You are on page 1of 121

DEGREE PROJECT IN VEHICLE ENGINEERING,

SECOND CYCLE, 30 CREDITS


STOCKHOLM, SWEDEN 2017

Degree Project in
Aeronautics
Conceptual design, flying and handling qualities
of a supersonic transport aircraft.

NIKOLAOS PERGAMALIS

KTH ROYAL INSTITUTE OF TECHNOLOGY


SCHOOL OF ENGINEERING SCIENCES
www.kth.se
Abstract

The purpose of this project is the design of a supersonic aircraft that is able to
meet the market’s requirements, be economically viable and mitigate the current
barriers. The initial requirements of the design have been set according to the
understanding obtained from a brief market research, taking into account the
market needs, in addition to the economical and environmental restrictions. The
conceptual design proposed is a supersonic transport able to execute transatlantic
flights carrying 15 passengers. The aerodynamics, propulsion data and weight of
the design have been estimated using empirical relations and experimental data
found in references. The design has been evaluated regarding its performance,
stability, flying and handling qualities. The relevant models have been created
using the software Matlab, while the flight testing has been executed at the
Merlin MP521 engineering flight simulator. Finally, a discussion is made about
the environmental impact of the supersonic transport, focusing on the aerodynamic
noise, generated by the sonic boom, and the air pollutants emissions.
Contents

1 Introduction 11

2 Conceptual Design 12
2.1 Market Research . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.2 Initial Sizing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.2.1 Desired Requirements . . . . . . . . . . . . . . . . . . . . 14
2.2.2 Initial takeoff weight estimation . . . . . . . . . . . . . . . 14
2.2.3 Mission profile and segments weight fractions . . . . . . . 15
2.2.4 Thrust-to-weight ratio . . . . . . . . . . . . . . . . . . . 17
2.2.5 Wing loading . . . . . . . . . . . . . . . . . . . . . . . . 18
2.2.6 Constraint analysis . . . . . . . . . . . . . . . . . . . . . 19
2.2.7 Initial sizing results . . . . . . . . . . . . . . . . . . . . . 20
2.2.8 Wing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.2.9 Tail . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.2.10 Fuselage . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.2.11 Landing gear . . . . . . . . . . . . . . . . . . . . . . . . 27
2.2.12 Propulsion . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.2.13 Aircraft model . . . . . . . . . . . . . . . . . . . . . . . . 35
2.2.14 Control surfaces . . . . . . . . . . . . . . . . . . . . . . . 37

3 Aerodynamics 38
3.1 Airfoils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.1.1 Airfoil selection . . . . . . . . . . . . . . . . . . . . . . . . 38
3.1.2 Subsonic aerodynamic coefficients . . . . . . . . . . . . . 39
3.1.3 Supersonic aerodynamic coefficients . . . . . . . . . . . . 40
3.2 Subsonic Lift-Curve Slope . . . . . . . . . . . . . . . . . . . . . . 41
3.2.1 Wing - Fuselage Assembly . . . . . . . . . . . . . . . . . . 41
3.2.2 Horizontal Tail . . . . . . . . . . . . . . . . . . . . . . . . 42
3.2.3 Total aircraft . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.3 Supersonic Lift-Curve Slope . . . . . . . . . . . . . . . . . . . . . 44
3.4 Maximum Lift Coefficient . . . . . . . . . . . . . . . . . . . . . . 45
3.4.1 Clean configuration . . . . . . . . . . . . . . . . . . . . . 45
3.4.2 High lift devices . . . . . . . . . . . . . . . . . . . . . . . 48
3.4.3 Horizontal tail . . . . . . . . . . . . . . . . . . . . . . . . 49
3.5 Subsonic Parasite Drag Coefficient . . . . . . . . . . . . . . . . . 50
3.5.1 Equivalent skin-friction method . . . . . . . . . . . . . . . 50

2
3.5.2 Component buildup method . . . . . . . . . . . . . . . . . 50
3.6 Supersonic Parasite Drag Coefficient . . . . . . . . . . . . . . . . 52
3.7 Critical Mach number . . . . . . . . . . . . . . . . . . . . . . . . 56
3.8 Drag due to Lift . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
3.9 Miscellaneous Drag . . . . . . . . . . . . . . . . . . . . . . . . . . 58
3.9.1 Flaps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
3.9.2 Spoilers . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
3.9.3 Landing gear . . . . . . . . . . . . . . . . . . . . . . . . . 60

4 Weights 61
4.1 Weights Estimation Refined Method . . . . . . . . . . . . . . . . 61
4.2 Center of Gravity . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

5 Stability and Control 64


5.1 Subsonic Static Longitudinal Stability . . . . . . . . . . . . . . . 64
5.1.1 Aircraft Pitching Moments . . . . . . . . . . . . . . . . . 64
5.1.2 Subsonic Neutral Point . . . . . . . . . . . . . . . . . . . 65
5.1.3 Longitudinal Control and Trim Analysis . . . . . . . . . . 66
5.2 Supersonic Static Longitudinal Stability . . . . . . . . . . . . . . 69
5.2.1 Supersonic Neutral Point . . . . . . . . . . . . . . . . . . 69
5.2.2 Trim Analysis . . . . . . . . . . . . . . . . . . . . . . . . . 69
5.3 Longitudinal Center of Gravity Location . . . . . . . . . . . . . . 70
5.4 Directional Stability . . . . . . . . . . . . . . . . . . . . . . . . . 71

6 Performance 73
6.1 Climb Performance . . . . . . . . . . . . . . . . . . . . . . . . . . 73
6.1.1 Minimum Time to Climb . . . . . . . . . . . . . . . . . . 74
6.1.2 Minimum Fuel to Climb . . . . . . . . . . . . . . . . . . . 76
6.2 Range . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
6.3 Descent and Loiter . . . . . . . . . . . . . . . . . . . . . . . . . . 77
6.4 Takeoff . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
6.5 Landing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
6.6 Total Mission . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

7 Test Flight 82
7.1 Simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
7.2 Flying and Handling Qualities . . . . . . . . . . . . . . . . . . . . 83
7.2.1 Modes excitation . . . . . . . . . . . . . . . . . . . . . . . 83
7.2.2 Dynamic Stability Requirements . . . . . . . . . . . . . . 85
7.2.3 Longitudinal Dynamic Stability . . . . . . . . . . . . . . . 87
7.2.4 Lateral-Directional Dynamic Stability . . . . . . . . . . . 90

8 Environmental Impact 95
8.1 Sonic Boom . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
8.2 Air Pollution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
8.2.1 Air Pollutants Identification . . . . . . . . . . . . . . . . . 99
8.2.2 Environmental Concerns of Supersonic Flight . . . . . . 101

3
9 Discussion - Conclusions 107

4
List of Figures

2.1 Mission profile division. . . . . . . . . . . . . . . . . . . . . . . . 16


2.2 Fuselage top view. . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.3 Fuselage side view. . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.4 Fuselage-wing assembly top view. . . . . . . . . . . . . . . . . . . 26
2.5 Fuselage-wing side view. . . . . . . . . . . . . . . . . . . . . . . . 27
2.6 Airbus A320 main landing gear retraction and stowage. . . . . . 27
2.7 Main landing gear logintudinal location [17]. . . . . . . . . . . . . 28
2.8 Diagram of nose landing gear location estimation. . . . . . . . . . 29
2.9 Supersonic air inlets. . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.10 Three-shock external inlet. . . . . . . . . . . . . . . . . . . . . . . 31
2.11 Oblique shock wave. . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.12 Concorde rectangular ramp intakes. . . . . . . . . . . . . . . . . 33
2.13 EJ200 turbofan engine. . . . . . . . . . . . . . . . . . . . . . . . . 35
2.14 Aircraft model top view. . . . . . . . . . . . . . . . . . . . . . . . 35
2.15 Aircraft model side view. . . . . . . . . . . . . . . . . . . . . . . 36
2.16 Aircraft model front view. . . . . . . . . . . . . . . . . . . . . . . 36
2.17 Rudder illustration. . . . . . . . . . . . . . . . . . . . . . . . . . . 37

3.1 NACA 64-006 lift curve for Re = 9·106 (XFOIL). . . . . . . . . . 39


3.2 NACA 64-009 lift curve for Re = 9·106 (XFOIL). . . . . . . . . . 40
3.3 Wing supersonic CN α for taper ratio of 0.2 [13]. . . . . . . . . . . 44
3.4 High aspect wing and airfoil maximum lift coefficient ratio at 0.2
Mach. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.5 Stall angle of attack increment at subsonic Mach numbers of 0.2
- 0.6. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.6 Trailing and leading edge high lift devices. . . . . . . . . . . . . . 48
3.7 Wing TE flaps (red), LE flaps (magenta) and ailerons (cyan). . . 49
3.8 Sear-Haacks body volume distribution [9]. . . . . . . . . . . . . . 54
3.9 Wing-body area rule design [9]. . . . . . . . . . . . . . . . . . . . 54
3.10 Aircraft cross-section area distribution. . . . . . . . . . . . . . . . 55
3.11 Wing critical Mach number in two-dimensional flow. . . . . . . . 57
3.12 Wing control surfaces (blue) and spoilers (red). . . . . . . . . . . 59

5.1 Wing-body and tail mean aerodynamic centers [28]. . . . . . . . 65


5.2 CG position influence on Cm at 0.5 Mach. . . . . . . . . . . . . . 66

5
5.3 Variation of δt to trim with the flight speed and the static margin
at SL flight. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
5.4 Variation of δt to trim with the flight speed and the flight altitude
for the subsonic Mach number range. . . . . . . . . . . . . . . . . 68
5.5 Variation of αtrim with the flight speed and the flight altitude for
the subsonic Mach number range. . . . . . . . . . . . . . . . . . . 68
5.6 Variation of δt to trim with the flight speed and the flight altitude
for the supersonic Mach number range and static margin of 0.1. . 69
5.7 Variation of δt to trim with the flight speed and the flight altitude
for the supersonic Mach number range and static margin of 0.35. 70
5.8 Variation of αtrim with the flight speed and the flight altitude for
the supersonic Mach number range. . . . . . . . . . . . . . . . . . 70

6.1 SEP contours diagram (dry thrust). . . . . . . . . . . . . . . . . 74


6.2 Flight path for minimum time to climb at cruise conditions. . . . 75
6.3 Flight path for minimum fuel to climb at cruise conditions. . . . 76
6.4 Illustration of takeoff path and distance. . . . . . . . . . . . . . . 78
6.5 Illustration of landing path and distance [13]. . . . . . . . . . . . 80

7.1 Short-period mode frequency requirements [39]. . . . . . . . . . . 86


7.2 Elevator impulse input flight recording of the short period mode
for 0.6 Mach at 30 kft. . . . . . . . . . . . . . . . . . . . . . . . . 88
7.3 Body axis pitch rate flight recording of the short period mode for
0.6 Mach at 30 kft. . . . . . . . . . . . . . . . . . . . . . . . . . . 88
7.4 Elevator step input flight recording of the phugoid mode for 0.6
Mach at 30 kft. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
7.5 True airspeed flight recording of the phugoid mode for 0.6 Mach
at 30 kft. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
7.6 Euler roll angle flight recording of the spiral mode for 0.6 Mach
at 30 kft. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
7.7 Aileron input flight recording of the roll subsidence mode for 0.6
Mach at 30 kft. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
7.8 Euler roll angle flight recording of the roll subsidence mode for
0.6 Mach at 30 kft. . . . . . . . . . . . . . . . . . . . . . . . . . . 92
7.9 Body axis roll rate flight recording of the roll subsidence mode for
0.6 Mach at 30 kft. . . . . . . . . . . . . . . . . . . . . . . . . . . 92
7.10 Rudder input flight recording of the dutch roll mode for 0.6 Mach
at 30 kft. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
7.11 Body axis roll rate flight recording of the dutch roll mode for 0.6
Mach at 30 kft. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
7.12 Body axis yaw rate flight recording of the dutch roll mode for 0.6
Mach at 30 kft. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

8.1 Drawing and specifications of the baseline configuration [41]. . . 96


8.2 Drawing and specifications of the low boom configuration [41]. . 97
8.3 Three view of a low boom SBJ concept [44]. . . . . . . . . . . . . 98
8.4 Air pollutants formation [48]. . . . . . . . . . . . . . . . . . . . . 100

6
8.5 Atmosphere ozone concentration and temperature till the altitude
of 100,000 ft [54]. . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
8.6 NO x emissions index for SST design (red) and for a typical subsonic
long haul airliner (blue). . . . . . . . . . . . . . . . . . . . . . . . 105

7
List of Tables

2.1 Summary of aircraft requirements. . . . . . . . . . . . . . . . . . 14


2.2 Summary of aircraft specifications. . . . . . . . . . . . . . . . . . 21
2.3 Summary of wing basic dimensions. . . . . . . . . . . . . . . . . . 23
2.4 Summary of tail basic dimensions. . . . . . . . . . . . . . . . . . 24
2.5 Fuselage basic dimensions. . . . . . . . . . . . . . . . . . . . . . . 26
2.6 Properties of the designed 4-shock external inlet for freestream
Mach number M1 =1.7. . . . . . . . . . . . . . . . . . . . . . . . . 32
2.7 EJ200 engine specifications. . . . . . . . . . . . . . . . . . . . . . 34
2.8 Propulsion system basic dimensions. . . . . . . . . . . . . . . . . 35
2.9 Aircraft exposed and wetted areas. . . . . . . . . . . . . . . . . . 36

3.1 NACA 64-006 aerodynamic coefficients ( Re = 9·106 ). . . . . . . 39


3.2 NACA 64-009 aerodynamic coefficients ( Re = 9·106 ). . . . . . . 40
3.3 Lift-curve slope of wing - body assembly for subsonic Mach number
range. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.4 Lift-curve slope of horizontal tail, downwash angle derivative and
horizontal tail CLα contribution to the aircraft, respectively, for
subsonic Mach number range. . . . . . . . . . . . . . . . . . . . . 43
3.5 Aircraft lift-curve slope for subsonic Mach number range. . . . . 43
3.6 Wing-body and horizontal tail lift-curve slope for supersonic Mach
number range. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.7 Horizontal tail downwash angle derivative and CLα contribution
to the aircraft, respectively, for supersonic Mach number range. . 45
3.8 Aircraft lift-curve slope for supersonic Mach number range. . . . 45
3.9 Subsonic aircraft maximum lift coefficient and stall angle of attack. 47
3.10 Subsonic horizontal tail maximum lift coefficient and stall angle
of attack. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.11 Subsonic parasite drag coefficients for the stated Mach number
and altitude flight conditions. . . . . . . . . . . . . . . . . . . . . 52
3.12 Supersonic skin-friction drag coefficient component for the stated
Mach number and altitude flight conditions. . . . . . . . . . . . . 53
3.13 Supersonic parasite drag coefficient component for the stated Mach
number and altitude flight conditions. . . . . . . . . . . . . . . . 56
3.14 Supersonic drag-due-to-lift factor. . . . . . . . . . . . . . . . . . . 58
3.15 Drag coefficients of high lift devices during takeoff and landing. . 59

8
4.1 Aircraft weights estimation. . . . . . . . . . . . . . . . . . . . . . 61
4.2 CG vertical location of basic aircraft components. . . . . . . . . . 63

5.1 Directional stability derivatives for the subsonic Mach number


range. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

7.1 Short-period mode damping ratio limits. . . . . . . . . . . . . . . 86


7.2 Phugoid mode stability requirements. . . . . . . . . . . . . . . . . 87
7.3 Minimum time to double amplitude limits for spiral mode. . . . . 87
7.4 Roll subsidence mode time constant limits. . . . . . . . . . . . . 87
7.5 Dutch roll mode damping ratio and frequency limits. . . . . . . . 87
7.6 Variation of short-period pitching mode characteristics with speed
and altitude. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
7.7 Variation of phugoid mode characteristics with speed and altitude. 89
7.8 Variation of spiral mode characteristics with speed and altitude. 91
7.9 Variation of roll subsidence mode characteristics with speed and
altitude. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
7.10 Variation of dutch roll mode characteristics with speed and altitude. 94

8.1 Carbon dioxide emissions of subsonic airliner and SST design


(6050 km distance flown). . . . . . . . . . . . . . . . . . . . . . . 103
8.2 Water vapor emissions of subsonic airliner and SST design (6050
km distance flown). . . . . . . . . . . . . . . . . . . . . . . . . . . 103

9
Nomenclature

A Area, m2 R Range, m
a Acceleration, m/s2 r Yaw rate, deg/sec
AR Aspect ratio r Radius, m
b Span, m Re Reynolds number
C Thrust SFC, g/KN s S Area, m2
c Chord. m T Thrust, N
CD Drag coefficient T Temperature, K
Cf Skin-friction coefficient T Period, sec
CL Lift coefficient t Time, sec
Cm Pitching moment coefficient t Thickness, m
CP Pressure coefficient TR Throttle ratio
D Drag, N V Speed, m s−1
D Diameter,m W Weight, N
d Distance, m x Longitudinal position, m
E Endurance, sec y Span-wise (lateral) position, m
EI Emissions index z Vertical position, m
e Oswald efficiency factor α Angle of attack, deg
g Gravitational acceleration, m/s2 γ Flight path angle, deg
h Height, m δ Control surface deflection, deg
I Area moment of inertia, m4  Downwash angle, deg
i Incidence angle, deg η Elevator deflection angle, rad
K Drag-due-to-lift factor ζ Rudder deflection angle, rad
L Lift, N ζ Damping ratio
L Length, m θ Angle, deg
l Length, m Λ Sweep angle, deg
M Mach number λ Taper ratio
m Mass, kg µ Rolling friction coefficient
ṁ Mass flow, kg/s ξ Aileron deflection angle, rad
n Load factor ρ Density, kg/m3
P Pressure, P a σ Sidewash angle, deg
p Roll rate, deg/sec φ Euler roll angle, deg
q Pitch rate, deg/sec ω Angular frequency, rad/s
q Dynamic pressure, P a

10
1. Introduction

In recent years, the prospect for an supersonic passenger jet operation has
risen again. The progress achieved in jet propulsion engines and the possibility
of composite materials usage at the aircraft structure are the main reasons that
the success of the supersonic flight has become much more feasible nowadays
compared to the past.
NASA has already began to work on the preliminary design for a low-boom
supersonic passenger with the project Quiet Supersonic Technology (QueSST)
[1], which has aroused its interest in the supersonic transportation, after the High
Speed Research (HSR) program that phased out in 1999. Moreover, a newly
founded company, named Boom Technology Inc., has started the development
of a supersonic transport aircraft being able to fly at a speed of 2.2 Mach and
carry up to 45 passengers [2]. The aim of the Boom Technology design is to
achieve fares comparable to the subsonic airliners business class and increase
the aircraft’s utilization and passengers load factors through the reduced seat
capacity of the aircraft, in comparison with the Concorde’s economic failure in
this aspect.
Apart from the economic feasibility for the supersonic passenger jet success,
the environmental restrictions and impact comprise important factors as well.
The overland flight ban, which has been implemented due to the sonic boom,
and the increased air pollutants emissions have been primary concerns as regards
the relevant regulation process and the mission accomplishment limitations.
In this project, an effort is made for a supersonic transport aircraft conceptual
design, which could be viable in the current market, as a result of the increasing
interest on this aeronautical domain. The whole conceptual design process is
demonstrated and the design is evaluated regarding its performance, stability,
and flying and handling qualities. Difficulties and problems that refer to the
design process and the requirements satisfaction are addressed, and improvements,
solutions are implemented or proposed, where possible.

