22 LowCostRegionalAircraft FinalReport Final1
22 LowCostRegionalAircraft FinalReport Final1
written by
Group 22
The DSE is the final assignment of the Aerospace Engineering bachelor curriculum. It is intended to demon-
strate skill and knowledge acquired during the bachelor program and apply it to a specific and realistic prob-
lem. This problem is solved by a group of 10 students in a multidisciplinary setting. The Design Synthesis
Exercise can be divided into four phases: a planning phase, a requirements phase, a concept phase and a fi-
nal design phase. This report represents the conclusion of the final phase.
We would like to thank our principal tutor, asst. prof. ir. J. Sinke, and our coaches, ir. Y. Zhang and ir. H. Koorn-
neef, for their continued support and assistance during our project. We also would like to thank the staff of the
faculty of Aerospace Engineering, without whom several questions would not have been answered. Lastly, we
would also like to thank the faculty of Aerospace Engineering for the use of their facilities during this project.
Group 22
Delft, June 2018
ii
Contents
Preface ii
List of Abbreviations vi
List of Symbols xii
List of Figures xiii
List of Tables xv
1 Executive Overview 1
1.1 Project Description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Organisation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.3 Market Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.4 Detailed Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.5 Production Plan. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.6 Operations & Logistics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.7 Finance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.8 Future. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2 Market Analysis 7
2.1 Aviation Market . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2 Regional Market. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.3 Customer Requirements & Needs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.4 Competition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.5 Market Share & Unit Cost . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.6 Requirements Overview. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
3 Configuration & Layout 12
3.1 Final Trade-Off . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.2 Class-I & Class-II Weight Estimations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.3 External Configuration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.4 Internal Configuration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
4 Flight Performance 16
4.1 Engine Selection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
4.2 Balanced Field Length . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
4.3 Flight Path . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
4.4 Flight Performance Optimisation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
4.5 Payload-Range . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
4.6 Flight Envelope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
4.7 V -n Diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
4.8 Turning Performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
4.9 Verification & Validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
4.10 Sensitivity Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
5 Aerodynamics 33
5.1 Wing Optimisation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
5.2 Tail Sizing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
5.3 Resulting Wing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
5.4 Verification & Validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
6 Flight Dynamics 42
6.1 Result of Previous Iteration & Sizing. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
6.2 Detailed Sizing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
6.3 Control Surface Sizing. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
6.4 High Lift Devices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
6.5 Aircraft Flight Dynamics Characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
6.6 Dynamic Stability Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
6.7 Eigenmotion Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
iii
iv Contents
AC Alternating Current
ADIRU Air Data Inertial Reference Unit
ADSEE Aerospace Design and Systems Engineering Elements
AHM Aircraft Health Monitoring
ann Anneald
ANoPP Aircraft Noise Prediction Program
APU Auxiliary Power Unit
ATC Air Traffic Control
ATM Air Traffic Management
ATR Aerei da Trasporto Regionale or Avions de Transport Régional
AVIC Aviation Industry Corporation of China
AVL Athena Vortex Lattice
C Composite
c.g. Centre of Gravity
CBS Cost Breakdown Structure
CO2 Carbon Dioxide
CRFP Carbon Fibre Reinforced Polymer
DC Direct Current
DOC Direct Operating Costs
vi
List of Abbreviations vii
LC Latest Completion
LS Latest Start
LT Longitudinal Transverse
M Metal
MAC Mean Aerodynamic Chord
MCC Maintenance Control Centre
MFD Multi-Function Display
MIT Massachusetts Institute of Technology
MLDT Mean Logistic Delay Time
MLW Maximum Landing Weight
MMT Mean Maintenance Time
MTBF Mean Time Between Failures
MTBMA Mean Time Between Maintenance Actions
Mth Months
MTOW Maximum Take-Off Weight
MTTR Mean Time To Repair
MZFW Maximum Zero Fuel Weight
SF Synthetic Fibre
viii
List of Symbols ix
xiii
xiv List of Figures
4.1 Climb gradient and required thrust for each take-off segment. . . . . . . . . . . . . . . . . . . . . 16
4.2 Engine trade-off ranges for all the criteria. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
4.3 Criteria weights for engine trade-off. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
4.4 Engine trade-off. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
4.5 Parameters after complete take-off. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
4.6 Aircraft action and configuration at each take-off segment. . . . . . . . . . . . . . . . . . . . . . . 21
4.7 Parameters after completion of cruise. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.8 Performance overview. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
4.9 Turn performance parameters calculated at MTOW at 7000 ft. . . . . . . . . . . . . . . . . . . . . . 29
4.10 Sensitivity of performance parameters to the zero-lift drag coefficient (C D 0 = 0.056). . . . . . . . 31
4.11 Sensitivity of performance parameters to the aspect ratio (A eff = 10.5). . . . . . . . . . . . . . . . 31
4.12 Sensitivity of performance parameters to the wing area (S = 61.10 m2 ). . . . . . . . . . . . . . . . 31
4.13 Sensitivity of performance parameters to the maximum take-off weight (MTOW = 30,200 kg). . . 32
xv
xvi List of Tables
11.1 Unit price of current large regional aircraft [87, 45, 52, 15]. . . . . . . . . . . . . . . . . . . . . . . . 93
11.2 Direct operating costs overview. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
11.3 Flight cancellations due to cause [26]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
11.4 Flight cancellations and delays due to different causes. . . . . . . . . . . . . . . . . . . . . . . . . . 98
13.1 Scale to measure the severity and likelihood of technical risks. . . . . . . . . . . . . . . . . . . . . 103
13.2 Pre-mitigation risk map. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
13.3 Post-mitigation risk map. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
Transport 50 to 75 passengers with the lowest possible operational costs over a distance of 1500 to
2500 km with an aircraft from regional airports with a paved runway length of 1500 m.
Develop a regional aircraft for the transportation of 50 to 75 passengers over a range of 1500 to
2500 km maximum with the key design goal of lowest possible operational costs in 11 weeks with
10 TU Delft BSc. Aerospace Engineering students. The aircraft at least must be able to operate from
regional airports with tarmac runways of only 1500 m, be optimised for sustainability and com-
ply with latest safety and environmental regulations. The design of the aircraft shall implement
current and near future technologies and be feasible for production to start in five years.
Project Requirements The project has several initial requirements, which can be found below. These re-
quirements, together with other requirements following from the market and requirement analysis were as-
sessed during the design phase and the result can be found in the compliance matrix in Appendix E.
1.2. Organisation
The team consist of 10 members who are all BSc. students at the faculty of Aerospace Engineering. The team
is organised in a way that each member has both a managerial and technical role. The organisation can be
found in Figure 1.1. The roles were divided at the beginning of the project. After five weeks into the project,
the roles were evaluated within the team. At the start of the detailed design phase in week six, the evaluation
was done once more. After careful consideration, it was decided that everybody would keep their role for the
remaining weeks.
1
2 1. Executive Overview
Content Officer
Robbin Veldhuis
Report Design Officer
Paulien Uijterwaal
Organisational organogram
Chief
Communications & Sustainability Officer Regulations Officer Risk Manager Marketing Officer
Quality Officers
Documentation Max Heilig Vince v/d Arend Adrian C.K. Loke Alexander Piva
Willem Hoekman
Technical organogram
The largest growth for the regional market is expected to be in the Asia-Pacific region, especially in China,
whereas the most aircraft will be sold in the Asia-Pacific and North American region. As the market will ex-
pand in the regions stated above, the amount of airports in those regions will increase as well.
As a new competitor in the market, it is essential to determine the expected market share. Based on the
market share of competitors, the MW 5 project will take up to 10% of the market, translating into 600 aircraft
at a predicted unit selling price of 41 million USD.
In Figure 1.2 the range capability of the MW 5 is given from two different airports, Minneapolis in Fig-
ure 1.2a and Paris in Figure 1.2b. The ranges of the CRJ700 and the MRJ70 are provided for the purpose of
comparison.
MRJ70 MRJ70
CRJ700 CRJ700
MW5 MW5
45°N
60°N
1880 km
2500 kkm
m
2270
30°N 45°N
188 70 k km
222500
0k m
m
30°N
15°N
15°N
120°W 90°W 30°W 0° 30°E
(a) Range capability from Minneapolis. (b) Range capability from Paris.
The aircraft is designed for 76 passengers in an economy class seating arrangement. Different seating ar-
rangements are offered to give versatility to the aircraft. The differences can be found in the table below. The
aircraft has a default seating arrangement of 2 + 2, but other arrangements are available upon request by the
customer.
Class Seat pitch [cm] Seat width [cm] Recline [deg] Rows Passengers
Economy 78.7 43.2 3 19 76
Economy+ 94.0 43.2 5 16 64
Mixed 78.7/94.0 43.2 3/5 18 72
Flight Performance To comply with requirements as set by the customer, a flight performance analysis is
performed. From this, multiple important aspects are analysed such as the take-off distance, the range, the
fuel burn and a proper engine selection. After an initial analysis, it is found that the balanced field length is
the driving design requirement for the engine. From this, it followed that an engine of at least 53.5 kN was
necessary to comply with a take-off length of 1500 m. The selected engine will be a ’scaled-down’ version of
the P&W1215G, called the P&W1200G, and can be seen in Figure 1.4.
With the selected engine, it was determined that the required runway length for take-off is 1490 m. From
further analysis, it is found that the fuel burn during a flight is approximately 4810 kg if 76 passengers were
transported over a range of 2500 km and that it would take a little under 3.5 hours. Furthermore, some typical
flight parameters are analysed to determine the performance. These parameters were the maximum rate of
4 1. Executive Overview
climb, the turning performance, the landing performance, the payload-range diagram and the flight enve-
lope. Also, recommendations are given to optimise the fuel burn.
Aerodynamics In order to optimise the aerodynamics of the aircraft, several different procedures are fol-
lowed. First, using the program AVL and Vortex-Lattice theory, the 2D main wing is optimised. This planform
is optimised for minimum drag using an iteration of 5 parameters (the aspect ratio, two taper ratios and two
sweep angles). The results of this optimisation can be found below.
Table 1.2: Starting values based on statistics compared to end values after optimisation.
A λ1 λ2 Λ1 Λ2 D
Before optimisation 10.0 0.80 0.40 10 deg 25 deg 3.9 kN
After optimisation 10.5 0.68 0.44 12 deg 25 deg 3.2 kN
Next, the airfoil is selected. Since the design cruise Mach number is 0.78, the focus is placed on supercritical
airfoils. The wing requires two different airfoils; one at the root and one at the tip. The airfoils at the interme-
diate points and the kink are determined using interpolation. The results are that the Whitcomb supercritical
airfoil is used at the root and kink location, while the NASA SC21010 airfoil is used at the tip.
Subsequently, the different tail surfaces are sized and an appropriate airfoil is selected. Using statistical
data and relationships, the stabiliser surfaces are determined. Each stabiliser has two airfoils, one at the root
and one at the tip. For the horizontal stabiliser, the NACA 0010 airfoil is selected at the root and the NACA
0009 airfoil for the tip. The vertical stabiliser has a NACA 0011 airfoil on both the root and the tip. Figures of
the airfoils can be found below.
(a) Whitcomb, root and kink airfoil. (b) NASA SC21010, tip airfoil.
(a) NACA 0011, vertical stabiliser. (b) NACA 0010, root airfoil and (c) NACA 0009, tip airfoil and
horizontal stabilser. horizontal stabilser.
Flight Dynamics In order to check the stability of the aircraft, first, the control surfaces are sized. This is
done using an iterative method starting from previous results. A loading diagram and scissor plot are con-
structed to determine the final size of the tail.
Next, the control surfaces are sized. The ailerons are sized using historical guidelines. To reduce adverse
yaw caused by aileron deflection, differential aileron control is chosen. Next, using data from XFLR5, the ele-
vator is sized. It is assumed that the elevator spans the entire horizontal stabiliser, making sure the contribu-
tion to the moment coefficient is sufficient. Finally, using XFLR5 as well, the rudder is sized to accommodate
changes in sideslip angle. It was concluded that a rudder spanning from 10 to 90 % of the chord is sufficient
to meet the requirements.
Subsequently, the high lift devices are sized. The high lift devices make sure the aircraft is able to produce
the increase in lift coefficient required for take-off and landing. The main location of high lift devices is at the
leading and trailing edges of the main wing. On the leading edge, slats are added. For the trailing edge, several
options are evaluated. By looking at the required increase in lift, the complexity of the system and several
other parameters, a trade-off produced the final flap type. Slotted flaps proved sufficient for the design and
are therefore chosen.
1.5. Production Plan 5
The second part of the Flight Dynamics assessment consists of the stability of the aircraft. First, the equations
of motion are determined which are then used to assess the stability of the aircraft in several different modes.
From the eigenmotion analysis, displayed in Table 1.3, it can be seen that the symmetrical eigenmotions have
a negative real part, meaning that the oscillations are damped. For the asymmetrical eigenmotions the real
part of the eigenvalue is only negative for the aperiodic and Dutch roll. The eigenvalues for the symmetric
and asymmetric case can be found in Table 1.3.
Structures and Materials In order to determine the proper material, the focus is placed on the two most
prominent structural elements. These are the fuselage and the wings. An initial material trade-off followed by
a second more in-depth trade-off resulted in the choice of Al-Li 8090-T851 as the main material. The benefits
mostly come from the low-density and relatively low increase in material cost compared to other materials
while maintaining good producibility and maintainability. Several additional considerations are also made
with respect to smaller aircraft components. For these components, different materials were chosen. It was
determined that the control surfaces would be made from composites because of their light weight.
The loading diagrams for the aircraft are generated and, in combination with the flight envelope, used to
determine the critical load cases for the structures mentioned above. For both the fuselage and the wing, the
appropriate stringers are chosen. This resulted in Z-stringers for the fuselage and a combination of Z- and
hat-stringers for the wing.
Next, stress analysis using the von Mises stress criterion is used to calculate the required thicknesses of
the different elements. For the fuselage sizing, crippling, buckling and shear are taken into account. For the
wing, bending and torsion are assessed. It is shown that both sections are able to carry the required loads.
1.7. Finance
A financial analysis is performed to investigate the costs and profits for both airlines as well as the manufac-
turer of the aircraft. For the total program cost, different costs are found. The development cost for the MW 5
is estimated to be 1 billion USD or 1.6 million USD per aircraft. Production cost is estimated to be 30.4 million
USD per aircraft. A unit aircraft price is determined to be 38.6 million USD including a 20% profit margin.
As operating cost is an important factor for the airliner, those costs are analysed as well. A distinction is
made between direct operating costs and indirect operating costs. The direct operating costs consist of flying,
maintenance, depreciation and some other minor parts. It is found that for an operator the MW 5 will cost
2875 USD/block hour. Indirect operating cost mainly depends on the airline itself but is estimated to be 55%
of the direct operating cost. A total cost reduction for airlines operating MW 5 is around 8 to 11% compared
to airlines operating current regional aircraft.
For the return on investment (RoI) and the break-even point again a distinction is made between manu-
facturer and operator. For the manufacturer the RoI is expected to be 19% and for the operator 12.2% with a
competitive ticket price. The break-even point for the manufacturer is reached after 270 aircraft are sold. For
the operator, the BEP is reached in the ninth year.
1.8. Future
The DSE project ends at the detailed design phase. However, the development of MW 5 is not yet completed
and many steps have to be made before release to the market. These future steps can be divided into three
different phases: detailed design, development and operational lifetime of the aircraft. The first of these phase
will take up about 1.5 years. This is followed by half a year of tool manufacture and facility set-up during the
development phase. After this, the prototypes are produced, tested and the certification process is started.
Once the aircraft receives its type certification, it is ready to entire service.
2
Market Analysis
The aviation market is a dynamic market which is constantly changing. While steadily growing, the composi-
tion of aircraft in use changes all the time.Throughout the years new markets rose as the demand for air travel
increased. In this chapter, a market analysis is performed in order to get a feeling on the feasibility of intro-
ducing a new low-cost regional aircraft. In Section 2.1 the aviation market as a whole is described, followed
by a more detailed analysis of the regional market in Section 2.2. Furthermore, the customer requirements
followed by the competitors in the market are discussed. The market share and target unit price are then
tackled. Finally, an overview of all the system requirements is presented.
Figure 2.1: Air traffic growth and world economic growth [67].
Although the increase in the past 20 years has been steady, there were some crises which influenced the
growth of the aviation market. In Figure 2.2 it can be seen that in the last 40 years some crises stagnated the
growth of the market. However, historical trends express the capacity by the world aviation industry to always
recover from these, which makes it a matter of when it will recover more than if it will [65, 67].
It can be further found in Figure 2.2 that the aviation market consisted of more than 5 trillion RPKs in 2012.
According to different market outlooks [9, 14, 33, 66], it is expected that for the next 20 years the RPKs will
increase to almost 20 trillion. The enormous growth in aviation offers opportunities to new manufacturers,
as the estimated amount of aircraft demand will steadily increase.
Another trend in the aviation market is the development of more sustainable aircraft [47]. Some of the
earlier experienced crises were related to oil dependencies. Due to the significant influence that an increase in
oil price has on the airline profits, the demand for more sustainable aircraft is increasing. In addition, because
of the positive effect on the environment, regulations are getting stricter and stricter on sustainability factors.
This shift leads to more light-weight designs, more efficient engines and the use of electrical systems.
7
8 2. Market Analysis
2.4. Competition
An analysis of the competition is essential in order to understand the market requirements and opportunities.
In fact, once these are identified, the design focus can be determined. Furthermore, through a competition
analysis, it is possible to understand the potential share market that can be taken by the new aircraft design.
This is crucial to determine a business plan.
There are three main aircraft manufacturers who are dominating the regional aircraft industry. As it can
be seen from Figure 2.3, these are respectively Embraer, ATR and Bombardier. Two other manufacturers are
also to be taken into account, which are Mitsubishi and AVIC, as they have not yet entered the market, but
already have a market share due to outstanding orders [50]. All values shown in the figure are expected market
shares for each company in 20 years.
ATR
Embraer
20.5 % 25.0 %
14.6 %
Bombardier
22.8 %
11.8 %
5.3 %
Other
Mitsubishi
AVIC
Embraer is the market leader for the regional aircraft market. Their expected deliveries by 2036 is 2300 aircraft
with a 70 to 90 passenger seating capacity [33], and their focus is mainly on North America and Asia-Pacific.
Between their aircraft, the ERJ-170 ER is the one with the most similarities with respect to the design goal.
ATR is a slightly different aircraft manufacturer with respect to the others. Instead of turbofan engines,
they use turboprop engines and focus more on a slightly lower passenger capacity. They expect to sell around
2000 aircraft in the next 20 years and they focus mostly on the upcoming Asia-Pacific region. Their ATR 72 is
the most comparable aircraft and, because of such, it is regarded as a strong competitor.
Bombardier is the third biggest manufacturer. They have turbofan powered aircraft and expect to sell
around the same amount of aircraft as the 2 other competitors, namely 2000, in the coming 20 years. Their
focus is not only the Asia-Pacific region but also Europe. The CRJ700 is the aircraft which mostly relates to the
requirements of the MW 5.
The two other manufacturers, Mitsubishi and AVIC, are entering the market. Mitsubishi has concluded its
test flights and expects to enter the market in 2020. They already have 288 [76] outstanding orders and 184
options. AVIC is producing the Comac ARJ21 and is already delivering aircraft. Of the 300 orders, they have
delivered 5 aircraft [86].
In Table 2.1 the comparable aircraft can be found from the different manufacturers. This gives an idea of how
some of the parameters, such as MTOW, will look like for the design.
10 2. Market Analysis
Sys.Perf.1: The aircraft shall be dynamically stable for the entire range of centre of gravity excursions.
Sys.Perf.2: The aircraft shall be controllable for the entire range of centre of gravity excursions.
Sys.Perf.3: The cruise speed of the aircraft shall be in the range of Mach 0.73 to 0.83.
2.6. Requirements Overview 11
Sys.Perf.5: The aircraft shall be able to make a steep approach with a flight path gradient of at least
−10%.
Sys.Perf.6: The noise produced by the aircraft shall be less than 264 EPNdB.
Sys.Perf.7: In landing configuration the steady climb gradient shall not be less than 3.2% at an airspeed
of no more than 1.3Vstall .
Sys.Perf.8: The climb gradient during the first take-off segment with one engine inoperative shall be
positive.
Sys.Perf.9: The climb gradient during the second take-off segment with one engine inoperative shall be
more than 2.4%.
Sys.Perf.10: The climb gradient during the final take-off segment with one engine inoperative shall be
more than 1.2%.
Sys.Perf.11: The climb gradient during the en-route segment with one engine inoperative shall be more
than 1.1%.
Sys.Perf.12: The climb gradient during the approach segment with one engine inoperative shall be more
than 2.1%.
Sys.Perf.13: There shall be no uncontrollable ground-looping tendencies in 90° crosswind, up to a wind
velocity of 20 kts.
Sys.Perf.14: The airframe structure shall not fail under ultimate load for a duration of 3 seconds.
Sys.Perf.15: The number of emergency exits present shall be sufficient to comply with CS-25 regulations.
Sys.Perf.16: The maximum take-off mass of the aircraft shall be less than 35,000 kg.
Sys.Perf.17: The operational empty mass of the aircraft shall be less than 20,600 kg.
Sys.Perf.18: The aircraft shall be able to fly to an alternate airport 100 km away after maximum range
flown.
Sys.Perf.19: The aircraft shall be able to loiter at the alternate for at least 30 minutes.
Sys.Perf.20: The landing distance on tarmac shall be less than 1500 m.
Sys.O&L.1: Turnaround time of the aircraft between maximum payload flights shall be less than 30 min-
utes.
Sys.O&L.2: The aircraft shall have a lifetime of at least 25 years.
Sys.O&L.3: The aircraft shall be able to perform three to four daily operations.
Sys.O&L.4: Lateral ground clearance of the aircraft shall be higher than 5°.
Sys.O&L.5: The seat pitch shall be equal to 31 in.
Sys.O&L.6: Maintenance intervals shall be increased due to a safe-life design of components.
Sys.O&L.7: The aircraft shall not tip over backwards for the entire range of centre of gravity locations
during ground loading.
Sys.O&L.8: The aircraft shall not turn over for any possible loading configuration and turning radius.
Sys.O&L.9: The aircraft doors shall be compatible with airstairs.
Sys.O$L.10: All essential aircraft systems shall be operative with ground power only.
Sys.O&L.11: The wing span shall not exceed 35 m.
Sys.O&L.12: The total length shall not exceed 35 m.
Sys.Sus.1: Noise generated by the aircraft shall be less than the Chapter 14 level EPNdB.
Sys.Sus.2: Fuel consumption per passenger-kilometre shall be less than 120 g.
Sys.Sus.3: Aircraft structure shall be made of at least 90% recyclable materials.
Sys.Sus.4: Manufacturing techniques shall be non-toxic.
Sys.Sus.5: Engine emissions shall comply with current regulations.
Sys.Sus.6: The fuel burned during operation shall be less than 8900 kg per flight.
Sys.Sus.7: Employees shall not be subjected to dangerous gases during the production of the aircraft.
3
Configuration & Layout
In this chapter, the reasoning of choosing the final concept is shown along with final internal and external
configurations. Firstly, the final six concepts and final trade-off summary table are presented in section Sec-
tion 3.1. Afterwards, the weight estimations are shown, for both, Class-I and Class-II in section 3.2. Lastly, the
external configuration and internal layout are presented in section 3.3 and 3.4 respectively.
12
3.2. Class-I & Class-II Weight Estimations 13
It can be seen from Table 3.2 that the most promising concept is concept 13. With a total score of 0.86, it is
clearly better than the other concepts. Design 13 therefore, enters the detailed design phase. Although the
other designs did not make it to the detailed design phase, there are some points that are worth considering
for the final design. Therefore the other five designs are taken into consideration when going into the detailed
design phase.
Excellent (++) Good (+) Average (0) Undesirable (−) Poor (−−)
V L Wbegin
µ ¶ µ ¶
Range = ln (3.1)
g · TSFC D Wend
where V is the cruise speed, TSFC is the thrust-specific fuel consumption in kg/(N s), L/D is the lift over
drag ratio during the cruise, and Wbegin and Wend are the weights at the beginning and the end of the cruise,
respectively. With a payload weight of around 6900 kg based on 76 passengers averaging a weight of 92 kg
including luggage [49], the MTOW can be approximated as
a + Wpayload
MTOW = ³ ´ (3.2)
1 − b + M ff + M fftrapped
where M ff and M fftrapped are the fuel fraction and the trapped fuel fraction as a fraction of the MTOW. The
variables a and b define the linear relation for the OEW as a function of MTOW based on reference aircraft,
where a is the offset value and b the slope. With the results from the Class-I weight estimation, a more de-
tailed weight estimation can be performed. The Class-II weight estimation is based on the Raymer method
[75], which consists of 20 equations and 70 input variables such as aspect ratio, MTOW, taper ratio, etc. The
output of the Class-II weight estimation are weights of different components (for example, fuselage, vertical
stabiliser, horizontal stabiliser, avionics etc.). By adding up all those weights, the operational empty weight is
obtained.
The results of the Class-I and Class-II weight estimations are given in Table 3.3 in tonnes.
14 3. Configuration & Layout
Class Seat pitch [in] Seat width [in] Recline [deg] Rows Passengers
Economy 31 17 3 19 76
Economy + 37 17 5 16 64
Mixed 31/37 17 3/5 18 72
Top view
Seats Lavatories Galley Scale: 1:50
(a) Economy.
