1977.

29. Giesing, J. P. , Kalman, T. P., and Rodden, W. P. “Subsonic Unsteady

Aerodynamics for General Configurations, Air Force Flight Dynamics Laboratory, AFFDL-

TR-71-5, Part 1, Vol. 1, Wright Patterson Air Force Base, Ohio, Nov. 1971, Appendix A
Chapter 7 – References 188

30. Giesing, J. P., Kalman, T. P., and Rodden, W. P., “Correction Factor

Techniques for Improving Aerodynamic Prediction Methods”, NASA-CR-144967, 1976.

31. Gordnier, R. E., and Melville, R. B., “Transonic Flutter simulations Using an

Implicit Aeroelastic Solver”, Journal of Aircraft, Vol. 37, No. 5, 2000, pp. 872-879.

32. Harder, R.L., and Desmarais, R.N., “Interpolation Using Surface Splines,”

AIAA Journal, Vol. 9, No. 2, 1972, pp. 189-191.

33. Hassig, H., J., “An Approximate True Damping Solution of the Flutter

Equation By Determinant Iteration”, Journal of Aircraft, Vol. 8, No. 11, pp. 885-889., 1971

34. Hong, M. S., Bhatia, K. G., SenGupta, G., Kim, Taehyoun, Kuruvila, G., Silva,

W. A., Bartels, R., and Biedron, R., “Simulations of a Twin-Engine Transport Flutter Model

in the Transonic Dynamics Tunnel”, Paper No. US-44 In: Proceedings of the International

Forum on Aeroelasticity and Structural Dynamics, Amsterdam, Netherlands, June 2003.

35. Houwink, R., Kraan, A. N., and Zwaan, R. J., “Wind-Tunnel Study of the

Flutter Characteristics of a Supercritical Wing”, Journal of Aircraft, Vol. 19, No. 5, 1982.

36. Hui, W.H. " Stability of Oscillating Wedges and Caret Wings in Hypersonic

and Supersonic Flows" AIAA Journal. Vol. 7, No 8, Aug 1969, pp. 1524

37. Ichikawa, I., “Doublet Strip Method for Oscillating Swept Tapered Wings in

Incompressible Flow,” Journal of Aircraft, Vol. 22, No. 11, Nov. 1985, pp. 1008 –1012.

38. Isogai, K., “Numerical Simulation of Transonic Flutter of a High-Aspect-Ratio

Transport Wing”, National Aerospace Laboratory, Japan, Report TR-776T, August, 1983
Chapter 7 – References 189

39. Jadic I, Hartley D, Giri J. “An Enhanced Correction Factor Technique for

Aerodynamic Influence Coefficient Methods”. MacNeal-Schwendler Corporation’s

Proceedings for 1999 Aerospace User’s Conference, 1999.

40. Jadic, I., Hartley, D., and Giri, J. “Generalized Aerodynamic Forces Based on

CFD and Correction Factor Techniques for AGARD wing 445.6”. AIAA-Paper-2001-1208,

In: Proceedings of the 42nd AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics,

and Materials Conference and Exhibit, Seattle, Washington, April. 16-19, 2001.

41. Jadic, I., Hartley, D., and Giri, J. “Improving the Linear Aerodynamic

Approximation in Linear Aeroelasticity”, AIAA-Paper-2000-1450, In: Proceedings of the 41st

Structures, Structural Dynamics and Materials Conference and Exhibit, Atlanta, Georgia, 3-6

April 2000.

42. Jameson, A., Schmidt, W. and Turkel, E., “Numerical Solutions of the Euler

Equations by Finite Volume Method Using Runge-Kutta Time-Stepping Schemes”, AIAA

Paper 81-1259, In: Proceedings of the AIAA 14th Fluid and Plasma Dynamics Conference,

Palo Alto, USA, 1981.

43. Jordan, P. F., “Aerodynamic Flutter Coefficients for Subsonic, Sonic and

Supersonic Flow (Linear Two-Dimensional Theory),” British Aeronautical Research Council,

London, ARCTR R&M 2932, 1957.

