Você está na página 1de 202

Integration of Aircraft Systems into

Conceptual Design Synthesis

Von der Fakultät für Maschinenwesen der Rheinisch-Westfälischen


Technischen Hochschule Aachen zur Erlangung des akademischen Grades
eines Doktors der Ingenieurwissenschaften genehmigte Dissertation

vorgelegt von

Tim Lammering

Berichter: Universitätsprofessor Dr.-Ing. Eike Stumpf


außerplanmäßiger Professor Dipl.-Ing. Rolf Henke
Universitätsprofessor Dr.-Ing. Dieter Moormann

Tag der mündlichen Prüfung: 09.05.2014

Diese Dissertation ist auf den Internetseiten der Hochschulbibliothek


online verfügbar.
Abstract

In the scope of this thesis, a methodology for modeling of entire system architectures of com-
mercial transport aircraft is presented. The author’s approach facilitates technology assessment
of innovative system technology on aircraft-level in conceptual and early preliminary aircraft
design. It closes the existing gap of sound systems modeling in early aircraft design synthesis.

The proposed methodology uses a strictly function-driven approach for sizing of the single air-
craft systems according to the dedicated aircraft-level functions. System weight and maximum
as well as mission dependent power off-takes are estimated. The principle of energy balance is
used to derive the power demand of the single systems. A modular and flexible approach has
been chosen that allows for fast variation of the system architecture. Thus, investigation of
large design spaces is possible. For integration and assessment of innovative systems technology,
a probabilistic framework is introduced that incorporates remaining uncertainties.

The proposed system model has been integrated into the Institute’s of Aeronautics and Astro-
nautics (ILR) Multidisciplinary Integrated Conceptual Aircraft Design and Optimization (MI-
CADO) environment. Thus, repercussions of system integration, e.g. on aircraft efficiency or
performance, can be captured in a highly integrated environment. The proposed systems model
is a valuable tool for sound assessment of innovative system technology on overall aircraft-level
already in conceptual aircraft design.

The proposed systems model has been validated against two commercial transport aircraft.
Additionally, comprehensive sensitivity studies show dependencies and identify the main design
drivers of the single systems. Two case studies show the added-value of integration of the
proposed systems model into early aircraft design synthesis. The first case study concentrates
on evaluation of the impact of engine power off-takes on aircraft performance and efficiency.
The second design study focuses on integration of innovative technology into the overall system
architecture and on application of the proposed probabilistic framework for considering the
remaining uncertainties in design and technology assessment.

V
Kurzfassung

Die vorliegende Arbeit beschreibt eine Methode zur Modellierung von Flugzeugsystemarchitek-
turen im Rahmen des Flugzeugvorentwurfs. Die Anwendung der Methode erlaubt bereits
frühzeitig innovative Systemtechnologie auf Flugzeugebene zu bewerten und schließt damit
eine existierende Lücke im heutigen Vorgehen.

Die beschriebene Methode nutzt einen streng funktionsbasierten Ansatz um die einzelnen
Flugzeugsysteme in Abhängigkeit der dedizierten Flugzeugfunktionen auszulegen. Gewicht
sowie maximale und missionsabhängige Leistungsaufnahmen werden für die einzelnen Systeme
bestimmt. Hierbei spielt das Prinzip der Energiebilanzierung eine entscheidende Rolle. Es
wurde ein variabler und modularer Ansatz gewählt, der die Untersuchung großer Parame-
terräume in Bezug auf die gewählte Systemarchitektur erlaubt. Weiterhin wurde ein proba-
bilistischer Ansatz eingeführt, der speziell die Einbindung innovativer Systemtechnologie er-
laubt und bestehende Unsicherheiten mit in die Technologiebewertung einfließen lässt.

Das entwickelte Systemmodell wurde direkt in die Flugzeugvorentwurfsumgebung MICADO des


Instituts für Luft- und Raumfahrtsysteme (ILR) integriert. Wechselwirkungen und Schneeball-
effekte der Systemintegration können so bereits früh im Flugzeugvorentwurf auf Flugzeugebene
abgebildet werden. Damit dient das entstandene Systemmodell zur umfassenden und ganz-
heitlichen Technologiebewertung.

Das Systemmodell wurde mit Hilfe von zwei Verkehrsflugzeugen validiert. Außerdem zeigen
umfassende Sensitivitätsstudien die Haupteinflüsse auf die Systemmodellierung sowie auf die
resultierende Systemarchitektur auf. Ferner wird der Mehrwert der Integration des Modells in
den Flugzeugvorentwurf mit Hilfe von zwei Anwendungsbeispiele aufgezeigt. Im Rahmen des
ersten Anwendungsbeispiels wird der Einfluss der Systemleistungsentnahme auf die Flugzeug-
effizienz und Flugleistung diskutiert. Das zweite Anwendungsbeispiel beschäftigt sich mit der
Integration innovativer Systemtechnologie in das Flugzeug sowie mit der Anwendung des prob-
abilistischen Modellansatzes. Der Nutzen einer Bewertung unter Berücksichtigung der beste-
henden Unsicherheiten wird dabei aufgezeigt.

VII
Acknowledgment

During my time at the ILR, I had the pleasure to work under Professor Rolf Henke. It is his
achievement that the ILR’s main expertise is now focused on conceptual aircraft design, systems
integration and optimization, and technology assessment. Without him, the presented research
would not have been possible. I thank Prof. Henke for his constant support and guidance also
in his current position as DLR executive board member. I thank Professor Dr. Eike Stumpf
for agreeing to supervise my thesis. In my last year at the ILR, I had the pleasure to work
under Prof. Stumpf. He continues to work towards the main research areas that were set by
Prof. Henke and successfully brings in his personal aspects. I am very grateful for many fruitful
discussions and his efforts in supervising this thesis. I also thank Professor Dr. Moormann and
Professor Dr. Jeschke for participating in the thesis committee.

I like to mention the ILR’s former academic counselor, Dr. Neuwerth, as well as the current one,
Dr. Hoernschemeyer. I thank both for their support and countless purposeful advices. Many
thanks go to my great colleagues at university. It has been a fantastic time and a pleasure
to work in such a fruitful, professionally challenging, and inter-disciplinary environment. I
personally thank Eckhard Anton, Kristof Risse, and Katharina Franz, which share the same
interest in aircraft design. I am grateful for their close collaboration. Without it, the presented
research would not have been possible; especially not its integration into the overall aircraft
design framework. I am glad for the fruitful culture of critically reviewing each other’s work
that we had established among each other. I am convinced that it has increased the quality of
all of our work significantly. Further, I thank Kristof Risse in particular for spending his leisure
time commenting my thesis. Further, research at university would not be possible without the
highly motivated and dedicated students, who assist the researchers at the institutes.

I am very grateful for the time that I have spent in Aachen and for the many close personal
friends that I have found during that time. Their friendship and support is priceless to me.
Thank you all for your warmth and support and for taking my mind of work through the time
that we spent together.

I thank my parents for bringing me up in an environment where I always felt beloved, and
where the pursuit of personal achievement and self-dependency was always encouraged. Thank

IX
you for your utmost support, first in my education and studies, and now in my professional
career. I apologize that my passion for aerospace has taken their grandchild as well as my wife
and me farer from their home than they like.

My warmest and biggest thanks goes to my wife Gesine and our newly born daughter. I thank
you for the unlimited support and love that you have granted me during our time together. I
thank you not only for allowing me to finish my thesis in our leisure time, but for encouraging
and driven by me to do so. Without you, I would have never been able to finish it.

My research at the Institute of Institute of Aeronautics and Astronautics (ILR) of RWTH


Aachen University has been financed by the European Commission as part of the 7th Framework
Program collaborative research projects SADE.
Contents

Abstract V

Kurzfassung VII

Acknowledgment IX

Nomenclature XXI

1 Introduction 1
1.1 Aircraft Systems Technology and Innovations . . . . . . . . . . . . . . . . . . . 1
1.2 Repercussions of Aircraft Systems Integration . . . . . . . . . . . . . . . . . . . 6
1.2.1 Direct Repercussions of Engine Off-Takes . . . . . . . . . . . . . . . . . . 6
1.2.2 Snowball Effects in Overall Aircraft Design Synthesis . . . . . . . . . . . 8
1.3 Motivation, Scope and Field of Application . . . . . . . . . . . . . . . . . . . . . 9
1.4 General Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

2 Related Work in Systems Modeling 13


2.1 Statistical Approaches for Estimating Systems Mass . . . . . . . . . . . . . . . . 13
2.1.1 Weight Fractions and Regression Analysis . . . . . . . . . . . . . . . . . 14
2.1.2 Trimmed Statistics in Industrial Application . . . . . . . . . . . . . . . . 16
2.2 Medium and High Fidelity Tools . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.3 Modeling of Entire Systems Architectures . . . . . . . . . . . . . . . . . . . . . 17
2.4 Probabilistic Approaches in Systems Design . . . . . . . . . . . . . . . . . . . . 19
2.5 The Need for a New Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . 20

3 Aircraft Systems Model 23


3.1 Proposed Methodology for System Modeling . . . . . . . . . . . . . . . . . . . . 23
3.1.1 Aircraft-Level Functions and Systems Dependencies . . . . . . . . . . . . 23
3.1.2 Generic Systems Description . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.1.3 Energy Balance within the Systems Architecture . . . . . . . . . . . . . . 31
3.1.4 Model Architecture Down to System-Level and Below . . . . . . . . . . . 33

XI
XII Contents

3.2 Proposed Systems Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34


3.2.1 Main Design Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.2.2 Environmental Control System – ATA-21 . . . . . . . . . . . . . . . . . . 37
3.2.3 Ice and Rain Protection – ATA-30 . . . . . . . . . . . . . . . . . . . . . . 43
3.2.4 Primary and Secondary Flight Control System – ATA-27 . . . . . . . . . 51
3.3 Implementation of the Proposed Systems Model . . . . . . . . . . . . . . . . . . 60
3.3.1 MICADO Program Module Structure . . . . . . . . . . . . . . . . . . . . 60
3.3.2 Modularity and Flexibility of the Program Structure . . . . . . . . . . . 61
3.4 Probabilistic Modeling of Innovative Systems . . . . . . . . . . . . . . . . . . . . 63

4 Integration into Conceptual Aircraft Design Synthesis 69


4.1 Conceptual Aircraft Design Synthesis . . . . . . . . . . . . . . . . . . . . . . . . 69
4.1.1 Design and Optimization Framework . . . . . . . . . . . . . . . . . . . . 69
4.1.2 Underlying Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
4.2 Impact of the Systems Model on Design Synthesis . . . . . . . . . . . . . . . . . 75
4.2.1 Systems Mass in Design Synthesis . . . . . . . . . . . . . . . . . . . . . . 75
4.2.2 Power Off-Takes in Design Synthesis . . . . . . . . . . . . . . . . . . . . 75
4.3 Conventional Reference Aircraft . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
4.3.1 Reference Aircraft: A320-200 . . . . . . . . . . . . . . . . . . . . . . . . 78
4.3.2 Reference Aircraft: Boeing 777-200ER . . . . . . . . . . . . . . . . . . . 80
4.3.3 Reference Aircraft: Single-Aisle SA185, 185-PAX . . . . . . . . . . . . . 81

5 Verification and Sensitivities of the Proposed Systems Model 85


5.1 Application to the Reference Aircraft . . . . . . . . . . . . . . . . . . . . . . . . 86
5.1.1 System Mass Estimates . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
5.1.2 Design and Mission Power Estimates . . . . . . . . . . . . . . . . . . . . 89
5.2 Sensitivities of the Proposed Systems Model . . . . . . . . . . . . . . . . . . . . 94
5.2.1 Sensitivities Towards Model Specific Input Parameters . . . . . . . . . . 96
5.2.2 Sensitivities Towards Aircraft Design Parameters . . . . . . . . . . . . . 99
5.2.3 Sensitivities Towards Top-Level Aircraft Requirements . . . . . . . . . . 102
5.3 Sensitivities of Overall Design Synthesis Towards the Proposed Systems Model . 104
5.4 Expected Level of Accuracy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
5.4.1 Expected Accuracy of Overall System Mass . . . . . . . . . . . . . . . . 107
5.4.2 Expected Accuracy of Bleed Air Demand . . . . . . . . . . . . . . . . . . 109
5.4.3 Expected Accuracy of Shaft Power Off-Takes . . . . . . . . . . . . . . . . 109
5.4.4 Limitations of the Proposed Systems Model . . . . . . . . . . . . . . . . 110
Contents XIII

6 Case Studies 113


6.1 Impact of Engine Power Off-Takes on Aircraft Performance and Efficiency . . . 113
6.1.1 Impact on Take-off Performance . . . . . . . . . . . . . . . . . . . . . . . 114
6.1.2 Impact on Fuel Efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . 116
6.1.3 Impact on Overall Design Synthesis and Optimization . . . . . . . . . . . 117
6.2 Innovative Morphing Leading Edge . . . . . . . . . . . . . . . . . . . . . . . . . 120
6.2.1 Laminar Aircraft Design . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
6.2.2 Innovative Morphing Leading Edge and Systems Architecture . . . . . . 122
6.2.3 Initially Derived Systems Characteristics . . . . . . . . . . . . . . . . . . 123
6.2.4 Technology Assessment with Probabilistic Approach . . . . . . . . . . . . 125

7 Conclusion 135

Bibliography 137

A Systems Definition 157

B Deviations in Power Off-Take Corrections 161

C Top-Level Aircraft Requirements for SA185 Aircraft 163

D Sensitivities of 777 Design 165


D.1 Sensitivities towards Model Specific Input Parameters . . . . . . . . . . . . . . . 165
D.2 Sensitivities towards Aircraft Design Parameters . . . . . . . . . . . . . . . . . . 166
D.3 Sensitivities towards Top-Level Aircraft Design Parameters . . . . . . . . . . . . 167
D.4 Sensitivities of Overall Design Synthesis . . . . . . . . . . . . . . . . . . . . . . 168

Curriculum Vitae of the Author 169

List of Publications by the Author 171


List of Figures

1.1 A generic aircraft system. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2


1.2 Classification of the different aircraft systems of civil transport aircraft [1, ch. 5]. 2
1.3 Schematics of conventional power systems architecture. . . . . . . . . . . . . . . 3
1.4 The technology S-curve [2]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.5 Schematics of all-electric power systems architecture. . . . . . . . . . . . . . . . 6
1.6 Influences of aircraft system integration on overall aircraft-level. . . . . . . . . . 7
1.7 Influence of power off-takes on SFC (35,000 ft and Mach 0.80) [3]. . . . . . . . . 7
1.8 Iterative design methodology. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.9 Knowledge during design. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

2.1 Regression analysis for estimating total systems mass. . . . . . . . . . . . . . . . 14


2.2 Allocation of available models to the design phases. . . . . . . . . . . . . . . . . 21

3.1 Function based systems domains of conventional systems architectures. . . . . . 24


3.2 Hydraulic power requirements for take-off and climb of an Airbus A320. . . . . . 26
3.3 Sizing mode and operational mode of the proposed systems model. . . . . . . . . 26
3.4 Power requirements in dependency on system states [4]. . . . . . . . . . . . . . . 27
3.5 Installation locations of standard bleed air system and its consumers (Airbus A320 top-
view taken from reference [5]). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.6 Schematics of the generic systems module. . . . . . . . . . . . . . . . . . . . . . 30
3.7 Generic systems architecture. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.8 Energy balance and systems boundaries. . . . . . . . . . . . . . . . . . . . . . . 31
3.9 Model architecture down to system-level and below. . . . . . . . . . . . . . . . . 34
3.10 Mass balance of cabin air flow (100 % cabin air inflow) [6]. . . . . . . . . . . . . 39
3.11 Required ACP power in dependency on ambient temperature and cabin utilization. 40
3.12 ACP cooling or heating in dependency on ambient temperature and utilization. 41
3.13 Required ECS bleed air flow for the design cases (A320 maximum bleed air flow [7]). 43
3.14 Simplified system schematics for leading edge anti-icing system. . . . . . . . . . 44
3.15 Structure and input parameters of the implemented wing anti-icing model [8]. . 46
3.16 Icing envelope for standard icing conditions [9, app. C]. . . . . . . . . . . . . . . 47
3.17 Required bleed air mass for variation of altitude. . . . . . . . . . . . . . . . . . . 48

XV
XVI List of Figures

3.18 Required bleed air for variation of TAS. . . . . . . . . . . . . . . . . . . . . . . . 50


3.19 Required bleed air for variation of droplet diameter. . . . . . . . . . . . . . . . . 50
3.20 Required bleed air for full vaporizing anti-icing of a typical single-aisle aircraft
in standard ISA conditions (dashed lines show iso bleed air mass flow). . . . . . 50
3.21 Required bleed air for maximum continuous icing and take-off icing conditions. . 51
3.22 Configuration of the conventional electronic flight control system (flap drive sys-
tem not shown, but analogue to slat transmission system). . . . . . . . . . . . . 53
3.23 Book-keeping approach for estimating FCS weight and center of gravity. . . . . . 56
3.24 Regression analysis for estimating power control unit weight (data source [10]). . 57
3.25 Derivation of FCS center of gravity from the single components. . . . . . . . . . 58
3.26 Weight break down of Airbus FCS [11] and estimates. . . . . . . . . . . . . . . . 59
3.27 A320 high lift component weights [10] and estimated weights. . . . . . . . . . . 60
3.28 Simplified UML class diagram of the implemented systems model. . . . . . . . . 61
3.29 System technology design space for flexible system architecture definition. . . . . 62
3.30 Definition of design space with the proposed framework. . . . . . . . . . . . . . 65
3.31 Continuous probability distributions. . . . . . . . . . . . . . . . . . . . . . . . . 65
3.32 Design space and parameter sweep. . . . . . . . . . . . . . . . . . . . . . . . . . 66

4.1 Methodology of ILR MICADO environment. . . . . . . . . . . . . . . . . . . . . 70


4.2 Fuel burn optimized mission profile. . . . . . . . . . . . . . . . . . . . . . . . . . 74
4.3 Change in fuel flow for CF6-80C2 engine model in different flight conditions. . . 77
4.4 Deviations in power off-take correction between MICADO and GasTurb for the
CF6-80C2, CFM56-5B4P, and GE90-115B engines. . . . . . . . . . . . . . . . . 78
4.5 General arrangement of A320 design. . . . . . . . . . . . . . . . . . . . . . . . . 79
4.6 Payload-range diagram. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
4.7 Comparison between Airbus A320 and A320 MICADO design. . . . . . . . . . . 80
4.8 General arrangement of 777 design. . . . . . . . . . . . . . . . . . . . . . . . . . 81
4.9 Payload-range diagram. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
4.10 Comparison between Boeing 777-200ER and 777 MICADO design. . . . . . . . . 81
4.11 Optimization of wing loading and thrust-to-weight ratio for the SA185 design. . 82
4.12 Comparison between ILR MICADO and PrADO design of SA185. . . . . . . . . 83

5.1 Comparison of aircraft weight break down of A320 and 777 MICADO designs. . 87
5.2 Deviations of estimated systems mass. . . . . . . . . . . . . . . . . . . . . . . . 87
5.3 Comparison of system weights of A320 and 777 MICADO designs. . . . . . . . . 88
5.4 Comparison of the author’s estimates with Koeppen’s and A320 weight data [11]. 89
5.5 Total estimated design power and bleed air (design engine off-takes). . . . . . . 90
5.6 Estimated design power and bleed air for the different consumer systems. . . . . 91
5.7 Simulated design mission of A320. . . . . . . . . . . . . . . . . . . . . . . . . . . 92
List of Figures XVII

5.8 Engine power off-takes of A320 design mission. . . . . . . . . . . . . . . . . . . . 92


5.9 Characteristic mission points of A320 mission profile. . . . . . . . . . . . . . . . 93
5.10 Mission dependent power off-takes for single A320 systems. . . . . . . . . . . . . 94
5.11 A320 – Sensitivity towards bleed air temperature. . . . . . . . . . . . . . . . . . 97
5.12 A320 – Sensitivity towards required cabin temperature. . . . . . . . . . . . . . . 97
5.13 A320 – Sensitivity towards recirculation of cabin air. . . . . . . . . . . . . . . . 97
5.14 A320 – Sensitivity towards span position of wing anti-icing. . . . . . . . . . . . . 97
5.15 A320 – Sensitivity towards number of modular concept units. . . . . . . . . . . . 98
5.16 A320 – Sensitivity towards deflection rate of the PFCS. . . . . . . . . . . . . . . 98
5.17 A320 – Sensitivity towards maximum take-off weight. . . . . . . . . . . . . . . . 100
5.18 A320 – Sensitivity towards operating weight empty. . . . . . . . . . . . . . . . . 100
5.19 A320 – Sensitivity towards number of passengers. . . . . . . . . . . . . . . . . . 100
5.20 A320 – Sensitivity towards cabin and fuselage length. . . . . . . . . . . . . . . . 100
5.21 A320 – Sensitivity towards wing span. . . . . . . . . . . . . . . . . . . . . . . . 101
5.22 A320 – Sensitivity towards wing reference area. . . . . . . . . . . . . . . . . . . 101
5.23 A320 – Sensitivity towards design range. . . . . . . . . . . . . . . . . . . . . . . 103
5.24 777 – Sensitivity towards design range. . . . . . . . . . . . . . . . . . . . . . . . 103
5.25 A320 – Sensitivity towards design payload. . . . . . . . . . . . . . . . . . . . . . 103
5.26 A320 – Sensitivity towards design passenger number. . . . . . . . . . . . . . . . 103
5.27 A320 – Sensitivity towards total system mass. . . . . . . . . . . . . . . . . . . . 104
5.28 A320 – Sensitivity towards total system mass within design synthesis. . . . . . . 105
5.29 777 – Sensitivity towards total system mass. . . . . . . . . . . . . . . . . . . . . 105
5.30 A320 – Sensitivity towards total bleed air demand. . . . . . . . . . . . . . . . . 106
5.31 A320 – Sensitivity towards total shaft power demand. . . . . . . . . . . . . . . . 106
5.32 Impact of inaccuracy of input parameters. . . . . . . . . . . . . . . . . . . . . . 110

6.1 Impact of engine power off-takes on TOFL for ISA +15 conditions. . . . . . . . 114
6.2 Changes in TOFL due to off-takes. . . . . . . . . . . . . . . . . . . . . . . . . . 115
6.3 Impact of engine power off-takes on available take-off thrust. . . . . . . . . . . . 115
6.4 Comparison of fuel efficiency with and without engine power off-takes. . . . . . . 116
6.5 Optimization of W/S and T/W of A320 with engine power off-takes. . . . . . . 118
6.6 Optimization of W/S and T/W of A320 without engine power off-takes. . . . . . 118
6.7 Impact of engine power off-takes on conceptual aircraft design synthesis. . . . . 119
6.8 General arrangement of laminar aircraft. . . . . . . . . . . . . . . . . . . . . . . 122
6.9 Turbulent and laminar flight polar for a drag reduction of 35 DC. . . . . . . . . 122
6.10 Kinematic chain concept (by courtesy of DLR & EADS IW). . . . . . . . . . . . 122
6.11 Possible system topologies for the SADE kinematic chain concept. . . . . . . . . 123
6.12 Influence of system integration on system mass. . . . . . . . . . . . . . . . . . . 125
6.13 Allocated design parameters in MICADO. . . . . . . . . . . . . . . . . . . . . . 127
XVIII List of Figures

6.14 Probability distributions for the identified parameter space. . . . . . . . . . . . . 129


6.15 Optimum design points for the parameter sweep. . . . . . . . . . . . . . . . . . 130
6.16 Potential for fuel savings. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
6.17 Fuel savings for NLF aircraft with 30 DC reduction and normalized parameter
space. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
6.18 Joint probability density distribution for normalized parameter space (p(e x, ye)). . 132
6.19 Probabilities of fuel savings for the 3D joint probability function. . . . . . . . . . 132
6.20 Potential for improvements in specific hourly productivity. . . . . . . . . . . . . 133
6.21 Probabilities for improved SHP for the 3D joint probability function. . . . . . . 134

D.1 777 – Sensitivity towards bleed air temperature. . . . . . . . . . . . . . . . . . . 165


D.2 777 – Sensitivity towards required deflection speeds. . . . . . . . . . . . . . . . . 165
D.3 777 – Sensitivity towards required cabin temperature. . . . . . . . . . . . . . . . 165
D.4 777 – Sensitivity towards recirculation of cabin air. . . . . . . . . . . . . . . . . 165
D.5 777 – Sensitivity towards span position of wing anti-icing. . . . . . . . . . . . . . 166
D.6 777 – Sensitivity towards number of MCUs. . . . . . . . . . . . . . . . . . . . . 166
D.7 777 - Sensitivity towards MTOW. . . . . . . . . . . . . . . . . . . . . . . . . . . 166
D.8 777 – Sensitivity towards OWE. . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
D.9 777 – Sensitivity towards number of passengers. . . . . . . . . . . . . . . . . . . 166
D.10 777 – Sensitivity towards cabin and fuselage length. . . . . . . . . . . . . . . . . 166
D.11 777 – Sensitivity towards wing span. . . . . . . . . . . . . . . . . . . . . . . . . 167
D.12 777 – Sensitivity towards wing reference area. . . . . . . . . . . . . . . . . . . . 167
D.13 777 – Sensitivity towards design payload. . . . . . . . . . . . . . . . . . . . . . . 167
D.14 777 – Sensitivity towards design passenger number. . . . . . . . . . . . . . . . . 167
D.15 777 – Sensitivity towards total bleed air demand. . . . . . . . . . . . . . . . . . 168
D.16 777 – Sensitivity towards total shaft power demand. . . . . . . . . . . . . . . . . 168
List of Tables

3.1 Overview of main design parameters for conductor systems and energy sources. . 35
3.2 Overview of main design parameters for consumer systems. . . . . . . . . . . . . 36
3.3 Design cases for the environmental control system. . . . . . . . . . . . . . . . . . 42

4.1 Key specifications and performance characteristics of A320 reference aircraft. . . 79


4.2 Key specifications and performance characteristics of 777 reference aircraft. . . . 80
4.3 Key specifications and performance characteristics of the SA185 reference aircraft. 83

6.1 Estimated masses of both smart leading edge system topologies. . . . . . . . . . 124
6.2 Margins and initial estimates for the design parameters. . . . . . . . . . . . . . . 128

A.1 Definition of ATA100-chapters [12]. . . . . . . . . . . . . . . . . . . . . . . . . . 157

B.1 Deviations in corrected fuel flow between MICADO and GasTurb for the CF6-
80C2 engine. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
B.2 Deviations in corrected fuel flow between MICADO and GasTurb for the CFM56-
5B4P engine. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
B.3 Deviations in corrected fuel flow between MICADO and GasTurb for the GE90-
115B engine. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162

C.1 Top-level requirements for the SA185 design synthesis. . . . . . . . . . . . . . . 163

XIX
Nomenclature

Acronyms
A/C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Aircraft
AC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Alternating current
ACP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Air conditioning pack
ad. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Addendum
ADIRS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Air data and inertial reference system
AEA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . All electric aircraft
AEA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Association of European Airlines
AiX . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Aircraft exchange XML file
ap. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Appendix
APPU . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Asymmetry pick-off unit
APU . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Auxiliary power system
ARINC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Aeronautical Radio Incorporated
ARP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Aerospace Recommended Practices
ATA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Air Transport Association
ATA21 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ATA-21 chapter: Environmental control system
ATA22 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ATA-22 chapter: Auto flight system
ATA23 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ATA-23 chapter: Communication system
ATA24 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ATA-24 chapter: Electrical power system
ATA25 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ATA-25 chapter: Equipment and furnishing
ATA26 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ATA-26 chapter: Fire protection system
ATA27 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ATA-27 chapter: Flight control system
ATA28 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ATA-28 chapter: Fuel system
ATA29 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ATA-29 chapter: Hydraulic power system
ATA30 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ATA-30 chapter: Ice and rain protection
ATA31 . . . . . . . . . . . . . . . . . . . ATA-31 chapter: Instrumentation, indication & recording system
ATA32 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ATA-32 chapter: Landing gear system
ATA33 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ATA-33 chapter: Lighting system
ATA34 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ATA-34 chapter: Navigation system
ATA35 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ATA-35 chapter: Oxygen system
ATA36 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ATA-36 chapter: Bleed air system

XXI
XXII Nomenclature

ATA38 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ATA-38 chapter: Water & waste system


ATA49 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ATA-49 chapter: Auxiliary power system
ATA70 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ATA-70 chapter: Propulsion system
ATC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Air traffic control
CAE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Computer aided engineering or design
cf. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . from Latin: confer, ’compare’
CFD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Computational fluid dynamics
ch. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Chapter
CS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . EASA Certification Specification
CSD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Constant speed drive
DC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Direct current
DLH . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Deutsche Lufthansa AG
DLR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . German Aerospace Center
DR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Dispatch reliability
EADS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . European Aeronautic Defense and Space Company
EASA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . European Aviation Safety Agency
EC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . European Commission
ECS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Environmental control system
EHA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Electro-hydraulic actuator
EMA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Electro-mechanical actuator
etc. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . from Latin: et cetera, ’and so on’
EU . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . European Union
FAME-S . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Fast and Advanced Mass Estimation - Systems
FAR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Federal Aviation Regulations
FCC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Flight control computer
FCS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Flight control system
FEM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Finite element method
HARLS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . High aspect ratio and low sweep wing
HLFC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Hybrid laminar flow control
i.e. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . from Latin: id est, ’it means’
IATA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . International Air Transport Association
ICAO . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . International Civil Aviation Organization
IFE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . In-flight entertainment system
ILR . . . . . . . . . . . . . . . . Institute of Aeronautics and Astronautics of RWTH Aachen University
IMAs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Integrated modular avionics
ISA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . International standard atmosphere
L/G . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Landing or landing gear
Nomenclature XXIII

LCC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Life-cycle costs


LEWIS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . NASA dynamic ice accretion software code
LTH . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Luftfahrttechnisches Handbuch
MBD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Multi-body dynamics
MCU . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Modular concept unit
MEA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . More Electrical Aircraft
MICADO . . ILR Multidisciplinary Integrated Conceptual Aircraft Design and Optimization
misc. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Miscellaneous
MOD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Multi-disciplinary design optimization
MOET . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . More Open Electrical Technologies
1
NACRE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . New Aircraft Concepts REsearch
NASA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . National Aeronautics and Space Administration
NLF . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Natural laminar flow
OEI . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . One engine inoperative
ONERA . . . . . . . . . . . . . . . . . . . . . . Organisation National d’Étude et de Recherche Aérospatiale
PAX . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Passengers
PCU . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Power control unit
Ph.D. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Doctor of Philosophy
POA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Power Optimized Aircraft
PrADO . . . . . . . . . . . . . . . . . . . . . . . . . . . . Preliminary Aircraft Design and Optimization Program
PTU . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Power Transfer Unit
R&T . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Research and technology
RAE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Royal Aircraft Establishment
RAT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Ram air turbine
Reqs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Requirements
2
SADE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . SmArt High-Lift DEvices for Next-Generation Wing
SAE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Society of Automotive Engineers
Specs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Specifications
THS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Trimable horizontal stabilizer
TL . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Torque limiter
TRU . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Transform rectifier unit
TsAGI . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Central Aerohydrodynamic Institute Russia
UML . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . unified modeling language
w/o . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Without
WTB . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Wing tip brake
1
EU 6th Framework Program Collaborative project
2
EU 7th Framework Program Collaborative project
XXIV Nomenclature

XML . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . extensible markup language

Parameter
q̄D . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Design dive dynamic pressure, N/m2
ṁ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Mass flow of ... (see acronyms/subscripts), kg/s
Q̇ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Heat flux of ... (see acronyms/subscripts), W/s
m̂ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Length specific mass of ... (see acronyms/subscripts), kg/m
x
e, ye, ze . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Normalized design parameters
A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Reference area of ... (see acronyms/subscripts), m2
C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Configuration of ... (see acronyms/subscripts), ❂
CD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Drag coefficient, –
CF . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Friction coefficient, –
CL . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Lift coefficient, –
CM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Pitching moment coefficient, –
cp . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Heat capacity of air, J/kg K
cw . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Heat capacity of water, Btu/lb F
Cbleed,Scholz . . . . . . . . . . . . Constant for estimating repercussions of bleed air off-takes [13], 1/K
Chinge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Control Surface hinge coefficient, –
cmean,chord . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Mean chord length of control surface, m
crudder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Mean control surface width, m
Cshaf t,Scholz . . . . . . . Constant for estimating repercussions of shaft power off-takes [13], N/W
CAS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Calibrated airspeed), kts
CG . . . . . . . . . . . . . . . . . . . . . . . . . . Center of gravity position in ... (see acronyms/subscripts), m
D . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Drag force, kN
d . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Diameter of ... (see acronyms/subscripts), m
DOC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Direct operating costs, USD/1000 PAX km
E . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Energy of/for ... (see acronyms/subscripts), J
f . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Functional requirement of ... (see acronyms/subscripts), –
F F . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Form factor for estimating viscous drag, –
g . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Gravity acceleration, 9.81 m/s2
h . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Altitude, ft
h . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Heat transfer coefficient, W m−1 K−1 or Btu h−1 ft−2
ka . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Airfoil technology factor for estimating wave drag, –
L . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Characteristic length of ... (see acronyms/subscripts), m
L/D . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Lift to drag ratio (L-over-D)
Le . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Heat of evaporation, Btu lb−1
Nomenclature XXV

LDR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Landing distance required, m


LW C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Liquid water content of atmosphere, g/m3
M . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Water catch of wing surface, lb/hr ftspan
m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Mass of ... (see acronyms/subscripts), kg
M a . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Flight Mach number, –
M F W . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Maximum fuel weight, kg
M LW . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Maximum landing weight, kg
M M O . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Maximum operating Mach number, kg
M T OW . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Maximum take-off weight, kg
n . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Rotation speed of ... (see acronyms/subscripts), 1/min
N, n . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Number of ... (see acronyms/subscripts), –
N1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Engine power setting resembled by shaft speed, %
OW E . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Operating weight empty, kg
P . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Power or power off-takes, W
p . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Coefficient of beta distribution
p . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Pressure, Pa
p(x) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Probability density function
pskin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Vapor pressure at wing surface temperature, psi
Q . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Heat load of ... (see acronyms/subscripts), W
Q . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Interference factor for estimating viscous drag, –
q . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Coefficient of beta distribution
q . . . . . . . . . . . . . . . . . . . . . . Specific heat load of ... (see acronyms/subscripts), W m−2 or W m−1
R . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Range of/at ... (see acronyms/subscripts), NM
Re . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Reynolds number of/at ... (see acronyms/subscripts), –
S . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Area of ... (see acronyms/subscripts), m2
SAR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Specific air range, kg/m
SF C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Specific fuel consumption, mg/Ns
SHP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Specific hourly productivity, NM/h
SLST . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Sea level static thrust, kN
T . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Temperature of/at ... (see acronyms/subscripts), K
T . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Thrust force, kN
t . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . time of ... (see acronyms/subscripts), sec
t/c . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Thickness to chord ratio of wing or airfoil, –
T /W . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Thrust-to-weight ratio, –
T AS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . True airspeed, kts
T ODR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Take-off distance required, m
XXVI Nomenclature

T T C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Time to climb, min


U . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Utilization of ... (see acronyms/subscripts), %
V . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Volume of ... (see acronyms/subscripts), m3
V M O . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Maximum operating speed, kts
W/S . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Wing-loading, kg/m2
x, y, z . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Design parameters

Symbols
α . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Angle of attack, deg
∆DC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Laminar drag savings, DC
γ̇ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Control surface deflection speed, deg/s
η . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Efficiency factor
γ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Control surface deflection, deg
Γ(r) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Gamma function
λ . . . . . . . . . . . . . . . . . . . . . . . . . . . Thermal conductivity of ... (see acronyms/subscripts), W/m K
λrudder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Width-depth ratio of control surface,
µ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Statistical mean
ρ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Density, kg/m3
σ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Statistical variance
ϕ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Angle of sweep of ... (see acronyms/subscripts), deg

Imperial Units and Other non-SI Units


Btu . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . British thermal unit, 1055.05 J
DC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Drag count, 0.0001 CD
F . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Degree Fahrenheit, (F + 459.67) · 5/9K
ft . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Feet, 0.3048 m
kts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Knots, 0.5144 m/s
lb . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Pounds, 0.453 kg
NM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Nautical miles, 1852 m
psi . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Pound-force per square inch, 6894.76 Pa

Subskripts
41 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Low pressure turbine entry station
45 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Engine core station
act . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Actuator
aw . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Adiabatic wall temperature
bleed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Bleed air
cabin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Passenger cabin
Nomenclature XXVII

comp . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Component
control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Control surfaces
conv . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Convection
crew . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Flight and cabin crew
crit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Critical
DD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Drag divergence
design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Design case
drop,mean . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Mean droplet diameter
elec . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Electric
eng . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Engine
fuel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Fuel
func . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . functional
gear . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Landing gear
heat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Heating
high lift . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . High-lift system
hydr . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Hydraulic
i . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Counter variable
init . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Initial
install . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Installation
L/G . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Landing
LE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Leading edge
lift . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Due to lift force
local . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . At local coordinate station
max . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Maximum
mission . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Transport mission
MO . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Maximum operating
off . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Off-takes
pump . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Pump
ref . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Reference wing area
req . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . requirement or required
sec power . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Secondary power system
shaft . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . From engine shaft
skin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Skin or surface of e.g. the wing
subsys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Subsystem
surf . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Control surface
sys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . System
vapor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Vaporating
XXVIII Nomenclature

vent . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Ventilation
visc . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Viscous
wave . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Due to shock wave effects
wet . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Wetted area
x . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . In longitudinal direction
y . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . In span direction
1 Introduction

Today, authorities and society call for leaps in aircraft efficiency for minimizing aviation’s
environmental footprint and for counteracting soaring oil prices. Ambitious research goals
have been set by the International Air Transport Association (IATA) [14, 15], the European
Commission (EC) [16, 17], and other authorities [18, 19, 20] for the year 2020 and beyond. Great
potential lies in the introduction of innovative and optimized technology on aircraft system-
level. At the same time, industry seeks for more integrated and optimized aircraft system
architectures to cope with steadily increasing vehicle complexity and extremely competitive
market environments [21].

1.1 Aircraft Systems Technology and Innovations

The term ’system’ is used in many contexts and not always in a technical one. In the scope of
this thesis, an ’aircraft system’ is defined as an assembly of different subsystem or components
that provide one or more predefined aircraft-level functions1 [22]. In figure 1.1, a generic
aircraft system is shown. The desired aircraft-level function is the typical output. The system
assembly of subsystems and components (items) is confined by the system boundary. Only
energy, information, and mass fluxes can be shared over the system boundary [23, ch. 4]. For
systems design it is mandatory to follow a clear definition of the single systems and to follow the
systems boundaries closely. This includes allocation of the different aircraft-level functions and
components to the single systems. The author follows the established Air Transport Association
(ATA) classification for the conventional systems architecture [12]. A short description of the
functions and components included in each ATA-chapter is provided in appendix A.

Together with increasing maturity of available technology and more multifarious requirements
regarding mission and operations, the complexity and capabilities of aircraft systems increases
steadily [24]. Today, aircraft systems fulfill essential functions for operations, safety, efficiency
and passenger comfort. A confined view on single systems is not sufficient for successful in-
1
Aircraft-level functions are top-level functions that describe overall capabilities, and that specify functional
requirements for the allocated aircraft system.

1
2 Chapter 1. Introduction

tegration of different systems into a complex overall architecture. A clear definition of the
interfaces between different systems, to other aircraft components, and also to the human op-
erators is required [22]. The term ’systems architecture’ refers to the overall systems structure
that provides the full functionality to the aircraft ([23, ch. 5], [1, ch. 5]). It includes not only
the different systems, but also their interfaces and their integration on overall aircraft-level. A
classification according to Moir [1, ch. 5] is used in order to give a brief overview of the conven-
tional systems architecture of transport aircraft. The systems are grouped in energy systems,
standard equipment, cabin systems, and operators equipment, see figure 1.2.


  
 
   
     
 
   




Figure 1.1: A generic aircraft system.



 

 
             
    $  
    $
     ) 
           






 


                
 
      
 !"    #!$%      
 &  '     &   
 (    )("
  *

   

Figure 1.2: Classification of the different aircraft systems of civil transport aircraft [1, ch. 5].

The propulsion system provides the required thrust; it is often referred to as ’primary power
system’. The ’secondary power systems’ 2 (hydraulic, electric and bleed air system) obtain
their energy from the engines3 (shaft and bleed air off-takes) and provide the single aircraft
systems with the required energy. A schematic overview of the conventional secondary power
distribution is shown in figure 1.3. Bleed air is mainly used for the ECS and the wing anti-icing
2
The terminology ’secondary power systems’ was formed throughout various NASA sponsored research ef-
forts ([25], [26], [27] or [28]); it will be also used throughout this thesis.
3
This is for standard flight only; failure cases are not discussed in this brief overview.
1.1 Aircraft Systems Technology and Innovations 3

system. Hydraulic power is provided by engine driven hydraulic pumps. Electric generators
provide the electrical system with power. Typical alternate current (AC) consumers are high
voltage systems such as galleys, lavatories, and other cabin systems. In normal flight, direct
current (DC) is provided by transformation of AC within the aircraft. DC is mainly used
by low voltage systems such as flight control computers, avionics, and cockpit systems. The
standard equipment (cf. fig. 1.2) includes all systems that are required for safe flight. These
systems are mainly defined by the different certification regulations of the various aircraft types.
Cabin systems are required for the typical mission of a commercial transport aircraft but not
for maintaining safe flight. Further, aircraft are often equipped with operator specific systems.
These systems are not required by certification regulation but increase the operational envelope,
the efficiency or reduce the pilot’s work load. For military aircraft the weaponry or systems
for electronic warfare and defense must be additionally considered [29, ch. 1]. For detailed
information on e.g. functions, power supply, or integration of those systems further reading in
e.g. Collinson [30], Fielding [31], Moir [32, 33], Roskam [34], or Scholz [35, ch. 12] is suggested.

 "#$  





    
  


 
 
 


       

  
  


  
 

  ! 


 

  
 

 
 

Figure 1.3: Schematics of conventional power systems architecture.

Technology improvements in aviation during the last decades, were mainly driven by step-
changes rather than revolution of aircraft technology [36, 37]. Steady improvements in turbofan
technology, and thus continuous improvements in fuel efficiency are an example for such evolu-
tions (e.g. [38], [39, p. 60]). After decades of continuous improvements, transport aircraft are
in a highly optimized state today. Hence, costs for further evolutionary improvements increase
steadily, whereas the overall impact on aircraft efficiency decreases [40, 21]. This phenomenon
of saturation in technology gain was first addressed by Foster [2]. He introduced an ’S-curve’
to describe the required costs and time for improving an arbitrary technology.

Conventional jet transport aircraft are reaching, or already have reached the upper part of the
’S-curve’ [40, 41, 42], see figure 1.4. In a medium to long term perspective, leaps in aircraft
performance and efficiency can only be achieved by introduction of innovative technology, which
4 Chapter 1. Introduction

is directly linked to aircraft systems technology and systems design [21, 37, 43, 44, 45]. Another
interesting aspect, which is shown in figure 1.4, is related to the lacking maturity of innovative
technology. It is commonly agreed that newly emerging technologies, at first, cannot compete
with existing and highly optimized ones. New technologies need investments and in-service
experience to mature and to show their real potential against well known and service proven
ones. The propfan or open-rotor technology is an example for a technology that failed first.
Already in the 1980s, a MD-80 was used to flight test an open-rotor engine [46]. Repercussions
of the technology4 gave the open-rotor engine no advantage over the common turbofan engines.
Today, with matured technology, improved design methods as well as different economical and
ecological constraints, the technology emerges again. From today’s perspective, the introduction
of new propulsion systems, for instance propfan or geared-turbofan engines, as well as of active
laminar flow control5 seem to be the most promising system-level technologies for enhancing
efficiency (e.g. [48], [49, ch. 2], [14, ch. 3], [50], [44, 51, 43, 44, 37, 52], [53] or [45]).

Evolution of system technology resulted in a


highly complex systems architecture with nu-
merous interfaces between the different sys-
   

 

tems. Current research does not only focus

on innovative systems technology in order to
meet the required environmental goals, but
also to reduce complexity and costs of new air-
craft programs. The U.S. Defense Advanced 



Research Projects Agency (DARPA) tries to
face increasing systems complexity, and thus
increasing time for design, integration, and 

testing with more flexible approaches to sys-
Figure 1.4: The technology S-curve [2].
tems design [21]. DARPA and others argue
that the classical strategy in systems design,
which is often represented by the V-model [54], cannot cope with highly integrated architec-
tures. According to DARPA [21] failures in design often occur at the many interfaces between
the different systems when the classical design approach is applied. The need for new design
approaches is directly mirrored in delays and significant budget overruns of current aircraft
programs. Examples are the Airbus A380 [55], and A400M [56] programs as well as the Boe-
ing 787-8 [57, 58], 747-8 [59], and F-35 [60] programs. Airbus also concluded with delays for
the A350XWB program [61, 62], the first flight was initially targeted for mid-2012. It was
shifted by approximately twelve months to June 2013 [63]. Systems design and integration is
the main cause for the delays and budget overruns in all of the cited programs. Changes in
4
Mainly noise emissions, vibrations and integration issues.
5
A technology that was also heavily researched in the 1980s [47].
1.1 Aircraft Systems Technology and Innovations 5

systems architecture towards more integrated architectures gain an increasing focus to tackle
such problems [64]. Until recently, efforts for optimization of current technology mainly con-
centrated on single systems solutions. For innovative systems design this has changed towards
optimization of the systems architecture as a whole. Current research aims at more modular,
more integrated, and especially at more-electrical approaches in systems design. Comprehen-
sive overviews of research targets, necessary technology, as well as challenges of such approaches
can be found for example by Emadi et al. [65], Faleiro [36], and Rosero et al. [66]. Integrated
modular avionics (IMAs)6 are already used in recent aircraft, e.g. Airbus A380 and A350,
Boeing 787 or F-22 Raptor [30, ch. 9]. Airframers have gained experience with integration of
several non-critical applications within common computing resources. IMA systems give an
idea how more integrated future systems architectures will look like in next-generation aircraft.
Key enabler for IMA technology and highly integrated avionic systems are matured digital data
bus technology, which allows for the required data integrity and speed, as well as modular and
standardized computer modules that are also applicable in safety critical applications [67] [68].

The trend towards more electric system architectures is evident. Modern aircraft replace hy-
draulic or bleed air systems with systems that rely on electric energy. Literature differs between
the more electric aircraft (MEA) and the all electric aircraft (AEA). The AEA concept implies
that only electric power is taken off the engines and distributed in a centralized power system
architecture ([32], [69]). Thus, both the hydraulic system as well as the bleed-air system are
removed in an AEA. Projected benefits are reduced complexity as well as increased efficiency
and reliability. In figure 1.5, schematics of a state of the art AEA systems architecture are
shown [66]. The main differences to the conventional power system architecture (c.f. figure 1.3)
are evident. Repercussions of the AEA concept are power electronics that have to provide power
densities, which are (yet) not available with current technology. According to Rosero et al. [66],
an electrical system that provides around 1.6 MW would be required for a next-generation wide-
body AEA that is designed for the 300 PAX class. Furthermore, power electronics produce high
thermal loads that have to be managed by the ECS. Size of the generators, redundancies, and
fire protection are other issues.

As an intermediate step towards the AEA, airframers concentrate on the MEA approach. The
Airbus A350 and A380 as well as the Boeing 787 follow this concept. The A350 and A380 are
equipped with only two hydraulic systems, and use electric actuation as alternate means of flight
control. The Boeing 787 goes a step further. It features a bleed-less system architecture [70],
in which the ECS and the ice protection system are powered from the electric system. Installed
engine generator capacity is boosted from 240 kW of the 777 to 1000 kW in the 787.

6
For a definition of IMA technology please refer to Moir [33, ch. 11].
6 Chapter 1. Introduction


 



   


 
 


  
   
 

   


 
    

 




 



Figure 1.5: Schematics of all-electric power systems architecture.