Top view
Seats economy+ Lavatory Galley Scale: 1:50
(b) Economy+.
Front view
Seats economy+ Seats economy Galley Lavatory Scale: 1:50
(c) Mixed.
2
W VLOF
TTOreq = + D TO + D g avg (4.1)
2g 0 s groundrollTO
where VLOF is the lift-off speed, s groundrollTO is the groundroll distance at take-off, D TO is the drag at take-off
experienced by the aircraft and D g avg is the average ground friction drag during take-off. The required thrust
for take-off until obstacle clearance was found to be of 84.4 kN, which gives 42.2 kN per engine. However,
looking into the balanced field length equation, which will be explained in more detail in Section 4.2, the
take-off thrust is found to be 53.5 kN per engine.
Looking into the take-off procedures segments, the aircraft performance at these phases are determined by
regulations. These are explained in Subsection 4.3.2. According to CS-25 [3], the aircraft should be able to fly
at specific climb gradients G at each of those segments with one engine inoperative (OEI). The equation used
to determine the thrust required is shown in Equation 4.2 [57]. The results are shown in Table 4.1.
Table 4.1: Climb gradient and required thrust for each take-off segment.
From Table 4.1, the second segment is the most critical. Comparing it with both the take-off until obstacle
clearance and the balanced field length, it can be seen that the balanced field length determines the mini-
mum required engine thrust. Therefore, one engine should at least provide a thrust of 53.5 kN at sea level.
Based on this, various available engines are identified. A maximum available thrust of 70 kN is set as the upper
limit, as otherwise the engine is considered to be overdesigned. The potential engines with their related thrust
can be seen in Table 4.4. To perform a proper trade-off, the following criteria and weights are set:
16
4.1. Engine Selection 17
• Thrust: as the selected engines provide a higher thrust than required, this criterion is included.
• TSFC: thrust specific fuel consumption determines the fuel consumption during the flight.
• Dry weight: the lower the weight of an engine the more favourable this criteria is.
• Flyover noise: for sustainability, a lower noise is considered to be more optimal, as noise plays an im-
portant role around regional airports. Between the various noise types, flyover noise is the most signif-
icant to consider at these airports. Nevertheless, if flyover noise is lower, normally also the other types
of noise are lower.
• NOx emissions: as CO2 emissions are directly related with TSFC and thus taken into account with that
criteria, NOx emissions need to be taken into account for sustainability as well. Other emissions types
are not considered for the trade-off as much less influential than the two above mentioned.
In order to evaluate the engines, criteria ranges are set for the trade-off in Table 4.2. Values in red are unac-
ceptable, which implies that the engines that have at least one of those are considered not fitting. This is so
determined as the aircraft should be comparable or better than competition for these criteria.
The thrust ranges are set as such because having a more than the minimum necessary thrust is advantageous.
In fact, for maintainability, it is better to not use full throttle, as expressed in [78].
For all the other criteria, an average value of the criteria of current engines was first identified and then the
ranges were created by increasing or decreasing by a certain percentage. For TSFC, 5% was chosen as there is
a small variation between different engines. Nevertheless, such percentage strongly decreases the fuel con-
sumption and thus the operational costs. For flyover noise, 1% was chosen for the small variations as well.
For dry weight and NOx emissions, a 15% range change was chosen. The reason is that the difference be-
tween the current engine values varies in a greater amount and that the influence below such % change is not
that strong.
The weighting of each criterion is defined as shown in Table 4.3. The TSFC, flyover noise and NOx emissions
were considered to be the most influential in terms of operational costs and sustainability.
Now that all the ranges are set and explained, the trade-off can be computed. This is shown in Table 4.4. All the
engines with a red value are eliminated as they are considered not fitting. By assigning 0 points for a yellow
value, 1 point for a light blue value and 2 points for a green value, it can be seen in Table 4.4 that the Pratt
& Whitney PW1215G is the most fitting engine for the aircraft design. An image of the engine can be seen in
Figure 4.1.
18 4. Flight Performance
However, as this engine delivers a maximum thrust of 66.6 kN, a proposal will be delivered to Pratt & Whitney
to manufacture a MW 5 customised scaled-down version of the PW1215G. In fact, due to the fact that 66.6 kN
of power is 20% higher than required, a scaled-down version could result in a reduction in engine dry weight
or a more efficient engine. Because the MW 5 is expected to deliver at least 600 aircraft, this translates into a
1200 engine order for Pratt & Whitney. This amount is expected to be high enough to make it a viable business
for the engine manufacturer. The proposal will be to design an engine with a maximum thrust of 58 kN. For
further calculations, the design thrust of 53.5 kN will be taken into account, as it is the minimum required
thrust.
1 Retrieved from:http://hatcheraviation.com/uploads/Pages%20from%20PHAK-chap%209.2%20Aircraft%20Performance.
pdf on 12/06/2018
4.3. Flight Path 19
The runway length is based on a possible engine failure at the most critical point, which is at V1 . The decision
speed V1 is chosen in such a way that when an engine failure occurs, the pilot is able to abort the take-off
and make a full stop on the runway or continue take-off to screen height with one engine out, resulting in the
same distance. This distance is defined as the balanced field length. The required take-off distance is equal
to the balanced field length, or 115% of the distance over screen height (10.7 m above the runway) with all
engines operative, whichever is greater [81].
In Figure 4.2, the balanced field length and the decision speed are shown graphically. It was found that the
balanced field length is 1490 m and that this is the limiting factor. The decision speed V1 is 70 m/s.
accelerate-stop
2,500 accelerate-climb
Distance required [m]
2,000
1,000
V1
500
0
0 10 20 30 40 50 60 70 80
Engine failure speed [m/s]
4.3.1. Take-Off
The take-off starts when the aircraft is at the runway and ends when the aircraft clears the screen height
which is set by regulations at 35 ft, or 10.7 m. The take-off consists of two parts: the ground roll and the air-
borne phase. The calculation of the take-off is based on Anderson [5]. Take-off is done at 1.1Vstall as required
by CS-25 [3] and is calculated with Equation 4.3. Here, C L maxTO is the maximum lift coefficient in take-off
configuration and is found to be 1.70. The stall speed was found to be 68.2 m/s.
s
MTOW 2 1
Vstall = (4.3)
S ρ C L maxTO
The acceleration during the ground roll determines how long it takes to reach the take-off speed. Drag and lift
can be calculated with Equation 4.4a and Equation 4.4b, where C D max is calculated with Equation 4.5. C L maxTO
is divided by 1.1 as this gives a safety factor for take-off. With the equations found in Anderson [5, 6], the
acceleration equation can be written as in Equation 4.6.
1 C L maxTO 1
L= ρ 0V 2 S (a) D = C D max ρ 0 V 2 S (b) (4.4)
2 1.12 2
20 4. Flight Performance
C L2max
TO
C D max = C D 0 + (4.5)
πAe
g0
a= (T − D − µr (W − L)) (4.6)
W
For the airborne phase of the take-off, a circular path is assumed. The radius of this path is calculated with
Equation 4.7. The load factor (n) is set to be 1.21 by regulations. The groundroll and airborne phase can be
seen in Figure 4.3 and the increase in speed can be found in Figure 4.4.
2
VLOF
R= (4.7)
g 0 (n − 1)
The fuel burned during this segment can be calculated using the thrust specific fuel consumption and the
thrust applied. Fuel burn per second can be calculated with Equation 4.8. Taking the time it takes to complete
the take-off, the total fuel burned can be calculated. The TSFC is given at sea-level and in coming stages, a
correction factor should be applied.
ṁ fuel = TSFC × T (4.8)
The fuel burn is calculated to be 62.3 kg. After clearance of the screen height, the following parameters for the
aircraft can be found in Table 4.5. Here, it can be seen that it takes approximately 28 s to take-off.
12 80
10
60
8
Speed [m/s]
Altitude [m]
6 40
4
20
2
0
0
0 200 400 600 800 1,000 1,200 0 200 400 600 800 1,000 1,200
Runway length [m] Runway length [m]
Figure 4.3: Runway covered to clearance of screen height. Figure 4.4: Speed increase during take-off.
First & second segment In the first segment, the landing gear is retracted. According to ESDU [68], the
retraction time is approximately 13 s. After completion, the aircraft steepens its flight path to climb to 122 m
(400 ft), which is set by regulations [3]. This short climbing phase is the second segment. The speed during
this segments increases 1.1 to 1.2 Vstall [57].
En-route climb segment In this phase, the high lift devices (HLDs) are retracted into cruise configuration
and the aircraft is accelerated. The limit on the acceleration is imposed by the maximum speed that the HLDs
can support while retracting. This speed is 90 m/s for HLDs that are retracted from a range between 18 to 24°
[42]. As an acceleration is performed and more thrust is required, the flight path angle for this phase was set
at 1° to minimise fuel consumption.
4.3. Flight Path 21
Final segment In the final segment, the aircraft is preparing for the climb to cruise altitude. The speed is
increased to the optimal speed to start climbing at the regulated minimum height of 457.2 m (1500 ft), which
is the altitude at which the take-off procedure is concluded. The optimal speed at this height to start climbing
is 120 m/s.
Depending on the configuration at which the aircraft is flying at every segment, the lift, drag and Oswald
factor change accordingly. The wing area does not change due to the used HLDs. These are based on the
analysis performed in Section 6.4. Furthermore, the equation to determine the thrust setting at each segment
is expressed as Equation 4.9.
Finally, the aircraft configuration for every segment is displayed in Table 4.6 and the displayed flight path in
Figure 4.5 and Figure 4.6. An overview of the characteristics can be found in Table 4.5. From this it can be
calculated that the total fuel burned up to this stage is approximately 108.4 kg.
150
400
100
Speed [m/s]
Altitude [m]
200
50
0
0
0 2,000 4,000 6,000 8,000 0 2,000 4,000 6,000 8,000
Distance covered [m] Distance covered [m]
Figure 4.5: Distance covered up to the climb. Figure 4.6: Speed increase during full take-off.
4.3.3. Climb
After the final take-off segment and the aircraft has climbed to an altitude of 457.2 m (1500 ft) above ground
level, the aircraft starts climbing to cruise altitude. For the climb to cruise altitude, it is important to look at
the time to climb and the amount of fuel burned.
22 4. Flight Performance
Time to climb In general, it is preferred to minimise the time to climb because the aircraft is much more fuel
efficient during the cruise. The time to climb can be minimised by climbing at the maximum rate of climb.
This is not the most optimum climb regarding fuel consumption, but it is, in fact, a good approximation
[5]. The airspeed at which the maximum rate of climb occurs increases with height. Therefore, a portion of
the available excess power must be used to accelerate [81]. In order to minimise the time to climb, the rate of
climb needs to be maximum at each altitude. The maximum rate of climb was calculated at different altitudes
between 457.2 m and 10,000 m altitude with increments of 100 m using Equation 4.10,
V dV P a − Pr
· ¸
ROC 1 + = , (4.10)
g dh W
where the right-hand side is the specific excess power and the expression between square brackets the kinetic
energy correction factor. The total time to climb is obtained by adding the increments in time between the
different altitude intervals.
X n · ∆h ¸
t= (4.11)
i =1 ROCi
The climb is an unsteady quasi-rectilinear climb: unsteady because the airspeed increases with increasing
height, and quasi-rectilinear, because the flight path angle remains more or less constant. During climb, the
flight path angle slowly decreases from 3.6° at the beginning of climb to 0.8° at cruise altitude. Because the
time required to climb to cruise altitude is approximately 20 minutes, the rate of change of flight path angle
is close to zero.
Fuel burned during climb The fuel burned during climb also needs to be taken into account. The fuel flow
at any moment in time depends on the thrust-specific fuel consumption and the thrust setting. The fuel flow
can be calculated using Equation 4.8 as mentioned earlier. When fuel is burned the aircraft becomes lighter
and the maximum rate of climb that can be achieved increases. The reduction in fuel weight is taken into
account when calculating the time to climb to cruise altitude and the fuel burned during the climb.
Limitations on the ROC The highest rate of climb that can be achieved is limited by ATC procedures which
restrict the calibrated airspeed of aircraft under 10,000 ft to 250 kts2 and the rate of change of cabin pressure.
Because the airspeed at which the maximum rate of climb occurs is generally higher than 250 kts, the maxi-
mum rate of climb can often not be achieved below 10,000 ft.
The other factor that limits the highest rate of climb is the rate of change of cabin pressure [81]. When the
cabin pressure changes too quickly, this can be unpleasant or even painful. In general, ambient pressure
changes are not perceived when the time rate of change of pressure is kept within the following limits:
dp
−30 < < 18 Pa/s
dt
The lower limit is applicable to climb and the upper limit applies to descent. At a normal cruise altitude of
10,000 m, the cabin pressure3 (p c ) is usually equal to the atmospheric pressure at a geo-potential altitude
of 1800 to 2400 m. The pressure at the highest geo-potential altitude was chosen because this results in the
lowest pressure difference between the inside and outside of the fuselage. However, this also results in the
largest difference in cabin pressure during climb, resulting in a lower allowable rate of climb than when a
geo-potential altitude of 1800 m was chosen for the cabin pressure. The minimum time to lower the cabin
pressure from the air pressure at sea-level to the air pressure at cruise altitude can be calculated using:
p 0 (1 − p c /p 0 )
t= (4.12)
d p/d t
When the air pressure in the cabin is equivalent to the outside pressure at an altitude of 2400 m, the pressure
ratio p c /p 0 = 0.75. This results in a minimum time to climb of 857 s. Consequently, the rate of climb should
never exceed 10,000 m / 857 s = 11.7 m/s.
The limitations due to the rate of change of cabin pressure and the airspeed restriction below 10,000 ft were
both taken into account when calculating the minimum time to climb. Below 10,000 ft the maximum rate of
climb is limited by the airspeed restriction and the rate of change of cabin pressure, and just above 10,000 ft
the maximum rate of climb is limited by the rate of change of cabin pressure only.
Maximum ROC By calculating the rate of climb at different altitudes and at different airspeeds, a contour
plot with lines of a constant rate of climb can be made. This is shown in Figure 4.7. The maximum rate of climb
is the highest at sea-level and decreases with increasing altitude. The altitude at which the maximum rate of
climb is zero is defined as the absolute ceiling. This is the highest altitude the aircraft can theoretically reach.
A more useful quantity is the service ceiling, conventionally defined as the altitude at which the maximum
rate of climb is 0.5 m/s. The service ceiling is a practical upper limit for steady, level flight [5]. It was found
that the calculated theoretical ceiling is 15,890 m.
However, these values are determined to be the upper limit as the decrease in engine thrust due to in-
creased airspeed was not taken into account. Therefore, the operational service ceiling is for now set to be
12,500 m and will be elaborated further in Section 4.6.
In Figure 4.8 the rate of climb is given as a function of airspeed at different altitudes. The maximum rate
of climb at each altitude is represented by the dashed line. Note that the ROC has not been calculated for
airspeeds below stall speed. Besides the rate of climb, the maximum airspeed that can be achieved at each
altitude can also be read from the graph. The maximum airspeed at each altitude can be found by looking at
the intersection point with the x-axis that corresponds with the highest true airspeed.
16 sea level
14 20
0
5000 m
2
12
Altitude [km]
ROC [m/s]
15
6
10 10,000 m
8
10
8
12 10
6 14
4 16
5
18
2 20
0 0
50 100 150 200 250 300 350 50 100 150 200 250 300 350
VTAS [m/s] VTAS [m/s]
Figure 4.7: Lines of constant rate of climb. Figure 4.8: ROC as a function of altitude and true airspeed.
4.3.4. Cruise
In terms of fuel burn and time the cruise stage the most significant stage. The cruise Mach number is set to
0.78 to be competitive with other regional aircraft and is elaborated in more detail in Section 4.6. Cruise is
defined as optimal under the following condition:
r
1
C L opt = C D πAe (4.13)
3 0
With this the thrust setting can be determined with Equation 4.14a and with the aforementioned Equation 4.8
the fuel burned during the cruise can be calculated. The thrust specific fuel consumption needs to be cor-
rected for the height, as the TSFC is given at sea-level. This is done by multiplying it with Equation 4.14b
to the power 0.75. After setting these parameters, the optimal height can be calculated by rewriting the lift
equation. This will output the optimal ρ which in turn gives a height to fly at.
CD ρ
T =W (a) σ= (b) (4.14)
CL ρ0
In Figure 4.9 and Figure 4.10 the flight path and speed changes can be found respectively. The range is set in
such a way that when landing is included the maximum set distance of 2500 km is reached. The parameters
24 4. Flight Performance
at the end of the cruise stage are shown in Table 4.7. In there it can be seen that the fuel burned during cruise
is approximately 3407 kg and that it takes 2.9 hours to reach that distance, which is 2280 km.
10 200
Altitude [km]
Speed [m/s]
5 100
0 0
0 500 1,000 1,500 2,000 2,500 0 500 1,000 1,500 2,000 2,500
Distance covered [km] Distance covered [km]
Figure 4.9: Distance covered up until the end of cruise stage. Figure 4.10: Speed changes up until cruise stage.
Descent At the end of cruise the aircraft starts to descent, which is done with the thrust levers set to idle
thrust. If it is assumed that the idle thrust can be neglected, the minimum angle of descent is constant and
can be calculated using
à !
CD q
γ̄ = arctan where C L opt = C D 0 πAe (4.15)
C L opt
where the descent angle γ̄ is independent of the aircraft weight.
Limitations on the ROD Just as with climb, the rate of descent is limited by the rate of change of cabin
pressure. For descent, the rate of change of cabin pressure is not allowed to exceed 18 Pa/s. If the same cabin
pressure and cruise altitude are assumed, i.e., p c /p 0 = 0.75, the minimum time to descent is 1428 seconds.
Consequently, the rate of descent should never exceed 10,000 m / 1428 s = 7.0 m/s.
It was found that for descent, the airspeed restriction of 250 kts below 10,000 ft does not impose any restric-
tions on the minimum angle of descent when the thrust levers are set to idle.
Landing After descent, the landing phase of the flight will be the final phase. The landing is the reverse of the
take-off and also consist of two parts; the approach and groundroll phase. The approach speed is determined
by 1.3Vstall where Vst al l is defined as in Equation 4.3. The approach is done at a descent angle γ̄ of 3°. After
the flare height is reached, a circular path is followed, which is calculated as:
1.3Vstall
R= (4.16)
∆n · g 0
4.4. Flight Performance Optimisation 25
In this equation, ∆n is set to be 0.10 by statistical data based on jets [91]. After the aircraft reaches the runway
a transition phase is incorporated in the path in which the aircraft is rotated to land; this takes approximately
2 seconds. After touchdown, the groundroll is done, in which the deceleration is given by:
g0
a br = (D + µbr (W − L)) (4.17)
W
The distance covered is calculated by the time it takes to decelerate from landing speed to standstill. The
result of the landing phase can be seen in Figure 4.11 and Figure 4.12. It takes approximately 1450 m to come
to completely stop, which is lower than the set requirement of 1500 m.
80
8
60
6
Speed [m/s]
Altitude [m]
40
4
2 20
0 0
0 500 1,000 1,500 0 500 1,000 1,500
Runway length [m] Runway length [m]
Figure 4.11: Runway covered from flare height to standstill. Figure 4.12: Speed decrease during landing.
After landing
W [N] 248990
Fuel burned [kg] 4808
s [km] 2498
h max [km] 11.0
t [hr] 3.34
Cruise climb Cruise climb was already implemented in the main flight path shown in Figure 4.13. This
ended up in 21 kg less fuel, which is about a 0.4% decrease in the total fuel. This is a small percentage but it
makes sense considering the short range and the fact that a bigger descent would be needed. A small decrease
in fuel can lead to enormous savings for the airline as multiple flights per day are executed.
Full runway use Making use of the full runway can be beneficial in various aspects. Due to the fact that the
engine can deliver more thrust than required to take off within 1500 m, a reduced throttle setting leads to less
26 4. Flight Performance
10
200
Altitude [km]
Speed [m/s]
5 100
0 0
0 500 1,000 1,500 2,000 2,500 0 500 1,000 1,500 2,000 2,500
Distance travelled [km] Distance travelled [km]
Figure 4.13: Runway covered to clearance of screen height. Figure 4.14: Runway covered to distance of screen height.
fuel consumed. Additionally, a lower thrust setting is better for the engine because this results in less engine
wear, which results in less maintenance. Besides this, a lower thrust setting also results in a decrease in noise
both from the airframe as from the engines.
Take-off segments A more continuous climb during the take-off segments results in less noise at the ground
and a faster climb with reduced fuel consumption than the step-wise climb. This leads to 68 kg of less fuel
burned.
Steep approach Finally, it was decided to adapt the aircraft design to a steeper approach. In fact, by improv-
ing the high lift devices to create a C L max at landing of 1.97, the aircraft is able to fly slower while approaching.
This results in fewer stresses on the landing gear while landing, which makes it possible to increase the ap-
proach angle. By increasing it 3 to 4°, the procedure ensures to decrease noise at landing [7]. Furthermore,
1 kg less fuel is consumed.
By considering all these optimisation actions, a much lower noise is produced at take-off and landing proce-
dures and 90 kg of fuel are saved each flight. This value represents a decrease of 1.5% decrease in the total fuel
consumption. This is a small saving for one single flight but when considering 1 year of operations it could
represent a significant increase in revenues for the aircraft operator.
4.5. Payload-Range
The payload-range diagram of the final aircraft design is shown in Figure 4.15. On the vertical axis the payload
is given and on the horizontal axis the still air range, which is the range with zero wind velocity.
With maximum payload, or 76 passengers, the aircraft taking off at its MTOW will be able to fly a distance
equal to 2500 km. After this point, the range can be increased further by reducing the number passengers and
carrying more fuel. When the maximum fuel capacity is reached, 62 passengers can be transported over a
distance of 2945 km. Finally, when the payload weight is reduced to zero and no passengers are present, the
aircraft can reach 4175 km.
8,000
M
MZFW = 21,615 kg TO
6,992 W
=
30
,20
0 kg
5,704
Payload [kg]
M
4,000
ax
.
Fu
e l=
85
85
2,000
kg
0
0 1,000 2,000 2,500 2,945 4,000
Still air range [km]
altitude is rare with large passenger aircraft as safety critical problems related to propulsion and control will
arise from the thinner air density.
Following that, the limit operating speeds are calculated. There are two limits that need to be observed in
the operation of the aircraft, which are the maximum operating equivalent airspeed and the maximum op-
erating Mach number. The maximum operating equivalent airspeed should not be exceeded as it may cause
structural damage due to aeroelastic effects, whereas the maximum operating Mach number should not be
exceeded as shock-waves will form on the wing as the airflow becomes locally supersonic. To be competitive
with the CRJ700 and E170, the maximum operating airspeed and Mach number is set to be at 335 m/s KEAS
(172 m/s EAS) and 0.78 respectively.
The flight envelope is presented in Figure 4.16 and is calculated for the MTOW of the aircraft. The stall
limits of the aircraft will move to the left if the aircraft is loaded lighter than the MTOW. It can be observed that
the true airspeed limit shifts from the equivalent airspeed limit to the Mach limit at approximately 8000 m.
Also, the true airspeed limit from 11,000 m upwards remains constant as the Mach number does not change
due to the constant temperature in the tropopause.
12,000 Vstallclean
Vstalllanding flaps
10,000 Max. VEASoperating
Max. Moperating
Altitude, H [m]
Hceiling
8,000
6,000
4,000
2,000
0
50 100 150 200 250
True airspeed, VTAS [m/s]
4.7. V -n Diagram
The V -n diagram is constructed in order to determine the maximum load factors that the aircraft needs to
be designed for, as well as to clearly define the operational limits of the aircraft for pilots. The constructed
V -n diagram is contained in Figure 4.17. To construct the V -n diagram, certification specifications from the
28 4. Flight Performance
European Aviation Safety Agency (EASA) CS-25 requirements are referenced [3].
From CS 25.337 [3], the maximum positive limit load factor of an aircraft of the weight class of the designed
regional aircraft is 2.5 while the negative limit load factor is −1. The left boundary of the manoeuvre limit is
determined by the load factor experienced by the aircraft in the maximum and minimum C L values. The
manoeuvre limit for flaps is also included to meet requirement CS 25.345 [3] that states that the flaps need
to meet a maximum positive limit load factor of 2. The maximum flap deployment speed, VF is constrained
by its structural strength and meets the requirement of being at least 1.8 times the stall speed at landing
configuration. The designed dive speed is the maximum speed the V -n diagram takes into consideration and
is driven by CS 25.335(b) [3], which states that it should be at least 1.25 times larger than the designed cruise
speed except when the speed is Mach limited.
In addition to manoeuvre loading, the gust loading also needs to be taken into consideration at the cruise
and dive speeds as stated by CS 25.335(d) [3]. The change in load factor due to a gust is calculated with
ρV C L α K U
∆n = 1 + (4.18)
W /S
where U is the gust velocities defined in CS 25.345(a)(5)(i) [3] and K is defined as
0.88µ
K= .
5.3 + µ
Manoeuvre
Manoeuvre (full flaps)
2 Gust (cruise)
Gust (dive)
Load factor, n [-]
−1
0 50 100 150 200
Equivalent airspeed, VEAS [m/s]
80 3
2
D n=1
40 D n=1.5
D n=2
D n=2.5 1
20
0 0
100 150 200 100 150 200
True airspeed, VTAS [m/s] True airspeed, VTAS [m/s]
Figure 4.18: Drag at different load factors Figure 4.19: Maximum achievable load factor at MTOW at 7000 ft.
and the available thrust.