11
2. Conceptual Design

2.1 Market Research


Flying supersonic has always been a very promising and tempting idea. A
flight duration could become two or three times less in comparison to the relevant
ones of subsonic aircraft. After the Concorde retirement in 2003,the supersonic
civil transportation is not existent any more. Although the required technology
to construct supersonic aircraft exists, the economic sustainment of the concept,
as the case of the Concorde, has shown is not feasible [3]. The supersonic
overland flight ban has also imposed one more important constraint. For the
above reasons, an excessive turn of the civil transportation from subsonic to
supersonic is highly improbable at least in the upcoming two or three decades.
However, the need for over-haul flights is expected to increase significantly
during the next twenty years. According to the Airbus forecast about the global
market for the period from 2016 to 2035 an increase of 95 % is anticipated
for the long-haul traffic [6]. Moreover, Boeing’s current market outlook (2015
- 2034) is also forecasting a grow of 5 % annually in the long-haul traffic over
the relevant time period [7]. These forecasts regarding the market growth and
the increased demand in intercontinental flights set a quite promising ground for
the supersonic flight involvement, where its biggest advantage of diminishing the
flight time could be maximally exploited.
Large supersonic transport aircraft, like the Concorde, will more likely not
be successful, since the added cost will be a significant hindering factor for the
common passenger to choose it. For this reason, the supersonic aircraft will
probably recruit exclusively business class passengers from the regular airline
flights, since the ticket price could be in this case competitive. Therefore,
operating full and smaller aircraft could be economically viable and profitable.
Small supersonic airliners carrying about 15 to 25 passengers could allow the
economic success of the concept, since the development risk and expenditure
will be significantly lower, as well as the potential market share [4].
Private individuals who are keen to buy business jets could be potential
customers for small supersonic aircraft too. In the upcoming ten years (2016 -
2025) period, Bombardier forecasts a total of 8300 new business aircraft deliveries,
when 2800 of them would be of medium size [8]. As the majority of them,
according to this prediction, is going to be delivered to North America and
Europe, fast and comfort transatlantic flights with a prestigious aircraft could

12
be an interesting choice for some wealthy individuals, not caring so much for
overall cost. Other potential customers could be government agencies, since
officials could save this way valuable time, and big corporations.
Regarding the overland ban, the solution of planning routes that cover long
overseas distances seems the more realistic. Such routes can be for example
the London – New York or the Paris – New York, which are now two of the
most busy intercontinental routes in operation. Moreover, routes in south-east
Asia are expected to increase over the upcoming years, since Asia is forecasted,
according to Boeing, to be the continent with the biggest demand in new aircraft
delivery with a global share of 38 % [7]. The growth that will be realized in this
area offers also new opportunities for long-haul supersonic flights. An example
is the Dubai – Singapore route, which can be performed almost overseas, by just
making a small detour (less than 5 % in total distance covered) compared to
the direct route [4]. The possibilities given for mostly overseas long-haul flights
decreases the necessity for an extreme low-boom design, which would reduce the
sonic boom sound intensity level into the acceptable range for human hearing.
Developing this kind of design would add extra development costs and would
deteriorate in general the aerodynamic performance of the aircraft, resulting in
reduced fuel efficiency, thus increasing further the cost, which has to be kept
as low as possible. The environmental consequences due to increased pollutant
emission to the atmosphere would be another important factor to be considered.
In conclusion, using all the above observations, the most feasible supersonic
transport design, considering mainly the financial and secondly the technical
limitations, would be the construction of a quite small supersonic aircraft that
could carry 15 passengers and have a range of 7200 km in order to execute
transatlantic flights. The cruise speed would be 1.7 Mach, which is about the
double of the regular cruise speed of a subsonic airliner. Although a higher Mach
number is preferred as more enticing for the clients, the choice of a moderate
speed would result in building an aircraft with reduced weight, having a positive
effect in fuel consumption and noise generation [5].

13
2.2 Initial Sizing
The initial design process includes the specification of the desired requirements
according to the aircraft mission and role. Then, the mission profile sections have
to be defined. According to this mission profile, an initial estimation will be done
for the maximum take-off weight.

2.2.1 Desired Requirements


The desired requirements for the supersonic transport to be designed are
presented in Table 2.1.

Cruise speed 1.7 Mach


Maximum speed 1.7 Mach
Minimum payload mass 1900 kg
Cruise altitude 15 km
Range 7200 km
Loiter time 20 min
Landing distance 3500 m
Takeoff distance (SL) 3500 m
Thrust specific fuel consumption (cruise) 26.9 g/KN s

Table 2.1 – Summary of aircraft requirements.

The payload mass corresponds to the 15 passengers plus the pilot, the co-pilot
and two crew members. An average of 100 kg is assigned for each person on-board
including the luggages.

2.2.2 Initial takeoff weight estimation


The first step is to make an initial estimation of the takeoff weight (W0 ).
The takeoff weight consists of the empty weight (We ), the payload weight (Wp )
and the fuel weight (Wf ) as shown in equation (2.1). The payload weight is
defined by the requirements, when the fraction of empty weight over the takeoff
weight is given by the empirical relation (2.2) found in reference [9].

We
W0 = Wp + Wf + W0 (2.1)
W0

0.06  −0.05
We T W0

= 0.32 + 0.66m−0.13
0 AR0.3 0.05
Mmax in fps units (2.2)
W0 W0 S

The fuel weight is computed after calculating the weight fractions for the
different mission segments, since the total aircraft weight drop corresponds only
to consumed fuel weight. The final fuel weight is then increased for safety reasons

14
by 6 %. The relevant weight fractions are calculated according to estimations
derived from historical data or with approximating relations. For each mission
segment i, the weight of the burned fuel is given by the relation

Wi
 
Wf i = 1− Wi−1 (2.3)
Wi−1

while the total amount of fuel for all n segments of the mission is estimated as

n
X
Wf = 1.06 Wf i (2.4)
i=1

Finally, the take-off weight can be computed using an iterative method for
solving the equations (2.1), (2.2) and (2.4), making an initial estimation for W0 ,
which will then converge to the actual value.
In order to estimate the empty weight faction (We /W0 ), the aspect ratio
(AR,), the wing loading (W0 /S) and the thrust-to-weight ratio (T /W0 ) have
to be defined. For the supersonic transport a low aspect ratio of 3.2 is selected,
since a long wing would result in a large value of the wave drag. Moreover, it is
not feasible to construct a light wing which could be structurally strong enough
to resist the relevant bending and torsional stresses in such high airspeeds and
to be free of the aeroelastic phenomena effects.
The thrust-to-weight ratio and the wing loading will be defined in the relevant
following subsections. It should be stated here that the selection of the appropriate
values for the two aforementioned parameters is a compromise. A large wing
loading is more preferred for the supersonic aircraft, which results in smaller
values of thrust-to weight ratio and thus lower thrust requirements. However,
the wing loading is limited from a reasonably low stall speed requirement and
consequently from the landing and takeoff distance restrictions.

2.2.3 Mission profile and segments weight fractions


In Figure 2.1 the different segments of the mission profile for the transport
aircraft are presented. The typical mission of the aircraft includes:

• taxi, warm-up and takeoff, 1


• climb, 2
• cruise, 3
• loiter, 4
• descend, 5
• landing and taxi back, 6

15
Figure 2.1 – Mission profile division.

The weight fractions for the different mission segments are calculated using
the following relations [9]

W1
= 0.97 (2.5)
W0

W2 2
= 0.991 − 0.007Mcruise − 0.01Mcruise (2.6)
W1

RC

W3
= e V (L/D)cruise (2.7)
W2

EC

W4
= e V (L/D)max (2.8)
W3

W5
= 0.99 (2.9)
W4

W6
= 0.992 (2.10)
W5

For the weight fraction estimation (eq. 2.7) of the cruise phase, the Breguet
range equation is used. The specific fuel consumption C for the cruise is the one
defined from the requirements before, while for the loiter the typical value of
the 22.7 g/KN s has been used [9]. The maximum lift to drag ratio (L/Dmax ) is
equal to the relevant value of the loiter phase for a jet aircraft. The correspondent
value for the condition of the most efficient cruise is 86.6% of the L/Dmax for a
jet aircraft [9] and is approximated using the relation [11]

L
 
= 11M −0.5 (2.11)
D cruise

16
2.2.4 Thrust-to-weight ratio
Cruise
The required T /W during cruise can be simply approximated using the rigid
equations of motion for steady flight. In this case, the thrust should equal the
drag and the lift should equal the weight, assuming that the thrust installation
angle is zero (thrust vector aligned with the velocity vector) and disregarding
the angle of attack influence. Thus
−1
T L
  
= (2.12)
W cruise D cruise

The above equation is a good initial approximation, although it underestimates


the drag, since during the level and unaccelerated flight the angles of attack
experienced are quite small.

Takeoff
Having estimated the thrust-to-weight ratio required for the cruise condition,
the correspondent value for takeoff can be calculated with the relation

T T Wcruise TT O
      
= (2.13)
W TO W cruise WT O Tcruise
where

Wcruise W2 W1
    
= = βc (2.14)
WT O W1 W0

The fraction (TTO /Tcruise ) can be computed using the installed engine thrust
lapse equations for maximum thrust presented in [10], for cruise at the desired
altitude and for sea level flight. These equations apply to the expected performance
of the advanced engines in the 2000 era and beyond. The flight altitude in these
relations are introduced as the ratio between the static temperature (θ) and the
static pressure (δ), respectively, at the flight altitude to the corresponding values
at sea level, as shown in equations (2.15) and (2.16).

θ = Talt /TSL (2.15)

δ = Palt /PSL (2.16)

The nondimensional temperature (θ0 ) and pressure (δ0 ) are defined using the
correspondent total temperatures and pressures at flight altitude, respectively,
as

γ−1 2
 
θ0 = T talt /TSL =θ 1+ M (2.17)
2

17
γ
γ−1 2
 
γ−1
δ0 = P talt /PSL =δ 1+ M (2.18)
2

The specific heat ratio (γ is equal to 1.4 for atmospheric air. The thrust
lapse (α) is defined as the maximum thrust at flight altitude over the maximum
thrust at sea-level flight and can be calculated from the equations (2.19) and
(2.20). For a low-bypass ratio turbofan engine, the so-called dry thrust lapse,
without the usage of afterburner, is estimated as
(
0.6δ0 for θ0 ≤ T R
αdry = (2.19)
0.6δ0 (1 − 3.8(θ0 − T R)/θ0 ) for θ0 > T R

and the so-called wet thrust lapse, using the afterburner, is estimated as
(
δ0 for θ0 ≤ T R
αwet = (2.20)
δ0 (1 − 3.5(θ0 − T R)/θ0 ) for θ0 > T R

The term TR appeared in the above equations is the throttle ratio, which
is equal to the so-called value θ0 ,break [10]. These equations are valid for values
of the theta break greater than one. The theta break relates two important
constraints of the turbine engine, the compressor pressure ratio and the turbine
inlet temperature, being actually the point where the engine control system must
switch from limiting the compressor pressure ratio to limiting the turbine inlet
temperature. The θ0 ,break should be chosen by the designer so that it provides the
best balance of engine performance over the expected range of flight conditions.
In this case of the supersonic transport, the throttle ratio should be greater than
one in order to fulfill the supercruise requirement. For this reason, it is set to
1.2, so that it operates closer to the optimal value of θ0 , as given by the cruise
condition, thus sustaining thrust to higher values of Mach number.

2.2.5 Wing loading


Having estimated the thrust-to-weight ratio at takeoff, the wing loading will
be estimated for the most critical conditions of the mission, which are the takeoff,
stall and landing. After computing the relevant values, the minimum wing
loading obtained will be used to calculate the wing reference area.

Stall
For the calculation of the wing loading for the stall condition, an initial
maximum lift coefficient (CLmax ) has to be estimated. For a supersonic aircraft
having a low aspect ratio wing and a large leading edge sweep, the value for
CLmax would be quite small. In this case, a CLmax of 1.6 will be assumed, which
is a typical value for supersonic wings. Since a moderate value for CLmax can
only be obtained for the supersonic aircraft configuration, in order to avoid a

18
low wing loading resulting in a large wing area, a very low stalling speed cannot
be achieved. However, it should be kept as low as possible so that the landing
distance requirement can be fulfilled too. For this reason, a stalling speed of 68
m/s has been selected at sea-level flight. Therefore the wing loading at stall can
be calculated as

W 1
 
2
= ρSL Vstall CLmax (2.21)
S stall 2

Takeoff
The wing loading at takeoff is given from the relation [9]

W T
   
= (T OP )σCLT O in fps units (2.22)
S TO W TO

where σ is the air density ratio (density at takeoff altitude over density at
sea-level), set in this case equal to one, assuming operation from airports at
almost zero altitude, like the Heathrow in London and the John F. Kennedy
in New York. The takeoff speed is specified from the FAR Part 25 regulations
for civil aircraft as 1.1 times the stall speed. Hence, the takeoff lift coefficient
(CLTO ) is obtained by dividing the maximum lift coefficient (CLmax ) with 1.21.
The takeoff parameter (TOP) can be obtained solving equation (2.22), where
increasing the TOP results in a requirement for increased takeoff distance. In
this case, the TOP should be chosen so that the aircraft fulfills the takeoff
distance requirement, while keeping the wing loading as high as possible.

Landing
The maximum wing loading at landing will be estimated using the maximum
landing distance requirement. A relation that connects the two parameters is
the following [9]

W 1 dland
   
= − da σCLmax in fps units (2.23)
S land 80 1.67

where the landing distance (dland ) is divided by 1.67 to provide the required
safety margin set by regulations (FAR 25) and da is the obstacle-clearance
distance, which is equal to 1000 ft for an airliner-type aircraft.

2.2.6 Constraint analysis


After estimating the thrust-to-weight ratio and the wing loading of the aircraft
for the aforementioned conditions, two important constraints, which relate the
two parameters, will be evaluated so that the final values for both parameters
are specified. These parameters are the takeoff distance and the supercruise

19
requirements. The two conditions will be evaluated using the relevant relations
from reference [10].

Takeoff distance
An initial approximation for takeoff distance, consisting of the ground roll
and the rotation distance, can be made with relation (2.24) assuming that the
thrust force is much larger than the resistance forces in the ground roll

!
kT2 O W

dT O =
gρSL CLT O αdry (TSL /WT O ) S TO
s ! s (2.24)
2 W

+ tr kT O in fps units
ρSL CLT O S TO

where kTO is the ratio of takeoff speed over stall speed (1.1 as defined in the
previous subsection) and tR the total rotation time needed for the aircraft to
move from the flat to the nose-up attitude needed for takeoff (normally 3 sec).
The wet thrust lapse parameter can be estimated from the equation (2.20).

Supercruise
An important restriction for the supersonic transport is that it should cruise
at the desired supersonic Mach number and altitude, without using the afterburner.
This is the so-called supercruise condition and it is essential in order to achieve
economically feasible supersonic flights at long range. The supercruise condition,
stated here, estimates the required thrust-to-weight loading needed at takeoff,
which is calculated as

!
TSL βc βc W qCD0
   
= K + in fps units (2.25)
WT O αdry q S TO βc (W/S)T O

where αdry and βc are given from the equations (2.19) and (2.14) respectively
and q is the dynamic pressure for the given cruise flight speed and altitude.
The zero lift drag coefficient (CD0 ) and the K coefficient of the lift-drag polar
equation (these two coefficients will be analyzed in more detail later on) are
initially approximated using typical values for the specific cruise conditions as
0.03 and 0.3 respectively.

2.2.7 Initial sizing results

The results obtained from the initial sizing analysis are presented in Table
2.2. The thrust-to-weight ratio corresponds to the maximum value obtained,
which is the one satisfying the supercruise condition.

20
Empty weight fraction 0.4715
Fuel weight fraction 0.4641
Takeoff mass 29499 kg
Wing loading (takeoff) 5110 N/m2
Wing loading (stall) 4539 N/m2
Wing loading (landing) 5629 N/m2
Maximum thrust-to-weight ratio 0.5851
Takeoff distance (SL) 2283 m

Table 2.2 – Summary of aircraft specifications.

2.2.8 Wing

Wing geometry
In this section, the geometry of the wing will be defined. The choice for
the wing is the trapezoidal with aspect ratio of 3.2, as stated previously. Other
parameters to be defined are the leading edge wing sweep (ΛLE ) and the taper
ratio (λ).
The wing sweep is being used in transonic flow in order to increase the critical
Mach number (the freestream Mach number at which the local Mach number
on the aircraft first reaches the sonic speed) and in supersonic flow in order to
decrease the loss of lift, associated with supersonic flight. Another important
feature for using swept wings at supersonic flows is the delay of the appearance
of aeroelastic divergence, which is critical for a light weight design for such high
values of dynamic pressure. For this aircraft configuration a LE wing sweep
angle of 45 deg has been chosen, which is although a bit larger compared to
the Mach cone angle (arcsin(1 /M )), so that the wing is capable of creating the
necessary lift for satisfying the takeoff and landing distance requirements.
The usage of a swept wing results in the requirement of having a highly
tapered wing in order to preserve the desired elliptical lift contribution over
the wing. Producing a lift distribution over the wing that resembles the ideal
elliptical one has the important effect of reducing the lift induced drag. However,
a very low taper ratio has the consequence of tip stalling tendency. Moreover,
it can be limited by requirements about adequate chord near wing tips for the
ailerons placement. Recommended values for highly swept wings in supersonic
flow are 0.2 - 0.3 [9][16]. For practical and structural reasons, the lower value of
0.2 is chosen, since results in a larger chord near wing root, which is needed for
engines placement, as well as grater internal volume due to increased maximum
thickness, making it possible to house the landing gear and fuel tanks.
The aircraft will have a low-wing configuration mainly for the practical reason
of placing the landing gear. A high-wing configuration is not feasible, since
for low aspect ratio supersonic wings where thin airfoils are used, there is not

21
enough space for their housing. An external blister would be unacceptable,
since it would increase the drag significantly. A mid-wing configuration is not
used in passengers aircraft for structural reasons, since the loads have to be
carried across the fuselage, reducing the internal usable volume of the fuselage
significantly. A low-wing configuration usually is accompanied with a dihedral
angle. However, the high wing sweep is contributing to the dihedral effect too,
creating a quite high effective dihedral angle, which diminishes the need of having
a wing geometric dihedral. For this reason, an initial zero dihedral angle will be
considered for this design.

Wing sizing

The first step of the wing sizing is to calculate the wing area. This can be
simply done using the estimation for the minimum wing loading in the following
equation

m0 g
SW = (2.26)
(W/S)min

Knowing the wing area, aspect ratio and taper ratio of the chosen trapezoidal
wing, the expressions about some important wing properties are presented.
Wing span:

bW = AR · S (2.27)

Chord at root:

2S
croot = (2.28)
b(1 + λ)

Chord at tip:

ctip = λcroot (2.29)

Mean aerodynamic chord (MAC):


!
2 1 + λ + λ2
c̄W = croot (2.30)
3 1+λ

Spanwise location of MAC with respect to the aircraft longitudinal axis of


symmetry:

b 1 + 2λ
 
ȲW = (2.31)
6 1+λ

22
Wing dimensions
In table 2.3 the basic wing dimensions are presented for the initial design.

Area 63.7487 m2
Span 14.2827 m
Root chord 7.4389 m
Tip chord 1.4878 m
Mean aerodynamic chord 5.1246 m
Spanwise location of MAC 2.7772 m

Table 2.3 – Summary of wing basic dimensions.

2.2.9 Tail
The tail area can be calculated using the tail volume coefficients presented in
[9] for the horizontal and the vertical tail. These coefficients are used to relate
the wing to the tail size, so that an initial estimation for the vertical (VT) and
horizontal (HT) tail area can be made from the following equations.

VV T bW SW
SV T = (2.32)
LV T

VHT c̄W SW
SHT = (2.33)
LHT

where VV T , VHT the volume coefficient, and LV T , LHT the moment arm for the
vertical and the horizontal tail respectively. Typical values for the cV T and cHT
are provided in [9]. The value selected for the horizontal tail is 0.4, which is
smaller than the recommended typical value, since the choice of a whole-moving
surface has been made in order to decrease the horizontal tail volume. Decreasing
the horizontal tail volume is critical in order to achieve a area-ruled design and
decrease the wave drag. The volume coefficient for the vertical tail has been set
to the typical value of 0.09 for the jet transport. The tail arm moment L , which
is initially approximated as the longitudinal distance of the quarter position of
the wing mean chord to the quarter mean chord position of the tail, has been
set to 40% of the fuselage length for both the vertical and horizontal tail arm.
In order to model the tail geometry, typical values have to be used for its
geometric properties. The horizontal stabilizer has been modeled as a straight
trailing edge tapered wing, with a sweep angle at the leading edge of 50 deg,
which is typically 5 deg larger than the wing sweep, so that it experiences a
greater critical Mach number compared to the wing. The vertical stabilizer has
been modeled as a trapezoidal wing having a 60 degrees leading edge sweep
angle. The selected values for the aspect ratio are 2.5 for the horizontal and 1.3

23
for the vertical stabilizer, while the taper ratio is 0.15 and 0.2 for the horizontal
and the vertical stabilizer respectively.
For computing the rest of the geometric properties of the horizontal and
vertical stabilizer, like the mean aerodynamic chord, the span etc., the expressions,
which have been presented for the wing in the previous subsection, can be used.