1.2 Repercussions of Aircraft Systems Integration

The main impacts of aircraft system integration on overall aircraft-level are shown in figure 1.6.
Aircraft design synthesis is always governed by a defined set of top-level requirements, design
specifications and certification regulations [71, ch. 2]. In aircraft design, the vehicle is sized and
analyzed within an iterative design synthesis (central block). Design of the different aircraft
systems is today often excluded from the early phases of aircraft design7 . When adding systems
design (left block) to the overall design synthesis, repercussions occur. Systems mass as well as
power off-takes influence block fuel, sizing of the propulsion system, and thus overall aircraft
gross weight and efficiency. Only integration of systems design into iterative design synthesis
allows for capturing such snowball effects8 . After achieving convergence in design synthesis,
the aircraft is assessed on overall aircraft-level. Assessment is based on different parameters
(e.g. block fuel or operating costs) as well as on scenarios and weighting functions.

1.2.1 Direct Repercussions of Engine Off-Takes

Dollmayer [72, 73] concludes that engine off-takes contribute to an increase in fuel consumption
of about 5 % for transport aircraft in the 150 PAX-class; Faleiro [36] and Scholz [74] have
reached the same conclusion. This corresponds well to research results that were published by
the author [3]. In figure 1.7, the influence of engine off-takes on specific fuel consumption (SFC)
for cruise condition of an A320-class aircraft is shown. The SFC increases approximately by 4
to 5 % for typical cruise loads. The increase in SFC is almost linear for both shaft power as
well as for bleed air off-takes. However, bleed air off-takes have a far more severe impact on
SFC, which is caused by their impact on the thermodynamic engine cycle [75, 76].

7
Mostly in conceptual and preliminary aircraft design
8
Amplifying of single impacts on an overall-level within iterative design synthesis.
1.2 Repercussions of Aircraft Systems Integration 7

 
       
  


 
   

   


      
       
 #    
  
 
 
   
 !   
   " 
  
  !!$#%
 "   
 
 &  
 #! 
 " 

 
   


Figure 1.6: Influences of aircraft system integration on overall aircraft-level.

Further, maximum available bleed air



off-takes depend on the core mass flow of
 
 

the engine ([36], [77, ch. 3]), which leads 

to an issue of modern high bypass ratio 


engines. The core mass flow is reduced in
 
favor of fan air mass flow with increasing 
     

    

bypass ratio. Thus, maximum available 

bleed air decreases with higher bypass ra-


tios. Modern engines do not provide suf- Figure 1.7: Influence of power off-takes on SFC
ficient bleed air when they are operated (35,000 ft and Mach 0.80) [3].
close to idle. Already for a Boeing 767
aircraft with a medium bypass ratio engine9 , the required bleed air exceeds the available supply
when the engines are operated in descend conditions10 [78]. In order to supply sufficient bleed
air, engines must operate at a higher thrust setting; higher fuel burn is the consequence. Thus,
evolution to bleedless system architectures is also facilitated by the trend towards ultra-high
bypass ratio engines.

9
General Electric CF6-80C2, bypass ratio 5.3
10
Idle descend, 22,000 ft, Mach 0.55
8 Chapter 1. Introduction

Influences of engine power off-takes in combination with overall system architectures are not
negligible for conceptual or preliminary design studies of fuel optimized aircraft and technology
assessment on overall aircraft-level. Hence, a fast and robust systems model is required for
integration already in conceptual and preliminary aircraft design synthesis. Such a systems
model must allow for estimating engine off-takes for different systems architectures and for
each single mission increment.

1.2.2 Snowball Effects in Overall Aircraft Design Synthesis

Commonly, conceptual and preliminary aircraft design mostly neglect snowball effects of system
integration. However, changes in systems architecture impacts system weights and engine power
off-takes, which leads to snowball effects in design synthesis. Higher fuel consumption inevitably
leads to an increase of mission fuel11 , and thus to a higher required maximum take-off weight
(MTOW). The higher MTOW influences the entire aircraft sizing and also itself recursively
within the iterative design synthesis, see figure 1.8. These snowball effects can be described by
mass growth factors to some extend [80]. Such mass growth factors, however, do not provide the
required sensitivities to design parameters that are mandatory for sound technology assessment.


   



 
$ %   #
  !"

   





Figure 1.8: Iterative design methodology.

Additionally, engine power off-takes reduce the available thrust. Again, bleed air off-takes have
the more severe effect. At the beginning of each design, a thrust-to-weight ratio (T/W) as
well as a wing-loading (W/S) is chosen12 . T/W is generally chosen to the lowest possible value
to improve fuel efficiency. Design studies have shown that with a chosen T/W, the available
thrust is often not sufficient to fulfill performance requirements13 when power off-takes are
incorporated. Hence, an increase of T/W and accordingly re-sizing of the installed power
plant is required. This, along with the corresponding snowball effects, leads to additional
significant impacts on overall aircraft-level within iterative design synthesis. This facilitates
11
Total loaded fuel at break release (including block fuel and total reserves), see Airbus [79, ch. 1].
12
Initial sizing from the conceptual step, see Dobrev [81] and Anton [82].
13
Those are either requirements for take-off, climb, ceiling or cruise performance. They are derived from
top-level aircraft requirements as well as from certification regulations.
1.3 Motivation, Scope and Field of Application 9

the conclusion that for correct aircraft design, systems architecture sizing must be considered
already in conceptual and preliminary design synthesis.

1.3 Motivation, Scope and Field of Application

Commercial aircraft are value creating products ([83], [84, ch. 4]). Hence, it is the main goal
to design the aircraft for maximum value rather than for instance for best performance (e.g.
speed or agility). Experience shows that high-performing absolutely state-of-the-art aircraft,
like the Concorde was at its time, do not necessarily turn out to be commercially successful.
Instead well balanced designs turn out to be the most successful ones. Measuring the actual
product value is not easy. Aircraft designs are assessed based on economics, and on direct
operating costs (DOC) in particular ([85], [84, ch. 6]). System architecture design is closely
linked to overall aircraft design characteristics such as aircraft gross weight, fuel consumption,
DOC and life cycle costs (LCC)14 . Hence, it has a strong influence on the market value of the
aircraft. According to Faleiro [36], three options exist to improve efficiency and to reduce costs:
optimization on system-level (equipment and engines), optimization of the systems architecture
for optimum fulfillment of required aircraft-level functions, and optimization of the aircraft as
a whole. The optimization on system-level is primarily targeted in detail systems design. The
other two options, however, are closely linked to conceptual and preliminary aircraft design
synthesis and multi-disciplinary design optimization (MDO), and are in the focus of this work.

According to Raymer [71, ch. 2.2], conceptual and early preliminary aircraft design is char-
acterized by the overall sizing of the aircraft’s configurational arrangement, size, weight, and
performance. Numerous alternative design concepts are prepared in response to the design
requirements, and numerous variations on those concepts are studied. Thus, a wide design
space is investigated for down-selection of the most promising concepts. The required accuracy
of e.g. predicted weight, performance or size is only of secondary importance. The expected
level of accuracy in conceptual and early preliminary aircraft design usually lies between 10
and 15 % [86, 71]. It is far more important to show the correct trends and effects within the
investigated design space for valid optimization and down-selection.

Today systems design has largely been neglected in conceptual and preliminary aircraft design.
Statistical relations are used to estimate only systems mass [87]; other repercussions of systems
integration are neglected. However, repercussions of systems integration must be captured for
gaining valid results and for sound decision-making. Accurate15 but also fast models are re-
quired in conceptual design to estimate relevant system parameters and to model repercussions
14
A detailed discussion on aircraft system integration on overall aircraft-level is provided in chapter 1.2.
15
Accurate especially in showing the right sensitivities to the relevant aircraft design parameters.
10 Chapter 1. Introduction

of systems integration. A particular difficulty lies in the low level of detail that is available,
compared to that required for sophisticated systems modeling. This is especially true for inno-
vative systems since many uncertainties in the design parameters remain rather than providing
accurate absolute values.

In the scope of this thesis, a methodology for modeling of aircraft systems and entire systems
architectures is presented that closes the current gap of sound systems modeling in conceptual
aircraft design synthesis. The proposed methodology allows for interdisciplinary optimization of
aircraft design by also taking into account repercussions of systems integration. Each system is
sized in terms of mass and mission dependent energy consumption by its functional performance
requirements. Also, installation location and system interfaces are considered. The modular
and flexible structure of the proposed methodology allows for fast integration of changes in the
overall system architecture, and thus for comparison and assessment of different design solutions
and technologies for sound decision-making. The proposed methodology further incorporates
an approach towards modeling of also innovative systems. This approach uses probabilistic
theory, and thus includes remaining design uncertainties in the assessment of technology. The
methodology has been implemented into a systems model. The required level of detail is suited
for application in conceptual design. Still, the model captures required influences and shows
required sensitivities to relevant requirements and aircraft characteristics. Furthermore, the
proposed systems model has been directly integrated into an automated overall aircraft design
synthesis: the ILR Multidisciplinary Integrated Conceptual Aircraft Design and Optimization
(MICADO) environment. Integration of the proposed systems model into MICADO allows
for repercussions of systems integration to propagate onto overall aircraft-level. Assessment of
technology in highly integrated design spaces becomes possible.

The available knowledge of a system or sys-


tems architecture also correlates with the dif-  !


 




 



ferent design phases [88]. As illustrated in 


  


  


figure 1.9, Moir [1, ch. 5] brings the available
knowledge of the aircraft systems architecture
into context as follows: In early conceptual 
   
aircraft design, only a basic idea of the sys-
tems architecture and its functionality exists.
  
This idea converges to a clear definition of   

the system architecture down to system-level 




during conceptual design. A definition of the 


   
most important interfaces as well as of the
Figure 1.9: Knowledge during design.
main functions of each system is provided.
Detailed knowledge of single subsystems or
1.4 General Overview 11

even systems components is not yet available. Hence, the author limits the proposed sys-
tems model to modeling of the systems architecture down to system-level. Also, only the main
aircraft-level functions as well as energy and mass fluxes are considered for sizing of the system
architecture. Further, the expected accuracy of the proposed systems model lies in the typical
range of that expected in conceptual design (see above). Emphasis is put on capturing of rele-
vant repercussions of systems integration and their propagation onto overall aircraft-level with
the proposed systems model for investigation of large design spaces.

This work concentrates on civil jet transport aircraft that are certified according to CS-25 [9]
or FAR-25 [89] regulations. The proposed systems model is tailored for application to such
aircraft. The modular structure along with the proposed methodology, however, allow for easy
adaption of the proposed systems model to other fields of application, e.g. also for military or
space applications.

1.4 General Overview

This introduction has given an overview of the research topic and aircraft system technologies.
The motivation of the author to present a new methodology for systems modeling in conceptual
aircraft design has been outlined.

In chapter 2, a literature survey outlines the different approaches and state-of-the-art in systems
modeling. Special emphasis is put on applicability of the available models in conceptual design
and to innovative system. The chapter concludes why a new methodology is required, and thus
underlines the author’s motivation.

The author’s approach towards system modeling is described in chapter 3. First, dependencies
of the proposed modeling approach on aircraft-level functions and on other design as well as
operational characteristics are outlined. Then, the proposed methodology is explained based
on generic system description and on interdependencies within a generic systems architecture.
Application of the principle of energy balance for sizing of the single systems and for estimating
overall power requirements is introduced. Adjacently, discussion of some of the underlying
mathematical system models as well as considerations regarding implementation of the proposed
systems model into a stand-alone software tool is provided. The chapter closes with the author’s
novel framework for probabilistic modeling of innovative systems.

In chapter 4, integration of the proposed systems model into overall conceptual aircraft design
synthesis and optimization is discussed; the MICADO environment is introduced. Emphasis
is put on integration of the proposed systems model into MICADO and on capturing of the
12 Chapter 1. Introduction

repercussions of system integration. The chapter closes with description of three reference
aircraft for the presented case studies.

Chapter 5 focuses on application, verification, and sensitivity studies of the proposed systems
model. First, results regarding sizing of the system architecture of the reference aircraft are
presented. Emphasis is put on verification of the proposed systems model with regards to
available data of commercial transport aircraft. Plausibility and feasibility of the proposed
model is further assessed by means of sensitivity studies. The chapter concludes with the
expected level of accuracy and limitations of the proposed systems model.

In chapter 6, two case studies are presented that show application of the proposed methodology
for sound technology assessment. The first case study concentrates on impact of engine power
off-takes on aircraft performance and efficiency, and thus concludes with the importance of
systems integration already in conceptual aircraft design. The second case study concentrates
on integration of innovative system technology into the proposed systems model. It concludes
with probabilistic assessment of possible improvements in efficiency and economics on overall
aircraft-level.

The thesis closes with main results and discussion of key benefits of the proposed approach
towards modeling of aircraft systems in conceptual aircraft design, see chapter 7.
2 Related Work in Systems Modeling

Different available approaches from literature towards the modeling of aircraft systems are
discussed in this literature survey. The survey starts with statistical approaches that are sug-
gested by literature for application in conceptual aircraft design (see ch. 2.1). Drawbacks of
these approaches are discussed. Then, different approaches for a more detailed modeling of sin-
gle systems are introduced (see ch. 2.2), which are applicable during preliminary and detailed
aircraft design. The literature survey closes with an overview of available models for sizing of
entire systems architectures (see ch. 2.3) and for incorporating uncertainties into design syn-
thesis (see ch. 2.4). For evaluation of the different available approaches, emphasis is put on
their applicability during conceptual aircraft design and on flexibility for integration of also
innovative system technology. The chapter concludes why a new methodology is required, and
thus underlines the author’s motivation for this work (see ch. 2.5).

2.1 Statistical Approaches for Estimating Systems Mass

System models that are commonly used in early phases of aircraft design are based on statistical
approaches and regression analyses. Such models were published amongst others by Luftfahrt
Technisches Handbuch [90], Howe [91, ad. 4], NASA [92], Torenbeek [93, ch. 8], Raymer [86,
ch. 15] Roskam [34, ch. 7], or Kundu [94]1 . They concentrate on systems mass and fail to
estimate engine power off-takes. Databases were used to derive relations between systems weight
and typical aircraft parameters such as maximum take-off weight (MTOW). Available data from
the 1st and 2nd generation2 of jet transport aircraft were mostly used for the regression analyses.
Hence, already the improvements in systems technology of today’s aircraft (4th generation) are
not captured correctly by these models. The available models offer a varying perspective on
the systems architecture. Most of them only allow to estimate total systems mass, others also
allow a mass breakdown onto systems-level.

1
This listing is not complete; each book on aircraft design offers a more or less detailed approach for estimating
systems weight.
2
Entry into service of aircraft between 1955-1975.

13
14 Chapter 2. Related Work in Systems Modeling

2.1.1 Weight Fractions and Regression Analysis

The simplest models use weight fractions of MTOW to estimate either total systems mass or
single systems masses. Howe [91, ch. 6] for example suggests to estimate total system mass of
jet transport aircraft to approximately 20% of maximum take-off weight. Weight fractions do
not show the required sensitivities towards relevant design parameters, and are thus not further
discussed in the scope of this thesis. Models that are based on regression analysis usually show
a non-linear dependency of systems mass on one or more overall aircraft parameters. Some only
estimate total systems mass. Berg [95, ch. 2] suggests a regression with MTOW for estimating
total systems mass, see equation 2.1. In figure 2.1, the regression function is plotted against
known total aircraft systems masses. Close correlation with the known data can be observed.
Nevertheless, equation 2.1 shows sensitivity towards only one design parameter.
(
1.6 · 10−7 · M T OW 2 + 0.11 · M T OW + 820 for M T OW ≤ 260 t
msys = (2.1)
0.026 · M T OW + 11, 600 for M T OW > 260 t
Total Systems Mass, 1,000 kg

25

20
A330-300
A340-300
A300-600 A330-200
15
A310-300
10 A321-200
737-200 A320-200
5
A319-100 aircraft data
A318-100
Saab 2000 regression
0
0 50 100 150 200 250 300 350 400
Maximum Take-Off Weight, 1,000 kg

Figure 2.1: Regression analysis for estimating total systems mass.

Other models estimate mass of single systems. This, however, is not done in a consistent way,
e.g. by following the ATA-chapter classification. Different models often combine single systems
to systems groups and give estimates for those only. Torenbeek [93] for example groups avionics,
electric system and instrumentation, while Roskam [96] groups avionics and instrumentation.
Roskam [96, ch. 5-7] also uses regression analysis for estimating systems mass. He provides
estimates either for single systems or for systems groups. Systems mass of the FCS (ATA-27)
is estimated as a function of MTOW and the design dive dynamic pressure (q̄D ):
 0.576
M T OW /0.4536 · q̄D /47.88
mAT A27 = 56.01 · . (2.2)
100, 000
This relation, however, is only valid for conventional mechanical flight controls; it is not appli-
cable to modern electrical FCS. It is not possible with Roskam’s methods to estimate systems
2.1 Statistical Approaches for Estimating Systems Mass 15

mass of the ECS and the anti-icing system separately. He groups both systems and suggests
equation 2.3 for estimating their combined mass. It is a function of cabin volume (Vcabin ) as
well as of the number of passengers (NP AX ) and crew (Ncrew ).
 0.419
Vcabin /0.02832 · (Ncrew + NP AX )
mAT A21+AT A30 = 469 · (2.3)
10, 000
The statistical models that were published by Luftfahrt Technisches Handbuch (LTH) [90] use
a wider range of input parameters than most others. It is also possible to estimate component
mass for some systems. Total hydraulic systems mass is, for example, accumulated by com-
ponent masses of pumps, ducting, reservoirs, fluid etc. [97]. As an example, the regression for
estimating hydraulic pump mass is given as follows:

0.711 LAT A29 0.987 0.467


mAT A29,pump = 0.61 · NAT A29,subsys · · Acontrol . (2.4)
10
It is a function of the number of hydraulic subsystems (NAT A29,subsys ), the characteristic length
of hydraulic ducting (LAT A29 ) and the sum of the control surface reference areas (Acontrol ).
Compared to the previous discussed regressions, this one is based on input parameters that
have a direct physical relation with pump weight. Hence, systems sensitivities can be indicated
by the LTH models to some extent.

A comparison between the different statistical approaches that are given in literature is difficult
due to the inconsistencies in the chosen system boundaries; some attempts for such comparison
exist. Berg [95] concludes that it is not a single model that shows the highest accuracy. He
suggests to rather use a mixture of different published models depending on the system that
is being modeled. Fernandes da Moura [98, ch. 6] also shows that results vary widely, and
that a direct comparison is not easy. Koeppen [11, ch. 2] provides a more detailed analysis
of deviations between different statistical models. He identifies two main reasons for large
deviations that were found: First, the limited number of input parameters does not allow a
functional modeling of the individual systems. Hence, systems are not satisfactorily represented
by regression analyses. Second, the regression analyses are mainly based on data of jet transport
aircraft from the 1st and/or 2nd generation. Major improvements in systems technology, such
as fly-by-wire control, are not mirrored in these regressions.

Hence, although simple statistical approaches are easily applicable in conceptual aircraft design,
they do not provide the required accuracies and sensitivities for sound technology assessment.
Furthermore, they do not allow for estimating energy off-takes, they only estimate systems
mass. Roskam [34, ch. 6] suggests to use known data of existing aircraft for estimating max-
imum bleed air off-takes, as well as maximum electrical and hydraulic power of the systems
architecture. Such data is, for example, published in Jane’s [99] at least for some aircraft.
However, characteristic off-takes over a flight mission are much lower. Hence, available data on
maximum off-takes cannot be used to estimate impacts on e.g. fuel efficiency.
16 Chapter 2. Related Work in Systems Modeling

2.1.2 Trimmed Statistics in Industrial Application

To keep the advantages of fast and easily applicable models in conceptual and preliminary
aircraft design, airframers use specially tailored statistical approaches for systems modeling.
Technology factors are implemented to account for evolutionary improvements in systems de-
sign. The models are also scaled and corrected by restricted data of the airframer’s fleet. The
Airbus FAME-S 3 code is an example for such a specially tailored approach in systems model-
ing [100, 101]. Accuracy of such tailored models is high [11, ch. 2], but only for a narrow product
line; namely for which they were tailor-made. High accuracy cannot be expected for arbitrary
aircraft designs. Since they are scaled with restricted data of the current airframer’s fleet, they
are not available for use by academia. Although evolutionary improvements can be accounted
for to some extent, they fail to account for innovative changes in systems architecture.

2.2 Medium and High Fidelity Tools

Since the mid-1990’s, computer aided engineering and design (CAE)4 tools for detailed modeling
of single systems have been developed. Such CAE tools forgo statistics and rely on physical
modeling instead. The approaches vary depending on the type of system that is investigated
and the available depth of information.

In an early approach, Scholz [102, 103] concentrates on modeling of the flight control and hy-
draulic power systems. He provides a tool for design and optimization on system-level. In the
proposed methodology, system components are sized by physical relations, e.g. by maximum
deflection rates and hinge moments. The entire system is virtually assembled from the com-
ponents. Koeppen and Carl [87, 104] use a similar approach for the FCS. In comparison to the
model by Scholz, they have reduced complexity for applicability in preliminary design. Calinski
et al. [105] have published a similar modeling approach for the electric system. For thermo-
dynamic systems, e.g. the ECS, methodologies for modeling and optimizing global heat fluxes
become more and more common. A comprehensive overview regarding exergy5 analysis and its
application is given by Tsatsaronis [106] as well as by Camberos and Moorhouse [107]. Exergy
analysis is a promising approach that allows for minimizing entropy, and thus for maximizing
efficiency. Figliola et al. [108, 109] and Vargas [110] have applied it first to design of ECSs.
Pellegrini et al. [111, 112, 113] published more recent results on aircraft systems analysis using
exergy. They show the theoretical advantage in energy efficiency of an electrical ECS over a
3
Fast and Advanced Mass Estimation - Systems
4
High-fidelity tools for detailed modeling, e.g. finite-elements methods (FEM), computational fluid dynam-
ics (CFD), or multi-body dynamics (MBD) simulations.
5
In thermodynamics, exergy is the energy that is available to be used for conversion to useful work.
2.3 Modeling of Entire Systems Architectures 17

conventional one by including the propulsion system into their analysis. Another interesting
approach in systems modeling are real-time simulations of system behavior including dynamic
and transient effects. Object-oriented modeling languages such as Matlab/Simulink and Mod-
elica are often used for such simulations. Krus [114], for example, has applied a Modelica model
for describing behavior of the primary FCS in dependency on the flight conditions. The focus
of such real-time simulations is rather on hydro or electro-mechanical systems [115]. Real-time
simulations are also used for modeling of the ECS [116, 117].

For re-sizing of components and systems to new functional requirements, the application of
scaling laws shows good results. Krus et al. [118] uses scaling laws for optimization of an existing
system architecture. Liscouet at al. [119] applied the principle to sizing of electro-mechanical
actuators; close agreement was achieved in validation. The principles of the application of
scaling laws in systems engineering are described by Pahl et al. [120, ch. 9]. Scaling laws
allow for fast re-sizing of existing components and are applicable already in preliminary design.
However, a reference system or component is always required as baseline.

The great advantage of these single system approaches is the high level of accuracy in estimat-
ing systems characteristics6 as well as system behavior in dependency on different operating
conditions. A detailed assessment on system-level is possible, but detailed knowledge about
subsystems and components is required. Application in early design is very limited, especially
when it comes to investigation of large design spaces.

2.3 Modeling of Entire Systems Architectures

The models that have been discussed in the previous sections, can only be applied to optimi-
zation on single system-level. For optimization of on aircraft-level, approaches for modeling of
entire systems architectures are required [36].

Optimization of aircraft systems architectures for minimum energy consumption was the focus
of the Power Optimized Aircraft (POA)7 research project [36]. Focus, however, was more on
system-level optimization of a pre-defined architecture; the design space was limited. An object-
oriented model library for Modelica was implemented as simulation framework for aircraft
systems architectures called Virtual Iron Bird [121]. Here, an inverse approach was used that
allows for sizing of power requirements from the functional requirements of the systems. Again,
required level-of-detail is high and the simulation framework relies on a predefined systems

6
For example, system mass, energy consumption or also the required installation space.
7
Funded by the European Union (EU) under the 5th Framework Program (2000-2006).
18 Chapter 2. Related Work in Systems Modeling

architecture. The More Open Electrical Technologies (MOET)8 directly followed POA [122].
MOET concentrated on the more-electrical aircraft concept alone and focus was even stronger
on system-level optimization. Nevertheless, integration of more-electrical systems into the entire
systems architecture was assessed. Van Driel et al. [123] present an optimization framework for
power systems using genetic algorithms. The aforementioned research relies on detailed system
models. Accuracy of results are high for the investigated systems architectures, but only for a
limited design space. Different systems architectures cannot be assessed with these approaches.
In the following, two methodologies for sizing of the entire systems architecture in preliminary
aircraft design are presented, which have been proposed by Koeppen and by Liscouet. Further,
a probabilistic approach that has been driven by Mavris is introduced. These models are the
closest to the one that is proposed by the author in this thesis.

Koeppen [87, 104, 11] presents a model for functional sizing of the entire systems architec-
ture. He concentrates more on hydro-mechanical systems, e.g. primary flight controls, than
on electrical or thermal systems. Nevertheless, a validated holistic model was developed that
uses the ATA-100 classification. Physical functions were established that allow for estimating
mass and mission dependent power consumption for each system, which accumulates to total
off-takes of the secondary power systems. Application of the derived physical functions requires
a high-level of input data, such as e.g. aerodynamic loads. The level of required input data
limits the application of this approach in conceptual design. Koeppen proposes to only con-
sider 80 % of the relevant system components for simplification of the model. This, however,
requires calibration of the overall model with empirical data of existing aircraft. For this pur-
pose, he uses scaling factors that are derived from current Airbus aircraft. This data is not
openly available and application of his model in academia is difficult. Also, the application of
his methodology to innovative systems is not possible, since data for calibration and scaling
is not available. Koeppen’s research is succeeded by Bornholdt et al. [124]. A more generic
approach towards modeling of the entire system architecture is targeted. They use functional
decomposition for distributing different aircraft-level functions within the system architecture.
It is currently targeted to improve accuracy of the model by integration of knowledge-based
meta models for modeling and sizing of the single systems. Their research is promising but also
in an early stage; no further details have been published yet.

A functional approach towards modeling of entire systems architectures is also proposed by


Liscouet [125, 126, 4, 127]. She avoids the ATA-based architecture and defines systems bound-
aries depending on the aircraft-level functions. Sizing of the single systems is derived from
performance requirements for fulfilling the specific aircraft-level functions. Power consumption
directly results from the aircraft-level functions. According to Liscouet, this approach allows
for application of the model to innovative and more integrated systems architectures. Although

8
Funded by the European Union (EU) under the 6th Framework Program (2006-2009).
2.4 Probabilistic Approaches in Systems Design 19

the methodology is based on simplified models that allow for integration into preliminary air-
craft design, it is not available to academia. Her model was developed in scope of an Airbus
internal research project; restricted data is used and models are not fully published.

Frischemeier [128] has published a first attempt towards assessment of integration of an inno-
vative system9 within the entire systems architecture. His approach requires a high-level of
detail and is only valid for a specific systems architecture. Further, snowball effects of systems
integration are not captured.

2.4 Probabilistic Approaches in Systems Design

Despite the application of mathematical models already in early design phases, it is inherent
that designers make certain decisions based on assumptions rather than on clear knowledge.
The need to make decisions based on assumption arises from uncertainties in the design itself
and future operating scenarios. Blanchard and Fabrycky [129] conclude that the lack of cer-
tainty makes decision making one of the most challenging tasks in design of complex systems.
Recently, probabilistic approaches were established to acknowledge the problem of uncertainty
and to support informed decision making in conceptual and preliminary design of aircraft and
system architectures. Such approaches are applicable to the full bandwidth of aerospace design
problems: from detailed single system design to holistic design and optimization of the entire
air transportation system. Bandte [130, ch. 1] identifies three aspects of design uncertainty
that can be supported by stochastic:

• Incomplete information about the operational environment: The future is hard to predict.
The value of a product, however, strongly depends on future operational environments.
For example, peaking fuel prices shift the optimum towards fuel efficiency rather than
to short travel times. Other aspects of the operational environment cover interactions of
the different systems within complex system architectures. Not all possible interactions
(wanted or unwanted) can be covered by analytical models, hence uncertainty remains.

• Availability of new technology: To cover the need for product improvements, new or
enhanced technology is introduced into the design. Questions always remain regarding
the availability of technology at the right time that provides the anticipated improvements.

• Uncertainty regarding the prediction accuracy of applied models: Engineering models are
not exact replicas of real physical situations [131]. Fidelity and prediction accuracy of
the various models varies and is often not known or transparent to the designer.
9
electro-mechanical actuation of the primary FCS
20 Chapter 2. Related Work in Systems Modeling

Mavris published numerous papers and supervises Ph.D. researchers on the topic of uncertainty
and application of probabilistic apporaches in aerospace system design (e.g. [88], [132], [130],
[133], [64], or [134]). His work covers all of the three aforementioned aspects; the full range of
his research cannot be cited in the scope of this literature survey. Amongst others, Economon
et al. [135] have applied stochastic for assessment of fuel burn trends of future aircraft that
arises from new technology and chances in payload-range capacity in a comprehensive design
study. Peoples and Willcox [136, 137] have applied a probabilistic approach for value-based
design optimization and risk assessment. West et al. [138] have applied stochastic expansion
for quantifying margins and uncertainties in the design of aerospace systems; they have targeted
application of their model to the design of a high speed transport aircraft. D’Ippolito et al. [139]
have applied a probabilistic methodology to the design process of composite materials. Smith
and Mahadevan [140] and Hanson et al. [141] have applied such methods to conceptual design
of launch vehicles.

When applying stochastic in aerospace system design, large design spaces need to be investi-
gated. For full coverage, each input and output parameter of the design process needs to be
varied and its influence on the overall design synthesis must be evaluated [130, ch. 1]. For large
and complex systems such a problem formulation cannot be handled. However, with today’s
computation power and down-selection of the governing influences, assessment of technology
based on probabilistic models allows for robust decision making while considering uncertainties.

Probabilistic approaches still require models for sizing and for estimating design characteris-
tics. For example, initial design solutions must be computed before parameters can be varied.
Most research in the field of application of probabilistic approaches in systems design does not
concentrate on the underlying models; integration of experts opinion and respond surfaces is
suggested for system modeling. Still the probabilistic approaches are independent from the
underlying design models. Hence, more sophisticated design approaches and methods can be
applied. However, computation time remains as limiting factor. Fast and robust models are
required for investigation of the large stochastic design spaces.

2.5 The Need for a New Methodology

Figure 2.2 shows allocation of the beforehand discussed approaches towards systems modeling
to the different design phases. It has to be kept in mind that boundaries between the different
design phases are soft, and application of the single models also in other design phases is not
categorically ruled out. Semi-empirical approaches (cf. ch. 2.1) are applied in very early design,
e.g. during concept definition. High fidelity models (cf. ch. 2.2) are not applied before late
preliminary or in detailed design. Hence, these two approaches frame the spectrum of models
2.5 The Need for a New Methodology 21

that have been depicted within this literature survey, see figure 2.2. Probabilistic approaches
can be combined with most of the provided models, and are thus applicable over the entire
spectrum. However, the most benefits can been drawn from probabilistic approaches in early
design phases; more and more uncertainties are eliminated while the design matures.

Also Koeppen [11] and Liscouet [4] approaches


to modeling of entire systems architectures com-

  
  
  


 

 
prise some disadvantages for application in con-

   


ceptual design and for integration into overall
aircraft design synthesis. The methodology pro-


posed by Koeppen relies on quite detailed sys-
   

tem descriptions. Detailed functional relations


   

  


are required for each system and component
   
that are not always available in early design.
!  

Further, scaling with current fleet data is re-


quired for application of his model. Considering  
 
the level-of-detail of the required input parame-
Figure 2.2: Allocation of available models to
ter as well as its limited flexibility to investigate
the design phases.
large design spaces, it is rather placed in prelim-
inary than in conceptual design, see figure 2.2.
Although the methodology proposed by Liscouet stays closer to early preliminary design, she
uses models that are based on restricted data, and that are not fully published. Hence, ap-
plication in academia is not possible. Further, both models only allow for integration of new
systems, if detailed sizing models for these exists. This is hardly the case when it comes to
innovative technology.

Yet, no aircraft system model is available that allows for sizing of the entire systems architecture
of an aircraft, and that is fully applicable in conceptual aircraft design. Thus, a new method-
ology is proposed by the author that closes the existing gap. This new approach concentrates
on modularity and flexibility, so that large design spaces on aircraft-level can be investigated.
The proposed methodology uses low to medium fidelity models for modeling of conventional
aircraft systems within the entire systems architecture. Hereby, the proposed methodology
intentionally uses single system models that are independent from each other as well as ex-
changeable for providing maximum modularity and flexibility. The approach towards modeling
of conventional systems architectures is further combined with a probabilistic approach, which
allows for integration of innovative system technologies. The probabilistic approach integrates
the remaining design uncertainties of innovative technology into overall design synthesis, and
makes the impacts of the design uncertainties transparent on overall aircraft-level. The pro-
posed methodology has been implemented in a systems model for application in the design of
22 Chapter 2. Related Work in Systems Modeling

commercial transport aircraft. Sound technology assessment on overall aircraft-level becomes


possible with the proposed systems model, if the model is integrated into an overall aircraft
design environment. It allows for capturing repercussions of systems integration also including
snowball effects of overall design synthesis. In the scope of this thesis, integration of the pro-
posed systems model into the ILR’s MICADO design environment is discussed. However, the
proposed systems model is as stand-alone tool, that could also be integrated in any other air-
craft design environment as long as the correct set of input parameters is provided to the model.
Thus, the proposed systems model closes the gap for sound systems technology assessment on
overall aircraft-level in conceptual design synthesis, see figure 2.2.
3 Aircraft Systems Model

The author’s approach towards systems modeling is described in this chapter. First, the de-
rived methodology is introduced in chapter 3.1. Dependencies of the approach as well as the
generic system architecture on which it is based are outlined. The author elaborates on the
importance of energy balancing. In chapter 3.2, the implemented systems model is described.
Three of the underlying mathematical models for systems modeling are described in detail.
Chapter 3.3 concludes with implementation of the proposed systems model into a software tool
for integration into MICADO. The author’s approach towards modeling of innovative systems
is illustrated in chapter 3.4; a strategy that incorporates design uncertainties is introduced.

3.1 Proposed Methodology for System Modeling

Allocation of different aircraft-level functions to the single ATA chapters is sufficient for de-
scription of the conventional system architecture. However, innovative and highly-integrated
systems require a multi-ATA approach, in which several aircraft-level functions are provided
by an highly integrated top-level system [36]. Hence, the classical ATA breakdown must be
widened for definition of systems boundaries of such innovative architectures. Therefore, the
proposed methodology concentrates on function-based modeling that is both valid for conven-
tional and innovative (more integrated) system architectures. Systems boundaries that are
defined by this functional approach still correspond to the classical ATA breakdown for conven-
tional architectures. Thus, ATA nomenclature is still used for description of the conventional
systems. The modularity of the proposed model, however, allows for fast adaption of the model
to highly integrated and accordingly to multi-ATA systems.

3.1.1 Aircraft-Level Functions and Systems Dependencies

The overall structure of the function-based systems domains is shown in figure 3.1. Each system
provides a specific physical function for which it requires a certain amount and form of energy.
The consumer systems use energy to perform specific aircraft-level functions. Actuators of the

23
24 Chapter 3. Aircraft Systems Model

FCS, for example, use hydraulic energy to perform mechanical work, which leads to deflection
of control surfaces, and thus provides flight control functions. The required energy for perform-
ing aircraft-level functions is provided by the secondary power systems. They extract power
from the primary power systems, convert it into the specific energy form1 , and distribute the
energy through the entire systems architecture. The primary power systems use fuel to propel
the aircraft2 ; a fraction of the primary energy is used to power the secondary power systems.
Definition and sizing of the overall systems architecture and the single systems is realized in a
top-down approach as illustrated in figure 3.1 (left-hand side). From functional requirements
of each aircraft-level function, the energy requirements for each consumer system are derived.
These energy requirements size the single consumer systems in terms of components and/or
system mass. Accumulated power requirements of the consumer systems size the secondary
power systems. Again, accumulated power requirements of the secondary power systems deter-
mine total power off-takes from the primary power systems. Re-sizing of the propulsion system
is not directly implemented in the presented systems model. It is covered by integration of
the systems model into MICADO. Systems are not only influenced by aircraft-level functions;
other dependencies and constraints are discussed in the following paragraphs.



           

 
  

 

  



    

    



      









   
 

Figure 3.1: Function based systems domains of conventional systems architectures.

1
Electric, hydraulic or pneumatic
2
With the exception of the auxiliary power unit and other emergency systems, which only provide the secondary
power systems with energy in failure cases or during ground operations.
3.1 Proposed Methodology for System Modeling 25

Time Dependency of Systems Operations

Performance requirements of each aircraft-level function change during the different phases of
flight; power requirements for each aircraft system change accordingly. While wing anti-icing
consumes a noticeable fraction of bleed-air during take-off and landing, it is switched off during
cruise and the ECS remains the only major consumer of bleed air. Bleed air requirements of
the ECS also change with the different phases of flight. Other examples for time dependency
are peaks in hydraulic power requirements during deployment or retraction of the landing gear
and high-lift devices. Power requirements for each system can be expressed as a function of
altitude (h), Mach number (M ), ISA condition, aircraft configuration(Chigh lif t & gear )3 and cabin
utilization (UP AX ). Hence, power requirements become a function of mission time (tmission ),
cabin utilization and ISA condition for a given mission profile4 :

Psys,req = f (h, M, ISA, Chigh lif t & gear , UP AX ) = f (tmission , ISA, UP AX ) . (3.1)

Hydraulic power requirements of an Airbus A320-200 aircraft that were obtained from flight
test data5 are plotted against estimated power requirements in figure 3.2. It is easy to observe
that dynamic and transient effects are neglected by the proposed model. Instead, a constant
base load is estimated for each system and for each flight phase. In addition, peak loads that
occur at specific time frames are superposed to the constant base loads. From figure 3.2 peak
loads that occur from retraction of the landing gear and the high lift system can be observed.
Total required system power (Psys,req ) for each flight increment is expressed as the sum of a
base load (Psys,reqbaseload ) and a peak load (Psys,reqpeakload ):

Psys,req (tmission , ISA, UP AX ) = Psys,reqbaseload + Psys,reqpeakload . (3.2)

Dependency on Sizing and Operational Mode

The proposed methodology differs between two modes: the sizing mode and the operational
mode. The systems architecture has to be sized before its operational characteristics can be
assessed. In the sizing mode, each system is sized by its design performance requirements6 and
typical safety factors that are used in aviation. Results of the system sizing are, for example,
number of required components, actuator size, or maximum power consumption. Thus, systems
mass as well as maximum performance capabilities are defined for each system within the sizing
mode. The sizing mode is only used within overall aircraft design synthesis and optimization as
3
Clean configuration during cruise, or deployed landing gear and/or high-lift devices for take-off and landing.
4
Defined flight path with altitude and speed profile as well as defined mission payload.
5
Data from personal correspondence.
6
For sizing of the systems architecture the aircraft’s design mission is used.
26 Chapter 3. Aircraft Systems Model

shown in figure 3.3.A detailed description of the overall preliminary aircraft design environment
and integration of the proposed model follows in chapter 4.
Normalized Hydraulic Power

Flight test data: required hydraulic power


0.8 20
Estimated data: required hydraulic power
Retraction of L/G

Altitude, 1000 ft
Flight altitude
0.6 15
Retraction of flaps/slats
0.4 10

0.2 5

0.0 0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Normalized Time

Figure 3.2: Hydraulic power requirements for take-off and climb of an Airbus A320.


 






 

 
   
 
  

 
 
 

 
  

 
 
 
 

   

 
 

  
 


   
 
     
 
 

 
   
   
   
 

Figure 3.3: Sizing mode and operational mode of the proposed systems model.

The operational mode is used for estimating operational system power off-takes during the
aircraft’s mission as discussed in the previous section. Thus, the required input is the specific
mission profile as specified by equation 3.1 and 3.2. The operational mode is used in overall
3.1 Proposed Methodology for System Modeling 27

design synthesis and optimization as well as for mission analysis for various missions of the
converged aircraft design, see figure 3.3. Assessment of the aircraft performance is based on
different study missions7 , see figure 3.3 (lower part). Each study mission analysis requires
a preceded calculation of the corresponding mission power off-takes by the proposed systems
model in the operational mode.

Power requirements do not only differ for variations in mission profile, but also for the various
system states such as normal flight or failure conditions. In case of system failures, non-essential
consumers are isolated to guarantee sufficient energy supply to essential systems. Also, essen-
tial systems may operate in a degraded mode with performance constraints. Different system
states and failure conditions are described in more detail by e.g. Moir [32, ch. 4-5]. The oper-
ational mode of the proposed model is currently implemented for assessing normal flight only.
Possible failure conditions are numerous and are neglected in conceptual design. The devel-
oped methodology, however, allows for expanding the design space by a third axis for different
operational modes to also cover failure conditions in the operational mode, see figure 3.4. En-
hancing the proposed model would allow for future applications to consider and assess off-design
performance characteristics of different systems architectures during operations.



   

     




    
 




  

Figure 3.4: Power requirements in dependency on system states [4].

Dependency on Installation Location

System design is also driven by installation location of the different systems. Installation
location influences sizing and energy efficiency of the secondary power systems, overall center of
gravity of the aircraft, and installation space. Further, system interfaces with other components,
with the airframe structure, and with human operators are impacted.

The bleed-air ducts as estimated by the proposed systems model for a standard single-aisle
aircraft are shown in figure 3.5. The available energy sources (engines and APU) are connected
with all consumers (ECS and wing anti-icing). The required length of conduction directly
influences systems mass as well as power losses within the secondary power systems, and thus
7
Typical mission that are flown in daily operations by the aircraft.
28 Chapter 3. Aircraft Systems Model

mass and efficiency are impacted by installation location of the energy sources and consumers.
Re-allocation of systems accumulates to a shift in the overall center of gravity of the aircraft.
Hence, wing and empennage sizing are directly impacted by systems allocation. The proposed
methodology includes calculation of the center of gravity of each system.

In the detailed design phase, installation location is driven by the available space. However,
detailed information regarding available space as well as required space must be available, which
is not the case in conceptual design. Hence, installation space is neglected for simplification of
the proposed methodology. Also installation location, and thus spatial layout of the systems
architecture is only considered in a two-dimensional coordinate system, see figure 3.5.


  


 
 

 

 





Figure 3.5: Installation locations of standard bleed air system and its consumers (Air-
bus A320 top-view taken from reference [5]).

System Interdependencies and Heat Loads

Possible interdependencies of single systems with the aircraft and/or other systems play an
important role for system design. The interdependencies with the aircraft are accounted for
by integration of the proposed systems model into overall aircraft design synthesis. Model-
ing of interdependencies between the different systems is covered by definition of the systems
interfaces. As defined in chapter 1, only energy, information, and/or mass fluxes can cross
system boundaries, and thus can interfere with other systems. Hence, all interfaces between
the different systems must be either based on energy, information, or mass fluxes. All defined
interfaces are purely function based, and thus can be described by physical function models.
3.1 Proposed Methodology for System Modeling 29

Further, unintended interferences that e.g. originate from systems failures8 are neglected.

Information fluxes and the according data interfaces are especially important for the different
avionic systems. Today, digital data buses are used in civil aviation for sharing information
between the different systems [33, ch. 2]. Since the impact of information fluxes on overall
system mass and energy consumption is negligible small, data interfaces are not further regarded
for in the proposed methodology. Instead, it concentrates on interdependencies within the
system architecture that are characterized by either mass or energy fluxes. One example is the
fuel flow from the fuel system to the single engines. Another example are power fluxes from
the secondary power systems to the single consumer systems.

Besides the required power fluxes, parasite heat loads are important especially for sizing and
operation of the ECS. Not only passengers emit heat loads inside the cabin, but also the various
aircraft systems. Thus, a parasite heat flux is assigned to each system in addition to the required
power flux. These heat loads directly correlate to the power requirements of the single systems.
For electrical systems, the emitted heat loads equal the energy that is lost within the system [32,
ch. 5], and thus can be estimated by an efficiency factor of the specific system. As identified by
e.g. Faleiro [36] or Rosero et al. [66], emitted heat loads of power electronics are one remaining
and not yet solved issue of the all electrical aircraft concept.

3.1.2 Generic Systems Description

A generic modular systems description is derived from the different systems domains, depen-
dencies, and constraints that were described beforehand. This generic systems description is
used throughout the proposed methodology and defines characteristics that each implemented
system model must have. The characteristics can be grouped as follows: (1) mass and center of
gravity, (2) power, and (3) heat load. The schematics of the generic systems module are shown
in figure 3.6: A mass (msys ) is allocated to each system. Mass distribution is neglected within
the model, and thus a point mass sufficiently describes mass characteristics of each system.
Additionally, position of center of gravity in longitudinal and span direction (CGx and CGy ) is
required. The required power for the system function (Psys ) is derived from power requirements
of the specific aircraft-level functions (Pf unc ). Power requirements are not constant but depend
on the aircraft mission as shown in equation 3.1 and 3.2. Mission power (Psys,mission ) require-
ments are further divided into base loads, that are constant for each mission segment, and into
time dependent peak loads (cf. ch. 3.1.1). Additionally to the mission power requirements,
design or maximum power (Psys,max ) is required for sizing of the systems, see figure 3.6. Each
system emits a parasite heat load (Qsys ) due to energy losses within the system. Such losses are,
8
For example, electro-magnetic interferences or fluid damage to electrical components.
30 Chapter 3. Aircraft Systems Model

for example, caused by mechanical friction or by electronic resistance. Hence, heat loads are
additionally assigned to each system. These heat loads are proportional to the required power
for providing the specific functions. Further, the model differs between maximum (Qsys,max )
and operational heat loads (Qsys,mission ).

For description of the overall systems architecture (puzzle) that is composed from single air-
craft systems (building blocks), a generic systems architecture is used. This generic systems
architecture is illustrated in figure 3.7. This example comprises a secondary power system,
an ECS, and two arbitrary consumers. Definition of system characteristics via the interfaces
allows for modular and flexible composition of different architectures. In the chosen example,
the secondary power system provides the required power to both consumer systems as well as
to the ECS. The consumer systems use that energy to perform their dedicated aircraft-level
functions. The ECS balances the heat loads of the three other systems within that architecture.
This simple example shows the importance of energy balance for the proposed methodology,
which is further described in the following section.

  

   
 
  


 
 

   
    
 
 
 
  

Figure 3.6: Schematics of the generic systems module.

 

   

  
 
  

 

  





Figure 3.7: Generic systems architecture.


3.1 Proposed Methodology for System Modeling 31

3.1.3 Energy Balance within the Systems Architecture

The first law of thermodynamics states the principal of conservation of energy. It is applied to
the overall systems architecture for balancing power and heat fluxes, which leads to the overall
required power off-takes from the engines for each flight increment. Conservation of energy is
also true for each system. According to figure 3.6, the system power requirement (Psys ) must
equal the sum of the function power (Pf unc ) and the emitted heat load (Qsys ):

Psys = Pf unc + Qsys . (3.3)

It can be seen from equation 3.3 that a strict separation of power and heat fluxes is not
possible; both parameters are linked. Nevertheless, the proposed methodology uses two different
envelopes for energy balancing: (1) the entire aircraft is used as envelope for deriving engine
off-takes and system power requirements, and (2) the fuselage is used for deriving ECS power
requirements from the corresponding heat fluxes. All power and heat fluxes are required for
deriving the overall engine power off-takes and system power requirements. Hence, the aircraft
is used as envelope for energy conservation. On the contrary, parasite heat fluxes of systems
that are located outside the fuselage, e.g. control surface actuators or wing anti-icing, are
emitted to the outer atmosphere, and do not impact heat balance of the fuselage and cabin.
Hence, for sizing of the ECS, the fuselage is used as envelope.

      


  
   
    

 
    

  

    

       

Figure 3.8: Energy balance and systems boundaries.

