You are on page 1of 104

Structural Design of a Modular Aircraft

Mafalda Joana Bayan Ferreira Magro

Thesis to obtain the Master of Science Degree in

Aerospace Engineering

Supervisors: Prof. Fernando José Parracho Lau


Dr. Frederico José Prata Rente Reis Afonso

Examination Committee
Chairperson: Prof. Filipe Szolnoky Ramos Pinto Cunha
Supervisor: Prof. Fernando José Parracho Lau
Member of the Committee: Prof. Afzal Suleman

February 2020
ii
Acknowledgments

I would first like to thank my thesis supervisors, Professor Fernando Lau and Dr. Frederico Afonso for the
the invaluable advice, wisdom and guidance. A special thanks to Dr. Frederico Afonso for his patience,
for the permanent availability to help and for believing in me and in my work.
I cannot forget to thank Professor Hugo Policarpo who was always available to help me and for
always making things look easier.
I also cannot forget to thank Student Support Office (NAPE) and to all the friends I met while working
on it, which played an important role in my academic life.
I want to thank all my dear friends, but principally to Inês, Nicolau and Samuel who had the patience to
support me and to listen to my dramas, not only during all this years, but specially during the development
of this thesis.
Finally, a huge gratitude to my family: to my Mother, to my Father, to my Sister and to Sérgio. This
was not possible without them! To my Mother and to my Father I want to express my gratitude for all
the opportunities they gave to me, the help they provide me to pursuit my dreams and my goals, for all
the love and patience. I hope I can make you proud! To my beloved Sérgio, thank you for the endless
support throughout the years, for all the patience, for the encouragement to finish this challenge, for
helping me dream and achieve my goals and, more importantly, for always being there for me!
To everyone I met along this path, that somehow helped, motivated and inspired me, I want to say
thank you. It was a hell of a climb, but the view is incredible!

iii
iv
Resumo

Nos dias de hoje, o sector dos transportes, em especial o sector aeroespacial, enfrenta diversos de-
safios para conseguir satisfazer a crescente procura por soluções de mobilidade, ao mesmo tempo
que tenta reduzir o trafego, melhorar a segurança e reduzir e/ou acabar com as emissões de gases
poluentes para a atmosfera. O interesse da sociedade por aviões com motores elétricos ou hı́bridos e
com capacidades VTOL tem aumentado, uma vez que este tipo de veı́culos tem o potencial para rev-
olucionar o transporte aéreo, permitindo uma maior flexibilidade no que diz respeito às infraestruturas
aeroportuárias e, consequentemente, uma maior facilidade de mobilidade. Dito isto, este trabalho con-
siste no desenvolvimento de um design do sistema de sustentação (asa e estabilizador) de um avião
modular, com capacidades VTOL, inserido no projeto Flexcraft, para as suas condições de voo crı́ticas.
Através da utilização do software comercial ANSYS, é possı́vel realizar análises de elementos finitos e
desenvolver uma metodologia de parametrização da estrutura de forma a minimizar o peso, sem com-
prometer a sua integridade estrutural, isto é, não ultrapassando a tensão de cedência do material em
estudo e a deflexão máxima definida para a asa e para o estabilizador.

Palavras-chave: VTOL, FEA, FEM, Parametrização, Flexcraft

v
vi
Abstract

Nowadays, the transportation sector faces a lot of challenges in meeting the growing demand for com-
mon passengers mobility while reducing traffic, improving safety and mitigating emissions. Automated
driving and electrification are disruptive technologies that may contribute to these goals, but they are
limited by congestion on existing roadways and land-use constraints. Due to this, the interest of society
in electric or hybrid Vertical Take-Off and Landing (VTOL) aircraft, that can basically act as flying cars,
has grown significantly. This peculiar type of aircraft may overcome the limitations of congestion and
land-use by enabling urban and regional aerial travel services with the need of much smaller infras-
tructures when compared with conventional aircraft. Said so, this work consists in the development of
a parametric design methodology of the lifting system (wing-tail) of a VTOL aircraft, on behalf of the
Flexcraft’s project, for its critical flight conditions. Using the commercial software ANSYS it is possible
to perform finite element analyzes and develop a methodology to perform structural parametric studies
with the main objective of minimizing the weight without compromising its structural integrity, i.e. not
exceeding the material yield strength and the maximum deflection defined for the wing and stabilizer.

Keywords: VTOL, FEA, FEM, Parameterization, Flexcraft

vii
viii
Contents

Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii
Resumo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v
Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii
List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi
List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xiii
Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xvii

1 Introduction 1
1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Topic Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.3 Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.4 Thesis Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

2 Historical Review 5
2.1 Brief History of Aviation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.2 STOL and VTOL Aircraft . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.2.1 Lighter-Than-Air Aircraft . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2.2 Rotorcraft . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.2.3 Powered Lift . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.2.4 Aircraft Representation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

3 Flexcraft Conceptual Design 15


3.1 Project Requirements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.2 Mission Profile . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.3 Concept Generation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.3.1 Utility Configuration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.3.2 Tandem Wing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.3.3 Flying Wing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.3.4 Configuration Choice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.4 Aircraft Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

ix
4 Structural Analysis 21
4.1 V-n and Gust Diagrams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
4.1.1 V-n Diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
4.1.2 Gust Diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
4.2 Finite Element Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.2.1 Finite Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
4.2.2 Finite Element Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
4.2.3 Aerodynamic Loading . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
4.2.4 Structural Improvement Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4.2.5 Parametric Constrains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
4.2.6 Material Selection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

5 Results 43
5.1 V-n and Gust Diagrams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
5.2 Convergence of the Geometric Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
5.3 Beam Geometric Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
5.3.1 Initial Variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
5.3.2 Parametrization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
5.3.3 Final Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
5.4 Shell Geometric Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
5.4.1 Initial Variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
5.4.2 Parametrization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
5.4.3 Final Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
5.5 Beam-Shell Geometric Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
5.5.1 Initial Variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
5.5.2 Parametrization Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
5.5.3 Final Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
5.6 Final Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

6 Conclusions 79
6.1 Achievements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
6.2 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

Bibliography 81

x
List of Tables

3.1 Project Requirements. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15


3.2 Utility VTOL concept characteristics. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

4.1 Inputs for the Beam Model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30


4.2 Cross Sections of the Beam structures in the BEAM model. . . . . . . . . . . . . . . . . . 33
4.3 Inputs for the Shell Model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
4.4 Inputs for the Beam-Shell Model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
4.5 Cross Sections of the Beam structures. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
4.6 Aluminium alloy properties. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

5.1 BGM - Initial Variables. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45


5.2 BGM - Constrains for Initial Variables. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
5.3 BGM - Variables for the optimization process. . . . . . . . . . . . . . . . . . . . . . . . . . 47
5.4 BGM - Final values for the variables. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
5.5 BGM - Final results. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
5.6 SGM - Initial Variables. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
5.7 SGM - Constrains for Initial Variables. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
5.8 SGM - Variables for the optimization process. . . . . . . . . . . . . . . . . . . . . . . . . . 56
5.9 SGM - Final values for the variables. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
5.10 SGM - Final results. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
5.11 BSGM - Initial Variables. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
5.12 BSGM - Constrains for Initial Variables. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
5.13 BSGM - Inputs for the optimization process. . . . . . . . . . . . . . . . . . . . . . . . . . . 66
5.14 BSGM - Final values for the variables. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
5.15 BSGM - Final results. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
5.16 Comparison of the results obtained for the 3 Geometric Models. . . . . . . . . . . . . . . 76
5.17 Time spent per simulation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
5.18 Advantages and Disadvantages of each Geometric Model. . . . . . . . . . . . . . . . . . 78

xi
xii
List of Figures

2.1 Helical air screw flying machine designed by Leonardo Da Vinci. . . . . . . . . . . . . . . 5


2.2 Wright brothers’ flight. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.3 Boeing 247. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.4 Heinkel He-178, the first aircraft powered by turbojet. . . . . . . . . . . . . . . . . . . . . . 7
2.5 Iconic aircraft through the years. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.6 Illustration of VTOL and STOL aircraft. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.7 Disaster of the Hinderburg dirigible. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.8 Convertiplane aircraft. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.9 Tail-Sitter aircraft. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.10 Vectored Thrust aircraft. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.11 Example of Lift Jet aircraft: Yakolev Yak-38. . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.12 Lift Fans aircraft. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.13 Schemes for several types of aircraft. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

3.1 Mission profiles. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16


3.2 Flexcraft Top View: Take-off on the left and Cruise on the Right. . . . . . . . . . . . . . . . 19
3.3 Flexcraft Front View: Take-off on the left and Cruise on the Right. . . . . . . . . . . . . . . 19
3.4 Flexcraft Dimensions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

4.1 Principal aerodynamic forces on an aircraft during flight. . . . . . . . . . . . . . . . . . . . 21


4.2 V-n diagram obtained for the Airbus A320. . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
4.3 Variation of total gust velocity with altitude. . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.4 Gust diagram obtained for the Airbus A320. . . . . . . . . . . . . . . . . . . . . . . . . . . 25
4.5 BEAM188 (on the left) and BEAM189 (on the right). . . . . . . . . . . . . . . . . . . . . . 27
4.6 SHELL181 (on the left) and SHELL281 (on the right). . . . . . . . . . . . . . . . . . . . . 28
4.7 MASS21. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
4.8 Beam Geometric Model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
4.9 Identification of variables RibsIn, RibsOut, t in, t out; Cross Sections 1, 2 and 3 and
MASS21 FE. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
4.10 Identification Cross Sections 3, 4, 5 and 6. . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
4.11 Identification of variables RibsIn, RibsOut, t in, t out,t wingshell and t spar. . . . . . . . . 34

xiii
4.12 Identification of variables t rearboom, t vertical,t skin tail and t spars tail. . . . . . . . . . 35
4.13 Beam-Shell Model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
4.14 Identification of variables RibsIn, RibsOut, t in, t out and Cross Section 1. . . . . . . . . . 37
4.15 Identification of variables t wingshell, t skin tail and Cross Sections 2 and 3. . . . . . . . . 37
4.16 MDO framework results. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4.17 Aerodynamic loads applied in the BGM (n=-1.8). . . . . . . . . . . . . . . . . . . . . . . . 40
4.18 Structural Design improvement model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

5.1 V-n and Gust Diagrams obtained for the Flexcraft at cruise altitude. . . . . . . . . . . . . . 43
5.2 Convergence study for the Beam, Shell and Beam-Shell Model . . . . . . . . . . . . . . . 44
5.3 BGM - Stress and Displacement Distribution for the Initial Variables. . . . . . . . . . . . . 46
5.4 BGM - variation of constrains and weight with t wingboom and chord percentage variables. 47
5.5 BGM - variation of constrains and weight with t rearboom and size rearboom variables. . 48
5.6 BGM - variation of constrains and weight with t in variable. . . . . . . . . . . . . . . . . . . 49
5.7 BGM - variation of constrains and weight with t spars tail and size spars tail variables. . . 50
5.8 BGM - variation of constrains and weight with t out variable. . . . . . . . . . . . . . . . . . 51
5.9 BGM - variation of constrains and weight with t vertical and size vertical variables. . . . . 51
5.10 BGM - variation of constrains and weight with t ribs tail variable. . . . . . . . . . . . . . . 52
5.11 BGM - Stress and Displacement Distribution, for n = −1.8, using the final values of the
variables. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
5.12 BGM - Stress and Displacement Distribution, for n = 4.4, using the final values of the
variables. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
5.13 SGM - Stress and Displacement Distribution for the Initial Variables. . . . . . . . . . . . . 56
5.14 SGM - variation of constrains with t wingshell variable. . . . . . . . . . . . . . . . . . . . . 57
5.15 SGM - variation of constrains with t skin tail variable. . . . . . . . . . . . . . . . . . . . . . 58
5.16 SGM - variation of constrains with chord percentage variable. . . . . . . . . . . . . . . . . 58
5.17 SGM - variation of constrains with t in variable. . . . . . . . . . . . . . . . . . . . . . . . . 59
5.18 SGM - variation of constrains with t out variable. . . . . . . . . . . . . . . . . . . . . . . . 60
5.19 SGM - variation of constrains with t vertical variable. . . . . . . . . . . . . . . . . . . . . . 60
5.20 SGM - variation of constrains with t rearboom variable. . . . . . . . . . . . . . . . . . . . . 61
5.21 SGM - variation of constrains with t spars tail variable. . . . . . . . . . . . . . . . . . . . . 62
5.22 SGM - variation of constrains with t ribs tail variable. . . . . . . . . . . . . . . . . . . . . . 62
5.23 SGM - Stress and Displacement Distribution, for n = −1.8, using the final values of the
variables. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
5.24 SGM - Stress and Displacement Distribution, for n = 4.4, using the final values of the
variables. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
5.25 BSGM - Stress and Displacement Distribution for the Initial Variables. . . . . . . . . . . . 66
5.26 BSGM - variation of constrains with t wingshell variable. . . . . . . . . . . . . . . . . . . . 67
5.27 BSGM - variation of constrains with t wingboom and chord percentage variables. . . . . . 68

xiv
5.28 BSGM - variation of constrains with t skin tail variable. . . . . . . . . . . . . . . . . . . . . 68
5.29 BSGM - variation of constrains with t reinforcement variable. . . . . . . . . . . . . . . . . 69
5.30 BSGM - variation of constrains with t rearboom and size rearboom variables. . . . . . . . 70
5.31 BSGM - variation of constrains with t spars tail and size spars tail variables. . . . . . . . . 71
5.32 BSGM - variation of constrains with t in variable. . . . . . . . . . . . . . . . . . . . . . . . 71
5.33 BSGM - variation of constrains with t vertical and size vertical variables. . . . . . . . . . . 72
5.34 BSGM - variation of constrains with t out variable. . . . . . . . . . . . . . . . . . . . . . . 73
5.35 BSGM - variation of constrains with t ribs tail variable. . . . . . . . . . . . . . . . . . . . . 74
5.36 BSGM - Stress and Displacement Distribution, for n = −1.8, using the final values of the
variables. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
5.37 BSGM - Stress and Displacement Distribution, for n = 4.4, using the final values of the
variables. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

xv
xvi
Nomenclature

APDL ANSYS Parametric Design Language.

BGM Beam Geometric Model.

BSGM Beam-Shell Geometric Model.

CAD Computer-aided Design.

CG Center of Gravity.

DOF Degree of Freedom.

FE Finite Element.

FEA Finite Element Analysis.

FEM Finite Element Method.

GUI Graphical User Interface.

HSP High Stress Point.

INEGI Instituto de Ciência e Inovação em Engenharia Mecânica e Engenharia Industrial

IST Instituto Superior Técnico.

MDO Multidisciplinary Design Optimization.

MTOW Maximum Take-off Weight.

SETsa Sociedade de Engenharia e Transformação S.A.

SF Safety Factor.

SGM Shell Geometric Model.

SIM Structural Improvement Model.

STOL Short Take-off and Landing.

VTOL Vertical Take-off and Landing.

Greek symbols

xvii
α Angle of attack.

∆ Variation of a parameter.

δ Displacement.

µ Mass Ratio.

ν Poisson’s Coefficient.

ρ Density.

σ Stress.

Roman symbols

û Total gust velocity.

c Average Wing chord.

a Size dimension of a hollow square cross section.

AR Aspect ratio.

b Wing Span.

brib Spanwise location of the Wing rib.

c Local wing chord.

CL Lift Coefficient.

Cm Pitching Moment Coefficient.

CLα Slope of the curve between the lift coefficient and the angle of attack.

CLmax Maximum Lift Coefficient.

Cmα Slope of the curve between the pitching moment coefficient and the angle of attack.

C nβ Yawing moment coefficient due to sideslip forces.

croot Wing root chord.

ctip Wing tip chord.

dwt Distance between the tail and the leading edge’s spar.

E Elasticity Modulus.

g Gravity Acceleration.

I Iteration.

K Aircraft Response Coefficient.

xviii
L Lift.

l Beam length.

m Mass.

mi Weight Variation Factor.

n Load factor.

Npar Number of Simulations performed.

npeak Peak Load factor.

S Wing area.

t Thickness.

t/c Thickness over chord ratio.

trib Wing rib thickness.

tspar Spar thickness.

u Normal component of the gust velocity.

V Velocity.

var Variable.

W Weight.

Subscripts

i Numeration of each variable.

max Maximum value of the parameter.

min Minimum value of the parameter.

x, y, z Cartesian components.

Superscripts

inc Incremental value of the variable.

max Maximum value of the variable.

min Minimum value of the variable.

xix
xx
Chapter 1

Introduction

1.1 Motivation

Nowadays, the transportation sector faces a lot of challenges in meeting the growing demand for com-
mon passengers mobility while reducing traffic, improving safety and mitigate emissions [1]. Automated
driving and electrification are disruptive technologies that may contribute to these goals, but they are
limited by congestion on existing roadways and land-use constraints.
Due to this, the interest of society in electric or hybrid Vertical Takeoff and Landing (VTOL) aircraft,
that can basically act as flying cars, has grown significantly. This peculiar type of aircraft has the potential
to overcome the limitations of congestion and land-use by enabling urban and regional aerial travel
services with the need of much smaller infrastructures and costs, when compared with the conventional
aircraft [2].
This increasing interest in VTOL aircraft has been pushing the boundaries of the science behind the
aerospace industry. These new ideas demand constant innovation and new knowledge, stronger and
lighter materials, among other solutions allied with new efficiency goals.
With this in mind, the motivation to perform a structural analysis in a VTOL aircraft can be essen-
tially resumed to the need of studying these new structures, identify critical points, design flaws and
propose feasible solutions that can withstand the needs and specifications of this type of aircraft, with
the minimum possible weight.