Tail dimensions
In table 2.4 the basic wing dimensions are presented for the initial design.

Horizontal stabilizer Vertical stabilizer


Area 12.0995 m2 7.5875 m2
Span 5.4999 m 3.1407 m
Root chord 3.8386 m 4.0265 m
Tip chord 0.5613 m 0.8053 m
Mean aerodynamic chord 2.6068 m 2.7738 m

Table 2.4 – Summary of tail basic dimensions.

2.2.10 Fuselage

Fuselage dimensions
The fuselage has been dimensionalized taking two important factors into
consideration. Firstly, the existence of enough usable volume to host the passengers
and secondly, the area ruling design, which will be further analyzed during the
wave drag calculation. Briefly the fuselage has to be squeezed at the area where
the wing is placed (”coke-bottling” design), so that a smooth volume distribution
and a smaller cross-sectional area is achieved.
An initial estimation for the fuselage length for a jet transport is given using
the relation (2.34) based on statistical data [9].

Lf us = 0.287m00.43 (2.34)

The typical compartment properties for first class seats are seat pitch of 1
m and seat width of 60 cm, where the aisle width should be about 60 cm and
its height more than 193 cm [9]. Thus, the maximum diameter has been set to
2.6 m, which is a value providing enough space, while keeping the cross-sectional
area as low as possible. The length of the fuselage has been enlarged by about
3 m compared to the stimation given from equation (2.34). This is done mainly
for two reasons. While the fuselage maximum diameter is enough to host two
passengers sitting side by side, for the portion of fuselage which is squeezed
this would not be possible and just one passenger could be hosted there. The

24
need to place also a lavoratory increases the compartment’s length requirements.
Additionally, the typical cabin compartment to overall length ratio for a supersonic
transport is about 0.55 [16], thus the lengthened fuselage should provide enough
space for hosting all the 15 passengers. Finally, the supersonic aircraft has a
sharper and longer nose in comparison to the subsonic aircraft, and a typical
value of its nose length to diameter ratio is about 4 [16]. The reason for that is
to avoid strong bow shock waves forming at the fuselage nose, which lead to a
significant increment of the wave drag. However, having a very low nose length
to diameter ratio would create practical problems with the cockpit housing and
the unhindered pilots’ visibility, especially during landing.

Fuselage lofting

The designed fuselage consists of three parts: the conical nose, the main
cylindrical fuselage and the fuselage tail. The main cylindrical fuselage consist
of three parts. The first one is a cylinder assembled with the nose till it reaches
the maximum diameter, the second part is a cylinder of the maximum diameter
but squeezed in the lateral direction (the fuselage width is manipulated while
the height remains constant) and the final part which is a cylinder assembled
with the fuselage tail. The fuselage has been lofted in Matlab and its model can
be seen in figures 2.2 and 2.3.

Figure 2.2 – Fuselage top view.

Figure 2.3 – Fuselage side view.

As it can be seen in fig. 2.3, the tail fuselage is converging towards the
higher side, so that the empennage can be placed higher up of the wing and
thus minimize the aerodynamic interactions due to the wing wake. The basic
dimensions of the fuselage lofting are presented in table 2.5. The fuselage total
volume has been computed to be 78.75 m3 .

25
Nose length 4 m
Nose maximum diameter 1.8 m
Cylindrical part length 19 m
Cylindrical part maximum diameter 2.6 m
Squeezed part minimum width 1.56 m
Fuselage tail length 4 m
Fuselage tail maximum diameter 2 m

Table 2.5 – Fuselage basic dimensions.

Fuselage-wing assembly

A low-wing configuration has been chosen for this aircraft design. The
position of the quarter mean aerodynamic chord, which corresponds to the
aerodynamic center in subsonic flight, has been placed in the longitudinal direction
14.07 m rear of the fuselage nose. The fuselage-wing assembly, incorporating the
area ruled design, can be seen in figures 2.4 and 2.5.

It has to be mentioned that the wing has been placed in that position due to
two factors. The first one is to have a long enough tail arm moment (L), which
results in a smaller tail size and demands the placement of the wing forward.
However, as it will be observed in the following subsection, the placement of the
wing too much forward, will have the consequence of a much longer and heavier
landing gear and perhaps an unacceptably high load acting on the nose landing
gear.

Figure 2.4 – Fuselage-wing assembly top view.

26
Figure 2.5 – Fuselage-wing side view.

2.2.11 Landing gear


In order to meet the requirements for the airframe layout and to avoid any
increment in the aerodynamic drag, the way of how the landing gear will be
stowed has to be decided. First of all, the landing gear arrangement has been
chosen to be the common ”tricycle” gear, with two main landing gears aft of the
aircraft center of gravity (CG) and one nose landing gear forward of the aircraft
CG. The number of wheels per strut are typically dependent on the aircraft
takeoff weight. For aircraft weighing 60,000 to 175,000 lbs, such as the current
design, two tires per struts are recommended [17]. The usage of two tires is also
preferable for safety reasons, like in the case of a flat tire.
One issue that is faced with the supersonic aircraft is the existence of very
thin wings, which does not give the opportunity for the main landing gear to be
stowed inside the wing easily, especially for configurations with two tires per strut
like the current. For this reason, the strut will be stowed inside the wing, while
the tires will be stowed inside the fuselage, after retraction, where the available
volume is much bigger (inward retraction). The aforementioned concept to be
used is visualized in fig. 2.6 for a subsonic airliner.

Figure 2.6 – Airbus A320 main landing gear retraction and stowage.

27
The main landing gear length and location can be determined using the
conceptual design method from reference [17]. The initial recommended location
for the main landing gear is about 55 % of the wing MAC. The main landing gear
location is then adjusted using the 15 deg angle rule in the static position shown
in fig. 2.7. For the current design, the correspondent location was estimated to
be at about 15.6 m aft of the fuselage nose. The length then should be set so
that the tail does not hit the ground during landing. Using the recommended 12
deg angle for fuselage tail tipping avoidance, which should not be much smaller
than the stall angle, the main landing gear length can be approximated as 2.45
m.

Figure 2.7 – Main landing gear logintudinal location [17].

The nose landing gear has been selected to be a forward retracting, which
is preferred since in the case of a failure the gravity and air drag assists the
gear to reach the down-and-locked position. The nose landing gear length
is initially set to be equal to the main landing gear length for the low-wing
configuration. The longitudinal position of the nose landing gear depends on the
aircraft position of the CG. For a first approximation the most aft location of
the CG is set just forward of the aerodynamic center (quarter-chord of MAC).
Assuming its longitudinal distance from the main landing gear location to be 12
m, the correspondent static loads on the main and the nose gear can be estimated
[17].

M ax static main gear load (per strut) % = (B − Bm )/2B (2.35)

M ax static nose gear load % = (B − Bn )/2B (2.36)

28
where the relevant distances are described in fig. 2.8. For the given distances,
the landing gear load is about 43.5 % per main landing gear strut and 13 % for
the nose landing gear strut, a loading contribution which is not optimal (8-10 %
preferable static loading for nose landing gear considering the most aft location
of the CG) but within the acceptable range for the initial layout.

Figure 2.8 – Diagram of nose landing gear location estimation.

The main landing gear tires can be sized too, using the obtained results and
the following statistical relations for transport aircraft [9]. Their diameter and
width are estimated from the equations (2.37) and (2.38) as 84.1 cm and 26.3
cm respectively.

0.315
Dtire = 5.3mtire (2.37)

wtire = 3.9m0.48
tire (2.38)

where mtire is the load per tire or wheel in kg.

2.2.12 Propulsion
For the propulsion system, a similar concept to Concorde will be followed,
with each engine mounted at the lower side of the wing. The engine location
is a quite complex task, mainly due to the engine intake, which especially for
supersonic flows has to be designed very carefully in order to control and provide
the airflow needed for the engine operation. Other things that deteriorate the
engine performance is the location of the engine inlet in a position where it
ingests distorted airflow, such as in the wing wake, which can cause stall to
the compressor. For supersonic fighters the most common configuration is to
use fuselage side-mounted inlets. This is a location that provides short ducts
and relatively clean air [9]. However, this is not possible for the low-wing
configuration transport aircraft for practical reasons, for instance due to the
need of placing the engines outside the fuselage and the need of landing gear

29
stowage. Using a configuration like the Concorde, has the advantages of having
a quite undisturbed airflow, of avoiding interferences between the two exhaust
nozzles for not being placed side by side, and for decreasing the structural stresses
on the wing, due to the inertial relief effect (the wing-root shear force can be this
way significantly reduced). A disadvantage of this configuration comes from the
spatial requirement of mounting the engines directly at the wing, since neither
the engine nor the inlet can be too long. Moreover, the nozzle placement at
the wing trailing edge will reduce the usable wing area for the placement of the
respective control surfaces (flaps, ailerons).

Air intake
In order to have a proper operation for the compressor blades, the airflow has
to be decelerated to about 0.4 Mach. The air intakes for supersonic aircraft are
sophisticated devices, which need to decelerate the supersonic flow to subsonic
through shock waves. The efficiency of the intake is expressed from the reduction
in total pressure loss, which is mainly dependent of the shock-wave pattern [12].
The air intakes used for supersonic flight are divided in two parts: the converging
supersonic inlet and the diverging subsonic diffuser. The principles followed for
a conceptual layout of the intake are presented in the following subsections.

1. Supersonic inlet
The choice of the supersonic inlet type is mainly determined from the design
Mach number. A ”pitot” or normal shock inlet, decelerating the supersonic flow
to subsonic through a single normal shock wave, results in big total pressure
losses for higher Mach numbers, which decrease significantly the inlet efficiency.
For example, for Mach number 1.7, which is the one of interest for this design,
the pressure recovery is just 85.57 %. In this type of intake the maximum inlet
mass flow is achieved when the normal shock is located right at the cowl lip
position (fig. 2.9(a)).

(a) Pitot or normal shock inlet (b) Internal shock inlet

Figure 2.9 – Supersonic air inlets.

Another type of air inlet is the internal compression, which is a ”two -

30
dimensional ramp” inlet. When this inlet operates at optimal design condition,the
flow is decelerated by two oblique shocks existing at the entrance of the facing
ramps, before a normal shock takes place at the minimum section position inside
the intake (fig. 2.9(b)). However, when operating at off-design conditions it is
possible the normal shock to take place outside the inlet (inlet not ”started”), a
situation that can stall the engine [9]. Moreover, the boundary layer on the inlet
walls can tolerate only a very modest adverse pressure gradient before separating,
a fact that leads to a smooth decrease of the inlet area and consequently to a
longer and heavier intake [23].
The type of supersonic inlet to be used in this design is the external compression
rectangular ramp inlet, where the weak oblique shocks are generated outside the
inlet and are followed by a normal shock at the cowl lip (fig. 2.10). In order to
design an efficient intake, the number of oblique shocks and the relevant wedge
angles have to be chosen. The external shock intake length can be initially
approximated using the shock angles computed from the correspondent ramp
angles. The cowl lip is then located just aft of the shocks [9]. In this case, a 5
% elongation of the computed external inlet length has been adopted as a safety
margin.

Figure 2.10 – Three-shock external inlet.

The inlet efficiency can be optimized through the appropriate selection of the
wedge angles. In order to achieve very high efficiency, the inlet will incorporate
three weak oblique shocks before the normal shock. The Mach number (M ), the
wedge angle (δ), the shock angle (θ) and the total pressure recovery (Pt2 /Pt1 ),
are related using the equations (2.39), (2.40) and (2.41) for compressible, adiabatic
flow of a perfect gas (for air γ=1.4) [18]. The correspondent notation can be
seen in fig. 2.11.

s
36M14 sin2 θ − 5(M12 sin2 θ − 1)(7M12 sin2 θ + 5)
M2 = (2.39)
(7M12 sin2 θ − 1)(M12 sin2 θ + 5)

M12 sin 2θ − 2 cot θ


tan δ = 5 (2.40)
10 + M12 (7 + 5 cos 2θ)

31
!7/2  5/2
Pt2 6M12 sin2 θ 6
= (2.41)
Pt1 M12 sin2 θ + 5 2
7M1 sin2 θ − 1

Figure 2.11 – Oblique shock wave.

The above equations can be used for the final normal shock too, setting θ=90
deg. In table 2.6, the wedge angles and the rest of the results of the four-shock
external compression inlet design are presented.

Wedge angle (deg) Mach number Shock angle (deg) Pressure recovery (%)
5.8960 1.5955 39 99.96
5.7782 1.4920 42 99.96
6.1309 1.2714 50 99.71
- 0.8009 90 98.40

Table 2.6 – Properties of the designed 4-shock external


inlet for freestream Mach number M1 =1.7.

The total pressure recovery for the designed supersonic inlet reaches 98.04
% for decelerating the airflow from 1.7 to about 0.8 Mach. During off-design
operation, the mass flow demand at lower speeds is reduced. This decrease
results in the static pressure rise at the compressor inlet, which forces the flow
to spill outside the compressor. Since this redundant subsonic flow has to spill
out, the normal shock moves farther upstream from the lip cowl, increasing the
so-called spillage drag significantly [22]. In order to avoid this effect and keep the
normal shock attached at the cowl lip, a by-pass door is opened in the diffuser,
so that the excess air can be thrown away before reaching the compressor. This
technique has been adopted in the Concorde’s air intake, as can be seen in fig.
2.12. It can be also observed in this figure the knife-edge shape of the intake
sections, which is a necessary feature of the supersonic inlet, so that large shock
angles and consequently increased wave drag can be avoided.

32
Figure 2.12 – Concorde rectangular ramp intakes.

Finally, a preliminary estimation can be made for the required capture area
size according to an approximation found in [9], which correlates the capture area
with the Mach number and the mass flow. For the freestream Mach number of
1.9, which corresponds to a safety margin of 0.2 Mach higher than the aircraft
maximum design condition, the capture area (Ac ) is approximated as 0.0054
times the maximum air mass flow (ṁ).

2. Subsonic diffuser
In the subsonic diffuser the airflow is further decelerated from the subsonic
Mach number obtained after the normal shock to 0.4 Mach to meet the compressor
operation conditions. This is achieved through a diverging duct, which means
that the area inside the intake should increase. The throat area to the engine
front face area can be estimated from the relation (2.42), which is used to express
the maximum area ratio between any two stations into the diffuser in terms of
the Mach number, corresponding to maximum air mass flow [23].
!3
Athroat M1 1 + 0.2M02
= (2.42)
Aengine M0 1 + 0.2M12

In this design, the objective is to slow down the flow from 0.8009 to 0.4 Mach
and thus the relevant ratio between the throat and the engine area is computed
as 0.6527, which means that the respective diameter ratio between the throat
and the engine would be 0.8079.
The subsonic diffuser has a typical efficiency of 97 - 98 %, which will result
to a total efficiency of the supersonic air intake of about 95 % for the given
design. In order to avoid high viscous forces on the duct walls and consequently
separation of the boundary layer flow, which would increase the total pressure
losses, the diffuser has to incorporate small expansion angles. The upper limit
of the slope is considered to be 10 deg [23]. Therefore, a slope of 6 deg has been

33
chosen for the diffuser design to ensure high intake efficiency. Knowing the duct
slope and the correspondent throat and engine diameters, an initial estimation
can be made for the subsonic diffuser length.

Engine selection

The main criteria for the engine selection is to provide enough thrust for
the aircraft operation (satisfy the maximum thrust-to-weight ratio), to have a
low thrust specific consumption, so that it it satisfies the requirements, a low
weight and a small size, to avoid increased parasite drag. For satisfying the
fuel consumption requirement, it is preferable to incorporate a low-bypass ratio
afterburning turbofan instead of a turbojet. Another benefit of the turbofan
engine is the reduced noise, which is of big interest especially during take-off.
After a research for the available engines and the relevant manufacturer data,
the engine selected was the EJ200 (fig. 2.13), a turbofan engine that is used
as powerplant for the Eurofighter Typhoon. The EJ200 has been developed
and produced in an international cooperation among Rolls-Royce, Avio, ITP
(Industria de Turbo Propulsores) and MTU Aero Engines. The relevant engine
specifications are shown in table 2.7. The presented data have been retrieved
from one of the engine manufacturers, particularly the MTU Aero Engines
product leaflet [24]. Two EJ200 engines will be incorporated in the supersonic
transport design.

Maximum thrust, reheated 90 kN


Maximum thrust, dry 60 kN
Bypass ratio 0.4
Overall pressure ratio 26:1
Specific fuel consumption, reheated 48 g/kN s
Specific fuel consumption, dry 21 g/kN s
Air flow rate 77kg/s
Length 4m
Maximum diameter 74 cm
Weight 1010 kg

Table 2.7 – EJ200 engine specifications.

Propulsion system dimensions

After having chosen the jet engine to be used for the supersonic aircraft,
the dimensions for the intake and thus for the whole propulsion system can be
quantitatively determined (Table 2.8). For simplicity, an initial rough estimation
of a rectangular nacelle with width equal to 1.05 times the engine diameter will be
incorporated in the design, for the total propulsion system length, excluding the
exhaust nozzle. The relevant ratio of intake length to engine diameter is about

34
3.19. The two nacelles have been placed in distance 2.9 m from the aircraft
longitudinal axis of symmetry.

Figure 2.13 – EJ200 turbofan engine.

Intake length 2.36 m


Capture area 0.4158 m2
Throat diameter 59.8 cm
Nacelle length 5.36 m
Nacelle width 77.7 cm

Table 2.8 – Propulsion system basic dimensions.

2.2.13 Aircraft model


The aircraft model has been created in Matlab, using the aforementioned
dimensions. The airfoils used for the wing and the tail modeling are discussed
in the following chapter. The relevant views of the model can be observed in
figures 2.14, 2.15 and 2.16.

Figure 2.14 – Aircraft model top view.

35
Figure 2.15 – Aircraft model side view.

Figure 2.16 – Aircraft model front view.

In table 2.9, some important geometric properties of the aircraft are presented,
in particular its computed exposed and wetted areas.

Wing exposed area (Sexp ) 48.2012 m2


Wing wetted area (Swwet ) 88.9441 m2
Fuselage wetted area (Sfwet ) 164.1938 m2
Horizontal tail wetted area (SHTwet ) 11.2816 m2
Vertical tail wetted area (SVTwet ) 15.5985 m2
Nacelles wetted area (Snacwet ) 26.8351 m2

Table 2.9 – Aircraft exposed and wetted areas.

36
2.2.14 Control surfaces
The primary control surfaces for this design are the aileron and the rudder.
The model will not incorporate an elevator, because the horizontal tail has been
modeled as a whole moving surface. A guidance for their initial geometry is
provided in reference [25].
The aileron spans the 28% of the wing, and in particular from 63 till 91%
of the semi-span. The outer 9% of the wing semi-span is excluded, since it is
placed at the wing tip vortex flow region and thus provides little effectiveness.
The aileron chord is modeled to include the aft 20% of the wing chord. The
aileron geometry and location is illustrated in the figure 3.8 of the upcoming
chapter.
The rudder spans the 90% of the vertical stabilizer, except for the outer 10%
near the vertical tail tip. The rudder chord is modeled as 35% of the vertical
stabilizer chord. The rudder geometry is depicted in figure 2.17.
The total aileron surface is 2.27 m2 , or 3.56% of the wing reference area. The
rudder surface is 2.55 m2 or 33.6% of the vertical stabilizer area.

Figure 2.17 – Rudder illustration.

37
3. Aerodynamics

3.1 Airfoils

3.1.1 Airfoil selection

The main feature of the aircraft to be designed is the efficient supersonic


flight. The dominant characteristic of the supersonic flow is the presence of
shock waves. In the case of a sharp-nosed wing section, oblique waves will be
formed on the nose with the flow remaining supersonic. On the other hand, if
a nose which is blunt is incorporated, the flow would detach from the nose tip
and would form a bow shock creating a region of subsonic flow behind the wave.
That is a very important factor to avoid the blunt wing section nose in order
to keep the wave drag low. Consequently, in the supersonic aircraft thin airfoils
are being used with typical thickness ratio 4 - 6 %.

Typical wing sections that are used for supersonic aircraft are the NACA
64-series. For instance, the fighters F-15 and F-16 are using 64A modified airfoils.
These airfoils are designed for maximizing laminar flow, decreasing drag and
increasing the critical Mach number.