(1) Application of Aircraft Envelope for Energy Balance

Engine power off-takes (Peng,of f ) accumulate from power requirements of the secondary power
systems (Psys,sec power ). With equation 3.3, engine off-takes can also be linked to the function
power requirements (Pf uncsec power ) and the emitted heat loads (Qsyssec power ) of the secondary
32 Chapter 3. Aircraft Systems Model

power systems as follows:


X X
Peng,of f = Psyssec power = (Pf uncsec power + Qsyssec power ) . (3.4)

For the conventional systems architecture, engine power off-takes can be divided into shaft
(Pshaf t,of f ) and bleed air (Pbleed,of f ) off-takes, which are defined by the corresponding secondary
power systems:
Pshaf t,of f = PsysAT A24 + PsysAT A29 = Pf uncAT A24 + QsysAT A24 + Pf uncAT A29 + QsysAT A29 ,
(3.5)
Pbleed,of f = PsysAT A36 = Pf uncAT A36 + QsysAT A36 .
Power requirements of each secondary power system are defined by the sum of the system power
requirements of the consumer systems:
X X
Pf uncsec power = Psysconsumer = (Pf uncconsumer + Qsysconsumer ) . (3.6)

For the conventional system architecture, the system power requirements of the three secondary
power systems can be derived from the specific consumers such as shown in equation 3.7 to 3.9.
Pf uncAT A24 = (PsysAT A21 + PsysAT A22 + PsysAT A23 + PsysAT A25
+ PsysAT A28 + PsysAT A30 + PsysAT A31 + PsysAT A32 (3.7)
+ PsysAT A33 + PsysAT A34 )electric
Pf uncAT A29 = (PsysAT A27 + PsysAT A32 )hyraulic (3.8)
Pf uncAT A36 = (PsysAT A21 + PsysAT A30 )bleed air (3.9)

Some systems use more than one form of energy. For example, the conventional ECS (ATA-21)
uses bleed air for cabin pressurization and for climate control. Additionally, it requires electric
energy for ventilation. Another example is the ice and rain protection system (ATA-30). It
uses bleed air for wing and engine inlet anti-icing, and electric energy for anti-icing of ducts,
valves, sensors, and antennas. Hence, some systems are included in more than one of the before
defined equations.

This energy balance approach is used in a reverse fashion within the proposed methodology.
Functional power requirements and heat loads of the different consumer systems are derived
from single system models and dedicated aircraft-level functions as illustrated in figure 3.1.
If power requirements of the different consumer systems are known (c.f. eq. 3.7-3.9), power
requirements of the secondary power systems can be accumulated first (c.f. eq. 3.6), which than
lead to overall engine power off-takes (c.f. eq. 3.4).

(2) Application of Fuselage Envelope for Energy Balance

The ECS has two main aircraft-level functions: (1) climate control of the fuselage and cabin, and
(2) pressurization of the cabin. The functional power requirements for the first one are driven by
3.1 Proposed Methodology for System Modeling 33

P
heat fluxes of systems ( Qsysf uselage ), passengers (QP AX ), and environmental conditions (Qsun ,
Qconvection , and Qlosses ), see figure 3.8. When using the fuselage as system boundary, the heat
flux balance can be written as follows:
X
Pf uncAT A21,climate = Qsun + Qconvection + QP AX + Qsysf uselage − Qloss . (3.10)

For the conventional system architecture of today’s transport aircraft, the following parasite
system heat fluxes are considered in the proposed model for determining the functional power
requirements for climate control of the ECS:
P
Qsysf uselage = QsysAT A21 + QsysAT A22 + QsysAT A23 + QsysAT A24
+ QsysAT A25 + QsysAT A28 + QsysAT A29 + QsysAT A31 (3.11)
+ QsysAT A33 + QsysAT A34 .

3.1.4 Model Architecture Down to System-Level and Below

The aircraft’s overall systems architecture is composed of an arbitrary number of single sys-
tems or building blocks (cf. fig. 3.6). The interdependencies and repercussions of the single
systems are accounted for by the defined interfaces of each system model (cf. ch. 3.1.1) within
the overall methodology. As aforementioned, the interfaces comprise mass estimation as well as
power requirements and emitted heat loads. For example, integration of power electronics into
the fuselage does not only lead to an increased overall system weight, but also to additional
heat loads within the fuselage that must be compensated by the ECS. Within the proposed
methodology, the ECS is then automatically sized according to its new performance require-
ments, which would directly impact sizing of the bleed air system and bleed air off-takes from
the engines. Repercussions of systems integration are considered by means of accumulating
single system weights and by energy balancing. Hence, this modular approach allows for easy
integration of a new system into an existing system architecture by simple adding a new system
or building block to the existing overall model.

For each specific system or building block, mathematical models are required that provide the
required data and systems characteristics. System models that provide the systems characte-
ristics on system-level are sufficient for the proposed methodology; information on component
characteristics of single systems are not required. However, for certain systems and/or certain
design studies it may be required or desirable to model single systems down to component-
level9 . As shown in figure 3.9, it is possible with the proposed modeling approach to choose
different model resolutions for different systems within the overall architecture as long as the
systems characteristics are provided for each system. Also the type of mathematical model
9
For example, actuator-level within the flight control system (ATA-27) for integration of EMAs or EHAs.
34 Chapter 3. Aircraft Systems Model

can vary. In the implemented version of the proposed systems model simple semi-empirical as
well as detailed functional based models are implemented coexistently for the different systems.
Some will be presented in chapter 3.2 in more detail.

 
  
    
 
 
 
  
  
   
   
 

 

 
  
  
  
  
 

    

 







   

Figure 3.9: Model architecture down to system-level and below.

3.2 Proposed Systems Model

The proposed systems model features different type of models: functional based physical mod-
els, semi-empirical models, as well as statistical ones are used throughout the systems architec-
ture. Since all implemented models cannot be presented in the scope of this thesis, the author
first concentrates on main design parameters of the conventional system architecture. Then,
the underlying models of three selected systems are presented as an example. The author chose
to use two system models for elaborating on the approach of deriving power requirements from
functional performance requirements of the system (see ch. 3.2.2 and 3.2.3). For the third
system model example, focus is put on deriving overall system weight from sizing of single
components, which again results from functional performance requirements (see ch. 3.2.4).

3.2.1 Main Design Parameters

The main design parameters that are used for estimating system mass and mission dependent
power consumption are listed below. Table 3.1 shows design parameters for the conductor
systems and energy sources10 , table 3.2 for the single consumer systems. Type and number
10
The fuel flow is no input parameter for the proposed system model since fuel consumption of the engine and
the APU is captured within the MICADO environment
3.2 Proposed Systems Model 35

of design parameters indicate the maturity of the implemented model for conceptual design
of the aircraft systems architecture. Special emphasis in systems modeling have been put
on systems that have significant influence on overall aircraft design synthesis, e.g. by means
of mission dependent power consumptions or system mass. For those systems sophisticated
models were implemented that rely on physical relations. For example, the ECS uses heat
balancing to estimate the required cooling or heating power for each flight phase, for each
specific environmental condition, and for each cabin utilization (cf. ch. 3.2.2). The FCS uses a
model for estimating hinge moments, and thus required actuation forces as well as an validated
NASA model [142] for estimating system mass (cf. ch. 3.2.4). It will be shown that the
proposed model mirrors the impacts of top-level aircraft design parameters and shows a non-
linear behavior, although it is intended for application within conceptual design. The overall
behavior of the proposed systems model for variations in different design parameters will be
discussed in greater detail in scope of the sensitivity studies that are presented in chapter 5.2.
The following sections will concentrate on single system models that are implemented within
the overall system model.

Design parameters for estimation of ...


System
system mass power consumption
ATA24 – Electric lcable , m̂cable , Pmax,elec , Psys,elec (t), ηAT A24
A/C geometry
ATA29 – Hydraulic OW E, lducts , Pmax,hydr , Psys,hydr (t), ηAT A29 ,
m̂ducts , m̂f luid , m̂pump , npumps,elec
A/C geometry
ATA36 – Bleed air lduct , m̂duct , ṁmax,bleed , ṁsys,bleed (t), ηAT A36 ,
A/C geometry Tbleed , Tcabin , cp
ATA49 – APU Pmax,elec , Pmax,hydr , ṁmax,bleed ,
felec,AP U , fhydr,AP U , fbleed,AP U , —
finstall,AP U
ATA70 – Propulsion — Pelec (t), Phydr (t), ṁbleed (t)

Table 3.1: Overview of main design parameters for conductor systems and energy sources.
36 Chapter 3. Aircraft Systems Model

Design parameters for estimation of ...


System
system mass power consumption
ATA21 – ECS P AX, ncrew,cabin+f light , P AX, ncrew,cabin+f light ,
Pmax,AT A21 , Sf uselage , nwindows , Swindow ,
OW E, Rdesign , Tbleed , Tcabin , cp , ηAT A21 ,
lcabin , neng , frecirculation , ṁP AX,min ,
f uselage geometry λf uselage , qsun , qhuman ,
Qsys,heat (t)
ATA22 – Auto flight OW E, Rdesign NM CU , PM CU
ATA23 – Communication OW E, Rdesign NM CU , PM CU
ATA25 – Furnishing M T OW , Rdesign , pdif f , P AX, Pbeverage (t), Poven (t),
P AX, ncrew,cabin+f light , Pf ridge (t), PIF E (t)
mmax,payload , Scabin
ATA26 – Fire protection M T OW —
ATA27 – Flight controls M T OW , V M O, M M O, M a(t), H(t), Tambient (t),
γcontr. surf. , γ̇contr. surf. , M T OW , V M O, M M O,
wing geometry, f in geometry, γcontr. surf. , γ̇contr. surf. ,
stabilizer geometry, wing geometry, f in geometry,
control surf. geometries, stabilizer geometry,
Pmax,actuators contr. surf. geometry,
ATA28 – Fuel MF W M T OW
ATA30 – Ice protection OW E, Pmax,AT A30 M a(t), H(t), Tambient (t),
LW C(t), λwing , cp ,
Tbleed , ηAT A30 ,
nengines , wing geometry
ATA31 – Instrumentation OW E, Rdesign NM CU , PM CU
ATA32 – Landing gear M LW mL/G , lstrut , tretraction
ATA33 – Lights M T OW lcabin , P AX
ATA34 – Navigation OW E, Rdesign NM CU , PM CU
ATA35 – Oxygen P AX, Rdesign —
ATA38 – Water & waste P AX, Rdesign —

Table 3.2: Overview of main design parameters for consumer systems.


3.2 Proposed Systems Model 37

3.2.2 Environmental Control System – ATA-21

The environmental control system pressurizes the aircraft and controls the cabine climate.
Cabin pressure is controlled by balancing of air outflow and ECS air inflow [32, ch. 7]. Further,
leakage of the fuselage must be considered. Regulations ([9, 25.841], [89, §25.841]) require cabin
pressure altitude not to exceed 8000 ft under normal operating conditions. Climate control
concentrates on temperature, contamination, humidity, and ozone control of the environment
within the pressurized areas of the aircraft. The temperature is controlled by means of chilled
or heated air that is ducted into the cockpit, cabin, cargo compartment, and avionic bays.

System Description

Main components of the ECS are the air conditioning packs (ACP), which provide the tem-
perature controlled and pressurized air. Different ACP designs are used in aviation. The most
common one is the bootstrap air cycle machine, which uses a turbine-compressor refrigerator
unit to cool the air [143]. Pressurized air passes through a turbine that drives the compressors.
Work is performed by the air flow, which results in reduction of temperature and pressure.
Such ACPs have conventionally been used in commercial transport aircraft together with hot
bleed air for the last decades. The trend towards bleedless systems architecture require electric
ACPs. The Boeing 787 uses such a design in which ram air is cooled by an electric driven
compressor [70]. For minimizing the required fresh air flow, current research suggests to use
a second cooling cycle, which provides cooling of the recirculated air in electric ECS architec-
tures [144]. For assessment of the two technologies, it has to be kept in mind that additional
drag due to the ram air inflow has to be considered for the electric ACP. Other components of
the ECS are mixer units, ducts, valves, and vents.

Modeling of System Power Consumption

The proposed model has to provide the required cooling or heating power for climate con-
trol, and the required air inflow for pressurization of the aircraft in different flight phases and
atmospheric conditions, as well as for different cabin utilization. Depending on the system
technology11 , the model either estimates the required bleed air mass flow, or the required elec-
tric power and ram air mass flow. Additionally to the ACP power consumptions, the required
electric power for ventilation is derived.

11
Either conventional bleed air ACPs or electric ACPs.
38 Chapter 3. Aircraft Systems Model

Literature provides several detailed models for simulation of environmental conditions inside
the cabin (e.g. [145], [146], [147], [148], [149], or [150]). Such detailed models assume dynamic
temperature gradients and also concentrate on detailed temperature distribution inside the
cabin. They are based on instationary CFD models and are required for assuring passenger
comfort in the scope of detailed cabin design. Application to conceptual aircraft design, how-
ever, is not feasible. Other models use exergy analysis not only to model the ECS but also to
asses efficiency on overall aircraft-level. Such exergy analyses include both the ECS and the
propulsion system (e.g. [151] and [110]). For the proposed systems model it is not required
to incorporate exergy. The proposed systems model is integrated into overall aircraft design
synthesis, which accounts for impacts of engine off-takes with a full thermodynamic engine
model (cf. ch. 4.1.2). Thus, impact of the ECS on engine efficiency is sufficiently assessed.

For sizing of the ESC it is sufficient to assume steady cabin temperature and pressure lev-
els [143]. Equation 3.10 provides the energy balance around the fuselage. It is rewritten for
considering the ECS functional power requirements (Pf uncclimate,AT A21 ), heat losses (Qloss ), and
heat sources (Qsources ) as follows:
dEf uselage
= Pf uncclimate,AT A21 − Qloss + Qsources = 0 . (3.12)
dt
For constant cabin temperature, the functional power that is provided by the ECS must balance
heat losses and heat sources. It can be written as the product of ECS mass flow (ṁAT A21 ), ECS
inflow temperature (TAT A21 ), and the specific heat capacity of air (cp ):

Pf uncclimate,AT A21 = ṁAT A21 · cp · TAT A21 . (3.13)

Hence, the air flow provided by the ECS can be derived from equation 3.13 for defined ECS
functional power requirements, which are derived from the different heat losses and heat sources
by the proposed model. The heat flux that is lost due to outflow and leakage is given by equa-
tion 3.14. The additional heat loads are broken down to single sources as shown in equation 3.15.

Qloss = (ṁoutf low + ṁleakage ) · cp · Tcabin (3.14)


X
Qsources = Qsun + Qconduction + QP AX + Qsys,f uselage (3.15)

Parasite heat loads that are emitted by various aircraft systems have been defined in equa-
tion 3.11. Sun radiation that enters through the windows provides a heat flux (Qsun ), which is
estimated from the specific heat load of sun radiation (qsun ), of the cabin window area (Awindow ),
and of the number of cabin windows (Nwindow ):

Qsun = qsun · Awindow · Nwindow . (3.16)

Conduction (Qconduction ) depends on the temperature difference between the cabin (Tcabin ) and
the outer fuselage skin (Tskin ), the fuselage’s wetted surface (Af uselage ), and the fuselage’s
3.2 Proposed Systems Model 39

heat transfer coefficient (hf uselage ). Skin temperature is derived from the free stream Mach
number (M a) and the ambient temperature (Tambient ) for boundary layer flow [152]. Heat
conduction is estimated from the following equation:

Qconduction = hf uselage · Af uselage · (Tskin − Tcabin ) with Tskin = Tambient · (1 + 0.18 · M a) . (3.17)

The biological heat load (QP AX ) depends on utilization of the cabin; a constant average heat
load per person (Qhuman ) is assumed [143]:

QP AX = Qhuman · (nP AX + nf light crew + ncabin crew ) . (3.18)

Equation 3.13 and 3.14 comprise mass fluxes. Additional to the aforementioned heat flux bal-
ance, a mass balance is used to set the relationship between the different mass fluxes. Figure 3.10
illustrates the cabin air flow of a common ECS. Only the cockpit is continuously supplied with
fresh air from the ACPs solely. A fraction of the cabin air is filtered and mixed with fresh
ACP air before it is recirculated into the cabin. Recirculation of cabin air reduces the required
fresh air flow significantly, and thus increases fuel efficiency of the aircraft. Typically, 40 % to
50 % of cabin air is recirculated within the mixing unit. Figure 3.10 shows typical proportional
mass flows of the ECS [6]. Within the proposed model, those mass flow fractions can be easily
adapted for capturing changes in the ECS architecture.

 
          
 
   

   
    
  
   
  
   
  

Figure 3.10: Mass balance of cabin air flow (100 % cabin air inflow) [6].

The mass balance around the fuselage states that inflows must equal outflows:

ṁAT A21 = ṁoutf low + ṁleakage . (3.19)

Hence, equation 3.12 is rewritten and solved for the required ECS fresh mass inflow:
P
Qsun + Qconduction + QP AX + Qsys,f uselage
ṁAT A21 = − (Tskin − Tcabin ) . (3.20)
cp

Additionally to the ACPs, ventilation of cabin air requires energy. Usually electric ventilators
are installed. Total required power accumulates from the single ventilators within the ECS.
40 Chapter 3. Aircraft Systems Model

Required power for a standard ventilator is given by equation 3.21 ([153, ch. 6.2], [11, ch. 4.1]).
If a mean pressure difference and mean air temperature is assumed for all ventilators within
the ECS, equation 3.21 can be simplified for estimating the total required power for ventilation
as shown in equation 3.22. No additional power is required for the outflow of cabin air due to
the negative pressure difference.
"  γ−1 #
X cp · Tair pvent,outi γ
i
Pvent = · − 1 · ṁventi (3.21)
i
ηventi pvent,ini
"  γ−1 #
cp · Tcabin pvent,out γ
⇒ Pvent = · − 1 · (ṁcabin + ṁf light deck ) (3.22)
ηvent pvent,in

Sensitivity and Plausibility of the Model

The complex relation between required ECS


power and aircraft mission will be discussed in 
  !


 
chapter 5.1 and 5.2. This section concentrates 

on the dependency of the required ACP po- 
wer for cooling or heating of the aircraft cabin 
on ambient temperature and cabin utilization,   
see figure 3.11. Cooling of the cabin is re-   
  
quired for negative ACP powers, and heating    
    
   

for positive ones accordingly. Equations 3.12       

to 3.15 show that the required power for cool-
ing or heating is a linear function of the single Figure 3.11: Required ACP power in depen-
heat loads. Thus, it is not surprising that fig- dency on ambient temperature and
ure 3.11 also reflects this linear relation. The cabin utilization.
results that are shown are based on a 150 pas-
sengers Airbus A320-like single-aisle aircraft design (cf. ch. 4.3.1).

The ACP power is provided by a fresh air flow with a defined temperature. Figure 3.12 shows
both the required fresh air flow as well as the temperature of that air flow. Multiplication
of both parameters leads to a linear relation that is directly proportional to the ACP power
shown in figure 3.11. Unsteadinesses that are observed from figure 3.12(a) and 3.12(b). They
are explained by the performance characteristics of the ECS. Cabin temperature and pressure
is controlled by the ACP air flow and its temperature. For minimizing loss in engine efficiency,
the air flow is minimized. The air flow then equals the minimum required fresh air flow that is
required by airworthiness regulations, and cooling or heating is only controlled by temperature
of the constant fresh air flow. The ACP air flow is constant and minimum accordingly for
specific cabin utilizations over a large range of ambient temperature, see figure 3.12(a). For
3.2 Proposed Systems Model 41

the same range of ambient temperature, temperature of the ACP air decreases with increasing
ambient temperatures, see figure 3.12(b). Hence, additional required cooling is provided by
lower temperature of the air flow. According to SAE ARP 85 [143] temperature of air that
enters the cabin must be in the range of 10 to 50 ◦C. Thus, the temperature range of fresh ACP
air is also limited and cannot be adjusted at will for temperature control of the cabin. Since fresh
air is first mixed with recirculated cabin air before it enters the cabin, ACP air temperature can
vary approximately between 0 ◦C and 70 ◦C for the chosen application, see figure 3.12(b). Only
if the ACP air temperature has reached its upper or lower limit12 , the ACP air flow is increased
for additional cooling or heating power, see figure 3.12(a). This is the case for very cool or hot
ambient temperatures and low cabin utilizations. The ECS operates in heating mode if the
temperature of the fresh ACP air is higher than the cabin temperature13 . Figure 3.12(b) shows
the crossover from cooling to heating mode (dashed line). Heating is required for low ambient
temperatures and low cabin utilization. The chosen application represents a flight during night
time, and thus no heat loads due to sun radiation were considered in this example.

The afore discussed system behavior shows plausibility of the implemented ECS model. The fol-
lowing section elaborates on the design point of the ECS that is used for determining maximum
power requirements, and thus also system weight.


   
  
 


  


  

 
 

     
   
     
  

 !      
 
 

   
 


         
   
 
(a) Fresh bleed air mass flow from ACP. (b) Temperature of fresh bleed air from ACP.

Figure 3.12: ACP cooling or heating in dependency on ambient temperature and utilization.

Design Point

Airworthiness regulations specify minimum required air ventilation rates for fresh air supply.
EASA regulations require a minimum fresh air supply of 0.28 kg/min (approx. 0.59 lb/min) per
occupant [9, 25.831]; FAA regulations require 0.55 lb/min [89, §25.831]. In normal operations, a
12
Constant ACP air temperature in fig. 3.12(b).
13
The cabin temperature was set to 24 ◦C for the chosen application.
42 Chapter 3. Aircraft Systems Model

higher fresh air supply is usually supplied to the cabin for improved passenger comfort [32, ch. 7];
typically 0.75 lb/min to 1.0 lb/min is supplied per occupant. The fresh air supply increases
further in case of high cooling or heating loads.

The ECS must be sized for maximum required fresh bleed air flow per ACP, and for total max-
imum air flow for ventilation [11, ch. 4.1]. The design point, however, depends on the aircraft
and the heat loads. The SAE ARP 85 [143] suggests investigation of different design cases for
preliminary ECS design. Cooling as well as heating cases shall be investigated according to
table 3.3 (cases 1 to 4). Degraded systems performance in case of ACP failure must also be con-
sidered (case 5 and 6). Since the provided design cases only consider minimum required fresh
air flow per occupant, the author defines a seventh case that corresponds to normal operations
with increased passenger comfort.

Cooling Heating Degraded Comfort


1 2 3 4 5 6 7
Flight phase cruise ground cruise ground cruise cruise cruise
Altitude, ft 33 000 0 41 000 0 33 000 41 000 33 000
Utilization, % 100 100 20 20 100 20 100
Cabin temperature, ◦C 24 27 24 24 27 21 24
ISA condition 0 +25 0 −55 0 0 0
Time of day day day night night day night day
Minimum air inflow
0.55 0.55 0.55 0.55 0.55 0.55 0.75
per PAX, lb/min
ACPs ACPs ACPs ACPs 1 ACP 1 ACP ACPs
System health
active active active active failed failed active

Table 3.3: Design cases for the environmental control system.

The implemented ECS model was applied to an A320 design (cf. ch. 4.3.1). The aforementioned
design cases have been calculated. Results for total required bleed air flow are illustrated in
figure 3.13(a), and for bleed air flow per ACP in figure 3.13(b). Results are normalized in
reference to maximum fresh bleed air flow of the ECS of an Airbus A320 [7]. It can be observed
that design case 2 results in highest required total bleed air flow, as well as in highest required
bleed air flow per ACP. Hence, for the chosen example, cooling on ground on a hot day sizes
the ECS. Cooling in cruise with one ACP inoperative results in required bleed air flow per ACP
that is only slightly lower. Further, figure 3.13 also shows close agreement between the Airbus
A320 ECS maximum performance and the results from the proposed model. The additional
design case that has been suggested by the author has no impact on sizing of the ECS of this
aircraft. It does not determine maximum required ECS energy, but defines typical operational
3.2 Proposed Systems Model 43

ECS power requirements during cruise flight.

For sizing of the ECS, the proposed model calculates required bleed air flows for the seven
defined design cases and choses the one that results in highest required bleed air flow. ACPs
and ventilation systems are sized accordingly. The mission dependent bleed air consumption
of the beforehand sized ECS is then calculated for each defined mission increment.
Normalized bleed air flow

Normalized bleed air flow


A320 maximum A320 maximum
1.0 fresh air flow
1.0 fresh air flow

0.8 0.8

0.6 0.6

0.4 0.4

1 2 3 4 5 6 7 1 2 3 4 5 6 7
Design cases Design cases
(a) Total required ECS bleed air flow. (b) Required ECS bleed air flow per ACP.

Figure 3.13: Required ECS bleed air flow for the design cases (A320 maximum bleed air flow [7]).

3.2.3 Ice and Rain Protection – ATA-30

The ATA-30 system provides protection against ice and rain (cf. app. A). For today’s com-
mercial transport aircraft, the functions concentrate on ice protection; rain protection is more
or less limited to windshield wipers [154]. Wing leading edges, engine inlets, as well as sensors
and antennas are de-iced or anti-iced by the system. Conventional jet aircraft are equipped
with anti-icing systems that use bleed air for ice protection of the wing leading edges as well
as of the engine inlets, and electrical heating for ice protection of other components such as
sensors or valves. The wing leading edge anti-icing subsystem is the most important and largest
consumer of energy within ATA-30. The implemented model of the wing anti-icing system is
described in the following sections.

System Description

A model for conventional (bleed air) and all electric ATA-30 systems are implemented in the
proposed systems model. Full vaporizing ice protection of the wing leading edges is implied
by the model. The heating power of such systems must be sufficient to fully evaporate any
44 Chapter 3. Aircraft Systems Model

water and moisture on the leading edge surface so that no runback14 ice forms [155]. For
commercial transport aircraft, ice formation on the inboard wing is acceptable and does not
pose a risk to safe flight [32, ch. 6.5]. Thus, only the outboard wing is protected by an anti-
icing system. The required span of the ice protection depends on detailed ice accretion and
aerodynamic characteristics of the wing, and is influenced by e.g. stalling characteristics and
thickness of the airfoils. A detailed prediction of the required extension of the ice protection
on the wing leading edges is not possible during conceptual aircraft design. Thus, the relative
span position, at which ice protection begins, is set as input for the proposed model15 . For
conventional transport aircraft it is sufficient to assume a starting point outboard of the inner
engine nacelle. The Airbus A320 aircraft is, for example, equipped with anti-ice protection on
the three outboard slats, cf. figure 3.5.

In case of conventional bleed air systems, Piccolo tubes are installed inside the slat elements to
distribute hot bleed air in span direction, see figure 3.14(a). Bleed air exits the Piccolo tube via
numerous openings and circulates inside the slat elements, where it heats the leading edge skin
of the wing. As illustrated in figure 3.14(b), electrical heater mats are used for heating of the
surface to the required temperature in case of an electrical anti-ice system. For both systems
a defined heat flux is required for ice protection.

  







(a) Conventional bleed air system. (b) Electric system.

Figure 3.14: Simplified system schematics for leading edge anti-icing system.

Modeling of Systems Power Consumption

The proposed model estimates the required heat flux at the outer wing skin in dependency
on wing geometry, atmospheric conditions, and flight conditions. Depending on the applied
system technology (cf. figure 3.14), the model results in either the required total bleed air mass
flow or in the required electrical power for heating.
14
Runback ice occurs if the system does not fully evaporates the moisture on the surface. The water runs back
from the heated zone and forms ice accretion downstream of the leading edge.
15
Input and output of the proposed model are discussed in the following section, cf. fig. 3.15
3.2 Proposed Systems Model 45

Available models from literature for estimating ice accretion and the required heat flux for
anti-icing are mostly based on detailed calculations with higher-order models. The most widely
spread model for determining ice accretion is the NASA LEWICE code [156, 157, 158, 159]. The
code incorporates dynamic icing models that allow for predicting ice accretion in dependency
on all relevant parameters. The code calculates the trajectory of single water droplets and
their impact position on the wing. Thus, for given flight and atmospheric conditions, ice
accretion can be dynamically simulated. Similar models were published by ONERA [160]
and RAE [161]. All three models are approved by authorities for support of airworthiness
qualification of ice protection systems. Correct modeling of the inner heat transfer is especially
important for estimating the required bleed air mass flow. Different heat transfer coefficients for
impinging jets were first published by Martin [162]. In a more recent work, Wright [163] used
the NASA LEWICE code and test results to derive heat transfer correlations for Piccolo tube
applications. Besides these semi-empirical approaches, different higher-order CFD simulations
were published that model the bleed air circulation and the corresponding heat transfer inside
a slat element [164, 165, 166, 167, 168]. Such approaches are, however, out of the scope of this
work, since their application in conceptual design is not feasible due to high modeling efforts
and computation times. The author chose to use a semi-empirical icing and heat transfer
model that was proposed by SAE AIR 1168/4 [169]. It has been implemented and enhanced
by Freund [8] for incorporation also of 3D wing geometries in scope of an ILR Bachelor thesis.
Another semi-empirical approach to modeling of ice accretion and heat transfer was published
by Sherif et al. [170]. Other than the SAE model, this one requires detailed information on
pressure distributions and velocity profiles over the wing segments.

The overall structure of the implemented model is illustrated in figure 3.15. The left-hand
side of the figure shows the model structure, and the right-hand side shows the required input
parameters accordingly. Input parameters are usually provided from integration of the proposed
systems model into overall design synthesis. However, the proposed systems model is a stand-
alone tool, and thus the required input parameters can also be provided by other means. In
a first step, the wing is discretized into sufficiently small segments for assuming 2D airfoil
characteristics of each segment. Also the liquid water content (LW C) of the air is estimated
from the atmospheric conditions and the flight level. Then, for each of the discretized wing
segments, the total required heat flux (Q̇i ) for full vaporizing anti-icing is estimated from the
energy balance of the outer wing skin:
Q̇i = hwingi · Sicingi · (Tskini − Taw ) + Mi · cw · (Tskini − Taw ) + M ·L . (3.23)
| {z } | {z } | i{z }e
energy flux convection Q̇convi energy flux heating Q̇heati energy flux vaporization Q̇vapori

Energy fluxes for convection, heating of the water, and vaporization are considered. For esti-
mating the different heat fluxes, it is required to calculate required outer wing skin tempera-
ture (Tskin ) and the adiabatic wall temperature (Taw ), the surface area that is anti-iced (Sicing ),
46 Chapter 3. Aircraft Systems Model

  & !(  



&

#
$    
 
    
 
   
 %#

 

 &

 
  
 


 
 !
  
'
    
     




    

  
    
 



    

 
 
 
 




 

 

 


 
  !
! 


  


  
    
 !

  

!

 
   !

 $(

    
 
   

     " # " $ % " % "$   


 !
! 


 
  




   "
+&


&



"

 +&

   (  &
 




  &  ) ,& 
   (  
 
*&   )

Figure 3.15: Structure and input parameters of the implemented wing anti-icing model [8].

the water catch (M ), the heat transfer coefficient of the wing structure (hwing ), the heat capa-
city of water (cw ), and the specific heat for vaporization of water (Le ). The different parameters
are estimated from the input data as shown in figure 3.15. Details regarding the single calcula-
tions and models for estimating the required energy fluxes are provided by Freund [8] or in the
SAE publication [169]. Summation of all required heat fluxes of each segment results in total
required energy flux for anti-icing of the wing (Q̇wing,AT A30 ):
X
Q̇wing,AT A30 = Q̇i . (3.24)

The required bleed air mass flow for each wing segment (ṁi ) is calculated with the enthalpy
equation from the required energy flux, the difference in bleed air (Tbleed ) and skin (Tskin )
temperature, the specific heat capacity of air (cp ), and the efficiency factor (ηbleed ):
X X Q̇i
ṁwing,AT A30 = ṁi = . (3.25)
cp · (Tbleed − Tskini ) · ηbleed
3.2 Proposed Systems Model 47

Summation of the single required bleed air mass flows results in total required bleed air for
anti-icing of the wing leading edge (ṁwing,AT A30 ). In case of electro-thermal anti-icing, total
required electric power flux (Ṗwing,AT A30 ) results from the required heat flux (Q̇wing,AT A30 ) and
the efficiency factor of the electrical heating devices (ηelectric ):

Q̇wing,AT A30
Ṗwing,AT A30 = . (3.26)
ηelectric

Efficiency of a bleed air anti-icing system averages to approximately 0.65 [4], whereas the
efficiency of an electro-thermal anti-icing system is approximated to 0.8 [169]. Please note that
by using approximated efficiency factors for estimating total required power flux, the absolute
value may comprise inaccuracies. However, as aforementioned it is the goal of the proposed
model to capture sensitivities to overall aircraft design synthesis correctly. This is still achieved
by the purely function based approach that is suggested in the scope of this chapter.

Icing Conditions

Icing on aircraft usually occurs in altitudes


below 22 000 ft and in ambient temperatures
Ambient temperature, ° C

0
between −30 ◦C and 0 ◦C. Supercooled water Icing Envelope
-5
droplets exist in saturated air (clouds) that -10
stay in the fluid state despite temperatures -15
below freezing. Such supercooled droplets -20
-25
form ice accretion on an aircraft. The sever-
-30
ity of icing depends on the atmospheric con-
0 2 4 6 8 10 12 14 16 18 20 22 24
ditions and the water content. Ice protec- Pressure altitude, 1000 ft
tion requirements are specified by airworthi-
ness regulations for commercial transport air- Figure 3.16: Icing envelope for standard icing
craft, e.g. by EASA CS-25 [9, app. C] and conditions [9, app. C].
FAR Part 25 [89, app. C]. Both regulations
provide the same limiting icing envelope for normal atmospheric conditions, see figure 3.16. In
case of adverse weather conditions, icing becomes also possible in higher altitudes.

Both EASA and FAA regulations consider two regimes for icing; maximum continuous icing and
maximum intermittent icing. The maximum continuous icing condition is defined by exposure
to low or medium liquid water content for an extended period of time. It is applicable to those
components such as wings that are affected by flight in continuous icing conditions, but that can
tolerate brief or intermittent encounter of conditions with greater severity [169]. The maximum
continuous intensity of atmospheric icing conditions is defined by the cloud LWC (0.05 g/m3
48 Chapter 3. Aircraft Systems Model

to 0.8 g/m3 ), the mean effective diameter of the droplets (15 ➭m to 40 ➭m), the ambient air
temperature (0 ◦C to −30 ◦C), and the interrelationship of these variables.

Maximum intermittent icing condition is defined by exposure to high LWC of up to 2.75 g/m3
for a brief period of time. It is applicable to those components such as engine inlets or sensors,
where ice accretion, even for a short period of time, cannot be tolerated. Maximum intermittent
icing conditions are defined by clouds with a standard horizontal extent of 3 NM. Other than
the FAA, the EASA defines a third icing regime; maximum take-off icing. It is defined by a
liquid water content of 0.35 g/m3 , a mean effective diameter of the water droplet of 20 ➭m, and
an ambient temperature on ground level of −9 ◦C [9, app. C]. The different icing conditions
and relations that are defined by EASA and FAA regulations will be used for evaluation of
plausibility of the implemented model as well as for definition of the design point of the wing
anti-icing system in the scope of the following sections.

Sensitivity and Plausibility of the Model

As aforementioned, the required energy for anti-icing results from the flight condition as well
as from the atmospheric conditions. In this section, the sensitivity of the model to these
parameters is discussed for showing plausibility of the model. Adjacently, the design point
for the wing leading edge anti-icing system is defined. Results are based on a 150 passengers
Airbus A320-like single-aisle aircraft design (cf. ch 4.3.1).

Figure 3.17 shows the impact of variation of


altitude on the required bleed air mass flow
ISA+5
for full vaporizing anti-icing of the wing lea- 0.4 ISA 0
Bleed air, kg/s

ding edge. Results are plotted for a con- ISA-5


ISA-10
stant true airspeed of 250 kts and an effec- 0.3
ISA-15
tive droplet diameter of 20 ➭m. The ambient
temperature has been varied. It can be ob- 0.2
served that icing conditions are not met for
TAS = 250 kts, ddrop,mean = 20 µm
the full regime of flight altitude. Ambient 0.1
temperature must first drop below freezing 0 5 10 15 20 25
Pressure altitude, 1000 ft
(cf. fig. 3.16). For standard ISA conditions
this is the case in approximately 8000 ft (solid Figure 3.17: Required bleed air mass for varia-
black line in fig. 3.17). Results show that the tion of altitude.
required bleed air mass flow decreases with
increasing flight altitude and decreasing ambient temperature. Also, a saturation effect can be
observed for each atmospheric condition in a different critical altitude (lower part of figure).
3.2 Proposed Systems Model 49

The counter intuitive characteristic of decreasing required bleed air flow for increasing alti-
tude, and thus decreasing temperature can be explained with the liquid water content of the
atmosphere. Extensive atmospheric studies by NACA [171, 172, 173]16 show the relationship
between atmospheric conditions, liquid water content and ice accretion. The studies show that
the highest liquid water content in icing conditions is found at the freezing point. Highest LWC
leads to the most severe ice accretion, and thus to maximum required energy for anti-icing.
The studies also show a saturation of LWC between approximately −20 ◦C and −30 ◦C ambient
temperature, and complete dry air for temperatures below −30 ◦C. These atmospheric cha-
racteristics are mirrored in the results of figure 3.17. Additionally, LWC increases for ambient
temperatures of 0 ◦C in lower altitudes. Thus, maximum required bleed air for anti-icing in-
creases with decreasing ISA temperature as shown in figure 3.17. For constant flight speed and
constant effective droplet diameter, highest energy for anti-icing is found for ISA conditions
of −15 ◦C (freezing at sea level).

Impact of variation of flight speed for a constant flight level and a constant mean effective
droplet diameter is shown in figure 3.18. It can be observed that the required bleed air mass
flow for anti-icing increases with increasing flight speed. This is caused by an increase in
required convection heating and an increase in the amount of water that accumulates on the
wing. Those two effects overcompensate the slight increase in boundary layer temperature with
increasing flight speed. Figure 3.18 also shows that the required bleed air mass flow increases
with increasing ambient temperatures in constant altitudes. This is again explained by the
lower LWC in lower ambient temperatures. Maximum icing conditions in terms of maximum
flight speed are further limited by the aircraft operational envelope and airspace limitations.

The mean effective droplet diameter directly influences LWC of the air [171, 172, 173], which
increases with smaller droplet diameters. Hence, the required bleed air mass flow for anti-
icing also increases with smaller droplet diameters, see fig. 3.19. According to EASA [9] and
FAA [89], icing conditions are limited to effective mean droplet diameters of no less than 15 ➭m.
Again, the required bleed air mass flow increases with temperature due to the increasing LWC.

The complex relation between flight speed, altitude, and required bleed air mass flow is ad-
ditionally illustrated in figure 3.20. Standard ISA conditions and a mean effective droplet
diameter of 20 ➭m have been used. Highest required bleed air mass flow is found for highest
flight speed at lowest possible altitude. The maximum flight speed is limited to 250 kts for oper-
ations below 10 000 ft by airspace regulations, and to the maximum operation speed (VMO) for
operations above 10 000 ft. A comprehensive sensitivity study of the icing model for all input
parameters was conducted by Freund [8] and is not further discussed in scope of this thesis.

16
Although the studies were conducted in the early 1950s, current EASA [9, app. C] and FAA [89, app. C]
regulations use the same data for definition of icing conditions.
50 Chapter 3. Aircraft Systems Model

The results show feasibility and plausibility of the implemented model.

ISA+5 ISA+5
0.4 ISA 0 0.4 ISA 0
Bleed air, kg/s

Bleed air, kg/s


ISA-5 ISA-5
ISA-10 ISA-10
0.3 0.3
TAS = 250 kts
h = 12,000 ft
0.2 0.2

h = 12,000 ft, ddrop,mean = 20 µm


0.1 0.1
160 180 200 220 240 260 280 300 320 15 20 25 30 35 40
True Airspeed, kts Effective droplet diameter, µm

Figure 3.18: Required bleed air for variation Figure 3.19: Required bleed air for variation
of TAS. of droplet diameter.


 





  
! 
!
 

 

 ! ! 


  !
 
 !  


" 

 
"! 
 

Figure 3.20: Required bleed air for full vaporizing anti-icing of a typical single-aisle aircraft in
standard ISA conditions (dashed lines show iso bleed air mass flow).

Design Point

It has been shown that the required energy for anti-icing depends on the specific atmospheric
as well as flight condition. Hence, an ideal system would continuously adjust its energy con-
sumption to match the specific condition. This is, however, not how today’s anti-icing systems
operate. Today’s systems operate digitally, which means that they are either switched off or
operate in a constant state. Thus, the system must be sized for worst conditions to guarantee
sufficient power in all relevant icing conditions such as specified by EASA [9] and FAA [89]. For
the wing anti-icing system, maximum continuous icing conditions, and for an EASA certification
also maximum take-off icing conditions must be considered.
3.2 Proposed Systems Model 51

Maximum required bleed air mass flow of 1.1 473


a conventional anti-icing system as well 1.0 430

Electric power, kW
as maximum required electric energy for

Bleed air, kg/s


0.9 Design Point 387
an electric one are plotted in figure 3.21
0.8 344
for maximum continuous icing conditions
0.7 301
in dependency on the ambient tempera-
ture. It was already shown beforehand 0.6 Max. Continous Icing 258
Max. Take-Off Icing
that maximum ice accretion can be ex- 0.5 215
-20 -15 -10 -5 0 5
pected for highest flight speeds at low- Delta ISA Temperature, ° C
est altitude, and minimum mean effec-
tive droplet diameter. Calculations show Figure 3.21: Required bleed air for maximum contin-
that all points in figure 3.21 correspond uous icing and take-off icing conditions.
to a flight speed that equals VMO, an al-
titude of 10 000 ft, and an mean effective droplet diameter of 15 ➭m. Hence, maximum required
energy for full vaporizing anti-icing of the wing leading edge in maximum continuous icing con-
ditions must be expected for ISA 5 ◦C conditions17 . Additionally, also the required energy for
maximum take-off icing conditions is plotted in figure 3.21. It can be easily observed that this
condition is not constraining the system design. Hence, the wing leading edge anti-icing system
of the A320-like aircraft is sized for a bleed air mass flow of approximately 1.07 kg/s by the
implemented model. For an electric wing anti-icing system with heating mats, the maximum
required power results in 460 kW.

3.2.4 Primary and Secondary Flight Control System – ATA-27

The flight control system (FCS) is further grouped into primary (PFC) and secondary flight
controls (SFC). PFC provide the aircraft-level function of 3-axis control and comprise ailerons,
elevators, rudders, and spoilers. SFC provide high lift and trim functions. It includes leading
and trailing edge high lift devices, as well as any means of trim such as trimable horizontal
stabilizers (THS). The flight control system comprises all components from cockpit controls to
actuation systems that move the single flight control surfaces. Flight control computers (FCC)
as well as actuators, gears and transmission systems are included. Not included are the aero-
dynamic control surfaces, which belong to the aircraft structure. The conventional FCS is a
main consumer of hydraulic energy. Hydraulic energy is converted by hydraulic-mechanical
actuators to deflection of control surfaces. In modern aircraft, more and more EHAs or EMAs
are used, which makes the FCS a significant consumer of electric energy.

17
Corresponds to 0 ◦C in 10 000 ft
52 Chapter 3. Aircraft Systems Model

System Description

The implemented FCS model concentrates on sizing of the different means of actuation in
dependency on the applied air loads. The model assumes an electronic FCS with fly-by-wire
technology. The focus is not put on the electronic flight controls with their different control
computers and data buses. Electronics do not contribute significantly to the overall system
weight and power consumption (cf. ch. 3.1.1). Typical architecture and features of electronic
flight control systems are described in recent publications by the author [174, 175, 176]. Fur-
ther, the implemented model assumes conventional configuration of the FCS with segregation
into primary and secondary flight control functions such as illustrated in figure 3.22. The
implemented FCS model for estimating system weight and power consumption of the FCS is
limited to certain component-level technologies. For example, the implemented model concen-
trates on conventional centralized high lift systems architectures that are driven by means of
power control units (PCU); distributed high lift systems are not yet covered. The flexible and
modular structure of the proposed model, however, allows for enhancing the model with specific
component-level technologies that are required for specific applications. For verification activi-
ties and sensitivity studies that are discussed in the scope of this thesis, the use of conventional
actuator technology within the flight control system is sufficient. Since the required power for
actuation of the single control surfaces is derived from the specific air loads of each control
surface, power modeling is independent from specific component-level technologies. The case
study in chapter 6.2 shows how easily the proposed flight control system model can be adapted
to also cover innovative technology. A distributed and electric-driven leading edge system is
integrated into the existing model for sound technology assessment.

The primary control surfaces are actuated by either hydraulic or electric actuators. The se-
condary control system comprises high lift devices as well as means for trim. Roll and yaw
trim are achieved by deflection of the ailerons and the rudder accordingly. For pitch trim,
however, the proposed model assumes a trimable horizontal stabilizer, which is common in all
modern transport aircraft. Further, the model concentrates on centralized high lift systems.
Both the leading and trailing edge devices comprise a centralized drive shaft, which is driven
by a power control unit (PCU). The PCU is of duplex motor design with a speed-summing
differential gearbox. The transmission shaft comprises several gearboxes and joints, as well as
torque limiters, wing tip brakes (WTB), and position sensors (APPU). The high lift devices
are driven by mechanical actuators, which transform the rotatory motion of the transmission
shaft into the desired surface deflection. For slat devices a track and pinion drive is assumed. A
flap track kinematic with rear link is assumed for the Fowler flap system, such as it is common
in current Airbus aircraft. Although the transmission system of the trailing edge devices are
not shown in figure 3.22, they are of the same design as the drive system of the leading edge
devices. A more detailed description of high lift systems and kinematics of various commercial
3.2 Proposed Systems Model 53

&&* 
+!
! "


 
        
 
   



! 

 
'( "
)     

&

 

 

 )

 
 
  # $# % 

Figure 3.22: Configuration of the conventional electronic flight control system (flap drive system
not shown, but analogue to slat transmission system).

aircraft can be found, for example, in a NASA contractor report by Rudolph [142].

Modeling of System Power Consumption

Modeling of the single consumers within the FCS is based on a physical approach. The re-
quired energy for deflection of the single control surfaces (Psurf ) is estimated from the hinge
moment (Mhinge ) and the deflection rate (γ̇):

Psurf = Mhinge · γ̇ . (3.27)

Hereby, hinge moments of the different flight control surfaces are directly proportional to the
dynamic pressure, and thus also to flight phase dependent air loads. The required energy
for surface deflection is distributed over the single hydraulic or electric actuators of that sur-
face. Total required power of the FCS accumulates from the single consumers for each flight
increment:
n m
! k
!
X X X
PF CS (t) = Psurf ,i (t) = Pact,i (t) + ... + Pact,i (t) . (3.28)
i=1 i=1 surf,1 i=1 surf,n

Today, two different concepts for actuation of the primary control surfaces exist. Some aircraft
use actuators in active-active configuration. The energy is then distributed equally among the
actuators. Other aircraft use actuators in active-standby configuration. Here, only one actuator
is active and carries the entire loads. For such a configuration, total energy per control surface
is not distributed among the actuators but each actuator is sized according to the maximum
54 Chapter 3. Aircraft Systems Model

load. Further, maximum power requirements result from maximum control surface deflections
under maximum air loads. It, however, must be kept in mind that control surface deflections
are limited with increasing dynamic pressure to protect the aircraft against excessive loads.
Koeppen [11, ch. 4.4] elaborates on the dependency of maximum control surface deflection on
flight speed and altitude. The proposed systems model uses deflection limitations that are
based on reference data by Koeppen.

A control surface can be simplified as a thin panel according to Scholz [102]18 . Further, this ap-
proach assumes inviscid, irrotational, and steady flow in dependency on the angle of attack (α),
and the control surface deflection (γ). For the thin panel approximation, the hinge moment is
estimated from the dynamic pressure (q), the hinge coefficient (Chinge ), and the reference area
of the control surface (Sref,rudder ) according to Truckenbrodt [152, ch. 12]:
1
· ρ · v 2 · Chinge (α, γ) · Sref,rudder
Mhinge = 2
. (3.29)
2
The hinge coefficient can be written as follows:
Z Z
∂Ch ∂Ch
Chinge = dα + dγ + Ch,0 = Ch,α · α + Ch,γ · γ + Ch,0 . (3.30)
∂α ∂γ |{z}
=0

For zero deflection (γ = 0), no hinge moment results on the control surface19 . Hence Ch,0 must
equal zero. According to Truckenbrodt [152, ch. 12], the control surface’s width-to-depth ratio
(λrudder ) can be used to estimate the hinge coefficient. The width-to-depth ratio is defined by
the mean control surface width (crudder ) and the mean chord length (cmean,chord ):
crudder
λrudder = . (3.31)
cmean,chord
The following estimates apply for deriving Ch,α and Ch,γ from λrudder :
 q p 
1
Ch,α = − 2 · (3 − 2λrudder ) · λrudder − λ2rudder − (3 − 4λrudder ) · asin λrudder (3.32)
λrudder
 (3/2)  p q 
4 1 − λrudder
Ch,γ = − · asin λrudder − asin λrudder − λ2rudder . (3.33)
π λrudder
This approach results in the maximum required power for control of the aircraft, and is thus
valid for sizing of the FCS and the secondary power systems. Operational loads, however, are of
much smaller order and mainly result from leakage within the hydraulics [11, 103]. Operational
loads of the PFCS are almost constant for the different flight phases. Hence, the proposed
systems model estimates operational loads of the PFCS from maximum control forces of the
specific flight phase and a constant factor (f ):

Pprimary, operational loads = f · Pprimary, max load (α, γ, v, ρ) . (3.34)


18
Koeppen [11] has enhanced this approach by also considering compressibility and control surface camber.
However, additional input is required that is generally not available within conceptual aircraft design.
19
Simplification of the model; both ailerons and elevators usually experience a hinge moment for zero deflection.
3.2 Proposed Systems Model 55

Figure 3.2 already showed a comparison of hydraulic power requirements between Airbus A320
flight test data and results of the proposed systems model. Close agreement in the data is
achieved. It can be observed from figure 3.2 that the assumption regarding almost constant
base loads is valid. Since the focus within this chapter is put on weight estimation rather than on
power estimation, the results are not further discussed. Greater details on estimation of power
consumption of the PFCS can be found in a ILR diploma thesis by Steinke [177]. The required
power of the SFCS is also derived from functional relationships. For the conventional high lift
system, PCU power requirements are estimated from aerodynamic loads, drive speeds, and the
different gear ratios within the transmission system. A detailed description of the implemented
model can be found in a ILR thesis by Groening [178]. His work orientates itself at the different
functional relations of the high lift system that are e.g. described by Koeppen [11].