V2
R= p (4.19)
g n2 − 1
From the information in Figure 4.19, the turn radius at every airspeed at its corresponding load factor is
calculated and is presented in Figure 4.20.
It should be realised however that the condition for minimum turn radius does not necessarily equal give
the condition for the minimum time to turn. Therefore, the minimum time to turn is found by the formula
2πR
T2π = (4.20)
V
The results of the calculations are summarised in Table 4.9. The results demonstrate that at maximum take-off
weight, the MW 5 is capable of meeting standard rate two turns (6 deg /s) at the upper-limit of typical holding
altitudes of 7000 ft.
Minimum time to turn 360 deg [s]
·104
Clean Clean
400
Flaps Flaps
Turn radius, R [m]
1
300
0.5 200
100
0
100 150 200 100 150 200
True airspeed, VTAS [m/s] True airspeed, VTAS [m/s]
Figure 4.20: Turn radius at MTOW at an altitude of 7000 ft. Figure 4.21: Turn radius at MTOW at an altitude of 7000 ft.
4.9.1. Verification
Verification can be divided into two parts: code verification and calculation verification. The purpose of code
verification is to make sure that the program does not contain syntax errors or other errors that prevent the
program from working correctly. Once the code is free of programming errors, the program should be checked
to make sure that the numerical model is correct. This was done on a small scale (unit tests) and on a larger
scale (system tests).
Different unit tests were performed to verify the correctness of the individual units of code:
• International Standard Atmosphere (ISA): For the calculation of the air temperature, air pressure and
air density, the International Standard Atmosphere was used. The calculations of the properties of the
ISA were checked by comparing these values to other Standard International Atmosphere tables [6].
• Take-off and landing: For the calculation of the required runway length for take-off and landing the
results were checked by using data from different textbook examples. The results of the textbook exam-
ples were then compared to the numerical results calculated by the program. For take-off, example 6.25
from Introduction to Flight by John D. Anderson [6] was used, and for landing example 6.26.
• Flight: For climb, cruise and descent the calculations were also verified by using data from textbook
examples and comparing the results of these examples with the results as calculated by the different
programs. For climb example 6.18 from Introduction to Flight by John D. Anderson [6] was used. For
cruise example 6.20 was used and for descent the example in section 13.1 from Elements of airplane
performance by Ger J.J. Ruijgrok was used.
• Balanced field length: For the verification of the balanced field length examples 6.25 and 6.26 from
Introduction to Flight by John D. Anderson [6] were used to verify the numerical results of the program.
After the individual units of code were tested and determined to be functioning properly, the program that
models the complete flight from take-off to landing was tested. This was done by o.a. changing some input
parameters and checking the influence of these changes on the output of the program. For example, when
the thrust-specific fuel consumption was set to zero, the aircraft weight remained constant, and when it was
increased, the fuel consumption increased.
4.9.2. Validation
After verification, validation was performed in order to determine that the main program accurately models
the complete flight. For the validation of the program that models the complete flight from take-off to land-
ing, the fuel consumption, maximum rate of climb, time to climb, cruise speed, time to descent, and other
parameters were compared to different regional aircraft. For the comparison, mainly the Bombardier CRJ700
was used. The reason for this is that the CRJ700 is very similar to the MW 5 and carries almost the same num-
ber of passengers. Besides the comparisons with other regional aircraft, the dimensions and characteristics
of these aircraft were used as inputs for the program. The outputs of the program, e.g. fuel consumption, rate
of climb, cruise speed, etc., were then compared to the actual values of the comparable aircraft to validate the
correctness of the calculations.
The effects of varying the values of these parameters can be found in Tables 4.10 through 4.13. In these tables,
the sensitivity to different output parameters can be found. These output parameters are the take-off length,
the maximum rate of climb at sea level, the time to climb to a cruise altitude of 10 km, the fuel burned during
the climb, the cruise time and the fuel burned during the cruise.
As can be seen in Table 4.10, the zero-lift drag coefficient has the biggest influence on the rate of climb and the
fuel consumption. The higher the zero-lift drag coefficient, the higher the fuel consumption and vice versa.
The influence on the time to climb is negligible, because the rate of climb is limited by the maximum rate of
change of cabin pressure that is still experienced comfortable by the passengers, and not by the maximum
rate of climb that can be achieved. This is the result of the slightly over-designed engines.
Table 4.10: Sensitivity of performance parameters to the zero-lift drag coefficient (C D 0 = 0.056).
%Difference %s take-off %ROCmax %t climb %(m fuel )climb %t cruise %(m fuel )cruise
−20 −1.11 14.47 0.00 −7.86 −1.40 −12.23
−10 −1.56 6.67 0.00 −3.67 −0.69 −5.93
+10 0.57 −5.79 0.00 3.31 0.56 5.52
+20 1.14 −10.88 0.00 6.57 0.92 10.65
The effective aspect ratio has a relatively small influence on the take-off distance, rate of climb and time to
climb. The influence on the fuel consumption is a bit larger. A higher effective aspect ratio results in a lower
fuel consumption, hence lower operational costs per block hour. The effective aspect ratio was increased by
adding winglets to the wing tips.
Table 4.11: Sensitivity of performance parameters to the aspect ratio (A eff = 10.5).
%Difference %s take-off %ROCmax %t climb %(m fuel )climb %t cruise %(m fuel )cruise
−20 1.79 −2.96 0.00 5.32 −1.08 8.31
−10 0.79 −1.32 0.00 2.37 −0.50 3.84
+10 −0.63 1.08 0.00 −2.04 0.38 −3.37
+20 −1.16 1.99 0.00 −3.82 0.58 −6.42
As can be seen in Table 4.12, changing the area of the wing has a relatively large impact on the take-off dis-
tance and the maximum rate of climb at sea level. Reducing the wing area increases the take-off length, but
reduces the fuel consumption and time spent during climb and cruise.
%Difference %s take-off %ROCmax %t climb %(m fuel )climb %t cruise %(m fuel )cruise
−20 18.24 11.80 −1.06 −4.72 −2.65 −4.94
−10 9.06 5.41 −0.34 −2.04 −1.23 −2.32
+10 −8.81 −4.61 −1.05 1.45 0.19 1.29
+20 −17.68 −8.71 −0.81 4.12 0.52 2.85
In Table 4.13 the sensitivity of different parameters to the maximum take-off weight can be found. The maxi-
mum take-off weight has the biggest influence on the take-off distance. It can be seen that a weight reduction
of 20% results in almost 50% less take-off distance required. This seems like a big difference, but it can be
explained. If the aircraft is lighter, it accelerates quicker to the lift-off velocity. Because the lift-off velocity is
proportional to the square root of the wing loading, the lift-off velocity is lower for a lower maximum take-off
weight. Furthermore, the ground run distance reduces even more due to the lower friction force between the
landing gear and the runway.
32 4. Flight Performance
Table 4.13: Sensitivity of performance parameters to the maximum take-off weight (MTOW = 30,200 kg).
%Difference %s take-off %ROCmax %t climb %(m fuel )climb %t cruise %(m fuel )cruise
−20 −48.45 30.38 −0.62 −18.82 −0.07 −17.71
−10 −21.00 13.63 −1.02 −9.85 −0.06 −8.86
+10 16.37 −11.36 −1.72 6.80 −1.57 7.23
+20 29.33 −21.00 −2.12 14.44 −2.45 14.81
5
Aerodynamics
In this chapter, the aerodynamic design of the aircraft will be discussed. First, the wing optimisation will be
discussed in Section 5.1, including both the 2D planform as well as the airfoils. Then, both the horizontal
tail as well as the vertical tail will be sized in Section 5.2. The results of the planform optimisation are pro-
vided in Subsection 5.1.5. Finally, the verification and validation of the models used for the wing planform
optimisation is done in Section 5.4.
Figure 5.1: Example of horseshoe vortices on a lifting surface. Figure 5.2: AVL vortex pattern.
Nonetheless, there are limits to the validity of VLM. AVL is only valid until Mach effects occur, i.e. when a
Mach number of 1 is reached locally. At this point the equations of VLM are not valid anymore. Unfortunately,
with a cruise Mach number of 0.78, the chances a of local sonic boom is relatively high. On the other hand,
methods that can take this into account are complex, hard to use and very time intensive, and thus out of
the scope of the design process at this stage. However, there are airfoils that are designed to avoid supersonic
effects, even at high Mach numbers: supercritical airfoils. With these airfoils, it is reasonable to assume that
no supersonic effects occur during the usual Mach regimes, especially since there are many other aircraft that
cruise at higher Mach numbers (0.82-0.85) with these airfoils.
1 H. Youngren, M. Drela, AVL, MIT, for more information see Program and Code Bibliography on page 115
1 S.J. Hulshoff, Non-Linear Lifting-Line Solver, Faculty of Aerospace Engineering, Delft University of Technology, (2007), retrieved on
06/06/2018 from https://aerodynamics.lr.tudelft.nl/~shulshoff/nlll/doc/manual.html
33
34 5. Aerodynamics
Thus, for this analysis, VLM is used, since they yield accurate results using relatively simple numerical
techniques. Furthermore, it is assumed that the local Mach number does not reach 1 at any point when using
supercritical airfoils.
Λ1
cr
cr λ1
Λ 2
cr λ 1 λ 2
x/b
b/2
Planform parameters
ρ [kg/m] Air density Fixed by fixed altitude
a [m/s] Speed of sound Fixed by fixed altitude
M [-] Mach number Set by performance
V [m/s] Cruise speed Set by M and a
S [m2 ] Surface area Set by altitude c l and M
A [-] Aspect ratio Iterated
b [m] Span Set by A and S
λ1 [-] Taper ratio root to kink Iterated
λ2 [-] Taper ratio kink to tip Iterated
Λ1 [deg] Sweep angle root to kink Iterated
Λ2 [deg] Sweep angle kink to tip Iterated
x k /b [-] Kink location Fixed based on reference aircraft
Now that the input parameters are defined, the next step is to define the output to be minimised. The best
indicator of the wing performance is the lift over drag ratio (L/D). Indeed, a maximum value of (L/D) is the
best indicator for minimum fuel consumption for a given weight, which is very important for operational
cost. However, since the lift is fixed since the aircraft weight is known, this is equivalent to minimising drag.
Using Python and AVL, a function that takes the parameters to be optimised as inputs, and the drag from AVL
as output is built. This is then minimised and the results are given in Table 5.2.
p 2·S 1 + λ + λ2
b= S·A cr = c mac = c r · (2/3) · (5.1)
(1 + λ) · b 1+λ
5.1. Wing Optimisation 35
The wing optimisation is based on SciPy’s basin hopping algorithm2 . This algorithm is a powerful all-round
optimiser that is adapted to non-convex optimisation, which the drag function is not, and even though it can
be powerful when lot’s is known about optimisation and the function behaviour, it can also figure out some
of the settings itself so is easy to use for beginners in the field of optimisation. Therefore it is well suited for
this project. The results of the optimisation are given in Table 5.2.
Table 5.2: Starting values based on statistics compared to end values after optimisation.
However, since the tip airfoil stalls at a lower angle of attack than the root airfoil as shown in Table 5.4, a twist
needs to be added to the wing. Indeed, the desired behaviour is for the root to stall first, so that control over
the aileron can be kept as long as possible. At sea level and cruise conditions, the difference is of 2.2° and 2.5°
respectively. To add a bit of a margin, the tip twist limit has been set at −3.5° in order to avoid tip stall. A more
detailed analysis is required in a later stage of the program, since the wing twist changes during the mission
due to aerodynamic loading.
Table 5.4: Angle of attack at maximum c l for main and tip airfoil in different conditions.
2 Basin hopping algorithm, SciPy, documentation of this algorithm can be found https://docs.scipy.org/doc/scipy/reference/
generated/scipy.optimize.basinhopping.html
36 5. Aerodynamics
Incidence angle With the new planform parameters, the optimum wing lift coefficient for cruise is found to
be 0.4149. From calculations in AVL it follows that the angle of attack of the wing corresponding to this wing
lift coefficient is 0.061°. The wing is installed with an incidence angle of the same amount to ensure that the
floor of the fuselage is level during cruise. The fact that the floor of the fuselage is level during cruise makes it
easier for flight attendants to push food carts through the aisle.
Dihedral angle In order to increase the lateral stability of the aircraft, the wings are installed with a dihedral
angle of 3°. The dihedral angle of 3° is an initial estimate based on reference aircraft and has to be checked in
the eigenmotion analysis.
Wingtip devices The induced drag component is reduced by installing winglets at the wing tips of the air-
craft. Winglets reduce the effect of wingtip vortices by blocking the cross-flow from the lower to the upper
side of the wing. Furthermore, lift spoiling effects at the most outboard section of the wing is also limited by
blocking the cross-flow. Finally, if installed properly the cross-flow component generates a thrust component
at the tip.
The main benefit of using winglets is that the effect is similar to having higher aspect ratio wings, without
the drawback of a higher bending moment and hence a heavier wing. A properly designed winglet can yield
an effective aspect ratio of about 20% higher than the geometrical aspect ratio of the wing [32]. The effective
aspect ratio is the value to be used for the next iteration of the performance calculations.
0.6
1
c l · c/c r e f [-]
0.4 0.5
c l [-]
0
0.2
−0.5
0 5 10 −10 −5 0 5 10
Distance from root [m] α [deg]
Figure 5.4: Spanwise lift distribution. Figure 5.5: c l α curve of the clean wing at M = 0.2.
1
−0.07
0.5
−0.08
c m [-]
c l [-]
0
−0.08
−0.5
−0.09
Figure 5.6: Drag polar of the wing. Figure 5.7: c mα curve of the clean wing at M = 0.2.
Taper and aspect ratio In general the aspect ratio and taper ratio of the horizontal tail are selected in such a
way that the structural weight of these components is reduced, keeping the aerodynamic performance of the
horizontal tail in mind. For the horizontal tail reduction of structural weight is preferred over the reduction of
induced drag. From historical data it follows that the aspect ratio of the horizontal tail is between 3 and 5 and
that the taper ratio of the horizontal tail is between 0.3 and 0.6 [90]. For the initial sizing of the horizontal tail
the average of the aspect ratio and taper ratio ranges are selected, which are 4 and 0.45 respectively.
Quarter chord sweep The last parameter required to determine the planform of the horizontal tail is the
sweep of the quarter chord line. Because the horizontal tail is located well above the main wing it is assumed
that the velocity ratio between the main wing and the horizontal tail is approximately 1.0. This means that
the horizontal tail experiences the same velocity, and hence Mach number, as the main wing. To postpone
the drag divergence due to Mach effects, the quarter chord line sweep of the horizontal tail is set the same as
the quarter chord line sweep of the main wing.
Airfoil selection Now that the planform of the horizontal tail surface is fully defined, the last element re-
quired to fully define the geometry of the horizontal tail surface is the airfoil used at the tip and root chord.
Because the horizontal tail has to be able to generate lift in both the upward and downward direction, a
symmetrical airfoil is employed on the horizontal tail surface. The drag divergence Mach number for the hor-
izontal tail should be higher than the drag divergence Mach number of the main wing. In order to achieve
this, the (t/c) of the horizontal tail surface is about 90% of the (t/c) of the main wing, both at the tip and at the
root.
The parameters of the horizontal tail surface are provided in Table 5.5.
Table 5.5: Horizontal tail geometry. Table 5.6: Vertical tail geometry.
S v C v · b 0.08 · 25.5
= = = 0.194 (5.2)
S lv 10.5
Taper and aspect ratio The aspect ratio and taper ratio of the vertical tail are sized in a similar fashion as
the aspect ratio and taper ratio of the horizontal tail. The ranges for the aspect ratio of the vertical tail for a
T-tail configuration are between 0.9 and 1.2, so an initial guess of 0.95 is used for the initial vertical tail design.
The taper ratio for T-tail configuration aircraft ranges between 0.6 and 1.0, which results in an average value
of 0.8 for the initial estimate of the taper ratio of the vertical tail[90].
Quarter chord sweep Similar to the horizontal tail design, the quarter chord sweep angle of the vertical tail
is set to the same value as the quarter chord sweep of the wing to postpone drag divergence due to Mach
number. The vertical tail is not located in the slipstream of any other element, so the velocity, and hence also
the Mach number, over the vertical tail is equal to that of the free stream.
Airfoil selection The vertical tail has to produce a side-force in the direction opposite to the angle of attack,
or side-slip in this case, it experiences. This means that the zero lift angle of attack of the airfoil selected
for the vertical tail should be zero. The zero lift angle of attack requirement and the direction of side-force
production requires a symmetrical airfoil for the vertical tail. As a guideline for the initial airfoil selection for
the vertical tail, the (t/c) of the airfoil of the vertical tail is approximately 98% of the (t/c) of the main wing.
The main parameters of the resulting vertical tail geometry are presented in Table 5.6.
Figure 5.8: Top view of the final wing planform including the airfoils.
All the airfoils used are given in Figure 5.9 for the main wing, and in Figure 5.10 for the horizontal and vertical
stabilisers.
5.3. Resulting Wing 39
(a) Whitcomb, root and kink airfoil. (b) NASA SC21010, tip airfoil.
(a) NACA 0011, used for the vertical (b) NACA 0010, root airfoil (c) NACA 0009, tip airfoil horizontal
stabiliser. horizontal stabilser. stabilser.
The horizontal tail mainly contributes to the longitudinal stability of the aircraft by changing the moment co-
efficient. The contribution of the horizontal tail to the moment coefficient curve is presented in Figure 5.11.
The main destabilising factor on the longitudinal stability is the contribution of the fuselage, which is esti-
mated by using the data available for the fuselage contribution of the Cessna Citation II [61].
The horizontal tail does not only influence the moment coefficient of the aircraft, but also the lift coeffi-
cient by producing additional lift if it experiences a positive angle of attack. The contribution of the horizontal
tail to the total lift coefficient generated by the aircraft is given in Figure 5.12. The contribution to the lift co-
efficient of the horizontal tail is notably less significant than the contribution to the moment coefficient due
to the large distance between the horizontal tail and the main wing.
Wing Wing
0.4
Tail 1 Tail
Fuselage Total
0.2 Total
0.5
C m [-]
C L [-]
−0.2 0
−0.4
−0.5
−10 0 10 −10 −5 0 5 10
α [deg] α [deg]
Figure 5.11: Contribution of various elements to the moment Figure 5.12: Contribution of the horizontal tail to the lift curve.
coefficient.
The tail does not only contribute to lift and moment generation, but also to the generation of induced and
parasitic drag. The contribution of the tail surfaces, both horizontal and vertical, can be seen in Figure 5.13.
The vertical tail does not contribute to the lift production, as it is oriented perpendicular to the wing and
horizontal tail, and hence only shifts the lift curve to the right. The contribution of the vertical tail to the drag
coefficient in cruise condition is found to be 0.001383. The horizontal tail does produce a changing lift with a
changing angle of attack, and hence a changing amount of induced drag. The contribution of the fuselage to
the total drag consists of friction drag and pressure drag. The drag contribution of the fuselage is given by the
estimation of the viscous drag provided by Raymer [75], as is presented in Equation 5.3. The drag contribution
of the fuselage depends on the form factor of the fuselage, which is calculated using Equation 5.4, the fuselage
area and the local Reynolds number over the fuselage. The form factor of the fuselage depends on the fuselage
length and diameter.
S w f us 0.455
C D fus = F F fus ¢2.58 ¡ ¢0.65 (5.3)
S
¡
log10 (Re) 1 + 0.114M 2
40 5. Aerodynamics
lf
60 df
F F fus = 1 + ³ ´3 + (5.4)
l f 400
df
Wing
1 Total
0.5
C L [-]
−0.5
Figure 5.13: Contribution of the tail and fuselage to the lift drag polar.
From Figure 5.13 it follows that the zero-lift drag coefficient of the entire design is 0.02128, which is 1.4%
higher than the initial estimate of 0.021. The Oswald efficiency factor of the wing planform is estimated using
the method presented by Sholz [84]. The method for estimating the Oswald efficiency factor is presented in
Equation 5.5. The resulting Oswald efficiency factor is 0.784, which is 2.05% lower than the initial estimate for
the Oswald efficiency factor of 0.8.
1
e= µ ¢0.33 ¶ (5.5)
f (λ)A 10 ct
¡
¡ ¢ 0.142+ 0.1
1 + 0.12M 6 1 + 2
cos (Λ )
+ (4+A) 0.8
0.25c
5.4.1. Verification
The goal of verification is to check whether the code does actually compute what it is built for, and there are
no mistakes in the code.
• AVL: It was checked that the input to AVL was correct by inspecting the geometry of the wing plotted
by AVL based on those inputs, as well as comparing the c l as predicted by AVL to the c l as predicted
by JavaFoil at the same angle of attack and Mach number, when taking into account some of the 3D
effects.
• Input to XFLR: The correctness of the input to XFLR is verified by comparing the result of wing and tail
planform calculations to the input in XFLR. Furthermore, a visual inspection is performed in XFLR to
ensure the correctness of the inputs provided.
• Code to compute C l α : A code was created to compute C l α automatically from the data points of the C l α
curve from JavaFoil. This was compared to a graphical estimate of the slope of the plot.
• Code to compute planform parameters: This code generates a full planform based on the iterated
inputs. It was tested by filling in the inputs of a reference aircraft and plotting the resulting planform,
checking things such as the wing area from the drawn planform against the output from the code, and
inspecting if it did not produce odd things.
5.4. Verification & Validation 41
5.4.2. Validation
The XFLR and AVL programs are validated by comparing the results to experimental data gathered in a wind
tunnel. The results from the wind tunnel experiment, XFLR and AVL computations are presented in Fig-
ure 5.14 and Figure 5.15 respectively. Both XFLR and AVL predict a linear increase in lift with increasing angle
of attack, which is valid for low angles of attack but becomes invalid as the angle of attack is increased. The
difference between the AVL computations and experimental data is below 10% for an angle of attack lower
than 13°. The angle of attack below which XFLR yields a reliable result, with less than 10% error, is also below
13°. The usage of XFLR and AVL is validated for angles of attack lower than 13°.
Experimental Experimental
XFLR AVL
1 1
C L [-]
C L [-]
0.5 0.5
0 0
0 5 10 15 20 0 5 10 15 20
α [deg] α [deg]
Figure 5.14: XFLR results versus experimental data. Figure 5.15: AVL results versus experimental data.
Wing optimisation For the wing optimisation, sensitivity is not relevant. The most important inputs are
the boundaries on the iterated parameters, and since the results are in the middle of the rages and not near a
limit, different limits should not change the outcomes.
Most of the other inputs were given as outputs by performance, no sensitivity analysis is performed on
them.
6
Flight Dynamics
The purpose of this chapter is to analyse and determine the behaviour of the aircraft during flight. First, the
tail and control surfaces are sized, followed by determining the aircraft characteristics. Lastly, the eigenmo-
tions of the aircraft are determined and checked for stability.
Fuselage Xlemac [m] MAC [m] Distance Nose Distance Nose Seating Seat
Length [m] to 1st Row [m] to Cargo Bay [m] Configuration Pitch [cm]
26.93 13.15 2.74 6.24 24 2+2 78.74
28.5 15.55 2.74 5.48 23.5 2+2 78.74
Most Front c.g. [m] %MAC Most Aft c.g. [m] %MAC c.g. Range [%MAC]
13.85m 0.53 14.60m 0.89 0.35
16.16m 0.20 17.54m 0.74 0.54
42
6.2. Detailed Sizing 43
By using the weights of the subsystems of both groups and their corresponding c.g. location from Table 6.3,
the location of the main wing w.r.t. the fuselage can be iterated and the optimum determined. The optimum
wing location is taken to be where the loading diagram provides the required range of c.g. locations, as re-
quired to minimise the total tail area.
30,000
Horizontal Tail Size [Sh/S]
Aircraft Mass [kg]
0.2
25,000
20,000 0
Figure 6.1: Final loading diagram & scissor plot of the aircraft.
44 6. Flight Dynamics
Ailerons are typically located from 50% of the span up to 90% of the span. Because only 33% of the wing span
is needed for the ailerons, the ailerons are located between 57% and 90% of the wing span. This leaves some
space for the placement of mechanisms that drive both the flaps and the ailerons. To reduce the adverse
yawing effect caused by a deflection of the aileron on one side of the aircraft, it is decided to implement
differential ailerons.
During cruise the outboard ailerons can not be used because of aeroelastic effects on the wing. In order to
provide lateral control during cruise part of the flaps can be deflected in opposite direction to function as
ailerons: flaperons. The flaperons are located between 16% of the wing span and 28% of the wing span.
C L h [-]
0
−1
−20 −10 0 10 20
δe [deg]
Figure 6.3: Elevator lift coefficient as function of elevator deflection and angle of attack.
lh Sh
C mh = −C L h (6.1)
S
6.3.3. Rudder
In order to increase stability when the aircraft experiences a sideslip angle with respect to the free stream
velocity a vertical tail is used to generate a yawing moment. The rudder can be used to limit the sideslip angle
by increasing the camber of the vertical tail in both directions. The rudder is positioned at the aft 25% of
the vertical tail and spans from 10% to 90% of the span of the vertical tail. The side force produced by the
vertical as a function of the rudder deflection angle and side force angle can be seen in Figure 6.4. The data
represented in Figure 6.3 is obtained using XFLR5.