For the wing, the NACA 64−006 section has been chosen. This is a symmetric
airfoil with 6 % thickness ratio.

For the horizontal stabilizer, the same symmetrical airfoil used for the wing
has been installed. Since the horizontal tail is swept back more than the wing,
using the same airfoil, ensures that it will have a greater critical Mach number,
as well as greater stall angle, which is necessary for recovery.The particular airfoil
offers a a low drag coefficient (Cd ) too.

For the vertical stabilizer, the symmetric airfoil NACA 64 - 009 has been
chosen. This airfoil offers a quite high Cla , which is an important factor for
having satisfying directional stability, while minimizing the weight, due to its
relative low thickness. The low drag coefficient (Cd ) of the airfoil is also another
reason for its selection.

The chosen airfoils and their coordinates, according to NACA, can be found
in Appendix A.

38
3.1.2 Subsonic aerodynamic coefficients
The aerodynamic coefficients for the selected airfoils in subsonic flow has been
calculated with the XFOIL airfoil design/analysis system, which is a viscous,
panel method program with boundary layer analysis. According to the program
creator, XFOIL is a tool that is particularly applicable to low Reynolds numbers
and works better for subcritical airfoils [26]. Thus, the calculated aerodynamic
coefficients have been compared to published experimental data for the particular
airfoils found in reference [15] and presented in Appendix B.
The zero angle of attack lift (Cl0 ) and pitch moment (Cm0 ) coefficients equal
zero for the used symmetrical airfoils.

NACA 64 - 006
In figure 3.1, the lift curve obtained with the XFOIL simulation is presented.
The resulting aerodynamic coefficients obtained from both aforementioned sources
are presented in table 3.1. A significant difference can be observed for the
Clmax and the stall angle. The data to be used for the ongoing analysis are
the experimental data, since for the angles of attack near stall the simulation
with XFOIL failed to achieve convergence of the solution.

Figure 3.1 – NACA 64-006 lift curve for Re = 9·106 (XFOIL).

XFOIL simulation Experimental data


Lift curve slope (Cla ), deg −1 0.1077 0.1040
Maximum lift coefficient (Clmax ) 0.73 0.83
Minimum drag coefficient (Cd ) 0.0049 0.0038
Stall angle (deg) 10 8.5

Table 3.1 – NACA 64-006 aerodynamic coefficients ( Re = 9·106 ).

39
NACA 64 - 009
In table 3.2 are presented both the simulation and the experimental data that
have been obtained. Significant differences for the Clmax and the stall angle can
be observed for this airfoil as well. Knowing the weakness of the supercritical
airfoil simulation, the experimental data will be used for the further analysis in
this case too.

Figure 3.2 – NACA 64-009 lift curve for Re = 9·106 (XFOIL).

XFOIL simulation Experimental data


Lift curve slope (Cla ), deg −1 0.1092 0.1080
Maximum lift coefficient (Clmax ) 1.62 1.17
Minimum drag coefficient (Cd ) 0.0040 0.0041
Stall angle (deg) 17 11

Table 3.2 – NACA 64-009 aerodynamic coefficients ( Re = 9·106 ).

3.1.3 Supersonic aerodynamic coefficients


For an airfoil in supersonic flow, an analytical expression is given from the
equations (3.1) and (3.2). These equations have been derived disregarding the
viscous terms and solving the relevant linearized velocity potential equation ??.
The results are valid for thin airfoils, which are typical for supersonic flight,
and for small angles of attack, where the linearization can be considered quite
accurate. Deriving the aforementioned equations, it can be observed that the cl
is independent of the airfoil shape and thickness, in contrast with the cd .

cl = √ (3.1)
M2 − 1

40
4(α2 + gc2 + gt2 )
cd = √ (3.2)
M2 − 1

In equation (3.2), gc and gt are functions of the camber and the thickness
along the chord, respectively. For a symmetric airfoil the camber term gc equals
to zero and for a double-wedge airfoil the thickness term gt equals the square of
the thickness ratio ( (t/c)2 ).

3.2 Subsonic Lift-Curve Slope

3.2.1 Wing - Fuselage Assembly

Reference [27] provides with an empirical equation for the wing’s subsonic
CLα estimation. This equation takes into account the compressibility effects, the
aspect ratio and the wing sweep. The effect of taper ratio is not directly included
in the relation, but it has been proved by the author that it can be diminished
by using the half-chord (Λc/2 ), instead of the quarter-chord (Λc/4 ) wing sweep.
The relevant equation is expressed as

clα · AR
(CLαw )M = r 2 in radians (3.3)
clα 2
clα
π + AR
cos Λc/2 + π − (AR · M )2

In order to account for the fuselage contribution, the CLα obtained from the
above relation, is multiplied with the factor (Sexp /Sref ) · (F ), where the fuselage
lift factor F is defined as [9]

F = 1.07 (1 + Df us /b)2 (3.4)

If the F obtained from relation (3.4) is greater than one, the result is physically
wrong, since it assumes that the fuselage produces more lift than the covered
portion of the wing. For that reason, the factor F is set to 0.98 [9]. Since
equation (3.3) can be considered valid for all subsonic values of Mach number
till the critical one, the results for a freestream Mach number range of 0.2 till
0.8 are presented in table 3.3.

41
Mach number CLαwb (deg −1 )
0.2 0.0543
0.3 0.0548
0.4 0.0557
0.5 0.0568
0.6 0.0584
0.7 0.0603
0.8 0.0630

Table 3.3 – Lift-curve slope of wing - body assembly for


subsonic Mach number range.

3.2.2 Horizontal Tail


The lift coefficient of the horizontal tail (CLt ) with respect to the horizontal
tail angle of attack (αt ) for the linear part of the curve is given from the following
relation

CLt = CLat αt (3.5)

The term CLt can be computed using the equation (3.3) and introducing the
correspondent properties of the horizontal stabilizer. In order to define a total
lift coefficient for the aircraft, the relevant coefficient of the horizontal stabilizer
should be written in terms of the wing-body angle of attack. The relation that
correlates the wing-body and the horizontal tail angle of attack is [28]

at = awb − it −  (3.6)

The horizontal tail incidence (it ) has been set to one degree. The downwash
angle (()) which is created by the wing wake and changes the effective angle of
the horizontal tail can be approximated by [28]

∂
 = 0 + αwb (3.7)
∂α

The downwash angle 0 at αwb = 0, is mainly dependent on the wing twist.


Since wing twist has been set to zero for the initial layout, the 0 can approximated
as zero. The downwash angle derivative can be approximated using the empirical
equation (3.8), based on wind-tunnel experimental data [29].

∂ 21o CLαw c̄W 10 − 3λw zHT


   
= 1− (3.8)
∂α AR0.5 LHT 7 bw

42
where the term zHT is the vertical distance of the horizontal stabilizer above of
the main wing plane equal to 2.2 m for this design. Therefore, the horizontal
tail contribution to the aircraft CLα can be written as

∂ SHT
 
∆CLαt = CLαt 1− (3.9)
∂α SW

In table 3.5, the calculated values of interest for the horizontal tail can be
observed.

Mach number CLαt (deg −1 ) ∂/∂α ∆CLαt (deg −1 )


0.2 0.0487 0.6104 0.0036
0.3 0.0492 0.6168 0.0036
0.4 0.0498 0.6262 0.0035
0.5 0.0506 0.6391 0.0035
0.6 0.0517 0.6562 0.0034
0.7 0.0531 0.6787 0.0032
0.8 0.0549 0.7083 0.0030

Table 3.4 – Lift-curve slope of horizontal tail, downwash angle derivative and
horizontal tail CLα contribution to the aircraft, respectively, for subsonic Mach
number range.

3.2.3 Total aircraft


The lift-curve slope for the total aircraft can be obtained by summing the
wing-fuselage and the horizontal tail contribution. The final results are presented
in table 3.5.

Mach number CLα (deg −1 )


0.2 0.0579
0.3 0.0584
0.4 0.0592
0.5 0.0603
0.6 0.0617
0.7 0.0636
0.8 0.0660

Table 3.5 – Aircraft lift-curve slope for subsonic Mach number range.

43
3.3 Supersonic Lift-Curve Slope
The wing is considered to be in purely supersonic flow when the Mach cone
angle is greater than the leading edge sweep. For the wing leading edge sweep
of 45 deg, that holds for Mach numbers greater than 1.4. The lift-slope curve at
supersonic speeds is quite hard to estimate without experimental measurements
or usage of a sophisticated fluid dynamics simulation model. An estimation
about the wing normal force curve-slope coefficient (CN α ) is given in figure 3.3.

Figure 3.3 – Wing supersonic CN α for taper ratio of 0.2 [13].

From the figure 3.3, the CN α can be approximated, taking into account the
leading edge wing sweep, the taper ratio and the aspect ratio.For small angles
of attack CN α can be considered equal to the supersonic wing CLα . In order
to account for the fuselage contribution the obatined coefficients are multiplied
with the factor (Sexp /Sref ) · (F ), like in the case of the subsonic flow.

Mach number CLαwb (deg −1 ) CLαt (deg −1 )


1.4 0.0663 0.0623
1.5 0.0597 0.0570
1.6 0.0545 0.0524
1.7 0.0499 0.0482

Table 3.6 – Wing-body and horizontal tail lift-curve slope for


supersonic Mach number range.

For the horizontal stabilizer, the supersonic lift curve slope is obtained applying

44
the same methodology with the aircraft wing. The correspondent values of CLαwb
for the wing-fuselage assembly and CLαt for the horizontal tail are presented in
table 3.6.
The relation (3.9) can be used to calculate the horizontal stabilizer contribution
to the CLα of the total aircraft. The downwash angle derivative can be roughly
approximated using the relation (3.10), provided in [9], with the relevant results
presented in table 3.7. In table 3.8, the supersonic lift-curve slope for the total
aircraft is included.

∂ 1.62CLαw
= in radians (3.10)
∂α πAR

Mach number ∂/∂α ∆CLαt (deg −1 )


1.4 0.6250 0.0044
1.5 0.5621 0.0047
1.6 0.5135 0.0048
1.7 0.4700 0.0049

Table 3.7 – Horizontal tail downwash angle derivative and CLα contribution to
the aircraft, respectively, for supersonic Mach number range.

Mach number CLα (deg −1 )


1.4 0.0708
1.5 0.0644
1.6 0.0593
1.7 0.0547

Table 3.8 – Aircraft lift-curve slope for supersonic Mach number range.

3.4 Maximum Lift Coefficient


3.4.1 Clean configuration
The aircraft maximum lift coefficient CLmax is very important for low flight
speeds and especially for the landing and the take-off flight phase. The estimation
of the CLmax in this section is for the stall speed at sea-level, which has been set
to 0.2 Mach and for a range till 0.6 Mach.
For the CLmax , the method from the USAF DATCOM has been used. Firstly,
the aspect ratio condition of the method has to be considered. The relations 3.11
and 3.12 are presenting the low and high aspect ratio condition, according to
USAF DATCOM, respectively [9], [21].

45
3
AR ≤ (3.11)
(C1 + 1)cos(ΛLE )
4
AR > (3.12)
(C1 + 1)cos(ΛLE )
where C1 is the taper ratio correction coefficient that can be calculated using
the following relation [21].
 0.4 (π(1−λ)2 )

C1 = 0.5 sin π(1 − λ)1.5+0.8 sin (3.13)

Figure 3.4 – High aspect wing and airfoil maximum lift coefficient
ratio at 0.2 Mach.

The wing used for this design does not satisfy neither of the conditions, since
it falls between the above calculated values. For that reason, both conditions
have been implemented and the results have been compared in the end. An
important parameter that is used for this method is the leading edge parameter
(∆y) of the wing section, defined as 21.3 times the thickness ratio (t/c) for a
NACA 64-series airfoil. Then the CLmax can be obtained from the following
relation in the case of high aspect ratio condition.

CLmax
 
CLmax = Clmax + ∆CLmax (3.14)
Clmax
where Clmax is the wing section maximum lift coefficient, the ratio CLmax /Clmax
is obtained from figure 3.4 for Mach 0.2, while ∆CLmax is the correction for
Mach numbers up to 0.6 and can be found using the relevant chart provided in
reference [9].
The correspondent stall angle of attack (αstall ) can be approximated from
the relation

46
CLmax
αstall = + α0L + ∆αCLmax (3.15)
C αw
where the zero lift angle of attack (α0L ) is equal to zero for the initially assumed
zero wing incidence and the angle of attack increment (∆αCLmax ) can be obtained
from figure 3.5.

Figure 3.5 – Stall angle of attack increment at subsonic Mach


numbers of 0.2 - 0.6.

The CLmax and αstall , in the case of the low aspect ratio condition, can be
obtained from the equations 3.16 and 3.17, where the correspondent parameters
can be found from the USAF DATCOM charts provided in reference [9].

CLmax = (CLrmmax )base + ∆CLmax (3.16)

αstall = (αCLmax )base + ∆αCLmax (3.17)

Mach number CLmax αstall


High AR Low AR Final High AR Low AR Final
0.2 095 0.91 0.93 25.6 22 23.8
0.3 0.93 0.89 0.91 25.1 21 23.0
0.4 0.90 0.88 0.89 24.3 21 22.6
0.5 0.87 0.85 0.86 23.5 20 21.7
0.6 0.86 0.83 0.85 22.9 20 21.4

Table 3.9 – Subsonic aircraft maximum lift coefficient and stall angle of attack.

47
The relevant values of CLmax and αstall obtained from both conditions are
presented in table 3.9. Since the values obtained are not very far, the mean
value of them has been computed and are implemented as values for this model.
It can be seen that the values obtained for the aircraft maximum lift coefficient
and stall angle of attack are greater than the relevant ones for the airfoil that
was implemented. This happens due to the fact that the low aspect ratio, swept
wing incorporates a sharp leading edge, which results in the leading-edge vortices
formation that enhances the maximum aircraft lift.

3.4.2 High lift devices


In order to achieve a CLmax high enough for landing and take-off, both trailing
edge (TE) and leading edge (LE) high lift devices have been incorporated to the
design. At the trailing edge the single slotted Fowler flap has been used, while
at the leading edge the slotted LE flap or so-called slat. Both devices aim at
the wing camber and wing area increment, with the slots being incorporated to
assist the avoidance of flow separation, with high pressure air being able move
through the slots from the lower to the upper side of the wing.

Figure 3.6 – Trailing and leading edge high lift devices.

The ∆Clmax increase of the wing section incorporating a TE slotted flap is


approximated as 1.3 times the airfoil chord ratio (c 0 /c) after and before the flap
rearward movement [9], as can be observed in figure 3.6. For this design a typical
value of 1.1 or 10 % chord increment, has been chosen.
Both TE flaps and slats are hinged having a chord 20% of the local wing
section chord. The TE flaps span the 35% of the wing span, while the slats its
62%. Due to the engines existence, the TE flaps and the slats are not continuous,
as can be seen in figure 3.7.
The maximum total aircraft ∆CLmax increment due to TE flap deflection
during landing can be estimated using the following equation [30]

Swf
∆CLmax = ∆Clmax K∆ (3.18)
Sw
where the empirical sweep correction K∆ is given from the relation

K∆ = (1 − 0.08 cos2 (Λc/4 )) cos0.75 (Λc/4 ) (3.19)

48
The area Swf corresponds to the wing area that are incorporating flaps and
the area Sw to the total wing area. The ratio (Swf /Sw ) equals to 39.7% and the
relevant ∆CLmax increment equals to 0.45.

Figure 3.7 – Wing TE flaps (red), LE flaps (magenta) and ailerons (cyan).

The ∆CLmax increment due to the LE device extension has been crudely
approximated as 0.4 [9], since there is no analytical method for its prediction.
The total ∆CLmax increment can be estimated as 0.85, meaning that the aircraft
CLmax using the designed high lift devices will be 1.78 at 0.2 Mach.

3.4.3 Horizontal tail


A requirement during the horizontal tail design is that it should stall later
than the wing for recovery reasons. Following the USAF DATCOM method for
a low aspect ratio wing, the CLmax and αstall for the horizontal stabilizer can be
estimated from equations (3.16) and (3.17). The obtained results are presented
in table 3.10.

Mach number CLmax αstall


0.2 0.89 24
0.3 0.88 24
0.4 0.87 23
0.5 0.85 23
0.6 0.83 22

Table 3.10 – Subsonic horizontal tail maximum lift coefficient and stall
angle of attack.

The values for the horizontal tail αstall are slightly higher than the relevant
values calculated for the aircraft previously. Taking into account that due to
the tail incidence and the downwash angle, which cause the horizontal tail to

49
experience an angle of attack lower than the wing, the horizontal tail will have
indeed a quite higher actual stall angle.

3.5 Subsonic Parasite Drag Coefficient


3.5.1 Equivalent skin-friction method
The equivalent skin-friction method is an approximative method, which uses
a equivalent skin-friction coefficient Cfe that is relevant to the aircraft class. This
coefficient includes both skin-friction and separation drag and for a supersonic
aircraft equals to 0.0025 [9]. This method is very generic and is presented here
just to get an initial estimate to use for comparison with the more accurate
results obtained later on.
The relevant equation for the parasite (zero-lift) drag coefficient CD0 estimation
is

Swet
CD0 = Cfe (3.20)
Sref

The total aircraft wetted area (Swet ) can be obtained from table 2.9 by just
summing the relevant wetted areas of each aircraft component. The resulting
CD0 is 0.0120.

3.5.2 Component buildup method


The component buildup method estimates the CD0 for each aircraft component.
Every component that has direct contact with air flow, is producing drag. The
basic contributing components of an aircraft in cruise are the wing, the fuselage,
the tail and the engine nacelles. The CD0 of each component is approximated
by computing a flat-plate skin-friction coefficient (Cf ) and multiplying it with
a component form factor (FF ), which is used in order to include the separation
drag. For the additional drag related to the components interference effects,
because of the thicker boundary layer formation at components intersections,
the interference factor (Q) is introduced. Therefore, the relation to calculate the
subsonic CD0 is expressed as [9]
P
(Cfc F Fc Qc Swetc )
CD0 = + CDmisc + CDL&P (3.21)
Sref

The factor CDL&P is related to drag created due to leakages and protuberances.
A carefully designed aircraft of this type can almost eliminate the specific drag
contribution, so that it can be considered negligible. The factor CDL&P corresponds
to miscellaneous drags, such as the flaps extension or the landing gear deployment.
This factor for the cruise condition can be set equal to zero. During different
flight phases, like for instance during landing, these drag contributions have to
be estimates and added to the total drag.

50
In order to compute the friction coefficient Cf , it has to be examined in what
extend the flow is laminar or turbulent. That was achieved through the Reynolds
number (Re) calculation for all the components for Mach number range of 0.2
till 0.8, which is just below the critical Mach number (see subsection 3.7), and
for the altitudes of 0, 5 and 10 km, respectively. It has to be mentioned that
the flight at the lowest Mach numbers and higher altitudes is not possible, thus
not of interest, but the relevant coefficients will be presented just for reasons of
completeness.
The characteristic length for the Reynolds number calculation has been
for the wing, the horizontal and the vertical stabilizer the respective mean
aerodynamic chord of each surface, while for the fuselage and the nacelle-engine
assembly their longitudinal total length. The Reynolds number has been computed
to be grater than 5 ·105 for the aforementioned flight conditions, so that the flow
can be considered fully turbulent in every case.
In order to account for the skin surface roughness, which can possibly increase
the viscous forces on the components’ surface, a fictitious Reynolds number, the
so-called cutoff Re, has been used. This Re can be computed from the equation
(3.22), using the characteristic length (l) of each component, as defined before,
and the skin roughness value k of polished metal, which is equal to 0.152·10−5
m. Therefore, in the case of an actual Re being higher than the relevant cutoff
Re, its actual value has to be replaced by the cutoff one in the equation (3.23)
for the Cf calculation.
 1.053
l
Recutof f = 38.21 (3.22)
k

The skin-friction coefficient (Cf ) can then be estimated for every component
using the relation

0.455
C fc = (3.23)
(log10 Rec )2.58 (1+ 0.144M 2 )0.65

The form factor (FF) for the wing, the horizontal and the vertical tail can be
calculated from the equation (3.24). The parameter (x/c)m corresponds to the
chordwise position of the airfoil maximum thickness, which is 0.4 for the selected
airfoils, while Λm is the sweep angle of the maximum thickness line.
The form factors (FF) of the fuselage and the nacelle assembly can be
computed from the equations (3.25) and (3.26), respectively. The factor f in
both equations is equal to the length over the maximum diameter ratio.

