Transcontinental Windowless Business Jet - Aerospace Vehicle Design

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 63

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/317090545

Transcontinental Windowless Business Jet - Aerospace Vehicle Design

Technical Report · March 2015


DOI: 10.13140/RG.2.2.23829.52960

CITATION READS

1 3,466

4 authors, including:

Sergiu Petre Iliev


Imperial College London
35 PUBLICATIONS 23 CITATIONS

SEE PROFILE

All content following this page was uploaded by Sergiu Petre Iliev on 24 May 2017.

The user has requested enhancement of the downloaded file.


Imperial College London

Aeronautical Engineering
AE3-403 Aerospace Vehicle Design

Aerospace Vehicle Design Final Report

Authors:
Joseph Dudley Supervisors:
Alix Frossard Dr Rob Hewson
Sergiu Iliev Dr Errikos Levis
Timothy Minshall

March 26th , 2015

Abstract
The following paper constitutes the final report on the design of a twin turbofan transcontinental
business jet. The cabin layout has been chosen and the fuselage dimensions deduced. Furthermore,
an optimum aerofoil design was identified and the dimensions of the wing and those of the high
lift devices were calculated. The incorporation of novel ideas such as a windowless CFRP fuselage
and fuselage-mounted engines are analyzed further in this paper. The selection and integration
of propulsive systems as well as their impact on aircraft drag and weight are considered. Control
surfaces have also been designed to be fit for safety and appropriate maneuverability. Finally, an
aerodynamic analysis as well as performance and cost preliminary evaluation were carried out. Both
the validity of the assumptions made and their impact on preliminary results were identified.
Contents
1 Design Summary 4

2 Wing Design 6
2.1 Reference Wing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2 Wing Sweep . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.3 Washout . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.4 Dihedral Wing Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.5 Wing Incidence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.6 Wing Tips . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

3 High Lift Device Selection & Sizing 9


3.1 3D Wing Aerodynamic coefficients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

4 Powerplant Selection & Integration 10


4.1 Thrust requirements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
4.2 Engine selection and sizing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
4.3 Augmented turbofan consideration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
4.4 Nacelle and intake design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
4.5 Engines pitch and cant angle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
4.6 Engine Placement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
4.6.1 Lateral and vertical position with respect to the fuselage . . . . . . . . . . . . . . 14
4.6.2 Longitudinal position along the fuselage . . . . . . . . . . . . . . . . . . . . . . . . 15
4.7 Installed engine parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
4.8 Engine drag estimation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
4.9 Nozzle design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
4.9.1 Noise and vibration reduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
4.9.2 Reverse thrust systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

5 Fuselage Design 19
5.1 Nose-Cockpit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
5.2 After Body . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
5.3 Aerodynamic Considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

6 Tailplane Sizing & Design 21


6.1 Tail Sizing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
6.2 Possible Improvements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

7 Aircraft System Layout Design 23


7.1 Avionics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
7.2 Emergency Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
7.3 Fuel Tanks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
7.4 Electronics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
7.5 Hydraulics & Actuators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
7.6 Environmental Control Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
7.7 Water & Waste . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
7.8 De-icing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

8 Undercarriage Design 25
8.1 Landing Gear . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

9 Weight & Balance Predictions - following page 25


9.1 Weight Reduction Concepts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
9.1.1 Advanced Composites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
9.1.2 Windowless CFRP Fuselage Design . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

1
10 Aerodynamic Analysis 29
10.1 Fuselage at Crusie . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
10.2 Wings at Cruise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
10.3 Wings at Landing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
10.4 Fuselage at Landing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
10.5 Wings at Take-off . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
10.6 Fuselage at Take-off . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
10.7 Horizontal Tail Plane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
10.8 Vertical Tail Plane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
10.9 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

11 Static Stability & Trim 34


11.1 Lateral Stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
11.2 Tail Lift-Curve Slope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
11.3 Downwash & Fuselage Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
11.4 Longitudinal Stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
11.5 Trim Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

12 Control Surface Design 37


12.0.1 Elevator & Rudder Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
12.0.2 Aileron Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

13 Performance Estimation 37
13.1 Take-off performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
13.2 Mission Performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

14 Cost estimations 46

15 Appendix 50
15.1 Fuselage Calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
15.2 Aerodynamic Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
15.2.1 Tail Plane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
15.3 Performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

2
List of Figures
1 SC(2)-0714 aerofoil geometry. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2 Leading edge sweep angle. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3 High lift device data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
4 Features of the engine that will be used, the Rolls Royce Ultrafan, successor of the BR725. 12
5 Preliminary capture area sizing [1]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
6 Nacelle weight fraction of empty weight [41]. . . . . . . . . . . . . . . . . . . . . . . . . . . 13
7 Relation between interference drag and distance from fuselage [1]. . . . . . . . . . . . . . . 15
8 TRL progression of NASA’s chevron nozzle design [42]. . . . . . . . . . . . . . . . . . . . 17
9 Figure shows typical relation between work done and landing run with respect to use of
thrust reverser in icy/wet runways, and it is seen that landing run is reduced about 2.5
km. [44]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
10 Left - engine and clam reverser diagram of main components [45], right - reverse thrust
in nominal position on a RR725 engine [46]. Both are an accurate representation of the
engine to be used. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
11 The fuselage layout. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
12 Graphs used to the after body design and fuselage aerodynamic considerations. . . . . . . 20
13 A T tail configuration from [18]. Note that this figure is to illustrate the T tail configura-
tion being used and not the number or position of the engines. . . . . . . . . . . . . . . . 21
14 The dimensions used for the tail calculations. . . . . . . . . . . . . . . . . . . . . . . . . . 21
15 IXION windowless concept interior. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
16 Details of the aerodynamic analysis calculations. . . . . . . . . . . . . . . . . . . . . . . . 33
17 Longitudinal static stability [18]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
18 NACA TR-711 The relationship between Kf and wing position from [20]. . . . . . . . . . 35
19 Historical data for aileron sizing from [19]. . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

3
1 Design Summary
To provide an outline of the preliminary report and a summary of the design parameters established
in this report, the table below and the 3-view drawing on the following page represent the baseline
configuration of our aircraft concept.
• Mission Profile
– T/W crit.=0.35 - from preliminary report constrains diagram.
– W/S crit.=386.15 kg/m2 -from preliminary report constraints diagram.
• Wings
– Sweep angle at quarter chord: Λc/4 = 24.76◦ .
– Span: b=35.72m.
– Reference area: S=150.15m2 .
– Double slotted fowler flaps and slats.
• Fuselage
– Type: windowless.
– Material: CFRP + GLARE radome.
– Cabin length 40.11m; max width 2.94m; max height 1.93m.
– Fuselage: D=3m, L=51.74m (6.375m cockpit and 5.25 afterbody).
• Powerplant

– Type: high bypass turbofan BPR=4.6.


– Fuselage-podded-mount, intake located 38.8m from nose.
– Installed/Uninstalled weight: 2×3234/4340Kg.
– Max installed: 117.1kN (26324.19lbf).
• Weight

– We , empty weight: 29.26 tonnes.


– W0 , MTOW: 58 tonnes.
• Undercarriage

– Type: breadth.

4
2 Wing Design
As a first approximation, the wing lift coefficient is assumed to be equal to the aerofoil lift coefficient
and if the aircraft is considered at equilibrium: CL = ( 1q )( W S ). The design lift coefficient is the lift
coefficient at which the aerofoil has the best lift to drag ratio. The main part of the mission being
the cruise, the design lift coefficient is calculated in mid-cruise configuration. At 10000m and M=0.79,
q=11590.8Kgm−2 s−1 and the wing loading ratio was calculated using Raymer’s formula given by equation
2.1. This gave ( W
S )cruise = 3788N m
−2
(with A=8.5, fixed in the first stage of the design). These values
lead to CL = 0.3268. In reality, due to 3D effects, the wing lift coefficient will probably be lower than
the aerofoil; thus it has been decided to look for an aerofoil with a design CL a bit higher than 0.33 to
ensure that the design CL will be around 0.33 in cruise.
  r
W πAeCD0
=q (2.1)
S cruise 3

During the first phase of the design, it had been decided to choose a supercritical aerofoil, designed
for a transonic speed, which implies, among other benefits, a laminar bucket on the drag polar. Following
these ideas, the SC(2)-0714 aerofoil was chosen, whose geometry is shown figure 1 and main properties
are given in the table 1.

Figure 1: SC(2)-0714 aerofoil geometry.

Some time was spent trying to find experimental data on this aerofoil, however, facing the lack of
significant results, the website airfoiltools.com giving Xfoil results was used. Considering how approx-
imate Xfoil results might be, it was decided to keep only the lift-related values and the aerofoil global
behavior. Some tests were made, using the aerofoil coordinates to calculate polar on Xfoil for Re=3×107
and M=0.79, however they were not converging, which is why the results for Re=1×106 have been used,
even though this will imply minor errors.

SC(2)-0714 Aerofoil for Re=106


Zero lift angle of attack (AoA) α0 = −4.8◦ (CL = 0)
Lift coefficient at zero (AoA) CLα0 = 0.574 (α = 0◦ )
Maximum aerofoil lift coefficient CLmax = 1.76 at α = 15.75◦
Maximum lift to drag ratio Max CD = 83.67 at α = −0.5◦
CL

Aerofoil lift curve slope CLα = 6.48

Table 1: SC(2)-0714 Aerofoil characteristics.

A laminar bucket is effectively visible for −5◦ < α < 0◦ which is coherent with the maximum of
CL /CD obtained for α = −0.5◦ where CL = 0.5288 (design lift coefficient). The α value could be an
indication for the incidence of the wing compared to the fuselage.
Typically, the optimal thickness to chord ratio is around 12%: such value has the higher CL,max
with a drag coefficient and a critical Mach number values (CD and MC ) acceptable. It must be noticed
that supercritical aerofoil are usually 1.1 times thicker than standard aerofoil. The SC(2)-0714 has a
maximum thickness to chord ratio of (t/c)max = 14%, a value coherent with the historical trend line for
a maximum design Mach number of 0.82 and it should give acceptable results. Indeed thick aerofoil tend
to separate first at the trailing edge with a point of separation moving forward with an increasing angle
of attack and such behavior is desired from handling quality point of view.

6
2.1 Reference Wing
Given the wing loading W/S, the aspect ratio A and the taper ratio λ, most of the reference wing
parameters may be calculated.
The aspect ratio was fixed on the first stage of this design: A=8.5. The taper ratio, which aim is to
get a lift distribution as close to the elliptical one as possible, has been chosen as λ = 0.25, because most
swept wings have a taper ratio between 0.2 and 0.3.
Considering that the constraint diagrams from the interim report were not complete (there was an
issue with the Take-off that was not resolved) hence not fully satisfactory; a W/S ratio was calculated
for cruise (and is given earlier): ( W S )cruise = 3788N/m or ( S )cruise = 386.15Kg/m . Such value is
2 W 2

coherent with the acceptable parts of the constraint diagrams and was then the one used to design the
wings.
The first parameter to be calculated is the wing area (S). It should be calculated using the take-off
gross weight (WMTOW = 57.98t) and the wing loading W/S calculated from the constraint diagram.
Thus S=W/(W/S) and S=150.15m
Such value seems high (it was estimated at the first stage that 100m2 < S < 150m2 ) but this is
due to the really high maximum take-off weight that resulted from the interim report and which was
conserved as suggested. √
Knowing S, the wing span b can then be found by: b = AS = 35.72m. Then the chord at the root
and at the tip can be calculated: Croot = B(1+λ) 2S
= 6.72m, and Ctip = 1.68m.The mean aerodynamic
chord c, which is required to find the aerodynamic center of the wing, is given as c̄ = 4.71m by equation
2.2 and y = 7.14m by equation 2.3.
2 (1 + λ + λ2 )
c̄ = Croot (2.2)
3 (1 + λ)
 
b (1 + 2λ)
y= (2.3)
6 (1 + λ)
The real Reynolds Number at cruise condition can then be calculated, using the mean aerodynamic
chord as length, Mach=0.79 and air properties at 10000m altitude. This is shown below.
ρV L 0.414 × 299 × 0.79 × 451
Re = = = 2.97 × 109
µ 1.4814 × 10−5

2.2 Wing Sweep


It has been decided to have a rear swept wing, since sweep can increase the critical Mach number hence
reduce adverse transonic effects and forward sweep produce aero elasticity issues. The Leading Edge
Sweep angle was chosen in accordance with the historical trend line on figure 2. Considering a maximum
Mach number of 0.82, the leading edge sweep angle was chosen as ΛLE = 28◦ . This leads to a sweep at
the quarter chord of Λc/4 = 25◦ . This was found using equation 2.4. Thus for a clean wing the maximum
lift coefficient of the wing will be: CL,max = 0.9CL,max cos(Λc/4 ) which leads to CL,max = 1.44.
1−λ
tan(ΛLE ) = tan(Λc/4 ) + (2.4)
A(1 + λ)

2.3 Washout
Washout is used, to prevent tip stall and, like taper ratio, to revise the lift distribution to an approximate
ellipse. It has been chosen to apply a −3◦ of twist, which is the value commonly used to provide reasonable
stall characteristics.