44. Jordan, P. F., ‘Reliable Lifting Surface Solutions for Unsteady Flow,” Journal

of Aircraft, Vol. 15, September 1978, pp. 626-633.

45. Kalman, T.P., Rodden, W.P., and Giesing, J.P., “Application of the Doublet

Lattice Method to Nonplanar Configurations in Subsonic Flow,” Journal of Aircraft, Vol. 8,

No. 6, June 1971, pp. 406 – 413.


Chapter 7 – References 190

46. Küssner, H. G., “General Airfoil Theory”, NACA – TM – 979, 1940.

47. Küssner, H. G., “General Lifting Surface Theory”, Luftfahrtförschung, Vol. 17,

1940, pp. 370-378.

48. Landahl, M. T., Unsteady Transonic Flow, Pergamon Press, New York, 1951

49. Landahl, M., T., and Stark, V.J.E., “Numerical Lifting Surface Theory –

Problems and Progress,” AIAA Journal, Vol. 6, No. 11, Nov. 1968, pp. 2049 – 2000.

50. Laschka, B., ‘Zur Theorie der Harmonishe Schwindingen Tragenden Flache

bei Unterschallanstromung”, Zetschrift für Flugwissenschaften, Vol. 11, No. 7, July 1963,

pp.265-292.

51. Lee-Rausch, E. M., and Batina, J., “Calculation of AGARD Wing 445.6 Flutter

Using Navier-Stokes Aerodynamics,” Paper AIAA-93-3476, In Proceedings. of the 11th

AIAA Applied Aerodynamics Conference, Monterey, California, 1983

52. Lessing, H. C., Troutman, J. L., and Menees, G. P., “Experimental

Determination on a Rectangular Wing Oscillating in the First Bending Mode for Mach

Numbers from 0.24 to 1.30,” NASA TND-33, December 1960.

53. Liu, D. D, “Lifting Surface Theory and Panel Methods”, Lecture Notes,

C.S.I.R., May, 1988.

54. Liu, D. D, and Winther, B. A., “Towards a Mixed Kernel Function Approach

for Unsteady Transonic Flow Analysis”, Presented at the AGARD Symposium on Unsteady

Aerodynamics, Ottawa, CA, September 26-28, 1977 ( Paper No. 12, AGARD-CPP-227).
Chapter 7 – References 191

55. Liu, D. D., “A Mixed Kernel Function Method for Unsteady Transonic Flow

Calculations”, Northrop Tech-Brief PC-20-59, March, 1976.

56. Liu, D. D., “Panel Method Lectures”, Lecture Notes, Arizona State University,

Tempe, Arizona, 1990.

57. Liu, D. D., “Computational Transonic Equivalent Strip Method for

Applications to Unsteady 3D Aerodynamics,” AIAA-83-0261, In: Proceedings of the 21st

AIAA Aerospace Sciences Meeting, Reno, Nevada, January 10-13, 1983.

58. Liu, D.D., Chen, P.C., Yao, Z.X., and Sarhaddi, D., “Recent Advances in

Lifting Surface Methods,” the Aeronautical Journal of the Royal Aeronautical Society, Vol.

100, No. 958, Oct. 1996, pp. 327-339.

59. Liu, D. D., and Hui, W.H. " Oscillating Delta Wings with Attached Shock

Waves" AIAA Journal., vol. 15, No 6, June, 1977, pp 804-811.

60. Liu, D. D., Kao, Y. F., and Fung, K. Y., “An Efficient Method for Computing

Unsteady Transonic Aerodynamics of Swept Wings with Control Surfaces”, Journal of

Aircraft, Vol. 25, No. 1, 1988, pp. 25-31.

61. Luber, W. Schmid, H., "Flutter Investigations in the Transonic Flow Regime

for a Fighter Type Aircraft", AGARD RN 703, September 1982.

62. Mabey, D., “The Physical Phenomena Associated with Unsteady Transonic

Flow”, in “Unsteady Transonic Aerodynamics”, Edited by David Nixon, American Institute

of Aeronautics and Astronautics, Washington, D.C., 1989, Chapter 1, pp. 1-55.