 4 ! 
0.6 t t
  
FF = 1+ + 100 1.34M 0.18 (cos Λm )0.28 (3.24)
(x/c)m c c

60 f
 
FF = 1 + 3 + (3.25)
f 400

51
F F = 1 + (0.35/f ) (3.26)

The interference factor (Q) for a well-filleted wing design and the fuselage is
negligible and can be set to unity. For a conventional tail, it can be assumed
as 1.04 for the horizontal and vertical stabilizer surfaces, while for the directly
mounted to wing nacelles is about 1.5 [9].
Making all the necessary calculations presented previously, the ”clean” subsonic
parasite drag coefficient (i.e. cruise condition configuration) can finally be computed
using the aforementioned relation (3.21). The obtained results are presented in
table 3.11.

Mach number CD0


0 km 5 km 10 km
0.2 0.0125 0.0132 0.0141
0.3 0.0121 0.0128 0.0136
0.4 0.0118 0.0125 0.0133
0.5 0.0115 0.0122 0.0130
0.6 0.0113 0.0119 0.0127
0.7 0.0111 0.0117 0.0124
0.8 0.0109 0.0115 0.0122

Table 3.11 – Subsonic parasite drag coefficients for the stated Mach
number and altitude flight conditions.

These results are close to the expected value of parasite drag coefficient, as
it has been obtained with the equivalent-skin-friction method. The respective
coefficient is also increasing at higher flight altitudes as it can be observed.
However, that does not mean a higher drag force, since the relevant dynamic
pressure at higher altitude is much smaller.

3.6 Supersonic Parasite Drag Coefficient


The supersonic parasite drag includes the skin-friction drag, the miscellaneous
drag and the drag due to leakages and protuberances, like in the case of the
subsonic flight. In addition to these contributions, it includes the so-called wave
drag, which is a drag increment resulting from the shock waves formation in
transonic and supersonic speed. The total drag coefficient using the components
buildup method, following the same principle like in the subsonic flow case, is
given from
P
(Cfc Swetc )
CD0 = + CDmisc + CDL&P + CDwave (3.27)
Sref

52
Since external stores are not incorporated in the SST aircraft design, the
CDmisc is zero. Moreover, the drag due to leakages and protuberances can be
assumed negligible, like in the case of the subsonic drag coefficient.

From the equation (3.27), it can be seen that for the supersonic skin-friction
drag component, the form (FF) and the interference factor (Q) are not included.
The skin friction coefficient (Cf ) is calculated using the equation (3.23), since the
flow is fully turbulent for the supersonic flight conditions. The flight conditions
that are evaluated here are flight at 10, 12 and 15 km altitude, with the supersonic
Mach number range to be from 1.4 to 1.7. Furthermore, in order to account
for the skin roughness, a cutoff Mach number is introduced and used in the
exactly same manner for the Cf calculation, such as for the subsonic flight. The
expression for the supersonic cutoff Mach number is

 1.053
l
Recutof f = 44.62 M 1.16 (3.28)
k

The total skin-friction drag coefficient (CDf ) can then be estimated by summing
the respective coefficients of every aircraft component for the aforementioned
flight conditions. The results are presented in table 3.12.

Mach number CDf


10 km 12 km 15 km
1.4 0.0084 0.0087 0.0093
1.5 0.0081 0.0084 0.0090
1.6 0.0079 0.0082 0.0088
1.7 0.0077 0.0079 0.0085

Table 3.12 – Supersonic skin-friction drag coefficient component for the


stated Mach number and altitude flight conditions.

The wave drag is the biggest drag contributor during supersonic flight. The
wave drag is dependent on the way the aircraft volume is distributed along its
longitudinal axis. The minimum wave drag appears for a so-called Sears-Haack
body, having an ideal volume distribution, like the one illustrate in figure 3.8.
The cross-section radius (r) of a Sears-Haack body can be given from the relation

r(x) = rmax (4x(1 − x))0.75 (3.29)

where x is the ratio of the longitudinal distance from the aircraft nose over the
tonal length, having thus values between 0 and 1.

53
Figure 3.8 – Sear-Haacks body volume distribution [9].

Figure 3.9 – Wing-body area rule design [9].

The Sears-Haack body distribution cannot be achieved for a real aircraft,


since the wing, the tail and the nacelles installation, are resulting in distortion
of this smooth cross-section area distribution. However, in order to make the
distribution smoother and as close to the ideal one as possible, the area ruled
fuselage is implemented. The area ruling of the fuselage has been firstly proposed
by Whitcomb, who presented this principle for the wing-body combination [31].
In order to achieve a smoother distribution and reduce the wave drag, the
fuselage has been squeezed at the location of the wing installation (see figure
3.9).

54
Figure 3.10 – Aircraft cross-section area distribution.

The area rule has been applied in this design, like it was stated in the
fuselage lofting section. In figure 3.10, the cross-section area of the designed
aircraft can be compared with the Sears-Haack ideal distribution. The presented
cross-section area of the aircraft corresponds to intersection with perpendicular
to the longitudinal axis and at zero roll angle. The nose location of the aircraft
corresponds to the zero longitudinal dimension. The inlet capture area has been
subtracted from the cross-section area distribution graph.
The most credible way to measure the wave drag, so that to test the design
and optimize it, is to conduct experiments on a wind tunnel. However, there
are also some analytical methods that can be used in order to make an initial
estimation during the conceptual aircraft design. Here two of them will be
illustrated, provided from references [9] and [29]. Both of them relate the aircraft
wave drag with the Sears-Haack body wave drag, through the relevant coefficient
given from relation (3.30), and an efficiency factor EWD . The typical value for
the EWD factor of a supersonic fighter and SST is about 1.8 [9]. However, a factor
of 1.4 is selected for this area-ruled design, since a very high wave drag would
deteriorate the aerodynamic efficiency significantly. The wave drag minimization
and optimization of the aircraft shape is a necessary procedure and one of the
most important challenges that has to be met for an efficient SST design, and
usually demands a lot of experimental testing of the model as well.
2
9π Amax

CDwS−H = (3.30)
2Sref l

In the above equation, the Amax is the maximum cross-section area, which
is about 6.42 m2 for this design, and l the aircraft’s overall length. The aircraft
wave drag coefficient can then be analytically estimated using the equations
(3.31) and (3.32). The final wave drag coefficient to be used is the arithmetic
mean of the values obtained from the two analytical relations. These value are

55
then added to the computed ones for the skin-friction drag, as presented before,
with the final supersonic parasite drag coefficients to be displayed in table 3.13.

πΛ0.77
!!
0.57 LE−deg
CDwave = CDwS−H EW D 1 − 0.386(M − 1.2) 1− (3.31)
100

p
CDwave = CDwS−H EW D (0.74 + 0.37 cos ΛLE ) (1 − 0.3 M − cos−0.2 ΛLE )
(3.32)

Mach number CD0


10 km 12 km 15 km
1.4 0.0216 0.0219 0.0226
1.5 0.0211 0.0214 0.0220
1.6 0.0205 0.0209 0.0214
1.7 0.0201 0.0204 0.0209

Table 3.13 – Supersonic parasite drag coefficient component for the


stated Mach number and altitude flight conditions.

3.7 Critical Mach number


A method for determining the critical Mach number of an infinite swept
wing is illustrated in reference [32]. The equation (3.33) relates the critical
Mach number (Mcrit ) for an infinite swept wing at zero angle of attack with the
critical pressure coefficient (CPcrit ).
 γ 
γ−1 2 2
!
γ−1
2 1+ 2 Mcrit cos Λc/4
CPcrit = 
γ+1 − 1 (3.33)
γMcrit 2 2

The pressure coefficient (CP ) at freestream Mach number M is then obtained


using the relation (3.34), where β is the compressibility correction factor. The
CP at M=0 is approximated as the minimum airfoil pressure coefficient. For
incompressible, inviscid flow using the Bernoulli equation the CP can be obtained
from relation (3.35).

CPM =0 CPM =0
CPM = =q (3.34)
β 1 − M 2 (cos2 Λc/4 − CPM =0 )

CP = 1 − (V /V∞ )2 (3.35)

The ratio (V /V∞ ) of the local velocity along the chord over the freestream
velocity can be obtained for the symmetrical airfoils NACA 64-006 and NACA

56
64-009 at zero angle of attack from reference [15]. Therefore, the minimum
airfoil pressure coefficient for the aforementioned airfoils is -0.171 and -0.267,
respectively.

Figure 3.11 – Wing critical Mach number in two-dimensional flow.

At the critical Mach number, the CP from the two equations (3.33) and (3.34)
is equal. Thus, the Mcrit can be found graphically by plotting the CP obtained
from both equation for a Mach number range of 0.6 to 1.0. The Mcrit is the one
where the two curves intersect. From figure 3.11, the Mcrit corresponding to the
wing can be obtained.
Using the same procedure the Mcrit for the horizontal and vertical stabilizer
can been computed too. The relevant values of the Mcrit are 0.92 for the wing
and the vertical stabilizer, and 0.94 for the horizontal stabilizer. It has to be
mentioned that these values for Mcrit correspond to two-dimensional linearized
flow based on the Weber’s compressibility correction [32] of equation (3.34). The
main reason for illustrating this method is to ensure that the Mcrit of the wing
does not exceed the Mcrit of the tail, for the selected wing sections and sweep
angles.
The aircraft Mcrit can be estimated through the relation [29]

0.6
t
  
Mcrit = 1 − 0.065 100 cos Λc/4 (3.36)
c max

For this SST aircraft design the Mcrit has been calculated as 0.84. Another
important Mach number related with the transonic flow, is the drag divergence
Mach number Mdd . The Mdd is the Mach number at which the formation of
shock waves begins to affect the aircraft drag significantly. As a rule of thumb
the Mdd is 0.08 Mach higher than the Mcrit [9], or 0.92 in this case.

57
3.8 Drag due to Lift
In order to calculate the total aircraft drag the drag produced due to lift has
to be added to the parasite drag. The total drag coefficient (CD ) is related to
the lift coefficient (CL ) through the following equation known as the drag polar.

CD = CD0 + KCL2 (3.37)

The so-called drag-due-to-lift factor K, in this case of a symmetric airfoil,


includes both the induced drag term and the viscous separation drag term. The
factor K for subsonic flight can be estimated through the equation

1
K= (3.38)
πARe
The coefficient e is the Oswald efficiency factor, which accounts for the
non-elliptical lift distribution over the wing. For a swept-wing aircraft, with
leading edge sweep over 30 deg, it can be estimated as [9]

e = 4.61(1 − 0.045AR0.68 )(cos ΛLE )0.15 − 3.1 (3.39)

Therefore, K for the subsonic flight range for this design is equal to 0.1181.
For the supersonic flight range, the factor K can be obtained from the relation
(3.40) [9]. The obtained values are presented in table 3.14.

AR(M 2 − 1) cos ΛLE


K= √ (3.40)
(4AR M 2 − 1) − 2

Mach number K
1.4 0.2061
1.5 0.2298
1.6 0.2524
1.7 0.2742

Table 3.14 – Supersonic drag-due-to-lift factor.

3.9 Miscellaneous Drag


3.9.1 Flaps
The flaps and slats are contributing to both the parasite and induced drag.
The relations (3.41) and (3.42) are giving a first estimation for the relevant drag
coefficients calculation [9].

58
Cf Swf
  
∆CD0f lap = Ff lap (δf lap − 10) (3.41)
C Sw

∆CDi = kf2 (∆CLf lap )2 cos Λc/4 (3.42)

In the above equations, the factor Fflap equals 0.0074 for slotted flaps and
the factor kf about 0.28. The Cf /C is the flap over wing chord ratio and the
δflap is the flap deflection in deg. Typical values for a Fowler TE flap is 40 deg
at landing and 20 deg at takeoff, while for a slat is 25 deg at landing and 15 at
takeoff. As a rule of thumb, the lift coefficient increment at take-off is about 65
% of the relevant value at landing. The approximated values for the flap drag
coefficients during takeoff and landing are presented in table 3.15.

Takeoff Landing
CD0 CDi CD 0 CD i
TE Flaps 0.0059 0.0053 0.0176 0.0124
Slats 0.0041 0.0042 0.0123 0.0098

Table 3.15 – Drag coefficients of high lift devices during takeoff and landing.

3.9.2 Spoilers
The spoilers are secondary control surfaces that are used as speed brakes
during flight, as lift dumpers during ground roll and takeoff abortion and as
roll control surfaces, mainly for higher speed, where the aileron effectiveness is
reduced.

Figure 3.12 – Wing control surfaces (blue) and spoilers (red).

59
The initial layout of the spoilers was based on guidance from [33]. The spoiler
is put just ahead of the flaps with a chord equal to 15% of the local wing chord.
The spoiler spans the 42% of the wing, and in particular from 19.6 till 61.6 % of
the semispan, as it can be seen in figure 3.12. The total spoilers surface (Ss ) is
about 4.52 m2 or 7.1% of the reference wing area.
The spoilers drag coefficient with respect to the spoiler deflection angle (δs )
can be roughly approximated through the relation [33]

Ss
∆CDspoiler = 1.9 sin (δs ) (3.43)
Sref

Using the above relation for a spoilers deflection of 45 deg, the increment of
the aircraft CD is estimated as 0.0952.

3.9.3 Landing gear


A quick estimation for the drag coefficient of a retractable landing gear can be
made from the empirical relation 3.44 [21]. This equation is valid for commercial
jet aircraft having with deployed flaps, which is the case during takeoff and
landing, when the landing gear is deployed too. The value obtained using the
following relation for the CD increment due to deployed landing gear is 0.0152.

3.099 · 10−4 m00.785


∆CDLG = (3.44)
Sref

60
4. Weights

4.1 Weights Estimation Refined Method


There are several approximative methods for estimating the weights of the
aircraft components. For a quick estimation, they can be computed as percentage
of the total aircraft weight. These methods are good for giving some initial
estimates for the basic aircraft components, which can be used as a guide.

Aircraft Component Weight (kg)


Wing 2360.2
Horizontal tail 244.4
Vertical tail 329.1
Fuselage 3340.3
Main landing gear 459.6
Nose landing gear 97.7
Engines 2590.3
Nacelle group 510.1
Engine controls 21.2
Starter (pneumatic) 88.3
Fuel system 207.7
Flight controls 285.3
APU installed 104.8
Instruments 69.3
Hydraulics 60.3
Electrical 329.3
Avionics 697.8
Furnishings 253.9
Seats 394.0
Air conditioning 176.2
Anti-ice 59.0
Handling gear 8.8

Table 4.1 – Aircraft weights estimation.

In this chapter, a more refined method, developed in reference [9], has been
implemented. The relevant weights are estimated using statistical equations

61
based on the aircraft geometric characteristics and initial layout. The analysis is
based on the equations correspondent to transport aircraft, which can be found
in [9] and will not be reproduced here. The results obtained from the weights
estimation analysis are registered in table 4.1.
One important parameter, used in the aforementioned equations and not
defined before, is the maximum load factor (nm ax), which has been set equal to
3.5 for the SST aircraft design.
The total aircraft empty weight after the sum of all the weights of the aircraft
components is calculated to be 12,541 kg. The initial estimated aircraft empty
weight has been 13,908 kg. Since the weight obtained is lower than the initial
estimation with the difference to be not very high, the aircraft dimensions will
not be refined. The initial takeoff weight will be considered the same for the
rest of analysis and they weight savings of the aircraft structure can be utilized
as payload increment. That means that the refined empty weight fraction is
dropping to 0.4251 from 0.4715 for this SST design. The relevant value of
the empty weight fraction for the Concorde has been calculated to be about
0.417, which means that the value obtained from this refined method is indeed
a reasonable estimate.

4.2 Center of Gravity


The center of gravity of the aircraft cannot be estimated so accurately during
the conceptual design, since the aircraft is susceptible of many changes till the
final actual design is obtained. For the symmetric aircraft design, the lateral
CG location is found on the aircraft longitudinal axis of symmetry. Moreover,
the location of the longitudinal CG has to be carefully placed from the designer,
since it strongly affects the aircraft longitudinal stability. An analysis, made
in the next chapter, makes an initial recommendation for the longitudinal CG
placement.
The vertical CG location can be calculated from the equation (4.1), where
all the N aircraft components have to be included.

N
X
mc zc
i=1
zCG = N
(4.1)
X
mc
i=1

An initial approximation of the vertical CG location has been made using


just the basic aircraft components, i.e. the wing, the fuselage, the horizontal and
vertical tail and the engine-nacelle groups. The estimations of the aforementioned
components’ vertical locations (zc ) are given in table 4.2. The aircraft CG
vertical location has been found to be at about 0.75 m with respect to the
fuselage centerline and defining the positive z direction as downwards.

62
Component zc (m)
Wing 1.1
Fuselage 0
Horizontal tail -1.1
Vertical tail -2.42
Engine - Nacelle assembly 1.69

Table 4.2 – CG vertical location of basic aircraft components.

63
5. Stability and Control

5.1 Subsonic Static Longitudinal Stability


The two criteria that have to be met to achieve static longitudinal condition
are that the aircraft pitching moment at zero angle of attack is positive (Cm0 > 0)
and that the slope of the pitching moment coefficient is negative (Cmα < 0). The
pitching moment (Cm ) is about the aircraft’s center of gravity (CG). Meeting
this two criteria is meaning that the aircraft can fly in stable equilibrium.
The assumptions that have been used for the analysis are presented here. The
wing aerodynamic center is modeled to be at 25 % of the wing mean aerodynamic
chord and the same holds for the horizontal tail. Moreover, the wing-body
assembly aerodynamic center has been assumed to be in the same location as
for the wing. The fuselage pitching moment about the wing-body aerodynamic
center has been set equal to zero and the propulsive system contribution has
been neglected.

5.1.1 Aircraft Pitching Moments


The pitching moment of the wing-body can be calculated from the following
equation [28]

Cmwb = Cmacwb + CLαwb αwb (h − hnwb ) (5.1)

In the above equation αwb is the wing-body angle of attack, hn and hnwb the
non-dimensional longitudinal distance of the center of gravity and aerodynamic
center from the fuselage nose divided with the wing’s mean aerodynamic chord,
respectively. Moreover, all the longitudinal locations presented in this chapter
are measured as distances from the aircraft nose.
The pitching moment Cmacw of the wing about the aerodynamic center is
strongly dependent on the airfoil pitching moment and it can be estimated
through the relation [9]
!
AR cos2 Λ
Cmacw = Cm0airf oil (5.2)
AR + 2 cos Λ

The airfoil that has been used for the wing is a symmetrical one with Cm0 = 0.
Thus, the resulting Cmacw is equals to zero too.

64
The tail pitching moment coefficient can be expressed as [28]

Sht
Cmt = −V̄H CLt + CLt (h − hnwb ) (5.3)
Sref

where V̄H = (l̄t SHT )/(c̄Sref ) with l̄t to be the distance between the wing body
and the horizontal tail mean aerodynamic center (see figure 5.1).

Figure 5.1 – Wing-body and tail mean aerodynamic centers [28].

The total aircraft Cm can then be written as

Cm = Cm0 + Cmα α (5.4)


where

!
CLαt SHT ∂

Cm0 = Cmacwb + CLαt V̄H (0 + it ) 1 − 1− (5.5)
CLα Sref ∂α

∂
 
Cmα = CLα (h − hnwb ) − CLαt V̄H 1− (5.6)
∂α

The parameter Cm0 corresponds to the pitching moment coefficient at zero


lift and is independent of the CG location.

5.1.2 Subsonic Neutral Point


The neutral point (NP) location of the aircraft can be found from the relation

CLαt ∂
 
hn = hnwb + V̄H 1 − (5.7)
CLα ∂α

The aircraft has a positive stiffness (Cmα < 0) when the center of gravity is
forward of the neutral point. This means that it should hold

Static margin (SM ) = hn − h > 0 (5.8)

65
From the equation (5.7) the aircraft’s neutral point location for the subsonic
aircraft range has been calculated to be 14.6 m, which determines the limit for
the center of gravity aft position.
The influence of the CG location on the aircraft Cm can be observed in the
figure 5.2. For positive pitch stiffness (h < hn ) the aircraft can fly in equilibrium,
when for zero (h = hn ) or negative pitch stiffness (h > hn ) the aircraft is unbalanced,
since the pitching moment cannot be zero for any angle of attack.