2.4 Dihedral Wing Design


It was decided on the first phase to have a dihedral wing design to increases the lateral stability. For a
low subsonic swept wing, the dihedral angle is usually between 3◦ and 7◦ . A medium value of δ = 5◦ has
been chosen. Indeed, having the engines attached to the fuselage and not under the wings, there is no
need for a high angle which would leave more space to the engine. Also 5◦ is a value in common with
other low wings, which means it should give reasonable results.

7
Figure 2: Leading edge sweep angle.

2.5 Wing Incidence


The incidence, which is the angle of the wing with respect to the fuselage, should be selected to minimize
drag. During cruise flight, to minimize drag, the aircraft must have an angle of attack as low as possible
(usually 1◦ or 2◦ ) while ensuring the wings provide enough lift. The aerofoil best lift to drag ratio is
obtained for an angle of attack α = −0.5◦ for Re = 106 . The incidence should then be chosen so that
in cruise, the angle of attack of the wing is about −0.5◦ . Thus the incidence could be of −1.5◦ , which
would give an aircraft angle of attack of 1◦ in cruise. However it must be highlighted that due to the
presence of twist and dihedral, just fixing the incidence this way does not ensure that every profile along
the wing will have an angle of attack of −0.5◦ . It is a compromise choice.

2.6 Wing Tips


Wing tip design is aimed to reduce the induced drag. The harder it is for the higher pressure air to
‘escape’ toward the upper part of the wing, the lower the induced drag. In the interim report, winglets
had been considered, however, winglets are more often used as add-ons to improve the aerodynamic
properties of an existing wing (usually a low aspect ratio wing while ours is of high aspect ratio) and
even though they are efficient they are detrimental to flutter property. Thus since the wing is being
created, another design was chosen, with upswept tips. Indeed it will act similarly to what a winglet
would do: increase the effective span and pressure resultant creating a forward force called ‘negative
drag’, adding less weight that what a winglet would do.

8
3 High Lift Device Selection & Sizing
A maximum lift coefficient (all flaps extended) of CL,max = 0.28 is required. According to the graph
shown in figure 3, with a 1/4chord sweep angle of 25◦ , to achieve CL,max = 0.28, a double slotted flap
and slat are needed. Double slotted flaps induce at landing a change in the 2D maximum lift coefficient
∆CL,max = 1.6c0F /c where slats induce ∆CL,max = 0.4c0s /c. Usually, Flap chord ratio c0F /c are around

Figure 3: High lift device data.

110% to 120% and the slat chord ratio c0s /c around 102% to 105%. For each device, the effect on the
wing lift coefficient is given by equation 3.1 where Sflapped is the area of the wing having flap/slat and
ΛHL is the sweep angle of the hinge line (corresponding to the leading edge of the flap/slat).
 
Stextf lapped
∆CL,max = 0.9∆CL,max cos ∆HL (3.1)
Sref

From the slat point of view, the assumption was made that Sslatted = 90% of Sref as Slats are usually
all along the span apart from the tips and the wing box. This gives Sslatted = 135m2 ; In order to get a
general idea of such result, and if the mean aerodynamic chord is considered, the Slat length might be
calculated as length = Sflapped /c = 28m for a 35.7m span. The Hinge Line for the slats is the Leading
Edge line, so ΛHL = ΛLE = 28◦ . With a chord ratio c0s /c = 2%, the benefit for the wing lift coefficient
is ∆CL,max slat = 0.29.
The aim of the high lift devices here is to achieve CL,max = 2.8. Since the clean wing coefficient is
1.44, the total benefit from the slats and flaps must be of ∆CL,max = 1.36. Thus the benefit from the
flap must be ∆CL,max flap = 1.07. This time, the reasoning is reversed: the unknown is the Sflapped /Sref .
The hinter line angle may be calculated using equation 3.2, where n is the chord fraction. The area ratio
is then given by 3.3.    
4n 1−λ
ΛHL = tan−1 tan ΛLE − (3.2)
A 1+λ
Sflapped ∆CL,max
= (3.3)
Sref 0.9∆CL,max cos ΛHL
The calculation was tried for c0F /c = 110%, it gives an area ratio of: Sflapped /Sref = 0.7 and Sflapped =
105m2 . Similarly, to get a general idea the Flap length would be around 22m for a 35m span. With
c0F /c = 120%, Sflapped would be 96m2 which gives a length of about 20.5m. Given the significant
difference (1.5m), it was decided to keep c0F /c = 120%.

9
3.1 3D Wing Aerodynamic coefficients
The Lift curve slope of the 3D clean wing will differ from that of the 2D aerofoil. In order to calculate
it, a semi-empirical equation can be used. However, according to [1, p. 265] it is ‘accurate up to the drag
divergent Mach number and reasonably accurate almost to M=1 for a swept wing’ the result in this case
should then be reasonably accurate. The Wing lift curve slope is given by equation 3.4 where β describes
the compressibility effects given by equation 3.5 and η is the aerofoil efficiency factor which is assumed
to be equal to 0.95 (Raymer’s value since CLα at M=0.79 is unknown). Sexposed /Sref is the exposed wing
area correction which is given by equation 3.6, where the part of the wing within the fuselage is assumed
to be rectangular.
2πA Sexposed
(CLα )clean = q F (3.4)
2+ 4+( A2 β 2
η2 )(1 + tan2 Λmax t
β2 ) Sref

β 2 = 1 − M 2 = 0.3759 (3.5)
Sexposed Sref − dfuselage croot
= = 0.866 (3.6)
Sref Sref
Λmax t is the sweep angle at the chord location where the aerofoil is thickest. Since t/cmax is is obtained
for x/c=0.37, Λmax t = 23.1◦ .
Aspect ratio is assumed to be the same as decided earlier (A=8.5), since the wing has a high aspect
ratio, the wing tips chosen will not affect the effective aspect ratio.
The wing lift curve slope is CLα = 6.474 slightly lower than CLα . The lift curve slope of the wings
will be affected by the flaps extension and will change according to equation 3.7.
  0  
c Sflapped
(CLα )flaps extended = (CLα )clean 1 + −1 (3.7)
c Sref

(CLα )flaps extended = 7.12 from figure 3 which is higher than the clean wing value, which is coherent since
the aim of slotted fowler flaps is to increase CL,max by increasing the slope. The change in wing zero lift
angle can also be estimated using equation 3.8.
Sflapped
(∆α0 )wing = (∆α0 )aerofoil cos ΛHL = −3.10◦ (3.8)
Sref

4 Powerplant Selection & Integration


This chapter presents the entire process of power plant selection and integration. The procedures used
are described in the lecture notes [22] and by Raymer [1] will be used throughout. This chapter is meant
to be a continuation of that presented in the preliminary report.

4.1 Thrust requirements


Using the constraint diagram, the maximum ratio of thrust loading was determined as T/W=0.35, the
most critical phase of flight being landing at London City (the flaps were considered to be deployed
increasing Clmax to 2.8) and without reverse thrust as per FAR 25 regulations. The MTOW was
recalculated to W=59.7 tonnes, using the weight reduction considerations presented in the previous
chapter, bearing in mind that a lower aircraft empty weight gives a lower required fuel weight. Knowing
T/W and W it is possible to obtain a required thrust of 117.1 kN per engine. Please note that this is
considered an ‘installed thrust’, after all the effects of losses are quantified, the functional engine should
output at least this amount of thrust in order to take off safely. Therefore, the ‘uninstalled thrust’ will
have a higher value that is to be determined.

4.2 Engine selection and sizing


The design specifications [23] restricts the engine type to ‘twin turbofan’. Therefore, a high bypass
ratio turbofan engine was chosen, starting from historical data on planes with similar applications and
requirements [29]. The selection was narrowed down to the Airbus A318 Series and the Gulfstream

10
G650. Although the A318 is designed for short to medium hauls, it carries up to 132 economy class seats
giving a good estimate for the upper bound of the dimensions and thrust requirements of our design.
The Gulfstream provides the lower bound of this interval. It is designed to carry up to 19 business
class passenger on medium-long flights. Following the engine analysis of these aircraft, it was found
that the G650 employs the use of two Rolls-Royce BR725 turbofan engines [32]. The A318 uses two
Pratt & Whitney PW6000 engines [31]. The engine parameters are provided in the table below and
are used in the ‘rubber engine’ model, to get a better estimate for the parameters (length, L; diameter,
D; weight, W) associated with an engine capable of delivering the thrust requirements. This approach
yielded better results than the empirical ‘fitted data’ model, these final values are presented in below
and full calculations are included in the Annex.

L = Lactual (SF )0.4 ; D = Dactual (SF )0.5 ; W = Wactual (SF )1.1 (4.1)

Engine Model Pratt & Whitney PW6000 Rolls-Royce BR725 Desired Engine
Maximum thrust output TT O ,SL (kN) 96 71.6 117.1
Bypass ratio, BPR 4.9 4.2 4.6
Engine Mass, We (kg) 2289 2229 3,329
Engine Length, Le (m) 2.74 5 4.53
Fan Diameter, Df (m) 1.44 1.27 1.61
SF Ccruise (lb/lbf hr) 0.63 0.52 0.55

Table 2: Parameters corresponding to current candidates and desired engine.

In order to obtain the maximum required thrust output from the desired engine, in addition to the
procedure using constraint diagrams, detailed in the preliminary report, the empirical equations below
were used to determine values for the thrust loading T/W under various flight situations: minimum
allowable climb requirements after take-off, thrust matching in cruising conditions and with weight at
the beginning of the cruise stage, respectively, maximum A/C required speed [4].

(4.2)
p
(T /W )F AR25 = Gmin + 2 CD0 /(πARe)

Where Gmin is the minimum climb slope given by FAR25 for a set segment, CD0 is the zero lift drag,
AR is the wing aspect ratio and e is the Oswald efficiency.

(T /W )Cruise = (D/L)Cruise (4.3)


c
(T /W )T.O. = AMmax (4.4)
Obtained from fitted data, where A=0.267, c=0.3.63, Mmax =0.79 as per design specifications.
All the above equations yielded a required thrust lower than that given by the constraint diagrams.
Therefore the latter was chosen as the conservative design scenario. Given the potential timescale of this
concept plane, it is reasonable to assume it will employ a future generation of the engines detailed above.
The new generation of the BR725 will likely borrow technology from the bigger ‘sister engine’, the future
Rolls Royce UltraFan. This is scheduled to enter production in 2020 [37]. It will have a very high bypass
ratio, whilst keeping a similar fan diameter. This allows for a smaller flow velocity difference between
the core and bypass flow, increasing efficiency. As it will be seen in the engine placement subsection, the
aircraft concept is able to accommodate larger engines, as ground-clearance is not an issue.
The next engine generation will feature “lightweight, high-efficiency compressors, turbines and a CTi
fan; advanced high-overall pressure ratio cycle; adaptive cooling; hybrid ceramic bearings and high
torque-density shafts. The combustor could provide a significant reduction in NOx with a margin that
anticipates future legislation” [38]. One of the main advantages is a power-reduction gearbox to drive
the low-speed fan. This allows for a much higher bypass ratio and significant improvements in SFC of
20-25%, decreasing the current estimate of SF Ccruise to 0.41.

11
Figure 4: Features of the engine that will be used, the Rolls Royce Ultrafan, successor of the BR725.

The Ultrafan thrust class will be perfectly suitable for the concept plane, as it can be designed to
produce thrust in the range of 111 kN to 200 kN perfect for these requirements [37].