1.2 Topic Overview

The Flexcraft project, whose name appeared as an abbreviation for Flexible Aircraft, arose exactly due
to the emerging interest in transforming the way people travel. This concept born from the NewFACE’s
project and is being developed by ALMADESIGN, Instituto Superior Técnico (IST), Sociedade de En-
genharia e Transformação S.A. (SETsa), Embraer Portugal S.A. and Instituto de Ciência e Inovação em
Engenharia Mecânica e Engenharia Industrial (INEGI).
Imagine that a person works in Lisbon, Portugal and needs to be at Porto in one hour, for a work

1
meeting, without delays and airport bureaucracies. That person can use the Flexcraft, with the VTOL
or the Short Takeoff and Landing (STOL) capabilities, without needing a runway to depart and with the
possibility of arriving basically anywhere.
The main objectives of the NewFACE’s project are the development of new concepts and solutions
for future aircraft, focusing on external configurations, interior design, materials, technologies and man-
ufacturing processes, promoting greater environmental and energy efficiency, and more importantly this
project aims to provide travel solutions focused on the passenger needs.
The Flexcraft combines the idea of modularity and flexibility, since this aircraft is composed by a
lifting system capable of disengage from the fuselage module, allowing the cabin reconfiguration for
different missions (commercial and recreational, emergency and assistance, among others), through
short runway operation, improving the overall performance.
Another different feature is the propulsive system, which is intended to be fully electric or hybrid. For
this last case, the fuel will be used to generate electricity to power the propulsive system.
Flight and Operation, Versatility and Usability, and Materials and Production Processes are the
three lines of development in the project that will be validated through three evolutionary demonstra-
tors, namely, a Remotely Piloted Vehicle, a Full Scale Mock-Up and Materials and Production Process
demonstrators.
Several innovative configurations have been proposed throughout the various phases of the project,
but this work is focused in a VTOL Utility configuration.

1.3 Objectives

The main purpose of this work is to study the structural behavior of the lifting system (wing-tail) of the
Flexcraft VTOL configuration when submitted to the normal flight loads, specially for the critical points of
the flight envelope, with the following objectives:

• Structural analysis using finite element method (FEM):

– development of geometric models with increasing complexity;

– parametric design of the structure;

– definition of all the forces acting in the structure;

– finite element analysis (FEA) for the critical points of the flight envelope;

– identify critical points in the structure;

• Development of a model that can be used to improve an initial estimation of the aircraft’s geometry.

1.4 Thesis Outline

The present document is divided in six main chapters:

2
• Chapter 1: Introduction - description, objectives and the aim of this work are presented.

• Chapter 2: Historical Review - a brief introduction of aircraft’ evolution and presentation of the
several types of VTOL and STOL aircraft.

• Chapter 3: Conceptual Design - the initial phase of the design process, whose main goal is to
select some workable concepts and optimize them as much as possible. During this phase a large
number of concepts are generated and evaluated against each other, in order to try to define the
main characteristics of the aircraft that better meets the requirements. Also, all the characteristics
and parameters of the Flexcraft VTOL configuration are presented.

• Chapter 4: Structural Analysis - the description and explanation of all the structural analysis per-
formed, as well as the need of using the FEM, is also presented in this chapter.

• Chapter 5: Results - all the results are shown.

• Chapter 6: Conclusions - the conclusions of the work done are presented, as well as the future
work that could be developed.

3
4
Chapter 2

Historical Review

In this Chapter, a brief history of aviation is presented, as well as a small review regarding the types of
VTOL and STOL Aircraft.

2.1 Brief History of Aviation

From the beginning of time, humanity has always looked up to the skies and wished to fly. This desire
led to numerous attempts to develop flying machines, being even some records attributed to Leonardo
Da Vinci [3]. In fact, Da Vinci was the first to approach heavier-than-air-flight in a, somehow, scientific
way - Figure 2.1. The solution proposed by Da Vinci was conceived to lift off the ground vertically, with
no ground roll or runway needed.

Figure 2.1: Helical air screw flying machine designed by Leonardo Da Vinci [3].

Nearly four centuries after, in 1903, the Wright brothers achieved what humanity was searching for
so long: they performed the first powered, heavier-than-air-flight, with an aircraft built by themselves [4].
For that, they have used a fixed surface to provide lift - the wings - Figure 2.2. Heavier-than-air-flight
was achieved using a totally different approach than the one proposed by Da Vinci and, since then, the
history of aviation has undergone a fast evolution, becoming one of the most popular transportation for
millions of people around the globe.

5
Figure 2.2: Wright brothers’ flight [5].

Only 11 years after the Wright brothers’ flight, in 1914, the first scheduled air service took place
between St. Petersburg, Florida and Tampa, Florida. Piloted by Tony Jannus and having only one
paying passenger, the flight lasted 23 minutes, being a considerable improvement over the two-hour
trip by boat. This was the first step towards the proliferation of commercial aviation around the world.
Despite this evolution, commercial aviation was being accepted very slowly by general public, since most
of them were afraid to embark in these new flying machines.
With the advent of World War I, aircraft undergone a period of great improvement, due to its military
value. The development of more powerful engines was the most significant event at the time, allowing
aircraft to achieve speeds of 210 Km/h, more than twice the speed of aircraft built prior to war [6]. This
was a problem for commercial aviation, since flying became associated with war. Also, there was a very
large surplus of airplanes at the end of the war such that the demand for new production was almost
nonexistent for several years and leading many aircraft builders to bankruptcy. The fact that aircraft were
not, by far, as safe as today, was also not allowing commercial aviation to evolve as quickly as it could.
However, this did not prevent major marks to be achieved during this period. On April 6, 1924, eight
U.S. Army Air Service pilots and mechanics in four airplanes left Seattle, Washington, to carry out the
first circumnavigation of the globe by air. They completed the journey 175 days later on September 28,
after making 74 stops and covering about 27550 miles [7].

Figure 2.3: Boeing 247 [8].

Some years later, in the 1930s, airliners concluded that, to attract passengers to the skies, they
needed both larger and faster aircraft. The manufacturers responded to the challenge, making the

6
1930s the most innovative period in aviation history, with a great improvement in aircraft design, engines
and components. The first modern airliners also appeared in this era, particularly with the introduction
of the Boeing 247 [9] - Figure 2.3 , the DC-1 and the DC-2.
The DC-3, introduced in 1936 and considered the plane that changed the world [10], was the first air-
craft to enable airlines to make money carrying passengers. It had 50% greater passenger capacity than
the DC-2, costing only 10% more to operate. It also was considered a safer plane, built of an aluminum
alloy stronger than materials previously used in aircraft construction, with more powerful engines.
In 1939, the biggest step for modern aircraft was given in Germany: the He-178, built by Heinkel,
was the world’s first aircraft to fly under turbojet power [11] - Figure 2.4. Fortunately, it took five years for
the German scientists to perfect the design, being too late to affect the outcome of World War II.

Figure 2.4: Heinkel He-178, the first aircraft powered by turbojet [12].

From here on, aircraft started to become bigger and bigger and more and more efficient. Boeing
707 was the first U.S. passenger jet, with a length of 38 m and four engines, it could carry up to 181
passengers and travel at speeds of 980 Km/h [13].
A great mark regarding supersonic flight was achieved in 1968, when Russia tested the Tupolev 144
with success [14]. Some months later, in 1969, the Boeing 747, often referred as the Jumbo Jet, was
the first wide-body commercial jet airliner and cargo aircraft, holding the passenger capacity record for
almost 40 years [15]. With seating for as many as 450 passengers, it was twice as big as any other
Boeing jet and 80 percent bigger than the largest jet up until that time, the DC-8. Due to its success,
Douglas built its first wide-body, the DC-10, while Lockheed introduced the L-1011.
The magnificent 747 was dethroned in 2005 by the A380, a double-deck, wide-body, four-engine jet
airliner manufactured by Airbus, which is stills, today, the biggest commercial jet [16].
A resume of some of the most iconic commercial aircraft, distributed by Maximum Take-off Weight
(MTOW) and year of introduction, can be seen in Figure 2.5.
Nowadays, flying is one of the most popular travelling methods around the world, due to its amazing
feasibility and security. In fact, 2017 was even considered the safest year in the history of commercial
aviation, with no jet crashes in passenger service anywhere in the world [18].
However, despite the great improvement in aircraft, a big constrain was always found - the need of
runways for taking-off and landing. Due to this issue, several attempts were made through the years to
achieve Da Vinci’s vision, i.e., to develop an aircraft, with the same performance, that could take-off in
short runways or even vertically: the STOL aircraft and the VTOL aircraft. As it will be seen in Section
2.2, this challenge proved to be really hard and complex.

7
Figure 2.5: Iconic aircraft through the years [17].

2.2 STOL and VTOL Aircraft

In 1937, 34 years after the Wright brothers’ first flight, a new remarkable achievement in aviation can
be identified: the first heavier-than-air vertical lift-off finally took place with the development of the first
helicopter, the Focke-Wulf FW-61, in Germany [19].
Despite being an essential vehicle for several civil and military applications, the helicopter remains
extremely limited in the speed and range that it can achieve. Due to this limitation, a quest for a new
type of aircraft started, one that could combine the vertical take-off and hover capability of the helicopter
and the speed and range of a fixed-wing aircraft - the VTOL aircraft [3].
The search for this type of aircraft led to numerous configurations using a wide variety of lifting and
propulsion devices. A summary of this configurations can be seen in Figure 2.6.
The several aircraft produced can be essentially grouped according to two different factors: the type
of propulsion and type of structural configuration. The categories can be defined as follows:

• Lighter-Than-Air Aircraft: Balloon and Dirigible;

• Rotorcraft: Helicopter and Autogyro;

• Powered Lift: Convertiplane, Tail-sitter, Vectored Thrust, Lift Jets and Lift Fans;

The next Sections will be focused in presenting, in short, all the listed types of aircraft.

8
Figure 2.6: Illustration of VTOL and STOL aircraft [20].

2.2.1 Lighter-Than-Air Aircraft

A Lighter-than-air Aircraft is, as the name says itself, an aircraft that can fly by using a gas less dense
than air, making use of the Buoyancy concept [21], which refers to the fact that an object tends to rise if
it is placed in a medium whose density is greater than its own.

9
Balloon

Balloons were the first aircraft to make use of this principle, being the first person carrying balloon built
in 1783. Despite being an enormous technical evolution for the time, these kind of aircraft have a major
and obvious disadvantage, which is the fact that the balloon goes wherever the wind carries it, having
the passengers almost no control regarding the direction and speed of the flight.

Dirigible

As the comprehension of flight physics and controls started to evolve, there was the need to convert
balloons in useful aircraft. This was achieved by adding to the newly developed balloon designs an
engine, to propel the aircraft, and a rudder, to control direction. This modified form of a balloon got
known as a dirigible. The first modern dirigible built was attributed to Count Ferdinand von Zeppelin [22],
which had its initial flight on July 2, 1900. The aircraft was 125 m long, had 12 m in diameter and it was
capable of lifting 12000 kg. At their peak, Zeppelin’s dirigibles were able of carrying 72 passengers.

However, with the continuing improvement of heavier-than-air aircraft, the slow dirigibles start to
become obsolete. Also, with the explosion and consequent destruction of the Hinderburg - Figure 2.7,
dirigibles have gone definitely into disuse for passenger transportation, being mostly used, nowadays,
for advertisements and similar activities.

Figure 2.7: Disaster of the Hinderburg dirigible [23].

2.2.2 Rotorcraft

There are two major types of rotorcraft planes - the Helicopter and the Autogyro. Both of them are
described next.

10
Helicopter

An helicopter is the most versatile aircraft type, since it is able to take-off and land vertically, to hover,
to fly forward, backward and laterally. These attributes makes helicopters capable of being used in
congested or isolated areas, where fixed-wing aircraft would not be able to take-off or land [24]. Despite
this, helicopters are still very slow when compared with fixed-wing aircraft, with cruise velocities around
300 Km/h, not making them attractive for long range passenger transportation.

Autogyro

An autogyro, also known as gyroplane or gyrocopter, is a type of rotorcraft that uses an unpowered rotor
operating in auto-rotation to develop lift and an engine-powered propeller to provide thrust necessary to
balance the gyroplane drag force [25].
The Magni M22, the Magni M16 and the Wallis Type WA-121 are some examples of autogyros.

2.2.3 Powered Lift

Convertiplane

A Convertiplane is a type of aircraft that takes-off by using a rotor, making the transition to fixed-wing for
forward flight.

(a) Bell Boeing V-22 Osprey [26] (b) Airbus Vahana [27]

Figure 2.8: Convertiplane aircraft.

There are three main types of Convertiplanes:

• Tiltrotor: this aircraft uses the rotating blades as rotors in vertical flight and as propellers in the
forward flight. The blades have cyclic pitch control for hover and the power shaft pivots from
vertical to horizontal. The main wing stays fixed.

• Tiltjet: Similar to tiltrotor concept, but with turbojet or turbofan engines instead of ones with pro-
pellers.

• Tiltwing: It consists in tilting the entire wing between the hover and the forward flight. Tilting
the entire wing increases aerodynamic flow over lifting and control surfaces during transition and

11
minimizes lift loss due to downwash in hover. An additional method of control during hover is
required, such as a tail jet or jet rotor. Ailerons must be used to control roll in forward flight and
yaw in hover. During hover control is very difficult, specially due to barn door effect of wing in the
vertical position.

The Bell XV-3 and the Bell Boeing V-22 Osprey are two good examples of this type of aircraft. While
the first one did not enter in production, the V-22 Osprey (Figure 2.8 (a)) is still used today in the United
States military. The futuristic project from Airbus, the Vahana, is also a great example of a tiltwing aircraft
(see Figure 2.8 (b)).

Tail-sitter

In the present design the entire aircraft points straight up, being the entire thrust converted directly to
vertical lift. It is easy to take-off in vertical position, but much more difficult to land facing in the opposite
direction of the aircraft travel. The same propulsion system is used for both hover and forward flight.
This aircraft never passed the prototype and project phases. Two known examples are the Convair
XFY Pogo and the Lockheed XFV (see Figures 2.9 (a) and (b), respectively).

(a) Convair XFY Pogo [28] (b) Lockheed XFV-1 [29]

Figure 2.9: Tail-Sitter aircraft.

Vectored Thrust

Essentially, this kind of aircraft is able to vector the jet engine exhaust to create a vertical or horizontal
motion. This characteristic can be used for VTOL, like the British Aerospace Harrier II (Figure 2.10(a)),
or for higher maneuverability, as performed by the Sukhoi Su-37 - Figures 2.10(b)).

Lift Jets

A lift jet is an auxiliary jet engine used to provide lift for the take-off operation but it is, usually, shut down
for forward flight.
One of the few operational military aircraft with auxiliary lift engines was the Soviet Yakovlev Yak-38.

12
(a) British Aerospace Harrier II [30] (b) Sukhoi Su-37 [31]

Figure 2.10: Vectored Thrust aircraft.

Figure 2.11: Example of Lift Jet aircraft: Yakolev Yak-38 [29]

Lift Fans

This type of aircraft configuration has lifting fans located in large holes in an otherwise conventional fixed
wing or fuselage, that are used for VTOL and STOL operations.
The take-off is made by using the fans to provide lift, and then a transition is made for the fixed-wing
lift in forward flight. The F-35 Lightning II (see Figure 2.12(a)) is an example of this kind of aircraft,
being even the only that entered into production. Also, the recent AgustaWestland Project Zero (Figure
2.12(b)) and the Vanguard C2 Omniplane are another two examples of this type of configuration that are
still under development.

(a) F-35 Lightning II (b) AgustaWestland Project Zero [32]

Figure 2.12: Lift Fans aircraft.

13
2.2.4 Aircraft Representation

To end Section 2.2, a scheme regarding several aircraft types discussed can be seen in Figure 2.13.

Figure 2.13: Schemes for several types of aircraft: a) Balloon, b) Dirigible, c) Helicopter, d) Tilt Rotor, e)
Tilt Jet, f) Tilt Wing, g) Tail Sitter, h) Vectored Thrust, i) Lift Fans.

In the following Chapter, the methodology followed during this work is presented.

14
Chapter 3

Flexcraft Conceptual Design

With all the historical considerations reviewed and presented in Chapter 2, the Flexcraft concept ap-
peared. On this Chapter, all the requirements and design considerations are presented.

3.1 Project Requirements

The initial requirements were defined by the consortium during the Technical Specification activity and
are presented in Table 3.1.

Table 3.1: Project Requirements.


Requirements
Range [km] 1000
Passengers 9
Payload [kg] 900
Specifications
Cruise Speed [km/h] 400
Cruise Altitude [ft] 6000
Take-off and Landing Distance [m] < 325
Rate of Climb [ft/min] 1200
Stall Speed [kts] < 61
Configuration
Power Train Series or Parallel
Lifting System Type STOL or VTOL

3.2 Mission Profile

Due to the fact that the aircraft can have STOL or VTOL capacity, two different types of mission profile
are initially defined - Figure 3.1. This was needed since the energy spent in these missions is calculated
differently in the take-off and landing segments.

15
Initially, jet fuel is considered for the Maximum Take-Off Weight (MTOW) estimation, being after
replaced by batteries for the possible all-electric and hybrid propulsion systems.

(a) VTOL (b) STOL

Figure 3.1: Mission profiles.

3.3 Concept Generation

After identifying all the requirements imposed by the project consortium and presented in Section 3.1, it
is mandatory to analyze other types of aircraft in the market and identify what can be improved.
Due to this, and to make sure that all the aforementioned requirements are met, some aircraft con-
figurations are considered, being them the Utility, the Tandem Wing and the Flying Wing. These three
configurations are presented next.

3.3.1 Utility Configuration

The Utility concept is a light airplane, normally used for people or cargo transportation, with low con-
sumption and good operational velocities, that allows an easy operation in difficult conditions, e.g. short
runways or pavements with obstacles, with low costs associated.
This aircraft category is limited to less than nine seats, excluding the pilot, a MTOW of 5670 kg,
and is intended for limited acrobatic operations. This may include intentional spins (if approved for the
particular type of airplane) as well as commercial maneuvers with higher bank angles - between 60 and
90 degrees [33].
In order to accommodate different types of payload, this aircraft must also be capable of quickly
changing and adapting its configuration, with reduced costs.