1
C Yv [-]
−1
−20 −10 0 10 20
δr [deg]
Figure 6.4: Rudder side force coefficient as function of rudder deflection and angle of sideslip.
In a similar fashion as the pitch moment contribution of the elevator, the yawing moment due to the side
force generated by the vertical tail is determined using Equation 6.2. From Figure 6.4 it can be seen that the
amount of side force generated by the vertical tail increases with increasing rudder deflection angle. The
negative correlation between the angle of sideslip and side force is explained by the fact that the angle of
sideslip is defined positive in the negative y-direction. Using Equation 6.1 it is found that for zero angle of
sideslip the yawing moment change due to a change in rudder deflection (C nδr ) is -0.0702.
46 6. Flight Dynamics
lv Sv
C n v = −C Yv (6.2)
S
Because the different trailing edge high lift devices have a different level of effectiveness, the resulting maxi-
mum wing lift coefficient depends on the selected types. The change in maximum wing lift coefficient due to
the high lift devices is given by Equation 6.3
Sw f
∆C L max = 0.9∆C l max cos(Λ0.7c ) (6.3)
S
Each high lift device type has a different value of ∆C l max , which also results in a different ∆C L max value for
each type of high lift device. The resulting flap area values for each type of high lift device are presented in
Table 6.4.
Each high lift device requires a different operating mechanism with a different amount of parts, and hence
a different mass. The mass, amount of parts and manufacturing costs are determined based on the methods
described in Rudolph [80]. The manufacturing cost, in USD, of the high lift devices is calculated using Equa-
tion 6.4.
Along with the maximum wing lift coefficient and production costs, the mass and the amount of parts for each
type of high lift devices are presented in Table 6.4. The part count is used as an indicator for maintenance of
the system, as more parts generally implies higher maintenance costs. Due to high manufacturing cost, high
maintenance and heavy weight the triple slotted and double slotted flaps are eliminated. As an initial estimate
the lightest and cheapest type high lift device is selected, so the simple slotted flap.
The influence of the high lift devices on the maximum lift coefficient in different configurations can be
found in Figure 6.5. The black line in Figure 6.5 indicates the lift curve of the wing in clean configuration. As
expected the leading edge high lift devices, indicated by the blue line, do not significantly influence the zero
lift angle of attack, but only increase the maximum lift coefficient. The flaps on the other hand, indicated with
the red line for landing configuration, have a large influence on the shift in zero lift angle of attack and the
maximum wing lift coefficient. The green line indicates the contribution of the flaps in take-off configuration,
which is lower compared to when the flaps are fully extended.
6.5. Aircraft Flight Dynamics Characteristics 47
Flap system Wing lift coefficient Mass [kg] Production cost [M USD] Part count [-]
− + ++ +
Plain and split
1.77 464 0.58 1715
0 ++ ++ ++
Slotted
1.97 439 0.51 1540
0 0 + +
Fowler
2.30 530 0.68 1800
+ − −− −
Double Slotted Fowler
2.59 651 1.04 2430
+ −− −− −−
Triple Slotted Fowler
2.80 740 1.33 2880
2
Slats
Clean
1.5 Flaps
Take-Off
C L [-]
0.5
−5 0 5 10 15
Angle of attack [α]
6.5.1. Requirements
CS-25 regulations require the aircraft to be both statically and dynamically stable for almost all possible c.g.
locations that can be encountered during aircraft operation, the only exception being that a lightly unsta-
ble roll instability is allowed. This information was found and documented in the Baseline Report [39], with
corresponding requirements set as following:
Sys.Perf.1: The aircraft shall be dynamically stable for the entire range of c.g. excursions.
Sys.Perf.2: The aircraft shall be controllable for the entire range of c.g. excursions.
An aircraft that is dynamically stable is by definition also statically stable, and therefore static stability was
not explicitly mentioned in the Performance Requirements.
6.5.2. Assumptions
In order to facilitate the mathematical analysis of the flight dynamics characteristics, several assumptions are
made. These assumptions either neglect several phenomena or slightly alter them [61]. All assumptions are
validated by analyzing the change in value of the parameters that are involved in each assumption.
48 6. Flight Dynamics
To determine the stability characteristics, one calculates the discriminant of the left-most matrix and solves
for the eigenvalues. These eigenvalues describe the symmetric eigenmotions, the phugoid and the short pe-
riod. The phugoid is characterised as a lightly damped motion with a long period. The short period on the
other hand is highly damped and has, as the name already suggests, a short period.
Lateral stability derivatives describe the behaviour of the aircraft regarding the asymmetric motion. Rolling
and yawing motion is part of the asymmetric motion, and described here. The linearised equations of motion,
in dimensionless form are shown in Equation 6.6
β
C Yβ + (C Yβ̇ − 2µb )D b CL C Yp C Yr − 4µb
0 − 21 D b 1 0 φ
pb = 0 (6.6)
Clβ 0 C l p − 4µb K X2 D b C l r + 4µb K X Z D b 2V
C nβ +C nβ̇ D b 0 C n p + 4µbK X Z D b C nr − 4µb K Z 2 D b rb
2V
Aircraft behaviour is determined in the same way as during the symmetric analysis. Setting the discriminant
equal to zero provides the eigenvalues, which when real and negative indicate stable behaviour of the eigen-
motions. The two eigenmotions for asymmetric motion are the Dutch Roll and the Spiral. Based on the cur-
rent aircraft geometry, the stability derivatives were determined with the use of the AVL program. The result
of running this program is shown in Table 6.5 up to Table 6.7.
Direction α β
Z force C Lα 6.458202 C Lβ 0
Y force C Yα −0.000068 C Yβ −0.36011
X moment Clα 0.000001 C lβ −0.12488
Y moment C mα −8.505626 C mβ 0
Z moment C nα 0.000014 C nβ 0.162044
while for longitudinal control the elevator is deflected to generate a moment about the Y-axis. While control
surfaces primarily generate rotation about one axes, they also indirectly influence rotation about other axes,
which is often undesired [61].
Aileron Control Derivatives Deflection of the ailerons is related to a total of three control derivatives. C Yδa
displays the effect of deflecting the ailerons on the side force of the aircraft. While it is non-zero for swept-wing
aircraft, it is small enough compared to the other derivatives that it can be omitted.The primary function of
the ailerons is to induce a rolling moment and subsequently rolling motion, described by the control deriva-
tive C l δa . Its value is negative, as a positive (deflection downwards on the right wing) deflection increases the
lift on the right wing and decreases that on the left wing. The result is the aircraft starts rolling towards the left,
which is considered negative rolling. C nδa related the yawing moment introduced by deflecting the ailerons.
For positive aileron deflection, more lift generated on the right wing also means that the drag is locally in-
creased, with the adverse being true on the left wing. Therefore, while the aircraft starts rolling to the left, the
locally increased drag introduces a positive yawing moment, counteracting the rolling motion. The value of
this control derivative is kept small by making the downward aileron deflection smaller than the upward one.
Hereby the change in drag is reduced, and the induced yawing moment lowered subsequently.
Rudder Control Derivatives Like for the aileron deflection, deflecting the rudder is also related to three
control derivatives. The force generated by deflecting the rudder is described by C Yδr . If δr is positive, the
lateral force it generates is also positive. The lateral force on the vertical tailplane will generate a moment
around the X-axis of the aircraft, and for positive rudder deflection, this moment is positive and the aircraft
starts rolling to the right, this motion is decribed by C l δr . The main purpose of the rudder is to introduce a
yawing moment, with effectiveness described by C nδr . It has a negative sign, so a positive rudder deflection
creates a negative yawing moment. C nδr is related to C Yδr by the negative ratio of vertical tail arm and the
wingspan.
Elevator Control Derivatives Once again, deflecting the elevator control surface has effects described by
three control derivatives. C X δe relates the deflection of the elevator with a change of force in the X-direction.
A non-zero elevator deflection alters the drag of the horizontal tailplane. While the total tailplane drag is not
negligible, changes due to elevator deflection are deemed so small, that they can be neglected. The main
purpose of the elevator is to alter the lift produced to the desired value required for either stability or ma-
noeuvring, depicted as C Zδe . Deflecting the elevator allows for altering the pitching moment of the aircraft,
as depicted in the final control derivative C mδe .
in Table 6.5 up to Table 6.7 are used as inputs in the equations of motion. The stability derivatives that could
not be determined at this point are estimated using the available information of the Cessna Citation II [21],
this aircraft has a geometry that is relatively close to the one designed here. A graphical representation of the
symmetrical and asymmetrical eigenmotions of the aircraft can be found in Figure 6.6 and Figure 6.7 respec-
tively. The eigenvalues related to the eigenmotions for the symmetrical and asymmetrical eigenmotions can
be found in Table 6.9.
The negative real part for the symmetrical eigenmotions implies that both the short period and phugoid mo-
tion are damped. As expected, the short period motion is strongly damped and has a time to half amplitude of
less than 2.4s. The phugoid motion takes significantly longer, approximately 155s, to reduce the amplitude of
the motion to half of the initial amplitude. For the asymmetrical eigenmotions the real part of the eigenvalue
is only negative for the aperiodic roll and Dutch roll.
52 6. Flight Dynamics
0 3
−5
1
−10 0
−1
0 50 100 150 0 50 100 150
Time [s] Time [s]
(a) Velocity change during the symmetrical eigenmotion. (b) Angle of attack change during the symmetrical
eigenmotion.
5
1
Pitch rate [deg/s]
Pitch angle [deg]
0 0
−1
−5
−2
0 50 100 150 0 50 100 150
Time [s] Time [s]
(c) Pitch angle change during the symmetrical eigenmotion. (d) Pitch rate change during the symmetrical eigenmotion.
10 10
sideslip angle change [deg]
0 0
−10 −10
0 5 10 15 20 25 0 5 10 15 20 25
Time [s] Time [s]
(a) Angle of sideslip change during the asymmetrical (b) Roll angle change during the asymmetrical eigenmotion.
eigenmotion.
10 10
Yaw rate [deg/s]
Roll rate [deg/s]
0 0
−10 −10
6.8. Static Stability Analysis 53
x c.g . − x w Sh lh
C m = C m ac +C L w −C L h (6.7)
c̄ S c̄
−1
δe = (C m0 +C mα (α − α0 ))
C
m δe
Sh lh
C mδe = −C L h (6.8)
δ S c̄
x − xw Sh lh
C mα = C L w c.g .
−C L hα
α
c̄ S c̄
What can be deducted from these equations, is that during flight the elevator deflection depends on airspeed
and c.g. location. To obtain a low elevator deflection and thereby low tail drag, it is desired to have these pa-
rameters cancel each other out. From the loading diagram it can be observed that the c.g. is located behind
the a.c. This combined with a high elevator effectiveness at cruise speed allows for relatively low elevator de-
flection for trim conditions. For the extreme c.g. locations as described in the loading diagram, and the lift
gradients of the airfoils, Equation 6.8 can be evaluated.
Based on the current aerodynamic parameters of the aircraft, combined with the information from the load-
ing diagram, a range of elevator deflection was determined for stability. The elevator deflection ranges from
0.25° at MTOW to 2.9° at most aft c.g. position. Such a large elevator deflection translates to a significant trim
drag, however, this c.g. position is only encountered on the ground. During flight it is virtually impossible to
achieve this c.g. position. At MLW the required elevator deflection is slightly higher than at MTOW, equalling
0.8° due to slightly more aft c.g..
the fuselage and horizontal tail are assumed to be negligible), with expressions for each of them from Corke
[24], its magnitude can be determined.
1 tan Λ A A2 x sin Λ
C nβw = C L2 − cos Λ − − +6 (6.10)
4πA πA(A + 4cosΛ) 2 8 cos Λ c̄ A
d θ qv t
µ ¶
C n βv t = V v t C L αv t 1 + (6.11)
dβ q
Equation 6.11 takes into account the effect of the fuselage on the vertical tail. This type of downwash is dif-
ficult to estimate. An empirical relation [36]is used to approximate the relation of sidewash and vertical tail
aspect ratio, and this value is taken. For an aspect ratio of 0.95, sidewash ratio is equal to 0.43. Applying all
the Aerodynamic parameters known at this point, the value for C nβ is found to be 0.16. Concluding is that,
since the value is positive, stability is found. If made more positive in the future, stability is improved but
controllability is worsened. Because C nβ is positive, C Yβ must also be positive.
6.9.1. Verification
The goal of verification is to check the code does actually compute what it is build for, and there are no mis-
takes in the code. The code used in this chapter was the AVL program made by MIT. Verifying the results was
done by manipulating the values entered in the program in such a way, that the correct result was known be-
forehand and significantly different from the initial result. For example, the tail arm was made negative, which
produced unstable symmetric motion, as expected. Other changes that were made include removing the ver-
tical tail and applying forward sweep. All changes in aircraft geometry were drastic in order to get significant
changes in program results. All subsequent results produced by the AVL file were correct and therefore the
initial results are deemed accurate and thereby verified.
6.9.2. Validation
The program is validated by checking if the results meet the requirements. According to Sys.Perf.1 & Sys.Perf.2
the aircraft has to be dynamically stable for all possible flying scenarios. If these requirements are met, the
results are validated. Looking at the eigenvalues of the aircraft in Table 6.9, it can be seen that four out of five
eigenvalues have negative real parts. Combined with a non-zero damping ratio, it can be said that dynamic
stability is achieved, and the requirements are met. The one eigenvalue that does not have a negative real part
is the one that corresponds to the spiral motion. The aircraft is unstable with regards to this eigenmotion,
and therefore technically not dynamically stable in all possible flight scenarios. However, CS-25 regulations
tolerate an unstable spiral motion as long as the degree of instability is not too severe. In Table 6.9 it can be
seen that the time taken to double the amplitude of the spiral motion is more than 47 seconds. This indicates
that the pilot has plenty of time to react to this disturbance, and this time to double amplitude is within the
bounds set by CS-25 regulations.
6.9. Verification & Validation 55
Considering all the eigenvalues and corresponding eigenmotions are either dynamically stable or not
too unstable according to regulations, the flight dynamics characteristics satisfy the requirements and are
therefore valid.
7
Structures
To make sure the aircraft is able to carry the loads subjected on it during operation, it is vital to have an ade-
quate structure which is able to carry these loads. In this chapter, two main structural elements of the aircraft
will be investigated in more detail. First, however, the material for these two structural elements is determined
in Section 7.1. Next, the fuselage sizing and wing sizing can be found in Section 7.2 and Section 7.3, respec-
tively. Finally, the results will be verified and validated in Section 7.4 and sensitivity analysis is presented in
Section 7.5.
Several material choices can be made with respect to metals. In the aerospace industry, aluminium and tita-
nium alloys are the most frequently used metals [43, 53]. In order for the material to be as light and sustainable
as possible (while retaining good manufacturing properties) a lot of research is conducted into the different
alloying elements. Recent innovations in alloying aluminium with lithium, magnesium, beryllium and scan-
dium resulted in several promising alloys [88, 44, 72, 37, 92]. Looking at Table F.1, the beryllium alloy can be
eliminated due to the high material costs. Also, beryllium is a toxic material, making it hard to manufacture
and maintain [55, 88]. Titanium is eliminated due to a combination of high material prices and difficulties in
manufacturing [71]. As titanium is a very hard metal, tool wear can drive tooling expenses to an unaccept-
able level. Aluminium-scandium alloys are also a new development promising a large reduction in weight
while maintaining key structural properties [44]. However, since this material is relatively expensive and new
(therefore unknown in terms of maintenance), it is also dropped. It might, however, be interesting for the fu-
ture.
The properties of the remaining materials are compared to determine the results of the initial trade-off. The
aluminium lithium alloys have similar strength compared to the regular Al-2xxx and 7xxx series. However,
they have a significantly lower density (almost 10%) making them really attractive in terms of weight reduc-
tion. Also, the maintainability is similar to the regular alloys, making them suitable as well. The drawbacks
56
7.1. Material Selection 57
compared to the regular materials are the high material costs and the difficulties in manufacturing. Com-
pared to aluminium alloys without lithium, alloys containing lithium can contaminate tools. This requires
a separate production line and equipment. Also, lithium is slightly toxic, meaning that handling should be
carefully considered. However, the good strength characteristics, low density and relatively low increase in
material costs outweigh the disadvantages. Therefore, aluminium-lithium alloys are chosen as the main ma-
terial of the aircraft.
The density of each material is used to give an initial estimate of the weight reduction by that material com-
pared to a reference material, in this case Al2024-T351. Because the analysis mainly focuses on the fuselage
and the wing, the weight reduction for these two components is determined in order to calculate a new OEW
and MTOW. Using the code from the Flight Performance and Propulsion (FPP) department, this translates
into a reduction of fuel which can be used to calculate the amount of money saved after the operational life-
time1 . The return on investment is subsequently calculated using an operational lifetime of 12.5 years with 4
cycles a day. All the results can be found in Table 7.1.
Table 7.1: Material weight saving rate of return based on fuel consumption.
The different materials also have different manufacturing properties and therefore influence the production
costs. The production costs can be as high as 70 to 85% of the purchase price of the aircraft [1]. The machin-
ing speed is used as an estimate to evaluate the effect of material choice on production costs. Assigning the
average value of 31.5 million EUR [1] to the reference material and calculating the differences, the results in
Table 7.2 can be calculated. Since the production costs are based on the weight, a correction for the density
between the materials and the reference material is also applied.
Table 7.2: Change in production cost with respect to the base material.
Extra
Machining Contin-
ρ Machining
Alloy Speed gency
[kg/m3 ] Cost
[m/s] (20%) [MEUR]
[M EUR]
Al 2024-T351 2750 67.1 - -
Al-Li 2090-T83 2600 42.7 15.3 7.8
Al-Li 8090-T851 2540 54.9 4.06 5.9
Looking at this table, it can be concluded that the first material increases the production costs dramatically.
The second material changes the production costs by a relatively small amount, making the total production
costs slightly higher than the upper bound of 35 million EUR presented in the Baseline Report [1].
Table 7.1 shows that the second material, Al-Li 8090-T851, has a faster ROI compared to the first material.
Looking only at the raw material influence, Al-Li 8090-T851 results in a total cost saving of 588,000 EUR over
1 Calculations based on A1 jet fuel for a fuel price of 0.4844 EUR/kg, determined from http://www.iata.org/publications/
economics/fuel-monitor/Pages/index.aspx, accessed on 12-06-2018
58 7. Structures
the total operational lifetime. Therefore, looking at the results from both evaluations, Al-Li 8090-T851 is cho-
sen as the material for the aircraft.
Assumptions and Parameters Several assumptions are made in order to construct the loading diagrams
for the fuselage. First, the load case needs to be determined. For this, the in-cruise condition was chosen.
Next, the fuselage is assumed to consist of two cantilever beams connected at the centre of gravity. The dif-
ferent structural weights, determined by the Class-II weight estimation method, are assumed to be constant
distributed loads. The main and tail lift acting on the structure are assumed to be point loads acting at their re-
spective aerodynamic centres (located at 25% of their respective MAC). The c.g. is assumed to be the average
of the c.g. range mentioned in the Stability & Control chapter (6). This c.g. value equals 16.87 m. A summary
of the distributed and point loads can be found in Table 7.3 below.
Distributed Point
Load Start [m] End [m] Magnitude [kN/m Load Location [m] Magnitude [kN]
Fuselage struc. 0 28.5 3.51 Cargo 23.5 16.77
Passengers 5.48 21.49 4.08 Main lift 16.27 272.81
Fuel weight 15.58 18.34 2.62 Tail lift 27.1 34.12
Wing struc. 15.58 18.34 8.58
Engines 21.46 24.96 6.67
Tail struc. 26.86 27.82 1.46
Loading diagrams The load distribution for the load case described above can be seen in Figure 7.1 and
Figure 7.2. From these graphs, it can be seen that the maximum shear force is equal to 154.23 kN and the
maximum bending moment to be 709.92 kN m.
These loads will be multiplied by the maximum load factor in the most critical load case and a safety factor to
get the design stresses and forces. The first factor follows from the flight envelope determined in Section 4.6
and equals 2.5. For the fuselage design, a safety factor of 2 is used.
7.2. Fuselage Sizing 59
0
100
Shear force [kN]
Moment [kNm]
−200
0
−400
−600
−100
0 5 10 15 20 25 0 5 10 15 20 25
Position [m] Position [m]
Figure 7.1: Shear force diagram along the fuselage. Figure 7.2: Moment diagram along the fuselage.
In this section, different stringers geometries and their properties will be discussed. In Figure 7.3 different
extruded stringers geometries are presented whereas in Figure 7.4 different formed stringers geometries are
shown. Integrally stiffened panels are not considered for the fuselage since they require complicated machin-
ing.
(a) Z-stringer. (b) Y-stringer. (c) I-stringer. (d) J-stringer. (e) Hat-stringer.
For fuselage application, all closed stringers (extruded Y-stringer, extruded hat-stringer, and formed closed-
hat stringer) cannot be used due to (corrosion) inspection issues, which would considerably increase the
maintenance of fuselage. J-stringer and I-stringer also eliminated due to added weight of extra rivets and
additional maintenance costs. The main disadvantage of extruded stringers is their cost and limitations on
the minimum thickness, therefore, formed Z-stringer will be used in the fuselage.
Assumptions
1. For crippling the Z-stringer is sized for Al7075-T6 due to lack of data for aluminium-lithium alloys.
Implications: Different stringer geometry might be optimal for aluminium-lithium alloys.
2. For crippling sizing the thickness across the stringer is constant.
Implication: The geometry of the stringer can further be optimised, for example for weight.
3. The total width of the stringer is the same as the total height of the stringer
Implication: The geometry of the stringer can further be optimised, for example for weight or area.
4. The skin carries the shear stress
5. The stringers carry normal stress (longitudinal stress)
6. Moments of inertia of stringers about their centroid is significantly smaller than their Steiner term,
therefore they are neglected.
Implications: Slightly lower moments of inertia, which would yield higher tensile stresses per boom.
Therefore the structure will be marginally overdesigned
7. For skin buckling, the skin between frames and stringers is considered flat.
Implications: Higher skin buckling stress would be achieved with a curved plate. Therefore the struc-
ture will be overdesigned.
8. For skin buckling, the skin is considered clamped between frames and hinged between stringers.
9. Stringer geometry is the same for all the sections
Implications: Different stringer types might be optimal for different loading conditions. Therefore there
is a possibility to reduce weight by introducing different stringers.
10. Stringers are equally spaced in the fuselage cross-section.
Implications: For different loading conditions, different sections of the fuselage will need a different
7.2. Fuselage Sizing 61
number of stringers. Therefore the amount of stringers will be overestimated as it is estimated for criti-
cal conditions.
11. Fuselage cross-section has a constant thickness.
Implications: The skin thickness will be designed to withstand the critical load case, thus it will be
overdesigned for all location other than critical location.
12. The analysis is done only for the cylindrical part of the fuselage.
Implications: Cone analysis is out of the scope for this report
Most assumptions used in this section are leading to more conservative design, leaving space to further opti-
mise the structure.
Structure idealisation The structure is idealised with equally spaced booms. Boom area is calculated using
Equation 7.1 from Megson [56].
t skin · b n,n+1 σn+1 t skin · b n,n−1 σn−1
µ ¶ µ ¶
B n = A stringer + · 2+ + · 2+ (7.1)
6 σn 6 σn
Where A stringer is the area of the stringer, t skin is the skin thickness, b is the distance between boom n and
boom n + 1. σ represents the direct stresses in the booms. The number of booms is equal to the number of
stringers needed. Once the boom areas and boom locations are known, moments of inertia are calculated
using Equation 7.2 [56]. Due to the symmetry of the fuselage, the centroid is in the geometrical centre of the
cross-section.
n n n
B n · x n2 B n · y n2
X X X
I xx = Iyy = Ix y = B n · xn · y n (7.2)
1 1 1
Due to symmetry, the I x y will be zero. All the stresses are calculated to withstand allowable loads. Allowable
loads are calculated as follows:
σtensileyield σcompressiveyield τmax
σmaxall = σminall = τmaxall = (7.3)
n load · n safety n load · n safety n load · n safety
The skin buckling is calculated to withstand the given loads per section, obtained from Figure 7.1 and Fig-
ure 7.2.
Crippling Compression in the aircraft fuselage can result in local instability failure, further referred to as
crippling. The crippling stress for entire geometry is computed by taking weighted average for each segment,
see Equation 7.4 [63]. Pn
(b n · t n · F ccn )
σcc = 1Pn (7.4)
1 (b n · t n )
Where b n is the segment width, t n is the segment thickness, and F ccn is crippling stress for a segment. Value
for F ccn can be found in appropriate graphs in the book of Niu [63].
Bending Normal stress due to bending is found using Equation 7.5 from Megson [56].
à ! à !
M y · I xx − M x · I x y Mx · I y y − M y · I x y
σz = ·x + ·y (7.5)
I xx · I y y − I x2 y I xx · I y y − I x2 y
Where M y and M x are bending moments in y and x-direction, respectively. x and y are the distances between
booms and x and y-axis, respectively.
Shear Shear flow due to shear force is calculated using the Equation 7.6 [56].
à ! à !
S x · I xx − S y · I x y Xn S y · I y y − Sx · Ix y Xn
qs = − 2
(B n · x n ) − 2
(B n · y n ) + q s,0 (7.6)
I xx · I y y − I x y 1 I xx · I y y − I x y 1
Where S x and S y are shear forces in x- and y-direction. Since the shear forces are assumed to be symmetric,
the resultant shear force will act through the centre of geometry which is also the shear centre. Therefore the
shear of the closed section q s,0 is zero. To obtain the shear stress, the shear flow needs to be divided by the
skin thickness.