4.3 Augmented turbofan consideration


Since the highest thrust requirement occurs in a limited flight regime, one could consider the use of an
augmented turbofan that uses an afterburner system to increase its thrust for that short period of time.
The engine operating at a lower nominal thrust that is suitable all of the other stages of flight. It can also
provide assistance in emergency situations such as taking off with one engine (as per FAR25). However,
it was found that in practice the weight of the fuel system as well as the considerable increase in engine
length, makes this idea impractical.

4.4 Nacelle and intake design


Engine nacelles play several important roles in the performance of an engine, mainly their purpose is to
keep as much of the uninstalled thrust as possible through a broad range of flight conditions. As can be
observed in Figure X (the one at the end of this subsection), long nacelles have been chosen, this is because
they provide better thrust at a lower noise (through the mixing of the outer and core flow). Our team
considers the extra weight these longer nacelles imply is well justified by the improved fuel consumption
and noise levels [40]. To size the nacelle, according to Raymer, it is reasonable to assume the nacelle
diameter will be 10% larger than that of the fan (Dnacelle = 1.1 ∗ f andiameter = 1.1 ∗ 1.61 = 1.77m)
and it’s length will be approximately Lnacelle = 4.53 + 1.5 = 6.03m[1]. Where 1.5m is the approximate
length of the thrust reverser presented later in this chapter.
Concerning the inlet geometry, due to the subsonic cruise Mach number of 0.79, a Pitot-type inlet
would be optimum. This should decelerate the flow to a M=0.5 in order to assure efficient compression
[22]. The capture area of the engine dictates the radius of the inlet. Considering the front face diameter
as 80/5 of the maximum engine diameter (that of the fan) [1] De = 1.61 ∗ 0.8 = 1.29m and the cruise
Mach number of 0.79, it is possible to determine this area using the graph in side figure, by estimating
the mass flow based on the procedure detailed on page 240 of Raymer [1].

12
Figure 5: Preliminary capture area sizing [1].

Acapture = 1.03 ∗ 0.005125 ∗ 127 ∗ De2 = 0.86m2 (4.5)


Notice the areapwas increased by 3% to compensate for the bleed-air systems [1]. Therefore the inlet
radius is Ri = Acapture /π = 0.3m
Cowl lip has a major influence on engine performance [22], the lip radius should be 6-10% of the inlet
radius. Larger inlet radius is good for high angles of attack and side-slip, accommodating additional air
at take-off thrust, but it produces more drag because of the larger blunt front. As the requirements are
for long flights, it is reasonable to prioritize for a smaller lip radius to reduce drag. It can be observed
that for the two main engines that were selected as candidates, one is designed for short-haul flights
(the PW6000 being employed on the regional A318), the other is designed for long-haul flights. These
differences are obvious when comparing the nacelle inlets. The design we opted for tends to resemble
the latter. The lower lip has a thicker radius, to reduce the effect of higher angles the attack, improving
performance in climbing regimes. It is typically taken as 50% larger [22]. Therefore the upper inlet radius
is chosen as 0.6Ri = 18cm and the lower radius is 0.09Ri = 2.7cm. Furthermore, according to Stington
[41] the weight of the nacelle and fuselage pylon represent on average 0.75% of the empty weight as seen
in the side figure i.e.219.47kg and on average it is half that of a wing-mounted engine.

Figure 6: Nacelle weight fraction of empty weight [41].

13
4.5 Engines pitch and cant angle
A positive pitch will cause the engines to produce lift. However, in order to obtain maximum engine
efficiency, the intake has to be aligned with the incoming flow. According to Raymer [1], an angle of -2°
will allow for this to occur optimally when taking a weighted average of the flow angle with the horizontal
in terms of flight time in each flight regime. The engine will be optimized for cruise, as the aircraft is
designed for long flights. The nacelles will be canted by 2° towards the nose to be better aligned with
incoming flow, considering the bending of streamlines by the fuselage. This empirical rule is also taken
from Raymer [1]. As an additional benefit of this angle is that the plane lateral stability is also improved
[4]. Both of these inclinations will also eliminate issues associated with the heating of the tail and the
fuselage aft of the engines, as the hot exhaust is directed slightly downwards and away from the fuselage.
Furthermore, it is worth mentioning that for safety, the engines will shear off the first heavy frame first
with the use of specialized bolts. This will result in a ‘peel back’ of the engines off the fuselage in case
of a critical emergency and they will not strike the tail.

4.6 Engine Placement


Having established the engine diameter, it is possible to determine its precise position. As presented in the
preliminary report, various potential locations were considered and their advantages and disadvantages
were weighted using a trade-off analysis matrix. Finally, podded engines that make use of the windowless
design are to be attached on three fuselage heavy, between the wing and tailplane. This is depicted in
the baseline configuration 3-view drawing at the beginning of this paper.

4.6.1 Lateral and vertical position with respect to the fuselage


From the perspective of interference drag, it is better to place the engines as far away from the fuselage
as possible. However, this will increase the length of the horizontal engine pillars and thus the weight of
the structure that would need to sustain higher bending moments. A compromise between these effects
must be reached. Interference drag can be found for the fitted curve of a large engine study (the figure
below) [1]. Knowing nacelle Dmax = 1.77m we chose an interference drag of 0.03 to obtain a distance
from the fuselage of Y = 1.6Dmax = 2.72m. Upon detailed sizing of the fuselage, a more informed
compromise between structural weight and drag can be made. One should note that this study refers
to the interference drag of an engine mounted under the wing, interference with the wing itself also has
a contribution in this case. Mounting the engine closer to the fuselage and further away from the wing
reduces interference with the wing significantly, yet increases that with the fuselage. Therefore, 0.03 is
a conservative estimate.

14
Figure 7: Relation between interference drag and distance from fuselage [1].

With respect to the vertical position of the engines, these need to be located high enough to prevent
the ingestion of the turbulent air from the wing at high angles of attack. At the same time, this height
needs to be small enough to avoid hot engine exhaust from hitting the tailplane. Furthermore, the
vertical position of the engine affects the location of the aircraft center of gravity and thus its stability.
For stability reasons the engines thrust line is desired to be below the vertical CG line as this would
produce a nose down moment at stall. This position is hard to be determined without using an iterative
procedure, therefore for preliminary design this will be set to ¾ of the fuselage height using a compromise
of the above criteria. The main advantage of the engine location is the fact there is no need to allow for
a 6‘’ ground clearance from the line drawn a 5° angle from the undercarriage-TARMAC contact. The
current position of the engine provides sufficient ground clearance to avoid engine strike at maximum
landing gear compression with deflated tires. It also reduces the chance of foreign objects damage from
items on the runway (the front wheel location has considered this aspect as well). Another advantage
for this concept is that there will be no need for stair-trucks in order to board/unboard the plane, as the
height is low enough to allow for multiple-layer in-door stairs to be deployed.

4.6.2 Longitudinal position along the fuselage


The combination of the windowless design and that of fuselage-mounted engines gives a high design
flexibility when it comes to mounting the engines longitudinally. From a stability point of view, shifting
the center of gravity forwards improves static margins and allows for a smaller tailplane structure. The
engines can have multiple optimum positions, depending on the location of the wings, tail size and
undercarriage. A NASA technical document [39] characterizes the trade-off of these parameters for
a STOVL. Based on this, it was possible to obtained a preliminary value for the engine location and
subsequently, considering the installed mass for each engine (4.3 tons), to adjust this value after obtaining
the CG and stability characteristics. The final engine position is 38.8m from the aircraft nose.

4.7 Installed engine parameters


Thrust losses associated with pressure recovery are a factor as the flow at the intake cannot be assumed
undisturbed by the wing and fuselage, especially at high angles of attack and low speeds, yet these

15
cannot be easily quantified at this stage. Therefore we assumed P1 /P0 )ref erence = 1, and reduced the
values associated with a Pitot intake by another 2% P1 /P0 )actual = 0.96 [1]. A ram correction factor
Cram = 1.3 must also be applied to account for losses at M=0.79 speed [4]. This was approximated from
experimental data for the GE-1 low BPR turbofan, thus it is a conservative value. Therefore, thrust loss
percentage is given by:

%Tloss = 100Cram [P1 /P0 )ref erence − P1 /P0 )actual ] = 100 ∗ 1.3 ∗ (1 − 0.96) = 5.2% (4.6)

Engine bleed losses are given by the removal of air from the compressor stage, to be used for cabin
pressurization, control surfaces actuators, de-icying and other systems. This typically gives a thrust losses
between 1-5% [22]. Due to the size of the plane we assumed 2% as an appropriate value. Furthermore,
power extraction losses should also be factored in especially as the virtual windows will consume a
significant amount of electricity and associated cooling is required. A reduction of 3.5% thrust was
considered [22]. The nozzle also has associated losses as the exhaust is not perfectly expanded. This was
considered as 2.5%. Chevron nozzles increasing this by a further 1%. Finally, required engine thrust, to
assure a 117.1kN installed thrust output would be:

Tuninstalled, required = Tuninstalled, no losses [1 + (5.2 + 2 + 3.5 + 2.5 + 1)/100] = 133.73kN (4.7)

Weight associated with engine installation can also be accounted for. According to Stington (p. 503,
[41]), the mass increase factor for installing the engines in fuselage pods is 1.3. Therefore the installed
engine mass will be 1.3 ∗ WEngine,U ninstalled = 4.2tons. Note that choosing the engine in a fuselage pod
as opposed to a wing pod decreases the nacelle weight, but would increase the installation weight, due
to the additional length of the fuel pipes and other systems[41]. In this case however, as most of the fuel
will be stored in the fuselage, this effect would be decreased.

4.8 Engine drag estimation


Following the procedures detailed in ’Aerodynamic Analysis’ [22], engine windmilling drag was estimated
as CDW E = 0.3A EF F
Sref = 0.004 per engine, and the parasitic drag associated with the engine and nacelle
as CD0 = 6.55E − 04 per engine. Finally, interference drag was calculated previously as 0.03.

4.9 Nozzle design


4.9.1 Noise and vibration reduction
One of the disadvantages of placing the engines on the fuselage is an increase of the noise perceived
by the passengers and acoustic fatigue of the fuselage. Therefore, for this class of planes, noise has a
much higher merit index than fuel efficiency. Having searched for the state of the art in engine noise
reduction it is possible to recommend the NASA chevron nozzle [43] to effectively solve this issue. This
technology is currently employed on the GEnx and Rolls-Royce Trent 1000 engines. Using the chevron
nozzle slightly reduces engine efficiency. At the timescale of this concept aircraft (2025) the technology
readiness level (TRL) of the new generation of variable geometry chevrons that utilize shape memory
alloys for immersion control, will make it commonly available [43]. Justifying this prediction is the TRL
progression of the current generation, shown in the figure below.

16
Figure 8: TRL progression of NASA’s chevron nozzle design [42].

4.9.2 Reverse thrust systems


The aircraft will make use of reverse thrust systems in order to reduce wear on the brakes and achieve
shorter landing distances. This system will be used immediately after touchdown, to increase deceleration
early in the landing roll when residual aerodynamic lift and high speed limit the effectiveness of the breaks.
This also allows for safer landing in icy/wet runway conditions.The effect on stopping distance can be
seen in the figure below.

Figure 9: Figure shows typical relation between work done and landing run with respect to use of thrust
reverser in icy/wet runways, and it is seen that landing run is reduced about 2.5 km. [44].

In order to improve business passenger experience further, active noise and vibration control will be
implemented. This technology enhances aircraft cabin comfort by using piezoelectric actuators applied to
the structure to minimize the radiating acoustic pressure field around the engine location. As developed
by the ’Aeronautics department’ for ’SEAT’ Project [49].
There are many types of reversers, some of which are: cascade system (cold stream), clamshell door

17
system (hot stream) and bucket target system (hot stream). ‘As per the geometry of jet engine, maximum
thrust is available only at the exit of the nozzle. It will be better to install reversers at the position where
maximum forward thrust is available’ [45]. The design chosen by this paper is that of the hot stream
clamshell door system because it provides a good compromise between simplicity (highest for cascade
type) and power (highest for bucket target type). This concept can be seen in the figure below.

Figure 10: Left - engine and clam reverser diagram of main components [45], right - reverse thrust in
nominal position on a RR725 engine [46]. Both are an accurate representation of the engine to be used.

In addition to the advantages above, thrust reversers of this type can also be used in the ‘power
back’ procedure i.e. for reversing without a pushback truck and for aborting take-off safely in adverser
weather. Furthermore, the system has negligible impact on maintenance costs [47].