Modeling of System Weight and Center of Gravity

The proposed systems model uses a book-keeping approach for estimating total FCS weight,
see figure 3.23. Hereby, total FCS weight (mAT A27 ) is accumulated from the single component
weights (mcomp ) and the number of the components (ncomp ):
m
X
mAT A27 = mP F CS + mSF CS = (mi · ni )comp ; (3.35)
i=1

control surface structure is not included in the FCS system weight. Further, the center of grav-
ity (c.g.) of the FCS is also accumulated from the centers of gravity of the single components:
Pm
i=1 (c.g.i · mi · ni )comp
c.g.AT A27 = c.g.P F CS + c.g.SF CS = . (3.36)
mAT A27

A NASA contractor report by Boeing [179] concludes with semi-empirical estimates for the
PFCS actuators of the single control surfaces as well as for the trimable horizontal stabilizer.
Actuator weight is given in dependency on the required actuation power, the control surface
geometry and other empirical parameters. The different estimates that are suggested by Boe-
ing [179] and that have been adopted into the proposed systems model shall not be discussed
in greater detail in the scope of this chapter. Instead, focus is put on the implemented method
for deriving component weights of the high lift system.

A different NASA contractor report by Rudolph [142] provides semi-empirical estimates for lea-
ding and trailing edge high lift system weights. The report also gives a comprehensive overview
on high lift system and component technology that is currently used onboard of commercial
transport aircraft. Rudolph distinguishes between different types of devices; e.g. single-slotted
Fowler flap with link and track kinematics, or slat with rack and pinion kinematics are covered.
56 Chapter 3. Aircraft Systems Model

 
 
   

  
    
 
 
  


 
 
 
  

      
  
    
  

 
  
 
 

   
 &
    &
   
 

     

 

 !     !    
 
   
  
 

 

 
 &
'   &
'  
 
 




 
   
 

 
   
 

 
  
(&) (&)
  
   


 
  
 
   


"
#  "
# 
$  $ 

%  % 
 
%  % 
 

Figure 3.23: Book-keeping approach for estimating FCS weight and center of gravity.

Also, Kruger slats and multiple-slotted Fowler flaps are included20 . Further, Rudolph breaks
the leading and trailing edge systems down into the following functional units: high lift panels,
kinematics, actuation and drive systems, and fairings. For each of the functional units the
weight is estimated from the slat or flap area and the specific technology. Total high lift sys-
tem weight accumulates from the single functional unit weights. Since only the functional unit
of actuation and drive system is allocated to the flight control system21 , Rudolph’s approach
has been enhanced in the scope of a ILR thesis by Chaves-Vargas [180].

In the proposed model, the functional unit of actuation and drive system is further broken down
into the single components that are listed in figure 3.23 (left-hand side). For each of the single
components, the weight is estimated from semi-empirical approaches that have been derived
by Chaves-Vargas and the author [180]. Data that has been published by the airframers in
Maintenance Facility Planning Documents (e.g. [181, 182, 183, 184, 185, 186, 187, 188, 189,
190, 191]) and other internally available documents (e.g. [192] and [10]) have been used. A
complete overview of all implemented models for estimating component weights of the high lift
20
For a complete overview of all technologies that are covered by Rudolph’s approach please refer to the
corresponding NASA report [142].
21
All other aforementioned functional units of trailing and leading edge high lift systems are commonly allocated
to the airframe structure and not to the flight control system (ATA-27).
3.2 Proposed Systems Model 57

system cannot be given in the scope of this thesis. Here, the approach is exemplified with the
help of the power control unit (PCU). Figure 3.24 shows the relation between installed PCU
power and PCU weight of selected aircraft. Due to the hooked track kinematic that was used
for the Airbus A300/310 high lift system, comparatively high PCU power is required (see top
right corner of fig. 3.24). Close agreement for the suggested regression analysis can be observed.
In the proposed model, the PCU weight is estimated from the installed PCU power as follows:

mP CU = 11.85 · PP0.353
CU . (3.37)

The required power is estimated by means of a physical approach from the air loads and the
required actuation forces. Similar relations, also for the other components of the high lift
system, have been implemented for estimating total high lift system weight from the weight of
the leading and trailing edge subsystems, and their components accordingly:

mSF CS = mLEdevice + mT Edevice with


mLE device or mT E device = n · mact + n · mtorque limiters + mtrans. shaf t (3.38)
+ mjoints + mgearboxes + mW T Bs + mAP P U s + mP CU .

60
A300/310 flap
A380 slat
50
A330/340 slat
A300/310 slat
PCU weight, kg

40
A320-family A330/340 flap
A340-600 slat
30

20
Sukhoi Superjet 100 A/C data
10
Regression
0
0 10 20 30 40 50 60
Installed PCU power, kW

Figure 3.24: Regression analysis for estimating power control unit weight (data source [10]).

So far, only estimates for deriving component weights of actuators and the high lift transmis-
sion and drive systems have been discussed. These estimates do not include additional FCS
system weight of the control inputs, FCCs and electronics, as well as of cabling between cockpit
controls and FCCs, and between FCCs and actuators. Literature does not provide a method
for estimating these missing component weights individually. Instead, the proposed model uses
weight fractions that are suggested by Koeppen [11]. He concludes that 24 % [11, ch. 4.4]
of total PFCS weight is not included in component weights of the single actuators, and that
58 Chapter 3. Aircraft Systems Model

2 % [11, ch. 4.5] are not included in the transmission and drive systems of the high lift system
accordingly. Koeppen accounts the actuation of the trimable horizontal stabilizer to the PFCS
weight for the suggested mass fractions. Hence, total FCS weight accumulates as follows:

mAT A27 = 1.24 · [(m · n)aileron + (m · n)spoiler + (m · n)elevator + (m · n)rudder ]act


+ 1.24 · [(m · n)horizontal stabilizer ]act (3.39)
+ 1.02 · [mLE device + mT E device ] .

The FCS’s center of gravity is also derived from the single FCS components. According to
chapter 3.1.1, the proposed model only considers the coordinate system in x and y dimension
for reasons of simplification. Further, for a symmetric aircraft configuration the single centers
of gravity are spread along the symmetry line22 . Nevertheless, for the FCS each compon-
ent is considered individually as illustrated in figure 3.25, which allows for covering arbitrary
configurations. Figure 3.25 shows the right-hand elevator actuators and their centers of grav-
ity. Positioning of the single actuators, and thus of their centers of gravity is derived from
the airframe structure and control surface geometry. Within overall aircraft design synthesis
(c.f. fig. 3.3 or 4.1), the airframe structure as well as the single control surfaces are sized prior
to the aircraft systems. Thus, layout of the airframe structure and of the single control surfaces
is available. This information is used by the proposed model for positioning of the single actu-
ators: Actuators are positioned aft of the rear spar and in front of the control surface (c.g.x,act. )
as well as close to the inboard and outboard end of the control surface (c.g.y,act. ). The single
components of the transmission and drive system of the high lift system are positioned accor-
dingly (cf. fig. 3.22). After positioning of the single components, the overall FCS center of
gravity is estimated according to equation 3.36.








 




 


 








 

Figure 3.25: Derivation of FCS center of gravity from the single components.

22
Positions in y-dimension must equal zero for symmetric configuration.
3.2 Proposed Systems Model 59

Plausibility of the Proposed Model

Again, results that are discussed are based on a A320-like design (cf. ch. 4.3.1). Koeppen
provides a weight break down of the PFCS [11, ch. 4.5] and SFCS [11, ch. 4.5]. He used restricted
data and averaged the weight break down for the entire Airbus fleet. This averaged weight
break down is compared against the estimated A320 FCS component weights in figure 3.26.
Figure 3.26(a) shows the weight break down for the PFCS, and figure 3.26(b) for the SFCS
accordingly. Close correlation can be observed, although estimated data for the A320 aircraft
are compared against average data of the entire Airbus fleet. Since the proposed model uses
the same mass fractions that Koeppen derived from this data for estimating FCS electronics
weight (c.f. eq. 3.39), it is not surprising that no deviation in estimated electronics weight is
found. However, also for the other components of the PFCS and SFCS deviations are below 5 %,
except for the flap actuation. Also, deviation of flap actuation weight of 6.8 % is well below
tolerances that are expected in conceptual design. Thus, the weight break down for both the
PFCS as well as for the SFCS show overall plausibility of the proposed model.

35 35
Average Airbus fleet Average Airbus fleet
Relative weights, %

Relative weights, %

30 30 -6.8%
Proposed model for A320 0% Proposed model for A320
25 25 +3.0%
-2.9%
20 20 +1.3%
+2.2% -0.1%
15 0% +0.1% 15
+0.6%
10 10 +1.4% +1.2%
5 5 0%
0 0
rudder spoiler aileron THSA elevator elec- flap slat flap slat flap slat elec-
control tronics actuators transmission PCU tronics

(a) Primary flight controls. (b) Secondary flight controls.

Figure 3.26: Weight break down of Airbus FCS [11] and estimates.

A more detailed mass break down of the slat and flap system of the A320 aircraft is shown in
figure 3.27; both A320 aircraft data [10] as well as estimated weights are plotted. Figure 3.27(a)
shows the weight break down of the slat system, and figure 3.27(b) the one of the flap system
accordingly. Close agreement can be observed from both figures. Total system weight of
both the slat and flap system shows deviation of approximately 1.5 %, where the proposed
model estimates the total system weights as slightly too high in both cases. Deviations of the
single component weights are below 5 %, except for the torque limiters of the slat system.
In this example, torque limiter weight for both the slat and flap system shows the largest
relative deviations. For the A320-family, torque limiters are integrated into the slat and flap
actuators, and are not listed seperately within the weight break down of the aircraft [10]. The
proposed model, however, assumes a standard high lift system with seperate actuators and
60 Chapter 3. Aircraft Systems Model

torque limiters. It can be summarized that the presented results for the A320 aircraft show
plausbility in component weights of the PFCS as well as of the SFCS.

120 120
+1.5% Airbus A320 slat system weight +1.6% Airbus A320 flap system weight
Relative weights, %

Relative weights, %
100 100
Estimated A320 slat system weight Estimated A320 flap system weight
80 80

60 60 +0.2%
0%
40 -4.0% 40
-3.5%
20 +1.9% 20 +1.8%
+5.3 -0.5%-1.8%+0.7% +4.4 -0.5%-1.1%+0.5%
0 0
total
weight

PCU
trans.
shaft

actuators
torque
limiters

WTBs

APPUs
gear
boxes

total
weight

PCU
trans.
shaft

actuators
torque
limiters

WTBs

APPUs
gear
boxes
(a) Slat system component weights. (b) Flap system component weights.

Figure 3.27: A320 high lift component weights [10] and estimated weights.

3.3 Implementation of the Proposed Systems Model

This section describes implementation and program structure of the proposed systems model.
It has been implemented in a C++ software module for integration into MICADO. Main focus
is put on modularity and flexibility regarding adaptability for modeling of variable and new
systems architectures with the provided software tool.

3.3.1 MICADO Program Module Structure

MICADO consists of loosely coupled software modules that each cover a specific discipline
within aircraft design synthesis. A particular strength of MICADO is that all implemented
modules are based on the same standardized data and program structure. Risse [193] elaborates
in greater detail on the control and data flow within MICADO. Here, only a brief overview is
given for providing understanding of the implementation of the proposed systems model. The
single program modules within MICADO are loosely coupled by data coupling, i.e. output of
one program module serves as input to another module. There is neither a strong dependency
between the single programs nor a direct communication or exchange of data. Each program
module only depends on a reasonable set of input data, which is read from a central data
repository XML file called Aircraft Exchange (AiX) and a specific program configuration XML
3.3 Implementation of the Proposed Systems Model 61

file. The AiX file contains the single aircraft design parameters, such as weights, geometry,
performance data, or requirements. The parameter values are calculated within the design
synthesis. As post-processing task, each program writes its results as output into the AiX
file, which are used by subsequent programs as input. Hence, only indirect dependency and
communication between the single tools (via the AiX file) is found in the MICADO environment.
This data coupling approach ensures modularity and flexibility by requiring every tool to feature
stand-alone capability as well as independence from other tools [193].

3.3.2 Modularity and Flexibility of the Program Structure

A simplified class structure of the implemented systems model is illustrated in figure 3.28.
The systemsArchicture class is responsible for generating and sizing of the aircraft systems
architecture. It is composed of object vectors aircraftSystem that define the overall systems
architecture (cf. fig. 3.1). The specific systems architecture is not rigidly implemented in the
program code. The object vectors are dynamic and can contain a variable number of different
systems without changes and re-compilation of the program code. It can be easily and efficiently
edited by the user for variable modeling of different systems architectures.

*+'
 
   



  $   , -
 ./
 "#   $   , -
 ./
"#   .  , -
 ./
$%
"# ,
. , -
 ./

 



 *+'

 "# '(()


"#
 
'  
$%
"#
 

*+'    

&  

 



'  
 

 
  *+'  

   

 "#  
   

"# 
   

$%
"# !   


Figure 3.28: Simplified UML class diagram of the implemented systems model.

All of the defined system object vectors are of the same type aircraftSystem, see center of
figure 3.28. The class aircraftSystem defines the generic aircraft system or building block
(cf. ch. 3.1.2). Characteristics and interfaces of the generic system (cf. fig. 3.6) are defined
by the variables of the class: the system mass (Mass), center of gravity (cgRefPoint), power
consumption (bleedAir, Hydraulic, Electric), and parasite heat loads (heatLoad ). Further, the
62 Chapter 3. Aircraft Systems Model

bool variable systemOperating defines whether the specific system is active during the aircraft
mission. For defining the specific system, an implemented system model (e.g. convention-
alATA21 see lower left part of figure 3.28) is allocated to each aircraftSystem element. These
system model classes contain the specific system models that have been described for the ECS),
ice and rain protection, and FCS in the previous sections. Hereby, each of the implemented
system model class contains a function for estimating power requirements (getPower()), system
mass (getMass()), and center of gravity (getCGRefPoint()).

This flexible and modular program structure allows for implementation of the different systems
models with various levels-of-detail. Further, implementation of the different models is inde-
pendent from each other, since system interdependencies are described by the system interfaces
and not by the specific system model code (cf. ch. 3.1.1). Thus, the overall program can be
freely enhanced by either new system models that describe innovative technology, or by system
models that add level-of-detail or accuracy to existing ones. This fast adaptability of the pro-
gram makes the proposed systems model to an especially valuable tool for sound technology
integration and assessment in conceptual aircraft design synthesis. For example, let us assume
a conventional reference aircraft and integration of an electric ECS and/or an electric anti-icing
system. The flexibility of the proposed systems model allows for automated investigation of
the design space that is shown in figure 3.29. For each option, different implemented aircraft
system models (e.g. conventionalATA21 vs. electricATA21 ) are used for composition of the
overall aircraft systems architecture (cf. fig. 3.28). Integration of the implemented systems
model into MICADO and automated design synthesis then allows for fast and efficient multi-
disciplinary design optimization of each concept, and thus for sound technology assessment and
down-selection of promising technologies.

        



  

      


  !
    
 

  
 
      


  
 !

 
  
!"#

Figure 3.29: System technology design space for flexible system architecture definition.
3.4 Probabilistic Modeling of Innovative Systems 63

3.4 Probabilistic Modeling of Innovative Systems

For integration and technology assessment on overall aircraft-level, innovative systems technol-
ogy often lacks the required maturity in important design parameters. The particular difficulty
of sound technology assessment lies in the low level-of-detail and the uncertainties that remain
within the design. No data or models are available. In the scope of this section, a framework23 is
presented that allows for integration and assessment of innovative systems technology on overall
aircraft-level already in the conceptual design phase by considering the most important design
uncertainties with a probabilistic approach. The proposed framework allows for assessing the
probability that a certain design goal can be achieved within a pre-defined parameter space. It
also concludes with definition of minimum performance margins for the innovative technology,
which are fed-back into the systems development process as performance requirements to the
novel system. In a case study, the proposed framework is applied to technology assessment of
an innovative morphing leading edge for commercial transport aircraft with natural laminar
flow technology, see chapter 6.2. This new framework allows for capturing the most relevant
impacts on overall aircraft-level and for considering design uncertainties. It is characterized by
the following 7 steps:

(1) Identification of Interdependencies of Technology Integration

In a first step, the interdependencies of technology integration are identified. These inter-
dependencies cover all benefits and repercussions of technology integration on both system-
architecture-level and on overall aircraft-level. Hence, the dimension of the parameter space of
the design problem is defined in this first step. Description of these interdependencies are still
qualitative, and are not yet translated to specific technical design parameters. In case of an
innovative high-lift system, an example for possible interdependencies are the impact on the
aircraft’s high-lift performance or on flight control system mass.

(2) Down-Selection of Relevant Interdependencies

The different identified interdependencies are weighted in regards to their overall impact on
design synthesis. Since computing resources are limited, only the most relevant ones are con-
sidered further. The most relevant interdependencies are defined to be those that impact overall
aircraft design on a noticeable scale. Interdependencies that have an impact of smaller order
can be neglected. Down-selection of the most relevant parameters is based on experience and
sound judgment, or on conceptual design studies, in which the impact of different parameters
23
The proposed framework was also presented in the scope of a conference paper by the author [194].
64 Chapter 3. Aircraft Systems Model

on overall design synthesis is evaluated by means of sensitivity studies. Hence, the parameter
space, which has been spanned over n-dimensions in the first step, is reduced to a more man-
ageable size of dimensions in the second step. In the case study that is presented in chapter 6.2,
the author down-selected the design problem to the most three relevant design parameters.

(3) Allocation of Design Parameters to Relevant Interdependencies

Adjacently the most relevant interdependencies are translated from a qualitative description
to a quantitative technical one. A specific design parameter from overall design synthesis is
allocated to each interdependency. For example, for a technology that enables laminar flow,
the design parameter that is mainly influenced is a reduction in friction drag. Changes in
the allocated design parameters propagate through design synthesis and impact overall design.
Relevant benefits and repercussions on aircraft-level are triggered by variation of the allocated
design parameters, if the interdependencies were allocated to the correct ones.

(4) Identification of the Parameter Space of Each Impacted Design Parameter

For conventional technology, the different design parameters can be estimated or computed
from various available models. Such models are, however, often not applicable to innovative
technology. In most cases, only initial estimates are available for an innovative technology,
which are not rarely based on ’rule-of-thumb’ assumptions. Thus, integration of innovative
technology in conceptual design introduces uncertainties into overall design synthesis and tech-
nology assessment. In this 4th step, a lower and upper margin for variations in each identified
design parameter is defined.

In the first four steps, the design parameter space has been fully described. It will be inves-
tigated for the specific design problem, see figure 3.30. In the following steps of the proposed
framework, uncertainties are introduced to the before defined parameter space.

(5) Allocation of a Continuous Probability Distribution to Each Parameter Space

Each design parameter space is then superposed with a continuous probability distribution,
which allows for computing the probability of occurrence of a specific parameter value. The
continuous probability distribution is given by an analytical probability density function p(x)
that links specific parameter values to a probability of occurrence [195]. The probability that
each design parameter lies within the specified parameter space must equal one. Thus, the
integral of the probability density function between the lower (a) and upper margin (b) must
3.4 Probabilistic Modeling of Innovative Systems 65

1. Step 2. Step
4. Step
Y

3. Step
Y

X X
xmin xmax

Figure 3.30: Definition of design space with the proposed framework.

also equal one accordingly:


Z b
P (a <= x <= b) = p(x) dx = 1. (3.40)
a

If no additional information regarding


the parameter space is available, a simple
rectangular probability distribution can




       


    
be applied that results in equal proba-
bility of each parameter value within the 
   
 
 
parameter space, see figure 3.31. In most
cases, however, an initial estimate of the  

design parameter is available. Hence, the
probability that the specific design para-
meter will result in a value that is close
to the initial estimate is higher, than that
  
the parameter results in a value close to 
  

the outer limits. Symmetric or asymmet-
ric distribution functions can be assigned Figure 3.31: Continuous probability distributions.
to the design parameter, see figure 3.31.
If the initial estimate lies in the middle
of the pre-defined parameter space, the probability distribution is resembled by a symmetric
distribution. For example, the Gaussian normal distribution can be applied [195]:
1 −(x−µ)2
p(x) = √ · e 2σ2 , (3.41)
2π · σ
where µ denotes the mean or expected value, and σ the variance of the probability distribution.
In case of an asymmetric probability distribution, the beta distribution can be applied [196]:
Γ(p + q) (x − a)p−1 · (b − x)q−1
p(x) = · , (3.42)
Γ(p) · Γ(q) (b − a)p+q−1
66 Chapter 3. Aircraft Systems Model

where p and q denote function coefficients, and Γ(r) the gamma function of the form [195]:
Z ∞
Γ(r) = tr−1 · e−t dt. (3.43)
0

Values of the identified design parameter within the allocated design space can vary by several
orders of magnitude24 . For better comparability and application of joint distribution functions,
the parameter space of each design parameter is normalized to a range of 0 to 1:
x−a
p(a <= x <= b) =⇒ p(0 <= x
e <= 1) with x
e= . (3.44)
b−a

(6) Overall Aircraft Design Synthesis and Optimization Parameter Sweep

After the design parameter space has been defined, a design and optimization synthesis is
conducted for a discrete combination of the different identified design parameters. Thus for n
design parameters, an n-dimensional matrix of fully re-sized and optimized aircraft is designed,
see figure 3.32. Each design is optimized towards a specific target, for example for minimum
fuel burn where the parameters wing loading and thrust-to-weight ratio are varied. For sound
technology assessment it is important that each design was fully re-sized and optimized for
each combination of the identified design parameters, so that gains and repercussions from
technology assessment can fully propagate through the design synthesis. Only this way, a fair
comparison of the different derived designs is possible. Since the required computing time grows
exponential with each additional design parameter, a fast and efficient overall aircraft design
synthesis is mandatory, which is available by means of the ILR MICADO environment.


   
      
 
 


       
       

  

 

  
  

  
 
  
  


   

      

 

  
  
  
 

 
   !"   !  ! 

Figure 3.32: Design space and parameter sweep.

24
For example, comparison of system mass and drag coefficient.
3.4 Probabilistic Modeling of Innovative Systems 67

(7) Technology Assessment within Provided Parameter Space

After the design and optimization synthesis is completed, the obtained n-dimensional aircraft
design matrix is assessed against a pre-defined reference aircraft, see figure 3.32. The assessment
is based on different criteria. Different scenarios as well as weighting functions can be applied
to the assessment. Usually, the same parameters are used that were also used within the multi-
disciplinary design optimization. Superposing of each design parameter with an n-dimensional
probability distribution allows for generating an n-dimensional joint probability distribution
of the results (cf. figure 3.32). For the proposed framework it is assumed by the author that
the different design parameters are independent of each other. Since the design parameters
are input for the overall design synthesis, independency of the parameters is assured for the
design process. For independent random variables, the joint distribution function results from
multiplication of the single probability functions [195]:

pjoint (e
x, ye, ...e
n) = p(e
x) · p(e
y ) · ... · p(e
n). (3.45)

For any region R of the n-dimensional design space, the overall probability of the joint distri-
bution function can be computed from the following equation:
Z Z Z
P ((e
x, ye, ...e
n) ∈ R) = ... pjoint (e
x, ye, ...e
n) de
x de
y ... de
n. (3.46)
R

Hence, information regarding the overall probability (P (e x, ye, ...e


n)) whether a certain design
goal can be reached is directly provided by the proposed framework. Integration of innovative
technology is only feasible if specific design goals are reached. The proposed framework allows
for identifying regions within the design space that provide the required design goals. Thus, the
framework concludes in minimum margins for the design parameters. The minimum margins for
each design parameter are fed-back into e.g. system design as design requirement. Chapter 6.2
shows detailed application of the proposed framework to technology assessment of an innovative
morphing high-lift system in the conceptual design phase.
4 Integration into Conceptual Aircraft
Design Synthesis

In this chapter, integration of the proposed systems model into overall conceptual aircraft
design synthesis is described. First, the ILR Multidisciplinary Integrated Conceptual Aircraft
Design and Optimization (MICADO) environment is introduced and a brief overview of the
underlying models is given (see ch. 4.1). Then, emphasis is put on integration of the proposed
systems model into MICADO, and on capturing the repercussions of systems integration (see
ch. 4.2). The chapter closes with presentation of 3 reference aircraft (see ch. 4.3) that are later
used for validation of the proposed model as well as for sensitivity and case studies.

4.1 Conceptual Aircraft Design Synthesis

The proposed systems model is an integral module of the overall MICADO design environment,
which has been developed at the ILR of RWTH Aachen University. MICADO is an overall
design methodology for multi-disciplinary conceptual aircraft design synthesis and optimization.

4.1.1 Design and Optimization Framework

From a defined set of top-level requirements and specification of the configuration, the different
aircraft components are sized to derive the overall configuration, see figure 4.1. Analysis tools
are then used to estimate mass and aerodynamic performance of the design. Based on the
configuration and the estimated performance characteristics, the design mission is simulated to
derive the required total loaded fuel and to check for convergence in maximum take-off weight.
The design loop is run until convergence in the design is achieved. If started from a ’white sheet’,
the design synthesis generally takes less than 6 minutes on a standard desktop computer. After
convergence, the design is assessed based on different criteria such as fuel efficiency, costs, or
emissions. Different scenarios as well as weighting functions are applied in the scope of design
evaluation. Subsequent to assessment of a single aircraft design, a multi-disciplinary design

69
70 Chapter 4. Integration into Conceptual Aircraft Design Synthesis

optimization (MDO) is run in which a defined parameter space is investigated. The overall
design methodology has been fully automized in MICADO. Optimized designs can be generated
with only a minimum of required user inputs and required pre-processing efforts. A more
detailed overview of the methodology as implemented in MICADO is given by Risse et al. [193]
and by the author [197].

 !
"#$ %  $

       


   
  
    
 

 


 


 

 

 
 

   
         

  
 
   " 
 

  "#   $ 

  !

 
 

   

           


       

 

  "#   $ 

##  %     

   


   

  

Figure 4.1: Methodology of ILR MICADO environment.


4.1 Conceptual Aircraft Design Synthesis 71

The main strength of the MICADO environment is its capability to capture variations of top-
level requirements or local parameter changes (such as wing geometry parameters), and to
evaluate the impact on typical parameters such as block fuel or costs. A powerful study manager
tool allows to perform parameter variations and optimizations with individually selected free
variables and objectives. The combined aircraft design and assessment methodology, together
with the flexible software architecture, enables detailed investigations of selected parameter
spaces. It is further characterized by low computational effort using parallelization as well as
by low pre- and post-processing effort for the user.

During the last years MICADE has been used for various aircraft design and operational studies.
Anton et al. [82] showed the impact on changes in top-level requirements on overall design syn-
thesis. Investigations regarding fuel efficiency on multi-stop long-haul missions were published
by Anton et al. [198] and by the author [199]. The ability to assess flying qualities in conceptual
design was shown by Risse et al. [200]. Case studies on assessment of economic parameters were
published by Franz et al. [201] and the author [202]. Sahai et al. [203, 204] showed results on
noise assessment. Operational constraints in design of laminar aircraft were investigated by
the author [205, 197]. Case studies on design and integration of innovative technology and/or
configurations were published by Braun [206] as well as by the author [207, 208, 3, 194].

4.1.2 Underlying Models

The underlying models for estimating mass, aerodynamics, propulsion characteristics, and flight
performance within MICADO are briefly described in the following paragraphs; detailed de-
scriptions can be found in publications by the author and his colleagues1 .

Operating Weight Empty and Component Weights

The operating weight empty is calculated from the estimated weights of the different system-
level components. Semi-empirical methods from literature are used to derive the component
weights. In an ILR master thesis by Berg [95], the different available methods from literature
were validated against known component weights of commercial transport aircraft. The me-
thods that showed the closest agreements have been implemented in MICADO. An adapted
higher-order model by NASA [209] has been implemented for estimating wing and fuselage
mass [210]. This analytical model allows to account for non-standard wing planforms such as
high aspect ratio laminar wings. Both works contain detailed verification and validation of the
implemented ILR mass predictions.
1
The corresponding publications are cited in the sections below.
72 Chapter 4. Integration into Conceptual Aircraft Design Synthesis

Aerodynamics and Flight Polars

An ILR tool is used to estimate trimmed full configuration flight polars of arbitrary aircraft
configurations in subsonic and transonic flight regimes. Results are sensitive to the relevant
design parameters. Detailed description and validation for application to today’s transport
aircraft configurations was published by the author [211]. Total aircraft drag is accumulated
from induced drag, viscous drag and wave drag. A similar approach was, for example, also
suggested by Gur et al. [212, 213].

The DLR multi-lifting-line tool LiftingLine (Release 2.3) [214, 215], which is based on the
potential theory, is used to estimate lift, induced drag, as well as pitching moment. It allows for
analyzing arbitrary non-planar wing configurations. Maximum lift coefficients are estimated
from semi-empirical approaches by Raymer [86, ch. 12]. To account for compressibility, a
Prandtl-Glauert [216, ch. 11.4] correction is applied. Influences of fuselage and engine nacelles
on induced drag are estimated from correction terms as suggested by Roskam [217, ch. 4].

The viscous drag (CD,visc ) is estimated from a component built-up approach and is computed
based on skin friction coefficient (CF ) and form factors (F F ) as described by Raymer [86,
ch. 12]. The skin friction is estimated from the flat plate solution for turbulent flow that
uses local Reynolds numbers (Relocal ) of each component, see equation 4.1. Additionally, lift
dependent viscous drag (CD,visc,lif t ) is estimated by a parabolic correction, which has been
derived during validation of the model [211] as shown in equation 4.2. Taking interference
effects into account, total viscous aircraft drag can then be estimated from equation 4.3.
0.455
CF = (4.1)
(log10 Relocal )2.58 · (1 + 0.144 · M 2 )0.65
CD,visc,lif t = a0 · CLa1 (4.2)
P
CF · F F · Q · Swet
CD,visc = + CD,visc,lif t (4.3)
Sref
Wave drag (CD,wave ) is determined by the difference of flight Mach number (M ) and critical
Mach number (Mcrit ) according to Lock’s approximation [218], see equation 4.4. Further,
the critical Mach number is estimated with the drag divergence Mach number (MDD ) from
equation 4.5. Finally, MDD is estimated from Korn’s equation, which was enhanced to include
wing leading edge sweep (ϕLE ) as shown in equation 4.6, where kA denotes a technology factor
describing the airfoil’s transonic characteristics. Other influences are neglected.
4
CD,wave = 20 · (M − Mcrit ) (4.4)
 1/3
0.1
Mcrit = MDD − (4.5)
80
kA t/c CL
MDD = − − (4.6)
cosϕLE cos2 ϕLE cos3 ϕLE
4.1 Conceptual Aircraft Design Synthesis 73

Engine Performance

A full thermodynamic engine model is used within MICADO for determining the engine per-
formance in various flight conditions. A wide-spread commercial available gas turbine cycle
analysis code called GasTurb [219] is used. Amongst others, it has been used in overall aircraft
design syntheses and performance analyses before by Diedrich et al. [220], Seitz et al. [221, 53],
Hall et al. [222], Vera-Morales et al. [223], Dollmayer [72, 73], as well as by Luedders et al. [224]
. The GasTurb model provides engine parameters such as available thrust or fuel flow as func-
tion of Mach number, altitude, engine power setting, as well as of bleed air and shaft power
off-takes. The available engine performance is hereby limited by e.g. maximum core tempera-
ture. Thus, MICADO uses a detailed engine model for correct modeling of the impact of flight
condition and power off-takes on engine performance and efficiency. Since it is not possible to
model every specific engine for each specific thrust requirement during MDO, a defined baseline
engine is slightly scaled linearly to match the specific thrust requirements. It is important to
choose a baseline engine close to the required one, so that linearly scaling stays valid.

Analysis of Flight Performance and Optimization of Flight Profile

MICADO incorporates a detailed flight performance model [198] used for detailed mission
simulation. The flight path is discretized into sufficiently small mission increments that are
analyzed using Newton’s laws of motion. The aircraft is hereby considered as a point mass.
The formulation of the laws of motion is expanded and written as power equilibrium of the
net power provided by the engines, and the sum of the derivatives of potential and kinetic
energy, see equation 4.7. It can be rewritten for aircraft performance analysis to equation 4.8,
which is also used by Oates [225, ch. 3]. Equation 4.8 states an initial value problem that
is solved by setting specific performance conditions according to the flight condition of each
mission increment. Thrust, drag, and gross weight vary with speed, altitude, and flight time.
Hence, iterative solving is required for each increment. Inputs for the simulation are aircraft
weight at each mission increment, flight polars as function of Mach number and altitude, a
thermodynamic engine model, as well as atmospheric conditions.

dh dT AS
(T − D) · T AS = mA/C · g · + mA/C · T AS · (4.7)
| {z } | {z dt} | {z dt }
power of thrust and drag
rate of change of potential energy rate of change of kinetic energy
 
dh  
1 + T AS · dT AS  = (T − D) · T AS (4.8)
dt  g dh  mA/C · g
| {z }
acceleration factor
74 Chapter 4. Integration into Conceptual Aircraft Design Synthesis

The initial cruise Mach number is a top-level requirement that is not changed during design
synthesis. Hence, the speed profile of the mission is already given by the design requirements.
The altitude profile, however, is unconstrained when neglecting operational requirements e.g. by
the air traffic control. Within the design synthesis, the mission altitude profile is optimized for
minimum fuel burn. It is achieved when flying on the altitude that leads to highest specific air
range (SAR) for a given gross weight [226, ch. 12]. Performance limitations such as minimum
climb performance, of course, have to be considered when optimizing the altitude. In the
scope of this thesis, typical step climb profiles are considered. Optimization of the altitude
profile with the flight performance module leads to mission profiles such as those exemplified
in figure 4.2. Due to the decrease in weight, SAR increases steadily with a constant gradient
for the optimized flight profile. The change in cruise lift coefficient shows a typical saw-tooth
profile, see figure 4.2.

50 1.0 Specific air range, m/kg 400 0.8

40 0.8
Altitude, 1000 ft

Lift coefficient
Mach number

300 0.6
30 0.6

20 0.4
200 0.4
10 Altitude 0.2 SAR
Mach CL
0 0.0 100 0.2
0 1 2 3 4 0 1 2 3 4
Range, 1000 NM Range, 1000 NM
(a) Altitude and Mach number. (b) Specific air range and lift coefficient.

Figure 4.2: Fuel burn optimized mission profile.

Monetary Assessment

Direct operating costs (DOC) are generally used for monetary assessment in conceptual aircraft
design as stated by e.g. Clark [84, ch. 6], Fielding [31, ch. 8], or Jenkinson [227, ch. 12]. DOC
give a direct figure of merit of the aircraft’s cost effectiveness. In the past, various DOC models
have been published (e.g. [228], [229], [230], [231],[232, 233], or [201]). Some of the models are
openly available to academia, others use restricted data of e.g. airlines. A comparison between
different DOC models was published by Gordon [234] and Meyer [235]. At the ILR, a new model
for estimating DOC in conceptual aircraft design has been recently developed [236, 201]. It
consolidates the strengths of the various available models into one. Details as well as verification
of the ILR DOC model was published by Franz and the author [201].
4.2 Impact of the Systems Model on Design Synthesis 75

The ILR MICADO environment also incorporates a tool for estimating aircraft list price as
well as recurring and non-recurring costs. Recurring costs are broken-down onto system-level
within the tool. A NASA model [92] was implemented for estimating recurring costs of the
single systems. Further, the tool allows for estimating the net present value for a given life cycle
scenario of the aircraft. Details and verification of the model were published by the author [202].

4.2 Impact of the Systems Model on Design Synthesis

Repercussions of systems integration on overall aircraft design synthesis have already been
discussed in chapter 1.2. System mass and secondary power off-takes have the main impacts
on design synthesis. This is both by direct impact as well as by snowball effects during design
iterations. The following paragraphs elaborate on capturing changes in system mass and engine
power off-takes within MICADO.

4.2.1 Systems Mass in Design Synthesis

During design synthesis, the proposed systems model runs in sizing mode. Thus, the entire
systems architecture is sized in accordance to the design input parameters of each design itera-
tion. The masses of the single systems are estimated and accumulated to total systems mass.
Accordingly, the centers of gravity are estimated for each system and for the entire architec-
ture. This information is fed-back into the Mass Estimation module (cf. figure 4.1), where the
operating weight empty and the aircraft’s center of gravity is re-estimated from the single air-
frame components and systems. Thus, increase in systems mass directly impacts the operating
weight empty and the weight and balancing of the aircraft. The increase in OWE and resizing
of the empennage lead to increased fuel consumption, and thus also to an increase in maximum
take-off weight. In the next design step, snowball effects from the increase in MTOW (e.g.
higher performance requirements) further impact systems mass until convergence is achieved.

4.2.2 Power Off-Takes in Design Synthesis

The almost linear increase of specific fuel consumption with shaft power and bleed air off-
takes was shown in figure 1.7. Scholz [13, 74] suggests to use semi-empirical corrections for
capturing the impacts of engine power off-takes. He uses equation 4.9 for correction of specific
76 Chapter 4. Integration into Conceptual Aircraft Design Synthesis

fuel consumption; changes in SFC are a linear function of shaft power off-takes (Pshaf t ):

Pshaf t
∆SF C = SF C · Cshaf t, Scholz · , Cshaf t, Scholz = 0.0155 N /W . (4.9)
Neng · SLST

The relation was derived from regression analysis. Only the engine’s SFC, the sea level static
thrust (SLST ) and the number of installed engines (Neng ) is required as input. A similar
correction function is given for capturing the impact of bleed air off-takes on fuel consumption,
eq. 4.10. Changes in fuel flow (ṁf uel ) are a linear function of bleed air off-takes (ṁbleed ) as well
as of the turbine entry temperature (T41 ):

∆ṁf uel = Cbleed, Scholz · T41 · ṁbleed , Cbleed, Scholz = 3.015 · 10−5 1/K . (4.10)

Prange [75] also gives linear functions for estimating the impact of shaft power and bleed
air off-takes on engine performance. He describes the reduction of thrust that is caused by
engine off-takes. None of the mentioned approaches distinguish between different engine types,
operating conditions, or between the engine station at which air is bled from the compressor or
fan flow. All of these characteristics, however, have significant influence on the engine efficiency
loss that is caused by shaft and bleed air off-takes [73, ch. 2].

In MICADO, GasTurb is used for capturing repercussions of secondary power off-takes on thrust
and fuel efficiency in combination with the proposed systems model. Due to time constraints
and modeling effort it is not feasible to run GasTurb calculations during mission simulation for
each mission increment and for each combination of flight condition, required thrust, and shaft
as well as bleed air off-takes. Instead, the author uses polynomial interpolation to estimate
the impact of secondary power off-takes on engine performance. The required coefficients for
interpolation are derived from parameter studies with GasTurb for each specific engine. Specific
engine characteristics, e.g. bypass ratio, core temperature, or the engine station from which
air is bled, are implicitly mirrored in the applied interpolations. This allows for fast mission
simulations and correct modeling of repercussions.

It can be shown with GasTurb that the main characteristic engine performance parameters,
which are impacted by power off-takes, are the fuel flow (ṁf uel ), the specific thrust, and the
engine core temperature (T45 ). Since maximum available engine thrust is limited by maximum
core temperature, an increase in engine core temperature by power off-takes simultaneously
leads to a decrease of maximum available thrust. The three identified parameters are corrected
within the Performance & Mission Analysis module (cf. figure 4.1) for each specific flight
condition by means of linear correction functions. The correction function for fuel flow is
exemplified in equation 4.11. The fitting coefficients (C1 and C2 ) are a function of flight
condition (h, M and ISA) and engine power setting (N1 ):

ṁf uel,corrected = ṁf uel,0 · [1 + C1 (h, M, ISA, N1 ) · Pshaf t + C2 (h, M, ISA, N1 ) · ṁbleed ] . (4.11)
4.2 Impact of the Systems Model on Design Synthesis 77

For each specific engine, six coefficients must be derived from multi-dimensional polynomial
curve fits. Practical use has shown that the derived coefficients differ only slightly between
different engines as long as the technology level is not significantly different, e.g. same bypass
ratio. Hence, the same correction coefficients can be used for today’s turbofan engines. The
correction coefficients that are currently used for common turbofan engines in MICADO, were
derived from a GasTurb model of the General Electric CF6-80C2 engine. In figure 4.3, devia-
tions for two different operating conditions between the corrected fuel flow that is determined
from MICADO (surface), and the one that is directly determined from GasTurb (dashed lines)
are shown. For both operating conditions the deviations are negligibly small for the engine.
Deviations of specific fuel consumption and engine core temperature are of similar order2 .

!" # 
 
 
 
  

$$ 
 %
$  %
 
 
 

 

   
 
    
      
 


 

  

 


 
(a) Sea level, Mach 0.25, 175 kN thrust. (b) 35,000 ft, Mach 0.80, 50 kN thrust.

Figure 4.3: Change in fuel flow for CF6-80C2 engine model in different flight conditions.

Two other turbofan engines haven been additionally used for verification of the applied cor-
rection functions: the CFM56-5B4P and GE90-115B engines. For all three engines, parameter
studies of engine power off-takes were conducted for two different operating conditions, and for
a typical range of shaft power and bleed air off-takes. In figure 4.4, relative deviations between
the applied correction functions in MICADO and GasTurb results are plotted. Results show
relative mean deviation, standard deviation as well as maximum deviation of fuel flow, specific
thrust, and engine core temperature. It can be observed that deviations lie below 2% for all
engines. Largest deviations are oberserved for the GE90-115B; maximum deviation are 1.7%
for that engine. Please note that these maximum deviations occur at the largest feasible power
off-takes. Typical operational off-takes are much lower, which also leads to lower deviations.
Mean deviation for the GE90-115B is 0.7 %. Deviations of all three corrected engine parame-
ters (ṁf uel , SF C, T45 ) are of the same order of magnitude, see figure 4.43 . For application in
conceptual aircraft design, the applied corrections in MICADO show sufficient accuracy.

2
For details see appendix B.1.
3
Detailed results are listed in appendix B.1-B.3.
78 Chapter 4. Integration into Conceptual Aircraft Design Synthesis

2.0
1.6 Mean deviation Standard deviation Max deviation
sea level, 35,000 ft, sea level, 35,000 ft,
Relative Deviation, %

1.2
M = 0.25, M = 0.80, M = 0.25, M = 0.75,
0.8 T = 175 kN T = 50 kN T = 70 kN T = 18 kN
0.4
0.0
-0.4
sea level, 35,000 ftl,
-0.8 M = 0.25, M = 0.83,
-1.2 T = 295 kN T = 65 kN
-1.6 CF6-80C2 CFM56-5B4P GE90-115B
-2.0
mfuel SFC T45 mfuel SFC T45 mfuel SFC T45 mfuel SFC T45 mfuel SFC T45 mfuel SFC T45

Figure 4.4: Deviations in power off-take correction between MICADO and GasTurb for the
CF6-80C2, CFM56-5B4P, and GE90-115B engines.

4.3 Conventional Reference Aircraft

For validation and sensitivity studies of the proposed systems model the author uses a typical
single-aisle and a long-haul aircraft with conventional technology. Top-Level requirements and
specifications follow the Airbus A320-200 and Boeing 777-200ER aircraft. For the sake of
simplification, those reference aircraft will be denoted as A320 and 777 aircraft throughout
this thesis, although they do not exactly represent the Airbus and Boeing aircraft but come
close to their characteristics and performance on a conceptual design level. Additionally, a 185
PAX single-aisle aircraft design is presented that will be used as baseline for the case study in
chapter 6.2.

4.3.1 Reference Aircraft: A320-200

The Airbus A320-2004 is designed for 150 passengers in a typical two-class layout and for
a maximum payload of 18.6 t. Its design mission is specified to a range of 2780 NM and a
payload of 13.6 t, which represents maximum cabin utilization without additional cargo. The
study mission is specified by the author to a range of 700 NM and a seat-load factor of 95 %,
which results in a payload of 12.8 t. The ILR MICADO environment was used for design of the
reference aircraft. Design synthesis resulted in a general arrangement as shown in figure 4.5.
Key design specifications and performance characteristics are summarized in table 4.1.

The reference aircraft design is verified against A320-200 data that was published by Jane’s [99]
as well as by Airbus [5]. The payload-range diagram was computed and compared against the
4
Basic version with a MTOW of 73.5 t that is powered by two CFM-56-5B3 turbofan engines [99].
4.3 Conventional Reference Aircraft 79

(a) Design specifications. (b) Performance characteristics.

MTOW 73 413 kg Design mission Study mission


OWE 41 863 kg
Take-off weight 73 413 kg 61 687 kg
MFW 18 712 kg
Payload 13 550 kg 12 825 kg
Maximum payload 18 633 kg
Mission range 2780 NM 700 NM
Wing loading 599.5 kg/m2
Block time 6.43 h 2.18 h
Thrust loading 0.333
Trip fuel 18 207 kg 7122 kg
Cruise Mach number 0.78
Taxi fuel 315 kg 249 kg
Passengers 150
Block fuel 15 372 kg 4640 kg
ULD devices 9 LD3❂45W
Reserve fuel 2835 kg 2482 kg

Table 4.1: Key specifications and performance characteristics of A320 reference aircraft.

one of the Airbus aircraft [5], deviations are illustrated in figure 4.6. Deviations in main design
parameters are shown in figure 4.7. Equal top-level requirements have been used for design
synthesis, see figure 4.7 (left-hand side). Deviations in MTOW, OWE, and MFW of the refer-
ence aircraft in comparison to the Airbus A320-200 are well below 1 % each. Close agreement
in the performance data can be observed from the payload range diagram. Only at low pay-
loads and long ranges (bottom right corner) slightly higher deviations of the two curves exist.
These, however, are not typical combinations of payload and range for standard operations.
Somewhat larger deviations can also be observed for the required take-off distance (TODR)
and for the required landing distance (LDR). Both parameters are not only sensitive to physi-
cal relations but also to semi-empirical coefficients, e.g. a breaking coefficient. The somewhat
larger deviations in field performance are likely caused by deviations in those semi-empirical
coefficients. Sufficient accuracy for the presented studies can be expected from the MICADO
design methodology.

20
design
Paylaod, 1000 kg

16 point
150 passengers
12

4 MICADO A320
Airbus A320
0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
Range, 1000 NM
Figure 4.5: General arrangement of A320
design. Figure 4.6: Payload-range diagram.
80 Chapter 4. Integration into Conceptual Aircraft Design Synthesis

8
results of design synthesis
on overall aircraft-level +7.0%
6

Rmax,payload
input parameters

LDRdesign
4
Deviations, %

for aircraft design synthesis +3.0%


2
+0.56% +0.57%
0% 0% 0% 0% 0% +0.14% +0.10%
0
W/S T/W
maximum
payload

design
payload

design
range

MTOW

OWE

MFW

Rmax,MTOW

Rferry

TODRdesign
-2 -0.96%

-4
-4.2%
-6

Figure 4.7: Comparison between Airbus A320 and A320 MICADO design.