3.3.2 Tandem Wing

The Tandem Wing configuration has two main wings with similar areas. One of the wings is located
forward, attached directly to the fuselage, while the other one is located in the back, separated from the
fuselage structure, but attached to the wing in the front. Both of them generate Lift.
This configuration should be quieter, due to the use of ducted fans, has the wings optimized for
cruise, the wing-carry-through structure is extended up front using a surface with an airfoil shape and

16
the distribution of the engines across the wing surface could allow the removal of some typical control
systems.
On the other hand, positioning the gravity center can be a problem, due to the higher number of
engines and the higher structural weight of the tail. All of this could lead to higher penalties for short
range missions.

3.3.3 Flying Wing

The last configuration, the Flying Wing is the most innovative one. It is a fixed-wing aircraft, without
tail, and with no defined fuselage. The payload, fuel and passengers are typically inside the main wing
structure and, theoretically, it is the most aerodynamic design configuration for a fixed wing aircraft.
Weight reduction due to the absence of tail, less aerodynamic drag and noise, optimized wings for
cruise and larger useful area to carry batteries are some of the advantages of this concept.
As for disadvantages, the control system can be a lot more complex, it could need mechanisms
to improve the behaviour at lower speeds, more engines will lead to more complexity and the higher
dimension ducted fans only work during the take-off, so it represents an extra weight during horizontal
flight.

3.3.4 Configuration Choice

After a careful analysis and deliberation of the characteristics of each configuration, the consortium is
deciding between the Utility VTOL or STOL configurations. Nevertheless, the development of this work
is focused in the Utility VTOL configuration.

3.4 Aircraft Design

As referred in Section 3.3, the Utility VTOL is the configuration in study. The aircraft top view and front
view are presented in Figures 3.2 and 3.3, respectively. It is possible to see that, in cruise conditions,
the propellers blades are aligned with the air flow.
The main characteristics of the aircraft are presented in Table 3.2. The fuselage is projected by Al-
maDesign and the pylon is intended to be developed in a future work but, for the theoretical calculations,
its height is assumed to be 40 cm.
The work developed during this study will mainly focus on the wing and tail structures, being their
dimensions represented in Figure 3.4.
Please note that the estimated structural weight for the wing-tail structure is about 1500Kg, i.e., 1/3
of the total MTOW value. This value is needed for the development of the computational models, since
it is the goal weight that must be achieved in the end.

17
Table 3.2: Utility VTOL concept characteristics. [34]
Part Parameter Value
MTOW [kg] 4500
Cmα -0.57
Cnβ 0.18
Fuel mass [kg] 440
Wiring mass [kg] 215.94
Global Batteries mass [kg] 419.78
Systems mass [kg] 95.20
2 Booms mass [kg] 92.46
Turbogenerator mass [kg] 121.30
Static margin [%] 10
Length [m] 6.2
Fuselage Diameter [m] 1.6
Capacity 9 pax
Mass [kg] 377.92
Span [m] 15
Wing area [m2 ] 35.8
Chord root [m] 3.4
Main wing Chord tip [m] 1.36
Mean chord [m] 2.4
AR 6,3
Airfoil NACA64A415
Twist NA
Mass [kg] 15.86
Span [m] 4
Horizontal Tail
Chord [m] 1.86
Airfoil NACA0009
Mass [kg] 16.36
Vertical Tail Span [m] 2
Chord 1.14
Propeller diameter [m] 1.4
Propeller power [kW] 410
Engines
Leading Edge 4 engines mass [kg] 411.07
Trailing Edge 4 engines mass [kg] 411.07

18
Figure 3.2: Flexcraft Top View: Take-off on the left and Cruise on the Right.[34]

Figure 3.3: Flexcraft Front View: Take-off on the left and Cruise on the Right.[34]

Figure 3.4: Flexcraft Dimensions. [34]

19
20
Chapter 4

Structural Analysis

Aircraft are usually made up of wings, fuselage, tail and control surfaces, such as ailerons, rudder and
elevator. There are some variations, e.g , a delta wing aircraft, like the Northrop Grumman B-2 Spirit,
that would not necessarily need a horizontal tail. Other example can be a canard configuration, such
as the Dassault Rafale B, which has the horizontal stabilizer near the nose. Basically, each component
must have specific functions and must be designed to ensure that it can carry out all the loads safely.
An aircraft structure is required to support ground and air loads. The first includes all the loads
encountered during its movement on ground (taxiing and landing loads, towing and raise up) and the
second one include loads imposed on the structure during flight by manoeuvres and gusts conditions.
Normally, when flying, direct loads, bending, shear and torsion occur in all parts of the structure in
addition to local, normal pressure loads imposed on the skin.

Figure 4.1: Principal aerodynamic forces on an aircraft during flight [35].

To ensure that an aircraft withstands all these loads, several types of computational analysis can and
should be performed prior to the manufacturing.
Due to the main scope of this thesis, only computational structural analyses are presented. There
are several types of structural analyses, such as:

• Static Analyses, which are used to determine the effects of steady loading conditions on a structure

21
while the inertia, damping and time-varying loads effects are neglected or can be approximated to
a static equivalent loads. It can be used to determine displacements, stresses, strains and forces
in structures or components (e.g tensile loading, shear and bending moment analysis).

• Modal Analyses, which is very important to determine the natural frequencies and mode shapes
of a structure or a component. These parameters are crucial to a structure design.

• Harmonic Analyses, that gives the ability to predict the steady-state response of a linear structure
to cyclic loads.

• Transient Dynamic Analyses, which can be useful to determine the time varying displacements,
strains, stresses and forces in a structure while it responds to any combination of static, transient
and harmonic loads. The inertia and damping effects are considered.

• Spectrum Analyses, being this type of analysis an extension of the modal analysis. It is used to
calculate stresses and strains due to a response spectrum or to random vibrations, both time-
dependent, e.g., jet engines thrust, rocket vibrations and many more.

• Buckling Analyses, which is a technique to obtain the critical loads that make the structure unstable
(buckling loads) and the result structure’s shape.

• Explicit Dynamic Analyses, type of analysis used to calculate fast solutions for large deformations
dynamics as well as to solve complex contact problems.

During this work, the types of analyses performed are essentially Static. The main goal is to make a
parametric study and change the thickness, size and number of several of the aircraft structural compo-
nents, and see how the structure behaves. These analyses should be performed for the critical points
of the flight envelope, which are the points in which the forces applied to the structure are higher.
Due to this need, the first step is to calculate the V-n and Gust Diagrams, which are presented in
Section 4.1

4.1 V-n and Gust Diagrams

4.1.1 V-n Diagram

The V-n Diagram, or flight envelope, shows, for a given velocity, the maximum expected load factor that
is used for the structural design and that the aircraft will experience during a specific flight condition. To
obtain the diagrams, the aircraft’s weight, velocity, altitude and air conditions are taken into account.
The load factor (n) is given by the ratio between the lift (L) and the aircraft weight (W), as follows

L
n= . (4.1)
W

With this basic relation, it is possible to obtain the V-n diagram for an aircraft. Figure 4.2 shows the
V-n Diagram obtained for an Airbus A320 [36].

22
Figure 4.2: V-n diagram obtained for the Airbus A320 [36].

The standard load factor for different types of aircraft is covered in Federal Aviation Regulations
[37], making it possible to select the maximum and minimum admissible values. With these values, the
horizontal lines AB and DE can be defined.
Lines OA and OE are defined using the stall conditions for the aircraft and the density (ρ), wing area
(S), maximum lift coefficient (CLmax ) and velocity (V). This relation is given by

L ρSCLmax 2
n= =± V , (4.2)
W 2W

which is deducted from the load factor definition. The positive sign is used for curve OA, while the
negative sign is used for curve OE. CLmax is assumed equal in both curves, which is an approximation.
Conditions at point B occurs for the maximum flight velocity, which is the dive velocity, VDive , obtained
by

VDive = 1.5VCruise . (4.3)

With this value, line BC can be drawn. Finally, the envelope is enclosed by drawing line CD.

4.1.2 Gust Diagram

When flying through turbulent gusts, the effective angle of attack of the aircraft changes during a short
time period. This change can increase or decrease the wing lift, changing the load factor as a conse-
quence. Regarding this effect, the aircraft must be designed to support this changes, being the gust
diagram a very important tool for it.
To plot the diagram, it is necessary to calculate the incremental load factor, ∆n. Considering a frontal
wind gust [38] and the slope of the Lift Coefficient curve (CLα ), the load factor is obtained by

∆n ρuSCLα
= , (4.4)
V 2W

being u the normal component of the gust velocity. This value must be calculated for different flight

23
conditions, using

u = K û, (4.5)

where K is the aircraft’s response coefficient and û the total gust velocity.
The total gust velocities, for different flight conditions, are obtained according to Figure 4.3. The flight
conditions analyzed are flight at high angle of attack, level flight and the dive condition.

Figure 4.3: Variation of total gust velocity with altitude [38].

The response coefficient, for a subsonic aircraft, is given by

0.88µ
K= , (4.6)
5.3 + µ

being the mass ratio, µ, obtained using the medium chord (c)

2W
µ= . (4.7)
ρgcCLα

The peak load factor (npeak ) is then the sum of the mean load factor at cruise and the incremental
load factor. Therefore,

npeak = 1 ± ∆n. (4.8)

Finally, enclosing all the peaks, the gust diagram for the aircraft can be presented. Figure 4.4 repre-
sents the Gust Diagram obtained, once again, for the Airbus A320 [36].

4.2 Finite Element Method

The Finite Element Method (FEM) is a numerical method used worldwide for solving complex problems
of engineering. In this method, the original body of the structure is approximated by the assemblage of
smaller similar regions/areas, known as the Finite Elements (FE), interconnected with each other [39].
The FEM can be divided in three main stages:

24
Figure 4.4: Gust diagram obtained for the Airbus A320 [36].

1. Pre-Processing - Define the problem with the FE Model;

2. Processing - Obtain the Solution;

3. Post-Processing - View and interpret the results.

Inside each stage, there are several steps performed. The Pre-Processing stage encloses steps
of Model Design, Meshing, Selection of FEs, Definition of Geometrical Properties, Material Selection,
Boundary Conditions, Loads and Type of Analysis. These steps can be done in a different order.
Regarding the Processing stage, there are two approaches that are widely used for structural prob-
lems:

• Flexibility Method [40] - The internal forces are the variables;

• Stiffness Method [41] - The displacement of the nodes are the variables.

The Stiffness Method is the most used on FEM software, including the one used during this disserta-
tion. On a structural analysis, there are displacement compatibility conditions that must be assured, i.e.,
elements that are connected to a common node, along with a common edge or common surface during
the Pre-Processing stage, must remain connected to that node, edge or surface during the Processing
stage. The compatibility conditions as well as the boundary conditions are written in terms of nodal
displacement and are essential to an adequate problem modeling.
Then, and using as an example a static analysis, the FEM software assembles the governing equa-
tions in matrix form, following

[K] [u] = [F ] , (4.9)

where [K] is the stiffness matrix, obtained from material properties and geometrical shape, [F ] are the
external forces applied to the structure and [u] are the node displacements, which are the unknowns of
the problem.

25
Finally, the Post-Processing stage, as the name itself illustrates, is the stage of visualizing and in-
terpreting the solutions, such as the visualization of the deformed structure, nodal displacement, nodal
stress, among others.
The FEM became so popular because it allows to simulate any system and to determine desired
properties with accuracy before experimental tests, fabrication and manufacturing, reducing costs. This
is also a very good approach for complex models, where exact or theoretical solutions are really difficult
to obtain.
The aforementioned points are some of the main reasons why the FEM was selected to carry out the
structural analysis of the wing-tail section of the Flexcraft’s model.
As stated previously, there are several commercial software systems available to perform compu-
tational simulations. For this specific work, the software chosen is ANSYS. It provides two different
environments to modulate the geometry and perform the computational analysis:

• Workbench - point and click operations in a Graphical User Interface (GUI), being a much more
visual approach. This option is more suitable for few load cases or complex CAD modifications.

• APDL - stands for ANSYS Parametric Design Language, meaning that the simulation must be
written in a script. For often repeated calculation, specially with changing variables (CAD or FEM
inputs), writing a script in APDL is a better approach.

Since the main goal of this work is to optimize and make parametric studies of the aircraft structure,
APDL is the environment chosen to make the computational simulations.
Next, the FEs selected for the structural analysis will be described in Subsection 4.2.1.

4.2.1 Finite Elements

The Flexcraft’s wing and tail are composed by several different components, such as the ribs, spars,
booms, skin, engines, and so on, meaning that there is the need to select proper FEs according to each
one of these structures. Also, and as explained below, it is necessary to define an element capable of
representing punctual masses.
For this work, and taking into account the structures that compose the Flexcraft wing and tail, three
different types of FEs are chosen: a Beam FE, a Shell FE and a Mass FE. All these elements are
described below.

Beam element - BEAM189

As stated in the beginning of Subsection 4.2.1, the Flexcraft’s wing-tail is composed by different compo-
nents. Some of them, such as the spars, can be modeled as Beam elements, so there is the need to
select a proper FE that is able to represent a Beam.
ANSYS has two different elements to represent Beams:

• BEAM 188 - 2-node linear element;

26
• BEAM 189 - 3-node quadratic element.

Both BEAM188 and BEAM189 FEs follow Timoshenko beam theory [42]. This theory includes shear-
deformation effects and provides options for unrestrained and restrain warping of the cross sections.
These type of FEs have 6 degrees of freedom in each node: 3 translations (in the x, y and z directions)
and 3 rotations (about the x, y and z directions). Since it can support warping, a seventh degree of
freedom can also be considered, which is the warping magnitude at each node. This element is well-
suited for analyzing slender to moderately thick beam structures, while making linear, large rotation
and/or large strain nonlinear studies [39].
The main difference between them is that BEAM188 is a linear element (with 2 or 3 nodes), while
BEAM189 is a quadratic element - Figure 4.5.

Figure 4.5: BEAM188 (on the left) and BEAM189 (on the right).[39]

This difference means that, for the same number of FEs, BEAM189 presents more accurate results
than BEAM188 [43]. To obtain the same level of accuracy using BEAM188, a more refined mesh is
needed, which would impact the computational performance of the solution.
With all the mentioned considerations, the BEAM189 is the FE chosen to model beam type compo-
nents.
To complete the FE characterization, it is necessary to state that elasticity, plasticity, creep and other
nonlinear material models are supported when using this type of Finite Element. The cross-section
associated with this element can be a built-up section referencing more than one material. Added mass
is also available for the modeling.

Shell Element - SHELL281

Thin-walled structures are widely used in many industries, since they can sustain high loads per weight
capacity. An aircraft is a type of structure that employees this type of thin-walled construction, being the
ribs and skin, of both wing and tail, examples of it.
Thin-walls present some challenges during the numerical simulation. They could be represented by
3D Solid FEs but this would make it difficult to construct accurate and efficient meshes that would also be

27
computationally efficient. To avoid this, Shell FEs can be used instead, since they create a mathematical
2D idealization of a 3D structure that does not requires the explicit modelling of the thickness.
There are two theories widely used in the formulation of Shell FEs:

• Kirchoff-Love theory;

• Reissner-Mindlin theory.

The difference between both theories can be explained as follows.


A FE that follows Kirchoff-Love theory does not calculate transverse shear stresses. This means that
the resulting deformations may be underestimated, being this effect specially noted in thick structures.
FEs following the Reissner-Mindlin theory [44] will take into account the shear stress distribution over
the thickness, showing a softer deformation behaviour due to the presence of the shear stresses [45].
Due to this extra calculation, using a FE that follow the Reissner-Mindlin theory might imply an higher
computational time but will present more accurate results.
Due to the evolution of technology, the newer versions of ANSYS APDL provide two suitable elements
for this kind of structure and both follow the Reissner-Mindlin theory:

• SHELL181: four-node low order FE;

• SHELL281: eight-node high order FE.

Figure 4.6: SHELL181 (on the left) and SHELL281 (on the right) [39]

As stated above, both elements follow the Reissner-Mindlin theory, but they have a different number
of nodes per side. Both are suitable for analyzing thin to moderately-thick shell structures, but SHELL181
is a four-node element with six degrees of freedom at each node and SHELL281 has eight nodes with six
degrees of freedom at each node. The degrees of freedom are translations in the x, y, and z directions,
and rotations about the x, y, and z-axes - Figure 4.6.
Since the Beam FE selected previously has 3 nodes, there is the need to select a Shell FE that
has also 3 nodes per side - the SHELL281. Despite the more accurate results, the previous reason is
fundamental for the FE selection. If a 2 node per side element was selected instead, compatibility issues
would occur and the results would be much less accurate.

28
Mass element - MASS21

In addition to the wing and stabilizer structures, there are other components attached to the structure that
influence its behavior. Components such as the engines, batteries and fuel tanks are some examples.
Since the main focus of this work is related with the wing and tail structure, there is no need to design
these components with geometric accuracy, but their effect on the structure must at least be simulated.
This way, the above components can be represented by punctual masses distributed throughout the
structure.

Figure 4.7: MASS21. [39]

MASS21 (Figure 4.7) is a point element that can have six degrees of freedom: translations in the X,
Y, and Z directions and rotations about the X, Y, and Z axes. A different mass and rotary inertia may be
assigned to each coordinate direction, to simulate a different behavior according to the direction.

4.2.2 Finite Element Models

As described, 3 types of FEs were chosen to represent the several components of the aircraft structure.
The next step is to implement these FEs in the FE Models which will simulate the aircraft structure.
Due to the complexity of the Flexcraft, two simplified models are tested first: one using only Beam
FEs and the other using only Shell FEs - the punctual masses will be present in both of the models. The
test of simplified models is important, since they consume a much lower computational power. With this
in mind, if the results obtained in these models are also accurate, they can be used for much quicker
simulations.
Finally, the more complex and complete model is developed, using a Beam and Shell FE according
to the structural component. These 3 models are described below.