18
5 Fuselage Design
The design requirements specify the aircraft must
accommodate ‘40 passengers in an all business
class layout with a minimum seat pitch of 60"
’. To accommodate this, it was decided to have Galley

only two seats per row in a 1-1 layout. A 2-1


layout was deemed undesirable as it would imply Crew Seat

the need for passengers to cross in front of other Passenger Seat

passengers when they leave/return to their seat. Seat Pitch

Taking a section of the fuselage interior, there are


two seats per width, separated by the aisle as can
be seen in the sketch shown in figure 11.
The seats and aisles dimensions were chosen as
the upper values for first class passengers. From
a regulation point of view, since the aircraft ac-
commodates less than 44 passengers, the require-
ments for exit doors symmetry and evacuation in
under 90s do not apply here. Therefore two exit Toilet Exit
doors are required (traditionally one every 10-20
rows) and have been placed on opposite sides of
the fuselage, one forward on the left side (the op-
erating one) and one forward on the right side of
the cabin. Putting the latest on the aft of the
cabin was considered but it would imply having Figure 11: The fuselage layout.
an exit door close to the engine, which could have
been detrimental (in case of a fire for example).
Regarding the 4 cabin attendants seats, economy-class size was chosen for reference. These seats
are placed forward and aft of the cabin, next to the galleys space, to improve the operating conditions.
The galley space length has been calculated choosing a galley volume per passenger of 0.23m3 and its
placement in the fuselage chosen by considering other constraints (aisle space, location of exits, and
height of the upper fuselage section). In the end our cabin disposition sketch is presented figure 11. This
final layout gives a cabin length of 40.11m with a width of 2.84m (assuming a circular fuselage) and a
maximal cabin height of 1.93m.
It was decided to store the luggage in bulk because it is consistent with a small aircraft (small in
term of passengers) and may be quicker at arrival.

5.1 Nose-Cockpit
The over nose vision is critical to ensure safety at Landing and taxing. Given an angle of approach of
αapproach = 5.5◦ (which is the case at London City Airport) and a velocity approach of Vapproach = 130kts,
the over nose angle is αover nose = αapproach + 0.07Vapproach = 14.6◦ .
The nose length was also to be fixed. To do that, a divergent Mach number was chosen. Since the
cruise is done at M=0.79 and the aircraft must be able to accelerate to a maximal speed of M=0.82 at
cruise altitude. Given that, if the aircraft design is done so that MDD = 0.83, wave drag will not affect
the aircraft performance in cruise and the resulting nose length is Ln = 6.375m.

19
(a) After body length calculation. (b) Fuselage drag.

Figure 12: Graphs used to the after body design and fuselage aerodynamic considerations.

5.2 After Body


The after body length was calculated using the data shown in figure 12a. The main objective was to
minimize the pressure drag whilst keeping a skin friction drag reasonable. Considering that, a fineness
ratio of 1.75 was found and since the diameter is fixed by the cabin configuration, (d=3m) the after body
length is x=5.25m. These figures are valid for an oval or conic after body shape only. It must be noted
that the after body will be swept.

5.3 Aerodynamic Considerations


Once the cabin, nose and after body lengths were calculated, the fuselage length could be deduced:
LF = 51.74m. Using the data given in figure 12b, one can conclude that such fineness is not optimal
from the frontal drag point of view. However, the volume and wetted drag would be satisfactory.

20
6 Tailplane Sizing & Design
6.1 Tail Sizing
The T tail was found to be optimum design as it does not enter the ‘slipstream’ of the wing and fuselage.
It gives better pitch control and has predictable and well documented design characteristics. Another
advantage is the high effective aspect ratio and low interaction drag (commonly used on gliders) that
prevents deep stall at high incidences and the vertical stabilizer will have to withstand significant forces.
Crucially, this design avoids the jet exhaust which is important due to the engines being positioned high
on the fuselage.
To calculate the required tail size, historical data of tail volume coefficients is used. The tail design
chosen in the preliminary report was a conventional tail design and so to size this we must calculate
both vertical and horizontal tails. The values for the vertical and horizontal tail volume coefficients
(cVT and cHT ) vary on the type of aircraft being designed and the specifications for the characteris-
tics of the aircraft. cVT and cHT were taken as 1 and 0.09 from a table of standard values in [19].
LVT SVT
cVT = (6.1)
bW SW
LHT SHT
cHT = (6.2)
c̄W SW
LVT and LHT are the vertical and horizontal
moment arms. The vertical moment arm is sized
slightly shorter than the horizontal moment arm
to aid spin recovery. During spin, the wake from
the horizontal tail could blanket the rudder, ef-
fectively making it useless. The values of LVT
and LHT were chosen as 23.23m and 24.63m re-
spectively. Using these and the values for tail vol-
ume coefficient, the vertical and horizontal tail ar-
eas, SVT and SHT were found to be 20.78m2 and
28.72m2 . All of these values are summarized in Figure 13: A T tail configuration from [18]. Note
table 3. that this figure is to illustrate the T tail configuration
being used and not the number or position of the
engines.

c Sv
S lv

Sh
xcg lh
xnp

Figure 14: The dimensions used for the tail calculations.

Equations 6.3 to 6.6 are used to calculate the


span (b), root chord (croot ), tip chord (ctip ) and mean aerodynamic chord (c̄) from the centerline (Ȳ ).
These equations are applied to both the vertical and horizontal tail sections:

b = AR × S (6.3)

2S
croot = (6.4)
b(1 + λ)
ctip = λcroot (6.5)

21
Variable Symbol Value
Horizontal Tail Volume Coefficient cHT 1
Vertical Tail Volume Coefficient cVT 0.09
Horizontal Moment Arm LHT 24.63m
Vertical Moment Arm LVT 23.23m
Horizontal Tail Area SHT 28.72m2
Vertical Tail Area SVT 20.78m2

Table 3: A summary of the tail plane sizing values.

2 (1 + λ + λ2 )
c̄ = croot (6.6)
3 (1 + λ)
The aspect ratio (AR) and taper ratio (λ) were chosen by averaging the values from the Gulfstream
G650 and Airbus A318. These values are given below in table 4.

Horizontal Tail Aspect Ratio ARHT 4.84


Vertical Tail Aspect Ratio ARVT 1.07
Horizontal Tail Taper Ratio λHT 0.47
Vertical Tail Taper Ratio λVT 0.95

Table 4: The tail aspect ratios and taper ratios taken from the Gulfstream G650.

b (1 + 2λ)
ȲHT = × (6.7)
6 (1 + λ)
b (1 + 2λ)
ȲV T = 2 × × (6.8)
6 (1 + λ)
The quarter-chord sweep angle for the horizontal tail plane was chosen as 4◦ greater than the main
wing. This is to prevent stall ahead of the wing. The values obtained from these calculations are
summarized in table 5 below.

Horizontal Tail
Span bHT 11.79m
Root Chord croot,HT 3.31m
Tip Chord ctip,HT 1.56m
Mean Aerodynamic Chord c̄HT 2.54m
Quarter Chord Sweep (Λc/4 )HT 33◦
M.A.C Distance from Centreline ȲHT 2.59m
Vertical Tail
Span bVT 4.72m
Root Chord croot,VT 4.52m
Tip Chord ctip,VT 4.29m
Mean Aerodynamic Chord c̄VT 4.41m
Quarter Chord Sweep (Λc/4 )VT 30◦
M.A.C Distance from Centreline ȲVT 2.34m

Table 5

6.2 Possible Improvements


The first way in which the tail design could be improved is by completing a number of design iterations.
This could be achieved using flight tests which would highlight performance issues in both the tailplane
and other aspects of the aircraft.

22
In reality, commercial airline designers such as Airbus and Boeing would use computational fluid
dynamics (CFD) software to analyze the performance of the tailplane to find potential improvements. It
must be noted though that the tailplane design is highly dependent on the general configuration of the
aircraft.

7 Aircraft System Layout Design


7.1 Avionics
The aircraft will make use of a fly-by-wire system with triplex redundancy.

7.2 Emergency Systems


Standard emergency systems including oxygen supplies, life belts, and other safety equipment. The cabin
design incorporates two emergency exits at the front to avoid the engines.

7.3 Fuel Tanks


For a total fuel capacity of 32,000 liters, with 5% surge capacity to account for changes in fuel density,
we obtain a required volume of 33.6m3 .

32000 × 1.05 × 0.001 = 33.6m3

Approximately 12m3 of the wing volume is available for fuel tanks based on aerofoil and wing dimensions
and an assumed capacity of 85% of structural volume [5]. Placing fuel in the wings helps to minimize
bending moments in flight and improves aerodynamic performance. These will serve as auxiliary tanks,
and provide independent fuel supplies for each engine.
The bulk of the fuel will be stored in the fuselage, requiring a volume of 23.5m3 assuming 8% is used
for structural purposes [6]. This will constitute approximately 40% of the available under-cabin volume
which is acceptable as very little cargo space is required in this aircraft. Firewalls will be placed between
each tank, and between the tanks and the cabin to prevent any fire spreading.
Fuselage fuel storage will be split between three tanks. One tank with a volume of 14 m3 will be
placed just forward of the wings, and one tank with a volume of 7m3 will be placed just aft of the wings.
These dimensions take account of the need to store other systems in the center of the aircraft.
Spreading the fuel throughout the fuselage will allow for control of the horizontal position of the
center of gravity, and increases the safe distance between the fuel and the engines. Analysis of the
change in center of gravity with fuel loading and consumption has shown that the aft tanks should be
used first and filled last to ensure a good stability margin.

7.4 Electronics
In place of a conventional gas turbine APU, the aircraft will make use of proton exchange membrane
fuel cells (PEMFCs). APUs are relatively inefficient and heavy, and produce a large amount of emis-
sions (particularly NOx). PEMFCs by contrast are not limited by thermodynamic efficiencies, have the
potential to be considerably lighter, and produce only water as a by-product. Waste heat will be used
for pre-heating of engine fuel, de-icing, and internal temperature regulation in order to improve overall
system efficiency [7].
Until PEMFC power density increases, the weight saving over an APU will be negligible due to the
additional system required [8], but as the technology continues to mature the cell can be upgraded to
further improve the performance of the aircraft.
Approximately 0.5L of water is produced per kWh of power [9], and with some filtering this can be
used for both potable and grey-water applications. As estimated 200L of water is required [10] during
the flight, and the cells will produce more than enough, giving a weight saving of 0.2 tonnes.
An additional advantage of using fuel cells is that the conventional 115V AC power grid can be
replaced with a 270V DC one. Through a reduction in the number of converters required in the system,
and lower duty cabling, a weight saving of about 140kg can be achieved [11].

23
It is anticipated that peak power requirement of the aircraft will be 850kW [12]. Commercial PEMFCs
can supply up to 120kW of power [13] so the aircraft will make use of 8 spread throughout the airframe,
with one largely redundant.
Four fuel cells will be placed at the aft of the fuselage and will provide power to start the engine,
and to power the actuators in the empennage, two in the middle of the plane will provide power for
the actuators on the wings, and two at the front of the plane will provide power to the cockpit. All
three banks will also provide power for the cabin lighting, air conditioning, pressurization system, and
passenger electronics. The load will be balanced to ensure that no one fuel cell is drawn on too much,
and non-essential loads do not adversely affect essential systems.

7.5 Hydraulics & Actuators


The traditional hydraulic power system used to control actuators will be replaced by a set of dual
redundant ElectroHydrostatic Actuators (EHAs). Additional electrical wiring will be required, but
removal of pumping systems, pipes, and fluid reservoirs will yield a weight saving of up to 450kg [14].
Redundancy is required in order to ensure the operation of control surfaces is not compromised by
hydraulic faults, and the inclusion of multiple independently located fuel cells ensures operation in the
case of electrical failure of one part of the system.
Use of EHA provides greater actuation force and greater redundancy than purely electronic systems,
whilst reducing electrical power usage by up to 25% [14], lower costs, and better environmental per-
formance than a purely hydraulic system. A reduction in the number of components also means less
opportunity for failure, and reduced maintenance costs [15].

7.6 Environmental Control Systems


Standard environmental systems will be used to provide control of cabin temperature, pressure, and
humidity. Air will be drawn from outside, heated by excess fuel cell heat, and filtered before entering
the cabin. Maximum air pressure will be equivalent to 2,800 feet in order to improve passenger comfort.