63. Malone, J.B., and Ruo, S.Y., “LANN Wing Test Program: Acquisition and

Application of Unsteady Transonic Data for Evaluation of Three-Dimensional Computational


Chapter 7 – References 192

Methods,” Air Force Flight Dynamics Laboratory, AFWAL-TR-83-3006, Wright Patterson

Air Force Base, Ohio, February 1983.

64. McCain, W. E., “Measured and Calculated Airloads on a Transport Wing

Model”, Journal of Aircraft, Vol. 22, No. 4,1985 .

65. Mello, O. A. F., and Azevedo, J. L. F., “Simulation of High Angle-of-Attack

Flow over a Hemisphere-Cylinder”, In: Proceedings of the 8th Brazilian Congress of

Engineering and Thermal Sciences, ENCIT 2000 -, Porto Alegre, Rio Grande do Sul, 2000.

66. Mello, O. A. F., and Sankar, L. N., “Computation of Unsteady Transonic Flow

over a Fighter Wing Using a Zonal Navier-Stokes/Full-Potential Method,” International

Journal for Numerical Methods in Fluids, Vol. 29, No. 5, March 1999, pp. 575-585

67. Mello, O., A., F., “An Improved Hybrid Navier-Stokes/Full-Potential Method

for Computation of Unsteady Compressible Viscous flows”, PhD Thesis, Georgia Institute of

Technology, November, 1994

68. Miles, J.W., “Potential Theory of Unsteady Supersonic Flow,” Cambridge

Univ. Press, London, 1959.

69. Palacios, R., Climent, H., Karlsson, A., and Winzell, B., “Assessment of

Strategies for Correcting Linear Unsteady Aerodynamics Using CFD or Test Results”, Proc.

of the CEAS/AIAA International Forum on Aeroelasticity and Structural Dynamics, Madrid,

Spain, June 2001, pp. 195-210

70. Pitt, D.M., and Goodman, C.E., “Flutter Calculations Using Doublet Lattice

Aerodynamics Modified by the Full Potential Equations”, Proc. of the 28th AIAA/
Chapter 7 – References 193

ASME/ASCE/AHS Structures, Structural Dynamics, and Materials Conference, Paper AIAA-

87-0882-CP, Monterey, California, USA, 1987.

71. Pulliam, T. H., and Chaussee, D. S., “A Diagonal Form of an Implicit

Approximate-Factorization Algorithm”, Journal of Computational Physics, Vol. 39, 1981,

pp.347-363.

72. Pulliam, T. H., and Steger, J. L., “Implicit Finite-Difference Simulations of

Three-Dimensional Compressible Flow”, AIAA Journal, Vol. 18, 1980, pp. 159-167.

73. Rodden, W. P., and Revell, J. D., “Status of Unsteady Aerodynamic Influence

Coefficients”, Institute of Aeronautical Sciences, Fairchild Fund Paper FF-33, 1962.

74. Rodden, W. P., Giesing, J. P., and Kalman, T. P., “New Developments and

Applications of the Subsonic Doublet Lattice Method for Non-Planar Configurations”,

AGARD Symposium – Unsteady Aerodynamics for Aeroelastic Analysis of Interfering

Surfaces, Paper no 4, Tonsberg, OsloFjorden, Norway, November 4-3, 1970. (see also

Douglas Paper 5826)

75. Rodden, W. P., and Bellinger, E. D., “Aerodynamic Lag Functions, Divergence

and the British Flutter Method,” Journal of Aircraft, Vol. 19, No. 7, July 1982.

76. Rodden, W. P., and Bellinger, E. D., “Unrestrained Aeroelastic Divergence in a

Dynamic Stability Analysis,” Journal of Aircraft, Vol. 19, No. 9, September 1982.

77. Rodden, W. P., Harder, R.L., and Bellinger, E., D., “Aeroelastic Addition to

NASTRAN,” NASA CR 3094, March 1979.

78. Rodden, W. P., and Johnson, E. H., “MSC/NASTRAN Aeroelastic Analysis -

User’s Guide Version 68”, The MacNeal-Schwendler Corporation, 1994, USA.