BGM - Beam Geometric Model

As described, the first model developed uses only the BEAM189 FE, which is based on the Timoshenko
beam theory [42], to represent all the structures, and the MASS21 to represent all the non-structural
components which are rigidly attached to the wing and stabilizer - Figure 4.8.
This is the simplest model developed during this work and it represents only the internal structure of
the Flexcraft. Since it is built only using Beam FEs, it is not possible to model the skins of the wing and

29
Figure 4.8: Beam Geometric Model.

stabilizer, as well as shaping the ribs with an airfoil format. The wing reinforcements can not also be
represented.
Due to the main goal of this work, there are several inputs that must be given to the APDL script,
before making the simulation. These values will change the geometry of the model, making it possible
to study their effect in the aircraft, and are presented in Table 4.1.

Input Description
RibsIn Number of wing ribs on the inner side
RibsOut Number of wing ribs on the outer side
NELEM Increasing this value increases the mesh granularity
t in Thickness of the wing root rib [m]
t out Thickness of the wing tip rib [m]
t wingboom Thickness of the wing spars’ cross section [m]
chord percentage Used to define the side of the spars cross section by d = clocal × chord percentage
size rearboom Define the cross section side of the boom connecting the wing to the tail [m]
t rearboom Define the cross section thickness of the boom connecting the wing to the tail [m]
size vertical Cross section side of the vertical boom [m]
t vertical Cross section thickness of the vertical boom [m]
t ribs tail Cross section thickness of the tail ribs [m]
t spars tail Cross section thickness of the tail spars [m]
size spars tail Cross section side of the tail spars [m]

Table 4.1: Inputs for the Beam Model.

For a better understanding of what these variables represent, they are identified in real images of the
BGM.

30
Starting by the wing (see Figure 4.8), it was considered that the wing has at least 3 ribs, being them
located at the wing root (Root Rib), at the connection with the tail structure (Middle Rib) and at the
wing tip (Tip Rib). The ribs between them can be inserted as inputs, being these inputs the RibsIn and
RibsOut variables.
Since the wing is facing much heavier loads near the root, a suitable initial approach is to model the
Root Rib with an higher thickness, while the Tip Rib can be modeled with a much lower thickness. The
values of these thicknesses are defined by the variables t in and t out, respectively.

Figure 4.9: Identification of variables RibsIn, RibsOut, t in, t out; Cross Sections 1, 2 and 3 and MASS21
FE.

The thicknesses between the Root and Tip Ribs follow a decreasing linear rule across the wingspan,
by the rule described in Equation 4.10, where brib represents the spanwise location of the rib.

t out − t in
trib = × brib + t in (4.10)
7.5

Since this model uses only beam FEs, and due to the fact that it is not possible to model the ribs using
an airfoil shape, the ribs are modeled as beams with a constant value, being this value the maximum
thickness value of the airfoil. Since the airfoil for the wing has a (t/c)max = 0.15, the height of each rib
can be modeled by Equation 4.11, where croot and ctip are the wing chord values at the wing root and
tip, respectively.

0.15ctip − 0.15croot
hrib = × brib + 0.15croot (4.11)
7.5

After the wing modulation, it is necessary to design and model the tail. The tail connects to the wing
by two Booms. The horizontal tail has basically the same components as the main wing: ribs and spars.

31
The horizontal tail is then connected to the booms by the vertical tail. Once again, since this model only
uses Beam FEs, it is not possible to module the tail skin and the ribs are only an approximation, since
they do not have an airfoil shape.

Figure 4.10: Identification Cross Sections 3, 4, 5 and 6.

The tail can be seen in Figure 4.10, where all the cross sections of its different structures are also
identified. Furthermore, in Table 4.2 all the sections of the Beam Model can be seen, as well as all the
inputs needed to each one.
As stated in the beginning of the Section, to obtain accurate results, it is necessary to modulate the
aircraft components that are built inside the wing. To do so, MASS21 FEs are used.
The components that must be modulated are the fuel tanks, the engines and the batteries. Following
the aircraft definitions presented in Chapter 3, the engines must be applied in the leading and trailing
edges of the Middle and Tip Ribs. Regarding the fuel tanks, their weight is distributed in the central part
of the wing, between the Middle Ribs of the left and right wings. Finally, the batteries are applied in the
leading edge’s spar. The application of these masses can also be seen in Figure 4.9.
Finally, regarding boundary conditions, the Beam model has all the Degrees of Freedom (DOF) con-
strained in the intersections of the wing spars with the wing Root Rib. This approach is an approximation,
in fact only the gravity centre (CG) should have all DOFs constrained. However, the parameterization of
the variables changes the location of the CG, which would make it difficult to fix it for each FEA.

SGM - Shell Geometric Model

The second model developed during this work is the Shell Geometric Model (SGM) and, as the name
itself refers, uses only Shell FEs to represent the aircraft structure. The Shell FE used is the SHELL281

32
Table 4.2: Cross Sections of the Beam structures in the BEAM model.
Cross Section Location Shape Variables

t → Equation4.10
1 Wing Ribs
h → Equation4.11

h = chord percentage × clocal


2 Wing Spars w = chord percentage × clocal
t = t wingboom

h = size rearboom
3 Booms w = size rearboom
t = t rearboom

h = size vertical
4 Vertical Tail w = size vertical
t = t vertical

h = size spars tail


5 Tail Spars w = size spars tail
t = t spars tail

t = t ribs tail
6 Tail Ribs
h = 0.09 × ctail

(element formulated based on the Reissner-Mindlin plate theory [44]) and, as previously stated, the
MASS21 FE is used to represent all the punctual masses. When compared with the previous model,
this one has the advantage of better representing the aircraft geometry, since it allows to design the ribs
with an airfoil shape and also allows the representation of the wing and tail skin.
Once again, the main goal of this work is to perform parametric studies on the model and so, as pre-
viously, it is necessary to provide several inputs to the APDL script. All the inputs needed are presented
in Table 4.3.
The internal structure of the wing is represented in Figure 4.11. The convention of the first model is
kept, since it is considered that the wing has at least three ribs: the Root Rib, the Middle Rib and the Tip
Rib. The ribs between them are defined as inputs, being the RibsIn and RibsOut variables.
Once again, and following the approach of the Beam Model, the ribs are modeled with different
thicknesses, being the Root Rib the thicker. The thickness of each rib is calculated using a linear
variation that follows the previously defined Equation 4.10. The spars’ thickness, tspar , varies in the
spanwise direction and depends on the local chord and the chord percentage variable. This value is

33
Table 4.3: Inputs for the Shell Model.
Input Description
RibsIn Number of wing ribs on the inner side
RibsOut Number of wing ribs on the outer side
aesize Reducing this value increases the mesh granularity
t in Thickness of the wing root rib [m]
t out Thickness of the wing tip rib [m]
t wingshell Thickness of the wing skin [m]
chord percentage Used to define the thickness of the spars by t spar(z) = c(z) × chord percentage
t rearboom Define the thickness of the booms connecting the wing to the tail [m]
t vertical Define the thickness of the vertical beams [m]
t spars tail Define the thickness of the tail spars [m]
t ribs tail Define thickness of the tail ribs [m]
t skin tail Define the thickness of the tail skin [m]

Figure 4.11: Identification of variables RibsIn, RibsOut, t in, t out,t wingshell and t spar.

obtained by Equation 4.12, where c(z) is the chord variation across the wing span.

tspar (z) = chord percentage × c(z) (4.12)

Wrapping the ribs and the spars, there is the wing skin, whose thickness is defined by the t wingshell
variable - Figure 4.11.
After designing the wing, it is necessary to develop the tail. The tail structures connects to the wing
in the Middle Ribs through the Booms ,being its thickness defined by t rearboom. The vertical tail, on its
side, it is defined by t vertical. The horizontal tail has essentially the same structure of the main wing,
being composed by wings and spars, wrapped by an external tail skin. The t spars tail, t ribs tail and
t skin tail variables define the thicknesses of the tail spars, tail ribs and tail skin, respectively - Figure

34
4.12.

Figure 4.12: Identification of variables t rearboom, t vertical,t skin tail and t spars tail.

As in the previous model, it is necessary to add the MASS21 FEs to represent the engines, batter-
ies and fuel tanks. The distribution is similar to the Beam Model, but taking now in consideration the
thickness of the wing - Figure 4.11.
Regarding the boundary conditions, similarly to the BGM, the SGM has all the DOFs constrained in
the lower intersections of the wing spars with the Root Rib, which is an aproximation.

BSGM - Beam-Shell Geometric Model

This is the last model developed and it is a mixture of the previous two - Figure 4.13 (where the beam
and shell elements are respectively shown). Since there are structures that clearly behave more as a
beam, such as the spars and booms and others that behave more as a shell, such as the skin and ribs,
this last model has the main goal of obtaining a structural behavior more similar to the real one.
This way, and following the work previously done, the FEs used are below identified:

• BEAM189: used to model the Spars and Connection Booms - Figure 4.13 (a);

• SHELL281: used to model the Ribs and Skins - Figure 4.13 (b);

• MASS21: used to represent all the components as mass points rigidly attached to the wing and
stabilizer - Figure 4.13 (a) and (b).

Once again, there are several inputs that must be given to the APDL script, before making the
simulation. These values will change the geometry of the model, in order to understand how each
one influences the structural behavior. These variables are presented in Table 4.4, along with a brief
description of each one.

35
(a) Internal Structure defined with BEAM189 FE (b) External Structure defined with SHELL281 after defining the
Internal Structure

Figure 4.13: Beam-Shell Model.

Input Description
RibsIn Number of wing ribs on the inner side
RibsOut Number of wing ribs on the outer side
aesize Reducing this value increases the mesh granularity
t in Thickness of the wing root rib [m]
t out Thickness of the wing tip rib [m]
t wingshell Thickness of the wing shell [m]
t wingboom Thickness of the wing spars’ cross section [m]
t reinforcement Thickness of the wing reinforcement’s cross section [m]
chord percentage Used to define the side of the wing spars cross section by d = clocal × chord percentage
size rearboom Define the cross section side of the boom connecting the wing to the tail [m]
t rearboom Define the cross section thickness of the boom connecting the wing to the tail [m]
size vertical Cross section side of the vertical boom [m]
t vertical Cross section thickness of the vertical boom [m]
t spars tail Define the thickness of the tail spars [m]
size spars tail Define the side of the tail spars [m]
t ribs tail Cross section thickness of the tail ribs [m]
t skin tail Define the thickness of the tail skin [m]

Table 4.4: Inputs for the Beam-Shell Model.

As described earlier, the wing ribs are modelled using the SHELL281 FE and the spars are modeled
using the BEAM189 FE. This can be seen in Figure 4.14, which represents the internal structure of the
wing.
The approach to model the ribs is the same as the one used in the previous models.
Wrapping the internal structure of the wing, there is the wing skin - Figure 4.15.
The wing skin is modeled using the SHELL281 element and its thickness can be inserted during the
inputs, being defined by the t wingshell variable. The skin modulation completes the basic structure of
the wing.
After designing the wing, the next step is to model the connection between the wing and tail, using
the Booms - Figure 4.15. The booms are modeled using the BEAM189, equally as the vertical tail.
However, there is the need to use two different cross sections for each component, being the booms

36
Figure 4.14: Identification of variables RibsIn, RibsOut, t in, t out and Cross Section 1.

defined by Cross Section 2 and the vertical tail defined by Cross Section 3.

Figure 4.15: Identification of variables t wingshell, t skin tail and Cross Sections 2 and 3.

Also, due the high stresses that the wing structure will face in the root and near the connection to
the tail, there was the need to reinforce the structure in this points. For the sake of simplicity, these rein-
forcements have the same side dimension as the spars near their placement. However, their thickness
can be changed by t reinf orcement. Then, the wing reinforcements are defined by Cross Section 4.
The horizontal tail has, one more time, the same internal structure as the wing. The internal structure
of the tail contains the tail ribs, whose thickness is defined by the t ribs tail variable, and by the tail spars,
defined by Cross Section 5. The internal structure is then wrapped up by the tail skin. Equally as the
wing, the thickness of the tail skin can be changed for each simulation and its value is defined by the

37
t skin tail variable.
In Table 4.5, it is possible to observe all the cross sections of the BEAM189 structures, as well as the
inputs needed to create their dimensions.

Table 4.5: Cross Sections of the Beam structures.


Cross Section Location Shape Variables

h = chord percentage × clocal


1 Wing Spars w = chord percentage × clocal
t = t wingboom

h = size rearboom
2 Booms w = size rearboom
t = t rearboom

h = size vertical
3 Vertical Tail w = size vertical
t = t vertical

h = chord percentage × clocal


4 Wing Reinforcement w = chord percentage × clocal
t = t reinf orcement

h = size spars tail


5 Tail Spars w = size spars tail
t = t spars tail

The final step to complete the BSGM is to add the masses representing the engines, fuel tanks and
batteries. Equally as the previous models, the engines are applied in the leading and trailing edges of
the Middle and Tip Ribs, the fuel tanks are applied in the central part of the wing and the batteries are
applied along the spar near the wing leading edge. The masses are modeled using the MASS21 FE.
A representation of the MASS21 elements is presented in Figure 4.14.
Regarding the boundary conditions, the model has all the DOFs constrained in the lower intersections
of the wing spars with the Root Rib. As explained in the BGM definition, this is an approximation, since
the model should only be constrained in its CG.

4.2.3 Aerodynamic Loading

To perform the structural simulations, it is necessary to apply the aerodynamic forces acting on the
aircraft. Calculating these forces is not under the scope of this study, so they are calculated using a
Multidisciplinary Design Optimization (MDO) framework [46] already developed, made for conceptual
design and analysis of new aircraft configurations. The aerodynamic model of this MDO framework

38
allows the preliminary analysis of the aerodynamic loading that an aircraft is facing.
In this work, the same aerodynamic model is used to generate the loads for the three structural
models (BGM, SGM and BSGM). This tool is based in a potential flow model - inviscid and irrotational (for
more details regarding this flow model see Katz and Plotkin [47]). Since the lift is the largest aerodynamic
load faced by the aircraft, in the upper and lower limits of the flight envelope, only the main wing and
horizontal tail are considered when applying the aerodynamic model, reducing the modeling complexity
of a full 3D mesh. The main limitation of this assumption is not considering the propellers mounted
ahead of the wing, which would slightly increase the lift on the wing (thus lift is slightly underestimated
in the aerodynamic model used). Considering the propellers would involve more complex aerodynamic
effects, which can not be predicted using this 3D panel method. Concerning the horizontal tail, this is
a conservative approach, since the effect caused by the wing wake is not accounted. This effect would
lead to a lower Lift force, meaning that the value obtained for the horizontal tail is slightly overestimated.
This model uses the 3D panel numerical method described by [47], with compressible corrections
(which in this case are not applied since the cruise speed of the Flexcraft is approximately Mach 0.3)
and viscous friction corrections.

Figure 4.16: MDO framework results: (a) Wing Mesh. (b) Wing (nmax ). (c) Horizontal tail (nmin )

An illustration of an aerodynamic mesh for the main wing can be seen on Figure 4.16 (a) alongside
with two illustrative simulations of the maximum and minimum load factors for main wing and horizontal
stabilizer, respectively - Figure 4.16 (b) and (c).
The aerodynamic loads for the maximum and minimum load factors are estimated such that the lift
generated by the main wings reaches 4.4 and -1.8 times the weight of the aircraft (the reference weight
is the one from Table 3.2, i.e. 4500 kg), respectively. These load factors are predicted to occur at the
cruise speed and sea level conditions for angle of attacks of 6 (nmax ) and -6.65 (nmin ) degrees.
The aerodynamic loads are then transferred to the closest nodes of the finite element mesh of the
aircraft’s structure by means of an equivalent forces/moments system, which ensures that all loads are
properly transferred. Figure 4.17 shows how the aerodynamic loads are applied to the aircraft’s BGM.

4.2.4 Structural Improvement Model

Since the Geometric Models are prepared for parametric studies, it is possible to take advantage of it
and try to optimize the variables defined for the BGM, the SGM and the BSGM. Due to this, a model that
runs on top of the Geometric Models, with the main goal of obtaining a lower weight for the structure, is

39
Figure 4.17: Aerodynamic loads applied in the BGM (n=-1.8).

built - the Structural Improvement Model (SIM) - Figure 4.18. With the objective of achieving a lightweight
structure, a upper limit target value of 1500 kg (which corresponds to 1/3 of the total MTOW defined) is
set.
Essentially, this model starts from a set of i variables, defined for each Geometric Model, with over-
dimensioned values that produce an over-weighted and over-robust structure. Having this initial set of
values, it is then necessary to define the minimum (varimin ), maximum (varimax ) and incremental (variinc )
values that will be analyzed for each variable, vari .
Since the parameter sequence influences the results obtained, it is necessary to find a way to decide
which variable to analyze first. Thus, as the main objective is to minimize the structural weight, W , it
is fundamental to verify the influence of each variable on it. Said so, the weight variation factor (mi ) is
calculated for each variable. This value represents the sensibility of the weight variation with the change
of each variable, in kg/m.
After defining this, it is necessary to define some constrains. These constrains can be geometric or
structural, according to the needs and specifications of the project.
After defining all the needed values and the constrains, the model increments one or two variables
at a time, depending on the structural component in study. For each increment, a FEA is performed
and all the values for the constrains are obtained. Then, during the post-processing phase, the value
for vari that minimizes the weight while fulfilling the constrains is fixed and used for the next variable’s
parametric study.
The total number of simulations performed, Npar , can be calculated by
" n  n
!
max min
varmax − varmin vari,1 − vari,1
X  X
i i
Npar = +1 + +1 ×
i=1
variinc i=1
inc
vari,1
!# (4.13)
max min
vari,2 − vari,2
inc
+1 ×3
vari,2

where the first block calculates the number of simulations for each variable studied individually, while the

40
Figure 4.18: Structural Design improvement model.

second block refers to the case when two variables must be studied together.