62 7. Structures
Torsion Shear stress due to torsion is calculated using Equation 7.7 [56].
T
τt or si on = (7.7)
2 · A en · t ski n
Where T is the torque applied, and A en is the enclosed area.
Where p o and p i are outside and inner pressure respectively, r o and r i are the outer and inner radii. However
due to the fact that skin thickness is more than thousand times smaller than the radius of the fuselage, the
radial stress will be insignificant for stress calculation, therefore it is neglected.
Axial stress The axial stress due to pressure is calculated using Equation 7.9 [63].
pi · r
σa = (7.9)
2 · t skin
where r is the fuselage radius.
Circumferential stress Circumferential stress, also known as hoop stress is calculated in Equation 7.10 [63]:
pi · r
σhoop = (7.10)
t skin
Skin buckling Skin buckling is used to size the frame pitch. Firstly the effective stringer width is calculated
using Equation 7.11 [63]. s
Ke · E
b eff = · t skin (7.11)
F st
where K e is a compression buckling constant, E is the elastic modulus of the material and F st is the stringer
allowable stress. Then the stringer pitch is calculated taking into account the effective width of the stringer.
Skin buckling under in-plane compression is calculated from Equation 7.12 [63] and skin buckling under
in-plane shear is calculated using Equation 7.13.
³ t ´2 ³ t ´2
skin skin
F c,cr = K c · η c · E · (7.12) F s,cr = K s · η s · E · (7.13)
b stringer b stringer
Where K c and K s are constants for flat plate buckling. η c and η s are plasticity reductions factors for in-plane
compression and in-plane shear, respectively. b st r i ng er is the stringer pitch. K c and K s are based on the ratio
of stringer pitch to frame pitch and the boundary conditions. The appropriate graphs can be found in the
book of Niu [63].
Von Mises criterion The von Mises criterion for multi-axial loading is also implemented. Equation 7.14
shows the formula used [64].
s
(σx − σ y )2 + (σx − σz )2 + (σ y − σz )2 + 6 · (τ2x y + τ2y z + τ2zx )
σy > (7.14)
2
One can see that 53 stringers are needed with a stringer pitch of 16 cm. From crippling calculation, the opti-
mal geometry for formed Z-stringer is found to be 54 mm by 54 mm (symmetric about its web) with 0.9 mm
thickness. It cripples at maximum allowable compressive stress. The skin thickness is equal to 1.5 mm. The
frame pitch is 108.8 cm with 14 frames. However, for passenger comfort, there will be 20 frames to assure each
window seat has a window. The visualisation of the cylindrical part of the fuselage can be seen in Figure 7.5.
In Figure 7.6 the optimised number of stringers of the cylindrical section is given. One can see that the for
the critical cross-section 53 stringers are needed, whereas for the least loaded cross-section 26 stringers are
needed. In Figure 7.7 the allowable stresses for aluminium-lithium 8090 are given (continuous lines), and the
actual stresses experienced by the stringers (dashed lines). The dashed lines do not go beyond their allowable
counterpart. It means that the structure can withstand loads from Figure 7.1 and Figure 7.2, and not fail. An
ideal situation would be if the dashed lines coincide with continuous lines. This would mean that properties
of the material and structure are used to the fullest.
200
50
Number of stiringers [-]
100
Stress [MPa]
40
0
30 −100
5 10 15 20 5 10 15 20
Longitudinal position [m] Longitudinal position [m]
σmi n al l σmax al l τmax al l
Figure 7.6: Number of stringers versus longitudinal position σmi n σmax τmax
(t skin = 1.5 mm).
Figure 7.7: Stresses experienced by stringers across fuselage
(t skin = 1.5 mm).
In Figure 7.8 the normal, shear and von Mises stress distribution are present for skin thickness of 1.5 mm
and 53 stringers (stresses are expressed in MPa). In Figure 7.8a one can see that critical loads are in top and
bottom of the fuselage (due to longitudinal bending). With the top being in tension and the bottom being in
64 7. Structures
compression. In Figure 7.8b the critical location is on the sides of the fuselage, this is precisely the longitudinal
location of the lift force. The last figure shows the distribution of the Von Mises stresses for multi-axial loading.
It can be seen that the normal stress has a major contribution.
(a) Normal stress. (b) Shear stress. (c) Von Mises stress.
As one can see in Figure 7.8 the structure is overdesigned and is thus too heavy. An ideal situation would be
when all the failure modes occur at the design loads. Therefore, there are still opportunities for enhancement:
• As mentioned previously for less loaded sections at the front of the fuselage or at the back, fewer
stringers are needed. Further research into optimising the structure is recommended. One way to op-
timise the structure is to reduce the number of stringers where possible. Hence reducing the weight of
the structure. However, cutting the stringer will result in stress concentrations. There need to be some
other structural element at that point to reduce it and carry the load from the missing stringer.
• A variable skin thickness can also reduce the weight, hence optimise the structure. One needs to re-
member, however, that there is a limitation to minimum skin thickness, mostly due to the need of fas-
tening the stringers and other structural elements.
• Different stringer geometries have different specific properties. For example, in the section primarily
loaded in torsion, closed stringer are superior to open stringers. Therefore, fewer stringers would be
needed.
Other methods of structure optimisation can also be considered. The most efficient practice would be to com-
bine all different methods. This would result in having variable skin thickness, variable number of stringers
and different stringers geometry for differently loaded sections. However, such calculations are out of the
scope for this project.
20
Lift
Fuel
Weight
Load distribution [kN/m]
Resultant
Wing geometry
10
−10
0 2 4 6 8 10 12
Spanwise distance from wing root [m]
From Figure 7.10 it is clear that an optimal design for bending is reached when the thickness is increased
at the location that is furthest from the neutral axis. Given this theoretical design it is possible to define a
stiffener area and determine the number of stiffeners in combination with a skin thickness which results in
acceptable normal stresses (i.e. below yield stress). This results in an initial layout which can be determined
for the entire wing box based on local wing geometry and wing loading.
H
(q b /G t )d s
q s,0 = − H (7.15)
d s/G t
When the shear centre has been found as a function of the span, the lift force can be moved to the shear
centre and an additional torque (which will be the local lift force multiplied by the local distance to the shear
centre) can be added to accommodate the move. Shear flows in between booms can then be determined for
the whole cross-section at all spanwise locations. Shear flows in the idealised structure are calculated using
Equation 7.6. For the torque the shear flow will be determined using Equation 7.7. Shear flows can then be
translated to shear stresses using the local thickness as seen in Equation 7.16. From these calculations the
design can be updated to make sure shear stresses do not exceed the maximum allowable shear stresses
given by the material selection as given in Section 7.1. Performing several iterations the cross-section of the
wing at the root is illustrated in Figure 7.12.
q
τ= (7.16)
t
For the stiffener selection, reference aircraft have been studied since local skin buckling and stiffener crip-
pling is not taken into account for this initial sizing. As these values greatly influence the stiffener type se-
lection it was decided to base that on reference aircraft. Therefore hat-stiffeners are used at the top skin and
7.4. Verification & Validation 67
Z-stiffeners at the bottom skin as that is found to be very common in reference aircraft. The stiffener dimen-
sions are based on what is needed to deal with the applied loads (i.e. keep the normal stresses below yield
stresses), they can be found in Figure 7.13. The top and bottom skin thickness at the root is determined to
be 9 mm and the front and aft spar are designed to be 5 mm and 7 mm respectively. The aft spar is slightly
thicker due to the increased shear stresses. These occur there because of stresses due to shear and stresses
due to torsion act in the same direction at the aft spar, while they are counteracting at the front spar.
7.3.4. Sweep
The sweep creates an additional torque applied to the structure. The shear stresses caused by this torque will
be determined using the principle of superposition. The shear stress will be a function of the enclosed area of
the wing box and local thickness as seen in Equation 7.7. Aside from the additional torsional loads there will
also be local increases in stress at the root. At the leading edge, the stresses will increase due to the shear-lag
effect [63]. To relieve these stresses additional material needs to be added at the trailing edge. This problem
is not solvable using beam theory and is therefore reserved for a later stage in the design.
π2 E I
P crit = (7.17)
L2
7.4.1. Verification
This subsection contains the verification for the structures chapter. More insight into verification and vali-
dation procedures can be gained by reading Appendix A. First, material selection is treated followed by the
loading diagrams. Next, the fuselage and wing sizing are discussed.
Material selection A MATLAB program was written for the calculations used during the second material
trade-off. In order to verify the outcome of this program, hand calculations are performed to make sure the
outcomes are correct. The methods used are based on linear comparisons between material properties, but
there is no reference as to the validity of this approach.
Loading diagrams Several steps are taken in order to verify the code used to generate the loading diagrams.
First, the code was debugged to make sure there were no syntax errors. After this, the results are verified
by comparing them to hand calculations as well as an online interpreter2 . These calculations match and
therefore the code can be considered verified.
Fuselage sizing The MATLAB code is verified by different means. As coding progressed debugging is done
on a regular basis. Once all the coding is done, the boom locations, areas and moments of inertia of the whole
idealised cross-section are tested against example 22.1 from Megson [56]. Then the shear stress and normal
stress are tested against example 22.2 from Megson [56]. For both, the MATLAB results are the same as the
results in the book. Therefore, those functions in the code are verified. The crippling stress function is tested
against example 1, page 441 [63], here the results are within a reasonable margin. The discrepancy occurred
due to the necessity of line approximation in MATLAB code. Skin buckling and stresses due to pressurisation
are verified by testing different inputs, limits and later comparing them to manual calculations. Skin buckling
is within the margin due to another line approximation. Stresses due to pressurisation are the same, when
calculated in MATLAB and when calculated by hand. Therefore, the code is verified.
Wing sizing For the verification of the wing sizing calculations the MATLAB code used is verified in several
ways throughout the whole process. For the loading diagrams manual calculations were done to verify that
the total wing loading equals the aircraft’s weight with load factors added. Also several manual calculations
are performed to approximate the torsion and shear loads at fixed locations and see if the results are close the
values given by the MATLAB program. For bending loads manual calculations were performed on a highly
simplified wing box, approximated as a square box with a boom at each corner. In this way also shear flow
calculations have been verified. Furthermore, the MATLAB program which calculates shear flow from boom
locations and sizes is applied to examples from Megson [56] to verify its outputs. With matching results in all
above examples the code is considered verified.
7.4.2. Validation
This section contains the validation of the loading diagrams, fuselage and wing sizing respectively.
Loading diagrams Several aspects of the results can be used to prove the validity of the program. For exam-
ple, it is known that free ends of a structure do not carry loads. Since this is also reflected in the outcome of
the program, the program is considered to be validated.
Fuselage To validate the results of fuselage sizing building a prototype is recommended. First, a scaled
model should be built and tested, when small-scale model passes all the tests and experiments, the actual
size model should be built and tested (only the fuselage). Once the second model passes all the tests, the
assembly can be made and tested.
Wing sizing For validation of the design it is recommended to build the wing box structure and then ap-
ply the in flight loads to it. Stresses at several locations can then be measured using strain gauges and the
calculations can be validated. This should be done during a later stage of the design process.
2 SkyCiv Cloud Engineering Software, https://platform.skyciv.com/login, first accessed on 07-06-2018
7.5. Sensitivity Analysis 69
Table 7.6: Sensitivity analysis for fuselage loads. Table 7.7: Sensitivity analysis wing.
Five types of manufacturing processes are discussed. In casting, a molten metal is poured into a mold which
contains a hollow cavity of the desired shape, then it is cooled down and separated of the mold. This process
allows for complex shapes. In the next process, forming, permanent deformation is introduced by grain dis-
location by applying external forces. This process can also be done at high temperatures to lower the forces
required, however, this disrupts the treatment of the alloy. The welding process joins parts by melting their
surface and press them together while cooling down. The suitability of Al-Li 8090 to each of these methods is
given in Table 8.1.
70
8.1. Manufacturing Processes for Metal Components 71
Besides the manufacturing processes themselves, pre-processes and post-processes are needed, such as sep-
arating processes and joining methods. For now, only pre-processing methods will be discussed. Four differ-
ent post-processing methods are compared in Table 8.2. Labour cost as well as raw material cost were left out
of the trade-off since they were similar for all methods. No trade-off can be made yet, since the relative weight
of each criteria is part dependent.
Batch size About 53 stringers are needed per aircraft for the fuselage alone, so the production process must
be able to handle relatively large batch sizes.
Shape: The parts are simple: thin-walled beams without taper, with a continuous cross-section. The toler-
ances are average (in the order of ±0.1 mm).
Adaptability Even though many stringers are needed, there will be some variation. For instance, the Z
stringers from the fuselage are going to have different dimensions as those from the wing. A manufactur-
ing technique that can be easily adapted for the different beams with low additional tooling cost is preferred.
Production cost The last factor taken into account is the production cost per unit, this includes labour, tool-
ing cost and the raw material cost, which may differ per process due to the form it is in or the amount of waste.
There are a few different methods that are able to achieve this shape: extrusion, deep-drawing, superplastic
deformation and bending, they are evaluated below based on J. Sinke [85].
Extrusion This technique is able to produce large batch sizes, and is well adapted for constant cross-section
beams. However, it becomes expensive quickly when lots of different cross-sections are needed, even if they
have the same shape but different sizes. Therefore, it is not sufficiently adaptable.
Deep drawing Deep drawing is a much used process since it allows a large number of different products.
However this process has a really high equipment cost and is therefore only suitable for really large batch
sizes, therefore not well adapted for this case.
Superplastic deformation This type of permanent deformation only occurs in very specific conditions,
leading to only grain boundary sliding without changing the shape of the grains. This is therefore a very ex-
pensive and time-consuming process. This process is eliminated due to complexity and cost.
Bending This is one of the most common and simple processes for simple prismatic elements. The equip-
ment cost is relatively low since the equipment used are more or less universal machines. The cost for differ-
ent sizes of the same cross section is negligible, since the same die can be used when the angle remains the
same.
From this, it can be concluded that bending is the best process to produce the stringers, since the material
cost is low and is well adapted for many different sizes and shapes.
Batch size Since each rib has a different size, thus only 600 of each rib need to be produced over the course
of 15 years,therefore the batch sizes are relatively small.
Shape The ribs are more complex compared to stringers, because they are curved and with many holes.
The tolerances are average (in the order of ±0.1 mm).
Adaptability Since each rib of the wing has a different size, a manufacturing technique that can be easily
adapted for the different beams with low additional tooling cost is preferred.
72 8. Production Plan
Production cost The last factor taken into account is production cost per unit. This includes labour, tool-
ing cost and the raw material cost, which may differ per process due to the form it is in or the amount of waste.
The following methods can be used to achieve the desired shape: rubber forming, milling and casting. These
will be compared based on [85]
Milling Milling allows for a high amount of complexity, even in 3D. The right shape for each rib is directly
achievable without any pre-processing, and the mold or programming cost is low. However milling ribs re-
quires quite some handling since it will have to be clamped twice, and thus must be outlined very accurately
the second time. Therefore this process is both expensive in equipment, and in labour. Furthermore, it creates
a lot of waste.
Casting There are two main types of casting processes, reusable and non reusable mold casting. Due to
the high cost of the mold, reusable mold casting is only well adapted for series larger that 1000 units. In the
perishable mold category, sand casting can be immediately eliminated due to the too high minimal product
thickness (3 mm for aluminium). This leaves investment casting, this process has low to moderate mould
cost, however, it has high labour cost.
Rubber forming Rubber forming is a press forming process involving two tools: a rigid tool and a soft tool.
The rigid tool is shaped for the particular geometry of the piece. The soft tool is a universal rubber tool that
forces the sheet in or over the rigid tool by pressure. About 50% of parts in the aerospace industry are made
using rubber forming due to the low tooling cost and simplicity of the process. However the long cycle time,
in the order of a few minutes, could be a disadvantage
To conclude, since on average 40 sets of ribs need to be produced a year, the simplicity and low costs of
the rubber forming process outweigh the long cycle cost. Therefore rubber forming is chosen to produce the
ribs.
Table 8.3: Manufacturing processes for long fibre carbon reinforced polymers.
The first technique is pulltrusion. This is a process that is a bit similar to extrusion. First the fibres are wind
on spools, than the are impregnated with resin and then pulled trough a hot die, curing the product and
8.3. Manufacturing Sequence 73
fixing its shape. The product now needs to be cut in pieces. This produces constant cross-section beams,
with continuous fibres1 .
Filament winding is a technique that is mainly used to manufacture both open and closed cylinder-like
structures such as pressure vessels. This process involves winding filaments impregnated in a bath with resin
under tension over a rotating mold, laying down fibres in the desired pattern or angle. The resulting product
can then be cured in a vacuum-bag, at room temperature or in a autoclave or oven.
Next, hand lay-up is the simplest technique for CFRP, involving laying pieces of woven of braided carbon-
fibre fabric on a mold. After each layer, the resin is brushed on manually in order to impregnate the fibres,
before being put in a vacuum-bag and cured, either at room temperature, in an oven or in an autoclave.
The production rate is usually slow, since it involved large amounts of hand-labour, however the equipment
investment required is minimal so is well adapted for small batch sizes.
For infusion, like hand lay-up, pieces of woven or braided carbonfibre fabric are laid dry a mold. Then a
vacuum mold is build around it, and the space between the bag and mold is emptied of air. Once all the air
has been removed from the bag, liquid epoxy resin is introduced to the bag through a pipe which then infuses
through the fibres. After the product is fully infused, the supply of resin is cut off (using a pipe clamp) and the
resin is left to cure, still under vacuum pressure at room temperature. Sometimes a post-cure is needed in an
oven or autoclave to change the resin properties, such as to increase the glass transition temperature.
Prepegs are fabrics of carbon fibres that are pre-impregnated with resin. These fabrics are sticky on one
side, and are carefully stuck on a mold, checking that there are no bubbles or airgaps between layers. This
method allows for a higher fibre-to-resin concentration compared to previous methods, yielding usually
lighter structures. However, the rolls of prepregs are very expensive and the technique is very labour intensive.
Batch size For each aircraft, there are four different parts for the ailerons, two on each side. It is expected to
sell 600 aircraft over the course of 15 years, therefore the batch sizes will be quite small. It is therefore probably
not interesting to invest in expensive machinery.
Surface finish As for all aerodynamic components, surface finish is an important criteria, especially at the
leading edge, which is also the best location for a mold line. However, since the leading edge of the aileron is
not exposed, the possibility of flashlines is not really an issue.
Shape The cross-section of the ailerons is not constant, they are treated as a tapered beam. No double cur-
vature is present.
First, the pulltrusion process is eliminated since it is impossible to make tapered beams using this process.
Next, tape laying is eliminated for being too expensive. Furthermore, filament winding is not well adapted to
the desired shape. For more complex shapes, the filament speed and angle control becomes too complex to
predict accurately, and leads to a decrease in quality.
The three procedures that are left are hand-layup, infusion and prepregs. Prepregs are by far more expen-
sive, due to raw materials and labour costs, while scoring similarly on all other criteria. Therefore this process
is eliminated. Between hand lay-up and infusion, infusion is slightly more expensive due to the fact that more
consumables are needed. However this process is faster, and better suited for our batch size and allows for a
more consistent product. Therefore, the ailerons will be made using infusion.
Next the tail is more complex, since it involves moving parts. From the figure, it is clear that there are many
steps that can be performed in parallel. Furthermore, the steps of the horizontal tail and vertical tail are sim-
ilar, so the work packages that will be performed at the same time will all be of the same size. Therefore, the
tail is an element that can be paralised well.
The next element in order of complexity is the fuselage. The fuselage can be subdivided in four main
components from rear to front: the cone, the tube, the flight deck and the randome. Contrary to the tail, these
elements are not as well balanced, since they are not as similar.
Finally the most complex part, the wing. The wing is composed of many subsystems itself. The work pack-
ages are difficult to balance. Building one wingbox, including making all the ribs, stringers and including the
fuel tanks more time consuming than making the ailerons. In a next stage the wing has to be investigated into
further detail in order to limit waiting times and delays as much as possible.
Critical path The most critical path follows the action succession 1-2-7-10-11-12 and it takes 27.5 minutes
when no delays are present. This can be seen in Figure 9.1 by following the thick-circled actions. Important to
consider is the fact that the EC of actions 4 and 7 are at the same time, which means that they both represent
the critical path. Depending on which action finishes first, the latest completed action becomes part of the
critical path. A total slack time of 9 minutes was found for maximum delays [38].
Operation actions time estimation The time that each action of the ground operations takes is based on the
analysis performed by Gok [38]. On the other hand, some actions have different completion time depending
on which equipment is used and in which method the action is performed. For these operations, further
analysis is performed to determine the most optimal option in terms of time and costs. These operations are
identified to be embarking and luggage loading and unloading. Some remarks are also made about fuelling
and the use of an electric push-back.
75
76 9. Operations & Logistics
Time
Activity ES EC LS LC Slack
[min]
1 Chocks on 1 0 0 1 1 0
2 Passenger disembarking 3.5 1 4.5 1 5.5 1
3 Routine maintenance checks 10 4.5 14.5 5.5 15.5 0
4 Fuelling 6 4.5 10.5 5.5 13.5 2
5 Luggage unloading 2.75 1 3.75 3 6 2.25
6 Luggage loading 3.75 3.75 7.5 6 9.75 0
7 Cleaning and water service 6 4.5 10.5 5.5 13.5 2
8 Catering unloading 1 4.5 5.5 5.5 7.5 1
9 Catering loading 1 5.5 6.5 7.5 9.5 1
10 Passenger embarking 11 10.5 21.5 13.5 29.5 5
11 Flight OPS and procedures 5 21.5 26.5 29.5 34.5 0
12 Chocks off 1 26.5 27.5 34.5 36.5 1
(0) (2)
4.5 14.5 4.5 10.5 (Slack)
ES EC
5.5 15.5 5.5 13.5
LS LC
3
(0)
(1)
0 1
4 26.5 27.5
0 1
34.5 36.5
1 11 12
5 6
(0)
21.5 26.5
2 7 10
(2.25) 29.5 34.5
1 3.75
(0)
3.75 7.5
3 6
(1) 8 9
1 4,5
12.5
6 9.75
13.5 29.5
5.5 13.5
9.1.2. Embarking
Embarking is considered to be one of the most time-consuming actions when the aircraft is on the ground.
Therefore, different types of boarding are considered in order to determine the optimal one in terms of time.
In addition, the presence of hand luggage in the cabin has an impact on the boarding time. Two other aspects
of the boarding procedures that are considered are the aircraft chair type and the platform type to get in and
out of the aircraft. The different options are shown in Figure 9.2.
9.1. Ground Operations & Logistics 77
Boarding
Block boarding Reverse pyramid Hand luggage Slide slip seats Airstairs
Random
The type of boarding depends heavily on the loading platform, which will be discussed first. The type of
chairs present in the plane will be discussed after that, followed by the presence of hand luggage. Finally the
different types of boarding will discussed and an advisement for the aircraft operator will be given.
Platform For the embarkment and disembarkment of the passengers, three types of loading platforms can
be used: airbridges, movable stairs and airstairs. The decision on which one to use depends on the aircraft
operator, but an analysis is performed to identify the optimal platform to minimise operational costs and
decrease TAT.
Airbridges are the most convenient loading platforms considering passenger comfort. Walking to the
plane or a bus service is not necessary and head count is not needed. However, in terms of operational costs,
airbridges are the most expensive and take the longest time to connect to the aircraft. Besides that, airbridges
can cause damage to the aircraft which decreases the maintenance intervals. Also the airbridge might not
even be capable of connecting to a low aircraft. Because regional airports do not necessarily have bridges,
this option is not used.
Movable stairs are available at most regional airports. They are cheaper in use than airbridges, but are not
faster in terms of connecting and have the same risk of causing damage. Delays may occur if stairs are not
present or used for other aircraft.
Airstairs are built-in stairs in the door of the aircraft. Positioning them does not take much time and due
to the short height of the plane they can be easily implemented on the door of the plane. Airstairs have the
disadvantage that they add weight to the plane. However, because of their small size this would only be an
extra weight of 65 kg [20]. The stairs increases 0.06% of the Direct Operating Cost (8 USD per trip) but are still
much cheaper than the movable stairs or airbridges, which commonly cost about 20 and 110 USD respectively
[20]. On the other hand, depending on where the aircraft is parked, people would need to walk to the aircraft
or use a bus service, which could cost time and money. Considering though that the airstairs need only 30
seconds to get ready to be used and to put them back, they need a decrease of 3 minutes in the TAT, which is
about 10% less than the total TAT. Due to the lower cost and shorter turnaround time, airstairs are selected.
Chairs The three types of chairs are shown in Figure 9.3. Side slip seats are only available for a 3 chairs
seating configuration, which means that they are not applicable on the MW 5. Furthermore, foldable chairs
could be advantageous for a 2 seat configuration but the advantages are not that big that it makes sense to
implement them. Therefore, conventional seats are advised to be the best seating option.
Luggage Having passengers without any hand-luggage in the cabin could drastically decrease the embark-
ing time to 3.5 minutes [19]. Although this seems optimal, the solution is not favoured for the sake of passen-
ger comfort. Also, the time to load the luggage as cargo would strongly increase. Therefore, an aircraft without
luggage in the cabin is not an option that is explored.