7.7 Water & Waste


Water and waste routing will follow conventional designs. Two 20 litre water tanks will store water from
produced by the hydrogen fuel cells, and will be positioned beneath the fore and aft ends of the cabin
to supply the toilets and galleys. The toilets will make use of vacuum waste removal systems.

7.8 De-icing
De-icing of control surfaces will be achieved using excess heat from the hydrogen fuel cells. In other
parts of the airframe, additional de-icing will done with electro-thermal circuits.

24
8 Undercarriage Design
8.1 Landing Gear
The landing gear will consist of four undercarriages in a diamond arrangement. The single-wheel nose
gear will support 8.8% of the MTOW, and will be positioned 6m from the nose. Two two-wheel bogey
main gears will be positioned 26m from the nose 3m on either side of the fuselage centerline under the
wings giving good turning performance. An additional two-wheel main gear will be positioned 32m from
the nose in order to provide a takeoff angle of 10◦ [1] preventing tail strike. A clearance height of 2.5m
was chosen to ensure a reasonable takeoff angle and provide clearance under the aircraft for maintenance.
A diamond gear layout allows for good ground stability and a large crab angle which makes crosswind
landings (often necessary at London City) easier. It also makes steering easier, and provides good
visibility over the nose during ground operation, allowing the jet to manoeuvre in the airport and avoid
having to drop business passengers off on the runway. Additionally, the floor can be kept level for easy
boarding.
The nose gear will have a positive trail, and a positive rake angle of approximately 7◦ , such that it
is dynamically stable and slightly statically unstable. The main gears will not have any rake or trail as
they operate on a bogey. All gears will make use of oleo-pneumatic shock absorbers.

Parameter Symbol Value Unit


Wheel track T 24 m
Wheel base B 3 m
Landing gear height H 2.50 m
Clearance angle αc 10.23 deg
Overturn angle φOT 20.56 deg
Tyre pressure Ptyre 200 PSI
Tyre diameter D0 43.25 "

Table 6: Landing gear parameters.

With a load growth factor of 1.25, three shock struts and twin-wheel arrangement, an Equivalent
Single Wheel Load (ESWL) for the main gears was obtained using formulae from [4]. Using this a 44 x
16 Goodrich tire was selected, with an inflation pressure of 200 PSI, and a maximum diameter of 43.25".
The frontal area of each gear is 0.97m2 , giving a drag coefficient of 0.0145 [17].
The landing gear will be retracted into the fuselage in order to minimize drag as this is an important
factor during a high speed, long distance flight. The increased cost and higher weight associated with a
retractable design will be offset by fuel savings, and cost is less of an issue for a business jet than for an
economy aircraft. The arrangement of the fuselage fuel tanks will be such that in the event of a rapid
landing the landing gear would not damage the tanks.

9 Weight & Balance Predictions - following page

25
Component Weight (kg) CG (% of L from nose) Sources
Wf us Fuselage 6,390 47.0 11.6% weight saving due to windowless design and use of composites
Ww Wings 4,712 55.4 15% weight saving due to use of composites
Wf urn Furnishings 3,817 50.0 Additional 150lb/seat added to Toreenbeek’s formula;
Assumed CG at center of plane based on historical patterns.
Wac Air conditioning 1,855 50.0 Assumed CG at center of plane based on historical patterns.
Winl Nacelle & Engine 1,503 83.3 10% weight saving due to use of composites
Wmlg Main landing gear 1,318 54.9 CG from positioning of landing gear.
Wf c Flight controls 1,186 2.0 Assumed CG at nose of plane based on historical patterns.
Wav Avionics 766 2.0 Assumed CG at center of plane based on historical patterns.
Wht Horizontal tailplane 478 90.5 17% weight saving due to use of composites
Wvt Vertical tailpane 473 90.5 17% weight saving due to use of composites
Wel Electrical system 331 2.0 140kg saving due to 270V power grid;
Assumed CG at nose of plane based on historical patterns.
Whf c Fuell Cells 640 2.0 Assuming a power density of 150W/kg [7];
Assumed CG at centre of plane based on historical patterns.
Wnlg Nose landing gear 232 11.7 CG from positioning of landing gear.
Winstr Instruments 183 2.0 Assumed CG at nose of plane based on historical patterns.
Wf s Fuel system 150 47.5 CG obtained from fuel tank positioning.

26
Whydr Hydraulic system 96 50.0 50kg saving due to use of EHAs;
Assumed CG at center of plane based on historical patterns.
Wai Anti-icing system 133 50.0 Assumed CG at center of plane based on historical patterns.
Wec Engine controls 100 83.3 Assumed CG same as engines,based on historical patterns.
Wes Engine starter 96 83.3 Assumed CG same as engines based on historical patterns.
Whg Handling gear 20 50.0 Assumed CG at center of plane based on historical patterns.
We Empty Weight 24,479 47.9 From above

Wc Cargo + Passengers 4,626 From preliminary report.

Wf Fuel (all tanks) 25,793 47.5 Weight and CG from fuel mass and tank position.
Wf f Fuel (fore tank only) 11,091 34.3 Weight and CG from fuel mass and tank position.
Wf a Fuel (aft tank only) 5,417 60.8 Weight and CG from fuel mass and tank position.
W0 MTOW 54,898 43.8 From We , Wc , and Wf .

Wx Landing weight 35,135 50.2 Based on design landing weight with 70%
of max. fuel consumed and wing fuel tanks emptied last.

Table 7: Unless otherwise stated, weight estimations are from [1], and CG estimations are from [16].
9.1 Weight Reduction Concepts
In the preliminary report the airplane was designed as if it is to be made from Aluminum Alloys. This
assumption was used in order to obtain a better estimate of empty weight using historical airplane data.
Composites have been used on a high proportion of structural components only recently, due to limited
knowledge of their properties. There are two means by which our design aims to reduce weight and
consequently improve fuel efficiency.

9.1.1 Advanced Composites


Firstly, composite materials will be used on the vast majority of load-bearing components, significantly
reducing weight by offering higher specific strength. This decision was made due to the anticipated
timescale of this concept plane (realistically, it would be expected to enter service after 2025). With
aircraft such as the Boeing 787 and recently, the Airbus A350 XWB, successfully demonstrating this
technology, it is highly likely it will become the norm among high-profile aircraft manufacturers as more
knowledge is gained.
From the preliminary report, MTOW was estimated as 64.4 tonnes and the empty weight as 33.1
tonnes. This was carried out using historical data of airplanes with similar characteristics [?]. According
to Raymer [1, p. 464], advanced composites would reduce the weight of the wing by up to 15%, that of
the tail by 17% and of the fuselage+nacelle by 10%.
Estimations of each component’s contribution to the empty weight of the aircraft were obtained from
historical data of ten similar-sized aircraft presented in the previous chapter an included in the Annex.
The final calculations are presented in the table below.

Component Empty Weight Mass (Al) Mass reduction Mass (C.) Mass Saved (C.)
Units % tonnes % tonnes tonnes
Fuselage + Nacelles 25 8.275 10 7.45 0.83
Wing 20 6.62 15 5.63 0.99
Tail 5 1.655 17 1.37 0.28
Total 50 16.55 12.70% 14.45 2.1

Table 8: Mass reduction through the use of advanced materials quantification.

Therefore, the total weight of the structural components would reduce empty weight by 12.7%,
resulting in a total mass saving of 2.1 tonnes.Typically, composite panels have a much higher post-buckle
performance than conventional materials. However, current airworthiness requirements do not allow for
this post buckling performance to be used in practice as it is still a subject of research at present time.
Applying this to the weight values obtained in the previous sections yields.

9.1.2 Windowless CFRP Fuselage Design


It is traditional for business jets to have windows that are much larger than those found on conventional
civil aircraft, as this is seen as a luxury feature on the market. This paper proposes the use of a concept
that takes this to the next level and allows for a better passenger experience whilst also providing a
reduction in fuselage complexity, weight and thus higher efficiency. Our proposal is to replace conven-
tional windows with virtual augmented-reality panoramic touch screens. This idea will be implemented
on the first supersonic business jet, the Spike Aerospace S-512 [26] that is currently under development
and expected to begin manufacturing within a decade. The concept was thoroughly presented at the
2013 NBAA business aviation show in the form of the IXION windowless concept plane [27] designed
by Technicon. Therefore, the second concept implemented for mass reduction is the windowless nature
of the fuselage. There is no need for window frame reinforcements, thick and complicated multi-layer
pressure seals also there will be no fatigue concerns associated with the windows. Currently, after the
fuselage is manufactured, holes are cut in the skin and reinforcements are added around them to dis-
tribute the stress around the mission material. This also adds mass to the structure, both in the form of

27
Figure 15: IXION windowless concept interior.

reinforcements and the porthole material itself, increasing operating costs. Eliminating windows would
simplify manufacturing and reduce associated costs. The fuselage would be manufactured in a similar
way as that of the A350 XWB, in sections that are limited in size only by the dimensions of the au-
toclave. Prepreg strips of CFRP would be rolled, through the use of a robotic arm, on a drum [33].
Subsequently, the section will be cured in an autoclave and reinforced with frames and stringers. Re-
placing conventional windows would significantly decrease manufacturing and development costs because
digital windows would not be a structural component, the system does not need to be designed to resist
any loads.

At present time, display technology has reached the level where this concept can be implemented easily.
Furthermore, considering the timescale of the aircraft, displays this will only become lighter, thinner and
cheaper before manufacturing starts. At CES 2013 LG unveiled the 84LM9600 TV, with the curvature
and size required by the aircraft [34]. At launch date, unit price was 17,000$, this decreased to 9,000
$ two years later. Considering this as the cost of a custom display with drivers located on one side (to
allow for 0-bezels) as screens are placed side by side) and the fact that 2X40 displays would be required,
the total cost would be 720k $. It is reasonable to assume that by 2025 the 55EM9700 next generation
of displays would become available at the same price. This screen is 4mm thick and each panel weighs
3.5 kg including display drivers [35]. Equating to total system mass of 280 kg.

Through the comparison of the cargo and passenger versions of 7 planes from Jane’s All the Worlds
Aircraft [29], a weight reduction of the fuselage of 28% was approximated. Given the fact the fuselage is
25% of the empty weight (50% of structural weight, i.e. 8.23 tonnes), and factoring in the system mass
of the digital windows as 350kg (including 2x6 cameras placed as seen on the main diagram), yields a
weight reduction of the aircraft of 24% of fuselage structural weight. Thus 1.74 tonnes are removed by
replacing conventional windows.

Having determined a significant mass reduction it is now possible to conclude that the use of tailored
CFRP composites in conjunction with virtual augmented-reality panoramic screens would be optimum
for this design, giving a reduction in the empty weight of the aircraft of 11.6% (relative to the advanced
composites A/C). According to the UK Centre for Process Innovation [36] ‘for every 1% reduction in
weight the approximate fuel saving is 0.75%’, leading to an estimated fuel savings of 8.7%. Further
analysis, considering that less fuel will be required for a lower structure, reveals a higher fuel reduc-
tion of 10.2%, thus a total reduction of the MTOW of 6.76 tonnes considering an advanced composite
windowless aircraft. All calculations are in the Appendix.

28
10 Aerodynamic Analysis
The drag will be divided into two main components: the parasite drag (CD0 ) and the lift induced drag
(CDi ). For each aircraft component, its contribution to both drag need to be estimated. For each
component, i, the parasite drag can be found using equation 10.1. The variables in the equation are
listed in table 9. P
(Cf i , F Fi , Qi , Swet i )
(CD0 )subsonic = (10.1)
SW + CD misc + CDL&P

Cf Skin friction coefficient


FF Form factor
Q Interference factor
Swet Wetted area
CD misc Miscellaneous drags
CDL&P Leakage and protuberance drag

Table 9: Variables used in equation 10.1.

10.1 Fuselage at Crusie


Since it is assumed the fuselage does not provide any lift, there is no fuselage lift induced drag. As for
the parasite drag, it is assumed the flow around the fuselage is turbulent, thus Cf is given by equation
10.2.
0.455
Cf = (10.2)
(log10 R)2.58 (1 + 0.144M 2 )0.65
The Reynolds number R used is the minimum value between those given by equation 10.3 and 10.4.
ρLV
Re = (10.3)
µ
 1.053
L
Recuttoff = 38.21 (10.4)
K
For the fuselage, the length parameter L used is the total fuselage length L=51.74m and the fuselage skin
is supposed to be covered with smooth paint, so k=2.08ft or k= 63399×10−6 m, which gives Recutoff =
7.25×108 . The properties of air at an altitude of 10000m are given in table 10. This gives Re = 3.42×108 .
So the Reynolds number used is Re and the fuselage skin friction coefficient is Cf,fuselage = 1.70 × 10−3 .

a 299 ms−1
ρ 0.414 Kgm−3
T 223 K
µ 1.48×10−5 Kgm−1 s−1

Table 10: Properties of air at 10000m as given by [3].