4.2.5 Parametric Constrains

As stated in Subsection 4.2.4, in order to run the SIM, it is necessary to define some constrains.
Even knowing that the main goal of the SIM is to reduce the weight of an initially oversized model,
that is still capable of withstanding all the aerodynamic loads, without structural failure, one must ensure
that the aircraft remains aerodynamic efficient, even when operating in critical conditions. This means

41
that, even for the limiting load factors, the structure must not present high deformations.
With that being said, it is necessary to define some constrains in order to obtain suitable values for
the structure. On this specific study, the main constrains are the maximum stress (σmax ), the Wing Tip
Displacement (δwingtip ) and the Tail Displacement (δtail ). Regarding σmax , the constrain is defined by

σyield
σmax < , (4.14)
SF

being σyield the yield stress of the material, i.e., the elastic limit of the material and SF a Safety Factor.
A SF = 1.5 is selected, since it is the normal value used for passenger aircraft structural design. This
safety factor alone is equivalent to a probability of failure of between 0.01% to 0.001% [48].
For δwingtip , a high wing stiffness is considered, meaning that its value can not be higher than 5% of
the wingspan [49]. Regarding the tail displacement, and due to its important role in ensuring the aircraft
stability, it was considered that its value could not also be higher than 5% of the distance between the
tail and the leading edge’s spar (dwt ) - Figure 3.4. Mathematically, these conditions are defined by

− 0.05b ≤ δwingtip ≤ 0.05b, (4.15)

− 0.05dwt ≤ δtail ≤ 0.05dwt . (4.16)

having b and dwt the values of 7.5 m and 9.25 m, respectively.

4.2.6 Material Selection

Choosing a material, in the aerospace industry, may require an extensive evaluation of its performance
under creep, tension, compression, bending, fatigue and corrosion. Nowadays, there are different op-
tions to fulfill the requirements of a new aircraft project: metals, composites, wood, among others [50].
The wing structure of a modern aircraft can be designed with a combination of different types of
materials, according to their specific structural function. Each of these components needs to support
different loads and so the right material needs to be selected. Steel or aluminium alloys can be used in
the ribs and spars manufacturing, while composite materials are common in the design of the wing skin
and the control surfaces.
For simplicity, an Aluminium alloy, which properties are presented in Table 4.6, is chosen to perform
all the structural analysis in this work. This metal is characterized by having lower density values when
compared to steel alloys, with good corrosion resistance properties [51].

Table 4.6: Aluminium alloy properties [52]


E [MPa] σyield [MPa] ν [-] ρ [Kg/m3 ]
70 276 0.33 2700

42
Chapter 5

Results

In this Chapter, all the results obtained are presented and discussed.

5.1 V-n and Gust Diagrams

Considering the data gathered for the Flexcraft and following the methodology presented in Section
4.1.1, it is possible to obtain the V-n and Gust Diagrams (Figure 5.1) for its flight conditions.

Figure 5.1: V-n and Gust Diagrams obtained for the Flexcraft at cruise altitude

The structural analysis presented in this work is performed for the critical flight conditions, n = 4.4
and n = −1.8, as defined in Corke [38], for an utility aircraft, with an air flow velocity of 400km/h and
an angle of attack of 6o and −6.65o , respectively. These critical points are the worse flight conditions
that the aircraft is expected to face. These aerodynamic loads are computed in the MDO framework
mentioned on Subsection 4.2.3.

43
5.2 Convergence of the Geometric Models

Before starting the computational simulations, it is mandatory to make a mesh convergence study, for
each of the FE models developed. This is a crucial step. If the FEM is applied using a non-converged
mesh, the solutions obtained might be erroneous, which can lead to wrong design decisions and disas-
trous consequences [53].
For the models in study, the two parameters that control the mesh refinement are the NELEM, for
the Beam Geometric Model, and the AESIZE, for the Shell and Beam-Shell Geometric Models. The
definition of these parameters is already stated in Section 4.2. To make the convergence study, a natural
frequency study is performed for each model - Figure 5.2. Please note that the natural frequencies
computed might not be representative of the real ones, since the results are obtained for a set of initial
inputs.

Figure 5.2: Convergence study for the Beam, Shell and Beam-Shell Model

Looking for the results, the values selected are the following:

• N ELEM = 5, for the Beam Geometric Model, which corresponds to 573 elements for the structure
in study. The average relative error, when compared with the more refined mesh, is 0.013%;

• AESIZE = 0.5, for the Shell Geometric Model, which is the equivalent to 3444 FEs. The average
relative error, when compared with the most refined mesh, is 5.36%;

• AESIZE = 0.5, for the Beam-Shell Geometric Model. This value corresponds to 32392 FEs and
has an average relative error of 2.93%.

Having defined the quality of the mesh, it is possible to proceed and retrieve the results from the
model.

44
A convergence study on the aerodynamic meshes of the main wing and horizontal stabilizer is also
done, resulting in meshes with 50 spanwise divisions and 32 chordwise divisions that have relative errors
in lift under 1%.

5.3 Beam Geometric Model

In this Section, the results related with the first Geometric Model developed, the BGM, are presented.
As described previously, the aircraft should be tested for both values of the maximum load factor, n = 4.4
and n = −1.8, and so results for these two flight conditions are presented.
To start the model it is necessary to define the initial values for each variable, described on the
following Subsection. The geometric variables RibsIn and RibsOut are not changed during this study,
being fixed in 2 and 5, respectively, and so their values are omitted during the entire Section.

5.3.1 Initial Variables

As described in the SIM definition, the initial set of variables must represent a robust and over-weighted
model, that satisfy all the constrains. The initial set of variables selected to start the model are described
in Table 5.1. The results for each constrain can be checked in Table 5.2

Table 5.1: BGM - Initial Variables.


Variable Value
t in [m] 0.0100
t out [m] 0.0050
chord percentage [-] 0.1000
t wingboom [m] 0.0100
size rearboom [m] 0.2500
t rearboom [m] 0.0500
size vertical [m] 0.1200
t vertical [m] 0.0050
t ribs tail [m] 0.0200
t spars tail [m] 0.0100
size spars tail [m] 0.2000

With this initial set of values, a structural weight of 3448.88Kg is obtained. Looking into the constrains
results, this weight can certainly be lowered, since their values are still really far from the boundaries
defined.

Table 5.2: BGM - Constrains for Initial Variables.


n σmax [MPa] δwingtip [m] δtail [m] Weight [Kg]
4.4 105.81 0.165 0.119
3448.88
-1.8 110.27 -0.126 -0.311

45
Figure 5.3 (a) and (b) represents the Stress and Displacement distributions obtained for the afore-
mentioned set of initial values and for the n = −1.8 flight condition. The value of σmax is 110.27M P a,
being almost half the maximum allowed. The Highest Stress Point (HSP) of the structure is located at
the root of the trailing edge spar. The connection between the boom and the vertical tail, as well as
the connection of the booms with the spars also presents high stress concentrations, indicating that this
could possibly be critical points.
Regarding the vertical displacement, and taking into account that n = −1.8 is being considered, the
structure behaves as expected, with the wing and tail bending downwards, since both W and L have the
same direction. The value of δtail is higher, in module, than δwingtip

(a) Stress Distribution (b) Displacement Distribution

Figure 5.3: BGM - Stress and Displacement Distribution for the Initial Variables

5.3.2 Parametrization

The main goal of this Subsection is to understand how each variable influences the behavior of the
structure and so a graphical representation of the variation of σmax , δwingtip , δtail and W with each
variable in study is presented. These variations are taken directly from Iteration 1 of the SIM. Note that,
due to the fact that some structures are affected simultaneously by two variables, these must be studied
together, to make sure that several combinations are analyzed.
To start the SIM, and as described in Subsection 4.2.4, it is necessary to define the vi variables in
study, as well as their minimum, maximum and increment values - Table 5.3. The variables are ordered
by their mi . In the cases were two variables must be studied together, the value that prevails is the
higher mi between both.

Var 1 and Var 2: t wingboom and chord percentage

According to the mi of each variable, the variables t wingboom and chord percentage are the first to be
analyzed. The results obtained for these variables, by the SIM, are presented in Figure 5.4. The results
are presented for both flight conditions.

46
Table 5.3: BGM - Variables for the optimization process.
i mi [Kg/m] vari varimin varimax variinc Increments
1 75373 t wingboom [m] 0.0020 0.0100 0.0020 5
2 1639 chord percentage [-] 0.0900 0.1300 0.0100 5
3 35413 t rearboom [m] 0.0100 0.0500 0.0100 6
4 10416 size rearboom [m] 0.2250 0.3500 0.0250 6
5 26296 t in [m] 0.0010 0.0100 0.0010 10
6 15725 t spars tail [m] 0.0050 0.0100 0.0010 6
7 864 size spars tail [m] 0.1750 0.3500 0.0250 8
8 14249 t out [m] 0.0005 0.0050 0.0005 10
9 14159 t vertical [m] 0.0010 0.0050 0.0010 5
10 602 size vertical [m] 0.1000 0.1600 0.0100 6
11 1681 t ribs tail [m] 0.0020 0.0200 0.0020 10
Total Number of Simulations 169

Figure 5.4: BGM - variation of constrains and weight with t wingboom and chord percentage variables.

By taking a first look at the results, the value of t wingboom = 0.002m can be clearly removed from
consideration, since it immediately overcomes σmax for the domain in study.
Globally, reducing chord percentage and t wingboom leads to an increase of δwingtip , in module, as
expected. For both flight conditions, since it is reducing the wing mass, the wing gets less rigid and the
aerodynamic loads will tend to increase the bending amplitude upwards, for n = 4.4, and downwards,
for n = −1.8. Regarding δtail , the variables in study have little to none influence on the positive load
factor. On the other hand, for n = −1.8, reducing chord percentage and t wingboom clearly affect δtail ,
which increases in module.
The progress of the σmax curves is practically identical for both values of n, being slightly lower for the
negative one. For the same value of chord percentage, decreasing t wingboom increases the σmax value

47
of the structure. The same behavior occurs by swapping the values. Also, it is clear that σmax always
varies, there are no horizontal levels. This is due to the fact that, as previously seen in Figure 5.3, the
HSP is located in the trailing edge spar, meaning that, even the smallest variation of chord percentage
and t wingboom will lead to a variation of the maximum stress of the structure.
All in all, considering all the constrains, the values that minimize the weight, in Iteration 1, are
chord percentage = 0.12 and t wingboom = 0.004m. These variations keep the highest stress point
at the spar’s root, for both flight conditions.

Var 3 and Var 4: t rearboom and size rearboom

The next component that is analyzed by the SIM are the booms, which are defined by the t rearboom
and size rearboom. The results obtained for these variables are exposed in Figure 5.5, for the two flight
conditions in study.

Figure 5.5: BGM - variation of constrains and weight with t rearboom and size rearboom variables.

Starting for the stress distribution, and looking into the positive load factor, it can be seen that,
decreasing size rearboom, produces little to no change to σmax , while decreasing t rearboom increases
it. On the other hand, t rearboom and size rearboom have a huge influence in the value of σmax . As
it could be seen in Figure 5.3, the HSP of the structure is at root of the spar. The spars are acting
as cantilever beams, supporting all the structural weight. Furthermore, since L and W have the same
direction, the stress distribution in the structure is clearly more affected by mass changes for n = −1.8
than it is for n = 4.4. This way, decreasing the boom’s mass decreases the stress concentration at the
spar’s root. Eventually, decreasing too much the size rearboom and t rearboom will move the critical
point of the structure to the booms, and a spike in the σmax variation should be observed. This scenario
can be clearly seen for t rearboom = 0.01m, where σmax highly increases if size rearboom decreases
past 0.25m.

48
Regarding the displacements, the behaviors are the expected. Since the rearbooms are mainly
responsible for supporting the tail structure, δwingtip remains essentially constant. For δtail , reducing
the boom robustness allows it to have an higher bending amplitude, so its values increases (in module),
either for the reduction of size rearboom and t rearboom.
Finally, and looking into the results, size rearboom = 0.3000 and t rearboom = 0.0100 are the values
selected, being the HSP still located at the wing spar’s root.

Var 5: t in

Continuing for the next variable, the results obtained for t in variation are presented in Figure 5.6. Once
again, the values are represented for both flight conditions.

Figure 5.6: BGM - variation of constrains and weight with t in variable.

Looking into the displacement constrains, δwingtip and δtail , it is clear that they are almost not affected
by the variation of t in. These results are expected since the spars are the structural components that
supports most of the aerodynamic loads and t in does not affect the tail in any way.
Regarding the stress distribution, it remains unaffected for the most part of the domain in study. The
peaks occur around t in = 0.003m, for n = −1.8, and t in = 0.002m, for n = 4.4, indicating that these
are the values for which critical stress point is transferred to the ribs.
Looking into the results, t in = 0.003m is the value selected and fixed for the following analyses.
Selecting this value moves the HSP to wing ribs, for the n = −1.8 flight condition. For n = 4.4, the HSP
is still located at the spar’s root.

Var 6 and Var 7: t spars tail and size spars tail

After the aforementioned variables already studied, the t spars tail and size spars tail are the next on
the list. Figure 5.7 shows the results obtained.
Changing these variables has almost no influence in the δtail and δwingtip constrains, which is ex-
pected, due to the really small variation in weight.

49
Figure 5.7: BGM - variation of constrains and weight with t spars tail and size spars tail variables.

Regarding σmax , it is interesting to observe two completely different behaviors for n = −1.8 and
n = 4.4. For the first one, there is exactly no change for the different values of t spars tail and
size spars tail. This is due to the fact that, for n = −1.8, the HSP is located in the ribs, and t spars tail
and size spars tail have completely no influence in the rib’s behavior. Since the variables were not
sufficiently reduced to move the HSP to the tail spars, it remains located at the wing ribs.

For n = 4.4, it is completely different, since the HSP is still located at the wing spars. Reducing
the weight on vertical tail increases the booms bending amplitude. These provokes a higher bending
moment on connection between the spar and the boom, which is then transmitted to the spar’s wing root,
increasing σmax . This way, decreasing t spars tail and size spars tail has a global effect of increasing
σmax , keeping the HSP located at the wing spar’s root, at least for the domain in study.

All in all, the values t spars tail = 0.005m and size spars tail = 0.175m are selected.

Var 8: t out

Proceeding, the next variable in study is t out, which has the results presented in Figure 5.8.

Equally as the t in, t out has little to none influence on the displacement constrains. Regarding σmax
variation, the value has a huge increase after t out = 0.004m. For n = −1.8, this is easy to explain, since
the HSP is already located at the wing ribs. For n = 4.4, lowering t out also moves the HSP to the wing
ribs.

Due to this, the fixed value is t out = 0.004m. This value moves the HSP to the wing ribs for both
flight conditions. For n = 4.4, it is important to refer that the spar’s root remain also highly stressed,
remaining a critical point of the structure.

50
Figure 5.8: BGM - variation of constrains and weight with t out variable.

Var 9 and Var10: t vertical and size vertical

The results obtained for the variables t vertical and size vertical are presented in Figure 5.9, as usual
for both flight conditions in study.

Figure 5.9: BGM - variation of constrains and weight with t vertical and size vertical variables.

As described previously, these two variables are responsible for changing the geometry of the vertical
tail. Since the vertical is supported by the booms, there is almost no noticeable variation in δwingtip .
Regarding δtail , the behavior is exactly the same as the other tail variables. Decreasing both t vertical
and size vertical reduces the tail weight, allowing the boom to bend further and increasing δtail , in
module, for both flight conditions.
For σmax , there are some combinations of t vertical and size vertical that do not produce any kind
of variation on it. This is due to the fact that, as explained, the HSP is located, at this point, on the wing
ribs, which are not influenced by t vertical and size vertical. Eventually, decreasing these variables will

51
lead to a loss of rigidity of the vertical tail, which can cause the HSP to move to this structure. This
behavior can be observed in the σmax distributions, for the lower values of t vertical.
Considering all the constrains, the values selected are t vertical = 0.002m and size vertical =
0.1400m. These values keep the HSP at the wing ribs.

Var 11: t ribs tail

Finally, the last variable of the BGM is the t ribs tail. The results are presented in Figure 5.10.

Figure 5.10: BGM - variation of constrains and weight with t ribs tail variable.

As defined, the tail ribs are the components that have the lowest impact on the global structural
weight. Due to this, varying t ribs tail almost does not change δwingtip and δtail . Regarding the stress
distribution, σmax remains constant for the higher thickness values. When t ribs tail is too reduced, the
HSP moves to the tail ribs, causing the spikes obtained in the graphical representation.
To wrap these results, t ribs tail = 0.008m is the value selected.