4.3.2 Reference Aircraft: Boeing 777-200ER

The Boeing 777-200ER5 is designed for 301 passengers in a typical three-class layout and for a
maximum payload of 56.9 t. It is an extended range version with a design range of 6800 NM and
a design payload of 27.7 t, which represents maximum cabin utilization with additional cargo.
The study mission is specified by the author to a range of 4000 NM and a load factor of 90 %,
resulting in a payload of 51.2 t. The ILR MICADO environment was used for design of the
reference aircraft. Design synthesis resulted in a general arrangement as shown in figure 4.8.
Key design specifications as well as performance characteristics are summarized in table 4.2.

(a) Design specifications. (b) Performance characteristics.

MTOW 268 502 kg Design mission Study mission


OWE 138 267 kg
Take-off weight 268 502 kg 256 870 kg
MFW 134 850 kg
Payload 27 700 kg 51 200 kg
Maximum payload 56 940 kg
Mission range 6800 NM 4000 NM
Wing loading 625.3 kg/m2
Block time 14.68 h 8.63 h
Thrust loading 0.287
Trip fuel 102 941 kg 67 310 kg
Cruise Mach number 0.84
Taxi fuel 811 kg 700 kg
Passengers 305
Block fuel 93 047 kg 58 056 kg
ULD devices 30 LD3
Reserve fuel 9083 kg 8554 kg

Table 4.2: Key specifications and performance characteristics of 777 reference aircraft.

Again, the reference design is verified against 777-200ER data that was published by Jane’s [99]
as well as by Boeing [237]. The payload-range diagram was computed and compared against the
5
Extended range version with a MTOW of 267.6 t that is powered by two GE90-85B turbofan engines [99].
4.3 Conventional Reference Aircraft 81

one of the Boeing 777-200ER [237] aircraft, see figure 4.9. Deviations in main design parameters
are shown in figure 4.10; again equal top-level requirements have been used. Deviations in
MTOW and OWE of the reference aircraft are well below 1 % each. Close agreement in the
performance data is also observed from the payload-range diagram6 . This characteristic is
observed for both the Boeing aircraft and the MICADO design. Also for the 777 design,
somewhat larger deviations are evident for the required take-off distance (TODR) and for the
required landing distance (LDR).

60
50

Paylaod, 1000 kg
design
40 point
30 305 passengers
20
10 MICADO 777-200ER
Boeing 777-200ER
0
0 1 2 3 4 5 6 7 8 9 10
Range, 1000 NM
Figure 4.8: General arrangement of 777
design. Figure 4.9: Payload-range diagram.

8
results of design synthesis
6 on overall aircraft-level

TODRdesign
Rmax,payload

input parameters

LDRdesign
4
Deviations, %

for aircraft design synthesis


+1.9%
2 +1.3%
0% 0% 0% 0% 0% -0.37% -0.12% +0.0%
0
W/S T/W
maximum
payload

design
payload

design
range

MTOW

OWE

MFW

Rmax,MTOW

Rferry

-2
-2.1%
-4 -2.9%
-5.6%
-6

Figure 4.10: Comparison between Boeing 777-200ER and 777 MICADO design.

4.3.3 Reference Aircraft: Single-Aisle SA185, 185-PAX

Top-level requirements of this third reference aircraft were specified by Airbus Germany for
academic purposes [238]. It is single-aisle aircraft with conventional configuration and tech-
nology that has close resemblance to an Airbus A321 or Boeing 737-900 aircraft. The aircraft
6
Other than the payload-range diagram of the A320, the one for the chosen 777 model does not comprise a
second kink. The second kink in payload-range diagrams corresponds to maximum fuel load.
82 Chapter 4. Integration into Conceptual Aircraft Design Synthesis

is designed for 185 passengers in a typical two-class cabin layout and for a maximum payload
of 23 t. Its design range is specified to 4000 NM with a design payload of 16.8 t, which represents
maximum cabin utilization but no additional cargo. The study mission has a range of 1000 NM
and is flown with a payload factor of 90 %, thus resulting in a payload of 20.7 t. The aircraft is
denoted as SA185 in the scope of this thesis. Key top-level requirements for design synthesis
with MICADO are listed in appendix C.1.

The A320 and 777 reference aircraft have been designed for a given set of key design parame-
ters such as wing loading and thrust-to-weight ratio. For the SA185 design, such key design
parameters are not known, and are thus derived from design optimization with MICADO. The
objective is minimization of fuel burn on the study mission. For every combination of wing
loading and thrust-to-weight ratio, a full design convergence was run that includes resizing of all
aircraft components and iterative execution of all design and analysis tools. Multi-disciplinary
design optimization resulted in a wing loading of 620 kg/m2 and in a thrust-to-weight ratio
of 0.323, see figure 4.11. Key design specifications as well as performance characteristics are
summarized in table 4.3.

, % -!  
  !
  

+
 ++
++
$'(


+
$ +
 "#
!
 
) *   
   
% % &   

    
 
      
    !



 
 

Figure 4.11: Optimization of wing loading and thrust-to-weight ratio for the SA185 design.

The SA185 reference design is verified against a preliminary aircraft design study that was
published by Werner-Spatz [238]. He used similar top-level requirements (c.f. table C.1) and
derived the design with the well documented and verified PrADO (Preliminary Aircraft Design
and Optimization Program) [239] tool. The PRADO design synthesis, however, was conducted
with a different wing loading and thrust-to-weight ratio than that the ILR MICADO optimi-
zation resulted in. Werner-Spatz did non conclude whether he conducted a design optimization.
For verification of the ILR MICADO design, the author used results that are obtained from
design optimization, as well as results that are obtained with the same W/S and T/W that
was used by Werner-Spatz. Deviations in typical overall aircraft design parameters between
the two ILR MICADO designs and the PrADO one are shown in figure 4.12.
4.3 Conventional Reference Aircraft 83

(a) Design specifications. (b) Performance characteristics.

MTOW 99 748 kg Design mission Study mission


OWE 50 527 kg
Take-off weight 99 748 kg 82 571 kg
MFW 42 973 kg
Payload 16 780 kg 20 730 kg
Maximum payload 23 033 kg
Mission range 4000 NM 1000 NM
Wing loading 620.0 kg/m2
Block time 8.96 h 2.35 h
Thrust loading 0.323
Trip fuel 32 608 kg 11 481 kg
Cruise Mach number 0.80
Taxi fuel 308 kg 294 kg
Passengers 185
Block fuel 28 721 kg 8129 kg
ULD devices 12 LD3❂45W
Reserve fuel 3579 kg 3058 kg

Table 4.3: Key specifications and performance characteristics of the SA185 reference aircraft.

10
results of design synthesis on overall aircraft-level
8
6 5.6% 5.9% +4.9% +4.5%
+3.7% +4.2%
L/Dcruise
mfuselage
mfuel,des
MTOW
Deviations, %

TOFL

LDN
2
0% 0% 0%
0
W/S T/W
maximum
payload

design
payload

design
range

OWE

mwing

mengine

SLST
-2 -1.2%
-4 -3.5% -3.0% -3.5%
-3.8%
-6 -5.3%
input parameters
-8 for aircraft design synthesis
-10

(a) ILR MICADO optimized design.

10
results of design synthesis
8 on overall aircraft-level +6.9%
+5.7%
6
L/Dcruise
mfuselage
MTOW
Deviations, %

4
SLST

LDN

2 +0.3% +0.9%
0.0% 0.0% 0% 0% 0% +0.0%
0
W/S T/W
maximum
payload

design
payload

design
range

OWE

mfuel,des

mwing

mengine

TOFL

-2 -0.7% -0.8%
-4 -3.0% -2.4%
-6
input parameters
-8 for aircraft design synthesis -7.6%
-10

(b) PrADO wing loading and thrust-to-weight ratio.

Figure 4.12: Comparison between ILR MICADO and PrADO design of SA185.
5 Verification and Sensitivities of the
Proposed Systems Model

Verification of models for application within aircraft design is always difficult. On the one hand,
only a few commercial transport aircraft exist on the market that share the same technologies.
Thus, sound statistical verification of the models is not possible. On the other hand, data of
the few existing aircraft is only available to a limited extend. Where global aircraft weight
data such as maximum take-off weight or operating weight empty is published, component
weights of either the airframe structure or single systems are not. The same is true for aircraft
fuel consumption and power requirements of the single systems. Verification of the proposed
systems model is achieved by means of discussion of plausibility of the results and by detailed
sensitivity studies. Further, it has to be kept in mind that a physical modeling approach was
chosen for modeling of the entire system architecture and also for the single systems (cf. ch. 3.1).
The system characteristics of weight and power requirements are derived from requirements of
the dedicated aircraft-level functions. Thus, it is expected that the single system models are
influenced and driven by the correct set of parameters. Furthermore, either available models for
single systems have been directly implemented into the proposed systems model, or available
models have been enhanced and updated prior to their integration. Hence, the proposed systems
model is based on single models that already have been verified in the past; either by the creator
or by the author in the scope of his research (cf. ch. 3.2.2, 3.2.3, or 3.2.4). Verification of the
single system models has already been addressed by means of a bottom-up approach in the
corresponding sub chapters. This chapter discusses propagation of possible errors within the
single system models onto overall system-level and aircraft-level, and assesses overall expected
level of accuracy as well as the main design drivers of the proposed model. Thus, a top-down
approach is chosen here.

For technology assessment in conceptual aircraft design it is generally not required to derive
correct absolute values, Required are correct trends and impacts of changes in the design. For
example, if two designs are assessed against each other, absolute estimated fuel consumption
for a given mission is only of secondary importance. It is by far more important to correctly
describe the relative differences between the two designs. Since a physical modeling approach

85
86 Chapter 5. Verification and Sensitivities of the Proposed Systems Model

was chosen, it can be expected that changes in the design parameters are correctly mirrored
in the systems design. Further, if a baseline aircraft is available for verification of the overall
design1 , also absolute estimated values for derivatives of that baseline aircraft keep validity.

The previous chapter has closed with description of the reference aircraft. Yet, no details re-
garding application of the proposed systems model to the reference aircraft have been presented.
Therefore, results regarding sizing of the systems architecture of the reference aircraft are dis-
cussed in chapter 5.1. Plausibility of the proposed model by means of its application to the
reference aircraft and expected trends in the results for the two aircraft are addressed. Plausi-
bility and feasibility of the proposed model is further addressed by means of sensitivity studies
in chapter 5.2. Sensitivities towards model specific input parameters, towards aircraft design
parameters, and towards top-level aircraft requirements are discussed. Chapter 5.3 concen-
trates on sensitivities of overall aircraft design synthesis on output parameters of the proposed
systems model. The chapter concludes with expected level of accuracy and limitations of the
proposed systems model in chapter 5.4.

5.1 Application to the Reference Aircraft

The Airbus A320 and the Boeing 777 are both aircraft with conventional systems architecture.
The main difference between both aircraft lies in their mission requirements. The A320 is a
single-aisle aircraft that is designed for short to medium range missions. The 777 is a wide body
aircraft that is designed for long range missions. This difference is also mirrored in estimated
systems characteristics that are discussed in the following paragraphs.

5.1.1 System Mass Estimates

The weight break down on overall aircraft-level is shown in figure 5.1 for both aircraft. Fig-
ure 5.1(a) shows relative masses, and figure 5.1(b) shows absolute ones. Both figures comprise
the weight break down into: mass of the airframe structure, mass of the propulsion and landing
gear system, and total systems mass. As aforementioned, mass of the propulsion and landing
gear systems is not estimated within the proposed systems model. it is covered by MICADO.
The proposed model estimates total systems mass to approximately 11.7 t for the A320, and
to 33.3 t for the 777 aircraft. The relative weight break down of both aircraft does not differ
significantly, see figure 5.1(a). The weight break down accumulates to OWE. For both aircraft,
deviation within estimated OWE are lower than 1 %, cf. fig. 4.7 and 4.10. Validated tools are
1
As it is the case for the A320 and 777 designs, cf. ch. 4.3.1 and 4.3.2.
5.1 Application to the Reference Aircraft 87

used within MICADO for estimating overall mass of the airframe structure as well as of the
propulsion and landing gear system. The accuracy of those tools lies within the one expected
in conceptual aircraft design. Hence, also validity of the accumulated total systems mass that
is estimated by the proposed model can be expected.

60 80
A320 A320
70
50 777 777
Relative mass, %

60

Mass, 1000 kg
40
50
30 40
30
20
20
10
10
0 0
e

g
r

ng
ar

on

s
a

m
ur

ur
in

sio
ge

ge
i

lsi
ste

ste
ct

nd

ct

nd
ul

pu
ru

ru
La

Sy

La

Sy
op
St

St

o
Pr

Pr
(a) Relative mass. (b) Absolute mass.

Figure 5.1: Comparison of aircraft weight break down of A320 and 777 MICADO designs.

Since no published data is available


for the A320’s and 777’s total systems
Deviations ILR estimates, %

6 +5.5% +5.7%
+4.7%
mass, the presented estimates are ve- 4
rified against estimates obtained from 2 +1.0%
semi-empirical handbook methods; chap- 0
ter 2.1.1 elaborates on such models. Es- -2 -0.6%
timates by Torenbeek [93] and by Bel- -4
... to Berg
... to Torenbeek -3.6%
tramo [92] have been used for verification -6 ... to Beltramo
of the proposed model. Figure 5.2 shows A320 777
deviations of the different handbook me-
thods relative to total systems mass as Figure 5.2: Deviations of estimated systems mass.
estimated by the proposed model. Max-
imum deviations of the estimated systems weight to the semi-empirical handbook methods are
below ±6 %. Further, no pattern within deviations to the semi-empirical handbook methods
can be observed from figure 5.2.

Figure 5.3 illustrates the system weight break down of the single systems according to the ATA
nomenclature. As before, relative and absolute masses are shown. The most striking difference
between the two aircraft can be observed from the relative mass fractions that are shown in
figure 5.3(a). For the long-range aircraft the relative mass fraction of the cabin and furnishing
system (ATA-25) is significant higher, which is explained by the enhanced cabin equipment
88 Chapter 5. Verification and Sensitivities of the Proposed Systems Model

of long-range aircraft. For the 777 it is estimated to approximately 26 % and for the A320
to 21 %. The furnishing system, however, is not the only system which’s mass fraction is higher
compared to the A320. Higher mass fraction of the furnishing system does not lead to lower
system mass fractions of the other 777’s systems accordingly. The fuel system (ATA-28) and the
lighting system (ATA-33) also show higher mass fractions. For the fuel system it is explained
by the much higher maximum fuel weight (MFW). MFW is estimated to 50 % of MTOW for
the 777, and to 26 % for the A320. This is, of course, also mirrored in fuel system weight.
Higher lighting system weight is explained by both increased weight of external lightning and
especially by increased weight of the cabin lightning. However, influences of the fuel system
and the lighting system on overall systems mass are small.
30
A320
25 777
Relative mass, %

20

15

10

0
21 22 24 25 26 27 28 29 30 33 35 36 49
ATA chapter
(a) Relative mass.

7
A320
6 777
5
Mass, 1000 kg

4
3
2
1
0
21 22 24 25 26 27 28 29 30 33 35 36 49
ATA chapter
(b) Absolute mass.

Figure 5.3: Comparison of system weights of A320 and 777 MICADO designs.

The implemented single system models are based on existing and already verified methods and
models; validity can be assumed. Koeppen [11, ch. 5] has published a comparison between
estimated system masses from his tool and those of the Airbus A320 aircraft. He provides
5.1 Application to the Reference Aircraft 89

relative system masses of: the ECS, the electric system, the FCS, the hydraulic system, and
the landing gear system. This data has been used by the author to show plausibility of the
proposed model. Figure 5.4 shows results obtained from the proposed model, and published
data of Koeppen’s tool as well as A320 weight data. The relative mass fractions are illustrated
in figure 5.4(a), and deviations of the published data from the results of the proposed model
are shown in figure 5.4(b). Deviations of the proposed model to the A320 weight data are
below ±4 %, which again shows close corelation for a conceptual design tool. Interestingly,
results obtained from the proposed model and those of Koeppen’s tool show the same trend
in deviating from the A320 weight data. Weights are either estimated as too low or too high
by both tools. Also, results by Koeppen show an even closer corelation to the A320 weight
data; deviations are approximately half of that of the proposed model. However, Koeppen has
chosen a far more complex modeling approach.

50 6
Deviations ILR estimates, %
Author’s estimate ... to A320 weights
4 +3.7%
40 Koeppen estimate +3.3% ... to Koeppen estimates
Relative mass, %

A320 weights
2
30 +0.7%
0
20
-2
10 -4
-4.0% -3.7%
0 -6
21 24 27 29 32 21 24 27 29 32
ATA chapter ATA chapter
(a) Relative system weights. (b) Deviations of the author’s estimates.

Figure 5.4: Comparison of the author’s estimates with Koeppen’s and A320 weight data [11].

5.1.2 Design and Mission Power Estimates

Total estimated design power requirements of the aircraft systems architecture are plotted in
figure 5.5. They correspond to the estimated design engine off-takes of the aircraft. Please note
that the estimated design power requirements are not typical mission off-takes but maximum
off-takes that govern sizing of the single systems.

The estimated design electric power demand results in 149 kW for the A320, and to 238 kW
for the 777 aircraft. According to Jane’s [99], the A320-family aircraft are equipped with two
engine-driven generators that provide a total installed generator capacity of 180 kW2 . Hence,
2
Hamilton Sundstrand 90 kW constant frequency generators providing 115/200 V three-phase AC at 400 Hz.
90 Chapter 5. Verification and Sensitivities of the Proposed Systems Model

estimated required generator capacity is 17 % lower than the installed one. However, keeping in
mind that the same generators are used for the entire family, it is not suprising that estimated
generator capacity is lower. For the Airbus A320-family, installed generator capacity is defined
by power requirements of the biggest aircraft of the family - the A321 - and not by the smaller
A320. The family concept has not been considered within the proposed model. MICADO has
been re-run with top-level requirements of the Airbu A321 as input. The design resulted in an
aircraft that closely resembles the Airbus A321. Generator capacity is estimated to 184 kW.
Hence, estimated generator capacity is 2 % higher than that of the real aircraft. For the Boeing
777, Jane’s [99] states a total installed generator capacity of 240 kW3 . Thus, deviation of the
estimated generator capacity is neglible. For both aircraft, estimated total electric design power
shows close agreement and feasibility.

For the chosen reference aircraft, increases in


overall power demand of the long-range air- 800 8.0
Hydraulic
craft are mainly mirrored within the bleed

Design bleed air, kg/s


Electric
Design power, kW

air and the hydraulic power demand, see fig- 600 Bleed air 6.0
ure 5.5. For the 777, estimated total hydraulic
design power increases by 450 %, and total de- 400 4.0

sign bleed air increases by 91 % accordingly.


Unfortunately, no reliable information on to- 200 2.0

tal installed hydraulic and bleed air power is


0 0.0
available for the aircraft. However, total re- A320 777
quired bleed accumulates from power require- Aircraft
ments of the ECS and the ice and rain pro-
Figure 5.5: Total estimated design power and
tection system. Both single system models
bleed air (design engine off-takes).
have been described in detail in chapter 3.2.2
and 3.2.3. Sensitivity studies have been con-
ducted that show plausibility of both models. The same holds for total estimated hydraulic
power demand. It is mainly accumulated from power requirements of the FCS and the landing
gear system; for both verified models are used. The increase in required hydraulic power is
caused by required actuation forces of the FCS and the landing gear of the large aircraft.

Estimated design power requirements for the single systems are plotted in figure 5.6. Fig-
ure 5.6(a) shows power requirements of the A320’s systems, and figure 5.6(b) of the 777’s
accordingly. Main consumers of hydraulic energy are the FCS (ATA-27) and the landing
gear (ATA-32), and of bleed air the ECS (ATA-21) and the ice and rain protection system (ATA-
30). For both aircraft, the main consumer of electric energy is the cabin (ATA-25). In case of

3
Hamilton Sundstrand 120 kW variable-speed, constant frequency generator providing 115/200 V three-phase
AC at 400 Hz.
5.1 Application to the Reference Aircraft 91

the A320, the third hydraulic system is pressurized with an electric hydraulic pump. Hence,
also the hydraulic system is a significant consumer of electric energy within the A320.
80 2.0
Hydraulic power

Design bleed air, kg/s


Electric power
Design power, kW

60 Bleed air 1.5

40 1.0

20 0.5

0 0.0
21 22 25 27 28 29 30 32 33 other
ATA chapter

(a) A320.

250 5.0
Hydraulic power

Design bleed air, kg/s


200 Electric power 4.0
Design power, kW

Bleed air
150 3.0

100 2.0

50 1.0

0 0.0
21 22 25 27 28 29 30 32 33 other
ATA chapter

(b) 777.

Figure 5.6: Estimated design power and bleed air for the different consumer systems.

The discussed estimated design power requirements are not typical mission loads, but max-
imum power demand that sizes the single systems. Mission power demand depends on the
specific mission increments. The design mission profile of the A320, as estimated by MICADO,
is plotted below; figure 5.7(a) shows the altitude profile and figure 5.7(b) the flight speeds
accordingly. Further, the particular lift coefficients are plotted in figure 5.7(c), and the thrust
and accumulated fuel consumption is illustrated in figure 5.7(d). A typical step-climb profile is
flown by the aircraft. Figure 5.8 illustrates the estimated engine power off-takes for the mission
profile. Bleed air as well as shaft power off-takes are plotted; shaft power off-takes accumulate
from total electric and hydraulic power demand. Peaks within the shaft power off-takes are
caused by extension or retraction of the high lift system and the landing gear. Switching off the
wing anti-icing in altitudes above 22 000 ft leads to significant reduction of required bleed air
off-takes as shown in figure 5.8. Other variations within engine off-takes result from changing
environmental and flight conditions over the various mission increments.
92 Chapter 5. Verification and Sensitivities of the Proposed Systems Model

40 1.0 500

0.8 400
Altitude, 1000 ft

30

Mach number

Speed, kts
0.6 300
20
0.4 200
Mach
10 TAS
0.2 100
CAS
0 0.0 0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 0.0 0.5 1.0 1.5 2.0 2.5 3.0
Range, 1000 NM Range, 1000 NM
(a) Altitude. (b) Mach and flight speed.

0.7 160 16
Thrust

Used fuel, 1000 kg


0.6 Fuel
Lift coefficient

120
Thrust, kN 12
0.5
80 8
0.4
40 4
0.3

0.2 0 0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 0.0 0.5 1.0 1.5 2.0 2.5 3.0
Range, 1000 NM Range, 1000 NM
(c) Lift coefficient. (d) Thrust and consumed fuel.

Figure 5.7: Simulated design mission of A320.

 

  







    
 

 
 

  
 

 

 


  
  
 
          
! "#$

Figure 5.8: Engine power off-takes of A320 design mission.


5.1 Application to the Reference Aircraft 93

The cruise segments dominate the mission profile. Those are, however, the most ’uninteresting’
mission segments for analyzing estimated system power requirements. In cruise, flight con-
ditions4 , and thus system power demands remain constant. The different climb and approach
segments are much more interesting for understanding system behavior in terms of power de-
mand. Resolution of these flight segments in figure 5.7 and 5.8 is low due to the dominating
cruise segments. Thus for further discussion, mission dependent power demands are not plotted
over mission range or time (x-axis) but over characteristic mission points. Each flight incre-
ment is hereby represented by a mission point, such as shown in figure 5.9 (x-axis) for the A320
mission profile. Please note that the x-axis is not scaled by either range or flight time.

In figure 5.10, power demand of the main con-


sumer systems is plotted for the characteris- 40
Profile
tic mission points. For the ECS (ATA-21) the
Altitude, 1,000 ft
Point
30
bleed air demand during cruise is reduced sig-
nificantly, when compared against the other 20
flight segments, see figure 5.10(a). This de-
10
crease in power demand is caused by the drop
in ambient temperature for cruise altitude. 0
0 2 4 6 8 10 12 14 16 18 20 22 24
Hence, less power is required for cabin cool- Mission Step
ing. Accordingly, the electric power that is
required for ventilation drops as well. Electric Figure 5.9: Characteristic mission points of
power demand of the cabin systems (ATA-25) A320 mission profile.
shows an inverse behavior, see figure 5.10(a).
It peaks and remains constant for cruise flight, which is mainly caused by power demand of
the galleys and the in-flight entertainment system. As shown in figure 5.10(c), the FCS (ATA-
27) requires high peak loads5 but for a short duration during departure and approach. Power
demand of the PFCS is lower and mainly depends on flight speed. Hydraulic power demand
increases during departure until the cruise Mach number is reached. Then, it decreases slightly
with each step-climb due to the decreasing atmospheric pressure, and decreasing dynamic pres-
sure. The ice and rain protection system (ATA-30) requires a constant bleed air mass flow for
anti-icing of the wing leading edges; wing anti-icing is switched off for altitudes above 22 000 ft.
The remaining bleed air demand during cruise is required for constant anti-icing of the en-
gine nacelles, which remains switched on during the entire flight. Further, the system requires
electric power for anti-icing of sensors, ducts, and vents. Power demand increases slightly
with decreasing ambient temperature. The discussed systems behavior shows plausibility and
feasibility of the proposed systems model in estimating the single system power demands in
dependency on the aircraft mission.
4
Constant altitude and speed as well as environmental condition.
5
Deployment and retraction of the high lift system.
94 Chapter 5. Verification and Sensitivities of the Proposed Systems Model

50 2.0 50 2.0
Electric Bleed air Electric
40 1.6 40 1.6

Bleed air, kg/s

Bleed air, kg/s


Power, kW

Power, kW
30 1.2 30 1.2

20 0.8 20 0.8

10 0.4 10 0.4

0 0.0 0 0.0
0 2 4 6 8 10 12 14 16 18 20 22 24 0 2 4 6 8 10 12 14 16 18 20 22 24
departure cruise approach departure cruise approach
Mission Step Mission Step

(a) Environmental control system (ATA-21). (b) Furnishing (ATA-25).

50 2.0 50 2.0
Hydraulic Electric Bleed air
40 1.6 40 1.6
Bleed air, kg/s

Bleed air, kg/s


Power, kW

Power, kW
30 1.2 30 1.2

20 0.8 20 0.8

10 0.4 10 0.4

0 0.0 0 0.0
0 2 4 6 8 10 12 14 16 18 20 22 24 0 2 4 6 8 10 12 14 16 18 20 22 24
departure cruise approach departure cruise approach
Mission Step Mission Step

(c) Flight control system (ATA-27). (d) Ice and rain protection (ATA-30).

Figure 5.10: Mission dependent power off-takes for single A320 systems.

5.2 Sensitivities of the Proposed Systems Model

For sound technology assessment it is important that the applied models capture the most
relevant influences of changes in the design. In the scope of this chapter, the impact of changes
in the various input parameters on the systems model are analyzed and discussed by means of
sensitivity analysis. According to Saltelli [240], sensitivity analysis show the reaction of a model
in its output parameters on changes of its input parameters. Both the direction6 as well as the
strength of the sensitivities of the output parameters are of interest. The theory assumes an
output parameter X, and a model Φ that depends on different input parameters i1 , ..., in [240]:

X = Φ(i1 , ..., in ) . (5.1)

6
Either decrease or increase of the model’s output parameters.
5.2 Sensitivities of the Proposed Systems Model 95

The sensitivity ∆X is defined by the change in one input parameter ∆i1 and the partial
derivative of the model:
∆X = ∂Φ/∂i1 · ∆i1 . (5.2)

Hence, sensitivity is defined by the slope of the sensitivity plots that are presented in the scope
of this section. In case of linear dependency, sensitivity is constant over the entire range of
variations of the input parameter. For non-linear dependencies, sensitivity is not constant but
depends on the specific values of the input parameter. In the scope of this thesis, sensitivity
analysis is used for:

• Identification of the most relevant input parameters: Sensitivities give a direct


measure of influences of the various input parameters. Hence, the input parameters with
the most significant influences on the model can be identified. Further, sensitivities also
show propagation7 of errors from the input parameters to the output parameters. This
information is required for verification of the model.

• Identification of model characteristics: The presented sensitivity analysis helps to


understand main characteristics such as linearities/non-linearities or additivity of the
proposed model. Also, influences within the complex model and of its integration into
the MICADO environment are assessed.

• Assessment of feasibility and plausibility of the proposed model: Also feasibi-


lity and plausibility of the proposed model are shown by means of sensitivity analysis.
Comparison of expected sensitivities and the obtained ones help to evaluate validity of
the proposed model.

The A320 and 777 have been used as reference for the presented sensitivity studies. Results
are only explicitly shown for both aircraft, if they differ significantly. If not, only those of the
A320 are shown in the following sections. Further, the sensitivity studies can be subdivided
into sensitivities of the systems model towards

• model specific input parameters (cf. ch. 5.2.1),

• aircraft design parameters (cf. ch. 5.2.2), and

• top-level aircraft requirements (cf. ch. 5.2.3).

Model specific input parameters are used as input to the proposed systems model but they
are not specific aircraft design parameters. For example, efficiency factors of hydraulic pumps
or electric generators are such parameters. Aircraft design parameters are also direct input to
the proposed systems model. For example, maximum take-off weight or the cabin volume are
7
Amplifying or damping of errors in the results.
96 Chapter 5. Verification and Sensitivities of the Proposed Systems Model

such parameters. The overall aircraft design synthesis is governed by a set of top-level aircraft
requirements. Such top-level aircraft requirements are input to the overall design synthesis and
influence the entire design and all other parameters. Amongst others, the design range or the
design payload are top-level aircraft requirements.

The sensitivity studies concentrate on estimated total system mass as well as on design electric,
hydraulic, and bleed air demand of the overall systems architecture. Mission dependent power
demand is not discussed in the scope of the sensitivity studies. The case study in chapter 6.1
focuses on impact of mission dependent power demand on aircraft performance and efficiency.

5.2.1 Sensitivities Towards Model Specific Input Parameters

In the following section, sensitivities of the proposed systems model towards selected model
specific input parameters are presented.

The air that is bled from the engines is pre-cooled before it is fed to the aircraft’s bleed air
system in common system architectures [32, ch. 6]. This guarantees a constant temperature
and pressure level within the bleed air system. However, exergy is lost by pre-cooling. Fig-
ure 5.11 shows the sensitivities of the proposed systems model to variations of bleed air tem-
perature (Tbleed air ) for the A320 design. Only the bleed air demand is influenced by variations
of Tbleed air . For decreasing Tbleed air , the demand for bleed air mass flow increases, which is
explained by lower energy content of the cooler air. The sensitivity of the model is non-linear
and the same behavior is found for the 777 design. Results also show that the proposed model
neglects impacts on the bleed air system mass. On the one hand, system mass should increase
with increasing Tbleed air due to additional isolation and installation weight. On the other hand,
system mass could also decrease due to decreasing mass flow demand. The proposed model
cannot predict, which influence dominates and which drives bleed air system weight.

In chapter 3.2.2, the author has elaborated on the impact of required cabin temperature (Tcabin )
and recirculation of cabin air on sizing of the ECS. Both parameters are direct inputs to the
proposed model; sensitivities of the model to these parameters are illustrated in figure 5.12
and 5.13. Cabin temperature mainly drives the power demand of the ECS, and thus the bleed
air demand of the aircraft. For lower required cabin temperatures more energy for cooling is
required. Also the required electric power demand for ventilation of the cabin increases slightly.
The sensitivities are hyperbolic; especially for low required cabin temperatures the impact on
bleed air demand is significant. The hyperbolic behavior is explained by the progressive increase
in required energy for cooling to low cabin temperatures. The slight increase in total estimated
system weight is caused by increased system weight of the ECS. Since the ECS only accounts
to approximately 15 % to overall system weight (see fig. 5.3(a)), the impact is of low order.
5.2 Sensitivities of the Proposed Systems Model 97

Variation of the fraction of cabin air that is recirculated within the ECS has the same impact
as variation of cabin temperature, see figure 5.13. A higher fresh bleed air mass flow is required,
if less cabin air is recirculated. However, the impacts are of a lower order of magnitude than
those of variation of cabin temperature, cf. fig. 5.12 and 5.13 (scale of y-axis).

10 40

5 20

∆ results, %
∆ results, %

0 0

-5 System mass -20 System mass


Bleed air Bleed air
-10 Electric power -40 Electric power
Hydraulic power Hydraulic power
-20 -10 0 10 20 -20 -10 0 10 20
∆ Tbleed air, % ∆ Tcabin, %

Figure 5.11: A320 – Sensitivity towards bleed Figure 5.12: A320 – Sensitivity towards re-
air temperature. quired cabin temperature.

20 5.0

10 2.5
∆ results, %

∆ results, %

0 0.0

-10 System mass -2.5 System mass


Bleed air Bleed air
-20 Electric power -5.0 Electric power
Hydraulic power Hydraulic power
-20 -10 0 10 20 -20 -10 0 10 20
∆ recirculation of cabin air, % ∆ span wing anti-icing, %

Figure 5.13: A320 – Sensitivity towards recir- Figure 5.14: A320 – Sensitivity towards span
culation of cabin air. position of wing anti-icing.

Wings of commercial transport aircraft are only protected against ice accretion up to a certain
span position, cf. ch. 3.2.3. Power demand for wing anti-icing increases with increasing leading
edge area that is protected against ice accretion. The relative span position of wing ice protec-
tion is direct input to the proposed systems model. Sensitivities of the model to variations of
the parameter are plotted in figure 5.14. Total bleed air demand increases linearly with relative
span position of wing ice protection. For conventional bleed air anti-icing system, no other
output parameters of the prosed model are influenced.
98 Chapter 5. Verification and Sensitivities of the Proposed Systems Model

The ARINC 600 standard defines an avionic packaging standard that makes use of modular
concept units (MCU) [33, ch. 2]. A MCU comprises a standardized outer dimension as well as
power and cooling requirements of the avionic components. Thus, the MCU standard allows
for efficient design and sizing of the avionic compartments by the airframer. For example,
in today’s commercial transport aircraft the air data and inertial reference system (ADIRS)
comprises approximately 8-10 MCUS [33, ch. 2]. Total number of installed MCUs defines
complexity of the avionic system. Figure 5.15 shows the sensitivities of the proposed systems
model to the number of installed MCUs. It can be seen that the MCUs impact the required
electric power and the bleed air demand. Both impacts are linear and directly proportional
to each other. Additional electronic units require additional electronic power, which again
is converted to parasitic heat loads. Hence, additional cooling power, and thus bleed air is
required. The impact on total electric and bleed air demand is, however, of a relatively low
order of magnitude. Further, it has been shown in chapter 5.1 that the avionic systems have
a comparable small impact on total system mass. Hence, the estimated total system mass is
only slightly influenced by variations in the number of installed MCUs, see figure 5.15.

The required deflection rate of the primary flight control surfaces is the last model specific input
parameter that is discussed in the scope of the here presented sensitivity studies; results are
shown in figure 5.16. An increase in required deflection rate mainly impacts the hydraulic power
demand of the FCS, which increases accordingly. Also the electric power demand increases,
since the A320 design uses an electric-driven hydraulic pump for pressurization of one hydraulic
circuit. Increase in power demand only results in a slight increase in total system mass. The
sensitivities of the model to required deflection rate of the PFCS are linear.

5.0 10

2.5 5
∆ results, %

∆ results, %

0.0 0

-2.5 System mass -5 System mass


Bleed air Bleed air
-5.0 Electric power -10 Electric power
Hydraulic power Hydraulic power
-20 -10 0 10 20 -20 -10 0 10 20
∆ number of MCUs, % ∆ control surface deflection rate, %

Figure 5.15: A320 – Sensitivity towards num- Figure 5.16: A320 – Sensitivity towards de-
ber of modular concept units. flection rate of the PFCS.

The sensitivities of the proposed systems model to the here discussed model specific input
parameters show feasibility in both the model’s reaction as well as in the order of magnitude
5.2 Sensitivities of the Proposed Systems Model 99

of the reaction. The sensitivities for the 777 design are comparable to the ones presented for
the A320 design. Results for the 777 design are shown in appendix D, see figure D.1 to D.6.

5.2.2 Sensitivities Towards Aircraft Design Parameters

Aircraft design parameters describe characteristics of the specific design. They result from the
different tools within overall design synthesis. In the scope of this sensitivity study, variation
of different aircraft design parameters has been conducted and the impact on sizing of the
overall system architecture has been analyzed. Please note that variation of a single aircraft
design parameter directly results in variation of also several other design parameters. For
example, if MTOW increases also OWE, component masses, and fuel consumption increases.
Those interdependencies within design synthesis are, however, not in the focus of the sensitivity
studies. Hence, only a specific parameter is varied and its influence on the system architecture
is analyzed without changes in any other design parameter.

Figure 5.17 and 5.18 show the impact of variations of MOTW and OWE on the system archi-
tecture. MTOW impacts total system mass as well as electric and hydraulic power demand, see
figure 5.17. Sensitivities of the systems model towards MTOW are linear. System mass also
shows linear sensitivity to changes in OWE; power requirements are not impacted by changes
in OWE, see figure 5.18. Both the sensitivities towards MTOW and OWE are of a relatively
low order of magnitude. For changes of 25 % in either MTOW or OWE, output of the system
model only varies by approximately 2 %. It has been shown in table 3.1 and 3.2 that rela-
tively few systems use either MTOW or OWE as input parameter. Hence, their impact on
the overall systems architecture characteristics is small. It is an entirely different picture, if
the entire design synthesis converges and changes in MTOW or OWE propagate through the
design (cf. ch. 5.2.3).

The number of passengers has a strong impact on sizing of the systems architecture. System
mass as well as bleed air and electric power demand increase with the number of passengers, see
figure 5.19. The strong increase in systems mass is caused by higher mass of furnishing and cabin
systems. It has been shown in chapter 5.1 that cabin and furnishing systems hold the largest
shares within overall system mass. Hence, their impact on overall system-architecture-level is
strong. Also, more bleed air is required for air conditioning of the cabin, which is accommodated
by additional passengers. Cabin and furnishing systems contribute to the increase in electric
power demand. The sensitivities show linear behavior.

Sensitivity of the proposed model towards changes in cabin and fuselage length are illustrated
in figure 5.20. Please note that for variation in the number of passengers also cabin length
would have to be adapted. For the results in figure 5.20, the number of passengers has not
100 Chapter 5. Verification and Sensitivities of the Proposed Systems Model

5.0 5.0

2.5 2.5
∆ results, %

∆ results, %
0.0 0.0

-2.5 System mass -2.5 System mass


Bleed air Bleed air
-5.0 Electric power -5.0 Electric power
Hydraulic power Hydraulic power
-20 -10 0 10 20 -20 -10 0 10 20
∆ MTOW, % ∆ OWE, %

Figure 5.17: A320 – Sensitivity towards max- Figure 5.18: A320 – Sensitivity towards ope-
imum take-off weight. rating weight empty.

10 5.0

5 2.5
∆ results, %

∆ results, %

0 0.0

-5 System mass -2.5 System mass


Bleed air Bleed air
-10 Electric powr -5.0 Electric power
Hydraulic powr Hydraulic power
-20 -10 0 10 20 -20 -10 0 10 20
∆ PAX, % ∆ cabin length, %

Figure 5.19: A320 – Sensitivity towards num- Figure 5.20: A320 – Sensitivity towards cabin
ber of passengers. and fuselage length.

been varied. Hence, superposition of the influences on systems architecture design that can
be observed from figure 5.19 and 5.20 must be expected for design synthesis. Changes in
cabin length alone impacts total estimated system mass as well as electric and bleed air power
demand. The increase in system mass is caused by higher mass of the cabin and furnishing
system as well as by increased length of cabling and ducting of the secondary power systems.
The slight increase in electric and bleed air power demand is caused by increasing cabin volume,
and thus by increasing demand of the ECS and the cabin systems. The impact of variation in
cabin length is of smaller order than for variation of the number of passengers.

Finally, sensitivities towards changes in wing geometry are analyzed. Figure 5.21 shows the
results for changes in wing span, where wing reference area has been kept constant. Thus, a
5.2 Sensitivities of the Proposed Systems Model 101

slender wing with higher aspect ratio results from the increased wing span. Total system mass
and electric power demand are hardly influenced. The strongest influence can be seen in the
bleed air demand. It has been shown before that bleed air demand increases with increasing ice-
protected leading edge area, cf. fig. 5.14. The same is observed for higher wing span. Relative
span position of ice-protection remains constant but the absolute position increases, hence the
bleed air demand increases accordingly. Hydraulic power demand decreases, which seems to
be counter intuitive first. However, leakage losses within longer ducting are over compensated
by changes in control surface size, and thus by changes in the required control forces. Control
surface width increases with wing span, but their depth decreases even more, which results in
less control surface area. Thus, less hydraulic power is required by the FCS. However, it has to
be kept in mind that in design synthesis the control surfaces would also be re-sized to maintain
required rudder efficiency.

Figure 5.22 shows variation in wing reference area. Wing aspect ratio has been kept constant,
and thus also wing span increases together with reference area. Also control surface size and the
ice-protected leading edge area increase. Power demand of the systems architecture increases
accordingly. Also hydraulic demand increases due to larger control surfaces, and thus higher
required control forces. Electric power demand increases slightly with the hydraulic demand,
since hydraulic power is partly generated by means of an electric pump. The increase in total
system mass is mainly driven by increased system mass of the FCS.

Again, the different results show feasibility of the proposed model. Results for the 777 design
are summarized in appendix D, see figure D.7 and D.12. No distinguishing differences can be
observed for the long-range aircraft.

10 10

5 5
∆ results, %

∆ results, %

0 0

-5 System mass -5 System mass


Bleed air Bleed air
-10 Electric power -10 Electric power
Hydraulic power Hydraulic power
-20 -10 0 10 20 -20 -10 0 10 20
∆ wing span, % ∆ wing reference area, %

Figure 5.21: A320 – Sensitivity towards wing Figure 5.22: A320 – Sensitivity towards wing
span. reference area.
102 Chapter 5. Verification and Sensitivities of the Proposed Systems Model

5.2.3 Sensitivities Towards Top-Level Aircraft Requirements

Changes in top-level requirements govern the entire design synthesis, and thus influence all
aircraft design parameters, which again leads to multiple impacts on input parameters of the
proposed model. In the scope of this thesis, the following top-level requirements have been
separately varied: design range8 , design payload9 , and design passengers10 . For each variation
of the parameters, the entire design synthesis has been re-run until convergence was achieved.
Further, the resulting MTOW has been plotted in the following sensitivity plots for showing
the impact of changes in design parameters on aircraft-level.

Figures 5.23 and 5.24 show sensitivities towards design range for the A320 and 777. It can be
observed that sensitivities show the same trend for both aircraft. However, impact on the long-
range aircraft is amplified when compared to the single-aisle design. An exponential increase in
MTOW for longer design ranges can be observed. Hence, the entire aircraft increases in terms
of weight and size. Accordingly, the estimated systems mass as well as the power demand of the
systems architecture increase exponential. Superposition of influences of the different aircraft
design parameters that have been discussed in the previous section (cf. ch. 5.2.2) is seen. The
sudden increase of estimated system mass for the A320 design is explained by the implemented
model for estimating furnishing mass. The model distinguishes between furnishing of short
to medium range aircraft and long-range aircraft. For long-range aircraft, furnishing mass is
estimated to be higher in order to satisfy passenger comfort on long-range missions. For the pre-
sented results, transition to long-range cabin design has been set to a design range of 3000 nm,
which corresponds to an increase in mission range for the A320 design of approximately 15 %,
cf. figure 5.23. The sudden increase in furnishing mass, and thus total system mass influences
all other parameters within overall design synthesis.

Figure 5.25 shows sensitivities towards design payload of the A320. With variation of design
payload, the design seat number remains constant. Hence, fuselage and cabin design is not
impacted by changes of design payload. MTOW increases linearly with design payload. Accor-
dingly, estimated systems mass as well as systems power demand increase as well. The strongest
impact is observed for the hydraulic power demand, which is mainly caused by stronger required
actuation forces of the landing gear system and of the FCS. For the presented results, wing
loading of the aircraft design has not been changed. Hence, with increasing MTOW also wing
reference area, and thus the entire wing as well as the control surfaces grow in their size. The
same is true for the empennage and the corresponding control surfaces. This results in strong
increase of hydraulic power demand that can be observed in figure 5.25.

8
Range of design mission.
9
Payload of design mission.
10
Cabin seat number for fuselage and cabin design.
5.2 Sensitivities of the Proposed Systems Model 103

MTOW shows similar sensitivity towards changes in the design passenger number, see fig-
ure 5.26. Other than before, variation of the design passenger number impacts also fuselage
and cabin design, which explains the stronger sensitivities of system mass as well as of electric
power and bleed air demand. The impact of changes in cabin design has been discussed in
the previous section, cf. ch. 5.2.2. While figures 5.19 and 5.20 have shown the sole influence of
the particular parameters, the here presented sensitivities are superposed by changes in over-
all design that result from overall aircraft design synthesis. Hydraulic power demand mainly
increases due to the increase of wing and empennage size. Total system mass as well as the
electric power and bleed air demand show an even stronger increase due to the re-design of the
fuselage and cabin.

Sensitivities of the 777 design to the design parameters are provided in appendix D, see fig-
ure D.13 and D.14; the same trends as for the A320 can be observed.

20 20

10 10
∆ results, %

∆ results, %

0 0
MTOW MTOW
-10 System mass -10 System mass
Bleed air Bleed air
-20 Electric power -20 Electric power
Hydraulic power Hydraulic power
-20 -10 0 10 20 -20 -10 0 10 20
∆ Rdesign, % ∆ Rdesign, %

Figure 5.23: A320 – Sensitivity towards de- Figure 5.24: 777 – Sensitivity towards de-
sign range. sign range.

10 10

5 5
∆ results, %

∆ results, %

0 0
MTOW MTOW
-5 System mass -5 System mass
Bleed air Bleed air
-10 Electric power -10 Electric power
Hydraulic power Hydraulic power
-20 -10 0 10 20 -20 -10 0 10 20
∆ mpayload,design, % ∆ PAX, %

Figure 5.25: A320 – Sensitivity towards de- Figure 5.26: A320 – Sensitivity towards de-
sign payload. sign passenger number.
104 Chapter 5. Verification and Sensitivities of the Proposed Systems Model

5.3 Sensitivities of Overall Design Synthesis Towards the


Proposed Systems Model

The previous chapter has focused on sensitivities of the proposed systems model towards differ-
ent input parameters. No sensitivities of aircraft design synthesis towards the proposed systems
model have been discussed. Repercussions of systems integration, which have been qualitatively
described (cf. ch. 1.2), are now quantitatively presented for the two reference aircraft.

Figure 5.27 shows sensitivities of the A320 design towards variation of estimated system mass.
Results in figure 5.27(a) show the direct impact of changes in system mass without re-sizing
of the design; no design synthesis has been run. Results in figure 5.27(b) also incorporate re-
sizing of the aircraft, and thus snowball effects from design synthesis. Variation of system mass
impacts OWE; a linear dependency can be observed. The increase in OWE leads to higher
fuel burn, and thus to an increase of required block fuel. This propagates to higher MTOW
of the aircraft. Sensitivities on aircraft-level towards the system mass are linear if no re-sizing
is conducted, see figure 5.27(a). If the entire design synthesis is re-run in order to achieve
design convergence, results are further impacted by snowball effects, see figure 5.27(b). For
the same variation of total system mass, the impact on aircraft-level increases by a factor of
approximately 2. Further, sensitivities are no longer linear. Where non-linearity is rather small
for OWE, it is noticeable for estimated block fuel and MTOW. Re-sizing of the aircraft also
impacts estimated design bleed air and shaft power off-takes. Design bleed air demand shows a
nearly linear behavior, it is non-linear for the design shaft power off-takes. Sensitivity of design
shaft power demand increases significantly with increasing total system mass, and thus with
global aircraft weights.

5.0 5.0

2.5 2.5
∆ results, %

∆ results, %

0.0 0.0
OWE
-2.5 -2.5 Block fuel
OWE MTOW
-5.0 Block fuel -5.0 Bleed air
MTOW Shaft power
-20 -10 0 10 20 -20 -10 0 10 20
∆ total system mass, % ∆ total system mass, %
(a) Without design synthesis. (b) With design synthesis.

Figure 5.27: A320 – Sensitivity towards total system mass.


5.3 Sensitivities of Overall Design Synthesis Towards the Proposed Systems Model 105

Figure 5.28 shows a further break


6 Bleed air
down of design shaft power demand Electric
into electric and hydraulic power de- 4 Hydraulic

∆ results, %
mand. It can be observed that the 2
non-linear sensitivity of design shaft 0
power demand is mainly driven by -2
a strong increase of hydraulic power
-4
demand. The increase in hydraulic -20 -10 0 10 20
∆ total system mass, %
power demand is again caused by
increasing aircraft weights and air- Figure 5.28: A320 – Sensitivity towards total system
craft size, cf. ch. 5.2. Bleed air and mass within design synthesis.
electric power demand show a linear
behavior. Further, the visual dip in estimated hydraulic power demand (see top right-hand side
of fig. 5.28) is explained by changes in the landing gear configuration of the A320 design.