Figure 5.2 – CG position influence on Cm at 0.5 Mach.

5.1.3 Longitudinal Control and Trim Analysis


The longitudinal control of the designed aircraft can accomplished through
the all moving horizontal tail surface. The change in the tail angle δt results in
the change of the tail incidence. The tail incidence has been defined as positive,
when it incorporates a negative angle of attack with respect to the wing-body
zero angle of attack reference axis, with the tail thus to create negative lift and a
positive (nose up) pitching moment. The tail deflection angle δt is then, in fact,
the tail incidence for the given condition, but with the opposite sign convection.
In order to present the δt as a change of tail incidence with respect to the
zero wing-body angle of attack line, the initial tail incidence (it ) for the trim
analysis has been set equal to zero. The results then for the δt are the opposite
signed it that are required for stable equilibrium. This has been done in order
to avoid conversions of the obtained from the trim analysis angles, which would
be with respect to the it , and keep the δt definition simpler.
The derivatives of the lift and the pitching moment with respect to the δt
can be expressed, respectively, as

SHT
CLδt = CLαt (5.9)
Sref

66
Cmδt = −CLαt V̄H + CLαt (h − hnw b ) (5.10)

So for the case of linear lift and linear pitching moment the set of equations
obtained is

CL = CL0 + CLα α + CLδt δt (5.11)

Cm = Cm0 + Cmα α + Cmδt δt (5.12)

where

SHT
CL0 = −CLαt (it + 0 ) (5.13)
Sref

The conditions for the aircraft trim are given in (5.14). Solving this system
of equations (3.11) and (5.12), the set of angle of attack and tail deflection angle
that trims the aircraft for each flight condition can be obtained.
It can be observed from the CLtrim condition that the relevant set of angles
is dependent on the flight speed V , flight altitude (through density ρ) and the
instant aircraft mass m. The variation of the δt with the flight speed and the
CG location can be seen in figures 5.3, for m = 27000 kg, which is a value
corresponding to the aircraft mass at the beginning of the cruise phase.

2W
CLtrim =
ρV 2 Sref (5.14)
Cmtrim = 0

Figure 5.3 – Variation of δt to trim with the flight speed and the static
margin at SL flight.

67
The δt and α variation with the flight speed for different flight altitudes is
presented in figures 5.4 and 5.5, respectively. The given results correspond to a
static margin of 0.1. The flight speed range of the presented results and of the
trim analysis in general, is restricted from the critical Mach number, which sets
the upper limit, and from the aircraft CLmax for each flight condition, which sets
the relevant lower limit.

Figure 5.4 – Variation of δt to trim with the flight speed and the flight
altitude for the subsonic Mach number range.

Figure 5.5 – Variation of αtrim with the flight speed and the flight
altitude for the subsonic Mach number range.

68
5.2 Supersonic Static Longitudinal Stability
The criteria for static longitudinal stability remain the same for the supersonic
flight, although the relevant aerodynamic coefficients are changing. The most
important change, during the supersonic flight compared to subsonic speed,
is the aft movement of the wing’s aerodynamic center. For this analysis, the
aerodynamic center of the wing-body has been assumed to be at 45% of the
mean aerodynamic chord, which corresponds to an aft movement of 20%. The
aerodynamic center of the horizontal tail has been place to the 45% location of
its mean aerodynamic chord as well [9].

5.2.1 Supersonic Neutral Point


The aforementioned change of the aerodynamic center position for supersonic
flight affects the aircraft neutral point, which moves aft too. From the equation
(5.7), the NP longitudinal location for supersonic flight has been computed to
be at about 15.9 m, which is 1.3 m aft in comparison with the subsonic NP,
which corresponds to 0.25 the wing’s aerodynamic chord, which is the value of
increment of the static margin for this design.

5.2.2 Trim Analysis


The aircraft total CL and Cm , like for case of the subsonic flight, are given
from the equations (5.11) and (5.12). These equations hold for angles of attack
lower than 5 deg, since for larger angles of attack, the linearized supersonic theory
that was used to obtain the relevant aerodynamic coefficients is not valid. The
trim conditions are given in (5.14) as well.

Figure 5.6 – Variation of δt to trim with the flight speed and the flight altitude
for the supersonic Mach number range and static margin of 0.1.

69
Similarly to the subsonic trim analysis, the mass of aircraft at trim is considered
27000 kg. In figures 5.6 and 5.7, the variation of the δt with the flight speed and
altitude is illustrated, for static margin 0.1 and 0.35, respectively. The relevant
variation of the angle of attack for static margin of 0.1 is illustrated in figure 5.8.

Figure 5.7 – Variation of δt to trim with the flight speed and the flight altitude
for the supersonic Mach number range and static margin of 0.35.

Figure 5.8 – Variation of αtrim with the flight speed and the flight altitude for
the supersonic Mach number range.

5.3 Longitudinal Center of Gravity Location


The longitudinal center of gravity location, like it has been explained before,
should be forward of the neutral point. The typical values for the static margin
are 0.05 - 0.15, since a CG shift of 10% that can take place during flight is

70
considered tolerable. Thus, setting a static margin of 0.1 the CG has to be
placed at 14.1 m for subsonic flight.
During the supersonic flight due to the neutral point aft movement, the static
margin will increase to 0.35. Comparing the values for δt from figures (5.6) and
(5.7), it can be seen that the tail deflection angle increases substantially due to
the relevant static margin increment at supersonic flight. This is an undesired
situation, since the larger δt is a factor that leads to increased drag during
cruise. For that reason, the aircraft CG has to be moved aft too, so that the
static margin is kept inside the desired range. This can be achieved through fuel
shifting, like in the case of Concorde. Therefore, having a static margin of 0.1
after the fuel shift, means that the location of the CG during supersonic flight
has to be at 15.4 m.

5.4 Directional Stability


An analysis for the directional or weathercock stability, for subsonic speeds
and fixed-rudder, has been accomplished in order to evaluate the vertical stabilizer
design. The requirement for static directional stability is that the directional
stability derivative, or so-called yaw stiffness, is positive (Cnβ > 0). That means
that on the aircraft will act restoring moments that will tend to decrease a
positive sideslip angle (β). For the normal case that both engines are operating,
the yaw stiffness contributions come from the wing, the fuselage and the vertical
stabilizer. The Cnβ can then be expressed as

Cnβ = Cnβf us + Cnβw + CnβV T (5.15)

The fuselage, the wing and the vertical tail contributions, for the subsonic
range of Mach numbers, can be estimated using the equations (5.16), (5.17) and
(5.18) [30]. In equation (5.16) Volf is the fuselage volume, df is the fuselage
mean depth and wf is the fuselage mean width.

V olf df
Cnβf us = −1.3 in radians (5.16)
Sref b wf

1 tan Λc/4
Cnβw = CL2 − ·
4πAR πAR(AR + 4 cos Λc/4 )
!! (5.17)
AR AR2 sin Λc/4
cos Λc/4 − − + 6(hn − h) in radians
2 8 cos λc/4 AR

∂σ qV T
 
CnβV T = VV T CFβV T 1+ (5.18)
∂β q
where
∂σ qV T 3.06(SV0 T /Sref ) zwf
 
1+ = 0.724 + + 0.4 + 0.009AR (5.19)
∂β q 1 + cos Λc/4 df

71
The wing-fuselage combination at a sideslip angle creates a sidewash σ, which
changes the effective angle of the vertical tail. The sidewash that comes from
the fuselage is the dominant factor, in comparison with the wing, and gives a
stabilizing air above the fuselage wake [34]. Hence, the effect of the sidewash is
stabilizing for a low-wing aircraft. The ratio (qVT /q) is actually equal to the
square of the velocity ratio (VVT /V ), and accounts for the propeller slipstream
effect [28]. For the case of the propeller absence this ratio equals to one.
The contribution of the two aforementioned parameters to the CnβVT can be
estimated analytically from the relation (5.19), where SVT 0 is the vertical tail
area including the area extended to the fuselage centerline and zwf is the vertical
distance of the root chord to the fuselage centerline, being positive for wing root
0
vertical location below it. For this design, the SVT equals to 11.47 m2 and the
zwf to 1.1 m.
The VVT is the vertical tail volume coefficient, which can be obtained from
equation (2.32), where LVT is the distance of the vertical tail aerodynamic center
to the aircraft center of gravity. Finally, the CFβVT is the lift-curve slope of the
vertical tail. This can be computed from the equation (3.3), like in the case of
the wing, introducing the relevant properties of the vertical tail. However, in
this equation it has to be introduced the effective and not the geometric aspect
ratio of the vertical tail. The effective aspect ratio of the vertical tail is 1.55
times greater than the geometric one, and this increment comes from the end
plate effect of the horizontal tail mounted below the vertical stabilizer [34].
The calculated subsonic directional stability derivatives (Cnβ ) are presented
in table 5.1. These values are greater than the suggested values of NASA TN
D-423, as presented in reference [9], meaning that the aircraft will be directionally
stable enough.

Mach number Cnβ (deg −1 )


0.2 0.0025 + 0.00094·CL2
0.3 0.0025 + 0.00094·CL2
0.4 0.0026 + 0.00094·CL2
0.5 0.0026 + 0.00094·CL2
0.6 0.0027 + 0.00094·CL2
0.7 0.0028 + 0.00094·CL2
0.8 0.0029 + 0.00094·CL2

Table 5.1 – Directional stability derivatives for the subsonic Mach


number range.

72
6. Performance

6.1 Climb Performance


For the climb performance analysis, the equations of motion for the rigid and
mass point geometry body have been utilized. The thrust model has been created
according to the thrust lapse expression from the equation (2.19), which gives
the variation of the thrust with the Mach number and the flight altitude. The
relation of the dry thrust has been used for this analysis, since the aim is the total
avoidance of using the afterburner for reasons of increased fuel consumption and
noise. However, the afterburner could be installed and used for safety reasons,
for instance in the case of an emergency, like the loss of one of the engines.
The aerodynamic model of the aircraft has been based on the aerodynamic
coefficients computed in previous chapter. The trim drag has been neglected
for the performance analysis of this section as well as for the following ones.
The aircraft center of gravity has also been assumed not to be influenced from
the fuel mass burned during flight. The equations describing this model, for the
engine installation angle of zero degrees, can be written as

mV̇ = T cos α − D − mg sin γ (6.1)

0 = T sin α + L − mg cos γ (6.2)

ḣ = V sin γ (6.3)

xE = V cos γ (6.4)

ṁ = −C · T (6.5)

With the assumption that the acceleration perpendicular to the flight path
is negligible, made in equation (3.2), the nonlinear equation can then be solved
to obtain the trim angle of attack. The thrust specific fuels consumption (C ),
in the absence of the engine data for the altitude and Mach range of interest,
can be approximated using the relations provided from reference [10], for the dry
and wet thrust of a low-bypass ratio turbofan engine, respectively.


Cdry = (0.9 + 0.3M ) θ (6.6)

73

Cwet = (1.6 + 0.27M ) θ (6.7)

The C in the above equations is given in fps units, i.e. 1 /hr, with the
parameter θ to have been already defined in equation (2.15).

6.1.1 Minimum Time to Climb


In order to find the optimal path that corresponds to the minimum time for
the climb and the acceleration phase, the flight envelope including the specific
excess power (SEP) contours for different altitudes has to be created. The SEP
is given from the formula

(T cos α − D)V
SEP = (6.8)
W

The SEP is defined as the excess power divided by the weight, where the
excess power is just the excess thrust, (T cos α − D), times the velocity, V. This
excess power can be used for altitude gain or for acceleration. For load factor
equal to one and considering steady and level flight, which means that the flight
path angle γ = 0 , the relevant flight envelope can be created for a specific aircraft
weight. The SEP contours of this SST design for maximum payload and 85 %
fuel mass (i.e. aircraft total mass of 27240 kg) are presented in figure 6.1. In
this figure, the stall angle of attack and maximum dynamic pressure limitations
have also been included. The stall angle of attack can be computed by dividing
the clean configuration CLmax with the aircraft CLα , with the relevant value to
be computed as 16 deg. The maximum dynamic pressure has been set to 85.82
KP a, which is the dynamic pressure corresponding to 1.1 Mach at sea level.

Figure 6.1 – SEP contours diagram (dry thrust).

74
The SEP contours of figure 6.1 are actually corresponding to the aircraft
climb rate. In order to obtain the minimum time to climb, the path to be
planned has to follow the maximum aircraft climb rates. For that reason the
aircraft has to accelerate at the same altitude till Mach 0.9, where the climb
with maximum climb rate will take place. The aircraft is climbing till about the
desired cruise height and then it dives, so that it accelerates and passes through
the transonic region, where the drag is too high that does not allow the aircraft
to accelerate any more, due to thrust deficiency. After the aircraft acceleration,
the altitude is regained, since the drag becomes lower in supersonic speeds. The
dive follows the relevant constant energy height curves, for which the dynamic
energy is purely converted to kinetic energy. The final part of the trajectory
includes the climb to the final cruise altitude and the acceleration to the cruise
Mach number.

In figure 6.2, the minimum time to climb trajectory is presented. The flight
path angle γ as a function of time has been defined as the control variable that
has been used to solve the set of differential equations (6.1) - (6.5), with the
initial conditions to be 0.1 km flight altitude and 0.3 Mach flight speed. The
fuel mass at the start of the climb has been set to 97 % of the total fuel capacity,
since a 3 % or 452 kg of the fuel has been assumed to be used during the taxiing
and takeoff. The time needed for this climb at 14.55 km and Mach 1.7, is about
1430 sec, where the fuel that has been burned has been estimated as 3399 kg
or 22.6 % of the total fuel capacity. It has to be mentioned that the initial
cruise altitude that can be reached is 450 m lower than the specified from the
requirements altitude of the 15 km.

Figure 6.2 – Flight path for minimum time to climb at cruise conditions.

75
6.1.2 Minimum Fuel to Climb
The flight path giving the minimum fuel consumption to climb is a bit
different in comparison with the minimum time to climb one. The minimum
fuel to climb trajectory follows the path that maximizes the fuel specific energy
(FSE) for each energy height. The fuel specific energy is defined as the change
in specific energy per change in fuel weight [9]. The FSE is defined as

SEP
F SE = (6.9)
C ·T
The FSE contours and the relevant flight path is presented in figure 6.3. The
γ(t) is the control variable in this case as well, while the same set of equations
has been solved, like for the minimum time to climb trajectory. The initial flight
and weight conditions have been considered the same too. The time needed to
reach the cruise flight conditions of 14.55 km and 1.7 Mach has been computed
as 1580 sec. For this climb, the mass of the burned fuel is 3205 kg or 21.3 %
of the total fuel capacity. The fuel savings following this trajectory compared
to the minimum time to climb one are about 1.3 % of the total fuel capacity.
The horizontal range covered during the climb and acceleration phase has been
calculated as 517.5 km.

Figure 6.3 – Flight path for minimum fuel to climb at cruise conditions.

6.2 Range
The horizontal range of aircraft during cruise can be calculated from the
Breguet range equation (6.10). During the cruise the aircraft is flying at constant
CL and speed. However, keeping these two parameters constant means that the
aircraft due to the fuel burn, which is reducing its weight during the cruise, will
gradually climb to higher altitudes. Moreover, the aircraft is possible to fly at

76
constant speed and Mach number, since the speed of sound for cruise altitudes
in stratosphere remains unchanged.
!
V L Winit
R= ln (6.10)
CD Wf inal

Since the target is the range maximization, the minimum fuel to climb flight
path is the desired one to be followed before the cruise. Knowing the relevant
fuel that has been burned during the climb the Winit can be easily obtained. At
the end of the cruise, it is assumed that the aircraft has a remaining 12% of the
total fuel capacity or Wfinal equal to 16248 kg.
For the cruise flight conditions of 1.7 Mach at the height of 14.55 km, the
lift-to drag ratio (L/D) has been calculated as 5.764. Therefore, the horizontal
range to be covered during cruise is estimated as 5068.9 km, while the flight
altitude at the end of the cruise as 17.49 km.

6.3 Descent and Loiter


After the end of the cruise phase, the aircraft needs to decelerate and descend
to a lower altitude, where the loiter before the final descent to land will take place.
For the descent the same model has been used, like in the case of the climb,
with the difference that the thrust provided from the engine is decreased. That
means that the excess power becomes negative and the aircraft decelerates, while
a negative γ initiates the dive. As it can be seen from the endurance equation
(6.11), the loiter should take place at the flight conditions where the product of
the (L/D)max with the C is maximum. Thus, the chosen conditions for the loiter
have been the 6.8 km altitude and the 0.4 Mach. The thrust has been reduced
to 20% of the maximum thrust for this descent. When the loiter phase ends,
the aircraft has to descend again to land. The final flight conditions before the
approach have been modeled to be the 0.2 km height and the 0.3 Mach speed,
using in this case the 10% of the engines maximum thrust.
!
1 L Winit
E= ln (6.11)
CD Wf inal

During the first descent the fuel burned and the horizontal range covered are
207 kg and 253.5 km, respectively. During the loiter, a total 441 kg of fuel has
been burned, corresponding to an endurance of 24 min and 14 sec. The final
descent has consumed 179 kg of fuel, covering an additional horizontal range of
159.4 km. The aircraft weight before the final approach has then been computed
to be 15421 kg, with 6.51% of the total fuel to be left for usage during the landing
and taxiing phases.

77
6.4 Takeoff
The takeoff distance, as can be seen from figure 6.4, consists of two basic
components, the ground run distance and the airborne distance.

Figure 6.4 – Illustration of takeoff path and distance.

The ground run distance includes the ground roll distance (dg ) and the
rotation distance (dr ). During the ground roll, the aircraft accelerates from
the zero velocity till the takeoff velocity (VTO ), which should be 1.1 times the
stall speed (Vstall ) [9]. The Vstall can be simply calculated setting the lift equal
to weight and using the CLmax for flaps in the takeoff position (about 80% of
the landing CLmax ). The relevant value obtained is 71 m/s. The dg can then be
calculated from the integral
Z VT O
V
dg = dV (6.12)
0 α

where the acceleration α is given from the relation

T ρ 
   
α=g −µ + −CD0 − KCL2 + µCL V 2 (6.13)
W 2W/S

For the rolling friction coefficient µ, the typical value of 0.05 for brakes off
has been used [13]. The total CD0 at the takeoff has been estimated as 0.0472,
including the flaps, slats and landing gear contributions. The lift coefficient CL
being based on the wing angle of attack during the ground roll is typically small.
The conservative assumption that is negligible has been made for this analysis.
Moreover, the trust provided during the ground run is not constant. For that
reason, in order to obtain more accurate results, the integral has been broken
into smaller segments for which average thrust has been used.
The rotation distance dr can be simply obtained multiplying the VTO with
the typical rotation time tr of 3 sec. The total distance during the ground run
has then been calculated as 1927 m.
In order to obtain the total takeoff distance, the dab has to be estimated
as well. During this airborne phase the aircraft is accelerating from the takeoff
speed (1.1Vstall ) to the climb speed (1.2Vstall ) [9]. Moreover, the aircraft gains

78
altitude, in order to achieve the 35 ft obstacle clearance (hob ), defined from the
FAR regulations.
The load factor (n) during the pull-up can be calculated from the equation
(6.14), where the 90% of the takeoff CLmax is used as a safety margin [13]. Then,
the correspondent turn radius (R) can be calculated from the equation (6.15).
Since the speed is not constant, the load factor and thus the relevant turn radius
is changing. For that reason, they have been computed for the given range of
speeds, and then the mean value of the turn radius has been used.
1 2
2 ρS(0.9CLmax )V
n= (6.14)
W

V2
R= (6.15)
g(n − 1)

Having estimated the pull-up R, the airborne distance can be calculated as


q
dab = R2 − (R − hob )2 (6.16)

The relevant value obtained for the dab is 289.2 m, which means that the
total takeoff distance is estimated as 2216.2 m.

6.5 Landing
The landing distance includes the approach distance (da ), the flare distance
(df ), the free roll distance (dfr ) and the ground roll distance (dg ). The landing
phases can be observed in figure 6.5.
The approach angle θa is usually small, and for a transport aircraft should
be θa ≤ 3 deg [9]. The θa can then approximated from the equation (6.17). In
order to satisfy the aforementioned constraint for the approach angle, the 63%
of the thrust has been used during landing, resulting in a θa equal to 2.79 deg.