5.3.3 Final Results

The final values of all the variables, for the BGM, are presented in Table 5.4. Their respective constrains
values and weight, for both flight conditions, are presented in Table 5.5. These results were obtained
after two iterations of the Structural Improvement Model.
Looking into the final results, the model was able to reduce the structural weight from 3448.88Kg to
811.85Kg, representing a reduction of 76.46%, without compromising the structural integrity or overcom-
ing the boundaries of the constrains previously defined. The W obtained is clearly underestimated, since
the BGM fails to represent several components, such as the wing skin, tail skin and wing reinforcements,
and makes an incorrect representation of the wing and tail ribs.
Looking for the variables, it is clear that all the thicknesses were extremely overestimated, since all
of them were reduced. However, it is curious to see that, to decrease W , all the size variables of the
hollow cross sections were increased. Considering a simple example of a cantilever beam, with length

52
Table 5.4: BGM - Final values for the variables.
i vari I0 I1 I2 Final Variation [
1 t wingboom [m] 0.0100 0.0040 0.0024 -76 %
2 chord percentage [-] 0.1000 0.1200 0.1600 60 %
3 t rearboom [m] 0.0500 0.0100 0.0040 -92 %
4 size rearboom [m] 0.2500 0.3000 0.3750 50 %
5 t in [m] 0.0100 0.0030 0.0030 -70 %
6 t spars tail [m] 0.0100 0.0050 0.0010 -90 %
7 size spars tail [m] 0.2000 0.1750 0.1750 -12 %
8 t out [m] 0.0050 0.0040 0.0040 -20 %
9 t vertical [m] 0.0050 0.0020 0.0020 -60 %
10 size vertical [m] 0.1200 0.1400 0.1400 17 %
11 t ribs tail [m] 0.0200 0.0080 0.0075 -63 %

Table 5.5: BGM - Final results.


n Constrain I0 I1 I2 Final Variation
δwingtip [m] -0.126 -0.159 -0.109 -13 %
-1.8 σmax [MPa] 110.27 177.47 180.82 64 %
δtail [m] -0.311 -0.419 -0.441 42 %
δwingtip [m] 0.165 0.234 0.164 -1 %
4.4 σmax [MPa] 105.81 182.03 176.74 67 %
δtail [m] 0.119 0.298 0.361 204 %
W [Kg] 3448.88 1237.58 811.85 -76 %

l, and with a hollow square cross section, with size a and thickness t. The cross section has a surface,
S, defined by

S = a2 − (a − 2t)2 = 4t(a − t) (5.1)

which can be used to obtain the weight of the beam using

W = S l ρ g = [4t(a − t)] l ρ g (5.2)

As it can be seen, due to being a second order variable, reducing t has a much bigger impact in
weight reduction when compared with a. This fact, along with the need of satisfying the constrains,
leads to the behavior identified in this study, which is the thickness reduction and size increase.
Finally, it is possible to analyze the stress and vertical displacements distributions obtained for each
flight condition. Figure 5.11 represents these distributions for n = −1.8, while Figure 5.12 represents
them for n = 4.4.
Starting by the negative flight condition, and for this final set of variables, the HSP is located at wing
ribs. However, looking into the distribution, it is clear that the tail spar is also facing a stress close to the
σmax allowed. The spars experience higher stresses near the root and the boom near the connection
with the wing. Both are expected behaviors, since both components act as cantilever beams. The
vertical displacement distribution is also expected, with negative values of δwingtip and δtail .

53
(a) Stress Distribution (b) Displacement Distribution

Figure 5.11: BGM - Stress and Displacement Distribution, for n = −1.8, using the final values of the
variables.

(a) Stress Distribution (b) Displacement Distribution

Figure 5.12: BGM - Stress and Displacement Distribution, for n = 4.4, using the final values of the
variables.

Equally as n = −1.8, for n = 4.4 the HSP is located at the wing ribs. However, it is possible to see
areas of high stress concentration at the wing and tail spar’s root. The wing spars are more stressed
than they are for n = −1.8 due to the higher magnitude of the aerodynamic loads, which provoke higher
bending moments across the structure. The displacement distribution is also expected, with positive
values of δwingtip and δtail .

Said so, and for the final set of variables, the critical points of the BGM are then the wing ribs, the
wing spar’s root and the tail spar’s root. Some other locations that present higher stress concentrations
are the connection points between the wing and the booms, between the booms and the vertical tails
and between the vertical tail and horizontal.

54
5.4 Shell Geometric Model

In this Section, the results related with the second Geometric Model developed, the SGM, are presented.
Following the same methodology, tests are performed for both values of the maximum load factor, n =
4.4 and n = −1.8.
Once again, the geometric variables RibsIn and RibsOut are not changed during this study, being
fixed in 2 and 5, respectively, and so their values are omitted during the Section.

5.4.1 Initial Variables

Equally as for the BGM, the initial set of variables must represent an over-weighted model that satisfies
all the constrains. This initial set of values is presented in Table 5.6, being the constrains values shown
in Table 5.7

Table 5.6: SGM - Initial Variables.


Variable Value
t in [m] 0.300
t out [m] 0.200
chord percentage [-] 0.100
t wingshell [m] 0.010
t rearboom [m] 0.300
t vertical [m] 0.020
t ribs tail [m] 0.005
t spars tail [m] 0.008
t skin tail [m] 0.005

Table 5.7: SGM - Constrains for Initial Variables.


n σmax [MPa] δwingtip [m] δtail [m] Weight [Kg]
4.4 71.62 0.006 0.112
17137.16
-1.8 164.21 -0.019 -0.390

This initial inputs satisfy all the constrains. However, the structure produced presents an enormous
weight value of 17137.16Kg when compared with the expected weight of 1500Kg. Nevertheless, this is
expected, since the Shell FEs are not indicated to represent beam type structures. However, this value
can certainly be reduced using the SIM, specially due to the reduced values of the constrains.
Since σmax is higher for the n = −1.8, the initial stress and displacement distributions are displayed
for this specific flight condition - Figure 5.13 (a) and Figure 5.13 (b), respectively.
As it can be seen, the stress distribution is zoomed near the wing and booms connection. This was
done to empathize the location of the HSP, which is precisely in the connection between the booms
and the wing. The SGM does not allow the representation of the wing reinforcement, and so the stress
caused by the bending and shear of the boom is passed also to the wing skin, causing the stress

55
(a) Stress Distribution (b) Displacement Distribution

Figure 5.13: SGM - Stress and Displacement Distribution for the Initial Variables

concentration represented. Regarding the vertical displacement distribution, the results are expected,
being all negative.
Regarding the flight condition n = 4.4, the HSP is located in the connection between the vertical and
the horizontal tail.

5.4.2 Parametrization

Following the same methodology, the vi variables in study are firstly defined, as well as their minimum,
maximum and increment values - Table 5.8. Once again, the variables are ordered by their mi . Please
note that for the SGM there are no variables studied together.

Table 5.8: SGM - Variables for the optimization process.


i mi [Kg/m] vari varimin varimax variinc Increments
1 198015 t wingshell [m] 0.0010 0.0100 0.0010 10
2 40656 t skin tail [m] 0.0005 0.0050 0.0005 10
3 21802 chord percentage [-] 0.0010 0.1000 0.0010 10
4 19565 t in [m] 0.0300 0.3000 0.0300 10
5 10059 t out [m] 0.0200 0.2000 0.0200 10
6 8740.4 t vertical [m] 0.0020 0.0200 0.0020 10
7 5603.4 t rearboom [m] 0.0300 0.3000 0.0300 10
8 2752.1 t spars tail [m] 0.0008 0.0080 0.0008 10
9 1152.9 t ribs tail [m] 0.0005 0.0050 0.0005 10
Total Number of Simulations 90

Var 1: t wingshell

Following the order defined by the mi ’s, the wing skin is the first component to be studied. The thickness
of the component is controlled by the t wingshell variable, being its results presented in Figure 5.14.

56
Figure 5.14: SGM - variation of constrains with t wingshell variable.

Regarding the displacements, the behaviors of δwingtip and δtail are quite clear to explain. Reducing
the wing skin thickness affects majorly the displacements of the wing, meaning that δtail remains essen-
tially the same for the domain in study. Decreasing t wingshell reduces the wing rigidity, meaning that
aerodynamic increase the bending amplitude which causes and increase, in module, of δwingtip .
Regarding the stress distribution, and for n = 4.4, σmax remains constant until t wingshell = 0.003m.
After that, σmax increases, indicating that the HSP moved from the tail to the wing skin. Checking the
FEA, the HSP is now on the upper side of the wing, located in the skin near the wing root.
For n = −1.8, the initial decrease in t wingshell actually reduces the σmax value, due to the decrease
of the inertial loads, which have the same direction as the L. However, decreasing its value too much
ends up having the contrary effect, which happens exactly for thicknesses below 0.002m.
Taking into account all the constrains, t wingshell = 0.002m is the value selected, indicated that the
initial value defined was clearly overestimated.

Var 2: t skin tail

The next variable to be studied is t skin tail, which represents the thickness of the tail skin - Figure
5.15.
Analyzing the results obtained, it is clear that δwingtip is almost not affected by the changes in
t skin tail, contrarily to δtail , which increases, in module, for lower values of t skin tail. The behav-
ior has already been explained - reducing the mass of the tail decreases its rigidity, which makes the
aerodynamic loads of the tail increase the bending amplitude of the boom.
Regarding the stress amplitude, it remains constant for the first decrements. Around t skin tail =
0.003m, it starts to increase. In the FEA, it can be see that the HSP moves to the tail skin, near the
connection between the vertical and horizontal tail, which explains this increase of σmax .
To conclude the analysis of the skin tail structure, the value t skin tail = 0.0025m is selected, since
it minimizes the structural weight while satisfying all the constrains.

57
Figure 5.15: SGM - variation of constrains with t skin tail variable.

Var 3: chord percentage

The next variable on the list is chord percentage, which is responsible for the geometry of the wing spars.
The results are illustrated in Figure 5.16.

Figure 5.16: SGM - variation of constrains with chord percentage variable.

The σmax distribution presents essentially the same progress for both flight conditions, being the
value for n = −1.8 always higher. Essentially, while decreasing chord percentage, it can be seen that
σmax faces a very small reduction from the initial value until around 0.01. For lower values, this values
faces a high increase, which happens when the HSP moves from the horizontal tail to the wing spar
location.
Regarding the vertical displacements, and as expected, δwingtip is highly influenced by this variable,
since it is changing the mass of the wing. On the other side, δtail is almost not influenced.
To finish the analysis of this variable, the value selected is chord percentage = 0.006. This value
moves the HSP to the wing skin, located near the root. The skin gets an higher stress concentration
due the loss of mass of the wing, which leads to a increase of the wing bending moment caused by
the aerodynamic loads, relatively to the wing’s root. This value causes a huge decrease in the global

58
structural weight, clearly showing that this variable was heavily over-dimensioned.

Var 4: t in

After chord percentage, the variable with the highest mi is t in, which is one of the variables responsible
for modelling the wing ribs geometry. The results are presented in Figure 5.17.

Figure 5.17: SGM - variation of constrains with t in variable.

For both flight conditions, it is possible to see that the δwingtip is almost not affected by t in’s variation,
since the spars are the responsible for withstanding the major loads of the wing. δtail , on its side, is more
affected for the negative n than it is for the positive. This is also reflected on the σmax distribution, which
faces a much higher increase for n = −1.8. Due to the negative load factor, the wing faces higher
stresses and displacements, since L and W have the same direction, explaining the bigger impact
caused by t in variation.
Said so, the value selected must be t in = 0.2700m. Lower values make σmax to overcome the
maximum allowed stress value during flight at n = −1.8.

Var 5: t out

To finish the analysis of the wing ribs, t out must also be analysed - Figure 5.18.
The behavior of changing t out is similar to the one described for t in: δwingtip and δtail almost don’t
vary, being the difference essentially in σmax . On this case, the inclination of the curves is similar for both
flight conditions, being higher for n = −1.8. As previously explained, the negative load factor causes the
aerodynamic and inertial loads to have the same direction, increasing the stress values.
The value of t out is then fixed in 0.1600m. This value, along with the value previously defined for
t in, keep the HSP in the wing skin. However, the HSP for the negative load factor presents now a
value of σmax = 183.98M P a, meaning that, even the smallest variation of a variable, can cause σmax to
overcome the maximum allowed stress.

59
Figure 5.18: SGM - variation of constrains with t out variable.

Var 6: t vertical

The next variable in study is responsible by defining the geometry of the vertical tail. This variable is the
t vertical and its results are presented in Figure 5.19.

Figure 5.19: SGM - variation of constrains with t vertical variable.

The most interesting behavior to observe here is that, reducing this variable, leads to a decrease of
σmax for the n = −1.8. As explained, for this flight condition, the inertial and aerodynamic loads have
the same direction. Reducing the mass on the tail reduce the bending moment acting on the wing,
which then, consequently, reduces the intensity of the resulting stresses. The value of δwingtip remains
constant, since this constrain is essentially affected by the wing variables. Regarding δtail , the behavior
of the curves is concordant with σmax . For the negative load factor, the bending moment caused by the
tail is reduced, which causes δtail to also decrease, in module. For the positive load factor, since the
direction of the aerodynamic forces points in the opposite direction of the inertial forces, reducing the
mass of the tail leads to an increase of the bending moment, consequently causing δtail to increase.
The value selected is then t vertical = 0.002m, the lowest value in study during this iteration. This
value keeps the HSP located at the wing, for both flight conditions.

60
Var 7: t rearboom

The model proceeds now to the analysis of t rearboom, which is the responsible to define the geometry
of the boom. The results for this variable can be seen in Figure 5.20.

Figure 5.20: SGM - variation of constrains with t rearboom variable.

The boom is the structure responsible for supporting the entire tail section, meaning that it should
be the main influence of the δtail constrain. This assumption is clearly correct, as it can be seen by the
great amplitudes of the results obtained for δtail . Since the rigidity of the boom is being reduced, and the
aerodynamic loads are still the same, the bending moment increases, increasing the tail displacement
value (in module). This increase in the bending moment increases the stress concentration on the wing
structure, which is where the HSP is located, being this the reason why the values for σmax also have
a huge increase. On the other side, and as expected, δwingtip remains essentially constant, since this
constrain is mainly driven by the wing variables.

Due to the δtail constrain for n = −1.8, it is not possible to lower t rearboom, meaning that it must
stay with the value of 0.3000m. The HSP location is still on the wing, for both n values.

Var 8: t spars tail

Proceeding with the parametrization process, the next variable to be studied is the t spars tail. The
graphical representation of the constrains is represented in Figure 5.21.

Starting for the flight condition of n = −1.8, there is a decrease of σmax with the reduction of
t spars tail, having this behavior the same explanation provided for the t vertical variable - reducing
the tail’s mass, decreases the bending moment caused by the tail, reducing the stress transmitted to the
wing structure. For n = 4.4 the effect is precisely the reverse being, however, the changes almost null.
Regarding the displacement constrains, they remain basically the same.

None of the values in the domain in study overcomes the maximum values allowed for he constrain,
so the value t spars tail = 0.0008m is selected. The HSP location still remains the same.

61
Figure 5.21: SGM - variation of constrains with t spars tail variable.

Var 9: t ribs tail

Finally, the last variable of the SGM is t ribs tail, which modules the thickness of the tail ribs. The results
for this variable are presented in Figure 5.22.

Figure 5.22: SGM - variation of constrains with t ribs tail variable.

The variable causes the exact same behavior on the structure as the t spars tail variable, so there is
no need to repeat the explanations. For the σmax variation, the stress increases when the HSP changes
its location to the tail ribs, due to the lower thicknesses.
The value selected t spars tail = 0.0035, which concludes the analysis of all the variables.

5.4.3 Final Results

The final values of all the variables, for the SGM, are presented in Table 5.9. Their respective constrains
values and weight, for both flight conditions, are presented in Table 5.10. Equally as for the results of
the BGM, these results were obtained after two iterations of the SIM.
Analyzing the final results obtained, the model reduces the structural weight from 17137.16Kg to
8890.79Kg, which represents a reduction of 48%. This value is extremely high when compared with the

62
Table 5.9: SGM - Final values for the variables.
i vari I0 I1 I2 Final Variation
1 t wingshell [m] 0.0100 0.0020 0.0019 -81 %
2 t skin tail [m] 0.0050 0.0025 0.0025 -50 %
3 chord percentage [-] 0.1000 0.0060 0.0056 -94 %
4 t in [m] 0.3000 0.2700 0.2560 -15 %
5 t out [m] 0.2000 0.1600 0.1540 -23 %
6 t vertical [m] 0.0200 0.0020 0.0015 -93 %
7 t rearboom [m] 0.3000 0.3000 0.2760 -8 %
8 t spars tail [m] 0.0080 0.0008 0.0005 -94 %
9 t ribs tail [m] 0.0050 0.0035 0.0031 -38 %

Table 5.10: SGM - Final results.


n Constrain I0 I1 I2 Final Variation
δwingtip [m] -0.019 -0.118 -0.123 548 %
-1.8 σmax [MPa] 164.21 177.52 183.90 12 %
δtail [m] -0.390 -0.425 -0.461 18 %
δwingtip [m] 0.006 0.094 0.101 1673 %
4.4 σmax [MPa] 71.62 156.98 169.53 137 %
δtail [m] 0.112 0.185 0.213 90 %
W [Kg] 17137.16 9405.75 8890.79 -48 %

one obtained for the Beam Geometric model, but expected. This model is built using only Shell FEs,
which are normally used to modules thin web structures. Examples of this type of components in the
aircraft are the wing and tail skins, as well as the ribs. Well, in this model, all the components are being
modeled using Shell FEs, which is clearly a non optimal approach, specially due to the representation of
the spars and the booms. Due to the nature of this FEs, the thickness of the Shell elements representing
the booms and the spars are clearly oversized, in order to compensate for the reduced capacity of the
Shell elements to withstand bending loads.
However, regarding the structural representation of the components, this model is more accurate
than the BGM, since it allows to represent correctly the wing and tail ribs and skin. Nevertheless, and
since it is independent from the Geometric Model used, the SIM presents some good results, since it
is able to reduce the weight in almost an half of the initial value, while keeping the constrains in the
boundaries defined.
Using this final values for the geometric variables, it is then possible to analyze the stress and vertical
displacements distributions obtained for each flight condition. Figure 5.23 represents these distributions
for n = −1.8, while Figure 5.24 represents them for n = 4.4.
Looking into the final distributions obtained, it is clear that the highest stress is obtained for the neg-
ative load factor in study. As explained previously, this is due to the fact that both the aerodynamic and
structural loads have the same direction. The HSP for both flight conditions is located at the wing, near
the boom connection for n = −1.8 and near the wing root for n = 4.4. Looking specifically to the wing,
and due to the aerodynamic loads, it is possible to observe clearly that the wing faces higher stress in

63
(a) Stress Distribution (b) Displacement Distribution

Figure 5.23: SGM - Stress and Displacement Distribution, for n = −1.8, using the final values of the
variables.