Chapter 7 – References 194

79. Rodden, W. P., Giesing, J. P., and Kalman, T.P., “New Development and

Applications of the Subsonic Doublet-Lattice Method for Nonplanar Configurations,”

AGARD Conf. Proc. No. 80-71, Part II, No. 4, 1971.

80. Rodden, W.P., Giesing, J.P., and Kalman, T.P., “Refinement of the Nonplanar

Aspects of the Subsonic Doublet Lattice Lifting Surface Method,” Journal of Aircraft, Vol. 9,

Jan 1972, pp. 69 – 73.

81. Rodden, W.P., Taylor, P.F., and McIntosh, S.C., Jr., “Further Refinement of

the Subsonic Doublet Lattice Method,” Journal of Aircraft, Vol. 35, No. 5, 1998, pp. 720–

727.

82. Rodden, W.P., Taylor, P.F., McIntosh, S.C., Jr., and Baker, M.L, “Further

Convergence Studies of the Enhanced Subsonic Doublet Lattice Oscillatory Lifting Surface

Method,” In: Proceedings of the International Forum on Aeroelasticity and Structural

Dynamics, Vol. III, Rome, Italy, 1997, pp. 401-408.

83. Rumsey, C., Biedron, R., Thomas, J., “CFL3D: Its History and some Recent

Applications”, NASA-TM 112861, May, 1997; presented at the “Godunov’s Method for Gas

Dynamics” Symposium, Ann Arbor, MI, May, 1-2, 1997.

84. Sankar, L. N, Bharadvaj, B. K., and Tsung, F.-L., “A Three Dimensional

Navier-Stokes/Full-Potential Coupled Analysis for Viscous Transonic Flows,” AIAA Journal,

Vol. 31, No. 10, October 1993 .

85. Sankar, L. N., and Kwon, O. J., “High-Alpha Simulation of Fighter Aircraft,”

Proceedings of the NASA High Angle-of-Attack Technology Conference, Vol. 1, NASA

Langley Research Center, Hampton, VA, NASA CP-3149, Pt. 2, 1990, pp. 689–702
Chapter 7 – References 195

86. Schulze, E., and Vogel, S., “Comparison of Two Potential Flow Methods for

Transonic Flutter Analysis”, Rept - 89-005, 1989

87. SenGupta, G., “Evaluation of the Dau-Garner Method for Predicting Unsteady

Pressures in Transonic Flow”, Paper presented at the International Symposium on

Aeronautical Science & Technology, Jakarta, Indonesia, June 24-27, 1996 – Paper ISASTI-

96-1.3.4

88. Silva, R. G. A., “State Space Aeroelastic Analysis Applied to Fixed Wing

Aircraft”, [“Análise Aeroelástica no Espaço de Estados Aplicada a Aeronaves de Asa Fixa”],

Master Thesis, University of São Paulo, São Carlos, 1994. (in Portuguese).

89. Silva, R. G. A., and Mello, O. A. F., “ Prediction of Transonic Flutter Using

NASTRANTM with Aerodynamic Coefficients Tuned to Navier-Stokes Computations” ,

CEAS/AIAA/ICASE/NASA Langley International Forum on Aeroelasticity and Structural

Dynamics, Williamsburg, Virginia, 1999.

90. Silva, R. G. A., Mello, O. A. F., and Azevedo, J. L. F “A Navier-Stokes Based

Study into Linearity in Transonic Flow fort Flutter Analysis, AIAA Paper 2002-2971, In:

Proceedings of the 32nd AIAA Fluid Dynamics Conference, St. Louis, USA, June 2002.

91. Silva, R. .G. A., Mello, O. A. F. and, Azevedo, J. L. F. , ”A Numerical Study

into Shock Displacement Effects for Aeroelastic Analysis in Transonic Flow” In: Proceedings

of the 9th Brazilian Congress of Thermal Engineering and Sciences - ENCIT 2002 ,

Caxambú, Minas Gerais, October 2002.

92. Silva, R. G. A., Mello, O. A. F., and Azevedo, J. L. F., “Transonic Flutter

Calculations Based on Assumed Mode Shapes Corrections”, In: Proceedings. of the


Chapter 7 – References 196

CEAS/AIAA International Forum on Aeroelasticity and Structural Dynamics, Madrid, Spain,

2001, pp. 183-194.