1 T
sin θa = − (6.17)
L/D W

During the approach and flare phase, the aircraft should decelerating from
the approach speed (Va = 1.3Vstall ) to the touchdown speed (Va = 1.15Vstall ).
The Vstall is the one obtained using the CLm ax corresponding to flaps in landing
position and is equal to 64.5 m/s. The approach turn radius (R) can then be
computed using the relations (6.14) and (6.15), like in the takeoff case, for the
previously stated variation of velocities. The approach distance, taking then into
account a clearance distance of 50 ft [13], is approximated as

15.24 − R(1 − cos θa )


da = (6.18)
tan θa

79
Figure 6.5 – Illustration of landing path and distance [13].

The flare and the free roll distance can be approximated using the relations
(6.19) and (6.20), respectively [13].

df = R sin θa (6.19)

df r = tr VT D (6.20)

Finally, the ground roll distance can be computed using from the integral
Z 0
V
dg = dV (6.21)
VT D α

where α is given again from the equation (6.13). The integral has been computed
using the same method, as for the takeoff. The total CD0 at the landing has
been estimated as 0.175, including the flap, slats, landing gear and spoilers
contribution. A spoilers deflection of 45 deg has been assumed. The value of
rolling friction coefficient used is the one corresponding to brakes usage. The
typical value of 0.5 has been used for the µ [13].
The total landing distance can be estimated by just summing all the relevant
contributions. The obtained value is then increased, so that the FAA requirements
that allow for pilot technique are met [9]. Hence, the final landing distance can
be estimated from the equation (6.22) as 2147.3 m .

80
5
dland = (da + df + df r + dg ) (6.22)
3

6.6 Total Mission


The evaluation of the above results shows that the aircraft is meeting the
requirements regarding the takeoff and landing distances. The total horizontal
range is 6053 km and the maximum endurance about 24 min. The maximum
range without any loiter can reach the 6354 km.
The obtained range is lower than the desired one, but yet enough to execute
transatlantic flights. The main focus for getting a bigger range should be the
improvement of the lift-to-drag ratio during cruise. A slight increment of the
(L/D)cruise to 6.2 from 5.764 would result to a total horizontal range of 6437
km or an increment of 384 km. This value assigned for the (L/D)cruise is quite
reasonable and feasible to be obtained, when compared with the relevant value
of the Concorde. According to reference [36], the Concorde had a (L/D)max of
7.5 for cruise at Mach 2, which corresponds to a (L/D)cruise of about 6.5. The
(L/D)cruise of the aircraft could be increased even more through camber and
wing twist optimization.
For a flight between London and New York, the total time of the climb, cruise
and descent is 3 hrs and 35 min. Assuming that no loiter is needed, which is the
case most of the times and including the time for taxiing, take-off and landing,
the time needed for the whole mission is 4 hrs maximum. Given that the normal
time of a subsonic airliner for that route is 7 hrs and 40 min to 8 hrs, the time
saving is very significant. Although the SST cannot follow the minimum fuel to
climb flight path for practical reasons, like the overland sonic boom ban, which
might impose an initial small subsonic cruise part on the mission, the time saved
remains still substantial and would be about 3 hrs and 30 min for the whole trip.

81
7. Test Flight

7.1 Simulation
The flying and handling qualities of the design have been flight tested using
the Merlin MP521 Engineering Flight Simulator. The MP521 Simulator comprises
a capsule with a six axis motion system, visual and instrument displays, touch
control panels, and hardware flight controls [37]. The software of the flight
simulator that has been run is the Excalibur II, while the design has been created
at the Excalibur Data Editor.
The model in the editor has utilized the geometrical data and mass properties
of the design, in addition to the aerodynamics and propulsion data, acquired
during the conceptual design. The mass moments of inertia are calculated by
the editor, using empirical relations, from the inserted values that correspond to
the aircraft mass parameters, center of mass and wing span, assuming one plane
of symmetry, i.e. Ixy = Iyz = 0.
The wing is split into panels for which the relevant data are inserted to the
editor. The wing is divided according to its control surfaces, particularly there
is a change when a control surface is met, so that there can be lifting surfaces
with or without control surface. In this case, the wing semi-span has been split
in a total six panels, where the other half includes six more panels, which are
automatically mirrored due to symmetry. For the horizontal tail, just one panel
has been used, since it has been modeled as a whole moving surface, while two
more panels have been considered for the vertical tail. For each panel the control
surfaces data, and the wing geometrical characteristics, in addition to the used
airfoil aerodynamic data, have been specified. The aerodynamic center of each
panel has to be independently calculated and set in the model editor too.
The model includes excessive data for the undercarriage as well, such as the
nose-wheel steering and the brakes, determining the aircraft ground performance.
These data have been specified using typical values recommended from the
software’s user guide manual, the Excalibur II Flight Model Editor Data Definitions.
The software of the flight simulator is based on the six degrees-of-freedom
equations of motion of a rigid body, and thus does not take into account any
aeroelastic effects. The translational positions are integrated in earth (NED)
axes and the angular positions (aircraft attitudes) are represented in the standard
four-parameter Quaternion format, in order to allow the attitudes be integrated

82
through the 90 degrees pitch attitude singularity.
The simulator includes a flight data recorder (FDR) with a recording frequency
that can be set to 1, 5 or 25 Hz. The flight data are held in an ASCII text file,
including speed and flight path variables, aircraft control deflections, rigid body
variables and atmosphere parameters. The recorded data have then been read
and processed using Matlab.

7.2 Flying and Handling Qualities


Aircraft handling is related with the aircraft response to the control inputs,
as for instance the elevator deflection, which can be discerned in short-term and
long-term response [38]. The short-term characteristics of the aircraft are very
essential, since poor behavior can make the aircraft very hard or even impossible
for the pilot to handle. Thus, the primary goal when evaluating the flight and
handling qualities of the design is to achieve satisfactory short-term response,
which are related to its short period dynamic modes. The long-term response is
substantial in maintaining steady flight of the aircraft and is determined from
the static stability and its long period dynamic modes [38]. However, these
modes having a long period, can be handled a lot easier by the pilot, so that
even marginally unstable modes can be considered satisfactory.
Taking the above into consideration, the flight test of the design aimed to the
dynamic modes evaluation, in terms of both longitudinal and lateral-directional
dynamic stability. The aircraft has been trimmed by determining the necessary
throttle setting and horizontal tail deflection for each flight condition being
considered. Then each mode has been excited and the results have been recorded
using the simulator’s FDR. For the executed measurements, the highest recording
frequency of 25 Hz has been implemented. Finally the obtained results have
been rated using the relevant guidelines and margins as specified in the Military
Specification MIL-F-8785C.
The aircraft flying and handling qualities have been evaluated for three
different altitudes, in particular at 10, 20 and 30 kf t, and for subsonic Mach
numbers. The design was tried to be assessed for the supersonic cruise conditions
as well, however the results obtained from the simulation have been considered
as inaccurate and hence will not be presented.

7.2.1 Modes excitation


The dynamic stability modes of the aircraft can be discerned in two categories,
the longitudinal and the lateral-directional dynamic modes, which are uncoupled.
Detailed guidance about the procedure of the excitation for both types of dynamic
stability modes during flight is given in reference [38], which is briefly described
in this subsection. These procedures help to measure and quantify the relevant
modes properties during test flight in a way that they can be comparable with
the analytical values.

83
The longitudinal dynamic stability modes consist of the short-period pitching
mode and the long-period or phugoid mode, which are both related to oscillatory
motion. The two modes have a significant difference as regards their frequency,
which makes possible their independent excitation. The short-period oscillation
can be excited by applying a short duration disturbance in pitch to the trimmed
aircraft. In order to achieve this a unit impulse is applied to the elevator
for about one sec. This impulse will more likely excite the phugoid mode
too. However, due to its low frequency, which requires significantly more time
in comparison with the short-period mode to develop the phugoid oscillatory
motion, the short-period measurements will not get affected.
In order to excite the phugoid mode, a small step input is applied to the
elevator, which causes the aircraft to accelerate, when descending, while the
thrust level setting is kept constant. The elevator is returned to its initial
position, when the aircraft speed has increased by about 5% compared to relevant
speed at the trimmed condition. It has to be mentioned that it is essential this
speed disturbance to be small for the results to be accurate, since the modes
are being evaluated using the small-perturbation model, which should not be
violated.
In contrast with the longitudinal dynamic modes, the lateral-directional ones
are more difficult to excite independently, due to mode coupling. For that reason,
the aileron or rudder input to excite these modes should be applied more carefully
and accurately for the relevant measurements to be taken. The lateral-directional
stability modes consist of the spiral, the roll subsidence and the dutch roll mode.
The first two are related to real eigenvalues, thus no oscillation is observed,
while the dutch roll mode is related to an imaginary eigenvalue, such as the
short-period and the phugoid mode, which includes oscillatory motion as well.
The spiral mode is excited when a small step input is applied to the rudder.
The aircraft then begins to turn and the inner wing of the turn side drops.
When the roll attitude is about 20 deg, the rudder returns to its initial position.
If the mode is stable, the aircraft will converge to its zero roll attitude angle,
and recover wings level. Differently, for the case of an unstable mode, the roll
attitude angle of the aircraft would contrarily increase and thus diverge from
the initial balance condition. The spiral mode is a weak mode with a big time
constant, and thus does not affect substantially the flying and handling qualities
of the aircraft, so that unstable modes can be acceptable too.
The roll subsidence mode can be excited when a square pulse is applied to the
aileron. In order to specify this dynamic mode during flight test, the roll attitude
angle is set initially at about -30 deg. Then the relevant aileron deflection is
applied, so that the aircraft rolls steadily to the +30 deg roll attitude, when the
aileron has returned to the neutral position. The roll subsidence mode is the one
that governs the transient exit of the steady part of the rolling motion. Its time
constant is short and its effect is visible just in the roll attitude response.
The dutch roll mode can be excited by applying a doublet to the rudder
pedals. The period of the rudder deflection should approximately match the

84
period of the mode for it to be excited. During this cyclical rudder deflection,
the aircraft is compelled to a forced oscillation. After returning the rudder
to the neutral position, the aircraft will continue to execute a free oscillation,
which corresponds to the oscillatory motion related to the dutch roll mode. The
dutch roll mode having short period should be stable and demonstrate adequate
damping, so that the aircraft flying qualities can be considered satisfactory.

7.2.2 Dynamic Stability Requirements


A common method for the flying and handling qualities evaluation of the
aircraft is the so-called Cooper-Harper rating scale. This is a qualitative rating
scale that expresses the pilot opinion. The pilot rates the aircraft behavior,
respective with the different flying phases that are needed to be tested, using
an 1 to 10 scale, where the lower the grade the better the flying and handling
qualities the aircraft exhibits.
However, in order for the results to be impartial, the modes obtained from
the flight test will be quantified, following the guidelines given in the previous
subsection. Then the calculated modes will be compared to the requirements
presented in reference [39], which correspond to specific values that determine
the margins for the aircraft’s flying and handling qualities assessment.
Before presenting the relevant requirements, the aircraft has to be classified
and categorized, according to instructions given in [39]. The classification is
related with the aircraft’s role, while its category with its flight mission profile.
For this case, the SST that needs to be evaluated is a medium weight and
low-to-medium maneuverability aircraft, which means that it belongs in Class
II. The mission profile of the SST is in accordance with the flight phases defined
for a category B aircraft, for which accurate flight-path control is required and
gradual, without precision tracking, maneuvers accomplishment.
The dynamic stability requirements, for both the longitudinal and lateral-
directional modes of an Class II and Category B aircraft, are presented in the
tables 7.1 - 7.5 and in figure 7.1. The time to double or half is the time required
for the initial perturbation of the trimmed condition to be doubled or halved
and is defined, for the case of an oscillatory motion, in equation 7.1, where the
symbol n corresponds to the real part of the relevant eigenvalue. The damping
ratio ζ is defined in relation 7.2, and the ωn is the undamped angular frequency
given from the relation 7.3 [28]. Therefore, obtaining the period (T ) of the
oscillation and the thalf or tdouble after processing the recorded flight data, both
ζ and ωn can be estimated using the aforementioned equations.

ln2 ln2
tdouble or thalf = = (7.1)
|n| | ζ | ωn

ζ = −n/ωn (7.2)

 0.5
ωn = ω 2 + n 2 (7.3)

85
Level Min ζ Max ζ
1 0.30 2.00
2 0.20 2.00
3 0.15 -

Table 7.1 – Short-period mode damping ratio limits.

Figure 7.1 – Short-period mode frequency requirements [39].

86
In figure 7.1, the parameter n/α correspond to the normal load factor per
unit angle of attack α.

Level Requirement
1 ζ > 0.04
2 ζ >0
3 T > 55 sec

Table 7.2 – Phugoid mode stability requirements.

Level tdouble
1 < 20 sec
2 < 8 sec
3 < 4 sec

Table 7.3 – Minimum time to double amplitude limits for spiral mode.

Level Time constant


1 < 1.4 sec
2 < 3.0 sec
3 < 10 sec

Table 7.4 – Roll subsidence mode time constant limits.

Level Min ζ Min ζ · ωn (rad/sec) Min ωn (rad/sec)


1 0.08 0.15 0.4
2 0.02 0.05 0.4
3 0 0 0.4

Table 7.5 – Dutch roll mode damping ratio and frequency limits.

7.2.3 Longitudinal Dynamic Stability

Short-period pitching mode


The short period pitching mode has been measured for the aforementioned
flight altitudes and for the Mach numbers of 0.5 and 0.8, except for the altitude
of 30 kf t, where due to the aircraft stall restriction, the lower Mach number has
been set to 0.6. The same flight conditions have been used for the other modes
measurements too.

87
In figures 7.2 and 7.3, the elevator impulse for the short period mode excitation
and the resulting oscillation as depicted in the body axis pitch rate are presented
for the stated flight conditions.

Figure 7.2 – Elevator impulse input flight recording of the short


period mode for 0.6 Mach at 30 kft.

Figure 7.3 – Body axis pitch rate flight recording of the short
period mode for 0.6 Mach at 30 kft.

The obtained values corresponding to the short-period pitching mode, after


processing the data of the flight recorder for all the tested flight conditions, as
well as their evaluation regarding the aircraft’s flying and handling qualities, can
be seen in table 7.6.

88
Mach Altitude (kf t) T (sec) thalf (sec) ωn (rad/sec) ζ Level
0.5 10 1.74 0.73 3.74 0.254 2
0.8 10 1.06 0.46 6.12 0.249 2
0.5 20 2.09 0.92 3.10 0.242 2
0.8 20 1.35 0.60 4.80 0.242 2
0.6 30 2.02 1.03 3.18 0.212 2
0.8 30 1.68 0.76 3.85 0.236 2

Table 7.6 – Variation of short-period pitching mode characteristics with


speed and altitude.

Phugoid mode
Similarly to the short-period mode, the relevant values of the phugoid mode
characteristics are presented in table 7.7, while in figure 7.5, the oscillatory
motion of the mode as regards the aircraft’s true airspeed is illustrated.

Mach Altitude (kf t) T (sec) thalf (sec) ωn (rad/sec) ζ Level


0.5 10 84.9 88.49 0.074 0.105 1
0.8 10 145.8 60.05 0.045 0.258 1
0.5 20 79.7 99.26 0.079 0.088 1
0.8 20 127.4 73.34 0.050 0.188 1
0.6 30 83.2 113.69 0.076 0.080 1
0.8 30 115.6 110.23 0.055 0.115 1

Table 7.7 – Variation of phugoid mode characteristics with speed and


altitude.

Figure 7.4 – Elevator step input flight recording of the phugoid


mode for 0.6 Mach at 30 kft.

89
Figure 7.5 – True airspeed flight recording of the phugoid mode for
0.6 Mach at 30 kft.

7.2.4 Lateral-Directional Dynamic Stability

Spiral mode

In figure 7.6, the flight recording for the roll attitude angle can be observed
for the stated flight conditions. It can be seen from the graph that the spiral
mode for this case is convergent to zero roll angle attitude, and thus stable.
In table 7.8, the spiral mode characteristics for the flight tested conditions are
presented.

Figure 7.6 – Euler roll angle flight recording of the spiral mode for
0.6 Mach at 30 kft.

90
Mach Altitude (kf t) thalf (sec) Level
0.5 10 26.60 1
0.8 10 43.05 1
0.5 20 27.51 1
0.8 20 47.92 1
0.6 30 32.44 1
0.8 30 52.95 1

Table 7.8 – Variation of spiral mode characteristics with speed and


altitude.

Roll subsidence mode

In figures 7.7 - 7.9 are presented the graphs obtained for the roll subsidence
mode from the flight test at the stated flight conditions, regarding the aileron
input, the roll attitude angle and the roll rate. In table 7.9, the roll convergence
time constant and evaluation regarding the previously specified flying and handling
qualities requirements are presented.

It can be observed in figure 7.9, that the the roll response stabilizes when
the moment due to damping in roll is the opposite to the disturbing moment in
roll caused by the aileron deflection [38]. Moreover, from figure 7.8 can be seen
that after the roll angle becomes steady, it starts decreasing. This is an effect of
the spiral mode, which tends to recover the aircraft’s wings level, after returning
the aileron to its zero deflection position.

Figure 7.7 – Aileron input flight recording of the roll subsidence


mode for 0.6 Mach at 30 kft.

91
Figure 7.8 – Euler roll angle flight recording of the roll subsidence
mode for 0.6 Mach at 30 kft.

Figure 7.9 – Body axis roll rate flight recording of the roll subsidence
mode for 0.6 Mach at 30 kft.

Mach Altitude (kf t) Time constant (sec) Level


0.5 10 1.81 2
0.8 10 1.53 2
0.5 20 2.17 2
0.8 20 1.92 2
0.6 30 2.92 2
0.8 30 2.36 2

Table 7.9 – Variation of roll subsidence mode characteristics with


speed and altitude.

92
Dutch roll mode

In figure 7.10, the rudder doublet for the dutch roll excitation can be seen.
In figures 7.11 and 7.12, the dutch roll oscillation is depicted in the roll and yaw
rate graphs, for the stated flight conditions. From these two graphs, it is visible
the expected phase shift between the roll and the yaw rate too. It can be seen
from the graphs that the dutch roll oscillation is the motion starting for this
case after the first 4 sec. In the first 4 sec, the observed oscillation is the one
enforced from the rudder input. In table 7.10, the dutch roll characteristics for
the tested flight conditions are presented.

Figure 7.10 – Rudder input flight recording of the dutch roll mode
for 0.6 Mach at 30 kft.

Figure 7.11 – Body axis roll rate flight recording of the dutch roll
mode for 0.6 Mach at 30 kft.

93
Figure 7.12 – Body axis yaw rate flight recording of the dutch roll
mode for 0.6 Mach at 30 kft.

Mach Altitude (kf t) T (sec) thalf (sec) ωn (rad/sec) ζ Level


0.5 10 1.35 1.04 4.702 0.142 1
0.8 10 0.84 0.76 7.535 0.120 1
0.5 20 1.65 1.53 3.835 0.118 1
0.8 20 1.04 1.21 6.069 0.094 1
0.6 30 1.74 2.00 3.628 0.095 1
0.8 30 1.28 1.66 4.927 0.085 1

Table 7.10 – Variation of dutch roll mode characteristics with speed and
altitude.

94
8. Environmental Impact

8.1 Sonic Boom


One of the most important environmental impacts of the SST operation is the
aerodynamic noise generated by the shock waves. The so-called sonic boom has
imposed a ban on the overland supersonic flight, which can now be performed
only overseas. NASA is conducting research in this field in order to create a low
boom design, which could be quiet enough to overcome the applied restrictions.
The early sonic boom research was conducted according to modified linear
methods, based on Whitham’ s theory for the sonic boom prediction. According
to the linear theory, the pressure signatures reaching the ground are N-waves,
which are typical for aircraft with high wing loading. However, later work has
shown that generation of non-N-waves on the ground was possible [40]. Moreover,
some important deficiencies has been recognized regarding the linear methods,
respective with the account for the three-dimensional nonlinear aerodynamics,
the propagation of the sonic boom waveforms through a real atmosphere, having
thus variable ambient conditions, and the atmospheric turbulence modeling,
which made necessary the development of more accurate methods and their
experimental validation through flight testing. Therefore, the later studies focused
on creating aircraft concepts with lift and volume distributions that would shape
non-N-waveforms at the ground, rather than trying to reduce the noise generated
by the N-waves, which created substantial limitations.
The focus of the High-Speed Research (HSR) program of NASA was the
creation of a big SST, which could carry more than 250 passengers. The idea
behind that concept was that the opening of more supersonic corridors overland
due to a low boom design, would result in a more excessive usage of supersonic
transports. Therefore, the development cost of the aircraft would be smaller
due to increased demand for its purchase. The SST design proposed in this
project refers to a small aircraft carrying 15 passengers. However, some sonic
boom minimization design guidelines could be implemented to smaller aircraft
and evaluated too.
In reference [41], several concepts are examined based on a reference design
as regards the sonic boom loudness. The delta baseline design wing has been
changed into a wing arrow design, having about the same aerodynamic efficiency
The relevant baseline and the low boom design and specifications are presented
in figures 8.1 and 8.2.