Boarding Types The type of boarding has the most impact on the time to board. All the different method-
ologies are discussed and subsequently analysed to identify the optimal boarding type. Boarding types are
78 9. Operations & Logistics
decided by the aircraft operators but an analysis with recommending purpose is performed. All these meth-
ods were studied by Van den Briel [41] and the times to embark were adapted to 76 passengers. The resulting
embarking time per type can be seen in Table 9.2. The values in the table are without considering the type of
platform used.
• Back-to-front: all the seats are filled from the back to the front, but there is no difference made between
window or aisle seats.
• Outside-in: with this method the passengers are boarding in columns, meaning first all the window
seats are filled and after that all the aisle seats.
• Block boarding: this is basically the same method as the outside-in method, but it is subdivided into
zones, where the zones are ordered back to front.
• By-seat: this method calls each passenger per seat. For larger aircraft, this is a time consuming and
unfeasible strategy.
• Flying carpet: this is a new method in which the passengers stand on a numbered carpet according to
their seat number; after this, the back located seats are allowed to enter first.
• Random: this boarding method does not have a particular order in which the passengers can embark
the aircraft.
• Reverse pyramid: this method is the combination of outside-in and back to front boarding.
Table 9.2: Embarking time of 76 passengers for each boarding methodology [41].
Back-to-front boarding is not advised as it is the slowest. Even though they are quick embarking types, reverse
pyramid and by-seat are also not advised as they are complex methods. This leads to high delay susceptibility,
which would potentially make the embarking procedure longer than other simpler methods. The flying carpet
has been studied to be potentially the fastest embarking methodology, but this would only be advantageous
if an airbridge would be used, which was discussed to be generally not available at the targeted airports. For
this reason, only block, random and outside-in are the embarking types that are considered most optimal.
9.1. Ground Operations & Logistics 79
For the PERT model analysis, the random embarking methodology was taken as it is the most commonly
used, and it was calculated to take about 11 minutes to embark 76 passengers. If an outside-in embarking
method would be used, 2 minutes could be potentially saved from the total turnaround time. The downside of
this method is that people travelling together might board at different times, which may be not the preferred
solution.
Regarding disembarking, 2 seconds per passenger are needed for the procedure [38]. For this reason, 2.5
minutes are needed to disembark all the passengers if no issue occurs. In case that it goes slower, 3 seconds
per passenger are taken into account.
• No conveyors;
• Classic belt loader;
• Ramp snake;
• Power stow;
• Sliding carpet.
Considering regional airports, ramp snakes and sliding carpets are predominantly not available. The "no con-
veyor" solution is not advised, unless it is the only possibility, as it is not favourable for workers. Lastly, the
classic belt loader and the power stow are left. These are both usually available in regional airports and can
be utilised depending on the preference of the operator. The normal belt loader is able to load one bag every
10 seconds, while the power stow one every 7 seconds [89]. Considering that the power stow is easily and
quickly put in place, the time advantage of the power stow is significant. For this reason, the power stow is
advised and used for the TAT analysis. Assuming that normally 25% of the passengers bring checked luggage
for short regional flights, about 20 luggage pieces need to be loaded and unloaded each cycle. This results in
2.5 minutes to load and 1.5 minutes to unload, as 5 seconds per luggage piece are needed for the latter.
9.1.4. Fuelling
For fuelling during the ground operations several assumptions are made. As most regional airports have dif-
ferent facilities to fuel an aircraft, a worst-case scenario was made. It is assumed that on regional airports,
tank trucks which are capable of refuelling at a rate of 1100 l/min are present. An estimate of 5000 kg of fuel is
needed, which is approximately 6350 l; it then takes 5.8 minutes to fill the tanks. Because the valve needs to
be attached and detached, the fuelling time is rounded up to 6 minutes. At bigger airports, fuel trucks or even
build-in fuel stations are capable of fuelling at a much higher rate. This will also mean that the fuel time will
be much lower for those cases. Finally, fuelling is crucial when considering the critical path. In fact, passen-
gers are not allowed to embark or disembark while this procedure is done, due to safety reasons. If fuelling
takes many minutes, it may become part of the critical path and increase the total TAT. For this reason, the
faster the refuelling rate the better it is.
9.2. Maintenance
The final main element of the operations and logistics of the design is maintenance scheduling and logistics.
Maintenance strategies and scheduling will be elaborated upon in this section.
• En-route servicing
• Terminating pre-flight checks
• Service checks
• Maintenance checks (A, B, C, D)
Besides scheduled maintenance, there is also unscheduled maintenance which is defined in the ATA-100
specification as:
Those maintenance checks and inspections on the aircraft, its systems and units which are dictated
by special or unusual conditions which are not related to the time limits specified. Includes in-
spections and checks such as hard landing, overweight landing, bird strike, turbulent air, lightning
strike, slush ingestion, radioactive contamination, maintenance checks prior to engine-out ferry,
etc.
In order to reduce the maintenance cost, and hence the operational cost for the airline, the maintenance has
to be performed as easily as possible. This is done by guaranteeing easy access to parts or subsystems which
are prone to maintenance. In terms of downtime of the aircraft scheduled maintenance is always preferred
over unscheduled maintenance.
The chosen logistical strategy is to regularly operate the aircraft at regional airports that have minimal tech-
nical support, only enabling routine checks, and to land at airports with appropriate maintenance facilities
when other types of checks and maintenance are required. The operator is advised to combine this strategy
with a commercial flight, such that the generated revenue is increased.
Corrective maintenance Fixing or replacing components either when they have failed or when they are
found to be failing. There are no interventions until a failure has occurred. The advantage of corrective main-
tenance is that the component lifetime is extended to the maximum potential. However, it also has some
severe implications on safety and grouping of maintenance because 100% of the maintenance which has to
be performed is unscheduled.
Preventive maintenance Replacing or repairing components or systems before failure occurs. The advan-
tage of preventive maintenance is that failures of critical systems and consequential damage as a result of
failure of critical systems is reduced. The downside of preventive maintenance is that it is hard to determine
the optimal moment for replacement of the part. Preventive replacement means that the full useful life of the
part is not used, which increases the maintenance cost. Furthermore, early replacement may push a compo-
nent back to the infant failure regime.
9.2.4. Accessibility
Because the fuselage is situated 1.4 m above the ground it is easily accessible from the ground for mainte-
nance checks of the lower part of the fuselage. The low wing configuration also allows for easy accessibility
for the wings from the ground without requiring any maintenance platforms.
Maintenance platforms are only required to provide accessibility to the upper side of the fuselage, the tail
and the engines. The low required amount of maintenance platforms means that most of the maintenance
activities can also be performed at workshops with limited facilities.
10
Aircraft Systems
In this chapter, the different systems in the aircraft will be described. First, the resource allocation will be
discussed in Section 10.1, consisting of the mass budget and the power budget. Thereafter, in Section 10.2,
the initial sizing of the landing gear covered and in Section 10.3 a description of the Auxiliary Power Unit is
given. The hardware and software diagram, data handling block diagram and electrical block diagram can be
found in Section 10.4, Section 10.5 and Section 10.6, respectively. The communication flow diagram is given
in Section 10.7. Next, the hydraulic system description, fuel system description can be found in Section 10.9
and Section 10.10, respectively. Finally, the chapter concludes with a section about the safety characteristics
of the aircraft.
82
10.2. Landing Gear 83
From Equation 10.1 it is found that the main wheels’ diameters are approximately 0.79 m and their width
is around 0.25 m. The nose gear carries a maximum of 15% of the weight [75]. For the nose gear a dynamic
load from breaking also needs to be taken into account. The dynamic breaking load can be calculated using
Equation 10.2 [75]. The maximum static load is represented by W , H is the aircraft c.g. height with respect
to the ground which is as a first approximation assumed to be in the midpoint of the fuselage and B is the
distance between the main and nose-wheels. To determine the values for H and B the location and height
of the main landing gear need to be determined. These will be determined in the next subsection. Using
those results and by calculating the maximum load on the nose-wheel, a nose-wheel diameter of 0.6 m with
0.14 m width is found to be required. This is 80% of the main wheel parameters which seems a reasonable
first estimation for nose-wheel sizing [75]. For the selection of tires a tire book from the manufacturer can be
requested. Then, the smallest tires that are able to carry the maximum loads are selected.
10 × H × W
dynamic load = (10.2)
g ×B
such that during operation the most aft c.g. tip back angle is minimally 15° and the most forward c.g. tip back
angle is maximally 30°. During ground handling (passenger embarking and cargo loading) the c.g. should stay
in front of the landing gear so it does not tip back. The location of the nose-wheel can than be determined by
assuming the nose-wheel carries less than 15% of the weight with the c.g. in the most forward position and
more than 5% for the most aft position (this is important during operation for good nose-wheel traction). We
then find the main gear is located 17.85 m from the nose of the aircraft and the nose-wheels are located at
2 m. The nose wheel then carries in between 6% to 10% of the aircraft weight during operation.
For determining the wheelbase the overturn angle is the main constraining factor. In Figure 10.2 [75] a visual-
isation is given of how to find the overturn angle. This angle should not exceed 63° [75], from which the static
ground line can be calculated to be 1.38 m. Then the angle between the c.g., the nose-wheel and the main
wheel can be calculated. This is found to be 4.5° giving a minimal wheel base of 2.82 m, meaning the wheel
track is slightly wider than the fuselage diameter. For this reason the main wheels will be attached to the wing,
aft of the wing box. The retracting system, struts, brakes and more will be designed in a later stage of the de-
velopment. Based on the main wheel height and location it can be found that the lateral ground clearance is
7.4° which is more than the required 5° [34].
advantage, however, is reduced total ground time, since no ground tug nor ground crew will be needed. An
estimated 2 to 7 minutes [23] per flight can be saved using wheel tug, this translates to roughly 172 USD/trip
saving [20].
Starter Motor
Actuator Cargo Bay
Wiring&Net
Power Supply Lines Volume Connection
Soundproofing
Mechanism Access Doors
Hinges/Skin
Container Fastening
Rudder
In Figure 10.3 observe that the five hardware components are marked with variously coloured dotted lines,
with the name of the block shown at the top. The software blocks are shown within the dotted block, with
their names shown in bold. Components of the software block are collected vertically, with names shown
in regular font. Relations between individual hardware or software blocks are shown by means of red arrow
86 10. Aircraft Systems
lines.
Flight Controls: The flight controls can be divided into primary flight controls and secondary flight controls.
The primary flight controls are used to control the aircraft in pitch, roll and yaw, and are absolutely
necessary to control the aircraft during flight. The secondary flight controls are primarily used to im-
prove the flight characteristics during certain flight phases and manoeuvres [54]. The aircraft uses an
electrical fly-by-wire control system, because this is lighter than a conventional manual flight control
system. Input from the primary flight computer goes to the control column to provide the pilots with
feedback. A stick shaker is also fitted into the control column, which shakes the control column when
the aircraft is about to stall.
Air Data Inertial Reference Unit: The Air Data Inertial Reference Units (ADIRUs) are one of the key compo-
nents as they collect air data and provide position and altitude information [8]. Normally the data of
ADIRU 1 is used, and the data of ADIRU 2 is used to verify that the data provided by ADIRU 1 is correct.
Flight Management System: The flight management system uses the data of the air data inertial reference
unit, the flight controls, the autopilot settings and weather data to determine the current position. It
also calculates the optimum course and vertical flight path to follow to minimise fuel consumption.
[58]
Electronic Flight Instrument Systems: The electronic flight instruments systems consists of a primary flight
display (PFD), a multi-function display (MFD) and the engine indicators and crew alerting system
(EICAS) display, which is shared by the captain and the co-pilot.
Flight Control Surface Actuators: The inputs on the flight controls and autopilot settings are processed by
the primary flight computer and send to the control surface actuators, which are subsequently used to
move the control surfaces.
The data flows between the other components can also be found in Figure 10.4.
Power Provision: These components are the ones responsible for generating the electrical power to be used
by the other systems. When the aircraft is parked, the main source of power is often ground power.
Many airports prefer the use of ground power over the APU in order to reduce the noise and pollution
produced. For this reason, the APU is only used when ground power is unavailable, to start the main
engines or during an emergency in flight.
Power Storage: The Power Storage department contains the components that hold the materials that are
converted into electrical energy by the Power Provision department. Electrical energy is either stored
directly in the main battery or indirectly in the form of jet fuel. The battery provides the power necessary
to start the APU, which in turn is used to start the main engines. The jet fuel is an indirect source of
electrical power, as it is first combusted, after which electrical power is generated.
10.7. Communication Flow Diagram 87
GPS
Flight Management
Electronic Flight System
Instrument Systems
Air Data Inertial
Reference Unit 2
Primary Flight Display
Elevator Actuators
Digital Weather Left Engine
Surveillance Radar
Aileron Actuators
Right Engine
Rudder Actuators
Transponder
Power Demand: Shown here are the main aircraft components that require electrical power. The functions
of these components range from monitoring the aircraft status, to provide passenger comfort or to
guarantee the safety of the people on board.
Power Conversion: This block represents the components that convert the power into the voltage/current
required by the components in the Power Demand department. The engine provides 115V /230V Al-
ternating current (AC), to be converted into 28V and for some components into direct current (DC) as
well.
The arrows connecting the components speak for themselves. In the Power Storage department, the fuel tanks
supply fuel to the main engines, and the batteries provide electricity to start the APU. The Power Provision
department then recharges the batteries through means of an alternator. On the ground, power is supplied
by either the APU or the airport supply, depending on which is available. In the air, the engines are the main
source of power. Only in an emergency, during which the engines are unable to supply power for whatever
reason, will the APU or the Ram Air Turbine take over and supply power to critical systems. The supplied
power is then converted into the desired type and voltage of the components in the Power Demand depart-
ment.
Safety Measures:
Engines Slides/Oxygen Masks Pressurization Flight Instruments
Provide Power Fire Supp/Fuse Box
Alternator
Recharge
Power Storage
Ground Power
Ground/Emergency
Power
Other Battery Fuel Tanks
aircraft
ypted Data
ata handler
Command aircraft
Check if input
subsystems
command is met
Supply Fuel
Aircraft
subsystems
GNSS
Legend:
al Continuous orange-Informative communication Pitot tube Fuel monitoring system
Dashed pink- Command communication
Dotted purple- Matter flow Fuel pressure Oil pressure
Sensors and indicators
Aircraft Health Vertical speed Fuel temperature Fuel quantity
Monitoring Airspeed indicator Altimeter
Gyroscopic instruments indicator Propulsion
(AHM) monitoring system
Hydraulic pressure Fuel flow
Attitude indicator Heading indicator Fire sensors Other sensors
Sensor
Ground services
Navigation system
Air Traffic Control
Communication (ATC)
avioncs
Operation Control
Autopilot Pilot Centre (OCC)
Yaw control Pitch control Roll control Propulsion system Fuel tanks
Engines An engine failure occurs if one of the engines suddenly stops producing thrust due to a malfunction
other than fuel exhaustion. An engine failure during one of the critical phases, such as take-off or landing, can
endanger the continuation of the flight significantly.
Hydraulic system Because the hydraulic system is an important safety-critical system, the aircraft has sep-
arate and redundant systems to power the control surfaces, flaps, landing gear, brakes and spoilers.Each of
the two engines is equipped with a hydraulic pump and there is also an electric power hydraulic pump as a
back-up in case of a double-engine failure.
Fuel system The fuel system includes the fuel tanks to store the fuel and the pumps to feed the engine from
the tanks to the engine. The failure of the fuel pumps that send fuel to the engines will result in the loss fuel
delivered to the engines as gravity feed isn’t possible because they are mounted higher than the fuel tanks.
Also, fuel pumps are also required to prevent large fuel imbalances in the wing that will make it difficult to
control the aircraft.
Avionics The avionics provide crucial information such as the altitude and airspeed to the pilot and flight
computer. The failure of systems related to the avionic system would result in the aircraft being unsafe to fly,
especially in poor visibility conditions.
Flight control system The flight control system is crucial for translating the pilot inputs to deflections on
the control surface. The fight computers between the pilot inputs and control surface deflections need to
protect the aircraft against manoeuvres that would bring it out of the flight envelope and expose the airframe
to large stresses.
Fire suppression system There are areas on the aircraft that are exposed to the combination of high temper-
atures and combustible materials such as the engines and landing gear bays. Areas identified to have a high
risk of catching fire will need to be equipped with fire suppression systems to contain the fire and preventing
it to spreading to other areas of the aircraft.
Engines There are two engines on the aircraft for redundancy. The aircraft is designed to be able to maintain
level flight in the event of a single engine failure. To provide redundancy in the remote case of a double engine
failure, a ram air turbine will need to be incorporated in the power subsystem to continue providing electricity
to power flight critical systems.
Hydraulic systems The redundancy philosophy applied for the hydraulic system is to have triple redundant
hydraulic systems. The goal of this is to eliminate single points of failure that could cripple the hydraulic
system.
Fuel system There will be multiple fuel pumps in the fuel system of the aircraft to ensure redundancy. Ad-
ditionally, there will be independent fuel lines from the tanks to the engine to prevent a blockage in one tank
to causing the entire fuel system to fail.
Avionics The entire avionic system from the pitot tubes to the displays will be redundant with independent
instruments for the pilot and co-pilot.
Flight control systems Given its importance to control the aircraft, the fly-by-wire control system and flight
computers will require many layers of duplication to ensure redundancy from failures from faulty or damaged
components. An extra form of redundancy will be achieved by allowing the flight computers to use alternate
or direct control laws when the avionics provide faulty data. There will also be a mechanical backup via the
elevator and rudder trim controls to give the pilots some pitch and lateral control in the very remote event
90 10. Aircraft Systems
of a complete failure of the flight control system or while resetting the flight control computers after a loss of
electrical power.
A B C
C
B
A
Brakes
Left Airbrakes A B Right Airbrakes
A B B C
Landing gear
B
Figure 10.7 displays the schematic layout of the hydraulic system as of now. Components like flaps, brakes,
landing gear and tail control systems have been given redundancy by being operated by two hydraulic sys-
tems. This is done because these components are seen as critical components, for which failure of one of
these results in significantly reduced aircraft performance or the aircraft even being unable to fly at all. Re-
dundancy for the components on the main wing, excluding HLDs, is provided in a slightly different way.
Instead of having each component powered by multiple hydraulic systems, each component is split into two
parts, with each part powered by a different hydraulic system. The components for which this is done are the
spoilers, airbrakes and ailerons. Should any of these components fail on one side of the wing, the component
on the other side is unaffected and will still be functioning properly. The result is a loss in performance of that
subsystem, but safe flight remains possible.
size in the wings is 5.37 m3 , within the fuselage another tank of 10 m in length below the floor is needed to
have a fuel tank size of 10.7 m3 . The fuel tank requirement was determined in [39] and is based on reference
aircraft data. This tank gives a maximum fuel weight of 8.5 t which is more than sufficient for a typical flight
and actually allows the MW 5 design to fly quite some extra kilometres when reducing the number of passen-
gers as can be seen in the payload range diagram Section 4.5.
The fuel system layout is shown in Figure 10.8. The red line in this figure is the wing of one of the reference
aircraft, the Fokker 902 . The dark green outline represents fuel tanks. Light green lines are crossfeed lines,
brown ones go to the engines/APU and the light blue piping is used for fuel jettison. Note: the fuel system
layout is assumed to be symmetrical with respect to the longitudinal axis and therefore only half is shown.
The total development costs for the entire program are projected to be 1 billion USD. The division between the
parameters can be found in Figure 11.1. The share of development cost per aircraft is calculated by dividing
the total development cost of the manufactured aircraft, which is estimated to be 600 in 15 years. This gives a
development cost per aircraft of 1.6 million USD.
DS & TC
9.9 % AE & DC
15.3 %
FTAC
47.3 %
15 %
10 %
2.5 % RDTE
FTOC RDP
92
11.2. Operating Cost 93
The total production cost for the entire program is projected to be 18.4 billion USD. The production cost
per aircraft is calculated by dividing the total production cost over the manufactured aircraft. This gives a
production cost per aircraft of 30.4 million USD.
Company Model Engine type Unit Price [Million USD] seating capacity
AVIC XIAN MA700 turboprop 25 68-86
ATR 72-600 turboprop 26 68-78
AVIC COMAC ARJ21 turbofan 32 78-105
Bombardier Q400 turboprop 32.2 82-90
Sukhoi Superjet 100 turbofan 35.4 87-108
Group 22 MW 5 turbofan 38.6 64-76
Bombardier CRJ700 turbofan 41 66-78
Embraer E170/E175 turbofan 41.5 70-88
Mitsubishi MRJ70 turbofan 46.3 69-80
The three turboprop aircraft are the ones with the lowest unit price. An exception is seen for the AVIC COMAC
ARJ21, which has a price that is comparable to turboprop powered aircraft even if it has turbofan engines.
Furthermore, by looking at Table 11.1, the MW 5 unit price is about middle priced compared to the com-
petitors. The unit price then results to be 9% more than the average of all the direct competitors and 1.4%
less if compared with turbofan mounted aircraft only. These results make the MW 5 attractive to airline op-
erators as the unit price is about median between competitors. Furthermore, the aircraft is designed to have
lower operational costs than comparable aircraft, making it additionally attractive. The operational costs are
analysed and displayed in Section 11.2.
which is partly due to the aircraft being lighter (Section 3.2) than competitors but also due to a more efficient
engine as found in Section 4.1. Furthermore, the maintenance cost of the engine will be significantly lower
than the ones that are currently operated. This is the case as the chosen geared turbofans have 1500 less air-
foils than current turbofans, which results in a predicted 40% decrease in maintenance costs [93]. As for the
landing fees, they are mostly directly related to the maximum landing weight of the aircraft which is signifi-
cantly lower for the MW 5 design in comparison to competitors [16]. The effect of fuel burn efficiency, engine
maintenance and landing fees on the total DOC is a reduction of 8-11% with respect to current competitive
aircraft. The values of cost reductions are given in percentages because the actual numbers are significantly
airline dependent. For example, fuel price and maintenance cost largely depend on the amount of operations
that an airline performs. The values given in Table 11.2 are typical values but those will be different depending
on the operator.
Cost MW 5
DOC component DOC element
[USD/block hour]
Crew $ 255
Flying Fuel $ 1267 (492.3 Gallons) −10% to −15%
Insurance $ 56
Airframe $ 410
Maintenance
Engine $ 51
Airfame $ 234
Depreciation Engines $ 196 −40%
Avionics $ 52
Landing fees $ 155 (based on $7.75/1000lbs) −15% to −30%
Landing &
Navigation fees $0
navigation fees
Registery taxes $6
Finance Finance $ 185
Total $ 2875 −8% to −11%
11.4.1. Manufacturer
The BEP for the manufacturing company is strongly related to the profit percentage at which each plane is
sold. The profit was decided to be 20% the program costs. The reason to chose this rate is explained by the
fact that the BEP is wanted to be reached after 45% of the aircraft batch is sold. The manufacturing company
wants to keep the risk of not having a positive RoI as low as possible. The 45% is estimated to be a value
that is for sure expected to be sold. The break even point is then reached once 270 planes have been sold.
Considering the production of 40 planes a year, this is reached after 6.75 years.
11.4.2. Operator
The BEP for the operators is calculated with equation Equation 11.2:
AEP
BEP = (11.2)
(REV − DOC − IOC)Uannbl Vbl
The outcome expresses a BEP after 9.25 years of the beginning of the aircraft operation. This would leave 5 to
10 years of revenues, if considering a 15 to 20 aircraft lifetime.
• Management: The management needs to be paid competitively in order to gain top talent to run the
program.
• Strategy: Money needs to be spent on marketing to bring awareness to the strengths of the aircraft to
potential operators of the aircraft with the goal of convincing them to buy our aircraft. Consultants
are hired to give a better view of the market and to provide insights and recommendation that the
management can act upon.
96 11. Financial Analysis
• Engineering: This includes the cost to design the aircraft and necessary tooling to produce it. The cost
to design the aircraft is the cost to hire engineers, build computing capabilities and carry-out testing.
• Certification: This is the cost to get the aircraft certified by airworthiness bodies.
• Office Expenses: The costs for providing an office for all employees to work in.
• Financing: These costs needs to be paid to the bank that is financing the program.
11.6.1. Reliability
One of the most important reliability parameters for aircraft is the dispatch reliability: the percentage of flights
that depart within a specified time window of the scheduled departure time. This time window is usually set
to 30 minutes, which is short enough that flight schedules are not affected too much and long enough to have
time to fix small mechanical problems, assuming spare parts and maintenance personnel are on hand. When
the aircraft does not depart within these 30 minutes, the aircraft does not depart "in time" and is considered
delayed.1
Program Cost
Marketing Overhead
Part Availability
Consultants Manpower Warehouses
Legal Jigs
Flight Operations
Prototyping Assembly Platforms
Planning
Testing Quality Control
Advising
Rent Warehouse
Insurance Transportation
Salaries Procurement
Utilities Materials
Overhead Engines
Financing Avionics
Interest Interior
Fees
Of all delayed flights, external factors and non-aircraft issues can be excluded. This results in the maintenance
dispatch reliability, which only includes delays and cancellations due to unplanned maintenance. External
factors that are not dependent on the aircraft used are then not taken into account. The maintenance dis-
patch reliability strongly depends on cancellations due to controllable reasons and delays due to unplanned
maintenance.