(a) Stress Distribution (b) Displacement Distribution

Figure 5.24: SGM - Stress and Displacement Distribution, for n = 4.4, using the final values of the
variables.

the root, decreasing progressively towards the tip. The booms have also higher stress concentrations
near the connections with the wing, indicating that these are also critical points for the structure. Fur-
thermore, the connection between the booms and vertical tail can also be considered a critical structural
point, as it can be seen by Figure 5.23 a) and Figure 5.24 a).

Regarding the displacement distributions, the results are more than expected . For the negative load
factor, all the displacements are negative, while for the positive load factor, all the displacements are
positive. For both flight conditions, δtail is higher, in module, than δwingtip .

5.5 Beam-Shell Geometric Model

In this Section, the results of the final model, the BSGM, are presented.

64
5.5.1 Initial Variables

Equally as it is done for the two first Geometric Models, it is necessary to define an initial set of values
as the starting point for the model to run.

The initial set of values are presented in Table 5.11 and its corresponding weight and constrain values
(δwing , δtip , σmax ) are presented in Table 5.12, for the two load factor limits.

Table 5.11: BSGM - Initial Variables.


Variable Value
t in [m] 0.100
t out [m] 0.050
chord percentage [-] 0.080
t wingboom [m] 0.005
t wingshell [m] 0.004
t reinforcement [m] 0.010
size rearboom [m] 0.250
t rearboom [m] 0.050
size vertical [m] 0.100
t vertical [m] 0.006
t ribs tail [m] 0.010
size spars tail [m] 0.100
t spars tail [m] 0.005
t skin tail [m] 0.007

Table 5.12: BSGM - Constrains for Initial Variables.


Load Factor δwingtip [m] δtail [m] σmax [MPa] W [Kg]
-1.8 -0.034 -0.212 105.89 6230.80
4.4 0.036 0.036 64.31 6230.80

With these initial values, the structure weights 6231.80Kg. This huge weight, as well as the constrain
values, indicates that the structure is clearly oversized, meaning that the SIM should be able to reduce
this structural weight.

The stress and vertical displacement distributions for the initial set of variables can be seen in Figure
5.25 (a) and (b) respectively. These distributions are obtained for n = −1.8.

Regarding the stress distribution, the wing root (Figure 5.25 (a)) is facing an higher stress when
compared with the tip. This is an expected behavior, due to higher aerodynamic loads and due to
the fact that the wing is withstanding all the weight from the boom and tail structure. The displacement
distribution (Figure 5.25 (b)) also presents an expected behavior. Due to the negative load factor, δwingtip
and δtail are also negative, being the value of δtail higher, in module.

For n = −1.8, and for this initial set of variables, HSP is located in one of the inner ribs. For the
positive load factor, HSP is located in leading edge wing spar.

65
(a) Stress Distribution (b) Displacement Distribution

Figure 5.25: BSGM - Stress and Displacement Distribution for the Initial Variables

5.5.2 Parametrization Results

To parametrize the structure, and as defined above, it is necessary to define the order of variables in
study. Said so, the initial inputs for the parametrization studies and to run the SIM are defined in Table
5.13.

Table 5.13: BSGM - Inputs for the optimization process.


i vari mi [Kg/m] varimin varimax variinc Increments
1 t wingshell 198462 0.0004 0.0040 0.0004 10
2 t wingboom 123141 0.0010 0.0050 0.0010 6
3 chord percentage 5078 0.05000 0.11000 0.01000 7
4 t skin tail 40656 0.0007 0.0070 0.007 10
5 t reinforcement 30204 0.0010 0.0100 0.0010 10
6 t rearboom 27535 0.0100 0.0500 0.0100 6
7 size rearboom 8099 0.1500 0.3500 0.0250 9
8 t spars tail 20103 0.0010 0.0050 0.0010 6
9 size spars tail 1093 0.0250 0.1750 0.0250 7
10 t in 19565 0.01 0.1000 0.01 10
11 t vertical 10740 0.001 0.0060 0.001 6
12 size vertical 716 0.07 0.1300 0.01 7
13 t out 9118 0.005 0.0500 0.005 10
14 t ribs tail 1153 0.001 0.0100 0.001 10
Total Number of Simulations 240

Following the order defined above, the parametrization results are then obtained.

Var 1: t wingshell

The first variable to be studied is t wingshell, since it has a mi = 198462Kg/m, making it the variable
with the highest weight variation sensibility.

66
Figure 5.26: BSGM - variation of constrains with t wingshell variable.

Starting by looking into the variations of σmax , and knowing the initial location of the HSP for each
load factor, it is possible to understand the behavior observed. Regarding n = 4.4, the HSP is located
at the root of the wing spars. Reducing t wingshell leads to a reduction of the wing’s mass and rigidity.
Since the aerodynamic loads remain the same, the wing bending moment increases, increasing the
stress concentration at the spar’s root. For n = −1.8, the HSP is located at one of the wing ribs. Since
the spars are the main responsible for withstanding the wing bending loads, decreasing t wingshell
almost does not impact the stress variation in the wing ribs. For values under t wingshell = 0.0012m,
there is a huge increase in σmax , meaning that, for lower values, the HSP moves to the wing skin.
Regarding the displacement constrains, the behavior is the expected - δtail presents little to none vari-
ation, while δwingtip increases, in module, due to increase of the wing bending moment when t wingshell
decreases.
Said so, and looking into the constrains, the value fixed is t wingshell = 0.0012m. For this variation,
the HSP remain in the same location for n = −1.8, moving to the wing skin for n = 4.4.

Var 2 and Var 3: t wingboom and chord percentage

Following the Structural Improvement Model, t wingboom and chord percentage are the next variables
in study. Their results are presented in Figure 5.27.
Analyzing the results, it can be seen that the influence of the t wingboom and chord percentage is
much similar to the described, for the same variables, but for the Beam Geometric Model. This is an
expected behavior, since the structures, on both models, are being modeled using the same type of FE,
the BEAM189.
In short, reducing chord percentage and t wingboom leads to an increase of δwingtip , in module.
The wing looses rigidity and the aerodynamic loads will tend to increase the bending amplitude up-
wards, for n = 4.4, and downwards, for n = −1.8. Regarding δtail , and as already stated, reducing
chord percentage and t wingboom clearly affect δtail , which increases in module. On the other side,
variables do not have much influence on δtail during flight with the limiting positive load factor.

67
Figure 5.27: BSGM - variation of constrains with t wingboom and chord percentage variables.

The σmax variation has already been described in Sub-subsection 5.3.2. Said so, the values selected
are chord percentage = 0.09 and t wingboom = 0.003m, keeping the location of the HSP in the wing
ribs, for n = −1.8, and the wing skin, for n = 4.4.

Var 4: t skin tail

As it is presented in Table 5.13, the wing skin has also a big impact in the structural weight. Said so, it is
the next variable to be studied - Figure 5.28.

Figure 5.28: BSGM - variation of constrains with t skin tail variable.

Once again, these results are quite similar to the ones described during Sub-subsection 5.4.2, for
the same variable but referent to the Shell Geometric Model - δwingtip remains constant for the different
values of t skin tail, while δtail increases, in module, for lower values of t skin tail.
Regarding the stress amplitude, it remains essentially constant for the first decrements. Around
t skin tail = 0.0021m, it starts to increase. Lowering the thickness increases the stress at the tail skin.

68
Eventually, for really low t skin tail, the HSP moves to the tail, causing the huge increases shown in the
σmax results.
Due to this, the value t skin tail = 0.0021m must be established for the next variations, causing the
HSP to move to the tail skin, for n = −1.8, while it remains located at the wing skin for n = 4.4.

Var 5: t reinforcement

The next variable to be studied is the t reinf orcement, which is responsible by the geometry of the wing
reinforcement structure. This variable was not possible to module in both previous models, due to the
limitations of each one. The results for this variable are present in Figure 5.29.

Figure 5.29: BSGM - variation of constrains with t reinforcement variable.

Looking into the results obtained, it is clear that, for n = 4.4, none of the constrains is affected by
varying t reinf orcement. For n = −1.8, however, it is different. For values lower than t reinf orcement =
0.004m, it is visible the increase of σmax , meaning that the reinforcement starts to be the higher stressed
component of structure. An higher stress is accompanied by an higher displacement. Since the booms
are connected to the wing in the location of the reinforcement component, their displacements are
affected. Since the booms displacements are affected, δtail is also affected, explaining the δtail variation
observed for lower thickness values.
As a result of that, the value t reinf orcement = 0.003m is established, causing the HSP to move to
the wing reinforcement, for the negative flight condition. For n = 4.4, the HSP remains located at the
wing skin.

Var 6 and Var 7: t rearboom and size rearboom

The model proceeds with the analysis of t rearboom and size rearboom, whose results are exposed in
Figure 5.30.
Once again, and as expected, the behavior of the constrains with the change of t rearboom and
size rearboom is quite similar to the description made for the Beam Geometric Model, in Sub-subsection
5.3.2.

69
Figure 5.30: BSGM - variation of constrains with t rearboom and size rearboom variables.

Starting with the stress distribution, and considering n = 4.4, there is almost no variation of σmax .
For n = −1.8, however, the behavior is different. Reducing the mass of the booms reduces the bending
moment and shear transmitted from all the boom and tail structure to the wing and reinforcements.
Since the HSP is located at the reinforcement, this causes a stress relief in this components, decreasing
the value of σmax . If t rearboom and size rearboom are heavily decreased, the stresses in the boom
component itself overcome the stresses in the wing, making it the critical component. This is a behavior
that can be observed, as an example, if one considers t rearboom = 0.01m and values of size rearboom
lower than 0.25m.
As for the displacements, is expected - δwingtip remains constant while δtail is widely affected, in-
creasing, in module, with the reduction of t rearboom and size rearboom.
Finally, and looking into the results, size rearboom = 0.25m and t rearboom = 0.01m are the values
selected. The HSP is still located at the wing reinforcement and the wing skin, for n = −1.8 and n = 4.4,
respectively.

Var 8 and Var 9: t spars tail and size spars tail

The following variables in study are t spars tail and size spars tail - Figure 5.31.
Starting by the displacement constrains, the results are the expected ones. Since these variables
are responsible for the geometry of a component from the horizontal tail, δwingtip remains constant. On
the other side, δtail present some minor changes, related with the weight variation of the tail.
The constrain that is most affected is clearly σmax . For the higher values of size spars tail, and
for all the values of t spars tail in study, there is no change of the σmax value. However, decreasing
size spars tail eventually leads to an increase in the tail spars’ stress distribution. Also, as it should, this
effect occurs earlier for lower thicknesses. As an example, for t spars tail = 0.001m, the stress increase
is observed for values of size spars tail lower than 0.13m. For t spars tail = 0.005m, the σmax only
increases for values of size spars tail lower than 0.08m.

70
Figure 5.31: BSGM - variation of constrains with t spars tail and size spars tail variables.

With this results, and considering the main goal of lowering W, the values t spars tail = 0.001m and
size spars tail = 0.15m are defined. For these values, the HSP remains located in the same structures
as before - the wing reinforcement, for n = −1.8, and the wing shell, for n = 4.4.

Var 10: t in

The variable t in is the one that follows. The results are presented in Figure 5.32.

Figure 5.32: BSGM - variation of constrains with t in variable.

Looking into the results, it is clear that t in, despite not making much changes in the geometric
constrains, has a huge impact in global weight of the structure. Also, taking into account the constrains
variation, it is clear that t in is clearly over-dimensioned.
The results for δwingtip and δtail are expected and have already been explained in Sub-subsection
5.3.2. The interesting behavior to analyze here is the σmax curve.
For n = −1.8, decreasing the t in variable leads to a decrease of σmax . As explained, for the negative

71
load factor, the weight has the same direction as the lift aerodynamic forces acting on the wing. So,
reducing t in has an initial stress relief effect on the structure. However, by further decreasing t in, the
rigidity of the wing also decreases, meaning that, due to its location, the reinforcement starts to withstand
more loads that were previously distributed for other components, specially the ribs nearby. Since the
wing reinforcement still remains the component with higher stress, this causes σmax to increase.
This way, looking essentially into the σmax variation, the value t out = 0.04m is fixed. The HSP still
remains located at the wing reinforcement, for n = −1.8, and at the wing shell, for n = 4.4.

Var 11 and Var 12: t vertical and size vertical

The next component in study is the vertical tail, which is defined by t vertical and size vertical.

Figure 5.33: BSGM - variation of constrains with t vertical and size vertical variables.

Once more, these results are extremely similar to the ones obtained for the same component, in the
study of the BGM.
Since these variables are responsible for the tail geometry, the values of δwingtip do not change.
Regarding δtail , the behavior is the same as the other tail variables. Decreasing both t vertical and
size vertical reduces the tail weight and rigidity, allowing it to have higher displacements. This causes
an increase of δtail , in module, for each one of the flight conditions in study.
For σmax , there are some combinations of t vertical and size vertical that do not produce any vari-
ation in σmax . As explained, the HSP is located, for both load factor, at the wing, which is almost not
influenced by t vertical and size vertical. Eventually, decreasing these variables can lead to a stress
increase in the vertical tail, which can cause the HSP to move to this structure. This behavior can be
observed in the σmax distributions, for the lower values of t vertical.
Considering all the constrains, the values selected are t vertical = 0.002m and size vertical =
0.12m. These values keep the HSP located at the wing skin, for n = 4.4, but moves it to the vertical tail,
near the connection with the boom, for the negative load factor.

72
Var 13: t out

To finish the analysis of the wing ribs, the t out variable is the one that follows. The results can be seen
in Figure 5.34.

Figure 5.34: BSGM - variation of constrains with t out variable.

This variable has little to none effect to the constrains. The only behavior worth noting is the increase
in σmax when decreasing t out. Reducing t out leads to a loss of weight and rigidity on the wing.
Since the aerodynamic loads, which are much higher than the weight, remain the same, this causes an
increase in the bending moment and amplitude of the wing. The wing skin strains even more which,
consequently, leads to higher stresses at the wing skin. Since the HSP was already located at the wing
skin, decreasing t out almost immediately provokes the increase of the σmax .
Due to this, and since the wing skin is almost at the maximum allowed stress, t out can only be
reduced to 0.045m. This value keeps the HSP at the wing skin, for n = 4.4, and at the vertical tail, for
n = −1.8.

Var 14: t ribs tail

Finally, the last variable in study is t ribs tail, responsible for the geometry of the tail ribs - Figure 5.35.
Due to its reduced weight, when compared with the rest of the components, t ribs tail has almost
no influence in the constrains. The results obtained for the SGM, in Sub-subsection 5.4.2, are similar,
so there is no need to a further analysis of this variable.

5.5.3 Final Results

The final values of all the variables, for the BSGM, are presented in Table 5.14. Their respective con-
strains values and weight, for both flight conditions, are presented in Table 5.15. Equally as for the
results of the BGM and SGM, these results were obtained after two iterations of the Structural Improve-
ment Model.
Looking at the final results, the SIM was able to reduce the structural weight from 5675.11Kg to
2190.46Kg, corresponding to a decrease of 61.40%. This value is quite near the goal weight of 1500Kg,

73
Figure 5.35: BSGM - variation of constrains with t ribs tail variable.

Table 5.14: BSGM - Final values for the variables.


i vari I0 I1 I2 Final Variation
1 t wingshell [m] 0.0040 0.0012 0.0012 -70 %
2 t wingboom [m] 0.0050 0.0030 0.0020 -60 %
3 chord percentage [-] 0.0800 0.0900 0.1160 45 %
4 t skin tail [m] 0.0070 0.0021 0.0018 -74 %
5 t reinforcement [m] 0.0100 0.0030 0.0020 -80 %
6 t rearboom [m] 0.0500 0.0100 0.0040 -92 %
7 size rearboom [m] 0.2500 0.2500 0.3250 30 %
8 t spars tail [m] 0.0050 0.0010 0.0004 -92 %
9 size spars tail [m] 0.1000 0.1500 0.2250 125 %
10 t in [m] 0.1000 0.0400 0.0390 -61 %
11 t vertical [m] 0.0060 0.0020 0.0012 -80 %
12 size vertical [m] 0.1000 0.1200 0.1500 50 %
13 t out [m] 0.0500 0.0450 0.0440 -12 %
14 t ribs tail [m] 0.0100 0.0020 0.0020 -80 %

Table 5.15: BSGM - Final results.


n Constrain I0 I1 I2 Final Variation
δwingtip [m] -0.053 -0.063 -0.060 13 %
-1.8 σmax [MPa] 125.19 159.92 182.50 46 %
δtail [m] -0.227 -0.434 -0.418 84 %
δwingtip [m] 0.058 0.079 0.076 31 %
4.4 σmax [MPa] 159.82 183.68 180.36 13 %
δtail [m] 0.061 0.305 0.325 432 %
W [Kg] 5675.11 2504.33 2190.46 -61 %

being this results discussed and compared with the previous ones in the final Section 5.6. However, as
expected, the weight obtained falls between the weight obtained for the BGM and the SGM, since this
model was built in order to use the pros from both and overcome most of the cons of each one of them.
This model should produce also the most accurate results, since it allows to represent correctly all the
aircraft structural components, with the right type of FE.

74
Using these final values for the geometric variables, it is then possible to analyze the stress and
vertical displacements distributions obtained for each of the load factors in study. Figure 5.36 represents
these distributions for n = −1.8, while Figure 5.37 represents them for n = 4.4.

(a) Stress Distribution (b) Displacement Distribution

Figure 5.36: BSGM - Stress and Displacement Distribution, for n = −1.8, using the final values of the
variables.

(a) Stress Distribution (b) Displacement Distribution

Figure 5.37: BSGM - Stress and Displacement Distribution, for n = 4.4, using the final values of the
variables.