93. Soviero, P. A. O., and Bortolus, M. V., “Generalized Vortex Lattice Method

for Oscillating Lifting Surfaces in Subsonic Flow”, AIAA Jounal, Vol. 30, No. 11, 1992, pp.

2723-2729.

94. Spalart, P., and Allmaras, S., “A One Equation Turbulence Model for

Aerodynamic Flows”, AIAA-Paper-92-0439, In: Proceedings of the AIAA 30th Aerospace

Sciences Meeting and Exhibit, Reno, Nevada, January 6-9, 1992.

95. Suciu, E., “A Comparison Between the Aerodynamic Pressures Factoring and

the Aerodynamic Derivatives factoring Methods for the Doublet lattice Program”, In:

Proceedings of the 3rd Worldwide Aerospace Conference and technology Showcase ,

September, 24-26, Toulouse, France, 2001.

96. Suciu, E., “A General Aerodynamic Derivatives Factoring Method for the

MSC.Nastran DLM Capable of Controlling Al Lifting Surfaces Aerodynamic Forces and

Moments, Including all Interference Effects”, Presented at the Aerospace Flutter and

Dynamics Council Spring 2003 Meeting, Dayton, Ohio, 7-9 May, 2003

97. Suciu, E., Glaser, J., and Coll, R., “Aerodynamic Derivatives Factoring

Scheme for the MSC/NASTRAN Doublet Lattice Program”, presented at the

MSC/NASTRAN Worlds Users’ Conference, Los Angeles – CA ,1990.

98. Sulaeman, E., “Effect of Compressive Force on Aeroelastic Stability of a Strut-

Braced Wing”, PhD thesis, Virginia Polytechnic Institute and State University, Blacksburg,

Virginia, USA, November 2001.


Chapter 7 – References 197

99. Theodorsen, T., “General Theory of Aerodynamic Instability and the

Mechanism of Flutter,” NACA Report No. 496, 1940.

100. Thomas, J. P., Dowell, E. H., and Hall, K. C., “A Harmonic Balance Approach

for Modeling Three-Dimensional Non-Linear Unsteady Aerodynamic and Aeroelasticity”,

Proceedings of the ASME International Mechanical Engineering Conference and Exposition,

November, 17-22, 2002, New Orleans, Louisiana, USA. Paper No. IMECE-2002-32535.

101. Thomas, J. L., Krist, S. L., and Anderson, W. K., “Navier-Stokes

Computations of Vortical Flows over Low-Aspect-Ratio Wings”, AIAA Journal, Vol. 28, No,

2, pp. 205-212, 1990.

102. Tijdeman, H., van Nunen, J. W. G., Kraan, A. N., Persoon, A. J., Poestkoke,

R., Roos, R., Schippers, P., and Siebert, C. M., “Transonic Wind Tunnel Tests on an

Oscillating Wing with External Stores, Part II: Clean Wing,” Air Force Flight Dynamics

Laboratory, AFFDL-TR-78-194, Part II, Wright-Patterson AFB, Ohio, 1979. (Also NLR-TR-

78106U, Part II)

103. Ueda, T., and Dowell, E. H., “A New Subsonic Method for Lifting Surfaces in

Subsonic Flow,” AIAA Journal, Vol. 20, No. 3, March 1982, pp. 348 – 355.

104. Van Zyl, L.H., “Arbitrary Accuracy Integration scheme for the Subsonic

Doublet Lattice Method,” Journal of Aircraft, Vol. 35, No. 6, 1998, pp. 975-977.

105. Van Zyl, L.H., “Application of the Subsonic Doublet Lattice Method to Delta

Wings,” Journal of Aircraft, Vol. 36, No. 3, 1999, pp. 609-611.

106. Vepa, R., “Finite State Modeling of Aeroelastic System”, NASA CR-2779,

Feb. 1977.
Chapter 7 – References 198

107. Watkins, C. E., Runyan, H. L., and Woolston, D. S., “On the Kernel Function

of the Integral Equation Relating the Lift and Downwash Distributions of Oscillating Finite

Wings in Subsonic Flow”, NACA Report – 1234, 1955, pp. 1-11.