In Table 11.3, only the percentage of cancelled flight operations are given per aircraft type for controllable
and uncontrollable reasons. Controllable reasons mainly include unplanned maintenance, whereas uncon-
trollable reasons include external factors and non-aircraft issues, such as severe weather, air traffic control
congestion, lack of catering or crew, security, etc. Cancellations due to unplanned maintenance are control-
lable, because they can be avoided to a certain extent by regularly performing maintenance checks.
Of the 2.6% of the flights with regional jets that were cancelled, 31% was caused by controllable reasons. This
means that approximately 0.8% all flights with regional jets are cancelled due to unplanned maintenance,
because unplanned maintenance is the largest contribution to cancellations due to controllable reasons.
However, for the maintenance dispatch reliability not only cancellations but also delays have to be taken into
account. In Table 11.4 the number of delays and cancellations for different aircraft types can be found. The
98 11. Financial Analysis
data in the table was found by looking at delays and cancellations of 1.95 million scheduled flights.2
From the table it can be seen that for small regional jets 4.5 percent of all flights were cancelled or delayed
due to maintenance. For large regional jets, 4.6 percent of all flights were cancelled or delayed due to main-
tenance. For the maintenance dispatch reliability, this results in a reliability of 95.5 percent for small regional
jets and a reliability of 95.4 percent for large regional jets. This means that large regional jets are almost as
reliable as small regional jets when looking at the maintenance dispatch reliability.
Small regional jets often carry less than 70 passengers, whereas large regional jets typically carry 70 to
100 passengers. Because the MW 5 can carry 76 passengers, it is a large regional jet and the maintenance
dispatch reliability is therefore targeted to be at least equal to the average maintenance dispatch reliability of
large regional jets, which is 95.4 percent.
11.6.2. Availability
For the design of the aircraft, the operational availability can be further split into two parts: the operational
availability A o and the inherent availability A i .
Operational availability is a measurement of the percentage of time the aircraft is available for flight. It
takes scheduled and unscheduled maintenance into account, as well as logistic delays. The operational avail-
ability depends on the Mean Time Between Maintenance Actions (MTBMA), the Mean Maintenance Time
(MMT) and the Mean Logistic Delay Time (MLDT). The operational availability is given in Equation 11.3 and
can be improved by increasing the time between maintenance actions, when allowed by regulations, and by
reducing the maintenance time and logistic delays.
MTBMA
Ao = (11.3)
MTBMA + MMT + MLDT
The components of the low-cost regional aircraft that are maintenance intensive i.e. the engines and high-lift
devices should have characteristics such as being easily accessible and removable which allows for the MMT
to be reduced. This contributes to reducing the direct operational cost of the aircraft.
The inherent availability is a performance parameter that depends on the Mean Time Between Failures (MTBF)
and the Mean Time To Repair (MTTR). The formula for the inherent availability is given in Equation 11.4 and
can be used as a first estimation, because internal transport and logistic delays are ignored.
MTBF
Ai = (11.4)
MTBF + MTTR
2 Retrieved from: http://www.worldtek.com/wp-content/uploads/2015/09/Measurement-of-Dispatch-Reliability-Tulinda-Larsen-Sept
pdf on 07-06-2018
11.6. Reliability & Availability 99
The inherent availability is improved over existing aircraft through reducing the MTTR by leveraging the air-
craft health monitoring systems proposed in Section 9.2. This allows the operator to better predict when
failures are going to occur and pre-allocate resources to address the failures more quickly, thus reducing the
MTTR.
12
Sustainable Development
Commuter air traffic is increasing with 5% per year and is therefore a fast growing sector, where airlines are
growing rapidly and air traffic is becoming more affordable. To reduce the environmental influence of growing
air traffic, the air traffic sector has to become more sustainable. Sustainable design is a crucial aspect to ensure
the future of air traffic. First a definition of sustainable development is provided in Section 12.1. Secondly, the
design choices that influence the sustainability of the design are provided in Section 12.2. Finally, the airframe
noise is determined and compared to reference aircraft in Section 12.4.
12.1. Definition
In order to implement sustainable development it is first necessary to define sustainable development. The
aim of sustainable development is to meet the needs of the current generation, without compromising the
ability of future generations to meet their needs [94]. The chosen definition of sustainable development is the
definition as is provided in the Oxford Dictionary1 :
"Development that is conducted with the aim to keep the depletion of natural resources and the
environmental impact as low as possible."
Aviation system
Aircraft Aircraft
Figure 12.1 shows the distinction between the subsystems of the aviation which can or can not be influenced
by the aircraft design. Not every aspect of the aviation system can be influenced by implementing new tech-
nologies in the aircraft design. However, measures increasing the level of sustainability of the airfield subsys-
tem and the air traffic management subsystem can be implemented. The airfield subsystem can be divided
in three different elements: structural, aerodynamic and propulsion.
100
12.3. Air Traffic Management Subsystem 101
Structural weight The structural weight of the aircraft is determined by the loads experienced by the air-
craft during operations as well as the selected materials. Because the loads that the aircraft has to be able
to withstand are set by the flight envelope and regulations, the only parameter to influence the structural
weight of the aircraft is the material selection. The influence of the maximum take-off weight, and hence the
operational empty weight, is presented earlier in Table 4.13.
The four main different types of materials considered for aircraft design are metal alloys, fibre reinforced
composites, metal matrix composites and fibre metal laminates. Each of the material types has its own range
of specific strength, determining the structural weight. In terms of structural weight it is beneficial to select
the material with the highest possible specific strength.
End of life solutions The possible end of life solutions depend on the selected materials, but the general
goal is to recycle the materials as much as possible. The materials considered for recyclability are the same
materials as mentioned for the structural weight, but metal matrix composites are discarded due to the low
technology readiness level.
Recycling metal alloys can be done quite easily by melting the parts that are to be recycled. A metal air-
craft can be recycled up to 85% by mass, which is quite attractive in terms of sustainability. Unlike for fibre
reinforced composites, the recycled metal alloys are allowed to be reused for critical structural parts.
There are significant challenges in recycling fibre reinforced composites. It is by legislation not allowed
to reuse fibre reinforced composite parts for critical structural elements. The most promising, but at this day
rather uncertain, method of recycling the resin of fibre reinforced composites is pyrolysis which is the thermal
decomposition of the material. However, development of this process is insufficient to consider it as an end
of life solution for the design on an industrial scale. A second possible recycling method for fibre reinforced
composites is burning the resin and only retrieving the remaining fibres. The fibres can then be used in a
different process, such as injection moulding.
According to Boom [17] it is possible to extract up to 98% of the original metal alloy with its original
composition, which is a relative high yield of recycled metal alloy. The first step is to separate the metal alloy
from the glass fibres making use of thermal delamination. After the metal alloy has been separated from the
glass fibres, the metal alloy can be remelted and refined.
system wide impact. During the different flight phases there are strategies that can be implemented to reduce
the environmental influence of the aircraft.
Some of the strategies discussed in [77] already are discussed in Chapter 4, resulting in a total fuel saving
of 37 kg per mission, which results in a CO2 emission reduction of 116.5 kg per operation.
The landing gear geometry for the reference aircraft is, within the scope of this project, nearly impossible to
obtain, so the same landing gear geometry is used for all aircraft in the ANoPP. The results from the ANoPP
show that the airframe noise produced by the design is lower than the airframe noise produced by comparable
aircraft.
75
MW 5
CRJ700
ERJ-170
70
Noise [dB]
65
60
55
0 1 2 3 4
log10 (frequency) [Hz]
Table 13.1: Scale to measure the severity and likelihood of technical risks.
OP-PR-2: Engine fire The three essentials for fire are all present at the engines, i.e. fuel, ignition source,
and oxygen. Therefore, the likelihood level of an uncontrolled engine fire is set as reasonably possible. In the
chosen design concept with fuselage mounted engines, the consequence of an engine fire has the potential
to be catastrophic as it can spread to the cabin and empennage.
OP-CS-1: Hydraulic system failure The total failure of the hydraulic system is catastrophic in severity as it
will result in the loss of flight controls but it is unlikely. Also, a special consideration to take into account for
the fuselage mounted engines at the rear of the aircraft is that an uncontained engine failure, even though
unlikely might result in shrapnel damaging the hydraulic pipes located in its vicinity.
OP-GO-1: Damage from ground operations Rough handling by the ground crew or collision of ground
equipment with the aircraft can cause damage that degrades its performance or creates the need for un-
scheduled maintenance. Hence, the severity of this event is marginal. Due to the configuration of the final
design, the wings and fuselage are close to the ground which make them vulnerable to collision from other
fixed and mobile objects which increases the likelihood of this to reasonably possible.
103
104 13. Technical Risk Assessment
OP-CS-1: Hydraulic system failure Due to the severity of this risk event, multiple mitigation actions must
be taken such as ensuring redundancy in the entire system and the elimination of single point of failures to
reduce the likelihood to remote.
OP-GO-1: Damage from ground operations To abate this risk, built-in flood lights could be incorporated
into the design of the aircraft to assist in ground operations at night. Cameras can also be installed at the tail
to allow the pilot to have a better view of the aircraft’s surroundings. These mitigation strategies should bring
the likelihood down, the event can then be considered unlikely.
DEV-PR-2: Engine integration The PW1215G has a fan diameter of 1.46 meters and is significantly wider
than the General Electric CF34-8C engine used on the Bombardier CRJ700 which has a fan diameter of 1.17
meters2 . Fuselage-engine integration issues such as aerodynamic interference and flutter could have a critical
impact on the program if fixes can not be found, requiring a change to an engine of a smaller diameter.
Nevertheless, this is deemed unlikely as there already exist business jets with similar configurations that have
engines almost as wide as their fuselages such as the Gulfstream G550.
DEV-AR-1: Shock-waves on wing The design of the wing utilise supercritical airfoils in order to delay the
onset of shock-waves on the wing when flying in the transonic region. CFD was not performed for the wing
design due to a lack of time, expertise, and equipment such as computing clusters and hence an assumption
was made that compressible effects can be ignored for supercritical airfoils in the transonic region. Neverthe-
less, there is a risk that the assumption may not be valid and that there are shocks on the wing at the target
cruise Mach number. This is accessed to be unlikely as similar regional aircraft such as the CRJ700 and E170
already operate at the designed Mach regimes. The occurrence of shocks would still be critical as it would
require the development of new airfoils or at worst a redesign of the wing.
DEV-ST-1: Structural weight change The structural design of the wings and fuselage make use of many
simplifying assumptions at this early stage of the design. The likelihood of a structural weight change is rea-
sonably possible as the structure is overdesigned to some extent for safety and lacks the finer analysis of the
smaller details such as cutaways and access hatches. Therefore, it would be possible to see weight changes
further down the program. This would have critical implications on the overall design of the regional aircraft
as the other subsystems will be affected as well.
1 https://www.wsj.com/articles/airbus-delivers-a320neo-to-lufthansa-1453306044
2 https://www.geaviation.com/sites/default/files/datasheet-CF34-8C.pdf
13.3. Risk Maps 105
DEV-ST-2: Production cost overrun The materials mix in the MW 5 includes modern materials namely the
Aluminium-Lithium alloy Al-Li 8090-T851 for the wing and fuselage and carbon fibre reinforced polymer
(CFRP) for the control surfaces and stabilisers. This carries the risk of ballooning production costs as special
considerations must be made such as specialised and separate tooling for AL-Li 8090-T851 to prevent lithium
contamination of other parts and increased complexity of quality control for CFRP. This risk is reasonably
possible with a marginal impact on the program.
DEV-PR-2: Engine integration Computational fluid dynamics (CFD) and analysis with advanced structural
models should be performed in the detailed design in order to identify any possible integration issues early
on so that more resources can be invested to address them without causing major program delays. This miti-
gation step should reduce the severity of this risk to marginal.
DEV-AR-1: Shock-waves on wing Expertise and computing resources for CFD analysis should be acquired
as soon as the detailed design phase begins, in order to validate the wing design or discover any potential
problems early on in the project to reduce the severity of this risk to marginal.
DEV-ST-1: Structural weight change To reduce the severity and likelihood of this risk, finite element mod-
elling with detailed part designs need to be executed. The structures department should also make predic-
tions on which directions the weight change of the different parts will be in and communicate them with the
other departments so that necessary preparations can be made.
DEV-ST-2: Production cost overrun Communication channels should be opened with experts on aircraft
production and suppliers for the material and tooling to gain a better understanding of the challenges with
producing an aircraft with Aluminium-Lithium and CFRP. This understanding needs to be translated into
better estimates of the production costs involved and implementable steps to reduce costs where possible.
Lean thinking should also be applied to reduce waste and cost and result in the severity level to be negligible.
106
14.2. Project D&D Gantt Chart 107
Determine Flight
Instruments/Avionics
Generate Cockpit
Lay-Out
Size Electrical
Determine Aircraft
Systems
Wiring Plan
Design Aircraft
Lighting
Analyze Stress
Concentrations
Analyze Aircraft Design
Structure Joints/Bonding
Design Cut-Outs
Determine Required
Final Finalise Detailed Tools
Design/Symposium Design Integrate Results
Determine Required
Manufacturing
Machinery
Guarantee Worker
Safety
Test Aircraft Safety
Register Aircraft
(Structural Strength) Design Wingtips
8-1-18 11-19-18 3-9-19 6-27-19 10-15-19 2-2-20 5-22-20 9-9-20 12-28-20 4-17-21 8-5-21 11-23-21 3-13-22 7-1-22 10-19-22 2-6-23 5-27-23
Detailed Design
Preparing final design
Aircra� structure
Aerodynamics
Flight performance
Size subsystems
Produc�on plan
Opera�ons and Logis�cs
Integrate results
Project development
Tool fabrica�on
Set-up produc�on facility
Produce prototype aircra�
Test prototype
Cer�fica�on process
Opera�onal life�me
Transport 50 to 75 passengers with the lowest possible operational costs over a distance of 1500 to
2500 km with an aircraft from regional airports with 1500 m tarmac runway maximum.
The project was split up into two major parts. The first phase focused on the aircraft design such as flight
performance and stability and control. The second part focused on the financial aspects of aircraft exploita-
tion. The first part focused on optimising the aircraft performance, with the main driver being cost to quality
ratio, whereas the second part focused on assessing the investment. In this chapter, the final conclusions and
recommendations are presented.
Conclusions
Conclusions for both both major parts of the report as explained above are present in this section. The con-
clusions are categorised to their relevant design department.
Structures For both the wing and the fuselage, critical load cases have been determined and designed for.
The whole process of designing has been reported and calculations have been performed for one critical lo-
cation. For the wing this location is at the root and for the fuselage it is at the location of the wing where the
lift acts. The main goal of the project is keeping the MW5 operating cost as low as possible. For the structural
design this mostly means keeping the structural weight to a minimum. The critical location for the fuselage is
at 16.27 m from the nose of the aircraft(location where the wing lift acts). For that cross-section 53 stringers,
14 frames and skin thickness of 1.5 mm are required to sustain the loads. For the wing, the number of stringers
needed at the root is 32 with a local rib pitch of 0.465 m. The goal of lowest weight was obtained by designing
a structure that would just withstand the maximum load. By iterating the design, meaning adding and remov-
ing material where possible or necessary, the design can be optimised further. The methods of designing the
wing and fuselage for critical load conditions can be used to determine an optimal design for the rest of the
wing and fuselage as well.
Aerodynamics The wing planform and airfoil selection has been performed with the aim to reduce the
induced drag during cruise as much as possible. The resulting lift over drag ratio of the clean wing during
cruise is 34.4, but this is excluding the drag caused by the fuselage and tail sections. The optimised wing has
an area of 61.1 m2 and a total taper ratio of 0.31. The lateral stability of the aircraft is increased by applying a
dihedral angle of 3°. The resulting wing can achieve a maximum lift coefficient in clean configuration of 1.07
at an angle of attack of 11.5°.
The horizontal and vertical tail are designed to provide the aircraft with sufficient longitudinal and lat-
eral stability. The computed horizontal tail surface of 10.1 m2 is sufficient to ensure both the stability and
controllability of the aircraft for each possible centre of gravity location. The vertical tail area of 11.8 m2
108
109
Flight performance Firstly, the engine selection was performed based on the thrust required over the entire
flight mission, its fuel efficiency, dry weight and environmental impact. Based on these criteria, a scaled-
down version of the PW1215G, which belongs to the most advanced geared turbofan family in production as
of 2018, was chosen to power the aircraft.
Secondly, the balanced field length was calculated to be 1490 m which fulfils the customer’s take-off re-
quirements of 1500 m. Following that, the relevant flight phases were investigated in order to propose opera-
tion optimisation strategies to reduce the fuel burn of each flight mission, and hence directly lower its opera-
tional cost. The requirement to fly a distance of 2500 km at maximum payload was then verified through the
construction of the payload-range diagram. With zero payload, the aircraft is able to travel a distance of 4175
km.
Moreover, the flight envelope and V-n diagram were made to determine the aircraft’s operational limits
and the structural strength the airframe has to be designed to withstand. To be competitive with comparable
regional aircraft such as the Bombardier CRJ700 and Embraer E170, the maximum operating Mach number
and service ceiling were set to be 0.78 and 41,000 ft respectively. The V-n diagram determined that the aircraft
needs to be designed to handle a maximum and minimum load factor of 2.5 & -1.
The turn performance was investigated to determine the aircraft’s maximum achievable turn rate at a
critical condition of 7000 ft at MTOW, which is the upper limit of typical hold altitudes. It was determined that
the MW 5 is capable of meeting the most demanding standard turns i.e. the rate 2 turns.
Stability and control Regulations require the aircraft to be dynamically stable in all expected aircraft con-
ditions and configurations. Based on the expected seating arrangement, subsystem placement and weights,
a suitable wing placement and tail size was determined to meet these requirements. As of now, controllability
is the most limiting condition, while a large stability margin is guaranteed during flight. Based on the geome-
try and specifications of the MW 5 design, a detailed analysis of the aircraft’s flight behaviour was performed.
This analysis resulted in the five eigenmotions that occur in flight. The conclusion that can be drawn from
this analysis is that the aircraft is stable in four of these motions, and the only unstable eigenmotion is stable
enough to allow plenty of time for the pilot to react, which is the driving factor to pass aircraft regulations for
this eigenmotion.
Finance The finance chapter discussed all the analyses that are money-related. Firstly, an estimation of the
program cost was calculated to be 19.4 billion USD. From these costs, the unit price of the MW 5 was set
as 38.6 million USD, such that a RoI of 19% was obtained for the manufacturing company. This could be
achieved by wanting a BEP of not later than 50% of aircraft sold. In fact, the BEP is now found as 45% of the
aircraft sold, which is at 270 sold planes. From the operating cost analysis it followed that a reduction of 8%
to 11% in operating cost is obtained. Based on that, RoI and BEP could be determined from the perspective
of an MW 5 operator. These were found to be 12.2% and 9.25 years respectively. In order to have a better
overview of the financial side of the program, a cost-breakdown analysis was performed. Finally, reliability
and availability were tackled. These are crucial to strengthen the interest of investors. In fact, the more the
MW 5 aircraft are expected to be reliable and available, the more revenues an aircraft operator makes.
Recommendations
Further research into the proposed low-cost regional aircraft, MW 5, is required in order to achieve even lower
operational costs. In this section, the recommendations per department are elaborated.
Structures Now that the preliminary sizing for the wing and fuselage is done, for the next phase in the de-
sign it is recommended to continue the design iterations and optimise the design for each specific section of
the aircraft. For the wing this means determining the design along the entire cross-section and for the fuse-
lage the next step would be to even further optimise for weight, which can be achieved by introducing stringer
discontinuity, variable skin thickness and different stringer’s sizes, geometries and placements. Furthermore,
it must be noted that the current design only deals with static loads. Once the design has been optimised
it is important to look further into the effect of dynamic loading to make sure the aircraft can withstand all
possible loading scenario’s. The landing gear structure needs to be designed and especially the connection
of the landing gear to the wing needs to be designed for relatively high dynamic loads. In the end phase of
the design several experimental tests should be performed as validation before the aircraft is able to start test
110 15. Conclusion & Recommendations
flights. At that moment loads can be applied to a prototype aircraft structure and strain gauges can be used
to read out stresses at all relevant locations.
Aerodynamics In a next phase of the project, it is recommended to look into CFD in order to test the as-
sumption that no shock-waves are present, and in order to have more accurate results. Also, it is interesting
to look into methods to delay transonic effects such as wing fences. Once the correct CFD models are avail-
able, it is also recommended to look into the design of wingtip devices. Wingtip devices can further decrease
the induced drag during cruise, which saves fuel and hence increases the sustainability of the aircraft and
reduces operational costs. The sizing of the vertical tail is based on a statistical method, but a more extensive
investigation to the dimensions of the vertical tail has to be performed in the next design phase.
Flight performance The flight performance can be improved with further and more detailed analysis. Firstly,
there should be a deeper look into the take-off and landing phases. More accurate data on tires, engine accel-
erating performance, take-off operations, reverse thrusters, optimal power and breaking capabilities should
be gathered. Secondly, more research can be done into optimal climbs and descents, as also established by
regulatory operations. Other flight operations need to be studied in more detail, if they are possible and how
they should be dealt with specifically, for example cruise climb. Furthermore, an in depth analysis can be
performed regarding noise minimisation. This also should be done by keeping in mind operative regulations.
Regarding engine thrust, its effective power varies with altitude and mach number. A detailed performance of
the selected engines could be requested to the manufacturing company in order to get more accurate results.
This could strongly improve the flight performance analysis. Finally, a flight path analysis with a specific route
including a diversion to an alternate airport would give a better understanding of a real-life scenario and the
resulting aircraft performance.
Stability and control While this requirement for the aircraft to be fully stable and controllable for the entire
range of c.g. positions is met, improvements can still be made. From the loading diagram it can be observed
that the most aft c.g. position occurs on the ground, and while the aircraft is in the air, the c.g. position is
significantly more forward. If the position of the aircraft components can be adjusted in such a way, that
the c.g. range becomes more centred, the overall stability and control performance of the aircraft can be
improved. Aircraft control and manoeuvrability can be further improved by a more detailed analysis of the
control surfaces design. Deploying control surfaces often produces forces and moments around undesired
axes, which is to be minimised. A possible approach in minimising this is analysing the effect of differential
control surface deployment in more detail.
Finance The financial analysis is based on the results of the current preliminary aircraft design. With the
increase of accuracy in further design phases, more reliable values can be found for finance. More analysis
can be done to determine the program cost more accurately and consequently the unit cost. More detailed
cost descriptions should be found for the various aircraft manufacturing parts. Regarding operating cost,
numbers can differ a lot between operators. Therefore it is important to work together with an airline to
determine the operating cost specific to their case.
Bibliography
[1] Group 22. Baseline Report. Tech. rep. Delft University of Technology, Faculty Aerospace Engineering,
May 2018.
[2] MTU Aeroengines. PW1000G. URL: http : / / www . mtu . de / engines / commercial - aircraft -
engines/narrowbody-and-regional-jets/pw1000g/. 2018.
[3] European Aviation Safety Agency. Certification Specifications for Large Aeroplanes, CS-25. Cologne,
Germany, 2007.
[4] R. Alderliesten. Fatigue and Fracture of Fibre Metal Laminates, Solid Mechanics and Its Applications.
Springer International Publishing AG, 2017.
[5] John D. Anderson. Aircraft Performance and Design. Singapore: McGraw-Hill, 1999.
[6] John D. Anderson. Introduction to Flight. 7th. New York, USA: McGraw-Hill, 2012.
[7] Nicolas E. Antoine and Ilan M. Kroo. Aircraft Optimization for Minimal Environmental Impact. reader
Vol. 41 No. 4. American Institute of Aeronautics and Astronautics, 2004.
[8] J.D. Aplin. “Flight Computer for the Boeing 777”. In: Microprocessors and Microsystems 20.8 (Apr. 1997),
pp. 473–478.
[9] ATR. Connecting the future, Turboprop Market Forecast 2016-2035. Tech. rep. ATR, 2015.
[10] Centre for Aviation. USD1 trillion for airport construction globally - but it’s not enough. Tech. rep. CAPA,
2017.
[11] GE Aviation. CF34-3 turbofan propulsion engine. URL: https : / / www . geaviation . com / sites /
default/files/datasheet-CF34-8E.pdf. 2018.
[12] GE Aviation. CF34-8C turbofan propulsion system. URL: https : / / www . geaviation . com / sites /
default/files/datasheet-CF34-8C.pdf. 2018.
[13] P. Balakrishnan et al. Natural Fibre Composites and their Applications in Aerospace Engineering. Vol. 23.
6. 2010, pp. 871–893.
[14] BCA. Market Forecast BCA 2017-2036. Tech. rep. Bombardier, 2016.
[15] Bombardier. Commercial Aircraft Prices. 2018.
[16] Bombardier. CRJ Series. Montreal, Canada, 2015.
[17] R. Boom. Recycling of aluminum from fibre metal laminates. Tech. rep. 2012.
[18] A. Boone. CRJ700 Aircraft Systems Study Guide. Boone and Rile publishing, 2000.
[19] Menkes van den Briel. Airplane Boarding. tech report. 2010.
[20] F.G. Carrasco and Scholz Dieter. Promising aircraft modifications for low handling costs. 2008.
[21] Cessna. Citation II.
[22] S. Chapple et al. “Flammability of Natural Fiber-reinforced Composites and Strategies for Fire Retar-
dancy: A Review”. In: Chemical Engineering and Processing 51 (2012), pp. 53–68.