The fuselage form factor was found 1.055 from equation 10.5, where f = L/d, L=51.74m and d=3m.
The fuselage interference factor is assumed to be equal to 1: Qfuselage = 1. The fuselage wetted area is
given by: Swet,fuselage = 452m2 .
60 f
F Ffuselage = 1 + 3 + (10.5)
f 400
D
= 3.83u2.5 Amax (10.6)
qunswept
The fuselage contribution to miscellaneous drag is given by equation 10.6, where Amax is the max-
imal cross sectional area and u is the upsweep angle of the aft fuselage (in radians). It was estimated
that u=14.4◦ so u=0.25rad. The maximal cross sectional area is assumed to be the fuselage maxi-
mal
P cross sectional: Amax = πR = 7m . Therefore D/qunswept = 0.838. This gives the value of
2 2

(Cf i , F Fi , Qi , Swet i ) for the fuselage at cruise as 0.812.

29
10.2 Wings at Cruise
Just like it was done for the fuselage, two Reynolds number must be calculated using the mean aerody-
namic chord as length L. This gives Rewing = 3.11 × 107 and the cut-off Recutoff,W = 5.81 × 107 , thus
the Rewing value will be the one used to calculate Cf,turbulent .
A supercritical aerofoil was chosen mainly because it could provide a laminar drag bucket, extremely
beneficial in cruise, because the flow around the wing is partly laminar. Conventional aerofoil lead to
have laminar flow over 10% to 20% of the wing while state of the art have up to 50%. Thus it is assumed
that our aircraft has laminar flow over 35% of the wing. Skin friction coefficients for both laminar
√ and
turbulent flow have then been calculated: Cf,turbulent = 2.38 × 10−3 and Cf,laminar = 1.328/ Re =
2.38 × 10−4 . The Wing skin friction coefficient was then found to be 1.63 × 10−3 using equation 10.7.

Cf = 0.35Cf,laminar + 0.65Cf,turbulent (10.7)

The wing Form Factor is given by equation 10.8, where (x/c)m = 0.37 and is the chord ratio of the
maximum thickness to chord. (t/c) = 0.14 is the maximum thickness to chord ratio. The wing form
factor was found to be 1.59.
     4  
0.6 t t
F FW = 1 + + 100 1.34M 0.18 (cos Λm )0.28 (10.8)
(x/c)m c c

The wing interference factor is assumed to be equal to 1, no matter what condition are considered
(take-off, cruise, and landing): Qwing = 1.
The lift induced drag at cruise condition is given by equation 10.9 and was found to be 4.04 × 10−3 .
The variables in this equation are listed in table 11.
 2
CDi = k CLα (α + iW − α0W ) (10.9)

Finally, the value of was found to be 0.6929. This value is lower than the fuselage one, which
P
wings,cruise

Symbol Description Value


k Induced drag coefficient 4.68 × 10−2
CLα Lift curve slope of clean wing 6.476
α Angle of attack of aircraft at cruise 1◦
iW Wing setting angle −1.5◦
α0W Zero lift angle of the wing −3.10◦

Table 11: The variables used in equation 10.9.

is coherent since the wing have been designed for their aerodynamics performance while the fuselage is
designed to accommodate the passenger.

10.3 Wings at Landing


At landing, the stall speed was calculated, assuming a load factor of 1 and using equation 10.10. W is the
landing weight estimated after the first 1st descent on the mission plan: W = 9.81 × 39100N, ρ is the Air
density at sea level (London City) (1.225Kgm−3 ), S is the reference area (150.15m2 ) and CL,max is the
maximum lift coefficient in landing condition (all flaps/slats extended) (CL,max =2.8). The stall speed on
landing was found to be 38.60ms−1 (75kts). The stall speed value is pretty low compared to the usual one
of airliner (100kts) however, a huge CL,max is used in purpose to lower this value and made the landing
at London City Airport possible. We will assume the landing speed to be 1.15stall,landing which gives
Vlanding = 86.3kts or Mlanding = 0.13. Then the Reynolds number, Re = 1.43 × 107 , is lower than Recutoff
and the laminar and turbulent drag coefficient are Cf,laminar = 3.51×10−4 and Cf,turbulent = 2.83×10−3 .If
the same assumptions are made in respect to the laminar/turbulent proportions then the wing skin
friction coefficient is Cf = 1.96 × 10−3 .Considering the extension of the flaps and slats, the flow around
the wing is obviously different than in cruise, since the slat and flap are slotted fowler, they are responsible

30
of re-energizing the boundary layer. Thus the proportions of laminar/turbulent flow might have changed
since cruise but there is no way of knowing it. Which is why it was conserved as 35-65%.
s
2W
Vstall,landing = (10.10)
ρSCL,max

Since the Mach has changed, the form factor at landing is different than the one at cruise: F FW =
1.15.
The lift induced drag will also change, and some lift induced drag created by the extending flaps need
to be taken into account. Indeed the lift curve slope, the angle of attack and the zero lift angle of the
wing will be changed by the extension of the flaps/slats and will be equal to CLα = 7.124, α = 9◦ (usual
value, it cannot be too high to prevent tail strike) and α0 = −18◦ (if ∆α0 due to flap is estimated of
about -15◦ at landing). Then CDi = 0.471.
Also the extending flap create Lift induced drag ∆CDi = kf2 ∆CL2 cos Λc/4 , where kf = 0.5, ∆CL lift
gained by the flap extension ∆CL = 1.07 which leads to ∆CDi = 0.26. Such value is relatively high but
the flap extended fully are also supposed to create drag so it remains coherent.
From the miscellaneous drag point of view, flaps will be extra drag when they are extended in other
words at take-off and landing and such drag will be given by: CD0 flaps = 0.0023(flap span/wing span)δflap ,
where δflap is the flap deflection. The deflection will be assumed to be 65% at landing. The extra drag
is then ∆CD0 flapsP= 8.95 × 10−2 .
The value of wings,landing is then found to be 0.6027. This value is lower than the one in cruise,
which could be surprising considering the extended flap. However the flap will add lift induced drag and
parasite drag on their own and this will be taken into account in the CD misc or CDi .

10.4 Fuselage at Landing


The only parameter that change when the flight phase changes is the skin friction coefficient. Indeed
it is the only parameter function of Reynolds number or Mach number. With the landing velocity, the
Reynolds number related to the fuselage can be calculated and Re=1.57×108 . So theP Reynolds number
used is Re and the fuselage skin friction coefficient is Cf,fuselage,L = 2×10−3 . This gives fuselage,L = 0.95.

10.5 Wings at Take-off


Similarly in take-off condition at MTOW the stall speed was calculated with W = 9.81 × 57000 and
CL,max = 1.96 which leads to Vstall,TO = 55.7ms−1 (108kts), which is a coherent value. The take-off
speed was assumed to be VT O = 1.10×Vstall,TO = 119kts and MT O = 0.18 which leads to Re = 1.97×107
lower than Recutoff . The laminar and turbulent drag coefficient are: Cf,laminar = 2.99 × 10−4 and
Cf,turbulent = 2.69 × 10−3 . Keeping the same assumptions about the laminar/turbulent proportion, the
wing skin friction drag is Cf = 1.85 × 10−3 .
The form factor in Take-Off condition is: F FwT O = 1.22. In the calculation of the lift induced drag,
the angle of attack will be assumed identical to the landing one since the tail strike issues are also to
be considered. However the zero lift angle of attack will be assumed as α0 = −8◦ (since a usual value
for ∆α0 due to flap at take-off is -5◦ ). These values lead to CDi = 0.132. Even though this value is a
bit high (drag should be minimize at take-off), it is much lower than the landing value which remains
coherent.
Compared to the landing condition, the flap and slats are not as extended for take-off. It is usually
assumed that the take-off CL,maxT O = 0.70CL,max landing which is why the lift gained with the flap
extension ∆CL is also assumed to be 70% of the landing value ∆CLT O = 0.749. Thus the lift induced
drag due to the flaps will be ∆CDi = 0.13.
From the miscellaneous drag point of view, similarly to the landing flaps will be extra drag when they
are extended. Usually the flap deflection at take-off is between 20◦ to 40◦ . In this case the deflection will
be assumed to be 20◦ at take-off (since the runway is short, P the priority is to reduce drag). The extra
drag is then CD0 flaps = 2.75 × 10−2 . All of these values gives wings,landing = 0.509.

31
10.6 Fuselage at Take-off
Similarly, the Reynolds number related to the fuselage at take-off was calculated and found to be equal
to Re=2.07×108 and is P lower than Recutoff so the fuselage skin friction coefficient is Cf,fuselage,T O =
1.92 × 10−3 . This gives fuselage,T O = 0.6028.

10.7 Horizontal Tail Plane


At cruise, similarly to the wing, the Reynolds number was calculated with l=MAC=2.54m: Re=1.38×107
and Recutoff = 8.68 × 106 so Recutoff will be used to calculate Cf,turbulent . The Laminar and Turbulent
skin friction coefficient are Cf,laminar = 3.24 × 10−4 and Cf,turbulent = 2.91 × 10−3 . Since a NACA profile
is used, the proportion of laminar flow was assumed to be lower than for the wing: 10%. So the skin
friction coefficient is given by equation 10.11.

Cf = 0.10Cf,turbulent + 0.9Cf,turbulent = 2.65 × 10−3 (10.11)

The form factor follows the same equation as the wingPso F FHT P cruise = 1.45. The interaction coefficient
is usually of Q=1.05 for a conventional tail. Finally HT P cruiseP= 0.2350.
Following the same principle as always, the calculations gave HT P TakeOff = 0.1930 and HT P Landing =
P
0.1863 (it can be noticed that for the landing conditions Re < Recutoff so it is Re which is used for the
Cf,turbulent calculation).
The Horizontal tail plane will generate lift induced drag that can be expressed using equation 10.12.
This leads to CDi HT P = 4.82 × 10−6 for cruise and 3.88 × 10−4 for Take-off and landing.

SH
CDi HT P = ηH 2
kH CLH (10.12)
Sref

10.8 Vertical Tail Plane


Following the exact same procedure, we ended with = 0.1223 and
P P
P V T P cruise = 0.1153, V T P takeoff
V T P landing = 0.1185.

10.9 Conclusion
Gathering the drag analysis carried out for each component and adding the extra drag such as under-
carriage, flaps, windmilling and fuselage upsweep drag when it is relevant i.e. the undercarriage and flap
drag are only for take-off and landing, and the flap drag vary between take-off and landing as it was
detailed before. This leads to a parasite drag of CD0 c = 0.0158 at cruise, CD0 T O = 0.1012 at take-off,
and CD0 L = 0.1632 at landing. Adding to that the Lift induced drag by the Wing and the horizontal
tail plane (and the extra ∆CDi for the flaps at Take-off and Landing) and the ‘leakage and protuberance
drag’ which is usually assumed as 0.03CD0 , the final drag coefficient of the aircraft is obtained for cruise:
CD,cruise =0.0203, at take-off CD,take-off = 0.3639 and at landing CD,landing = 0.8989. Details are given
in figure 16.
Considering that the Lift coefficient in cruise is given by 10.13, finally the Lift to drag ratio in cruise
will be L/D=CL /CD =16.1.
Wmidcruise
CL = 1 2
= 0.33 (10.13)
2 ρSV
These values of drag coefficient seem reasonable, the Landing drag is really high compared to the
other but is something we are looking for to help the aircraft decelerate and reduce the landing distance.
However it is only by studying the aircraft performance that we will see if the aircraft does meet the
design requirements.

32
Figure 16: Details of the aerodynamic analysis calculations.

33
11 Static Stability & Trim
An aircraft is subject to minor changes in the forces that act on it - these changes are called perturbations.
If the tendency of an aircraft is to return to this original speed and orientation without input from the
pilot, the aircraft is described as statically stable. This implies positive stability. In comparison, if
an aircraft’s orientation and speed diverge from the original values after a perturbation, the aircraft is
described as statically unstable and hence has negative stability. The other case is neutral stability,
where a perturbation would cause neither the orientation or the speed to diverge or converge to the
original values - this is zero stability. These concepts are illustrated in figure 17.

pitch-down disturbance

statically stable

statically neutral

statically unstable

Figure 17: Longitudinal static stability [18].