108. Whitlow, W., Jr., “Computational Unsteady Aerodynamics for Aeroelastic

Analysis,” NASA TM-100523, December, 1987.

109. Winzel, B., "Recent Applications of Linear and Nonlinear Unsteady

Aerodynamic Aerodynamics for Aeroelastic Analysis", AGARD CP 507 - Transonic

Unsteady Aerodynamics and Aeroelasticity, San Diego, California, October 7-11, 1991

110. Yates Jr., E. C., “AGARD Standard Aeroelastic Configurations for Dynamic

response I-Wing 445.6”, AGARD Report No. 765, 1988.

111. Yates, E. C. Jr., “Modified Strip Analysis Method for Predicting Wing Flutter

at Subsonic to hypersonic Speeds”, Journal of Aircraft, Vol. 3, No. 1, Jan-Feb, 1966

112. Yates, E.C., Jr., Land, N.S., and Foughner, J.T., Jr., “Measured and Calculated

Subsonic and Transonic Flutter Characteristics of a 45º Sweptback Wing Planform in Air and

in Freon-12 in the Langley Transonic Dynamics Tunnel,” NASA TN-D-1616, Mar. 1963.

113. Yonemoto, K., "A practical method predicting transonic wing flutter

phenomena", 14th International Council of the Aeronautical Sciences Congress (ICAS-1984),

Toulouse, France, 1984, pp 724-732.

114. Zona Technology, “ZAERO Software System- Theoretical Manual – V. 6.5”,

Scottsdale, Arizona, October, 2003.

115. Zwaan, R. J., “Verification of Calculation Methods for Unsteady Airloads in

the Prediction of Transonic Flutter”, Journal of Aircraft, Vol. 22, No. 10, 1985.
APPENDIX A

GIESING´S FORCE MATCHING METHOD

The procedure to be revisited herein is the force matching method presented by

Giesing et al.(1976). The objective of this method is to perform the correction of the AIC

matrix to match experimental loading distributions of the lifting surface under investigation.

This method was originally proposed based on the requirement of a safe prediction of the

control surface flutter.

This method assumes that there is a weighting function operator which matches the

experimental pressures based on the theoretical ones as:

{∆C p }e = Wp  {∆C p } . (A.1)

In many cases what is available is experimental loads instead of the pressure distributions. In

such case, the experimental loads can be written as follows:

{Ce } = [ S ]{∆C pe } , (A.2)

where {Ce } is the load vector and [S] the integration matrix which relates the pressures to

the loads. Replacing (A1) in (A2) yields:

{∆Ce } = [ S ] W p  {∆C p } . (A.3)

This equation will be solved for the correction matrix [Wp].

As there are many more unknowns than equations, the way to solve this system is

using the Least Squares Method. Considering that the deviation can be written as :

{ε p } = W p − I  . (A.4)

The weighted sum of the square of the deviation {ε p } is:


Appendix A - Giesing´s Force Matching Method. 200

= {ε p } Tp  {ε p }
H
∑T ε 2
p p , (A.5)

shall be a minimum, and the error frictional fp is defined as:

1
{ε p } Tp  {ε p }
H
fp = . (A.6)
2

Replacing eqn. (A4) into eqn. (A1), the loads vector can be written as:

{Ce } = [ S ]{∆C p } + [ S ] ∆C p  {ε p } = {C } + [ S ] ∆C p  {ε p } . (A.7)

One should note that the integration matrix [S] relates pressure to loads, and these

loads can be either panel or strip loads. In a practical sense, the experimental loading data is

acquired in a stripwise way; the integration matrix needs to be properly defined as an operator

which relates panel pressures to strip loads. Then, rewriting the pressures vector as a diagonal

matrix, integrated load matrix will be defined as:

 S p  = [ S ]  ∆C p  . (A.8)
   

Recalling equation (A7), one should note that

{∆Ce } = {Ce } − {C } , (A.9)

{∆Ce } =  S p  {ε p } . (A.10)

since the variation of the error function as:

δ f p = {ε p } Tp  {δε p }


H
, (A.11)

the variation of the experimental constraints is

{δ∆Ce } =  S p  {δε p } = 0 . (A.12)

One should note that equations (A11) and (A12) compose a linear system of equations,

δ f p + {λ p } {∆Ce } = 0
H
, (A.13)

with the vector {λ p} as the associated Lagrange multipliers. Thus, rewriting this system of

equations substituting (A11) and (A12) into (A13) we get:


Appendix A - Giesing´s Force Matching Method. 201

(Tp ) + {λ p }  S p  {δε p } = 0
H H
δε p  , (A.14)
   

as δε p is an arbitrary non zero value, we have:

{λ p } = 0
H H
 ε p  Tp  +  S p  , (A.15)
     

or

{λp } = − Tp  ε p  .


H
S p  . (A.16)
 

Recalling equation (A10) given by {∆Ce } =  S p  {ε p } , the Lagrange rule replace is given by:

{λ p } = −( S p  Tp  )


−1 H −1
S p  {∆Ce } , (A.17)
 

and the {Wp} is obtained from


−1
{ε p } = − Tp  {λ p }
H
S p  , (A.18)
 

W p  = [ I ] +  ε p  . (A.19)
   

The weighting function Tp is arbitrary, at least near to the positive. The authors suggested in

their work that

Tp  = [ AIC ][ I ] , (A.20)


 

with [Wp] matrix defined, it is possible to restore the loads, but not necessarily the pressures.
APPENDIX B

THE DAU-GARNER METHOD

The Dau-Garner method (Dau, 1992, and Garner, 1977) is a procedure whose main

concern is to look for a rational form to compute the transonic unsteady pressure amplitudes

and phases. This method is one of the most employed in the aeronautical industry. However

here the objective is to show mathematically its failing with regard to the pressures phase

computation.

This method is based on a set of assumptions added by semi-empirical hypothesis,

which were introduced by Garner (1977). It is assumed that the flow around the lifting surface

can be described with sufficient accuracy by perturbation velocity potentials, and adiabatic

relationships. These assumption are valid for the cases of transonic flows with weak shock

waves. The displacement amplitudes are sufficiently small, thus the time dependent

perturbations of the velocities are much smaller that the steady state streamwise component of

the velocity, in the “x” direction. Furthermore, it is assumed that the relationship between the

velocity potentials, remains the same whether the flow is transonic or purely subsonic. Thus,

the ratio between the unsteady and steady gradients of the velocity potential should be written

as:

φ xnl ( ik ) φ x ( ik )
R ( ik ) = = . . (B.1)
φ xnl ( 0 ) φ x ( 0 )

And finally, the ratio of the quasi-steady transonic pressures to its corresponding linear

pressures is the same for all comparable vibration modes of the lifting surface.
Appendix B - The Dau-Garner Method 203

With the aid of these assumptions and considering the Euler momentum equation it is

possible to write the unsteady pressure distribution as a function of the velocity potential

gradients, after linearization for small disturbances, leading to:

C pnl ( ik ) = −2 (φ xnl ( ik ) + ikφ nl ( ik ) ) , (B.2)

where C pnl ( ik ) and φ xnl ( ik ) are the complex amplitudes of the first harmonic of the pressure

and velocity potential gradient, respectively. The complete derivation of Eq. (B.2), departing

from the Euler momentum equations is found in the work of Dau (1992). Recalling the Euler

momentum equation, one should note that the steady pressures C pnl ( 0 ) are related to thee

steady velocity potential φ xnl ( 0 ) as:

C pnl ( 0 ) C pnl ( 0 )
φ (0) = −
nl
x ⇒ φ (ik ) = −
nl
x R(ik ) . (B.3)
2 2

Looking at Eq. (E.2), is possible to see that the computation of the unsteady pressure

distribution should be performed from the following relationship:


X

( ik ) = R ( ik ) C ( 0 ) + ik ∫ R ( ik ) C pnl ( 0 )dξ
nl nl
C p p . (B.4)
X L. E .