The 777 design shows the same sensitivities towards changes in total system mass, see fig-
ure 5.29. However, for re-sizing of the aircraft, the impact of higher system mass on overall
aircraft-level is more severe than for the single-aisle aircraft, see figure 5.29(b).

5.0 5.0

2.5 2.5
∆ results, %

∆ results, %

0.0 0.0
OWE
-2.5 -2.5 Block fuel
OWE MTOW
-5.0 Block fuel -5.0 Bleed air
MTOW Shaft power
-20 -10 0 10 20 -20 -10 0 10 20
∆ total system mass, % ∆ total system mass, %
(a) Without design synthesis. (b) With design synthesis.

Figure 5.29: 777 – Sensitivity towards total system mass.

Variation of bleed air demand directly impacts required block fuel and MTOW accordingly,
see figure 5.30(a). OWE is not directly influenced by variation of bleed air demand. Both
required block fuel and MTOW show a linear dependency. For re-sizing of the aircraft, and
thus design convergence, snowball effects also impact system mass and shaft power demand,
see figure 5.30(b). These impacts are, however, of low order. The strongest impact of variation
of bleed air demand can be observed for required block fuel and take-off field length (TOFL).
The strong impact of bleed air off-takes on available thrust, and thus on aircraft performance
106 Chapter 5. Verification and Sensitivities of the Proposed Systems Model

and efficiency will be further elaborated on in the scope of the case study that is presented in
chapter 6.1. The results that are illustrated in figure 5.30(b) show linear sensitivity of overall
design synthesis on variation of bleed air demand.

5.0 5.0

2.5 2.5
∆ results, %

∆ results, %
0.0 0.0
OWE
Block fuel
-2.5 -2.5 MTOW
OWE System mass
-5.0 Block fuel -5.0 Shaft power
MTOW TOFL
-20 -10 0 10 20 -20 -10 0 10 20
∆ total bleed air demand, % ∆ total bleed air demand, %
(a) Without design synthesis. (b) With design synthesis.

Figure 5.30: A320 – Sensitivity towards total bleed air demand.

Impacts of variation of shaft power off-takes are of a smaller order than variations of bleed air
demand, see figure 5.31(a). An increase in shaft power off-takes directly influences estimated
system mass, which increases accordingly. This again has a slight impact on OWE, block fuel,
and MTOW. If the design synthesis is re-run, only a slight amplification of the sensitivities can
be observed, see figure 5.31(b). Impact on bleed air demand is not noticeable, and thus not
plotted in figure 5.31(b). Also aircraft performance and required TOFL are hardly impacted
by variations of shaft power off-takes.

5.0 5.0

2.5 2.5
∆ results, %

∆ results, %

0.0 0.0
OWE
-2.5 OWE -2.5 Block fuel
Block fuel MTOW
-5.0 MTOW -5.0 System mass
System mass TOFL
-20 -10 0 10 20 -20 -10 0 10 20
∆ shaft power demand, % ∆ shaft power demand, %
(a) Without design synthesis. (b) With design synthesis.

Figure 5.31: A320 – Sensitivity towards total shaft power demand.


5.4 Expected Level of Accuracy 107

The same sensitivities towards bleed air demand and shaft power off-takes are observed for the
777 design, see figure D.15 and D.16. Again, due to the longer design mission the impact on
block fuel, and thus on the entire design synthesis is stronger than for the single-aisle aircraft.

The proposed systems model also impacts the aircraft’s overall center of gravity, and thus the
weight-and-balancing of the aircraft. Both, wing allocation and empennage sizing are impacted
within overall design synthesis by shifts in the estimated aircraft’s center of gravity. Sensitivity
of overall design synthesis towards the systems architecture’s center of gravity are, however,
of a smaller order than those towards systems mass and system power demand. Therefore,
sensitivity towards the systems architecture’s center of gravity will not be discussed further in
the scope of this thesis.

5.4 Expected Level of Accuracy

Two different sources of error must be considered for assessing the expected level of accuracy of
the proposed systems model: (1) The single implemented system models contain simplifications,
and thus inaccuracies in their output parameters. (2) The different input parameters of the
proposed systems model also contain errors that propagate through the model. In the scope of
this failure assessment, the output parameters of the systems model11 are considered separately.
Adjacently, limitations of the proposed systems model are briefly discussed.

5.4.1 Expected Accuracy of Overall System Mass

As already discussed in the introduction to this chapter, it is not easy to verify models in
conceptual design and to assess the accuracy due to limited available data sources. For the
estimated mass on system-level, maximum deviations to averaged weight data of the Air-
bus fleet were shown to be below 3 % for the PFCS (cf. fig. 3.26(a)), and below 7 % for the
SFCS (cf. fig. 3.26(b)). Deviations in estimated weight of single components and sub-systems of
the A320 high lift system are even of an smaller order of magnitude (cf. fig. 3.27). Deviations to
single system weights that have been published by Koeppen [11] are in the same order of mag-
nitude; they do not exceed 4 % (cf. fig. 5.4). Estimated overall system mass can only be verified
against results from different semi-empirical handbook methods. Verification shows maximum
deviation of 6 % for the proposed model towards the different handbook methods (cf. fig. 5.2).
These examples show that single system masses as well as overall system mass are estimated
with an accuracy of 5 to 10 %. The expected level of accuracy is thus within the desired range
11
Overall system mass, bleed air demand, and shaft power off-takes.
108 Chapter 5. Verification and Sensitivities of the Proposed Systems Model

for application of the proposed model in conceptual aircraft design.

As aforementioned, propagation of errors of the different input parameters of the model must
be considered. It has been shown in figure 4.7 and 4.10 that deviations in characteristic design
parameters of the two reference aircraft are below 7 %. Maximum deviations, however, were
found for the estimated take-off field length and for the landing distance. Those performance
parameters are no input parameters for the proposed systems model. Deviations in the other
characteristic parameters lie below 2 %.

According to Montgomery and Runger [195], failure propagation for a function y = f (xi ) is
given by accuracy or standard deviation (σ) of the single input parameters (xi ), and the partial
derivation of the function towards the specific input parameter as follows:

s 2  2  2
∂f ∂f ∂f
σy = · σx21 + · σx22 + ... + · σx2n , (5.3)
∂x1 ∂x2 ∂xn

∂f
where the partial deviations of the function ( ∂x i
) correspond to the sensitivities12 of the pro-
posed model that have been discussed in chapter 5.2. Different input parameters with different
sensitivities of the proposed systems model influence the estimated overall system mass. Hence,
the expected propagation of errors within the proposed systems model is approximated by av-
eraged values:
 ¯ 
√ ∂f
σy = ninput parameters · · σ¯x , (5.4)
∂x

¯ denotes the mean sensitivity and σ¯ the mean error in the input parameters.
where ∂f
∂x x

Tables 3.1 and 3.2 list the most relevant input parameters of the proposed systems model for
estimating overall system mass. The model uses approximately 40 different input parameters for
estimating overall systems mass from the single system models; thus ninput parameters ≈ 40. The
strongest sensitivity of overall system mass has been found towards the number of passengers,
cf. fig. 5.19 and 5.26. Here, failures in the input parameter mitigate to failures in the results
that are damped by a factor of 0.513 . Sensitivities towards most other input parameters are
smaller, thus the author assumes a mean sensitivtiy of 0.2 for the input parameters. Together
with the expected accuracy of the different input parameters of 2 %, equation 5.4 yields an
expected error in the results due to error propagation of approximately 2.5 %. This must be
considered as the lowest boundary of accuracy of the proposed systems model.

12
Slope of the sensitivity plots.
13
For an variation of the input parameter of 20 %, the overall system mass varies by approximtaly 10 %.
5.4 Expected Level of Accuracy 109

5.4.2 Expected Accuracy of Bleed Air Demand

For verification of system power off-takes even less data for sound verification of the model
exists. Bleed air demand accumulates from energy demand of the ECS and the ice and rain
protection system in case of conventional systems architectures. It has been shown in figure 3.13
that estimated maximum air conditioning pack flow only deviates by 3.3 % from the A320 ACPs.
For verification of the ATA-30 bleed air demand no data exists. However, it has to be kept
in mind that both system models (ATA-21 and ATA-30) purely use functional relations for
estimting bleed air demand. Also, the ATA-30 model is based on a validated and proven model
by SAE [169]. Hence, the author expects that total bleed air demand is also estimated with an
accuracy of 5 to 10 %, which again lies within the desired range for application of the proposed
systems model in conceptual aircraft design. Total estimated bleed air demand depends on
approximately 20 input parameters, see tables 3.1 and 3.2. Further, estimated total bleed air
demand shows its strongest sensitivity towards the required cabin temperature, see figure 5.12.
Here, failures in the input parameters mitigate to failures in the results that are amplified
by an factor of approximately 214 . Again, sensitivities towards most other input parameters
are of a smaller order, but still higher than for the estimated total system mass. Hence, the
authors approximates mean sensitivity of bleed air demand to 0.4. According to equation 5.4,
the expected error in the results due to error propagation results in approximately 3.6 %, which
again is considered as the lowest boundary of the model’s accuracy in estimating bleed air
demand.

5.4.3 Expected Accuracy of Shaft Power Off-Takes

Estimated shaft power off-takes accumulate from the total electric and hydraulic demand of the
various aircraft systems. For the electric demand, deviations of 17 % and of 1 % towards installed
generator capacity of the A320 and the 777 were identified in chapter 5.1.2. The higher deviation
for the A320 aircraft has been explained with the over-sized generators, which are required to
faciliate the A320 family concept. For hydraulic demand no data for verification exists. For
the conventional systems architecture, total hydraulic demand results from the demand of the
FCS and the landing gear system. Again, it must be emphasized that both system models
rely on functional relations for estimating power demand. Hence, the author expects that total
hydraulic and electric demand, and thus also accumulated shaft power off-takes, are estimated
with an accuracy of 5 to 10 %. Estimated shaft power off-takes rely on approximately 40 input
parameters, see tables 3.1 and 3.2. The electric demand shows its strongest sensitivity towards
the number of passengers, see figure 5.26. Failures in the input parameters mitigate with an
14
For an variation of the input parameter of 20 %, the overall system mass varies by approximtaly 40 %.
110 Chapter 5. Verification and Sensitivities of the Proposed Systems Model

factor of approximately 0.5. Hydraulic demand shows its strongest sensitivity towards changes
in aircraft weight and aircraft size, see figures 5.23 and 5.24. Here, superposition of varies
influences results in a sensitivity of the hydraulic demand of approximately 1.5. In average, the
author assumes a mean sensitivity for both the electric and hydraulic demand of 0.4 Thus, the
expected error in the results due to error propagation results in approximately 5 %.

5.4.4 Limitations of the Proposed Systems Model

The previous paragraphs have shown that the expected level of accuracy of the proposed model
lies within 5 to 10 %. For application in conceptual design, this accuracy is more than suffi-
cient. Still the expected accuracy depends on different factors that are further discussed in the
following paragraphs. The level of accuracy and especially propagation of errors in the input
parameters has been assessed by means of aircraft designs that feature conventional technology.
Further, with the Airbus A320 and the Boeing 777-200ER baseline aircraft exist for verification
and scaling of the overall design synthesis that resulted in the reference aircraft. Thus, scaling
of the overall design synthesis was possible.

For verified and scaled designs, the expected


errors in the input parameters lie below 2 % as
Propagation of errors, %

12 Error in system mass


aforementioned. If verification of the overall Error in bleed air
Error in shaft power
aircraft design is not possible, e.g. if no base- 9

line aircraft exists, errors in the input para- 6


meters may be greater. Figure 5.32 shows the
impact of inaccuracies of the input parame- 3
reference of
ters on error propagation within the model; 0
failure assessment

a strong impact can be observed. For exam- 1 2 3 4 5


Error of input parameters, %
ple, if errors in the input parameters increase
to 5 %, errors propagate to an increase in Figure 5.32: Impact of inaccuracy of input
failures of estimated system mass from 2.5 % parameters.
to approximately 6 %. The increase is even
worse for bleed air demand and for the shaft power off-takes. Thus, although the proposed
systems model does not influence the accuracy of its input parameter, it still strongly impacts
the expected accuracy of the results. Highest accuracy in the input parameters is achieved, if
baseline aircraft exists for verification and scaling.

Further, the afore derived expected level of accuracy is only valid for aircraft that feature
system technology and systems architectures that are mirrored within the implemented single
system models. If technology of the aircraft deviates from the implemented models, errors in
5.4 Expected Level of Accuracy 111

the results will be of a higher order of magnitude. For example, close agreement between the
estimated FCS weight of the A320 was achieved. For an Airbus A300 or A310 larger deviations
must be expected since those aircraft feature mechanical flight controls. One possibility to cope
with different technologies is to enhance the existing models by implementation of additional
system models. As discussed in chapter 3.3, the flexible and modular structure of the proposed
methodology allows for fast and easy adaption. This approach, however, is only feasible if
models for the specific technology exist. This is generally the case for older technology. For
innovative technology this approach is not feasible. A different approach to cope with uncer-
tainties that arise from innovative technology has been presented in chapter 3.4. The case study
in chapter 6.2 will show application of the proposed methodology to a specific design case.

Also maximum range of validity of the implemented model must be considered. The proposed
systems model has been implemented for application to commercial transport aircraft, and the
sensitivity studies have shown that parameter variation of ±25 % still leads to valid results for
a single-aisle and for a long range aircraft. However, the model may reach its limits for larger
variations. For example, for application to design synthesis of a business jet or to a future
extreme large airliner (1000+ PAX) validity is not assured. For application to such ’extreme’
designs, the results must be critically checked in regards to feasibility.

Figure 5.23 and 5.27(b) have shown discontinuities in some of the output parameters that result
from predicted changes in configuration of the aircraft or the systems architecture. Although
trends from those changes in configuration show feasibility in the results, results must be
critically assessed in the near proximity of those discontinuities. For example, it is certainly
not correct that overall systems mass increases by 20 % because of a change of the aircraft’s
design range from 2950 NM to 3050 NM (cf. fig. 5.23). It is, however, correct that furnishing
mass would significantly increase at some point if the aircraft design implies a change from a
typical single-aisle aircraft to a long-range aircraft.

Another source of deviations of the results of the proposed systems model to existing aircraft lies
in the airframer’s concept of aircraft families and possible derivatives. Implemented functional
relations for deriving performance requirements of the different systems from the aircraft-level
functions only consider the specific design and no other aircraft family members. Results do
not account for possible derivatives of an aircraft, e.g. by stretching or shrinking. Often, all
members of an aircraft family, such as the A320-family, feature the same system components for
reducing complexity and increasing commonality for MRO activities. For example, all A320-
family aircraft use the same PCU on both the slat and flap system despite different airloads.
The proposed systems model, however, correctly estimates different power requirements of the
PCUs, and thus implies different sizings accordingly. Hence, when comparing results against
existing aircraft and also for re-designing existing aircraft, results must be critically analyzed.
6 Case Studies

The last chapter has concluded with plausibility and validity of the proposed systems model and
its integration into conceptual aircraft design. This chapter presents two case studies that show
the benefits from integration of the proposed systems model into conceptual aircraft design. The
first case study concentrates on the impact of engine power off-takes on aircraft performance and
efficiency (see ch. 6.1). It underlines the author’s motivation of capturing systems power demand
already in conceptual design. The second case study focuses on integration of innovative systems
technology into the proposed systems model (see ch. 6.2), and thus concludes with propagation
of repercussions that arise from system integration to overall aircraft-level for sound technology
assessment. Integration of an innovative morphing leading edge device is presented. Further,
the case study concludes with probabilistic assessment of possible aircraft fuel savings and
improvements in aircraft economics. It shows the application of the author’s framework for
modeling of innovative system technology that has been presented in chapter 3.4.

6.1 Impact of Engine Power Off-Takes on Aircraft


Performance and Efficiency

In conceptual aircraft design engine power off-takes (shaft power and bleed air) are mostly
neglected, although the impact on design synthesis and on design optimization is significant.
Repercussions of system integration have been introduced in chapter 1.2 and 4.2. In the scope
of this case study, the identified repercussions will be quantitatively confirmed for overall design
synthesis of the A320 reference aircraft, cf. ch. 4.3.1. Focus is put on impact of engine power
off-takes:

• on aircraft performance, and thus on the required take-off field length,

• on aircraft efficiency, and thus on mission fuel burn, and

• on aircraft design synthesis and on multi-disciplinary optimization.

113
114 Chapter 6. Case Studies

6.1.1 Impact on Take-off Performance

Engine power off-takes during take-off reduce the available thrust significantly, and thus in-
creases the required take-off field length. Take-off field length is an important top-level aircraft
requirement that must be met by the corresponding aircraft design. Figure 6.1 shows the im-
pact of engine power off-takes on take-off field length. The straight lines show the dependency
of TOFL on airport altitude for a take-off gross weight that equals MTOW. The dashed and
dotted lines show the dependency for lower take-off gross weights accordingly. It can be ob-
served that TOFL increases exponential with airport altitude and with take-off gross weight.
The straight lines in figure 6.1 show TOFL with estimated bleed and shaft power off-takes of the
A320 reference aircraft. The diamond-symboled lines show TOFL without bleed air off-takes
but still with estimated shaft power off-takes. It can be easily observed that TOFL decreases
significantly if no bleed air off-takes are considered. For example, estimated TOFL decreases
by approximately 11.5 % (cf. fig. 6.1: from 1.0 to 0.885) for the design case1 . The squared-
symboled lines in figure 6.1 show TOFL without any engine power off-takes. Differences to the
estimtated TOFL without only bleed air off-takes can be neglected. Hence, impact of shaft
power off-takes on available thrust is also neglectable.

with off-takes (MTOW)


1.6
with off-takes (0.83·MTOW)
TOFL, normalized

1.4 with off-takes (0.74·MTOW)


w/o bleed off-takes (MTOW)
1.2
w/o bleed off-takes (0.83·MTOW)
1.0 w/o bleed off-takes (0.74·MTOW)
w/o off-takes (MTOW)
0.8 w/o off-takes (0.83·MTOW)
w/o off-takes (0.74·MTOW)
0.6
0 2 4 6 8 10 12
Altitude, 1,000 ft

Figure 6.1: Impact of engine power off-takes on TOFL for ISA +15 conditions.

Changes in estimated TOFL for a take-off gross weight of MTOW, in comparison to the design
case with bleed air and shaft power off-takes, are shown in figure 6.2. The decrease in TOFL
of 11.5 %, if no bleed air off-takes are considered, is easy to observe. If also no shaft power
off-takes are considered, TOFL decreases by additional 0.5 %. Further, it can be observed from
figure 6.2 that changes in TOFL increases with increasing airport altitude.

1
0 ft airport altitude and gross weight that equals MTOW.
6.1 Impact of Engine Power Off-Takes on Aircraft Performance and Efficiency 115

The impact of engine power off-takes on TOFL is caused by a reduction of available thrust
during take-off. Maximum available thrust for take-off engine setting is illustrated in fig-
ure 6.3 for the reference A320 design with power off-takes (straight line) as well as for the A320
without bleed air off-takes (diamond-symboled line), and the one without any engine power
off-takes (squared-symboled line).

Figure 6.3(a) shows available take-off


thrust in dependency on the flight Mach
-11.5
number. For the turbo-fan engine, the

∆ TOFL, %
typical decrease in available thrust with -12.0

increasing flight speed is observed. If -12.5


no bleed air off-takes are considered, -13.0
available thrust increases by approxima- w/o bleed off-takes (MTOW)
-13.5 w/o off-takes (MTOW)
tely 4.7 % for a Mach number of zero,
0 2 4 6 8 10 12
and by approximately 7.1 % for a typi- Altitude, 1,000 ft
cal cruise Mach number of 0.8, see fig-
ure 6.3(b). Again, the impact of shaft Figure 6.2: Changes in TOFL due to off-takes.
power off-takes on avaible thrust is ne-
glectable. The impact of engine power off-takes on available thrust explains the increase in
required TOFL. Also other aircraft performance characteristics, e.g. minimum climb gradients
and maximum cruise Mach number, are impacted by the reduction in available thrust. The
significants of the impact on those performance requirements will be further addressed in the
scope of the design synthesis and optimization (cf. ch. 6.1.3).
T-O Thrust, normalized

w/o bleed off-takes


1.0 7.0
∆ T-O Thrust, %

w/o off-takes
0.9 6.5

0.8 6.0

0.7 with off-takes 5.5


w/o bleed off-takes
0.6 5.0
w/o off-takes
0.0 0.2 0.4 0.6 0.8 0.0 0.2 0.4 0.6 0.8
Mach Mach
(a) Impact on t-o thrust. (b) Changes in t-o thrust.

Figure 6.3: Impact of engine power off-takes on available take-off thrust.


116 Chapter 6. Case Studies

6.1.2 Impact on Fuel Efficiency

Mission fuel efficiency is often measured by means of specific fuel consumption (SFC)2 . It de-
scribes efficiency of an engine in regards to the delivered thrust. It is not constant over the
mission but varies with flight conditions. Figure 6.4(a) shows SFC of the A320 reference design
(straight line) for the design mission. Peaks in SFC represent phases of higher thrust settings or
idle setting e.g. in climb or in final approach. For the cruise segments, SFC averages to appro-
ximately 16.5 mg/Ns if engine power off-takes are considered. Again, the diamond-symboled
line shows the resulting SFC if no air is bled from the engines. SFC decreases significantly
by approximately 5.5 % in comparison to the reference design. SFC is further reduced slightly
if also no shaft power off-takes are considered (squared-symboled line). However, bleed air
off-takes have the most significant impact.

The impact on SFC is also observed from accumulated fuel consumption that is plotted in
figure 6.4(b) over the design mission. Peaks in SFC propagate as steps to consumed fuel.
Cruise phases dominate fuel consumption, which are shown as segments of almost constant
gradients. Lower SFCs propagate to lower gradients of fuel consumption in figure 6.4(b). If
no bleed air off-takes are considered, fuel consumption is reduced by 5.1 % for the reference
aircraft, which equals approximately 714 kg of fuel on the design mission. Fuel consumption is
reduced by another 0.7 % if also no shaft power off-takes are considered. The significance of the
results for overall design synthesis and design optimization will be disucssed in the following
section.

20
with power off-takes 14 with power off-takes
Consumed fuel, 1,000 kg

w/o bleed air w/o bleed air


19 w/o power off-takes 12 w/o power off-takes
SFC, mg/Ns

10
18
8
17 6
4
16
2
15 0
0.0 0.4 0.8 1.2 1.6 2.0 2.4 2.8 0.0 0.4 0.8 1.2 1.6 2.0 2.4 2.8
Mission range, 1000 NM Mission range, 1000 NM
(a) Specific fuel consumption. (b) Fuel consumption.

Figure 6.4: Comparison of fuel efficiency with and without engine power off-takes.

2
Specific fuel consumption is defined by the quotient of thrust and fuel flow.
6.1 Impact of Engine Power Off-Takes on Aircraft Performance and Efficiency 117

6.1.3 Impact on Overall Design Synthesis and Optimization

In chapter 4.3.3, the author has elaborated on multi-disciplinary design optimization by means
of variation of wing loading and thrust-to-weight ratio. To show the importance of considering
engine power off-takes already in conceptual aircraft design synthesis, the author has conducted
parameter variations of wing loading and thrust-to-weight ratio for the A320 reference aircraft.
The parameter variations allow for identifying the optimum design point. Figure 6.5 shows
design optimization for the A320 reference aircraft with engine power off-takes, and figure 6.6
without engine power off-takes accordingly. For each combination of wing loading and thrust-
to-weight ratio, a full design loop with MICADO has been run that concluded in a converged
design case each. The design space is limited by top-level performance requirements. Anton
and the author [82] have elaborated on these constraints; here only a brief overview is given:

• Maximum cruise Mach number is constrained by both available thrust and cruise drag.
Available thrust increases with T/W, and drag decreases with W/S3 . Hence, the critical
combination of W/S and T/W that limit the design space show a parabolic boundary for
maximum cruise Mach number, see figures 6.5(b) and 6.6(b) (lower left-hand corner).

• Minimum climb gradients are only constrained by T/W. If a critical T/W is reached,
requirements regarding minimum climb gradients cannot be met. Hence, requirements
regarding minimum climb gradients limit the design space by means of an horizontal
boundary, see figures 6.5(b) and 6.6(b) (lower part).

• Take-off field length depends on available thrust and high lift characteristics of the
wing. For increasing W/S, the required maximum lift coefficients for achieving equal
field performance also increase. Hence, higher T/W are required for increasing W/S in
order to satisfy the TOFL requirements. The TOFL boundary shows an exponential
characteristic, see figures 6.5(b) and 6.6(b) (lower right-hand side).

• Required landing distance only depends on high lift characteristics, and thus on wing
loading. Hence, an almost vertical boundary line constraints the design space regarding
LDN requirements, see figures 6.5(b) and 6.6(b) (right-hand side).

For the A320 reference design with engine power off-takes, minimum T/W is constrained
to approximately 0.32 by climb requirements. Maximum W/S is constrained to approxima-
tely 630 kg/m2 by LDN requirement, see figure 6.5(b). The optimum design point that results
in minimum fuel burn almost coincide with the Airbus A320 design point, see figure 6.6(b).Since
engine power off-takes impact available thrust, constraints of the design space are significantly

3
Wing reference area decreases with increasing W/S; less surface area leads to lower friction drag. Wave drag
increases with lift coefficients due to smaller wings. Here, friction drag has the dominating impact.
118 Chapter 6. Case Studies

relaxed for the A320 design without engine power off-takes (see figure 6.6(b)) with the ex-
ception of LDN constraint. For example, lowest possible T/W now decreases from 0.32 to
approximately 0.295. Also the optimum design point is shifted towards lower T/W and higher
W/S if no engine power off-takes are considered. It results in a W/S of 610 kg/m2 , and a T/W
of 0.312. If no engine power off-takes are considered, the fuel consumption for the new design
point decreases by 8.6 % in comparison to the reference design, see figure 6.6(b).

,!" ,! ,!,


!" !"

# $%&'&'() #& & 



 ,!"
!" +"
 ,!,

 !"

)% 
   * &
!" 

 
  !"

   
   
   
        

 
   
(a) 3D view. (b) 2D view.

Figure 6.5: Optimization of W/S and T/W of A320 with engine power off-takes.

-!- -! !,


!"  !"
# $%&'&'() #& & 



 !"
!,"

 !" +"




!"
)% 
# * &
  !,-
  !" !,

   

  

 
   
        

 
   
(a) 3D view. (b) 2D view.

Figure 6.6: Optimization of W/S and T/W of A320 without engine power off-takes.

Repercussions of engine power off-takes on conceptual aircraft design synthesis and multi-
disciplinary optimization within the MICADO environment are summarized in figure 6.7.
Changes in characteristic aircraft design parameters in comparison to the A320 reference de-
sign (straight line) are plotted. The diamond-symboled dashed line shows results if the mission
6.1 Impact of Engine Power Off-Takes on Aircraft Performance and Efficiency 119

analysis is conducted without considering any engine power off-takes. However, no re-sizing of
the aircraft is conducted. Hence, only estimated block fuel for the design and for the study
mission are influenced. The results coincide with the ones that have been shown in figure 6.4(b).
Fuel consumption decreases by 5.8 % for the design mission and by 6.3 % for the study mission.
The triangle-symboled dotted line shows results for re-sizing of the reference aircraft without
engine power off-takes; the design point (WS and T/W) has not been changed yet. These re-
sults coincide with the A320 design point that is indicated in figure 6.6(b). A further decrease
of fuel consumption can be observed. Also the entire design synthesis is now impacted, which
results in decrease of most design weights.

A320 reference design


4 No engine power off-takes: no re-design
Changes to reference design, %

No engine power off-takes: re-design


No engine power off-takes: optimization
0

-4

-8

-12
wing loading

thrust-to-weight

wing mass

fuselage mass

empennage mass

landing gear mass

propulsion mass

systems mass

OWE

block fuel (design)

block fuel (study)

MTOW

Figure 6.7: Impact of engine power off-takes on conceptual aircraft design synthesis.

Finally, results for the new design point that is shown in figure 6.6(b) are given by the squared-
symboled straight line in figure 6.7. Changes in the design point propagate to fuel savings of
approximately 8.6 % on the design mission, and to savings of approximately 10.8 % on the study
mission. Further, both OWE and MTOW decrease by approximately 3.5 %. Weight savings
on components level vary between 0 and 9 %. Hence, engine power off-takes have a significant
impact on overall aircraft design synthesis and design optimization, which further propagates to
assessment of the design space. Correct selection of the design point is especially important in
aircraft design synthesis, which is significantly impacted by engine power off-takes as shown in
the scope of this section. Please note that the before discussed results must not be interpreted
as a possibility to optimize the design by increasing fuel efficiency through neglecting engine
120 Chapter 6. Case Studies

power off-takes. Power off-takes are required for the systems to fulfill their dedicated aircraft-
level functions. Results show the importance of already considering engine power off-takes
during conceptual aircraft design synthesis, when large parameter spaces are investigated. A
wrongfully chosen design point propagates through the different design phases and either leads
to a non-optimum product or to costly iteration loops within later design phases.

6.2 Innovative Morphing Leading Edge

Today, application of laminar flow technology is among the most discussed approaches towards
improved efficiency. Fuel savings of approximately 10-15 % are targeted with this technol-
ogy [43]. Laminar flow technology targets the delay of transition from laminar to turbulent
flow, which results in a reduction of skin friction drag, and thus in lower fuel burn. Two tech-
nologies for achieving laminar flow are currently in the focus of research [47]: active or hybrid
laminar flow control (HLFC), and natural laminar flow (NLF) technology. HLFC uses active
suction or blowing to stabilize the boundary layer; NLF uses only passive means where sta-
bilization of laminar flow is achieved by airfoil shaping. This case study concentrates on an
enabling technology for the latter. Natural laminar flow technology requires high contour and
surface smoothness, which cannot be achieved with current high-lift devices. Today’s slotted
leading edge devices inevitably lead to unsteadiness in surface contour even when they are
retracted. Integration of morphing structures as droop nose devices could provide the required
smoothness to enable natural laminar flow for an aircraft’s wing.

Conceptual design and feasibility studies of such smart high-lift devices were targeted by the
EC funded 7th Framework Program Collaborative project SADE4 [241, 242, 243, 244, 245].
Research included selection of morphing structures [246, 247, 248] and actuation concepts [249,
250], hardware testing of prototype systems [251], wind tunnel testing of a full-scale wing
segment, and integration and technology assessment of the concept on overall aircraft-level [208].
In the scope of SADE, the proposed morphing concept reached a conceptual maturity, which
leaves open uncertainties in important design parameters such as system mass or high-lift
performance. Research on the SADE concept is currently continued within the EC funded 7th
Framework Program Collaborative project SARISTU5 . It is the goal to reach higher technology
maturity, and to target impact on other supporting systems such as the anti-icing system, or
bird strike and lightning strike protection. In this case study, the author’s framework for
integration of innovative system technology is applied to assessment of the SADE innovative
morphing leading edge device. Results are based on two publications by the author [208, 194].
4
SmArt High-Lift DEvices for Next-Generation Wing
5
Smart Intelligent Aircraft Structures
6.2 Innovative Morphing Leading Edge 121

6.2.1 Laminar Aircraft Design

Contrary to the conventional reference aircraft (cf. ch. 4.3.3), the NLF aircraft features a high-
aspect ratio and low sweep wing design. Accordingly, the engines are moved aft to the rear
fuselage to provide undisturbed air flow over the wings, see figure 6.8. Due to the lower angle
of sweep, and due to requirements regarding maximum Reynolds number for NLF wings [252],
the cruise Mach number is reduced from Mach 0.80 to Mach 0.74. This adaption in top-level
requirements changes the aircraft’s economics due to longer trip times, and thus lower utilization
in a given time frame. Other top-level requirements remain unchanged in comparison to the
conventional reference aircraft.

In the past, numerous research was published that concentrate on aerodynamic airfoil and wing
design for NLF technology, which is not the focus of this research. The author uses an existing
NLF wing design within the overall MICADO design synthesis. Within the EU 6th Framework
Program Collaborative project NACRE6 [253], a High Aspect Ratio and Low Sweep (HARLS)
wing was designed and optimized for application to single-aisle aircraft. The HARLS wing is
designed for a cruise Mach number of 0.74 [254]. It features a leading edge sweep of 18.3 degrees
and an aspect ratio of 13.8. In design synthesis of the different NLF concepts, the HARLS wing
is slightly scaled7 in each iteration to match the required W/S.

Turbulent flight polars are estimated with a validated program module [211] within MICADO.
Prediction of laminar flight polars in dependence on the transition line is often inaccurate
because current semi-empirical or higher-order CFD models contain uncertainties and have
undue sensitivity to outside influences. To avoid over-optimistic results, the author choses
not to estimate drag savings from a prediction of the transition line in the design synthesis of
NLF aircraft. Instead, drag savings are treated as an input parameter, so that quasi laminar
flight polars are estimated from the turbulent flight polars in conjunction with a pre-defined
reduction in drag counts (DC). Dependency of laminar drag reduction on Mach number and
lift coefficient is neglected; the same is true for changes in pitching moment. Since laminar flow
can only be assumed for clean configuration, lift coefficient and Mach number variations are
narrow. It can be assumed that additional trim drag due to higher pitching moments of NLF
airfoils is already incorporated in the assumed drag savings. Thus, the approach is sufficient for
conceptual design. Figure 6.9 shows exemplarily the lift over drag (L/D) polar for turbulent
and laminar airflow of the NLF aircraft design. The chosen approach towards modeling of
laminar flight polars as well as its validity for the design case is described in greater detail in
publications by the author [199, 205, 194].

6
New Aircraft Concepts REsearch
7
Leading edge sweep and aspect ratio are held constant; only the wing planform is scaled.
122 Chapter 6. Case Studies

24
Mach 0.74
20 turbulent
laminar
16

L/D
12

0
0.0 0.2 0.4 0.6 0.8 1.0 1.2
CL

Figure 6.8: General arrangement of laminar Figure 6.9: Turbulent and laminar flight polar
aircraft. for a drag reduction of 35 DC.

6.2.2 Innovative Morphing Leading Edge and Systems Architecture

The SADE morphing droop nose is based on a concept by the German Aerospace Center (DLR)
and EADS Innovation Works [246, 247, 248, 249, 250, 251]. It is realized in a full-scale wind
tunnel demonstrator8 that was tested at the Central Aerohydrodynamic Institute (TsAGI,
Russia). The concept is based on a kinematic chain that is combined with a flexible composite
skin with variable skin thickness as illustrated in figure 6.10. This concept relies on already
certified composites for use in aviation. It has been designed for integration into a 185-PAX
single-aisle aircraft (cf. ch. 4.3.3).

Two optional system topologies were un-


der investigation in SADE. The first op- 

tion relies on a conventional central-   
ized drive system with a power control
unit (PCU) and a torque shaft, see fig-
ure 6.11(a). The main lever, which intro-
duces the displacements at several points
    
 

into the skin, is driven by a torque shaft  

 
  
 

and gearboxes. This system topology
uses similar components for the trans- Figure 6.10: Kinematic chain concept (by courtesy
mission system as a conventional high lift of DLR & EADS IW).
system. Hence, the advantage lies in ap-
plication of proven technology. Certification considerations such as a guaranteed synchronized
droop nose movement (CS 25.701 [9] and FAR §25.701 [89]) are easily met. The second sys-
tem topology relies on distributed electro-mechanical actuators for each actuation station, see
8
Rectangular wing segment of 5 m morphing span and 3 m chord.
6.2 Innovative Morphing Leading Edge 123

figure 6.11(b). The EMAs are powered by the aircraft’s 115 V/400 Hz AC electrical system
(ATA-24). Other than for the mechanical drive, the use of distributed EMAs implies a sophis-
ticated control unit to meet certification requirements. This more upstream approach follows
the trend towards more electrical aircraft architectures.

 
       
 

   !"   
  
 

  ! "  #   

 

 
 
  
 !" 
# $  
%%& ' ())*+  



 

(a) Option (A) – conventional drive. (b) Option (B) – electrical drive.

Figure 6.11: Possible system topologies for the SADE kinematic chain concept.

6.2.3 Initially Derived Systems Characteristics

The morphing leading edge system is further subdivided onto component-level for estimating
the required systems characteristics: (1) high-lift surface, (2) drive system, and (3) kinematics.
High-lift surface characteristics are the same for both system topologies, whereas the drive
system as well as the kinematics differ. Glass fiber composites are suggested by DLR for the
flexible skin. Studies show that a unit mass of approximately 15.3 kg/m2 per reference leading
edge area is required; stiffeners and attachments are included. Thus, for a leading edge reference
area of 21.9 m2 of the reference aircraft, panel weight results in 355 kg.

FE-analyses imply that a total of 72 actuators stations are required to hold the targeted contour
of the flexible skin also under high aerodynamic loads in e.g. cruise flight. Additionally, a rib
has to be attached to the front spar at each actuation station as shown in figure 6.10. For
comparison, a typical single-aisle aircraft like the Airbus A320 features a total of 10 slats. Each
slat is actuated by two geared-rotatory actuators. Thus, total required actuators increase by
a factor of 3.6 for the morphing leading edge concept due to the flexibility of the skin. FE-
analyses further show that higher actuation torques are required than for conventional slat
systems. Additional torque is required for deformation of the skin besides encountering the air
loads. Thus, components of the drive system must be sized for higher torques.
124 Chapter 6. Case Studies

Initial estimates result in two different system masses, which are accumulated from the different
components of topology (A) and (B) as listed in table 6.1. Comparison to the conventional slat
reference system shows an increase in total system mass of 72 % for system topology (A) and
of 100 % for system topology (B), see figure 6.12(a). This significant increase in overall system
mass results from higher weights for kinematics and the drive system, despite the decrease in
structural high-lift surface weight. Further, FE-analyses imply an increase in maximum power
requirements of 38 %, and of 19 % accordingly for an average drive cycle. Hence, system mass
of the secondary power systems also increases for the proposed innovative high-lift system.
(a) PCU and torque shaft drive system. (b) Electromechanical actuator drive system.

Components Units Mass, kg Components Units Mass, kg


Total – 1,806 Total – 2,100

High-lift surface – 335 High-lift surface – 335


Panels 10 334.9 Panels 10 334.9
Drive system – 226 Drive system – 296
Power control unit 1 55.2 Control unit 72 216.0
Torque shaft 1 91.9 Power & wiring 2 80.0
Wing tip brakes 2 15.7
Kinematics – 1,469
Gearboxes 2 29.3
Torque limiters 2 12.6 Actuators 72 504.0
Bearings 2 20.8 Lever & struts 72 583.2
Front spar ribs 72 381.6
Kinematics – 1,246
Actuators 72 280.8
Lever & struts 72 583.2
Front spar ribs 72 381.6

Table 6.1: Estimated masses of both smart leading edge system topologies.

For integration of the innovative morphing leading edge devices, secondary power systems have
to be resized accordingly for the changes in mission dependent power requirements and peak
loads during high lift deployment. Since non-essential electric consumers can be easily isolated
during deployment of high lift devices, total required generator capacity remains unchanged
also for the distributed drive architecture. However, additional cabling for the 72 drive sta-
tions increases ATA-24 system weight as shown in figure 6.12(b). For the centralized drive
architecture, maximum hydraulic power increases with higher torques for deployment of the
morphing leading edges. System weight increases accordingly as shown in figure 6.12(b). Those
6.2 Innovative Morphing Leading Edge 125

changes in systems weight accumulate to an increase in OWE of the aircraft, see fig. 6.12(c).
The indicated changes in total OWE, however, do not account for snowball effects of systems
integration in overall design synthesis yet.

1000 110 106


Reference System
Relative masses, %

System Topology (A) 104

Relative masses, %

Relative masses, %
System Topology (B) 105
102
100
100

100 95
98

90 96
Total High-Lift Actuation + Drive ATA-24 ATA-29 OWE
System Surface Kinematics System (no snowball effects)

(a) Changes in high lift component mass (logscale). (b) Changes in secondary (c) Changes in OWE.
power systems mass.

Figure 6.12: Influence of system integration on system mass.

6.2.4 Technology Assessment with Probabilistic Approach

The proposed framework for technology assessment considering uncertainties is used for in-
tegration and assessment of the innovative morphing leading edge devices on overall aircraft-
level. System characteristics that were beforehand derived from the initial estimates are used as
starting values for the probabilistic approach. Application of the different steps of the proposed
methodology (cf. ch 3.4) are described in the following paragraphs. The case study evaluates
technology integration on overall aircraft-level.

(1) Identification of Interdependencies of Technology Integration

The morphing leading edge is an enabling technology for application of natural laminar flow
technology. Integration directly influences the aircraft’s high-lift system performance as well
as its systems characteristics such as mass and power consumption. Also other systems are
affected by changes in the overall aircraft systems architecture. The flexible skin structure is
incompatible with current bleed air anti-icing systems. Integration of an electric wing anti-
icing system will most likely lead to a bleed-less systems architecture, which again results in
an electric ECS. The required electric power increases, and thus the electric system is affected
126 Chapter 6. Case Studies

directly from systems integration by increased power loads. A more comprehensive overview of
the interdependencies was already given by the author [208, 194].

(2) Down-Selection of Relevant Interdependencies

Required changes in other systems were not scope of the conceptual design activities within
SADE research. These repercussions of systems integration are directly targeted in the follow-
on project SARISTU9 , which is currently active research. It was also decided within SADE
to concentrate on the more upstream approach with distributed EMAs, rather than on the
conventional drive concept with a central PCU. Thus, the morphing leading edge devices are
driven with electric power, which leads to increased power requirements of the electric system
and accordingly to decreased power requirements of the hydraulic system. It was, however,
shown in figure 6.12 that the corresponding re-sizing of the secondary power systems only
results in slight increase of system mass, which has an overall impact on fuel efficiency of less
then one percent [208]. Further, more power is required for morphing than for deployment
of conventional slats. This, of course, leads to increased engine power off-takes, and thus to
increased fuel burn. Since the devices are only deployed/retracted once per mission cycle,
overall impact on mission fuel is negligible [208]. Thus, re-sizing of the aircraft’s power systems
as consequence of technology integration can be neglected for the morphing leading edge devices.
The following three most relevant interdependencies remain:

• Laminar drag savings and fuel efficiency

• High-lift systems mass and operating weight empty

• High-lift performance and wing loading

(3) Allocation of Design Parameters to Relevant Interdependencies

Laminar flight polars are estimated from turbulent ones in conjunction with a pre-defined
reduction in drag counts. Thus, possible laminar drag savings (∆DC) are allocated to the
design parameter that describes the reduction in overall drag. The interdependency with system
mass and OWE is covered by variations of the leading edge high-lift system mass (mLE,system ).
This propagates first to overall system mass and then to OWE within design synthesis. The
interdependency with high-lift performance and wing loading is covered by variations of the
maximum lift coefficient for landing configuration (CL,max,L/G ). Again, the impact propagates
through the design, so that the identified interdependencies are sufficiently covered within
9
Smart Intelligent Aircraft Structures: project started in September 2011; it will end in August 2015.
6.2 Innovative Morphing Leading Edge 127

design synthesis and optimization. Figure 6.13 shows the allocated design parameters and the
modules of MICADO, in which variations of these parameters are introduced.


   

      

   

           !  


        

 "# $

  

     


& 



% 
    # 

  


  
 
$  '  

&
#(  )
  

  


  

Figure 6.13: Allocated design parameters in MICADO.

(4) Identification of the Parameter Space of Each Impacted Design Parameter

In the following paragraph, lower and upper boundaries as well as ’most-likely’ estimates for
each of the three identified design parameters are derived.

A. Laminar Drag Savings (∆DC)

It is extremely improbable that integration of morphing leading edge devices will lead to an
increase in overall drag. Thus, zero drag reduction can be assumed as lower margin. According
to literature [252, 255, 256] drag savings between 20 and 40 drag counts seem to be most realistic
for application laminar flow technology to transport aircraft. Therefore, a drag reduction
of 30 DC was chosen as initial estimate. Further, literature also states that drag savings of 50
to 60 drag counts are quite optimistic, which leads to the assumed upper margin of 60 DC.
128 Chapter 6. Case Studies

B. High-Lift System Mass (mLE,system )

Today’s conventional high-lift systems are highly optimized for weight and performance. Hence,
it is also extremely improbable that application of a first-generation morphing leading edge
system results in lower systems weight and better high-lift performance. System weight of the
conventional system, when applied to the NLF configuration, was estimated to 1051 kg, which
serves as lower margin. The aforementioned estimates result in a systems weight of 2100 kg for
the distributed systems architecture, which is regarded as initial estimate within this probabilis-
tic approach. The conceptual system design was yet not optimized for weight, which leads to
the conclusion that additional weight can be saved. On the other hand, it was also not designed
for airworthiness qualification. Further, lightning strike, bird strike, and ice protection is yet
not integrated into the morphing leading edge, all of which will add additional weight to the
system. The author chose to assume additional 20 % of system weight as worst case scenario,
and thus as upper margin for the design parameter.

C. High-Lift Performance (CL,max and L/Dtake−of f )

It can be expected that high-lift performance will level between the one of an aircraft without
leading edge devices and the one with a conventional high-lift system. Maximum lift coefficient
for landing configuration was estimated to 3.12 for the conventional system, and to 2.60 for the
system without any leading edge devices with MICADO. This corresponds to the chosen lower
and upper boundaries of the design parameter. Internal two-dimensional CFD calculations
for shape optimization of the innovative system by DLR were extrapolated to a full three-
dimensional configuration. Extrapolation has resulted in an initial maximum lift coefficient for
landing of 2.78 for the morphing leading edge configuration. Impacts on L-over-D in take-off
configuration are covered by the estimated drag savings that result from laminar flow. Since
most aircraft use means to seal the gap between the slats and the wing in take-off configuration,
no significant difference in high-lift performance between a conventional slat and the innovative
droop nose device is expected for take-off configuration.

The identified margins and initial estimates of the three design parameters are summarized in
table 6.2.

parameter lower margin upper margin initial estimate


Laminar drag savings (∆DC) 0 DC 60 DC 30 DC
LE high-lift system mass (mLE,system ) 1051 kg 2519 kg 2099 kg
max. lift for landing (CL,max,L/G ) 2.60 3.12 2.78

Table 6.2: Margins and initial estimates for the design parameters.
6.2 Innovative Morphing Leading Edge 129

(5) Allocation of a Continuous Probability Distribution to Each Parameter Space

According to the approach that has been suggested in chapter 3.4 the parameter space is normal-
ized and probability functions are applied. The three resulting probability functions are plotted
in figure 6.14 for the normalized parameter space. Only the reduction in drag counts shows a
symmetric probability distribution over the design space. Hence, a Gaussian normal distribu-
tion is used as normalized probability density function (p∆DC (e x)), see equation 6.1. For the
two other design parameters asymmetric beta distributions (pmsystem,LE (e y ) and pCL,max,L/G (e
z ))
are applied, see equation 6.2 and 6.3.
√ x−µ)2
(e
x) = 1/( 2π · σ) · e− 2·σ2 with σ = 0.166 and µ = 0.5
p∆DC (e (6.1)
p−1 q−1
pmsystem,LE (e
y ) = 50.154 · ye · (1 − ye) with p = 4.7 and q = 2.483 (6.2)
z ) = 85.959 · zep−1 · (1 − ze)q−1 with p = 2.959 and q = 4.7
pCL,max,L/G (e (6.3)

(6) Overall Aircraft Design Synthesis and Optimization Parameter Sweep

For variations of the three identified design


parameters, design synthesis and optimi-
~), p(z~)

~) ~) (z~)
zation parameter sweeps were run with MI- p∆DC(x pm (y pC
2.5 LE,sys L,max,L/G

CADO. A matrix of 420 aircraft were fully


~), p(y

2.0
designed and optimized for the discrete pa-
probability density: p(x

rameter space with 7 steps in ∆DC, 6 steps 1.5

in CL,max,L/G , and 10 steps in mLE,system . 1.0


The optimizations show that morphing lea- 0.5
ding edge system weight has significant 0.0
influence on overall design synthesis, but 0.0 0.2 0.4 0.6 0.8 1.0
normalized design parameter: ~x, ~y, ~z
no influence on the optimum design point
for wing loading and thrust-to-weight ra- Figure 6.14: Probability distributions for the iden-
tio. Optimum wing loading and thrust-to- tified parameter space.
weight ratio are only influenced by laminar
drag reductions and by maximum lift coef-
ficient. Figure 6.15 illustrates the resulting optimum design points (W/S and T /W ) for the
investigated parameter space. Since friction drag decreases with increasing drag savings, op-
timum W/S shifts to lower values. Correspondingly, required T/W also decreases, due to
improved take-off and climb performance that results from lower W/S. For decreasing maxi-
mum lift coefficients, W/S also shifts to lower values. Higher approach speeds, which would lead
to longer landing distances, must be compensated by lower W/S. Loss in take-off performance
due to lower maximum lift is partly compensated by higher T/W.
130 Chapter 6. Case Studies

  



 

 



  
  
     
       

    
!"    
       