A tail plane is introduced to aircraft design in order to efficiently compensate for the large positive
pitching moment caused by the main wings. In addition, the tail plane can be used to modify the center
of gravity. This allows us to manipulate the static margins so we can achieve the longitudinal stability
characteristics required in the aircraft design specification. The vertical tail plane provides desirable
lateral stability characteristics by providing yaw stability. The design chosen utilities swept wings which
will also improve the lateral stability due to differential drag: as one wing is pushed forward, the effective
sweep angle decreases and so drag increases while drag is reduced on the opposite wing. The resulting
drag distribution creates a restoring moment.

11.1 Lateral Stability


It is important that the aircraft will return to its original state after a perturbation in roll or yaw. A
perturbation in yaw could be caused by a gust or due to single engine failure. The yaw and roll motion
is coupled and so the affects of both roll and yaw must be investigated.

11.2 Tail Lift-Curve Slope


Using the tail design we can now find the lift curve slope of the tail. This is required for the longitudinal
stability analysis. Equation 11.1 gives the lift curve slope of the tail where FHT , β and η are variables
given by equations 11.2, 11.3 and 11.4.
2πAHT Sexp,ht
CLα ht = s  2   Sref,ht Fht (11.1)
AHT β tan2 Λmax,ht
2+ 4+ η 1+ β2

FHT is the fuselage spillover factor which is caused by airflow being directed around the fuselage which
increases the airflow reaching the tail; β corrects for compressibility effects, and η is the aerofoil efficiency

34
factor.  2
dht
FHT = 1.07 1 + (11.2)
bHT
p
β = 1 − M2 (11.3)
βCLα hta
η= (11.4)

The values calculated for the tail lift curve slope are summarized in table 12. To calculate these values
the lift curve slope for the tail was required and so a NACA 0011 and NACA 0008 were selected for the
horizontal tail and vertical tail respectively. The vertical tail section is thinner to reduce drag.

Symbol Description Value


Fht Fuselage spillover factor 1.257
β Compressibility effects 0.6131
η Aerofoil efficiency factor 0.95
CLα ht Horizontal lift-curve slope 3.10 1/rad
CLα hta Horizontal tail lift-curve slope 6.30 1/rad
Λmax,ht Horizontal tail maximum thickness sweep 32.7◦
α0h Tail zero lift angle 0 rad

Table 12: Results of the tail lift slope calculations.

11.3 Downwash & Fuselage Effects


The tail efficiency factor (ηh ) is less than 1 due to the loss in dynamic
pressure behind the main wing compared to the free-stream flow. For
conventional tails this value is about 0.9 but because we are using a
T-tail, we have assumed ηh = 1.
Kf is an empirical constant which accounts for the position of the
wing along the fuselage. Kf was found from figure 18 as 1.6. Using
this, the pitching moment about the center of gravity can be found
using equation 11.5.
Lf Wf
CMα f = Kf (11.5)
c̄w Sw
The derivative of downwash angle with respect to α is found using
the empirical relation given by equation 11.6, where the values of KA ,
Kλ and Kh are given by equations 11.7, 11.8 and 11.9 respectively.
C |
Note that CLLα WWM|M=0.79
=0
can be assumed equal to 1.2. dαdε
was found
α
to be 0.307 1/rad.
 1.19
dε q CLα W |M =0.79
= 4.44 KA Kλ Kh cos Λc/4 (11.6)
dα CLα W |M =0

1 1
KA = − (11.7)
AR 1 + AR1.7
10 − 3λ
Kλ = (11.8)
7

Figure 18: NACA TR-711 The 1 − |hHT /b|


Kh = (11.9)
relationship between Kf and
p
3
LHT /b
wing position from [20].

35
11.4 Longitudinal Stability
For the longitudinal stability analysis, values to be found include the neutral point, the static margin
and the pitching moments about the center of gravity and the neutral point. The neutral point is the
point where the pitching moment does not change with angle of attack, the static margin is the difference
between the neutral point and the center of center of gravity. We know that this value must be positive
∂Cmα
due to ∂α being negative for a longitudinally stable aircraft.
A more stable aircraft is achieved by making the neutral point as far aft of the center of gravity as
possible. This will also increase the static margin which will increase the corrective pitching moments
caused by a perturbation. We must also note though that it can be dangerous for an aircraft to be too
stable as it will be difficult to maneuver quickly if something needs to be avoided. This is the reason
why some fighter jets are sometimes made unstable on purpose.
Table 13 gives the values which are required to calculate these variables. The calculations themselves
were completed using equations 11.10 to 11.13.

Symbol Description Value


xach Aerodynamic center of horizontal tail 53.97m
hht Height of horizontal tail above wing at zero lift angle 9.42m
xacw Aerodynamic center of wing 29.34m

Table 13: The inputs required for the longitudinal stability analysis.

∂ Sh xach
xnp CLα w CM αf + ηh CLα h (1 − ∂α ) Sw c̄
= ∂ Sh
(11.10)
c̄ CLα w + ηh CLα h (1 − ∂α ) S w

xnp − xcg
SMpower-off = (11.11)

 
∂CM cg (xacw − xcg ) ∂ Sh (xach − xcg )
= −CLα w + CMα f − ηh CLα h 1 − (11.12)
∂α c̄ ∂α Sw c̄
∂CMα ∂CLα w
=− SMpower-off (11.13)
∂α ∂α
It is also necessary to calculate the static margin when the engines are in use. This is found using
equation 11.14 from [20]. The outputs of the longitudinal stability analysis are given by table 14.

SMpower-on ≈ SMpower-off − 0.02 (11.14)

Symbol Description Value


xnp Neutral point 34.22m
SMpower-off Static margin power-off 0.21
SMpower-on Static margin power-on 0.19
xcg Aircraft center of gravity 33.2m

Table 14: Longitudinal stability calculation results.

11.5 Trim Analysis


So far, we have selected an appropriate wing and tailplane position as well as the tail areas. We must now
find the tail setting angle (ih ) for trimmed flight. Trimmed flight is achieved when the sum of pitching
moments about the center of gravity is zero and the sum of lift forces is equal to the weight. We can plot
CM cg against −(CLw )cruise by using equation 11.15, and by varying the angle of attack and tail setting
angle in equations 11.16 to 11.18, a number of characteristic lines can be plotted. A vertical line can
also be plotted corresponding to the total lift required given by equation 11.19. The intersection of the

36
characteristic lines with the total lift when the pitching moment about the center of gravity is zero gives
the tail setting angle.
   
xacw − xcg Sh xach − xcg zt T
CM cg = −(CLW )cruise + CM0 W + CMα f α − ηh CLh + (11.15)
c̄ SW c̄ qcruise SW c̄

Sh
(CL )cruise = (CLW )cruise + ηh CLh (11.16)
SW
A cos2 Λc/4
   
CM0 W = CM0 airf − 0.01 × 1.3 (11.17)
A + 2 cos Λc/4
(CLW )cruise = (CLα W )cruise (α + iW − α0W ) (11.18)
 
d
CLh = CLα h (α + iW )(1 − ) + (ih − iW − α0h ) (11.19)

12 Control Surface Design


12.0.1 Elevator & Rudder Design
The elevator and rudder sizes are determined by horizontal and vertical tail sizes. The tailplane does not
need to carry any fuel and so the elevator and rudder can be made as large as possible. Aerodynamic
flutter effects are minimised by hinging both of these control surfaces at 20% chord. The trend for jet
transport aircraft is for the elevator chord length to be equal to 25% of the mean aerodynamic chord of
the horizontal tail, and the rudder 35% of the vertical tail.
The equation of normal forces for the rudder is given by equation 12.1 where static stability is achieved
when ∂N∂β ≥ 0. A summary of the variables in equation 12.1 is shown in table 15. This equation has
been included to illustrate the point that the rudder dimensioning carried out in this report is subject
to optimization during flight testing.

N = Nwing + Nwδα δα + Nfuselage + Fv LVT − T Yp − Fp (xcg − xp ) (12.1)

Nwing , Nwδα δα , Nfuselage Normal forces from the wing, ailerons and diffuser
Fv Vertical tail lift
Yp Horizontal distance from the centerline of the engine
T Net thrust of the jet engines
Fp Force perpendicular to the thrust line
xp Distance from the nose to the front of the engines.

Table 15: Variables used in equation 12.1.

12.0.2 Aileron Design


The position of the ailerons are determined by the trailing edge flaps developed in the aerodynamics
section. The trailing edge flaps span from 0% to 60% of the wing and so it was chosen that the ailerons
should span from 62% to 98% of the wing span to fit the historical data given in figure 19. This gave the
aileron chord ratio as 0.25 based on the mean aerodynamic chord of the main wing. The hinge position
was chosen to be 20% for mechanical reasons. These values, along with the elevator and rudder values
are summarized in table 16.

13 Performance Estimation
NB: the notation used are the same as the one given in the lecture notes. The detailed Calculation tables
are available in the Appendix.

37
Figure 19: Historical data for aileron sizing from [19].

Control Surface Span Chord Hinge Position Area


Elevator 11.79m 0.635m 20% 7.49m
Rudder 4.72m 1.54m 20% 7.27m
Ailerons 6.43m 1.18m 20% 7.57m2

Table 16: A summary of the control surface dimensions calculated.

13.1 Take-off performance


First of all, the take-off distance will be calculated. In order to do so, it is divided into 4 different
phases: Ground roll, Rotation, Transition, Climb to hobs (obstacle height hobs = 35f t according to FAR
25 regulations).

Ground Roll
The ground roll distance is given by equation 13.1. Values used are summarized in table 17. Using these
values gives KT = 0.389, KA = −1.6 × 10−5 and Sg = 543.5m.

1 ln(KT + KA V22 )
Sg = (13.1)
2gKA KT + KA V12

T
KT = (13.2)
W −µ
CL2
  
ρ W
KA = µCL − CD0 − (13.3)
2 S πAe

Rotation Distance
The rotation time depends on the pilot, however, it is commonly assumed that the rotation lasts 3s
[1] and if the acceleration is assumed to be negligible then the rotation distance is SR = 3VT O where
V2 = VR which gives SR = 167.1m.

38
Symbol Description Value
g Acceleration due to gravity 9.81ms2
T/W Thrust to weight ratio 0.419
µ Runway dry friction coefficient 0.03 (FAR25)
ρ Air density at sea level 1.225kgm−3
W/S Wing loading 3724.1Nm−2
CL Lift Coefficient 0.186
CD0 Parasite drag 0.1012
A Wing aspect ratio 8.5
e Oswald efficiency factor 0.8
V1 0
V2 61.3ms−1

Table 17: Values used for the ground roll calculations.

Transition & Climb Segment


The radius climb is assumed to be constant and the transition speed is assumed as VT R = 1.15Vstall VT R =
64.06ms−1 . The manoeuvre radius is given by equation 13.4.

VT2R
R= (13.4)
(n − 1)g

Usually at take-off the load factor is about n=1.2 which leads to R=2091m. The manoeuvre angle is
given by equation 13.5.  
T L
γCL = sin−1 − (13.5)
W D
The Lift over Drag ratio needs to be calculated at this condition: L/D = CL /CD and the lift coefficient
can be assumed as 90% of the maximum take-off lift coefficient so CL = 1.764. The Drag coefficient is
the one calculated with flaps extended for take-off and landing gear out: CD = 0.3639 so L/D=4.85.
Given the radius and the manoeuvre angle, the vertical distance traveled may be calculated as: hT R =
R(1 − cos γCL ) and hT R = 47.78m.
The take-off phase of the flight is considered as ended as soon as the aircraft reaches hobs . Thus if
hT R > hobs the climb distance will be equal to zero SCL = 0 since there is no need to climb (to reach
the end of the take-off phase) and the transition distance is given by equation 13.6.

(13.6)
p
ST R = R2 − (R − hobs )2

If hT R < hobs then the transition distance is ST R = R sin(γCL ) and the climb distance is defined as the
distance travelled while the aircraft climb to hobs and is given by equation 13.7.
hobs − hT R
SCL = (13.7)
tan γCL
In this case hT R > hobs since hobs = 10.7m so ST R = 211m and SCL = 0m.