Finally, the unsteady pressures computation is performed by the Dau-Garner method (Dau,

1992) satisfying relation (B.4).

The failing with regard to the pressures phase computation, identified by SenGunpta

(1996), has as explanation. Introducing the approximation of the potential for low reduced

frequencies as φ = φR + ikφ I , that is, a simple first order expansion series, the unsteady

pressure relationship between the potential and the pressures can be written as

C pnl ( ik ) = −2 φ xnlR + ikφ xnlI + ik (φRnl + ikφInl )  ⇒


(B.5)
( )
C pnl ( ik ) = −2 φ xnlR + ikφ xnlI + ikφ Rnl − k 2φInl  = −2 φ xnlR + ik φ xnlI + φ Rnl 
 
Appendix B - The Dau-Garner Method 204


where φR = φ xR dx . That is, the out of phase component contribution vanishes because the

low reduced frequency value and the out of phase correction is simply reduced to a phase shift

of 180 degrees, since the velocity potential gradient ratio does not change the imaginary part

of the potential. This is probably the main cause of the pressure phases differences detected

by SenGupta (1996), in about 50 degree.


FOLHA DE REGISTRO DO DOCUMENTO
1. 2. 3. 4.
CLASSIFICAÇÃO/TIPO DATA DOCUMENTO N° N° DE PÁGINAS

TD 02 de Julho de 2004 CTA/ITA-IEA/TD-001/2004 233


5.
TÍTULO E SUBTÍTULO:
A study on correction methods for aeroelastic analysis in transonic flow
6.
AUTOR(ES):

Roberto Gil Annes da Silva

7. INSTITUIÇÃO(ÕES)/ÓRGÃO(S) INTERNO(S)/DIVISÃO(ÕES):
Instituto Tecnológico de Aeronáutica. Divisão de Engenharia Aeronáutica – ITA/IEA
8.
PALAVRAS-CHAVE SUGERIDAS PELO AUTOR:

1. Aeroelasticidade . 2. Aerodinâmica não estacionária. 3. Escoamento transônico. 4. Métodos de


correção. 5. Aerodinâmica computacional.
9.PALAVRAS-CHAVE RESULTANTES DE INDEXAÇÃO:
Aeroelasticidade; Aerodinâmica não-estacionária; Escoamento transônico; Modelos matemáticos;
Correção; Dinâmica dos fluidos computacional; Sustentação aerodinâmica; Mecânica dos fluidos; Física
10.
APRESENTAÇÃO: X Nacional Internacional
ITA, São José dos Campos, 2004, 233 páginas
11.
RESUMO:

The work presents a study of correction techniques to compute unsteady transonic pressure
distributions and aeroelastic stability in this flow regime. The methodologies herein investigated are
based on corrections of pressure distributions by the weighting of the lifting surface self-induced
downwash, resulting from aeroelastic structural displacements or prescribed motions. A number
approaches were investigated.
An investigation into the linear/nonlinear behavior of unsteady transonic flows was also conducted. It
was concluded from such investigation that unsteady transonic flows present a linear behavior with
respect to small aeroelastic structural displacements around a steady nonlinear mean flow. Such behavior
is the basis for further development of downwash correction methods.
The correction of pressure distributions through the weighting of the lifting surface self-induced
downwash is also known as downwash weighting method. This method has been enhanced leading to a
new downwash correction technique. The procedure may be divided in two steps, where the first step is a
nonlinear steady mean flow correction, with nonlinear pressure differences considered as reference
conditions to correct the self induced downwash. The second step is the correction of the unsteady
component of the downwash, where the corresponding reference unsteady pressure differences are
predicted by a linear aerodynamic model, based on the potential flow equations.
This extended downwash correction method led to a rational formulation named as “successive
kernel expansion method” (SKEM). The unsteady pressures and aeroelastic stability boundaries
computations using such method led to good agreement with experimental measurements. This procedure
is a rapid form to compute the transonic flutter speed boundaries, compared to computational
aeroelasticity and experimental techniques.

12.
GRAU DE SIGILO:

(X ) OSTENSIVO ( ) RESERVADO ( ) CONFIDENCIAL ( ) SECRETO

Você também pode gostar