    

(a) Optimum wing loading. (b) Optimum thrust loading.

Figure 6.15: Optimum design points for the parameter sweep.

(7) Technology Assessment within Provided Parameter Space

The presented technology assessment is based on technical parameters alone. Due to additional
uncertainties in scenario parameters no monetary assessment, e.g. for direct operating costs,
is conducted in scope of this thesis. The assessment is based on the aircraft’s efficiency on
a 1000 NM study mission. This study mission is flown with a payload factor of 95 %, which
corresponds to a payload of 21.9 t (cf. ch. 4.3.3). The design synthesis and optimization para-
meter sweeps provide the required block fuel (mblock f uel ) on the study mission for each design.
Since cruise Mach number is reduced from 0.8 to 0.74, economics are changed for the NLF
aircraft in comparison to the conventional reference aircraft. Hence, for assessment against the
conventional reference aircraft it is not sufficient to use block fuel alone. Besides block fuel,
specific hourly productivity (SHP ) is additionally introduced into the presented assessment. It
is suggested by Doganis [257] for evaluation of aircraft economics and is defined as the product
of payload (mpaylaod ) and flight speed (T AS), divided by block fuel. Flight speed can easily be
substituted by the quotient of mission range (Rmission ) and block time (tblock time ), thus:

mpaylaod · T AS mpaylaod · Rmission


SHP = = . (6.4)
mblock f uel mblock f uel · tblock time

A. Assessment of Block Fuel

Results for block fuel consumption are plotted in figure 6.16(a). The different surfaces cor-
respond to a specific drag reduction each, and show the relative fuel savings in comparison
to the conventional reference aircraft in dependence on leading edge system weight and on
maximum lift coefficient. Not surprisingly, overall fuel consumption increases with lower drag
savings, higher system mass, and less high-lift performance. For drag savings of 30 DC, which
is assumed to be the most realistic case, fuel savings vary between 5.7 % and 11.5 % over the
6.2 Innovative Morphing Leading Edge 131

conventional reference aircraft. Due to repercussions of system integration, not all designs
within the relevant parameter space show improvements in fuel efficiency. As illustrated in fig-
ure 6.16(a), for drag savings of 10 DC and specific combinations of mLE,system and CL,max,L/G ,
the designs show higher fuel consumption (limit shown by dashed line). Designs that correspond
to the parameter space margins from table 6.2 are plotted in figure 6.16(b). The most probable
case (rectangle-solid line) corresponds to initial estimates for mLE,system and CL,max,L/G ; drag
savings are varied. The worst case (triangle-dashed line) corresponds to highest system weight
and worst high-lift performance. The best case (diamond-dashed line) corresponds to lowest
system weight and best high-lift performance accordingly. Additionally, the dashed horizontal
line shows the reference to the conventional reference aircraft. The most probable case and
laminar drag savings of 30 DC lead to an improvement in fuel efficiency of approximately 7 %.

    &' & (  5



  conventional
 refernce A/C
$ ) 0
∆ Block Fuel, %

$%
-5
$! $# )
$ -10
$)
$ worst case
-15 probable case
 # best case
#
  " -20

 

!  
10 20 30 40 50
     
  Savings in Drag Counts
(a) Fuel savings for design space. (b) Comparison to laminar reference aircraft.

Figure 6.16: Potential for fuel savings.

So far, uncertainties in the design parameters have not been incorporated in the results that
are illustrated in figure 6.16. In figure 6.17, potential fuel savings of the NLF aircraft with a
drag reduction of 30 DC are shown in the normalized parameter space. It is easy to observe
that iso-fuel savings (dashed iso-lines) show a linear trend, and that fuel savings vary between
approximately 6 and 11 %, cf. figure 6.16(b). The joint probability density distribution is plotted
in figure 6.18 for the same normalized parameter space. As expected from equation 6.2 and 6.3,
it is an asymmetric distribution with a peak at higher system weights and lower maximum
lift coefficients. For assessing the probability that e.g. the different levels of fuel savings from
figure 6.17 can be achieved, the joint probability density function is integrated numerically over
the entire parameter space10 :
Z x
emax Z 1
P (0 <= x
e <= x
emax , yemin (e
x) <= ye <= 1) = p(e
x, ye) de
y de
x. (6.5)
0 yemin (e
x)

10
According to fig. 6.17, the integral runs from the top right corner to the bottom left one.
132 Chapter 6. Case Studies

Results (P (ex, ye)) that are illustrated in figure 6.17 and 6.18 do yet not incorporate the probabi-
lity for variations in drag savings (P (e z )). Hence, they must be superposed with the probability
density function to derive the overall probabilities (P (e x, ye, ze)) for the full three-dimensional
design space. Figure 6.19 shows the overall joint probabilities and discrete margins for fuel
savings. Hence, for integration of innovative morphing leading edge systems and application
of NLF technology, fuel savings of more than 4 % are very probable (P = 85%) and of more
than 2 % almost certain (P = 95%). The steepest gradient is found for fuel savings between
6 and 8 %. Thus, fuel savings in this range seem to be realistic (P = 42% to P = 66%). Fuel
savings of more than 12 %, however, are quite unprobable (P = 8.7%), and of more than 14 %
almost impossbile (P = 2.6%).

 

     

 
 
   

   

 

 

  

  
 
 
   
 
     

#
 
        
    "
      ! 

Figure 6.17: Fuel savings for NLF aircraft Figure 6.18: Joint probability density distri-
with 30 DC reduction and nor- bution for normalized parameter
malized parameter space. space (p(e
x, ye)).

100
95%
y, ~z), %

85%
80
66%
~, ~

60
Probability P(x

42%
40
22%
20
8.7%
2.6%
0
0 2 4 6 8 10 12 14 16
Improvements in Block Fuel, %

Figure 6.19: Probabilities of fuel savings for the 3D joint probability function.
6.2 Innovative Morphing Leading Edge 133

B. Assessment of Specific Hourly Productivity

Results for SHP are plotted in figure 6.20(b). As evident from equation 6.4, the longer mission
time reduces the benefits from NLF technology that were seen for fuel consumption significantly.
For assumed laminar drag savings of 10 DC, the resulting designs show worse economics than for
the conventional reference aircraft, see figure 6.20(a). Even for the best case in figure 6.20(b),
at least 15 DC are required to meet equal performance compared to the conventional reference
aircraft. Repercussions of the NLF aircraft even become more significant when considering the
uncertainties. In figure 6.21, the overall probabilities for improvements in block fuel and SHP
are compared against each other. Whereas it is almost certain that improvements in block fuel
can be achieved, improvements in SHP only result in a probability of 65 %. Hence, it is not
unrealistic that the NLF aircraft with innovative morphing leading edge devices will show worse
economics11 than the conventional reference aircraft. Improvements in SHP of more than 6 %
are quite unprobable (P = 11%), and of more than 10 % almost impossible (P = 1.2%).


15
∆ Specific Hourly Productivity, %

   worst case


    probable case
 
10
  best case

 5

   0
 conventional
reference A/C
-5
 '

  '( -10
!
"

  # 
'  %
 
$ 10 20 30 40 50
$ ')  %&
 ' Savings in Drag Counts
(a) SHP for design space. (b) Comparison to laminar reference aircraft.

Figure 6.20: Potential for improvements in specific hourly productivity.

C. Conclusion

Results show that fuel savings can be achieved for almost the entire parameter space. Hence,
the corresponding probability for achieving better fuel efficiency than the conventional refer-
ence aircraft is nearly one (P = 99%). Fuel savings of more than 6 % are still quite proba-
ble (P = 66%). Fuel savings of more than 12 % are unprobable (P = 8.7%). SHP incorporates
the additional repercussion of longer required block times of NLF aircraft into the assessment.
It was shown that this repercussion leads to worse overall improvements over the conventional
reference aircraft. A large area of the parameter space even show worse SHP than the con-
ventional reference aircraft. This was also mirrored in the assigned probabilities. An improved
11
Captured by means of specific hourly productivity in the scope of this thesis. Please refer to Peter et al. [258]
for assessment of DOC; his results yield the same conclusion.
134 Chapter 6. Case Studies

SHP for the introduced innovative technology is still probable (P = 65%), but far from certain.
Improvements in SHP of more than 4 % are rather unprobale (P = 24%), and more than 10 %
almost impossible (P = 1.2%).

Direct operating costs have not been considered in the scope of this thesis. However, it is
important to also consider DOC for capturing additional repercussions of the NLF aircraft. In
a recent publication, Franz and the author [201] showed that for moderate fuel price scenarios,
additional repercussions lead to a further worsening of economics when compared to SHP.
Peter et al. [258] showed in a recent publication that incorporating DOC into the assessment
of the morphing leading edge devices indeed leads to a worse perspective of the technology.

This case study has shown the added-value that is yielded by the author’s approach towards
integration of innovative systems technology and incorporating design uncertainties for sound
technology assessment in conceptual aircraft design.

100
SHP
y, ~z), %

80 block fuel
65%
~, ~

60
Probability P(x

43%
40
24%
20
conventional 11%
reference A/C 4.1%
0
-8 -6 -4 -2 0 2 4 6 8 10 12 14 16
Improvements in Block Fuel and SHP, %

Figure 6.21: Probabilities for improved SHP for the 3D joint probability function.
7 Conclusion

In the scope of this thesis, the author has presented an approach towards modeling of entire
aircraft systems architectures within conceptual aircraft design. The methodology allows for
sizing of the single systems from their dedicated aircraft-level functions. Further, single system
weights as well as the maximum and mission dependent power demands are estimated for
capturing repercussions of systems integration in overall aircraft design synthesis. The proposed
methodology is based on a universal approach that uses energy balancing for determining energy
fluxes within the overall systems architecture. This approach provides the required modularity
and flexibility by means of composition of the overall systems architecture from single systems
or generic building blocks. Hence, existing system architectures can be easily adapted, which
allows for investigation of large design spaces within conceptual aircraft design synthesis.

Further, the author has introduced a novel framework for integration of also innovative system
technologies for which no reliable models exist. A probabilistic approach has been chosen
that incorporates the remaining design uncertainties into the results, and allows for assessment
of innovative system technology on overall aircraft-level. By using probabilistic theory, this
approach provides probabilities whether specific design goals can be achieved with the new
technology, e.g. a certain increase in aircraft efficiency or economics.

Both the proposed methodology for modeling of entire systems architectures as well as the
new framework for integration of innovative systems technology by means of a probabilistic
approach, rely on generic models that can be used for any application that features complex
systems, e.g. aviation or space applications. In the scope of this thesis, the proposed method-
ology has been implemented into a system model for today’s commercial transport aircraft.
Single system models have been chosen for implementation that rely on functional relations for
estimating the single system characteristics from the aircraft-level functions. Hence, main de-
pendencies of system sizing on performance requirements and on overall aircraft characteristics
are captured correctly within this conceptual design tool.

Further, the proposed systems model has been integrated into the MICADO environment,
which allows for capturing repercussions of systems integration in overall design synthesis and
for assessing impacts on overall aircraft-level. Both mission dependent engine power off-takes

135
136 Chapter 7. Conclusion

as well as system weights propagate through design synthesis and snowball effects are captured
by the MICADO environment. Hence, sound technology assessment of also innovative system
technology becomes possible.

Comprehensive sensitivity studies have shown the main interdependencies of the proposed sys-
tems model; main design drivers of the systems architecture have been identified. Further the
sensitivity studies have allowed for showing plausibility and feasibility of the obtained results,
and thus support verification of the proposed model. The author has shown that the expected
level of accuracy lies in the range of 5 to 10 %, and thus completely satisfies the required
accuracy for application in conceptual and early preliminary design synthesis. The proposed
systems model also yields limitations. For example, the concept of communality of installed
systems and components for reducing complexity and MRO efforts in aircraft families is not
regarded for by the proposed systems model. It has been shown that some deviations between
results of the proposed systems model and the Airbus A320 aircraft can be traced back to such
communality considerations. The proposed systems model is curently also limited to applica-
tion to commercial transport aircraft by its implemented single system models. When using
the tool for modeling of other aircraft types, close attention must be paid to assure feasibility
of the results.

The two case studies that have been presented facilitate the benefits of the proposed systems
model for overall aircraft design synthesis. The author emphasized the importance of captur-
ing repercussions of system engine power off-takes on the aircraft performance and efficiency
in design synthesis. Only by incorporating system power demand, and thus also engine po-
wer off-takes, multi-disciplinary design optimization concludes in the correct optimum design
point. Neglecting repercussions of system integration leads to not-optimum design points that
propagate through the different design phases, and lead to exponential growing costs for cor-
rection as the design matures. Hence, application of the proposed model already in conceptual
design synthesis allows for sound decision making and results in reliable overall aircraft design
optimization. The benefits that arise from the novel framework for integration of innovative
system technologies have been shown in the second case study. The framework has been used
for assessing integration of innovative morphing leading edge devices on overall aircraft-level.
The approach allows for capturing design uncertainties that still exist in this new technology.
The case study has concluded with probabilistic technology assessment regarding the impact
on fuel efficiency and economics of a new single-aisle aircraft that features the aforementioned
technology. Results have shown the benefits of this novel approach. The proposed systems
model offers the capability for sound technology assessment and corresponding down-selection
of the most promising technologies already in conceptual aircraft design.
Bibliography

[1] Moir, I. and Seabridge, A. G., Design and development of aircraft systems: An introduc-
tion, AIAA education series, American Institute of Aeronautics and Astronautics and
Professional Engineering Pub., Reston VA , Bury St Edmunds, 2004.

[2] Foster, R. N., Innovation: The Attacker’s Advantage, Summit Books, 1st ed., April 1986.

[3] Lammering, T., Anton, E., Risse, K., and Franz, K., “Impact of Systems Integration on
Fuel Efficiency in Preliminary Aircraft Design,” 3rd International Workshop on Aircraft
System Technologies, Shaker, Hamburg, March 2011, pp. 171–180.

[4] Liscouet-Hanke, S., A Model-Based Methodology for Integrated Preliminary Sizing and
Analysis of Aircraft Power System Architectures, Doctoral thesis, Universite de Toulouse,
Toulouse, France, Sept. 2008.

[5] Airplane Characteristics - A320 , Airbus S.A.S., Customer Services, 31707 Blagnac Cedex,
France, 27th ed., July 2012.

[6] Schiek, T., “ATA21 Environmental Control Systems Function and Design,” HAW lecture,
Airbus Operations S.A.S., Hamburg, Germany, 2007.

[7] “Manual A319/A320/A321,” Tech. Rep. 21.00 - AIR CONDITIONING - GENERAL,


Airbus S.A.S, Aug. 2011.

[8] Freund, R., Entwicklung eines Modells eines Fluegelenteisungssystems, Bachelor thesis,
RWTH Aachen, Institute of Aeronautics and Astronautics (ILR), Aachen, Germany, 2011.

[9] EASA, Certifications Specifications for Large Aeroplanes, No. CS-25, Europena Aviation
Safety Agency, Brussels, Belgium, 12th ed., July 2012.

[10] Natzer, B., Bentele, R., Perrin, U., and Huth, S., “A318, A319, A320, A321 - Slat and
Flap Weight Report,” Design Report TK1065 Issue H, Liebherr-Aerospace Lindenberg
GmbH, Lindenberg/Germany, May 2012.

[11] Koeppen, C., Methodik zur modellbasierten Prognose von Flugzeugsystemparametern im


Vorentwurf von Verkehrsflugzeugen, Schriftenreihe Flugzeug-Systemtechnik, Shaker Ver-
lag, Hamburg, 1st ed., 2006.

[12] “ATA Specification 100 - Specification for Manufacturers’ Technical Data,” Revision 1999
99100P, Air Transport Association of America (ATA), Inc., Washington D.C. 20004, 1999.

137
138 Bibliography

[13] Scholz, D., “Betriebskostenabschaetzung von Flugzeugsystemen als Beitrag zur Entwurfs-
optimierung,” DGLR Jahrbuch 1995 , Deutsche Gesellschaft fuer Luft- und Raumfahrt
(DGLR), Bonn, 1995, pp. 50–61.

[14] Banbury, J., Behrens, D., and Browell, Q., “The IATA Technology Roadmap Report,”
Tech. rep., International Air Transport Association (IATA), Switzerland, June 2009.

[15] Bisignani, G., “A global approach to reducing aviation emissions,” Tech. rep., Interna-
tional Air Transport Association (IATA), Switzerland, Nov. 2009.

[16] A Vision for 2020 - Meeting Society’s Needs and Winning Global Leadership, European
Aeronautics, Office for Official Publications of the European Communities, Luxembourg,
1st ed., Jan. 2001.

[17] Flightpath 2050 - Europe’s Vision for Aviation, Office for Official Publications of the
European Communities, Luxembourg, 2011.

[18] “Report of the Committee on Aviation Environmental Protection,” Tech. Rep. Doc 9938,
International Civil Aviation Organization (ICAO), Montreal/Canada, Feb. 2010.

[19] Holdren, J. P., “National Aeronautics Research and Development Plan,” Tech. rep., Na-
tional Science and Technology Council, Washington D.C. 20502, Feb. 2010.

[20] Wall, R., “Research Road Map,” Aviation Week & Space Technology, Vol. 172, No. 44,
Dec. 2010, pp. 40.

[21] Warwick, G. and Norris, G., “Designs for Success,” Aviation Week & Space Technology,
Vol. 172, No. 40, Nov. 2010, pp. 72–75.

[22] Certification Considerations for Highly-Integrated or Complex Aircraft Systems, No.


ARP4754 in SAE Aerospace Recommaned Practices, Washington D.C., Nov. 1996.

[23] Bruns, M., Systemtechnik - Methoden zur interdisziplinaeren Systementwicklung,


Springer-Verlag, Berlin, 1991.

[24] Anderson Jr., J. D., The Airplane: A History of Its Technology, Amercian Institute of
Aeronautics and Astronautics, Inc., Reston, Va., 1st ed., 2002.

[25] “An Advanced Concept Secondary Power Systems Study for an Advanced Transport
Technology Aircraft,” Tech. Rep. NASA CR-112103, The Boeing Company, Seattle, WA
98124, Aug. 1972.

[26] Cronin, M., Hays, A., Green, F., Radovcich, N., Helsley, C., and Rutchik, W., “Inte-
grated Digital/Electric Aircraft Concepts Study,” Tech. Rep. NASA CR-3841, Lockheed-
California Company, Burbank, California, 1985.

[27] Hoffman, A. C., Hansen, I. G., Beach, R. F., Plencner, R. M., Dengler, R. P., Jefferies,
K. S., and Frye, R. J., “Advanced Secondary Power System for Transport Aircraft,”
Technical Paper NASA TP-2463, Lewis Research Center, Cleveland, Ohio, 1985.
Bibliography 139

[28] Murray, W., Feiner, L., and Flores, R., “Evaluation of All-Electric Secondary Power
for Transport Aircraft,” Tech. Rep. NASA CR-189077, McDonnell Douglas Corporation,
Long Beach, California 90846, Jan. 1992.

[29] Moir, I. and Seabridge, A. G., Military Avionics Systems: Aerospace Series, AIAA Edu-
cation Series, John Wiley & Sons, Ltd, West Sussex PO19 8SQ, UK, 1st ed., Feb. 2006.

[30] Collinson, R. P. G., Introduction to Avionics Systems, Springer, Dordrecht, 2nd ed., 2006.

[31] Fielding, J. P., Introduction to Aircraft Design, Cambridge University Press, Cambridge
CB2 2RU, UK, 1st ed., 1999.

[32] Moir, I. and Seabridge, A. G., Aircraft systems: Mechanical, electrical, and avionics
subsystems integration, John Wiley & Sons, Inc., West Sussex PO19 8SQ, UK, 3rd ed.,
2008.

[33] Moir, I. and Seabridge, A., Civil avionics systems, AIAA Educational Series, American
Institute of Aeronautics and Astronautics and Professional Engineering Publ., Bury St.
Edmunds, 2002.

[34] Roskam, J., Layout of Landing Gear and Systems, Vol. Part IV of Airplane design / Jan
Roskam, DARcorporation, Lawrence, Kan., 5th ed., 2007.

[35] Davies, M. and Scholz, D., The Standard Handbook for Aeronautical and Astronautical
Engineers, McGraw-Hill Professional, New York, NY 10121-2298, 1st ed., Oct. 2002.

[36] Faleiro, L., “Beyond the More Electrical Aircraft,” Aerospace America, Vol. 43, No. 9,
Sept. 2005, pp. 35–40.

[37] Warwick, G., “Choose the Future,” Aviation Week & Space Technology, Vol. 173, No. 26,
Aug. 2011, pp. 44–47.

[38] Prange and Wieland, “Triebwerstechnologie-Trends fuer zivile Turbostrahltriebwerke,”


Tech. Rep. AT 12 000-02, Luftfahrttechnisches Handbuch (LTH), Jan. 1983.

[39] Daly, M., Jane’s Aero-Engines, Vol. 26, Jane’s Information Group Limited, Couldsen,
U.K., Sept. 2009.

[40] Norris, G., “Next Steps,” Aviation Week & Space Technology, Vol. 173, No. 14, April
2011, pp. 40–42.

[41] Warwick, G., “Changing the Game,” Aviation Week & Space Technology, Vol. 172, No. 40,
Nov. 2010, pp. 70–71.

[42] Norris, G., “Building Blocks,” Aviation Week & Space Technology, Vol. 173, No. 14, April
2011, pp. 36–41.

[43] Wall, R. and Warwick, G., “Cleaner Skies,” Aviation Week & Space Technology, Vol. 173,
No. 22, June 2011, pp. 70–74.
140 Bibliography

[44] Warwick, G., “Time for Change,” Aviation Week & Space Technology, Vol. 173, No. 26,
Aug. 2011, pp. 50–52.

[45] Jupp, J. A., “21ST Century Challenges for the Design of Passenger Aircraft,” ICAS 2012
Proceedings, Brisbane, Australia, 2012.

[46] Warwick, G., “UHB: the acid test,” Flight International , Vol. 132, No. 4075, Aug. 1987,
pp. 22–25.

[47] Joslin, R. D., “Overview of Laminar Flow Control,” Tech. Rep. TP-1998-208705, NASA,
Langley Research Center, Hampton, Virginia, Oct. 1998.

[48] Alibrandi, W., “Power Shift,” Aviation Week & Space Technology, Vol. 178, No. 4, Jan.
2011, pp. 116–118.

[49] “Environmental Report 2010,” Tech. rep., International Civil Aviation Organization
(ICAO), Montreal/Canada, 2010.

[50] Heidmann, J., “Nasa’s Current plans for ERA Propulsion Technology,” Jan. 2010.

[51] Norris, G., “Smooth Operators,” Aviation Week & Space Technology, Vol. 172, No. 35,
Sept. 2010, pp. 61–61.

[52] Norris, G., “Real Rotors,” Aviation Week & Space Technology, Vol. 173, No. 22, June
2011, pp. 74–77.

[53] Seitz, A., Advanced Methods for Propulsion System Integration in Aircraft Conceptual
Design, Ph.D. thesis, Technische Universitaet Muenchen, Muenchen, Jan. 2012.

[54] “Systems Engineering for Intelligent Transportation Systems,” Tech. Rep. FHWA-HOP-
07-069, U.S. Department of Transportation, Washington D.C. 20590, Jan. 2007.

[55] Wall, R. and Sparaco, P., “Time To Deliver,” Aviation Week & Space Technology,
Vol. 165, No. 13, Oct. 2006, pp. 52–53.

[56] Regan, J. and Hepher, T., “EADS wants A400M contract change, adds delay,” Tech. rep.,
Reuters, Jan. 2009.

[57] Mecham, M. and Norris, G., “87 Monts and Counting,” Aviation Week & Space Technol-
ogy, Vol. 173, No. 4, Jan. 2011, pp. 24–26.

[58] Schofield, A., “Dreams Meets Reality,” Aviation Week & Space Technology, Vol. 172,
No. 6, Dec. 2010, pp. 24–25.

[59] Norris, G. and Mecham, M., “Jumbo Questions,” Aviation Week & Space Technology,
Vol. 173, No. 7, Feb. 2011, pp. 25–26.

[60] Royce, D., “The JSF Question,” Aviation Week & Space Technology, Vol. 173, No. 4,
Jan. 2011, pp. 50–53.
Bibliography 141

[61] Kingsley-Jones, M., “Early Warning,” Aviation Week & Space Technology, Vol. 172,
No. 42, Nov. 2010, pp. 39–40.

[62] Flottau, J., “Reality Bites,” Aviation Week & Space Technology, Vol. 173, No. 40, Nov.
2011, pp. 30–31.

[63] Flottau, J., “Testing Times,” Aviation Week & Space Technology, Vol. 175, No. 20, June
2013, pp. 62–64.

[64] Mavris, D. N., de Tenorio, C., and Armstrong, M., “Methodology for Aircraft System
Architecture Definition,” American Institute of Aeronautics and Astronautics (AIAA),
Reno, NV, Jan. 2008.

[65] Emadi, A. and Ehsani, M., “Aircraft Power Systems: Technology, State of the Art, and
Future Trends,” Aerospace and Electronic Systems Magazine, IEEE , Vol. 15, No. 1, Jan.
2000, pp. 28–32.

[66] Rosero, J., Ortega, J., Aldabas, E., and Romeral, L., “Moving Towards a More Electric
Aircraft,” Aerospace and Electronic Systems Magazine, IEEE , Vol. 22, No. 3, March
2007, pp. 3–9.

[67] Sghairi, M., Aubert, J.-J., de Bonneval, A., and Crouzet, Y., “Distributed and Reconfi-
gurable Architecture for Flight Control System,” 28th Digital Avionics Systems Confe-
rence, IEEE, Orlando, FL, Oct. 2009.

[68] Bieber, P., Boniol, F., Boyer, M., Noulard, E., and Pagetti, C., “New Challenges for
Future Avionic Architectures,” Jorunal AerospaceLab, Vol. 11, No. 4, May 2012, pp. 1–
10.

[69] Bosson, J., “Key Enablers for Power Optimized Aircraft,” 27th Congress of the Inter-
national Council of the Aeronautical Sciences (ICAS), Nice, France, Sept. 2010, pp.
ICAS2010–7.4.1.

[70] Sinnett, M., “787 No-Bleed Systems: Saving Fuel and Enhancing Operational Efficien-
cies,” Aero Quarterly, , No. Qtr. 4, 2007, pp. 6–11.

[71] Raymer, D. P., Enhancing Aircraft Conceptual Design Using Multidisciplinary Optimi-
zation, PhD thesis, Kungliga Tekniska Hoegskolan, SE-100 44 Stockholm, Sweden, May
2002.

[72] Dollmayer, J. and Carl, U., “Einfluss des Leistungsbedarfs von Flugzeugsystemen auf den
Kraftstoffverbrauch,” Deutscher Luft- und Raumfahrtkongress 2004 , 2004.

[73] Dollmayer, J., Methode zur Prognose des Einflusses von Flugzeugsystemen auf die
Missionskraftstoffmasse, Schriftentreihe Flugzeug-Systemtechnik, Shaker Verlag, Ham-
burg/Germany, 1st ed., 2007.
142 Bibliography

[74] Scholz, D., Seresinhe, R., Staack, I., and Lawson, C., “Fuel Consumption due to Shaft
Power Off-Takes from the Engine,” 4th International Workshop on Aircraft System Tech-
nologies, Shaker, Hamburg, April 2013, pp. 169–179.

[75] Prange and Sommer, “Einflussfaktoren auf die Triebwerksleistungen,” Tech. Rep. AT 24
000-01, Luftfahrttechnisches Handbuch (LTH), Dec. 1980.

[76] Herrmann and Biehl, “Rechnerische Ermittlung der installierten Leistungen von Tur-
bostrahltriebwerken,” Tech. Rep. AT 21 000-06, Luftfahrttechnisches Handbuch (LTH),
Oct. 1989.

[77] Braeunling, W. J., Flugzeugtriebwerke, VDI Buch, Springer Verlag, Germany, 2001.

[78] Tagge, G. E., “Secondary Electric Power Generation with Minimum Engine Bleed,” Air-
craft Electric Secondary Power , NASA, Cleveland, Ohio, Sept. 1982, pp. 37–50.

[79] Getting to Grips with Aircraft Performance, No. 1 in Flight Operations Support & Line
Assistance, Airbus S.A.S., Customer Services, 31707 Blagnac Cedex, France, Jan. 2002.

[80] Fuerst and Sellner, “Massezuwachsfaktor (growth factor),” Tech. Rep. MA 230 00-03,
Luftfahrttechnisches Handbuch (LTH), Nov. 1999.

[81] Dobrev, Y., Initial Sizing von Airlinern mit Jet- und Turbopropellerantrieb, Thesis,
RWTH Aachen, Institute for Aeronautics and Astronautics (ILR), Aachen, Germany,
2008.

[82] Anton, E., Lammering, T., and Henke, R., “Fast Estimation of Top-Level Requirement
Impact on Conceptual Aircraft Designs,” 10th AIAA Aviation Technology, Integration,
and Operations (ATIO), Vol. AIAA-2010-9220, AIAA, Fort Worth, Sept. 2010.

[83] Markish, J., Valuation Techniques for Commercial Aircraft Program Design, Department
of aeronautics and astronautics, Massachusetts Institute of Technology, Boston, June
2002.

[84] Clark, P., Buying the Big Jets, Ashgate Publishing Limited, Hampshire/UK, 2nd ed.,
2007.

[85] Stuessel, R., “Der Luftverkehr und seine Anforderungen an zukuenftige Verkehrs-
flugzeuge,” DGLR Jahrbuch 1982 , Vol. I, Sept. 1983, pp. 024/01–024/32.

[86] Raymer, D. P., Airplane Design: A Conceptual Approach, AIAA Education Series, Ame-
rican Institute of Aeronautics and Astronautics, Inc., Washington, D.C. 20024, 2nd ed.,
1992.

[87] Koeppen, C. and Carl, U., “Erfassung und Bewertung von Systemen im Flugzeugent-
wurf,” Deutscher Luft- und Raumfahrtkongress 2002 , Technische Universitaet Hamburg-
Harburg, 2002.
Bibliography 143

[88] Mavris, D. N., DeLaurentis, D. A., Bandte, O., and Hale, M. A., “A Stochastic Approach
to Multi-disciplinary Aircraft Analysis and Design,” Amercian Institute of Aeronautics
and Astronautics (AIAA), Reno, NV, Jan. 1998.

[89] FAA, Airworthiness Standards: Transport Aircraft Category Airplanes, No. Part 25 in
Federal Regulations, Federal Aviation Administration (FAA), Washington D.C., July
2010.

[90] “Massenanalysen,” Tech. rep., Luftfahrttechnisches Handbuch (LTH), Jan. 2008.

[91] Howe, D., Aircraft Conceptual Design Synthesis, Professional Engineering Publishing
Limited, London and Burs St Edmunds, UK, 1st ed., 2000.

[92] Beltramo, M. N., Trapp, D. L., Kimoto, B. W., and Marsh, D. P., “Parametric Study
of Transport Aircraft Systems Cost and Weight,” Tech. Rep. NASA CR 151970, Science
Application, Inc., Los Angeles, California, April 1977.

[93] Torenbeek, E., Synthesis of Subsonic Airplane Design, Delft University Presss, 1982.

[94] Kundu, A. K., Aircraft Design, Cambridge Aerospace Series, Cambridge University Press,
New York, NY 10013-2473, USA, 1st ed., 2010.

[95] Berg, F., Vergleich und Erweiterung verschiedener Methoden zur Massenabschaetzung
einzelner Komponenten im Flugzeugvorentwurf , Master thesis, RWTH Aachen, Institute
of Aeronautics and Astronautics (ILR), Aachen/Germany, 2009.

[96] Roskam, J., Component Weight Estimation, Vol. Part V of Airplane design / Jan Roskam,
DARcorporation, Lawrence, Kan., 2nd ed., 1989.

[97] vom Baur and Hippe, “Statistische Massenabschaetzung: Hydraulische und pneumatische
Anlagekomponenten Transportflugzeuge,” Tech. Rep. MA 704 22-02, Luftfahrttechnisches
Handbuch (LTH), March 1984.

[98] Fernandes da Moura, E. J., Vergleich verschiedener Verfahren zur Massenprognose von
Flugzeugbaugruppen im fruehen Flugzeugentwurf , Diploma thesis, Hamburg University of
Applied Science, Hamburg/Germany, Feb. 2001.

[99] Jackson, P., Munson, K., and Peacock, L., Jane’s All the World’s Aircraft 2007-2008 ,
Vol. 98, Jane’s Information Group Limited, Couldsen, U.K., 1st ed., May 2007.

[100] Schmidt, A., Lapple, M., and Kelm, R., “Advanced Fuselage Weight Estimation for the
New Generation of Transport Aircraft,” 56th Annual Conference, Bellevue, Washington,
May 19-21 , Society of Allied Weight Engineers, Inc., Bellevue, Washington, May 1997,
p. 48, L. R. ”Mike” Hackney Award.

[101] Keilig, T. and Schmidt, A., “Gewichtsprognose von CFK-Ruempfen fuer zukuenftige
Passagierflugzeuge,” DGLR Jahrbuch 2002 , Vol. II, 2002.
144 Bibliography

[102] Scholz, D., Entwicklung eines CAE-Werkzeuges zum Entwurf von Flugsteuerungs- und
Hydrauliksystemen, Vol. 262 of Fortschritt-Berichte VDI , VDI, Duesseldorf, Germany,
1997.

[103] Scholz, D., “Computer Aided Engineering for the Design of Flight Control and Hydraulic
Systems,” Society of Automotive Engineers (SAE), Vol. 105, 1997, pp. 203–212.

[104] Koeppen, C. and Carl, U., “Kopplung und Sensitivitaet des Einflusses von Systemen auf
den Flugzeugentwurf,” Deutscher Luft- und Raumfahrtkongress 2003 , Technische Univer-
sitaet Hamburg-Harburg, 2003.

[105] Calinski, D., Carl, U., and Koeppen, C., “Prognose des Leistungsbedarfs und der Masse
Elektrischer Bordnetze im Flugzeugentwurf,” Deutscher Luft- und Raumfahrtkongress
2003 , Technische Universitaet Hamburg-Harburg, 2003.

[106] Tsatsaronis, G., “Thermoeconomic analysis and optimization of energy systems,”


Progress in Energy and Combustion Science, Vol. 19, No. 3, 1993, pp. 227–257.

[107] Camberos, J. A. and Moorhouse, D. J., Exergy Analysis and Design Optimization for
Aerospace Vehicles and Systems, No. 238 in Progress in Astronautics and Aeronautics
Series, AIAA, 1st ed., Sept. 2011.

[108] Figliola, R. and Tipton, R., “An Exergy-Based Methodology for Decision-Based design
and Integrated Aircraft Thermal Systems,” 2000 Word Aviation Conference, AIAA/SAE,
San Diego, CA, Oct. 2000.

[109] Figliola, R., Tipton, R., and Li, H., “Exergy Approach to Decision-Based Design of
Integrated Aircraft Thermal Systems,” Journal of Aircraft, Vol. 40, No. 1, Jan. 2003,
pp. 49–55.

[110] Vargas, J. and Bejan, A., “Integrative thermodynamic optimization of the environmental
control system of an aircraft,” International Journal of Heat and Mass Transfer , Vol. 44,
No. 20, Oct. 2001, pp. 3907–3917.

[111] Pellegrini, L., Gandolfi, R., da Silva, G., and de Oliveira Jr., S., “Exergy Analysis as a
Tool for Decision Making in Aircraft Systems Design,” Reno, Jan. 2007.

[112] Gandolfi, R., Pellegrini, L., da Silva, G., and de Oliveira Jr., S., “Exergy Analysis Applied
to a Complete Flight Mission of a Commercial Aircraft,” Reno, Jan. 2008.

[113] Gandolfi, R., Pellegrini, L., and de Oliveira Jr., S., “More Electrical Aircraft Analysis
Using Exergy as a Design Comparison Tool,” Reno, Jan. 2010.

[114] Krus, P., “Complete Aircraft Simulation for Distributed System Design,” Proceedings of
the International Conference on Recent Advantages in Aerospace Actuation Systems and
Components, Toulouse, France, June 2001.
Bibliography 145

[115] Ros, T., Budinger, M., Reysset, A., and Mare, J.-C., “Modelica Preliminary Design
Library for Electomechanical Actuators,” 4th International Workshop on Aircraft System
Technologies, Shaker, Hamburg, April 2013, pp. 149–158.

[116] Jordan, P. and Schmitz, G., “Development of a Modelica Library Covering Multiple De-
sign Phase for Environmental Control Systems,” 3rd International Workshop on Aircraft
System Technologies, Shaker Verlag, Hamburg, March 2011, pp. 123–131.

[117] Jordan, P. and Schmitz, G., “Progress in TEMPO Project: A Modelica Library for Scal-
able Modelling of Aircraft Environmental Control Systems,” 4th International Workshop
on Aircraft System Technologies, Shaker, Hamburg, April 2013, pp. 71–80.

[118] Krus, P., Johansson, B., and Austrin, L., “Concept Optimization of Aircraft Systems
Using Scaling Models,” Recent Advances in Aerospace Actuation Systems and Compon-
ents, Toulouse, France, Nov. 2004.

[119] Liscouet, J., Budinger, M., and Mare, J.-C., “Technology Assessment on Aircraft-Level:
Modeling of Innovative Aircraft Systems in Conceptual Aircraft Design,” 3rd Interna-
tional Workshop on Aircraft System Technologies, Shaker Verlag, Hamburg, March 2011,
pp. 55–64.

[120] Pahl, G., Beitz, W., Feldhusen, J., and Grote, K.-H., Engineering Design: A Systematic
Approach, Springer-Verlag, London, 3rd ed., 2007.

[121] Bals, J., Hofer, G., Pfeiffer, A., and Schallert, C., “Virtual Iron Bird - A Multidisciplinary
Modelling and Simulation Platform for New Aircraft System Architectures,” Deutscher
Luft- und Raumfahrtkongress 2005 , 2005.

[122] Jomier, T., “MOET - Final Technical Report,” Public Report MOET-0.02-AF-DEL-
PublicReport-0001-09-R1.0, Airbus Operations S.A.S., Dec. 2009.

[123] van Driel, E., Slingerland, R., and Antoine, N., “A Framework for Aircraft Power Sys-
tems Optimization,” Structures, Structural Dynamics, and Materials Conference, AIAA,
Newport, Rhode Island, May 2006.

[124] Bornholdt, R., Luedders, H., and Thielecke, F., “Entwicklung einer Ganzheitlichen
Entwurfs- und Analyse-Methodik fuer Innovative Flugzeugsystem-Architekturen,” 61.
Deutscher Luft- und Raumfahrtkongress, Deutsche Gesellschaft fuer Luft- und Raum-
fahrt (DGLR), Berlin, Germany, Sept. 2012.

[125] Hanke, S., Pufe, S., and Mare, J.-C., “A Methodology for Aircraft Power System Archi-
tectures Analysis,” 1rd International Workshop on Aircraft System Technologies, Shaker
Verlag, Hamburg, March 2007, pp. 59–68.

[126] Liscouet-Hanke, S., “A Simulation Framework for Aircraft Power System Architecting,”
26th Congress of the International Council of the Aeronautical Sciences (ICAS), Anchor-
age, Alaska, 2008.
146 Bibliography

[127] Liscouet-Hanke, S., Mare, J.-C., and Pufe, S., “Simulation Framework for Aircraft Power
System Architecting,” Journal of Aircraft, Vol. 46, No. 4, July 2009, pp. 1375–1380.

[128] Frischemeier, S., “Power-by-Wire-Technologie fuer die Flugsteuerung - Systeme und Be-
wertung,” DGLR Jahrbuch 1997 , Vol. I, 1997, pp. 705–714.

[129] Blanchard, B. S. and Fabrycky, W. J., Systems Engineering and Analysis, Pearson Edu-
cation Limited, Essex/England, 5th ed., 2013.

[130] Bandte, O., A Probalistic Multi-Criteria Decision Making Technique for Conceptual and
Preliminary Aerospace Systems Design, Ph.D. thesis, Georgia Institute of Technology,
Sept. 2000.

[131] Dixon, J. R., Design Engineering; Inventiveness, Analysis, and Decision Making,
McGraw-Hill Book Company, New York, NY, 1966.

[132] Mavris, D. N. and Kirby, M. R., “Technology Identification, Evaluation, and Selection
for Commercial Transport Aircraft,” Vol. SAWE Paper No. 2456, SAWE, Inc., La Mesa,
CA 91942-2110, May 1999.

[133] Biltgen, P. T. and Mavris, D. N., “Technique for Concept Selection Using Interactive
Probabilistic Multiple Attribute Decision Making,” Journal of Aerospace Computing, In-
formation, and Communication, Vol. 6, No. 1, Jan. 2009, pp. 51–67.

[134] de Tenorio, C., Methods for Collaborative Conceptual Design of Aircraft Power Architec-
tures, Ph.D. thesis, Georgia Institute of Technology, Atlanta, GA, Aug. 2010.

[135] Economon, T. D., Copeland, S. R., Alonso, J. J., Zeinali, M., and Rutherford, D., “Design
and Optimization of Future Aircraft for Assessing the Fuel Burn Trends of Commercial
Aviation,” 49th AIAA Aerospace Sciences Meeting, AIAA, Orland, Jan. 2011.

[136] Peoples, R. and Willcox, K., “A Value-Based MDO Approach to Assess Business Risk
for Commercial Aircraft Design,” 10th AIAA/ISSMO Multidisciplinary Analysis and Op-
timization Conference, AIAA, Albany/NY, Aug. 2004.

[137] Peoples, R. and Willcox, K., “Value-Based Multidisciplinary Optimization for Commer-
cial Aircraft Design and Business Risk Assessment,” Journal of Aircraft, Vol. 43, No. 4,
July 2006, pp. 913–921.

[138] West, T., Hosder, S., and Winter, T., “Quantification of Margins and Uncertainties
for Aerospace Systems using Stochastic Expansions,” 52nd Aerospace Sciences Meeting,
AIAA, National Harbor/MD, Jan. 2014.

[139] d’Ippolito, R., Donders, S., Tzannetakis, N., Auweraer, H. V. d., and Vandepitte, D.,
“Integration of Probabilistic Methodology in the Aerospace Design Optimization Pro-
cess,” 46th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and Mate-
rials Conference, Structures, Structural Dynamics, and Materials and Co-located Confe-
rences, AIAA, Austin/TX, April 2005.
Bibliography 147

[140] Smith, N. and Mahadevan, S., “Probabilistic Methods for Aerospace System Conceptual
Design,” Journal of Spacecraft and Rockets, Vol. 40, No. 3, May 2003, pp. 411–418.

[141] Hanson, J. M. and Beard, B. B., “Applying Monte Carlo Simulation to Launch Vehicle
Design and Requirements Verification,” Journal of Spacecraft and Rockets, Vol. 49, No. 1,
Jan. 2012, pp. 136–144.

[142] Rudolph, P. K. C., “High-Lift Systems on Commercial Subsonic Airliners,” NASA Con-
tractor Report 4746, Ames Research Center, Moffet Field, California, Sept. 1996.

[143] SAE, Air Conditioning Systems for Subsonic Airplanes, No. ARP 85F in Aerospace
Recommaned Practices, Society of Automotive Engineers (SAE), 2012.

[144] Mevenkamp, C. and Galzin, G., “Climate Control and HVAC Simulation for Occupied
Spaces Using Open Source software,” 3rd International Workshop on Aircraft System
Technologies, Shaker, Hamburg, March 2011, pp. 113–122.

[145] Mueller, D., Schmidt, M., and Mueller, B., “Impact of Systems Integration on Fuel Effi-
ciency in Preliminary Aircraft Design,” 3rd International Workshop on Aircraft System
Technologies, Shaker, Hamburg, March 2011, pp. 279–288.

[146] Romani, R. and Goes, L. C., “Cabin Temperature Control Model for Commercial Air-
craft,” AIAA Modeling and Simulation Technologies Conference, American Institute of
Aeronautics and Astronautics, Aug. 2012.

[147] Schumacher, T., de Villiers, E., and Jackson, A. P., “Climate Control and HVAC Simu-
lation for Occupied Spaces Using Open Source software,” 3rd International Workshop on
Aircraft System Technologies, Shaker, Hamburg, March 2011, pp. 309–317.

[148] Tu, Y. and Lin, G., “Dynamic Simulation of Humid Air Environmental Control System,”
40th International Conference on Environmental Systems, International Conference on
Environmental Systems (ICES), American Institute of Aeronautics and Astronautics,
July 2010.

[149] Webel, O., Ruetten, M., and Wagner, C., “Impact of Systems Integration on Fuel Effi-
ciency in Preliminary Aircraft Design,” 3rd International Workshop on Aircraft System
Technologies, Shaker, Hamburg, March 2011, pp. 289–295.

[150] Wu, C. and Ahmed, N. A., “Aircraft cabin flow pattern under unsteady air supply,”
29th AIAA Applied Aerodynamics Conference, American Institute of Aeronautics and
Astronautics, Honolulu, HI, June 2011.

[151] Slingerland, R. and Zandstra, S., “Bleed Air versus Electric Power Off-Takes from a
Turbofan Gas Turbine over the Flight Cycle,” 7th AIAA Aviation Technology, Integration,
and Operations (ATIO), Amercian Institute of Aeronautics and Astronautics (AIAA),
Belfast, Northern Ireland, Sept. 2007.

[152] Truckenbrodt, E., Fluidmechanik , Vol. 2, Springer, Berlin, Germany, 4th ed., 1996.
148 Bibliography

[153] Huenecke, K., Die Technik des modernen Verkehrsflugzeuges, Motorbuch Verlag,
Stuttgart/Germany, 2nd ed., 2000.

[154] “ATA-30: Ice and Rain Protection,” Training Manual JAMF ATA 30, Lufthansa Tech-
nical Training GmbH, Frankfurt/Main, Germany, July 2000.

[155] Whalen, E. A., Broeren, A. P., and Bragg, M. B., “Characteristics of Runback Ice Accre-
tions and Their Aerodynamic Effects,” Tech. Rep. DOT/FAA/AR-07/16, Federal Avia-
tion Administration (FAA), Washington D.C. 25091, April 2007.

[156] Shaw, R. J., “NASA’s Aircraft Icing Analysis Program,” Technical Memorandum TM-
88791, NASA Lewis Research Center, Cleveland, Ohio, Sept. 1986.

[157] Bidwell, C. S. and Potapczuk, M. G., “User’s Manual for the NASA Lewis Three-
Dimensional Ice Accretion Code (LEWICE 3D),” Tech. Rep. NASA TM 105974, NASA
Lewis Research Center, Cleveland, Ohio, Dec. 1993.

[158] Wright, W. B., “Validation Results for LEWICE 3.0,” Contract Report CR-2005-213561,
NASA QSS Group, Inc., Cleveland, Ohio, March 2005.

[159] Wright, W. B., “User’s Manual for LEWICE Version 3.2,” Contract Report CR-2008-
214255, NASA QSS Group, Inc., Cleveland, Ohio, Nov. 2008.

[160] Blitzer, L., Hedde, T., and Guffond, D., “ONERA three-dimensional icing model,” AIAA
Journal , Vol. 33, No. 6, June 1995, pp. 1038–1045.

[161] Myers, T. G., “Extension to the Messinger model for aircraft icing,” AIAA Journal ,
Vol. 39, No. 0, Jan. 2001, pp. 211–218.

[162] Martin, H., “Heat and Mass Transfer between Impinging Gas Jets and Solid Surface,”
New York, Academic Press, Inc., Vol. 13, 1977, pp. 1–60.

[163] Wrig, “An Evaluation of Jet Impingement Heat Transfer Correlations for Piccolo Tube
Application,” 42nd AIAA Aerospace Sciences Meeting and Exhibit, AIAA, Reno, NV,
Jan. 2004.