Looking into the final distributions obtained, the higher stress for both flight conditions is quite sim-
ilar, being 182.50M P a for n = −1.8 and 180.36M P a for n = 4.4, indicating that the structure is well
dimensioned for both load factors.
For the negative load factor, the areas that face higher stresses are the connections between the
wing and the boom, the wing reinforcement, the connection between the boom and the vertical tail and,
finally, the connection between the vertical tail and the horizontal tail, where the HSP is located. For
n = 4.4, the areas are the same, with the addition of the wing spar’s root and the wing skin near it.
Concerning the displacement distributions, the results are clearly the expected. For n = −1.8, all the
displacements are negative, while for n = 4.4, all the displacements are positive. For both load factors,
δtail is higher, in module, than δwingtip .

75
5.6 Final Remarks

To conclude Chapter 5, the results obtained after applying the SIM to each Geometric Model are com-
pared and discussed. To a better visualization, they can be seen in Table 5.16.

Table 5.16: Comparison of the results obtained for the 3 Geometric Models
n Constrain BGM SGM BSGM
δwingtip [m] -0.109 -0.123 -0.060
-1.8 σmax [MPa] 180.82 183.90 182.50
δtail [m] -0.441 -0.460 -0.418
δwingtip [m] 0.164 0.101 0.076
4.4 σmax [MPa] 176.74 169.53 180.36
δtail [m] 0.361 0.213 0.325
W [Kg] 811.85 8890.79 2190.46

The Beam Geometric Model, and as defined, is clearly the model that has the lowest weight, being
way below the design goal weight of 1500Kg. This is completely expected, since this is not capable of
representing several components of the structure, such as the wing and tail skin and the wing reinforce-
ment. Also, the ribs are represented with an approximation, making it quite inaccurate.
Regarding the Shell Geometric Model, the weight obtained is clearly oversized. This model allows
a better structural representation of the aircraft in study, but it can be considered the worst of the 3
Geometric Models developed. To better explain it, this model uses only Shell FE, even for representing
components that clearly behave as a beam. Examples of these components are the booms, the spars
and the vertical tail. This fact leads to a mandatory increase of the FE’s thickness of these components,
to compensate the bad capability of these elements to behave as a beam. This way, it is not possible to
lower more the weight without compromising the constrains defined, specially δtail .
Finally, the Beam-Shell Geometric Model was developed in order to try to take advantage of both
models advantages - it allows an accurate representation of the entire structure’s geometry and it uses
Beam FEs for the components that behave as beams and it Shell FEs for the components that behave
as shells. Running the SIM on top of this Geometric Model produces a weight of 2190.46kg, which falls
in between the values obtained for the other two Geometric Models, as expected. Among the 3, this is
definitely the model that should produce the more accurate results, due to the reasons defined.
However, even using the BSGM, the weight obtained is yet a bit higher than the goal weight of
1500kg. This is due to the fact that, for simplicity, only 14 variables were considered for the definition
of the Geometric Model, meaning that only 14 geometric parameters could be changed. By defining
other variables, a lower weight could have been achieved. As an example, considering a higher wing
skin thickness near the wing root and a thinner thickness near the tip, using a t wingshell root and a
t wingshell tip variables, could lead to a decrease of the weight (i.e. tapering the skin thickness). Also,
using the same approach on the booms could have also been possible. In short, the increase of the
model complexity would benefit of an increase of the variables number, in order to obtain a reduced
structural weight. However, this increase felt of the scope of the project, but it is definitely something that

76
could be tested in future work.
One other parameter that is worth evaluating is the time per simulation of each Geometric Model, to
understand the efficiency of each one. The results collected are presented in Table 5.17.

Table 5.17: Time spent per simulation.


Model Static [s] Mass [s]
Beam 16.21 6.61
Shell 14.59 6.42
Beam-Shell 33.62 12.76

The Beam and the Shell Geometric Models have quite low simulation times, either for the Static and
Mass simulations. The higher values of the Static simulation are mostly due to the need for applying
the aerodynamic forces on the wing and tail components, which is a process that really slows down the
simulation.
The Beam-Shell Geometric Model has much higher values when compared with the other two. This is
essentially due to the complexity of the model, which needs to couple 3 different types of FE (BEAM189,
SHELL281 and MASS21) and has also a huge number of FEs: 32392 against 573 for the BGM and
3444 for the SGM. This adds an extra computational effort when defining and resolving the matrices that
define the FEM problem, leading to an higher simulation time.
Regarding the SIM itself, it presents really good results. It was capable of reducing the initial weights
of Beam, Shell and Beam-Shell Geometric Models in 76.46%, 48.12% and 61.40% indicating that it is a
reliable model. Also, the model can be applied to any type of Geometric Model and with any type of
constrains, making it appealing for future studies.
To wrap the present Chapter, Table 5.18 is presented, which contains a listing of the advantages and
disadvantages of Geometric Models developed.

77
Table 5.18: Advantages and Disadvantages of each Geometric Model.
Modes Advantages Disadvantages
• Does not allow a correct repre-
sentation of the Geometry, since
• Simple to develop some components can not be
represented
• Computationally efficient
• Weight obtained is underesti-
BGM • Less Variables mated
• Allows a first simple and quick vi- • Model might produce inaccurate
sualization of the forces acting on results
the structure
• Wrong usage of Beam FE in the
ribs
• Wrong usage of Shell FEs in sev-
eral components, such as spars
• Simple to develop
and booms
• Computationally efficient
• Weight obtained is clearly over-
SGM • Less Variables estimated, due to the wrong us-
age of Shell FEs in all the com-
• It allows a better representation ponents
of the geometry
• Model might produce inaccurate
results
• Accurate representation of the
Geometry and all the compo-
nents • More complex model, leading to
BSGM more variables
• More accurate results
• Higher time spent per simulation
• Correct usage of different FEs
according to the type of compo-
nent

78
Chapter 6

Conclusions

In this Chapter the main conclusions are presented.

6.1 Achievements

After all the work developed, it is possible to conclude that the main objectives of this work were achieved.
Three different Geometric Models were developed, using ANSYS APDL, each one of them with a dif-
ferent degree of complexity, that allow the study and analysis of the Flexcraft’s structure. These models
can be used to determine the behavior of the structure, due to the effects of inertial and aerodynamic
loading. They were also prepared for parametrization processes, since several geometric values are
defined as variables and must be introduced beforehand.
Due to the above, a SIM was also developed, which can run on top of each Geometric Model. This
is an algorithm that changes the input variables, according to the values defined in the start. A FEA is
made for each different variation, being the needed results exported for analysis.
With this algorithm, is then possible to obtain an approximation of an optimized structure, that with-
stands the aerodynamic and inertial loads, without overcoming some geometric and structural constrains
that are defined during the project phase and must be passed in the beginning of the SIM.
The SIM can be implemented in any other Geometric Model built using ANSYS APDL, as long as
the Geometric Model is prepared for parametrizations. This fact shows the great versatility of this SIM,
which is one of its biggest assets, along with the ease of use.

6.2 Future Work

In order to continue the work developed, further improvements are suggested:

• Since aluminium was the only material used in this study, it would be very interesting to consider
the behavior of the wing-tail structure with a composite material on the skin parts, as it is used in
the majority of aircraft built nowadays. Considering its properties, it would surely help in reducing
the weight without overcoming the constrains or compromising the structural integrity of the aircraft.

79
However, a further study is required, since the composite material is not an isotropic material and
it does not behave the same as aluminium, adding an extra level of complexity to the structural
analysis. Also, the layers of the composite would have to be parametrized, instead of only varying
the total thickness [54].

• Also, it would be extremely interesting to compare and validate the results obtained by the models
with a Flexcraft prototype. This would allow to understand which model is more accurate, in order
to use it in future developments. Another option would be to apply the model to an already existing
aircraft of aircraft prototype, and compare the computational and experimental results.

• Performing an aeroelastic analysis would also be important, in order to understand if the constrains
applied are not underestimated. For example, increasing the boundaries of δtail and δwingtip would
certainly allow the rigidity of several components of the wing-tail structure of the Flexcraft, leading
to a reduction of the global estimated weight.

• Also, the Geometric Models can also be improved, by defining new variables to parametrize and
change the way some variables are defined. As an example, it would be interesting to vary the
wing skin thickness along the span or vary the cross section of the booms, making it more robust in
areas of larger stress intensity, and less robust in the other areas. This could also allow to reduce
the estimated global weight.

80
Bibliography

[1] E. Commission. White paper on transport — roadmap to a single european transport area —
towards a competitive and resource-efficient transport system, 2011. ISBN 978-92-79-18270-9.

[2] U. Elevate. Fast-forwarding to a future of on-demand urban air transportation, 2016.

[3] M. D. Maisel, D. J. Giulianetti, and D. C. Dugan. The history of the xv-15 tilt rotor research aircraft:
From concept to flight. Monographs in Aerospace History 17, National Aeronautics and Space
Administration, Office of Policy and Plans, NASA History Division, Washington D.C, 2000. NASA
SP-2000-4517.

[4] L. K. Loftin. Quest for Performance - The Evolution of Modern Aircraft. NASA SP-468. NASA
Scientific and Technical Information Branch, Washington, D.C., 1985.

[5] Wright brothers first flight. https://www.nasa.gov/multimedia/imagegallery/image_feature_


976.html, 2008. Accessed: 15/12/2017, 17h45.

[6] F. L. Faurote. The Aircraft Year Book For 1919. Manufacturers Aircraft Association, 501 Fifth
Avenue, NYC, 1919.

[7] M. J. F. Sunderman. Early Air Pioneers 1862-1935. The Watts Aerospace Library. Franklin Watts,
Inc., 575 Lexington Avenue, New York 22, 1961.

[8] Quest for performance: The evolution of modern aircraft. https://history.nasa.gov/SP-468/


ch4-4.htm. Accessed: 15/12/2017, 18h10.

[9] H. M. Holden. The boeing 247 - the first modern commercial airliner. World Airnews, pages 28–30,
April 2016.

[10] D. J. Ingells. The Plane That Changed the World: A Biography of the DC-3. Aero Publishers, 1st
edition edition, June 1966. ISBN-10: 0816875006.

[11] S. Fahlström and R. Pihl-Roos. Design and construction of a simple turbojet engine. Msc thesis,
Uppsala Universitet, Box 536 751-21 Uppsala, September 2016.

[12] Heinkel he 178. http://www.aviation-history.com/heinkel/he178.html, 2012. Accessed:


15/12/2017, 18h30.

[13] Boeing 707 revelations. Flight Magazine, P 115, January 1956.

81
[14] R. A. Rivers, E. B. Jackson, C. G. Fullerton, T. H. Cox, and N. H. Princen. A qualitative piloted
evaluation of the tupolev tu-144 supersonic transport. Technical Report TM-2000-209850, National
Aeronautics and Space Administration, Langley Research Center Hampton, Virginia 23681-2199,
February 2000.

[15] M. W. Bowman. Boeing 747: A History. Pen and Sword Aviation, 47 Church Stress, Barnsley,
South Yorkshire, S70 2AS, 1st edition edition, 2014.

[16] G. M. Simons. The Airbus A380: A History. Pen and Sword Aviation, November 2014. ISBN-10:
1783030410.

[17] The evolution of airplanes. https://publishing.aip.org/publishing/journal-highlights/


evolution-airplanes, 2014. Accessed: 15/12/2017, 18h50.

[18] E. A. S. Agency. Annual safety review 2018. Report, European Aviation Safety Agency, 2018, Post-
fach 1012 53, 50452 Cologne, Germany, 2018. Safety Intelligence and Performance Department.

[19] Focke-wulf fw 61. https://www.militaryfactory.com/aircraft/detail.asp?aircraft_id=922,


2018. Accessed: October 2018.

[20] S. B. Anderson. Historical overview of v/stol aircraft technology. Technical Memorandum A-8511,
NASA, Washington D. C. 20546, March 1981.

[21] D. Neilson, T. Campbell, and B. Alllred. Model-based inquiry in physics: A buoyant force module.
77:38–43, 11 2010.

[22] F. I. Petrescu and R. V. Petrescu. The Aviation History. Books on Demand GmbH, Norderstedt,
USA, 2012. ISBN 978-3-8482-3077-8.

[23] Hindenburg crash: The end of airship travel. https://www.livescience.com/


58959-hindenburg-crash.html, 2017. Accessed: 15/12/2017, 19h00.

[24] J. Fay. The Helicopter : History, Piloting, and How it Flies. Dvd and Chrlrs PLC, 4th edition edition.
ISBN-10: 0715389408.

[25] W. Stalewski. Aerodynamic design of modern gyroplane main rotors. Transactions of the Institute
of Aviation 1(242), Institute of Aviation, al. krakowska 110/114, 02-256 Warsaw, Poland, 2016. DOI:
10.5604/05096669.1202204.

[26] Boeing v-22 osprey. http://www.boeing.com/defense/v-22-osprey/, 2019.

[27] P. PradeepPeng and P. Wei. Energy optimal speed profile for arrival of tandem tilt-wing evtol aircraft
with rta constraint. Septemmber 2018. DOI: 10.13140/RG.2.2.19438.87367.

[28] J. R. Chambers. Modeling flight - the role of dynamically scaled free-flight models in support of
nasa’s aerospace programs. Nasa sp 2009-575, NASA, 300 E Street SW, Washington, DC 20546,
2009.

82
[29] A. Intwala and Y. Parikh. A review on vertical take off and landing (vtol) vehicles. International
Journal of Innovative Research in Advanced Engineering (IJIRAE), 2(2):186–191, February 2015.
ISSN: 2349-2163.

[30] L. O. Nordeen. Harrier II: Validating V/STOL. Naval Institute Press, February 2007. ISBN-13:
978-1591145363.

[31] H. Magazines, editor. Popular Mechanics, volume 178, January 2001. Hearst Magazines. ISSN
0032-4558.

[32] K. Desmond. Electric Airplanes and Drones: A History. McFarland and Company, Inc., 2018.
eISBN: 978-1-4766-3341-1.

[33] EASA. Easy access rules for normal, utility, aerobatic and commuter category aeroplanes (cs-23)
(initial issue). Technical report, EASA, June 2018.

[34] F. Afonso, H. Policarpo, F. Lau, and A. Suleman. Desenvolvimento - Projeto de VRP, FLEXCRAFT
E.3.1. Report, Instituto Superior Técnico, Lisbon, Portugal, 2018.

[35] T. H. G. Megson. Aircraft Structures for Engineering Students. 4th edition edition, 2007. ISBN-10:
0-750-667397.

[36] J. W. Rustenburg, D. A. Skinn, and D. O.Tipps. Statistical loads data for the airbus a-320 aircraft
in commercial operations. Final Report UDR-TR-2001-00080, University of Dayton Research Insti-
tute, Structural Integrity Division, 300 College Park, Dayton, OH 45469-0120, April 2002.

[37] Part23 – Small Airplane Certification Process Study. Federal Aviation Administration, Washington
DC, USA, ok-09-3468 edition, July 2009.

[38] T. C. Corke. Design of Aircraft. Pearson Education, Upper Saddle River, N.J.: Prentice Hall,
November 2002. ISBN 0-13-089234-3.

[39] Theory Reference for the Mechanical APDL and Mechanical Applications. ANSYS, Inc, Southpoite
275 Technology Drive Canonsburg, PA 15317, peter kohnke edition, April 2009. Release 12.0.

[40] A. Kaveh, K. Koohestanib, and N. Taghizadiehb. Force method for finite element models with
indeterminate support conditions. Asian Journal Of Civil Engineering, 8(4):389–403, January 2007.

[41] E. Hearn. Mechanics of Materials 2 - An Introduction to the Mechanics of Elastic and Plastic
Deformation of Solids and Structural Materials. Butterworth-Heinemann, 3 edition, 1997. ISBN:
978-0-7506-3266-9.

[42] C. C. Ike. Timoshenko beam theory for the flexural analysis of moderately thick beams – variational
formulation, and closed form solution. -Italian Journal of Engineering Science, 63(1):34–45, March
2019.

83
[43] V. Debnath and B. Debnath. Deflection and stress analyses of a beam on different elements using
ansys apdl. International Journal Of Mechanical Engineering AND Technology, 5(6):70–79, June
2014. ISSN 0976 – 6359.

[44] G. Liu and S. Quek. The Finite Element Method - A Practical Course. Butterworth-Heinemann, 2nd
edition, 2014. ISBN 978-0-08-098356-1.

[45] T. Nelson and E. Wang. Reliable fe-modeling with ansys. Report, CADFEM GmbH, Munich, Ger-
many, 2004.

[46] F. Afonso, J. Vale, F. Lau, and A. Suleman. Performance based multidisciplinary design optimization
of morphing aircraft. Aerospace Science and Tecnology, 67:1–12, August 2017.

[47] J. Katz and A. Plotkin. Low-Speed Aerodynamics. Cambridge University Press, 2nd edition edition,
2001. ISBN 978-0-521-66219-2.

[48] A. Kale, E. Acart, R. T. Haftka, and W. J. Stroud, editors. Why are Airplanes so Safe Struc-
turally? Effect of Various Safety Measures on Structural Safety of Aircraft, April 2004. 45th
AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and Materials Conference.

[49] J. J. M. da Silva. Design and optimization of a wing structure for a UAS Class I 145 kg. Master’s
thesis, Técnico Lisboa, July 2017.

[50] T. M. Salih. Nature of the materials for modern airplane parts. 12 2010.

[51] Y. Kaushik. A review on use of aluminium alloys in aircraft components. Journal on Material
Science, 3(3):33–38, December 2015.

[52] J. G. Kaufman and E. L. Rooy. Aluminum Alloy Castings: Properties, Processes, and Applications.
ASM International, December 2004.

[53] K.-J. Bathe. Finite Element Procedures. K.J. Bathe, Watertown, MA, 2nd edition edition, 2016.
ISBN 978-0-9790049-5-7.

[54] M. S. Muttulingam. Composites modeling with ansys - modeling layered composites the simple
way. ANSYS, Inc, October 2011.

84

You might also like