Take-Off Distance
The final Take-off distance is then ST O = 1.15(SG + SR + ST R + SCL )S = 1049.4m. Such distance
is reasonable considering our relatively high MTOW, it should allow the aircraft to be operated on
many airports, which is an advantage, especially for a business jet. However to complete the study, the
Balanced Field Length must be evaluated.

Balanced Field Length (BLF)


The FAR25 states that if the aircraft suffer an engine failure it must be able to either continue take off
and climb to hobs if the engine failure happens after V1 (decision speed) or abort take-off and decelerate

39
to a full stop safely if it happens before V1 . The runway length required to do so is called the balanced
field length and may be calculated using a semi-empirical equation given by 13.8. The value of the
balanced filed length (BLF) was found as 4542ft (1384m).
  
0.863 W/S 1 655
BFL (in ft) = + hobs + 2.67 + (13.8)
1.2.3G ρgCL,climb T /W − U σ

G = γCL − γmin (13.9)


 
T − DOEI
γCL = sin−1
(13.10)
W0
To fulfill both the FAR25 requirements and the design requirements to operate at London City Airport,
the maximum of STO and BFL must be lower than the Take-Off Distance Available (TODA) at London
City. In this case the maximum distance is the BFL: BFL=1384m and at London city, the TODA
is of 1385m. Objectively, this means that the aircraft can legally operate at London City, respecting
the FAR25 regulations. However, it must be noticed that the margin is extremely small, but since the
maximum take-off weight was overestimated in the preliminary report, it is likely that the BFL with a
correct value of MTOW will be lower and thus there should not be any trouble operating at London
City Airport.

40
γ γ
γ γ

γ
μ
13.2 Mission Performance
Based on an empty weight 24.2 tonnes from the component weight estimations, fixed empty weight and
fuel fractions, a MTOW of 57 tonnes, and a MZFW of 38.5 tonnes, the following mission performance
was obtained. Cruise SFC was taken to be 0.41, based on a 25% improvement from using the Rolls Royce
Advance engine (see powerplant section).

43
ρ

γ
γ

ρ
ρ ρ ≤ ρ ρ
ρ
Flight Envelope
14000

12000

10000
Altitude (m)

8000

6000

4000

2000

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
Mach Number
14 Cost estimations
’Cost is the ultimate figure of merit in aircraft design and optimization’ [22]. Therefore, although this
is not one of the required topic of this paper, and the process of cost evaluation is difficult and ever-
changing as material and manufacturing processes are improved (considering the plane would likely be
manufactured in 2025), our team carried out a preliminary costing assessment. Adding to the already
determined cost of the windowless system (483k£), an empirical rule for estimating unit life-cycle cost
for a business class plane is to multiply the empty weight by 1000£/kg. This gives a cost of 29.75 million
£ (based on 2012 figures, [22]) or 38 million £ (in 2025, based on a 1.5% inflation rate).

46
References
[1] Raymer D. Aircraft design a conceptual approach. Reston, VA: American Institute of Aeronautics
and Astronautics; 2012.
[2] Airfoiltools.com. NASA SC(2)-0714 AIRFOIL (sc20714-il) [Internet]. 2015 [cited 26 March 2015].
Available from: http://airfoiltools.com/airfoil/details?airfoil=sc20714-il
[3] Fr.wikipedia.org. Air [Internet]. 2015 [cited 26 March 2015]. Available from: http://fr.wikipedia.
org/wiki/Air
[4] Roskam J; Airplane Design: Part IV: Layout Design of Landing Gear & Systems; University of
Kansas; Lawrence; 1989.
[5] Aerospace Engineering e-Mega Reference, Butterworth-Heinemann; 1 edition; 2009.
[6] Levis E. Systems Packaging Weight & Balance Estimation. Lecture presented at; 2015; Imperial
College London.
[7] Klebanoff L, Pratt J, Munoz-Ramos K, Akhil A, Curgus D, Schenkman B; PEM Fuel Cell Systems
for Commercial Airplane Systems Power; Sandia National Laboratories; Arlington, Virginia; 2011.
[8] Hill et al; Modelling of Fuel Cell APU Utilisation for Aircraft Applications; 46th AIAA/AS-
ME/SAE/ASEE Joint Propulsion Conference & Exhibit; 2010.
[9] Renouard-Vallet G, Saballus M, Schmithals G, Schirmer J, Kallo J, Friedrich K A; Improving the
environmental impact of civil aircraft by fuel cell technology: concepts and technology progress;
Energy & Environmental Science; 2010.
[10] Guide to Hygiene and Sanitation in Aviation (third edition); World Health Organization; Geneva;
2009.
[11] Brombach J, Schröter T, Lücken A, Schulz D; Optimizing the Weight of an Aircraft Power Sup-
ply System through a +/- 270 VDC Main Voltage; IEEE 7th International Conference-Workshop
Compatibility and Power Electronics CPE 2011; Tallinn , Estonia; 2011.
[12] Spencer K M; Investigation of Potential Fuel Cell Use in Aircraft, Institute for Defence Analyses,
December 2013.
[13] Corcau I J, Dinca L; On using PEMFC for Electrical Power Generation on More Electric Acraft;
International Journal of Electronics and Electrical Engineering; 2012.
[14] Charrier J, Kulshreshtha A; Electric Actuation for Flight & Engine Control: Evolution & Challenges;
45th AIAA Aerospace Sciences Meeting and Exhibit; January 2007.
[15] Rongjie et al; Design and Simulation of Electro-hydrostatic Actuator with a Built-in Power Regu-
lator: Chinese Journal of Aeronautics, Volume 22, Issue 6, December 2009
[16] Torenbeek E. Synthesis of subsonic airplane design. Dordrecht [u.a.]: Kluwer; 1990.
[17] Levis E. Systems Aerodynamic Analysis. Presentation presented at; 2015.
[18] Wikipedia. Longitudinal static stability [Internet]. 2015 [cited 10 March 2015]. Available from: http:
//en.wikipedia.org/wiki/Longitudinal_static_stability
[19] Levis E. Aerospace Vehicle Design - Fuselage and Flying Surface Design. Lecture presented at; 2015;
Imperial College London.
[20] Levis E. Aerospace Vehicle Design - Stability, Control & Handling Qualities. Lecture presented at;
2015; Imperial College London.
[21] Usatoday30.usatoday.com. USATODAY.com [Internet]. 2015 [cited 13 March 2015]. Available from:
http://usatoday30.usatoday.com/weather/wstdatmo.htm

47
[22] Dr E. Levis, Dr R. Hewson, AE3-403 Aerospace Vehicle Design Lecture Notes. Imperial College
London; 2014/2015.
[23] Dr E. Levis, Dr R. Hewson, AE3-403 Aerospace Vehicle Design Coursework Assignment Briefing.
Imperial College London; 2015.
[24] [Internet]. 2015 [cited 2 March 2015]. Available from: http://www.ukvirtual.co.uk/dl/charts/eglc.pdf
[25] Hickey S., The windowless plane set for take-off in a decade [Internet]. the Guardian. 2014
[cited 2 March 2015]. Available from: http://www.theguardian.com/business/2014/oct/26/
innovations-windowless-plane
[26] Spike Aerospace - Home Page [Internet]. 2015 [cited 2 March 2015]. Available from: http://www.
spikeaerospace.com/
[27] Nbaa.org. NBAA - National Business Aviation Association [Internet]. 2015 [cited 2 March 2015].
Available from: http://www.nbaa.org/
[28] NASA. The Double Bubble D8 [Internet]. 2015 [cited 2 March 2015]. Available from: http://www.
nasa.gov/content/the-double-bubble-d8-0/
[29] Jackson P. IHS Jane’s all the world’s aircraft.
[30] Rolls-royce.com. BR725 [Internet]. 2015 [cited 2 March 2015]. Available from: http://www.
rolls-royce.com/civil/products/smallaircraft/br725/
[31] 12. Wikipedia. Pratt 0̆026 Whitney PW6000 [Internet]. 2015 [cited 2 March 2015]. Available from:
http://en.wikipedia.org/wiki/Pratt_%26_Whitney_PW6000
[32] Rolls-royce.com. Small aircraft engines [Internet]. 2015 [cited 2 March 2015]. Available from: http:
//www.rolls-royce.com/civil/products/smallaircraft/
[33] Airbus making first composite barrel for A350 XWB [Internet]. 2015 [cited 20
March 2015]. Available from: http://blog.seattlepi.com/aerospace/2010/12/07/
airbus-starts-making-first-composite-barrel-for-a350-xwb/
[34] LG Press Release [Internet]. 2015 [cited 21 March 2015]. Available from: http://www.lg.com/au/
press-release/lg-to-unveil-worlds-first-105-inch-curved-ultra-hd-tv-at-ces-2014
[35] OLED Information [Internet]. 2015 [cited 22 March 2015]. Available from: http://www.oled-info.
com/introduction
[36] Center for Process Innovation Windowless Fuselage Concept [Internet]. 2015 [cited 22 March 2015].
Available from: http://www.uk-cpi.com/windowless-fuselage/#.VRQ6VxA_iSq
[37] Rolls Royce ACARE Workshop Presentation [Internet Document]. 2015 [cited 24 March
2015]. Available from: http://ec.europa.eu/research/transport/pdf/aerospace_engineers_
between_innovative_en.pdf
[38] AIN Online: Rolls-Royce Advances Toward UltraFan [Internet Article]. 2015 [cited 24 March
2015]. Available from: http://www.ainonline.com/aviation-news/air-transport/2014-07-14/
rolls-royce-advances-toward-ultrafan
[39] Charles c. Smith, Effect of Engine Position and High Lift Devices on Aerodynamic Characteristics.
NASA. Langley Reseach Center, Hampton Va. 1971 [Internet Document]. 2015 [cited 24 March 2015].
Available from: http://ntrs.nasa.gov/archive/nasa/casi.ntrs.nasa.gov/19710013648.pdf
[40] Kundu, A.K., Aircraft Design. Cambridge: Cambridge University Press; 2010.
[41] Darrol Stington, The Design of the Aeroplane p.518, 2nd edition, Blackwell Publishing 2001
[42] NASA.gov All Rights Reserved. [Image]. 2015 [cited 24 March 2015]. Available from: http://www.
nasa.gov/images/content/476644main_trl_static_image_1300x900.jpg

48
[43] Clayton J. Bargsten, Malcolm T. Gibson, NASA Innovation in Aeronautics; Washington DC
[Ebook]. 2011 [cited 24 March 2015]. Available from: http://www.aeronautics.nasa.gov/ebooks/
downloads/nasa_innovation_in_aeronautics.pdf
[44] Rodger L. Modglin, Frederick H. Peters, U.S. Patent (6,487,845 B1), Pivot Fairing Thrust Reverser,
Cleveland, Tulsa, 2002
[45] Mohd Anees Siddiqui and Md Shakibul Haq, Review of Thrust Reverser Mechanism used in Turbo-
fan Jet Engine Aircraft; Integral University, Lucknow, India [Online Paper]. [cited 24 March 2015].
Available from: http://www.ripublication.com/irph/ijert_spl/ijertv6n5spl_18.pdf

[46] Reducing Landing Distance, Aerospaceweb.org [Internet]. 2010 [cited 24 March 2015]. Available
from: http://www.aerospaceweb.org/question/propulsion/q0181.shtml
[47] Creative Commons. Anonymous Author. [Internet Image]. [cited 24 March 2015]. Available from:
http://i.ytimg.com/vi/ASqbmx5TfHA/maxresdefault.jpg
[48] Mohd Anees Siddiqui, UTILIZATION OF THRUST REVERSER MECH-
ANISM IN TURBOFAN ENGINES – A REVIEW, India [Online Pa-
per]. [cited 24 March 2015]. Available from: http://www.ijtra.com/view/
utilization-of-thrust-reverser-mechanism-in-turbofan-engines-a-review.pdf
[49] Green Aviation Conference Proceedings, Aeronautical Department, Imperial College London
[Online Paper]. 2012. [cited 24 March 2015]. Available from: https://workspace.imperial.ac.uk/
greenaviation/Public/2012%20Symposium%20material/Green%20Aviation%20booklet%202012.
pdf

49
15 Appendix
15.1 Fuselage Calculations

50
15.2 Aerodynamic Analysis

51
52
53
15.2.1 Tail Plane

54
55
15.3 Performance

56
Engine Calculations
Weight Calculations
59
60
61

View publication stats

You might also like