[164] Saeed, F., Morency, F., and Paraschivoiu, I., “Numerical Simulation of a Hot-Air Anti-
Icing System,” 38th Aerospace Sciences Meeting and Exhibit, AIAA, Reno, NV, Jan.
2000.

[165] Papadakis, M. and Wong, S.-H. J., “Parametric Investigation of a Bleed Air Ice Protection
System,” 44th AIAA Aerospace Sciences Meeting and Exhibit, AIAA, Reno, NV, Jan.
2006.

[166] Hua, J., Kong, F., and Liu, H. H., “Unsteady Thermodynamic CFD Simulation of Aircraft
Wing Anti-Icing Operation,” .
Bibliography 149

[167] Hua, J. and Liu, H. H., “Fluid Flow and Thermodynamic Analysis of a Wing Anti-
Icing System,” Canadian Aeronautics and Space Journal , Vol. 51, No. 1, March 2005,
pp. 35–40.

[168] Liu, H. H. and Hua, J., “Three-Dimensional Integrated Thermodynamic Simulation for
Wing Anti-Icing System,” Journal of Aircraft, Vol. 41, No. 6, 2004, pp. 1291–1297.

[169] Weiner, F., “Ice, Rain, Fog, and Frost Protection,” Tech. Rep. SAE AIR 1168/4, SAE
International, Warrendale, PA 150596, July 1990.

[170] Sherif, S., Pasumarthi, N., and Bertlett, C., “A semi-empirical model for heat transfer
and ice accretion on aircraft wings in supercooled clouds,” Cold regions sciences and
technology, Vol. 26, No. 3, 1997, pp. 165–179.

[171] Jones, A. R. and Lewis, W., “Recommended Values of Meteorological Factors to be


Considered in the Design of Aircraft Ice-Prevention Equipment,” NACA Technical Note
NACA-TN-1855, NACA Ames Aeronautical Laboratory, Moffet Field, California, March
1949.

[172] Hacker, P. T. and Dorsch, R. G., “A summary of Meteorological Conditions Associ-


ated with Aircraft Icing and a Proposed Method of Selection Design Criterions for Ice-
Protection Equipment,” NACA Technical Note NACA-TN-2569, NACA Lewis Flight
Propulsion Laboratory, Cleveland, Ohio, Nov. 1951.

[173] Lewis, W. and Bergrun, N. R., “A Probability Analysis of the Meteorological Factors
Conductive to Aircraft Icing in the United States,” NACA Technical Note NACA-TN-
2738, NACA Ames Aeronautical Laboratory, Moffet Field, California, July 1952.

[174] Lammering, T., Weber, G., Thierer, S., Schaedler, P., Ried, G., and Schneider, T., “Lieb-
herr State-of-the-Art Fly-by-Wire Flight Control System for Commercial Transport Air-
craft,” 4th International Workshop on Aircraft System Technologies, Shaker, Hamburg,
April 2013, pp. 127–137.

[175] Lammering, T., Marka, C., Dreyer, N., and Grom, T., “Liebherr Distributed Electronic
Flight Control System and Smart Components,” Deutscher Luft- und Raumfahrtkongress
2013 , Stuttgart/Germany, 2013.

[176] Weber, G., Lammering, T., Thierer, S., Schaedler, P., Ried, G., and Schneider, T., “The
Liebherr Fully Integrated FCS Design - A Case Study,” 13th AIAA Aviation Technology,
Integration, and Operations (ATIO), Amercian Institute of Aeronautics and Astronautics
(AIAA), Los Angeles, California, Aug. 2013.

[177] Steinke, T., Entwicklung einer Methodik zur Modellierung und Analyse von Systemstruk-
turen im Flugzeugvorentwurf , Diploma thesis, RWTH Aachen, Institute of Aeronautics
and Astronautics (ILR), Aachen/Germany, 2010.

[178] Groening, S., Systemmodell Hochauftrieb - Modellierung der Systemkomponenten und


Abschaetzung der Einfluesse auf das Gesamtsytem Flugzeug, Thesis, RWTH Aachen,
Institute of Aeronautics and Astronautics (ILR), Aachen/Germany, June 2009.
150 Bibliography

[179] Anderson, R. D., Flora, C. C., Nelson, R. M., Raymond, E. T., and Vincent, J. H.,
“Development of Weight and Cost Estimates for Lifting Surfaces with Active Controls,”
NASA Contractor Report 144937, Boeing Commercial Airplane Company, Seatle, WA,
March 1976.

[180] Chaves-Vargas, M., Aufbau einer Datenbasis zur Modellierung konventioneller


Hochauftriebssysteme, Thesis, RWTH Aachen, Institute of Aeronautics and Astronau-
tics (ILR), Aachen/Germany, Jan. 2010.

[181] “A300 - Maintenance Facility Planning MFP,” Tech. Rep. A.MFP E 00A, Airbus Indus-
tries (Product Support), 31707 Blagnac Cedex, France, Sept. 1972.

[182] “A300-600 - Facility Planning Manual (Maintenance Facility Planning MFP),” Tech. Rep.
C. MFP, Airbus Industries (Customer Service Directorate), 31707 Blagnac Cedex, France,
March 1983.

[183] “A310 - Maintenance Facility Planning MFP,” Tech. Rep. B.MFP E 00A, Airbus Indus-
tries (Product Support), 31707 Blagnac Cedex, France, Sept. 1980.

[184] “A318 - Facility Planning Manual (Maintenance Facility Planning MFP),” Tech. Rep. I.
MFP, Airbus Industries (Customer Services), 31707 Blagnac Cedex, France, July 2002.

[185] “A319 - Facility Planning Manual (Maintenance Facility Planning MFP),” Tech. Rep. H.
MFP, Airbus Industries (Customer Services), 31707 Blagnac Cedex, France, Jan. 1996.

[186] “A320 - Facility Planning Manual (Maintenance Facility Planning MFP),” Tech. Rep. D.
MFP, Airbus Industries (Customer Services), 31707 Blagnac Cedex, France, Sept. 1985.

[187] “A321 - Facility Planning Manual (Maintenance Facility Planning MFP),” Tech. Rep. G.
MFP, Airbus Industries (Customer Services), 31707 Blagnac Cedex, France, Sept. 1992.

[188] “A330 - Maintenance Facility Planning MFP,” Tech. rep., Airbus Industries (Product
Support), 31707 Blagnac Cedex, France, Feb. 2009.

[189] “A340-200/-300 - Facility Planning Manual (Maintenance Facility Planning MFP),” Tech.
Rep. F. MFP, Airbus Industries (Customer Services), 31707 Blagnac Cedex, France, Feb.
1992.

[190] “A340-500/-600 - Facility Planning Manual (Maintenance Facility Planning MFP),” Tech.
Rep. F. MFP, Airbus Industries (Customer Services), 31707 Blagnac Cedex, France, April
2001.

[191] “A380 - Maintenance Facility Planning MFP,” Tech. rep., Airbus Industries (Product
Support), 31707 Blagnac Cedex, France, Oct. 2009.

[192] Muenz, T., “High Lift System Weight Break Down,” Tech. rep., Liebherr-Arospace Lin-
denberg GmbH, Lindenberg/Germany, Nov. 2010.
Bibliography 151

[193] Risse, K., Lammering, T., Anton, E., Franz, K., and Hoernschemeyer, R., “An Integrated
Environment for Preliminary Aircraft Design and Optimization,” 8th AIAA Multidisci-
plinary Design Optimization Specialist Conference, AIAA, Honolulu, HI, April 2012.

[194] Lammering, T., Risse, K., Franz, K., Peter, F., and Stumpf, E., “Assessment of Innova-
tive Leading Edge Devices Considering Uncertainties in Conceptual Design,” 12th AIAA
Aviation Technology, Integration, and Operations (ATIO), AIAA, Indianapolis, IN, Sept.
2012.

[195] Montgomery, D. C. and Runger, G. C., Applied Statistics and Probability for Engineers,
John Wiley & Sons, Inc., 4th ed., 2007.

[196] Johnson, N. L., Kotz, S., and Balakrishnan, N., Continuous Univariate Distributions,
John Wiley & Sons, Inc., New York, NY 10158-0012, 2nd ed., 1995.

[197] Lammering, T., Anton, E., Risse, K., Franz, K., and Hoernschemeyer, R., “Influence of
Off-Design Performance on Design Synthesis of Laminar Aircraft,” Journal of Aircraft,
Vol. 49, No. 5, Oct. 2012, pp. 1324–1335.

[198] Anton, E., Lammering, T., and Henke, R., “A comparative analysis of operations to-
wards fuel efficiency in civil aviation,” Applied Aerodynamics: Capabilities and Future
Requirements, Bristol, BS8 1TR, UK, July 2010.

[199] Lammering, T., Anton, E., Risse, K., Franz, K., and Hoernschemeyer, R., “Gains in Fuel
Efficiency: Multi-Stop Missions vs. Laminar Aircraft,” 11th AIAA Aviation Technology,
Integration, and Operations (ATIO), AIAA, Virginia Beach, Vi, Sept. 2011.

[200] Risse, K., Anton, E., and Henke, R., “Methodology for Flying Qualities Prediction and
Assessment in Preliminary Aircraft Design,” AIAA Aviation Technology, Integration, and
Operations (ATIO), AIAA, Sept. 2010.

[201] Franz, K., Lammering, T., Risse, K., Anton, E., and Hoernschemeyer, R., “Economics
of Laminar Aircraft Considering Off-Design Performance,” 8th AIAA Multidisciplinary
Design Optimization Specialist Conference, AIAA, Honolulu, HI, April 2012.

[202] Lammering, T., Franz, K., Risse, K., Hoernschemeyer, R., and Stumpf, E., “Aircraft
Cost Model for Preliminary Design Synthesis,” 50th AIAA Aerospace Sciences Meeting
including the New Horizons Forum and Aerospace Exposition, AIAA, Nashville, TN, Jan.
2012.

[203] Sahai, A., Anton, E., and Stumpf, E., “An Interdisciplinary Approach for a Virtual
Air-Traffic System Simulation for Subjective Assessment in VR Environments,” 28th
International Congress of the Aeronautical Sciences (ICAS), ICAS, Brisbane, Australia,
Sept. 2012.
152 Bibliography

[204] Sahai, A., Wefers, F., Anton, E., Stumpf, E., and Vorlaender, M., “Interdisciplinary
Auralization of Take-off and Landing Procedures for Subjective Assessment in Virtual
Reality Environments,” 18th AIAA/CEAS Aeroacoustics Conference (33rd AIAA Aeroa-
coustics Conference), Amercian Institute of Aeronautics and Astronautics (AIAA), Den-
ver, Colorado, June 2012.

[205] Lammering, T., Anton, E., Risse, K., Franz, K., and Hoernschemeyer, R., “Influence of
Off-Design Performance on the Design of Aircraft with Laminar Flow Technology,” 11th
AIAA Aviation Technology, Integration, and Operations (ATIO), AIAA, Virginia Beach,
Vi, Sept. 2011.

[206] Braun, S., Lammering, T., Risse, K., and Hoernschemeyer, R., “Assessment of Poten-
tial for Drag Reduction Using a Box-Wing Configuration,” Deutscher Luft- und Raum-
fahrtkongress 2011 , Bremen, Sept. 2011.

[207] Henke, R., Lammering, T., and Anton, E., “Impact of an Innovative Quiet Regional
Aircraft on the Air Transportation System,” Journal of Aircraft, Vol. 47, No. 3, June
2010, pp. 875–886.

[208] Lammering, T., Anton, E., and Henke, R., “Technology Assessment on Aircraft-Level:
Modeling of Innovative Aircraft Systems in Conceptual Aircraft Design,” 10th AIAA
Aviation Technology, Integration, and Operations (ATIO), Vol. AIAA-2010-9072, AIAA,
Fort Worth, Sept. 2010.

[209] Chambers, M. C., Ardema, M. D., Patron, A. P., Hahn, A. W., Miura, H., and Moore,
M. D., “Analytical Fuselage and Wing Weight Estimation of Transport Aircraft,” NASA
Technical Memorandum TM-110392, NASA, Ames Research Center, 1996.

[210] Dern, P., Ouriaghi, Z., Schueltke, F., and Wittgens, S., Methode zur Bestimmung der
Fluegel- und Rumpfmasse im Flugzeugvorentwurf , Project thesis, RWTH Aachen, Insti-
tute of Aeronautics and Astronautics (ILR), Aachen/Germany, 2011.

[211] Lammering, T., Anton, E., and Henke, R., “Validation of a Method for Fast Estima-
tion of Transonic Aircraft Polars and its Application in Preliminary Design,” Applied
Aerodynamics: Capabilities and Future Requirements, Bristol, BS8 1TR, UK, July 2010.

[212] Gur, O., Mason, W. H., and Schetz, J. A., “Full Configuration Drag Estimation,” 27th
AIAA Applied Aerodynamic Conference, AIAA, San Antonio, TX, June 2009.

[213] Gur, O., Mason, W. H., and Schetz, J. A., “Full-Configuration Drag Estimation,” Journal
of Aircraft, Vol. 47, No. 4, July 2010, pp. 919–939.

[214] Horstmann, K.-H., Ein Mehrfach-Traglinienverfahren und seine Verwendung fuer Ent-
wurf und Nachrechnung nichtplanarer Fluegelanordnungen, DFVLR forschungsbericht,
German Aerospace Center (DLR), Dec. 1987.

[215] Horstmann, K. H., Engelbrecht, T., and Liersch, C., “A Multiple Lifting Line Procedure
for Design and Analysis of Non-planar Wings,” April 2007.
Bibliography 153

[216] Anderson, J. D. J., Fundamentals of Aerodynamics, Mechanical Engineering Series,


Mcgraw-Hill Higher Education, 3rd ed., 2001.

[217] Roskam, J., Airplane Design, Part VI: Preliminary Calculation of Aerodynamic, Thrust
and Power Characteristics, Vol. 6, Roskam Aviation and Engineering Corporation, Ot-
tawa, Kansas, 2nd ed., 1985.

[218] Malone, B. and Mason, W. H., “Multidisciplinary optimization in aircraft design using
analytic technology models,” Journal of Aircraft, Vol. 32, No. 2, April 1995, pp. 431–438.

[219] Kurzke, J., “GasTurb Software Package, Vers. 11,” Tech. rep., Munich, Germany, 2009.

[220] Diedrich, A., Hileman, J. I., Tan, D., Willcox, K., and Spakovszky, Z., “Multidisciplinary
Design and Optimization of the Silent Aircraft,” 44th AIAA Aerospace Sciences Meeting
and Exhibit, AIAA, Reno, NV, Jan. 2006.

[221] Seitz, A., Broichhausen, K., Seifert, J., and Donnerhack, S., “An Integrated Para-
metric Model for Engine and Aircraft Design and Performance Optimization,” 44th
AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit, AIAA, Hartford, CT,
July 2008.

[222] Hall, C. A., Schwartz, E., and Hileman, J. I., “Assessment of Technologies for the Silent
Aircraft Initiative,” Journal of Propulsion and Power , Vol. 25, No. 6, Nov. 2009, pp. 1153–
1162.

[223] Vera-Morales, M. and Hall, C. A., “Modeling Performance and Emissions from Aircraft
for the Aviation Integrated Modeling Project,” Journal of Aircraft, Vol. 47, No. 3, May
2010, pp. 812–819.

[224] Luedders, H., Kirchner, F., and Thielecke, F., “SysFuel+ - Eine Bewertungsplattform fuer
ein Multifunktionales Brennstoffzellensystem auf Gesamtflugzeugebene,” 60. Deutscher
und Luft- und Raumfahrtkongress, Deutsche Gesellschaft fuer Luft- und Raumfahrt
(DGLR), Bremen, Germany, Sept. 2011.

[225] Oates, G. C., Aircraft Propulsion Systems Technology and Design, AIAA Education Se-
ries, AIAA, Jan. 1989.

[226] Scheiderer, J., Angewandte Flugleistung: Eine Einfuehrung in die operationelle Flug-
leistung vom Start bis zur Landung, Springer-11774, Springer-Verlag Berlin Heidelberg,
Berlin, Heidelberg, 2008.

[227] Jenkinson, L., Simpkin, P., and Rhodes, D., Civil Jet Aircraft Design, Arnold, Lon-
don/UK, 1st ed., 1999.

[228] “Standard Method of Estimating Comparative Direct Operating Costs of Turbine Po-
wered Transport Airplanes,” Tech. rep., Air Transport Association of America (ATA),
Inc., 1967.
154 Bibliography

[229] “Direct Operating Costs in Air Transport,” Tech. rep., College of Aeronautics of the
Cranfield Institute of Technology, Cranfield/UK, 1975.

[230] “DHL Method for Definition of the Performance and Direct Operating Costs of Com-
mercial Fixed Wing Aircraft,” Tech. rep., Deutsche Lufthansa AG, Cologne, Germany,
1976.

[231] “A New Method for Estimating Current and Future Transport Aircraft Operating Eco-
nomics,” Contract Report NASA CR-145190, American Airlines, Tulsa, Ok, March 1978.

[232] “Long Range Aircraft AEA Requirements,” Tech. rep., Association of European Airlines
(AEA), Brussels, Belgium, Dec. 1989.

[233] “Short-Medium Range Aircraft AEA Requirements,” Tech. rep., Association of European
Airlines (AEA), Brussels, Belgium, Dec. 1989.

[234] Gordon, P. and Habibie, I. A., “Analyse Massgeblicher DOC-Berechnungsmethoden fuer


Regional-, Kurz-, Mittel- und Langstreckenflugzeuge,” DGLR Jahrbuch 1992 , Vol. II,
March 1992, pp. 655–664.

[235] Meyer, S., Ein Vergleich von DOC-Methoden hinsichtlich der Kosten fuer Gebuehren,
Thesis, Hamburg University for Applied Sciences, Hamburg/Germany, Nov. 2004.

[236] Schuermann, T., Entwicklung eines Betriebskostenmodells fuer Verkehrsflugzeuge, Thesis,


RWTH Aachen, Institute of Aeronautics and Astronautics (ILR), Aachen/Germany, Nov.
2008.

[237] 777-200/300 - Airplane Characteristics for Airport Planning, Boeing Commercial Air-
planes, 31707 Blagnac Cedex, France, d ed., Oct. 2002.

[238] Werner-Spatz, C., Flugzeuggesamtentwurf mit Zirkulationskontrolle am Hochauftriebssys-


tem, Ph.D. thesis, Technische Universitaet Braunschweig, Shaker Verlag Aachen, 2010.

[239] Strohmeyer, D. and Seubert, R., “Improvement of a Preliminary Design and


Optimization Program for the Evaluation of Future Aircraft Projects,” 7th
AIAA/USAF/NASA/ISSMO Symposium on Multidisciplinary Analysis and Optimi-
zation, AIAA, St. Louis, MO, Sept. 1998.

[240] Saltelli, A., Chan, K., and Scott, E. M., Sensitivity Analysis, John Wiley & Sons, Ltd,
West Sussex PO19 8SQ, UK, 2000.

[241] Monner, H. P. and Riemenschneider, J., “Background and recent results of the European
project ’Smart High Lift Devices for Next Generation Wings’,” 1st EASN Association
Workshop on Aerostructures, Paris, France, Oct. 2010.

[242] Monner, H. P. and Riemenschneider, J., “Morphing high lift structures: Smart leading
edge device and smart single slotted flap,” 6th European Aeronautics Days - Aerodays
2011 , Madrid, Spain, March 2011.
Bibliography 155

[243] Monner, H. P., “SADE - Project Periodic Report 2009,” Tech. rep., German Aerospace
Center (DLR), June 2009.

[244] Monner, H. P., “SADE - Project Periodic Report 2010,” June 2010.

[245] Monner, H. P., “SADE Newsletter 2011,” Tech. rep., German Aerospace Center (DLR),
Brunswick/Germany, June 2011.

[246] Amiryants, G., Ishmuratov, F., and Malyutin, V., “Selectively Deformable Structures
for Design of Adaptive Wings ”Smart” Elements,” 1st EASN Association Workshop on
Aerostructures, Paris, France, Oct. 2010.

[247] Thuwis, G., Abdalla, M., and Guerdal, Z., “A Variable Stiffness Skin for Morphing High-
lift Devices,” 1st EASN Association Workshop on Aerostructures, Paris, France, Oct.
2010.

[248] Kintscher, M., Heintze, O., and Monner, H. P., “Structural Design of a Smart Leading
Edge Device for Seamless and Gapless High Lift Systems,” 1st EASN Association Work-
shop on Aerostructures, Paris, France, Oct. 2010.

[249] Kirn, J., Lorkowski, T., and Bair, H., “Development of Flexible Matrix Composites
(FMC) for Fluidic Actuators in Morphing Systems,” 1st EASN Association Workshop on
Aerostructures, Paris, France, Oct. 2010.

[250] Ameduri, S., “A Smart Leading Edge Architecture aimed at Preserving Laminar Regime,”
1st EASN Association Workshop on Aerostructures, Paris, France, Oct. 2010.

[251] Kintscher, M., “Experimental Testing of a Smart Leading Edge High Lift Device for
Commercial Transportation Aircrafts,” 27th Congress of the International Council of the
Aeronautical Sciences (ICAS), Nice, France, Sept. 2010.

[252] Schrauf, G., “Status and Perspective of Laminar Flow,” The Aeronautical Journal ,
Vol. 109, No. 1102, Dec. 2005, pp. 639–644.

[253] Nicholls, K., “NACRE - New Aircraft Concepts Research,” Aviation & the Environment:
Where are we Going? , RAeS Greener by Design, Oct. 2008.

[254] Wong, P., Maina, M., and Nicholls, K., “High Aspect Ratio Low Sweep (HARLS) Wing
Parametric Design Investigations,” Applied Aerodynamics: Capabilities and Future Re-
quirements, Bristol, BS8 1TR, UK, July 2010.

[255] Seitz, A. and Horstmann, K.-H., “Design Studies on NLF and HLFC Applications at
DLR,” ICAS 2010 Proceedings, Nice, France, 2010.

[256] Allison, E., Kroo, I., Sturdza, P., Suzuki, Y., and Martins-Rivas, H., “Aircraft Conceptual
Design with Natural Laminar Flow,” ICAS 2010 Proceedings, Nice, France, 2010.

[257] Doganis, R., Flying Off Course: airline economics and marketing, Routledge, New York,
NY 10016, 4th ed., 2010.
156 Bibliography

[258] Peter, F., Lammering, T., Risse, K., Franz, K., and Stumpf, E., “Economic Assessment of
Morphing Leading Edge Systems in Conceptual Aircraft Design,” 51th AIAA Aerospace
Sciences Meeting including the New Horizons Forum and Aerospace Exposition, AIAA,
Grapevine\TX, Jan. 2013.
A Systems Definition

In the following section a brief overview and description of the different aircraft systems ac-
cording to the ATA-100 chapters are given. The brief description includes main functions,
components and power sources. It is oriented at today’s civil transport aircraft with modern
but conventional technology. For a more detailed information further reading is suggested, for
example in Collinson [30], Fielding [31], Moir [32, 33], Roskam [34] or Scholz [35, ch. 12].

ATA chapter Group Topic


00 - 18 Aircraft General
21 Environmental control system
22 Auto flight
23 Communication
24 Electrical power
25 Equipment and furnishing
26 Fire protection
27 Flight controls
28 Fuel
29 Hydraulic power
Aircraft Systems
30 Ice and rain protection
31 Instrumentation, indication and recording
32 Landing gear
33 Lights
34 Navigation
35 Oxygen
36 Bleed air
38 Water and waste
49 Auxiliary power unit
51 - 57 Structure
60 - 67 Propeller/Rotor
70 - 84 Power Plant
91 - 116 Miscellaneous

Table A.1: Definition of ATA100-chapters [12].

157
158 Chapter A. Systems Definition

ATA-21 – Environmental control system (ECS) The environmental control system com-
prises all systems for climate control (temperature, humidity and sterilization of inducted air)
and pressurization of the cabin, flight deck, cargo compartment and avionics bay. Main compon-
ents are the air conditioning packs, mixing units and the ducting systems for ventilation. Also
temperature sensors, control units as well as valves for both inflow and outflow are components
of the ECS. The conventional ECS’s main power source is bleed air.

ATA-22 – Auto flight The auto flight system comprises the avionics for auto flight and flight
management functions. Main components are the flight control computers, which have direct
interfaces with the flight control system, navigation system, air data system, and the cockpit.
The auto flight system uses mainly electric DC power.

ATA-23 – Communication Both the systems for internal and external communication
belong to ATA-23. This includes internal intercom systems, analog and digital antennas, as
well as receiver and transmitter units. The main human-machine interface is the flight deck.
The communication systems generally use electric DC power.

ATA-24 – Electric power The electric power system is responsible for generation and dis-
tribution of electric power (both AC and DC) in the aircraft. During standard operations,
AC power is generated by electric generators from the engine shafts. AC power is then trans-
formed into DC via transform-rectifier units. During failure cases, the electric system can be
additionally supplied from batteries, auxiliary power units (APU) or ram air turbines (RAT).
All components for generation, distribution and control of electric power are allocated to the
ATA-24 chapter. Exceptions are the APU, which forms an own ATA chapter (ATA-49), as well
as the RAT, which is included in the hydraulic system (ATA-29).

ATA-25 – Equipment and furnishing The equipment and furnishing system comprises
all elements of cabin equipment such as seats, carpets, overhead bins, partition walls, and
isolation. Also galleys, lavatories, as well as in-flight entertainment systems belong to this ATA
chapter. If the aircraft is equipped with a cargo loading system, it is also categorized as ATA-25
subsystem. Often different elements of the equipment and furnishing system are also grouped
as standard and operator’s items. Standard items comprise all items that each aircraft of a
specific type is equipped with, whereas operator’s items comprise special and customized items
that vary from operator to operator.

ATA-26 – Fire protection The fire protection system is responsible for detecting, annun-
ciating, and fighting open fire or smoldering fire. The system comprises smoke detectors in all
critical areas of the aircraft such as engines, avionic bays, cabin, galleys or lavatories. If smoke
is detected, flight and cabin crew are warned. Active firefighting systems are integrated with
the engines that allow the flight crew to extinguish an engine fire. Mobile fire extinguishers are
positioned in all relevant areas of the aircraft for manual operations.
159

ATA-27 – Flight controls The flight control system (FCS) or ATA-27 can be further grouped
into primary (PFC) and secondary flight controls (SFC). Primary flight controls provide the
aircraft-level function of 3-axis control and comprise ailerons, elevators, rudders, and spoilers.
Secondary flight controls provide high lift and trim functions. It includes leading and trailing
edge high lift devices, as well as any means of trim such as trimable horizontal stabilizers (THS).
The flight control system comprises all components from cockpit controls to actuation systems
that move the single flight control surfaces. Thus, flight control computers (FCC) as well as
actuators, gears and transmission systems are included. Not included are the aerodynamic
control surfaces, which belong to the aircraft structure. The conventional flight control system
is a main consumer of hydraulic energy. Hydraulic energy is converted by hydraulic-mechanical
actuators into deflection of the control surfaces. In modern aircraft, more and more electro-
hydraulic actuators (EHA) or electro-mechanic actuators (EMA) are used, which makes the
flight control system also a significant consumer of electric energy. However, electric actuation
technology is required that provides larger power densities in order to match performance of
hydraulic actuators in order to further facilitate the more-electric aircraft concept.

ATA-28 – Fuel The fuel system provides the propulsion system and the APU with sufficient
fuel; also in some aircraft the fuel system is used for trim during cruise. It comprises tanks,
ducting, valves, vents, pumps and the fuel management system. The main power consumers
are the fuel pumps, which generally work with electric AC power. The fuel system has direct
interfaces e.g. with the propulsion system, the flight control system and the auto flight system.

ATA-29 – Hydraulic power The hydraulic power system is responsible for generation
and distribution of hydraulic power in the aircraft. Historically hydraulic systems of transport
aircraft used a working pressure for 3000 psi, whereas the current trend goes to higher pressures.
The Airbus A380 and A350 aircraft, for example, use working pressures of 5000 psi. All
components for generation, distribution, and control of hydraulic power are allocated to the
ATA-27 chapter. Main components are pumps, ducts, reservoirs, accumulators, servojacks, and
hydraulic fluid. During standard operations hydraulic pressure is generated by engine driven
pumps in two or three independent hydraulic circles. Further generation of hydraulic power
by electrical pumps and/or the ram air turbine is possible. Also cross-feed of hydraulic energy
between the single circuits is possible via a power transfer unit (PTU), although no hydraulic
fluid is shared or transferred.

ATA-30 – Ice and rain protection All systems for ice and rain protection of the aircraft
are summarized in ATA-30. Main components are wing, engine and sensor anti-icing systems.
Bleed air is conventional used for wing and engine anti-icing; electrical heating on the other
hand for keeping sensors and valves free of ice. Where wing anti-icing is only switched on
during operating below 22,000 ft under icing conditions, engine and sensor anti-icing is always
on. Rain protection is mainly limited to the pilot’s windshield wipers.

ATA-31 – Instrumentation, indication and recording The ATA-31 comprises all cockpit
instrumentations and indication devices. Further, means for data recording such as flight data
and cockpit voice recorders belong to this ATA chapter. Todays commercial transport aircraft
are generally equipped with glass cockpits, and thus instrumentation and indication are mostly
digital display devices.
160 Chapter A. Systems Definition

ATA-32 – Landing gear All landing gear systems are summarized in ATA-32. This includes
the landing gear structure, actuation, tires, breaks, steering and control systems, e.g. the anti-
skid. Landing gear systems are today a major consumer of hydraulic energy, which is required
for deployment and retraction, breaking, and for steering. Additionally, landing gear structural
weight is a notable share of overall operating weight empty.

ATA-33 – Lights All internal and external lighting is comprised in ATA-33. Most important,
this includes e.g. cabin lighting, cockpit lighting, as well as stroke and navigation lights.
Depending on the technology, the systems use either electric DC or AC power. Recently, LED
technology gains an increased importance also for application in aeronautics. This shifts the
power requirements further towards DC current.

ATA-34 – Navigation All navigation aids are comprised within ATA-34. This includes
antennas, transmitter and receiver units, as well as avionic computers. The navigation systems
have direct interfaces with e.g. the flight control system, auto flight system and the air data
system. Navigation avionics used electric DC power.

ATA-35 – Oxygen Civil transport aircraft carry emergency oxygen for sudden loss in cabin
pressurization. For passengers emergency oxygen masks are stored in the overhead bins, which
are released automatically in an emergency situation. Emergency oxygen for passengers is
obtained from a chemical reaction within individual cartridges. For the flight and cabin crew
oxygen tanks are provided additionally. Please note that ATA-35 does not comprise cabin
pressurization and climate control functions, which belong to the environmental control system
(ATA-21, see above).

ATA-36 – Bleed air For transport aircraft with conventional systems technology, bleed air
is mainly used for environmental control and ice protection. The bleed air system comprises all
subsystems and components that are responsible for bleeding the air from the engine, and for
ducting and controlling the bleed air flow to the single consumers. Generally hot air is bled from
the hot engine core and pre-cooled with cool fan air to a temperature level of approximately
200-250 ◦ C before it is further distributed to the consumers.

ATA-38 – Water and waste Drinking water is used in both lavatories and galleys. Systems
for storage (fresh and waste), ducting and control are comprised in ATA-38. Interfaces for
refilling and draining are required. Electric pumps that work with AC power provide the
necessary pressure to the system; overall power consumption, however, is generally negligible.

ATA-49 – Auxiliary power unit Today, the auxiliary power unit is a small gas turbine
that is installed in the rear of the fuselage and that can provide the aircraft with power during
ground operations and in emergencies. It has a direct interface with the fuel system and with
the cockpit controls. For the future the integration of fuel cell technology as possible APU is
likely. Also the fuel cell could then be used as primary power source for electric DC power.
B Deviations in Power Off-Take
Corrections

The following table lists deviations in fuel flow, specific thrust and engine core temperature
between MICADO and GasTurb for 3 exemplarily engines, and 2 operating conditions each.

General Electric CF6-80C2 (e.g. Airbus A310 or McDonnell Douglas MD-11)


Operating Condition 1 Operating Condition 2
Altitude 0 ft 35,000 ft
Mach 0.25 – 0.80 –
Thrust 175 kN 50 kN
Shaft power off-takes 0 - 350 kW 0 - 350 kW
Bleed air off-takes 0 - 2.5 kg/s 0 - 2.5 kg/s
Mean deviation ṁf uel 0.144 % 0.140 %
Standard deviation ṁf uel 0.059 % 0.052 %
Maximum deviation ṁf uel 0.247 % 0.204 %
Mean deviation SF C 0.141 % 0.141 %
Standard deviation SF C 0.059 % 0.051 %
Maximum deviation SF C 0.244 % 0.202 %
Mean deviation t45 0.252 % -0.008 %
Standard deviation t45 0.150 % 0.327 %
Maximum deviation t45 0.465 % -0.773 %

Table B.1: Deviations in corrected fuel flow between MICADO and GasTurb for the CF6-80C2
engine.

161
162 Chapter B. Deviations in Power Off-Take Corrections

CFM International CFM56-5B4P (e.g. Airbus A320 family)


Operating Condition 1 Operating Condition 2
Altitude 0 ft 35,000 ft
Mach 0.25 – 0.75 –
Thrust 70 kN 18 kN
Shaft power off-takes 0 - 600 kW 0 - 600 kW
Bleed air off-takes 0 - 3.5 kg/s 0 - 3.5 kg/s
Mean deviation ṁf uel -0.168 % -0.427 %
Standard deviation ṁf uel 0.164 % 0.412 %
Maximum deviation ṁf uel -0.493 % -1.293 %
Mean deviation SF C -0.164 % -0.428 %
Standard deviation SF C 0.165 % 0.419 %
Maximum deviation SF C -0.492 % -1.297 %
Mean deviation t45 0.174 % 0.205 %
Standard deviation t45 0.079 % 0.100 %
Maximum deviation t45 0.309 % 0.342 %

Table B.2: Deviations in corrected fuel flow between MICADO and GasTurb for the CFM56-
5B4P engine.

General Electric GE90-11B (e.g. Boeing 777-300ER)


Operating Condition 1 Operating Condition 2
Altitude 0 ft 35,000 ft
Mach 0.25 – 0.83 –
Thrust 295 kN 65 kN
Shaft power off-takes 0 - 200 kW 0 - 200 kW
Bleed air off-takes 0 - 1.5 kg/s 0 - 1.5 kg/s
Mean deviation ṁf uel 0.168 % 0.681 %
Standard deviation ṁf uel 0.129 % 0.563 %
Maximum deviation ṁf uel 0.407 % 1.749 %
Mean deviation SF C 0.171 % 0.710 %
Standard deviation SF C 0.131 % 0.561 %
Maximum deviation SF C 0.415 % 1.756 %
Mean deviation t45 0.263 % 0.914 %
Standard deviation t45 0.128 % 0.496 %
Maximum deviation t45 0.507 % 1.678 %

Table B.3: Deviations in corrected fuel flow between MICADO and GasTurb for the GE90-
115B engine.
C Top-Level Aircraft Requirements for
SA185 Aircraft

The following table lists top-level aircraft requirements that were used for design of the SA185
reference aircraft in scope of this thesis.

Design mission
Design payload mpaylaod,design 16780 kg
Design range Rdesign 4000 nm
Take-off distance T ODRdesign 2100 m
Landing distance LDTdesign 1650 m
Approach speed vapproach 145 KCAS
Time to climb TTC 25 min
Initial cruise altitude hinit,cr 33000 ft
Initial cruise Mach number Minit,cr 0.80
Design flight envelope
Maximum operating speed V MO 350 KCAS
Maximum operating Mach number MMO 0.84
Maximum operating altitude hM O 41000 ft
Maximum operating altitude OEI hM O,OEI 13000 ft
Accommodation
Passengers (business class) P AXbusiness 16
Seats abreast (business class) 4
Passengers (economy class) P AXeconomy 169
Seats abreast (economy class) 6
ULD type LD3-45W
Number of ULDS 12
Propulsion and configuration
Number of engines neng 2 (turbofan)
Conventional tube & wing configuration with under wing engines

Table C.1: Top-level requirements for the SA185 design synthesis.

163
D Sensitivities of 777 Design

This section shows the sensitivities of the 777 design according to the ones shown for the A320
design in scope of chapter 5.2.

D.1 Sensitivities towards Model Specific Input Parameters

10 10

5 5
∆ results, %

∆ results, %

0 0

-5 System mass -5 System mass


Bleed air Bleed air
-10 Electric power -10 Electric power
Hydraulic power Hydraulic power
-20 -10 0 10 20 -20 -10 0 10 20
∆ Tbleed air, % ∆ control surface deflection rate, %

Figure D.1: 777 – Sensitivity towards bleed Figure D.2: 777 – Sensitivity towards re-
air temperature. quired deflection speeds.

40 20

20 10
∆ results, %

∆ results, %

0 0

-20 System mass -10 System mass


Bleed air Bleed air
-40 Electric power -20 Electric power
Hydraulic power Hydraulic power
-20 -10 0 10 20 -20 -10 0 10 20
∆ Tcabin, % ∆ recirculation of cabin air, %

Figure D.3: 777 – Sensitivity towards re- Figure D.4: 777 – Sensitivity towards recircu-
quired cabin temperature. lation of cabin air.

165
166 Chapter D. Sensitivities of 777 Design

5.0 5.0

2.5 2.5
∆ results, %

∆ results, %
0.0 0.0

-2.5 System mass -2.5 System mass


Bleed air Bleed air
-5.0 Electric power -5.0 Electric power
Hydraulic power Hydraulic power
-20 -10 0 10 20 -20 -10 0 10 20
∆ span wing anti-icing, % ∆ number of MCUs, %

Figure D.5: 777 – Sensitivity towards span Figure D.6: 777 – Sensitivity towards number
position of wing anti-icing. of MCUs.

D.2 Sensitivities towards Aircraft Design Parameters

5.0 5.0

2.5 2.5
∆ results, %

∆ results, %

0.0 0.0

-2.5 System mass -2.5 System mass


Bleed air Bleed air
-5.0 Electric power -5.0 Electric power
Hydraulic power Hydraulic power
-20 -10 0 10 20 -20 -10 0 10 20
∆ MTOW, % ∆ OWE, %

Figure D.7: 777 - Sensitivity towards MTOW. Figure D.8: 777 – Sensitivity towards OWE.

10 5.0

5 2.5
∆ results, %

∆ results, %

0 0.0

-5 System mass -2.5 System mass


Bleed air Bleed air
-10 Electric power -5.0 Electric power
Hydraulic power Hydraulic power
-20 -10 0 10 20 -20 -10 0 10 20
∆ PAX, % ∆ cabin length, %

Figure D.9: 777 – Sensitivity towards number Figure D.10: 777 – Sensitivity towards cabin
of passengers. and fuselage length.
D.3 Sensitivities towards Top-Level Aircraft Design Parameters 167

10 10

5 5
∆ results, %

∆ results, %
0 0

-5 System mass -5 System mass


Bleed air Bleed air
-10 Electric power -10 Electric power
Hydraulic power Hydraulic power
-20 -10 0 10 20 -20 -10 0 10 20
∆ wing span, % ∆ wing reference area, %

Figure D.11: 777 – Sensitivity towards wing Figure D.12: 777 – Sensitivity towards wing
span. reference area.

D.3 Sensitivities towards Top-Level Aircraft Design


Parameters

10 10

5 5
∆ results, %

∆ results, %

0 0
MTOW MTOW
-5 System mass -5 System mass
Bleed air Bleed air
-10 Electric power -10 Electric power
Hydraulic power Hydraulic power
-20 -10 0 10 20 -20 -10 0 10 20
∆ mpayload,design, % ∆ PAX, %

Figure D.13: 777 – Sensitivity towards design Figure D.14: 777 – Sensitivity towards design
payload. passenger number.
168 Chapter D. Sensitivities of 777 Design

D.4 Sensitivities of Overall Design Synthesis

5.0 5.0

2.5 2.5
∆ results, %

∆ results, %
0.0 0.0
OWE
Block fuel
-2.5 -2.5 MTOW
OWE System mass
-5.0 Block fuel -5.0 Shaft power
MTOW TOFL
-20 -10 0 10 20 -20 -10 0 10 20
∆ total bleed air demand, % ∆ total bleed air demand, %
(a) Without design synthesis. (b) With design synthesis.

Figure D.15: 777 – Sensitivity towards total bleed air demand.

5.0 5.0

2.5 2.5
∆ results, %

∆ results, %

0.0 0.0
OWE
-2.5 OWE -2.5 Block fuel
Block fuel MTOW
-5.0 MTOW -5.0 System mass
System mass TOFL
-20 -10 0 10 20 -20 -10 0 10 20
∆ shaft power demand, % ∆ shaft power demand, %
(a) Without design synthesis. (b) With design synthesis.

Figure D.16: 777 – Sensitivity towards total shaft power demand.


Curriculum Vitae of the Author

Tim Lammering born on the 22nd of January 1981 in Moenchengladbach/Germany.

Professional Career

since 06/2012 Liebherr-Aerospace Lindenberg GmbH


Flight Controls and Actuation Systems
Team Leader System Verification and Integration
05/2007 - 05/2012 Institute of Aerospace Systems
researcher and Ph.D. candidate with focus on
preliminary aircraft design and systems integration
04/2011 - 05/2012 RWTH Aachen University
teaching assignment for the lectures:
- Aerospace Systems (Master’s Curriculum)
- Computation-based Aircraft Design (Master’s Curriculum)
- Aircraft Systems (Bachelor’s Curriculum)
02/2006 - 05/2007 CFD Consultants GmbH
internship and freelance project engineer

Education

09/2001 - 05/2007 RWTH Aachen University


Mechanical Engineering Diploma (Dipl.-Ing.)
with specialty in Aeronautical Engineering
08/2005 - 02/2006 Queen’s University of Belfast
1991 - 2000 St. Wolfhelm Gymnasium Schwalmtal
university entrance diploma (Abitur)
08/1997 - 02/1998 Bishop Fenwick High School Middletown/Ohio

169
List of Publications by the Author

Journal Publications

1. T. Lammering, E. Anton, K. Risse, K. Franz, and R. Hoernschemeyer, Influence of


Off-Design Performance on Design Synthesis of Laminar Aircraft, Journal
of Aircraft, ed. 49, vol. 5, p. 1324-1335, Oct. 2012

2. R. Henke, T. Lammering, and E. Anton, Impact of an Innovative Quiet Regional


Aircraft on the Air Transportation System, Journal of Aircraft, ed. 47, vol. 3, p.
875-886, June 2010

Conference Publications and Proceedings

1. T. Lammering, A. Sauterleute, B. Hauber, T. Schneider, J. Becker, C.Schaeper, and D.-U.


Sauer, Conceptual Design of a Battery-Powered High Lift System for Single-
Aisle Aircraft, in 52th AIAA Aerospace Sciences Meeting including the New Horizons
Forum and Aerospace Exposition, National Harbor/MD, 2014, AIAA-2014-0378

2. J. Becker, C. Schaeper, J. Muennix, D.-U. Sauer, T. Lammering, A. Sauterleute, B.


Hauber, and T. Schneider, Design of a Safe and Reliable Li-ion Battery System
for Applications in Airborne System, in 52th AIAA Aerospace Sciences Meeting
including the New Horizons Forum and Aerospace Exposition, National Harbor/MD,
2014, AIAA-2014-0380

3. T. Lammering, G. Weber, S. Thierer, P. Schaedler, G. Ried, and T. Schneider, Lieb-


herr State-of-the-Art Fly-by-Wire Flight Control System for Commercial
Transport Aircraft, in 4rd International Workshop on Aircraft System Technologies,
Hamburg, 2013, p. 127-137, ISBN 978-3-8440-1850-9

4. T. Lammering, C. Marka, N. Dreyer, and T. Grom, Liebherr Distributed Electronic


Flight Control System and Smart Components, in Deutscher Luft- und Raum-
fahrtkongress, Stuttgart, 2013

5. F. Peter, T. Lammering, K. Risse, K. Franz, and E. Stumpf, Economic Assessment


of Morphing Leading Edge Systems in Conceptual Aircraft Design, in 51th
AIAA Aerospace Sciences Meeting including the New Horizons Forum and Aerospace
Exposition, Grapevine/TX, 2013, AIAA-2013-0145

171
172 List of Publications by the Author

6. G. Weber, T. Lammering, S. Thierer, P. Schaedler, G. Ried, and T. Schneider, The


Liebherr Fully Integrated FCS Design - A Case Study , in 13th AIAA Aviation
Technology, Integration, and Operations (ATIO), Los Angeles/CA, 2013, AIAA-2013-
4332

7. T. Lammering, K. Risse, K. Franz, F. Peter, and E. Stumpf, Assessment of Innova-


tive Leading Edge Devices Considering Uncertainties in Conceptual Design,
in 12th AIAA Aviation Technology, Integration, and Operations (ATIO), Indianapolis/IN,
2012, AIAA-2012-5588

8. T. Lammering, K. Franz, K. Risse, R. Hoernschemeyer, and E. Stumpf, Aircraft Cost


Model for Preliminary Design Synthesis, in 50th AIAA Aerospace Sciences Meet-
ing including the New Horizons Forum and Aerospace Exposition, Nashville/TN, 2012,
AIAA-2012-0686

9. K. Risse, T. Lammering, E. Anton, K. Franz, and R. Hoernschemeyer, An Integrated


Environment for Preliminary Aircraft Design and Optimization, in 8th AIAA
Multidisciplinary Design Optimization Specialist Conference, Honolulu/HI, 2012, AIAA-
2012-1675

10. K. Franz, T. Lammering, K. Risse, E. Anton, and R. Hoernschemeyer, Economics of


Laminar Aircraft Considering Off-Design Performance, in 8th AIAA Multi-
disciplinary Design Optimization Specialist Conference, Honolulu/HI, 2012, AIAA-2012-
1760

11. T. Lammering, E. Anton, K. Risse, and K. Franz, Impact of Systems Integration on


Fuel Efficiency in Preliminary Aircraft Design, in 3rd International Workshop on
Aircraft System Technologies, Hamburg/Germany, 2011, p. 171-180, ISBN 978-3-8322-
9904-0

12. T. Lammering, E. Anton, K. Risse, K. Franz, and R. Hoernschemeyer, Influence of Off-


Design Performance on the Design of Aircraft with Laminar Flow Technol-
ogy , in 11th AIAA Aviation Technology, Integration, and Operations (ATIO), Virginia
Beach/VI, 2011, AIAA-2011-7017
173

13. T. Lammering, E. Anton, K. Risse, K. Franz, and R. Hoernschemeyer, Gains in Fuel


Efficiency: Multi-Stop Missions vs. Laminar Aircraft, in 11th AIAA Aviation
Technology, Integration, and Operations (ATIO), Virginia Beach/VI, 2011, AIAA-2011-
6885

14. T. Lammering, Impact of Innovative Quiet Regional Aircraft on the Air Trans-
portation System, Active High Lift and Impact on Air Transportation Symposium,
invited talk, Brunswick/Germany, April 2011

15. S. Braun, T. Lammering, K. Risse, and R. Hoernschemeyer, Assessment of Potential


for Drag Reduction Using a Box-Wing Configuration, in Deutscher Luft- und
Raumfahrtkongress 2011, DGLR, Bremen/Germany, 2011

16. T. Lammering, E. Anton, and R. Henke, Technology Assessment on Aircraft-Level:


Modeling of Innovative Aircraft Systems in Conceptual Aircraft Design, in
10th AIAA Aviation Technology, Integration, and Operations (ATIO) Conference, Fort
Worth/TX, 2010, AIAA-2010-9072

17. E. Anton, T. Lammering, and R. Henke, Fast Estimation of Top-Level Require-


ment Impact on Conceptual Aircraft Designs, in 10th AIAA Aviation Technology,
Integration, and Operations (ATIO) Conference, Fort Worth/TX, 2010, AIAA-2010-9220

18. T. Lammering, E. Anton, and R. Henke, Validation of a Method for Fast Es-
timation of Transonic Aircraft Polars and its Application in Preliminary
Design, in RAeS Applied Aerodynamics: Capabilities and Future Requirements, Bris-
tol, BS8 1TR, UK, 2010

19. E. Anton, T. Lammering, and R. Henke, A comparative analysis of operations to-


wards fuel efficiency in civil aviation, in RAeS Applied Aerodynamics: Capabilities
and Future Requirements, Bristol, BS8 1TR, UK, 2010

Você também pode gostar