[23] WheelTug Company. WheelTug technology. 2017.
[24] T.C. Corke. Design of Aircraft. 4th. University of Notre Dame: Pentice Hall, 2003.
[25] Granta Design. CES EduPack. comprehensive database of materials and process information. 2017.
[26] Dr. Tulinda Larsen. The Connected Aircraft: Improving Dispatch Reliability. Sept. 2015.
[27] T. Dursu et al. “Recent developments in advanced aircraft aluminium alloys”. In: Materials and Design
56 (2014), pp. 862–871.
[28] E. Mooij, E., Papp, Z.„ van der Wal,W. Simulation, Verification and Validation:Lecture Notes. internal
publication. Feb. 2018.
[29] EASA. EASA type certificate data sheet E.018. URL: https://www.easa.europa.eu/sites/default/
files/dfu/EASA%20TCDS%20E%20018%20issue%2012.pdf. 2018.
[30] EASA. EASA type certificate data sheet E.063. URL: https://www.easa.europa.eu/sites/default/
files / dfu / EASA - TCDS - E . 063 _ Rolls -- Royce _ Deutschland _ Tay _ Series _ engines - 04 -
18062013.pdf. 2013.
111
112 Bibliography
[31] EASA. EASA type certificate data sheet IM.E.096. URL: https : / / www . easa . europa . eu / sites /
default/files/dfu/EASA%20TCDS%20IM%20E96_PW814GA-PW815GA_01.pdf. 2017.
[32] K.E. Eickmann. Committee on assessment of aircraft winglets for large aircraft fuel efficiency. The na-
tional academies press, 2007.
[33] Embraer. Market Outlook Embraer 2017-2036. Tech. rep. Embraer, 2016.
[34] FAR. Airworthiness Standards: Transport Category Airplanes. Tech. rep. Feb. 1965.
[35] Antonio Filippone. Advanced Aircraft Flight Performance. Cambridge University Press, 2012.
[36] Lewis R. Fisher and Herman S. Fletcher. EFFECT OF LAG OF SIDEWASH ON THE VERTICAL-TAIL CON-
TRIBUTION TO OSCILLATORY DAMPING IN YAW OF AIRPLANE MODELS. Langley Field, VA, 1955.
[37] C. Giummarra et al. “NEW ALUMINUM LITHIUM ALLOYS FOR AEROSPACE APPLICATIONS”. In: Pro-
ceedings of the Light Metals Technology Conference (2007).
[38] Y.S. Gok. Scheduling of aircraft turnaround operations using mathematical modelling: Turkish low-cost
airline as a case study. reader. Coventry University, 2014.
[39] Group 22. Baseline Report. Tech. rep. Delft University of Technology, Faculty Aerospace Engineering,
May 2018.
[40] Group 22. Midterm Report. Tech. rep. Delft University of Technology, Faculty Aerospace Engineering,
June 2018.
[41] S. Hartjes. Operations Control and Flight Planning. internal publication. 2018.
[42] Lance Sherry Houda Kerkoub Kourdali. A comparison of two takeoff and climb out flap retraction stan-
dard operating procedures. Tech. rep. URL: https://ieeexplore.ieee.org/document/7486344/
authors. Delft, Netherlands: Delft University of Technology, 2016.
[43] I. Inagaki et al. “Application and Features of Titanium for the Aerospace Industry”. In: NIPPON STEEL
& SUMITOMO METAL TECHNICAL REPORT 106 (2014).
[44] Aleris International Inc. AEROSPACE ALUMINUM AA5028 AlMgSc. Tech. rep. 2015.
[45] Korea Aerospace Research Insitutey. MRJ(Mitsubishi Regional Jet)의 현황과 개발과정. URL: https :
//www.kari.re.kr/download/viewer/1471022116937/index.html. 2016.
[46] S. Jackson. Systems engineering for commercial aircraft. Ashgate, 1997.
[47] Ebad Jahangir. Aviation and Sustainable Development: Trends and Issues. Tech. rep. ICAO, 2009.
[48] Roskam Jan. Part 8: Cost estimation, Design development, manufacturing and operating. Ed. by DAR-
corporation. Basingstoke, UK, 1999.
[49] Roskam Jan. Part I: Preliminary Sizing of Airplanes. Ed. by DARcorporation. Basingstoke, UK, 1999.
[50] Ray Jaworowski. Long-Term Growth Projected for Regional Aircraft Market. Tech. rep. Forecast Interna-
tional, 2017.
[51] Lloyd R. Jenkinson. Civil jet aircraft design. 1st ed. Great Britain: Arnold, 1999.
[52] David Kaminski-Morrow. IndiGo tentatively signs for 50 ATRs. 2017.
[53] S. Krishna et al. “A Comparative Study of Aluminium Alloy and Titanium Alloy”. In: International Jour-
nal on Recent Technologies in Mechanical and Electrical Engineering (IJRMEE) 2.8 (2015), pp. 85–88.
[54] Wolfgang Langewiesche. Stick and Rudder: An Explanation of the Art of Flying. 1st ed. New York, NY:
McGraw-Hill Professional, 1990.
[55] Granta Design Limited. CES EduPack 2017, version 17.1.0. accessed via TU Delft subscription. 2017.
[56] T.H.G Megson. Aircraft Structures for engineering students. Fourth edition. Great Britain: Elsevier Aerospace
Series, 2008.
[57] J. Melkert. Aerospace design and systems engineering elements slides, Design of the fuselage. Internal
publication. Delft, Zuid-Holland, 2018.
[58] Sam Miller. “Contribution of Flight Systems to Performance-Based Navigation”. In: AERO Magazine.
Seattle, Washington: Boeing Commercial Airplanes, 2009, pp. 21–28.
[59] L. R. Miranda, R. D. Elliot, and W. M. Baker. A generalized vortex lattice method for subsonic and super-
sonic flow applications. Contractor Report 2865. Lockheed-California Co for N. A.S.A., 1997.
Bibliography 113
[60] MIT lecture outlines. Unified lecture #2: The Breguet Range Equation. online. Oct. 2008.
[61] J.A. Mulder. Flight Dynamics. TU Delft, 2013.
[62] N.V. Nayak et al. “Composite Materials in Aerospace Applications”. In: International Journal of Scien-
tific and Research Publications 4.9 (2014).
[63] M.C.Y Niu. Airframe stress analysis and sizing. 2nd ed. Granada Hills, CA: Conmilit Press, 1999.
[64] NPTEL Mech. Eng. lecture outlines. Plasticity. online. Oct. 2004.
[65] Marius G. Oprea. THE EFFECTS OF GLOBAL ECONOMIC CRISIS ON THE AIR TRANSPORT OF PASSEN-
GERS IN EUROPE AND IN ROMANIA. Journal. GeoJournal of Tourism and Geosites, 2010.
[66] International Civil Aviation Organization. Global Air Transport Outlook To 2032 And Trends To 2042.
Tech. rep. ICAO, 2017.
[67] International Civil Aviation Organization. World Aviation and the World Economy. report. 2012.
[68] A.N. Page. ESDU 87018. Tech. rep. there are many authors but only the chairman is written. 1987.
[69] Y. Park. “Fuel burn rates of commercial passenger aircraft: variations by seat configuration and stage
distance”. In: Journal of Transport Geography 41 (2014), pp. 137–147.
[70] K.L. Pickering et al. “A review of recent developments in natural fibre composites and their mechanical
performance”. In: Composites: Part A 83 (2016), pp. 98–112.
[71] A. Pramanik. “Problems and solutions in machining of titanium alloys”. In: The International Journal
of Advanced Manufacturing Technology 70.5-8 (2014), pp. 919–928.
[72] N. Prasad et al. “Mechanical behaviour of aluminium–lithium alloys”. In: Sadhana 28 (2003), pp. 209–
246.
[73] Ivchenko Progress. D36, Series A Turbofan. URL: http://ivchenko-progress.com/?portfolio=d-
36-4&lang=en. 2018.
[74] Ivchenko Progress. D-436-148 Bypass Turbofan Aero Engine. URL: http://ivchenko-progress.com/
?portfolio=d-36-4&lang=en. 2018.
[75] Daniel P. Raymer. Aircraft design: A conceptual approach. 3th. Reston, Virginia: AIAA education series,
1999.
[76] Nikkei Asian Review. Mitsubishi Aircraft loses order for 40 regional jets. Tech. rep. Nikkei, 2018.
[77] T. G. Reynolds et al. EVALUATION OF POTENTIAL NEAR-TERM OPERATIONAL CHANGES TO MITI-
GATE ENVIRONMENTAL IMPACTS OF AVIATION. Tech. rep. 2010.
[78] T.G. et al. Reynolds. Evaluation of potential near-term operational changes to mitigate environmental
impacts of aviation. Tech. rep. 2010.
[79] J. Roskam. Airplane Design Part IV:Layout of landing gear and systems. DARcorporation, 2000.
[80] P. K. C. Rudolph. High-Lift Systems on Commercial Subsonic Airliners. NASA, 1996.
[81] Ger J.J. Ruijgrok. Elements of Airplane Performance. 2nd ed. Delft: Delft Academic Press (VSSD), 2009.
[82] Safran. SaM146: performance and adaptability. URL: https://www.safran- aircraft- engines.
com/commercial-engines/regional-jets/sam146/sam146-1s17. 2018.
[83] Schrenk. “A simple approximation method for determining spanwise lift distribution”. In: NACA Tech-
nical Memorandum 948 (1940).
[84] D. Sholz. Estimating the Oswald factor from basic aircraft geometrical parameters. Hamburg, Germany,
2012.
[85] Jos Sinke. Production of Aerospace Systems. reader AE3211-II. internal publication. TU Delft, 2010.
[86] Matthew Miller Siva Govindasamy. Exclusive: China-made regional jet set for delivery, but no U.S. certi-
fication. Tech. rep. REUTERS, 2015.
[87] Richard Smith. Mitsubishi’s jet set for push into Middle East. 2016.
[88] W. Speer et al. “Applications of an aluminum–beryllium composite for structural aerospace compo-
nents”. In: Journal of Air Transport Management 8 (2004), pp. 895–902.
[89] Power Stow. Time measurement for loading with Power Stow Rollertrack. 2011.
114 Bibliography
[90] E. G. Tulapurkara. Airplane design(Aerodynamic). Dept. of Aerospace Engg., Indian Institute of Tech-
nology, Madras.
[91] M. Voskuijl. Flight & Orbital Mechanics slides, Airfield Performance. internal publication. Delft, Zuid-
Holland, 2018.
[92] R.J.H. Wanhill. “Status and prospects for aluminium-lithium alloys in aircraft structures”. In: Fatigue
16.1 (1994).
[93] Graham Warwick. Civil engines: Pratt & Whitney gears up for the future with GTF. 2007.
[94] WCED. Our Common Future; report of the UN commission on Sustainable Development. Tech. rep. 1987.
[95] Pratt & Whitney. Pure Power PW1200G Engine. URL: https://www.pw.utc.com/Content/Press_
Kits/pdf/ce_pw1200g_pCard.pdf. 2018.
[96] Y. Yang et al. “Recycling of composite materials”. In: Chemical Engineering and Processing 51 (2012),
pp. 53–68.
Program and Code Bibliography
In order to be able to design an aircraft, not only methodologies and techniques as described in books or
articles are used, computers and numerical tools were used as well. However, just like design methods, it is
important to build on others and not invent everything yourself, while giving credit where it is due. Therefore,
this section is dedicated to reference all external programs and libraries used in the design.
All code written by the team, as well as some of the open source libraries used can be found on GitHub,
https://github.com/paulienuij/DSE_group22_Low-Cost-Regional-Aircraft.
• MATLAB: version R2018a was used, student licences provided by the TU Delft.
• Python: use is made of Anaconda 3, a scientific Python distribution that comes with Python 3.6 and
most commonly used scientific packages. In this project, especially SciPy and NumPy are used.
Aerodynamic Toolbox
The following external programs and libraries and programs have been used by the aerodynamics team. Spe-
cial thanks to our coach Yi Zhang, and his colleague Yu Zhang of the aerodynamic department that helped us
find AVL and discover PARSEC and sent us some examples that were good starting point to understand how
to use them.
• AVL: AVL is a program for the aerodynamic and flight-dynamic analysis of rigid aircraft of arbitrary
configuration using vortex-lattice theory. It is an open sourced project created by MIT, and can be found
on http://web.mit.edu/drela/Public/web/avl/
A good guide to start with AVL written by Kinga Budziak of Hamburg University of Applied sciences can
be found on http://www.fzt.haw-hamburg.de/pers/Scholz/arbeiten/TextBudziak.pdf
• Parsec-airfoils: This Python library generates and plots the contour of an airfoil using the PARSEC
parameterization, it was created Dimitrios Kiousis. It is open source and free to use for non commercial
purposes and can be found on https://github.com/dqsis/parsec-airfoils
115
A
V&V Procedures
During a design process, verification and validation of used methods, code and results is vital to make sure
that the process produces the correct results. In this chapter, the verification and validation procedures used
during this design process will be explained. In Section A.2 the verification aspects will be discussed whereas
Section A.3 will focus on the validation aspects. First, in Section A.1 a short introduction to V&V methods will
be given.
A.1. Introduction
Verification and validation are used to monitor the design process and make sure all computations provide
the desired results. It is also a useful tool for other users to assess the quality of a method. Lately, V&V proce-
dures get more and more recognition.
Verification answers the question: Are you solving it right?. It mainly deals with mathematical issues and
proves that a model is solved correctly. It checks whether a product, a service or a systems meets a set of
design requirements. Validation answers the question: Are you solving the right thing? and deals with entire
process. It provides evidence that the correct model for the physical process is solved. The simulation of a
physical process can be visualised in Figure A.1. This figure also shows the V&V steps and where they impact
the simulation process.
Test A test type of verification which requires instrumentation, for example, wind tunnel test or flight test.
Testing is essential in the certification process. Testing assures that every aircraft element performs the func-
tion it was intended to perform, performs it to the expected level of performance and does not perform func-
tions it was not intended to perform [46]. Since there is no instrumentation available, this verification method
116
A.3. Validation 117
Analysis An analysis is any kind of mathematical, computational or logical task performed to verify a re-
quirement that cannot be verified in any other manner or other verification methods are too expensive. Anal-
ysis can be also used in conjunction with testing to extend the envelope of the test results. For example,
computer simulation can be used to predict aircraft behaviour in regimes beyond flight data points. Analysis
verification includes simulation, in-service data and similarity and system safety assessment [46].
Demonstration Demonstrations are similar to tests, but contrary to test, do not require instrumentation.
However, since there will be no real-life model present, Demonstration method will not be used.
Examination Examinations are easiest types of verification, they are simply a visual confirmation that a re-
quirement has been met, for example, hardware inspection to check if the required hardware component is
installed in the aircraft [46]. This method might be used to verify some of the requirements using the techni-
cal drawings.
For the code verification, the analysis method is chosen. Below the verification procedures for codes are de-
scribed.
Debugging First step in code verification is debugging. In order to ensure a code works, all the syntax errors
due to typos, missing semicolons or others syntax errors have to be identified and fixed. Once all the syntax
errors are gone, the code should compile and run, however, it still might not give the correct results.
Unit tests Next step in verification procedures is unit tests. The code is divided into smaller functions
(chunks) and run each function separately. To ensure the chunks are working correctly different input val-
ues are tested or different limitations such as boundary conditions, to check whether different output values
are obtained. Then the hand calculations for each function are performed. Finally the program outputs are
tested against hand calculation outputs.
System test Once all the functions are verified, they are, again, connected into one code for the system test.
The whole system is tested for different input values, to check if different output values are produced, but
also to check if all the units interact with each other in desired manner.
A.3. Validation
Validation is testing program outputs against the real values obtained through experiment or measurements.
If the model predicts the experimental outcomes within the predetermined accuracy requirements, the model
is considered validated for its intended use [28]. If the model is not matching the real-life data, one has to
re-consider assumptions used. Once the program gives anticipated results, the discrepancies need to be dis-
cussed and explained. If the outputs cannot predict the real-life data the program cannot be validated, oth-
erwise, the program is validated.
Since for this report no real-life data is attainable for the aircraft nor the experiment can be conducted, the
group decided to validate the outcome parameters against statistical data obtained from different aircraft.
For example, plot MTOW and OEW for similar aircraft, and the designed aircraft thus see if designed aircraft
fits in the regression line well. If it does, it can be said that the MTOW and OEW combination was validated.
B
Renders and Technical Drawings
118
119
A
4
4
3
3
Isometric view
2700 Scale: 1:200
29811.16
6367.73
Top view
Scale: 1:200
2
2
28500
6370
Right view
Scale: 1:200
20-6-2018
CHECKED BY DATE SIZE DRAWING NUMBER REV
Front view XXX
Scale: 1:200
XXX A3 MW5 - 001 X
DESIGNED BY DATE
DSE Group 22 SCALE 1:200 WEIGHT(kg) XXX SHEET 1/5
XXX
H G B A
H
G
F
E
D
C
B
250 A
400
4
510 1100 4
1850
2700
920
Detail A
3
Scale: 1:50 3
Detail D Top view
Scale: 1:40 Scale: 1:400
6370 4357.49
D
2
29811.16 Rear view 2
Scale: 1:400
Right view
Scale: 1:400
2000
16525
1491.6
3800
DRAWING TITLE
C DRAWN BY DATE
1
Willem 20-6-2018 Detailed three view 1
3000 CHECKED BY DATE SIZE DRAWING NUMBER REV
Bottom view
Detail C XXX XXX A3 MW5 - 002 X
Scale: 1:400
Scale: 1:100 DESIGNED BY DATE
DSE Group 22 SCALE 1:200 WEIGHT(kg) XXX SHEET 2/5
XXX
H G B A
C
Aircraft Data
121
122 C. Aircraft Data
Perform
pre-flight
operations
Perform Perform
passenger & cargo normal
Perform operations operations
Take-off
operations Perform Perform
×
Freighter × abnormal
(D.1)
Perform operations operations
flight
operations Perform Perform
non-revenue emergency
Perform operations operations
post-landing
operations
With the first matrix being operational phase functions, second being mission functions and last one being
situations functions. The functional flow diagram is presented in Figures D.1 through D.3. It was worked out
for operational phase functions with passenger and cargo operations under normal conditions as this is most
popular profile for the passenger typical aircraft. However, one needs to remember that some other, perhaps
critical, function can be present in other configuration. The functional flow diagram is worked out up to 3rd
level of detail. Further levels of detail are possible but not necessary at this stage of the design.
TOP level
Perform abnormal Perform
Perform pre-flight Perform take-off Perform flight Perform landing Perform post flight
operation maintenance
1.0 2.0 3.0 4.0 6.0
5.0 7.0
Second Level
TOP level 123
Facilitate
Determine Perform visual Perform pre-flight
Perform pre-flight passenger Load aircraft
required fuel inspections checks
1.0 embarking 1.5
1.1 1.2 1.3
1.4
Second Level
TOP level
Facilitate
Determine Perform visual Perform pre-flight
Perform pre-flight passenger Load aircraft
required fuel inspections checks
1.0 embarking 1.5
1.1 1.2 1.3
1.4
Perform with
Perform abnormal Handle Perform Perform with
control system
operations emergencies redirection engine failure
failure
5.0 5.1 5.2 5.3
5.4
Perform
Perform Perform interior Perform exterior Perform structural Perform electrical Perform engine Perform sensor
mechanical
maintenance inspections inspections maintenance maintenance maintenance maintenance
maintenance
7.0 7.1 7.2 7.3 7.5 7.6 7.7
7.4
Functional
Perform pre-flight
breakdown
Perform pre-start
diagram
Perform start up
(FBD) represents all the functions that the aircraft needs to perform. The
Perform before
checks checklist checklist taxi checklist
functional 1.3 breakdown1.3.1 does not
1.3.2 indicate 1.3.3 at what time the functions need to performed or in what sequence.
It is presented
Facilitate
passenger
in Figure
Allow for D.4. The
Allow for breakdown
Instruct cabin is performed up to 4th level of detail. Further levels of detail are
gate/stairs passenger doors crew
embarking
possible 1.4 but not necessary
1.4.1 at
1.4.2 this stage 1.4.3of the design.
Load passenger
Load aircraft Load cargo Load fuel
facilities
1.5 1.5.1 1.5.2
1.5.3
Allow Facilitate
Provide steering
groundcontrol pushback
2.2.2
2.2 2.2.1
Taxi to runway
2.3
Deploy parking
Configure aircraft Deploy HLDs
brakes
2.4 2.4.2
2.4.1
Perform
3.4 - Perform Allow pitch control Allow roll control Allow yaw control
maneuvers
maneuvers 3.4.1 3.4.2 3.4.3
3.4
Meteorological
Avoid hazards Terrestrial hazards
hazards
3.5 3.5.1
3.5.2
Monitor airborne
Monitor health
health
3.6
3.6.1
Monitor ground
health
3.6.2
Provide electrical
Provide power
power
3.7
3.7.1
Provide hydraulics
power
3.7.2
D.2. Functional Breakdown 125
Change
Retract
4.7 - Change
HLDs
configuration
Start spinning Ignite main configuration
4.7.1
Start engines Start APU 4.7
2.1 - Start engines 2.1 - Start engines
engines 2.1 - Start engines
engines
2.1 2.1.1
2.1.2 2.1.3
Engage thrust Engage disc
Perform braking 4.8 - Perform Engage
4.8.2 air brakes
- Engage air 4.8.3 - Engage disc Engage
4.8.4 spoilers
- Engage
reversers brakes 4.8.2 brakes
brakes spoilers 4.8.4
Allow ground Facilitate 4.8 braking
Provide steering 4.8.1 4.8.3
control pushback
2.2.2
2.2 2.2.1
Taxi
4.9
Taxi to runway
2.3
Deploy parking Handle Cope with internal Cope with Allow for medical
Configure aircraft Deploy HLDs 5.1 - Handle 5.1 - Handle 5.1 - Handle
brakes emergencies fires external fires aid
2.4 2.4.2 emergencies emergencies emergencies
2.4.1 5.1 5.1.1 5.1.2 5.1.3
Communicate Request take-off Confirm free air Perform Allow second Allow second Allow second
Allow loiter
with tower clearance space redirection climb cruise descent
5.2.3
2.5 2.5.1 2.5.2 5.2 5.2.1 5.2.2 5.2.4
1.0 - Perform pre-flight 2.0 - Perform take-off 3.0 - Perform cruise 4.0 - Perform landing 5.0 - Perform abnormal operation 6.0 - Perform post flight operations 7.0 - Perform maintenance
127
128 E. Compliance Matrix
Sys.Perf.15 The amount of emergency Verified Chapter 9 No idea what section this is
exits present shall be suffi- described in
cient to comply with CS-25
regulations.
Sys.Perf.16 The maximum take-off Deffered To be verified after the air-
mass of the aircraft shall be craft is built
less than 35,000 kg.
Sys.Perf.17 The operational empty Deffered To be verified after the air-
mass of the aircraft shall be craft is built
less than 20,600 kg.
Sys.Perf.18 The aircraft shall be able to Verified Subsection 4.3.6
fly to an alternate airport
100 km away after maxi-
mum range flown.
Sys.Perf.19 The aircraft shall be able to Verified Subsection 4.3.6
loiter at the alternate for 30
minutes.
Sys.Perf.20 The landing distance on Deleted Deleted due to the removal
tarmac shall be less than of Cus.3
1200 m.
Sys.Perf.21 The landing distance on Verified Subsection 4.3.5
tarmac shall be less than
1500 m.
Sys.O&L.1: Turnaround time of the Verified Chapter 9 Gantt Chart
aircraft between maximum
payload flights shall be less
than 30 minutes.
Sys.O&L.2 The aircraft shall have a Deferred No fatigue analysis per-
lifetime of at least 25 years. formed on the main struc-
tural elements of the air-
craft
Sys.O&L.3 The aircraft shall be able to Verified Chapter 11 When only long range
perform three to four daily flights are performed
operations. (2500km) the aircraft is
able to perform 3 daily
operations, when the range
is shorter 4 is also possible
Sys.O&L.4 Lateral ground clearance of Verified Section 10.2
the aircraft shall be more
than 5 degrees.
Sys.O&L.5 The seat pitch shall be Verified Section 3.4 Verified
equal to 31 inches.
Sys.O&L.6 Maintenance intervals shall Deferred No fatigue analysis per-
be increased due to safe- formed on the structure
life design of components.
Sys.O&L.7 The aircraft shall not tip Verified Section 10.2
over backwards for the en-
tire range of center of grav-
ity locations during ground
loading.
Sys.O&L.8 The aircraft shall not turn Verified Section 10.2 Based on the main gear
over for any possible load- track requirement
ing configuration and turn-
ing radius.
Sys.O&L.9 The aircraft doors shall be Verified Chapter 9
compatible with airstairs.
130 E. Compliance Matrix
(a) Concept 1. (b) Concept 2. (c) Concept 3. (d) Concept 4. (e) Concept 5.
(f ) Concept 6. (g) Concept 7. (h) Concept 8. (i) Concept 9. (j) Concept 10.
(k) Concept 11. (l) Concept 12. (m) Concept 13. (n) Concept 14. (o) Concept 15.
(p) Concept 16. (q) Concept 17. (r) Concept 18. (s) Concept 19. (t) Concept 20.
(u) Concept 21. (v) Concept 22. (w) Concept 23. (x) Concept 24. (y) Concept 25.
132