95
Figure 8.1 – Drawing and specifications of the baseline configuration [41].

It can be seen from the two drawings that the low boom design incorporates
a bigger wing. It has been found that a bigger wing is beneficial since the
lower wing loading results in reduced pressure levels to sonic boom for the
lift contribution [41]. The result is a reduction of about 7 PLdB between
the two designs. However, this sonic boom loudness reduction comes with a
significant price, which is an increase of about 12 % on the maximum takeoff
weight per passenger. The modified wing planform with the larger area and
the higher wing sweep demands higher structural weight, which makes the low
boom configuration heavier. Thus, it has been become clear that the low boom
design is related with a performance penalty, as an effect of the increased aircraft
structural weight. Moreover, the high wing sweep, in addition to the low aspect
ratio, makes the low speed performance of the aircraft very challenging. The
above observations conclude that there is a trade-off between the aircraft performance
and the sonic boom loudness, which has to be balanced [42].
Furthermore, the low boom design usually leads to the aircraft wave drag
increment. The sonic boom minimization concepts adopt blunt nose, which is
far from the optimal configuration for minimal wave drag. A blunt nose creates
a strong bow shock, so that the secondary shocks are weak and do not overtake

96
and enhance the front shock. The far field pressure signature produced in this
way, thus, comes to be much weaker than in the case of a sharp nose, where
the front bow shock coalesces with the stronger secondary shock waves [43]. In
this case, the nose shape is a trade-off between the sonic boom loudness and the
wave drag magnitude as well.

Figure 8.2 – Drawing and specifications of the low boom configuration [41].

The examination of low boom concepts for aircraft with high payload has
shown that it was very difficult to achieve theoritical ground overpressure substantially
less than 1 psf [44]. For that reason a lighter and smaller aircraft would be
perhaps a better candidate for decreasing the sonic boom loudness. A lighter
aircraft demands a lower amount of lift to sustain level flight, which is a factor
that decreases the sonic boom intensity [45]. It has to be mentioned here that
the sonic boom is measured in psf or Pa of overpressure. For shaped pressure
signatures, reference [46] provides with a method to obtain the relevant perceived
level of noise loudness (PLdB).
As it has been stated before, the wing planform is an important factor for
achieving reduced sonic boom. In the case of the small SST design, still reduced

97
wing loading is needed and thus a large wing. It is important too that the wing
incorporates a long wing root, which will result in the gradual development of
the area and lift. Moreover, the span cannot be decreased too much, in order
to maintain an acceptable performance during the low-speed flight. Normally,
the wing in a low boom design is usually placed well aft in comparison with the
conventional one, so that its interaction with the aircraft nose shock is reduced.
However, this creates serious problems with the aircraft stability. The center of
gravity of the fuel placed in the wing can be at a long distance with respect to
the empty aircraft center of gravity, which can result in large shifts of the total
aircraft center of gravity location [47]. For that reason as much fuel as possible
has to be placed in the front portion of the wing, which due to its increased size
and thus volume will perhaps offer this opportunity. Finally, this fact would not
allow the wing to incorporate very thin airfoils, in order to achieve the desired
volume for the fuel storage, which would subsequently increase the drag.
The fuselage design of a small SST design has also the disadvantage of
the decreased fineness ratio, since the length of the body is smaller but the
maximum diameter cannot be decreased too much in order to be able to house
the passengers and of course the aircraft systems. That has anyway an important
wave drag penalty on the design. Furthermore, the optimal position of the
engines, according to [44], would be the aft fuselage behind the wing trailing
edge. This wing-nacelle interference would be in this way avoided and the flow
field disturbances of the nacelles would correspond just to volume and not to
lift contribution effects. Another advantage would be increased space for the
trailing edge devices placement on the wing. However, the engines support in
this location would add on structural weight as well.

Figure 8.3 – Three view of a low boom SBJ concept [44].

In figure 8.3, a concept for a low boom supersonic business jet (SBJ), capable
of carrying 8-10 passengers, is illustrated. This SBJ design incorporates a
canard, instead of a horizontal tail, and all the relevant specifications of the
concept can be found in reference [44]. Due to its decreased size and weight

98
compared to the bigger SST concepts, which is also a result of flying at Mach
numbers below 2, this design was able to generate significant lower ground
overpressure of about 0.5 psf at cruise start. However, the drag and weight
penalties that are accompanied with the low boom design are becoming even
more apparent. The consequence is an aircraft with lower performance, which
although makes the design environmentally viable, its economical viability is
questioned. Furthermore, the increased weight of the low boom design creates
some difficulties regarding the incorporation of engines with reasonable size and
weight, that would be capable of producing the necessary thrust for satisfying
the aircraft supercruise requirement. The development of engines with improved
performance and the usage of composite materials would be substantial factors
in achieving an economic viable concept as well.

8.2 Air Pollution

8.2.1 Air Pollutants Identification


During the flight, various pollutants substances are being emitted into the
atmosphere and are primarily due to the combustion gases from the propulsion
system. During the combustion process in general the most important pollutant
emissions that can be identified are the carbon dioxide (CO 2 ), the nitrogen oxides
(NO x ), the water vapor (H2 O), the carbon monoxide (CO), the hydrocarbons
(HC ), the sulfur oxides (SO x ) and the soot particles (C ). In reference [48], a
detailed analysis is presented regarding the chemical mechanisms of the pollutants
creation. Here only a brief description of them and their environmental impact
will be made.
Carbon dioxide is the product of complete combustion of hydrocarbon fuels,
like kerosene in this case. Carbon in fuel combines with oxygen in the air to
produce CO 2 . It is the most significant gas that contributes to the greenhouse
effect. Carbon dioxide emissions from aircraft can be calculated from a knowledge
of the amount of fuel consumed during the flight. Fuel consumption does not
scale linearly with distance traveled due to the extra fuel burn required to lift
the plane up to cruising altitude, and the necessity to carry large quantities of
fuel for long distance flights. The highest fuel burn rate, thus the highest rate
emission of gases, occurs during the take-off and climb section, because of the
increased thrust needed to climb to cruise altitude and the heavier configuration
of the aircraft comparing to the other stages of the flight. The cruise is the most
fuel-efficient stage of the flight because the air is less dense and the aircraft is
flying at its most efficient operating speed, so the emissions are less than the first
stage of the flight. However, for the intercontinental flight that will be executed
in this case, the cruise time and distance is a lot longer than the corresponding
ones of the take-off and climb part. Thus the biggest part of the CO 2 emissions
are carried out at the cruise altitude. These emissions from an individual flight
depend mainly on the distance traveled, the weather conditions (head or tail
wind), the flight altitude, the cargo and passengers load.

99
Nitrogen oxides are produced when air passes through high temperature/high
pressure combustion, where nitrogen and oxygen present in the air combine to
form NO x . The nitrogen oxides are one of the most dangerous and toxic air
pollutant of aviation activity. They contribute to various environmental effects
and have significant impact throughout the whole flight, on both higher and lower
altitudes. The NO x are produced in higher engine power settings in contrast to
the CO and the HC, and its emissions are maximum near the stoichiometric
condition. In figure 8.4, the various air pollutants emissions dependence on the
equivalence ratio (Φ), can be observed. The Φ is defined as the fuel-to-air ratio
of the mixture over the relevant air-to-fuel ratio for stoichiometric combustion.
Therefore, for Φ <1 the mixture is characterized as lean, and for Φ >1 as rich.

Figure 8.4 – Air pollutants formation [48].

Water vapor is the other product of complete combustion as hydrogen in


the fuel combines with oxygen in the air to produce H2 O and is released by the
propulsion system into the atmosphere after the combustion process. For the
low layers of the atmosphere these emissions can be neglected, since they are
too low compared to the natural emissions. However, at high altitude, under
certain atmospheric conditions, because of the very low temperature of the air,
the vapor condenses into droplets to form contrails and cirrus clouds. So, the
impact of this pollutant should be taken into account only on the cruise part of
the flight. It is considered to have an effect to global warming and possibly to
precipitation inducement.
Carbon monoxide is formed due to the incomplete combustion of the carbon
in the fuel. It is a short-lived greenhouse gas (2 months) and its concentration is
extremely variable. In the atmosphere it is eventually oxidized to carbon dioxide.
It is very toxic and poisonous in the ground level, and in the higher layers of the
atmosphere can contribute to the increase of the tropospheric ozone, through its

100
photochemical oxidation. The CO is a result of the incomplete combustion of
the fuel and is mainly formed because of the absence of sufficient O2 during the
combustion. Thus, the emissions of CO are maximized at lean mixture operation
at low power settings, such as during idle. The presence of carbon monoxide can
be detected and measured with CO detectors, in order to prevent poisoning.
Hydrocarbons are emitted due to incomplete fuel combustion. They are
also referred to as volatile organic compounds (VOCs). Many VOCs are also
hazardous air pollutants. The formation of unburned hydrocarbons (UHC) are
a result of engines with low pressure increase in the compressor and relatively low
temperatures in the combustor. Hence, the largest amount of UHC are produced
during the lean mixture operation, i.e. during idle, like in the case of the CO.
Sulfur oxides are produced when small quantities of sulfur, present in essentially
all hydrocarbon fuels, combine with oxygen from the air during combustion.
SO x emissions are directly related to the sulfur content of the fuel, so it can
be estimated from the burned fuel and the relevant sulfur content of kerosene.
These oxides are corrosive and responsible for the formation of acid rain.
Soot particles that form as a result of incomplete combustion, and are small
enough to be inhaled, are referred to as particulate matters. They can be solid
or liquid. Particulate emissions were a problem on the earlier jet engines when
operating on high thrust settings. For normal operating conditions of the modern
engines, the production of smoke in every stage of the flight has been radically
reduced, so that the amount of particulate emissions can be considered negligible.

8.2.2 Environmental Concerns of Supersonic Flight


The major environmental impacts of aviation include primarily the climate
change and the ozone layer depletion. The two more prominent differences
between supersonic and subsonic cruise are the increased fuel consumption,
which leads to an increase in combustion products, and the higher cruise altitude
of the supersonic compared to the subsonic aircraft. The relevant influence of the
supersonic transport on the aforementioned environmental impacts is examined
in the following subsections.

Climate Change

The climate change, which is particularly being referred as enhanced greenhouse


effect or global warming, is one of the most important environmental concerns.
The uniqueness of the aircraft operation, compared to the other human activities
affecting the climate change, is the direct emission of air pollutants into the
higher levels of the atmosphere. The gases, which are being emitted from the
jet engines and contribute to the greenhouse effect, are the carbon dioxide,
the nitrogen oxides and the water vapor. The greenhouse effect increases the
temperature of the Earth by trapping heat in the atmosphere. This heat is
a result of the sun radiation absorption from the greenhouse gases, which are

101
hindering the heat absorbed from the ground to bounce back to space. The
global average surface temperature of the earth has increased by 0.6◦ C during
the 20th century, while a increment of at least 1.8◦ C is expected for the next
one [49]. Apart from the global temperature rise, this trapping of heat in the
atmosphere can also affect the weather conditions on the planet, such as the
appearance of heavier rainfall, floods, tornadoes, thunderstorms. Investigations
have also shown that the oceans are possible to expand, the ice on the poles to
melt and the surface of the sea level to rise covering parts of the existing land.
Although the contribution of the aviation emissions is only a small portion
of the total greenhouse gases emissions, the increment of the air traffic over the
last years and the forecasts for the upcoming ones, show that can develop in a
rather serious factor as regards the climate change. According to the European
Environmental Agency [51], the emissions of CO 2 have increased about 80 %
between 1990 and 2014, while the prediction is for a further grow of 45 % between
2014 and 2035. There are currently no requirements for the engine certification
respective with the greenhouse gases. However, the recent trend of the aviation
emissions increment, make their influence on the enhanced greenhouse effect
more substantial.
From the emitted gases the most problematic for the greenhouse effect is
considered to be the CO 2 , which among else has a long life cycle. The H2 O
can be considered a significant emission too, especially for flying vehicles in the
stratosphere, like the supersonic transport. Flying in such high altitudes, with
very low temperatures, the water vapor produced is converted into persistent
contrails, which evaporate very slowly. These contrails may be very long (dozens
of kilometers), forming the so-called cirrus cloud fields, which can potentially
have a strong influence in the climate change. The water is produced as a fixed
ratio to fuel which is consumed for complete combustion of kerosene, like in
the case of the carbon dioxide. In particular, the combustion of 1 kg kerosene
produces 3.16 kg CO 2 and 1.24 kg H2 O [48].
In tables 8.1 and 8.2, the emissions of CO 2 and H2 O of the supersonic
transport design are compared to the relevant emissions of a commercial subsonic
airliner. The results correspond to a transatlantic flight of 6050 km with a 100%
passenger load factor. The emissions of the SST design per km are lower, which
is a consequence of its smaller payload (just 15 passengers instead of the 416
of the Boeing 747) and thus size. However, the emissions of CO 2 and H2 O
per km per seat of the SST are about 5.85 times greater than the subsonic
airliner’s. That comes from the fact that the SST cruises at a significantly
smaller drag-to-lift ratio and with a larger thrust specific fuel consumption. In
particular, the specific fuel consumption of the RB211-524H installed in Boeing
747 during cruise is 16.14 g/KN/s [52], which is a value considerably lower.
From the obtained results, it can be inferred that the influence of small
supersonic aircraft flight would not be so environmentally problematic concerning
the greenhouse effect. However, an excessive growth of the supersonic transportation,
especially in the case of large supersonic transports replacing part of the current

102
subsonic flee, could essentially affect the total aviation emissions. Moreover,
the phenomenon of the persistent contrails and cirrus clouds formation at the
stratosphere and its consequences on the environment would need to be further
investigated.

Aircraft CO 2 (kg/km) CO 2 (kg/km/seat)


Boeing 747 33.6960 0.0810 [53]
SST 7.1191 0.4746

Table 8.1 – Carbon dioxide emissions of subsonic airliner and SST design
(6050 km distance flown).

Aircraft H2 O (kg/km) H2 O (kg/km/seat)


Boeing 747 13.2225 0.0318
SST 2.7936 0.1862

Table 8.2 – Water vapor emissions of subsonic airliner and SST design
(6050 km distance flown).

Ozone Layer Depletion

The influence of the supersonic flight on the ozone layer constitutes the
biggest environmental concern. Nearly 90% of the ozone exists in the stratosphere,
forming the ozone layer. Since the SST cruises at high altitudes from 47,700 to
about 57,300 f t, it is obvious that the aircraft will directly emit the produced
NO x in the stratosphere and thus in the ozone layer. The subsonic airliners
executing long haul flights at a cruise altitude of around 35,000 f t are emitting
NO x in the low level of the stratosphere as well. The NO x in the stratosphere
are participating in a catalytic chemical reaction, which leads to the ozone
destruction. The ozone layer breakdown could allow the ultraviolet B radiation
from the sun to pass through this ozone shield and reach the Earth, which
could cause among else skin cancer and cataract in humans, but could harm the
animals too. However, both the subsonic and supersonic aircraft emit NO x in
the stratosphere, the impact of the SST flying at higher flight altitudes is more
significant due to the increased ozone concentration. It is shown in figure 8.5
that the highest ozone concentrations are observed between 60,000 and 80,000 f t,
which comprise the typical flight altitudes of supersonic aircraft at speeds equal
to Mach 2 and higher. Thus, the excessive flight of large supersonic transports
with speeds greater than Mach 2, cruising at altitudes near the maximum ozone
concentration while burning big amount of fuel, could potentially be the most
problematic SST concept regarding the ozone depletion environmental impact.

103
Figure 8.5 – Atmosphere ozone concentration and temperature till the
altitude of 100,000 ft [54].

The NO x are produced during the combustion of the kerosene. In reference


[55], it is shown a simple correlation between the NO x emission index (EINO x )
and the combustor inlet total temperature (Ttc ). The equation (8.1) is an
empirical relation for the prediction of the NO x emissions based on the so-called
Lipfert correlation, stated previously [48], where δ is defined in equation (2.16).

EIN Ox = 10(1+0.0032(Ttc −581.25)) δ (8.1)

The above equation demonstrates that for high combustor inlet temperatures
and thus high engine pressure ratios, the NO x emissions of the engine are
significantly increased. The pressure ratio, determining the total temperature
at the compressor exit, influences the actual primary zone temperature in the
combustion chamber [48], and thus the NO x production as well. In figure 8.6, the
NO x emissions index variation is presented with respect to the overall pressure
ratio for a subsonic airliner and the SST design. The example of the subsonic
aircraft that was used is the Boeing 747-400, and the relevant cruise conditions
are 0.85 Mach at a flight altitude of 11 km. For the SST the cruise speed is
1.7 Mach at an average flight altitude of 16 km. The upper limit of the overall
pressure ratio, including the fan, of the RB211-524H engine of the Boeing 747-400
is set to 33 [52], while for the EJ200 is set to 22, so that the total temperature
at the compressor exit does not exceed 900 K, which is a practical limit of the
compressor materials and its cooling requirements [50]. For the RB211-524H a

104
diffuser isentropic efficiency of 0.97 has been assumed, while the intake isentropic
efficiency for the SST would be 0.95 as estimated in Chapter 2. For both engines,
the compression isentropic efficiency has been set to the typical value of 0.85.

Figure 8.6 – NO x emissions index for SST design (red) and for a typical
subsonic long haul airliner (blue).

From the calculated values illustrated on figure 8.6, the maximum EINO x of
the subsonic aircraft is about 16.2 g/kg fuel at its maximum compression ratio of
33, which is a value that agrees with the relevant ones stated in reference [56] for
long range subsonic transports. On the contrary, the pressure ratio effect on the
EINO x of the SST is much more prominent with an emissions index of 36.58 g/kg
fuel at the maximum set pressure ratio of 22. The increased fuel consumption
of the SST is also a parameter that contributes to even higher NO x emissions,
creating great concerns about the environmental viability of an excessive turn
in supersonic transportation in the near future.
In order to keep NO x emissions within acceptable limits, an emissions index
as low as 5 during cruise, which corresponds to about a 80% reduction of the
above calculated values, would be necessary, which has been the goal set to
be investigated during the NASA’s High Speed Research Program as well [54].
In order to achieve such low NO x emissions, new engine concepts have to be
developed emphasizing on this reduction, while maintaining the other engine
requirements of low thrust specific fuel consumption, high thrust-to-weight ratio
and reliability. Moreover, the economic viability of the undertaking to build a
new engine capable to be incorporated at supersonic aircraft has to be examined
as well.
In order to reduce the NO x emissions, the equivalence ratio Φ has to be
controlled, so that the engine operates in the low emissions region of Φ, according
to figure 8.4. Moreover, it is important to reduce the time of the gases remaining
in high temperatures. There are different types of combustor concepts, referred

105
to as dry low NO x combustors, which focus on more efficient, as regards the
relevant emissions, combustion process. Three types of them are the lean premixed
pre vaporized combustor, the staged combustor type and the rich burn quick
quench lean combustor. The concepts behind the aforementioned combustion
chambers operation, in addition to their pros and cons, are more thoroughly
described in reference [48].
Finally, the water emissions in the stratosphere is believed to have some
influence on the ozone destruction too, since it can affect the affect the composition,
growth and aerosol reactions and provide a source of HO x radicals that enhance
ozone loss [54]. However, this is an effect that has to be further investigated, so
that the consequences of the water vapor emissions in the ozone layer become
more certain.

106
9. Discussion - Conclusions

Since the Concorde retirement, there is no supersonic transport aircraft in


operation, but the possibility of a new SST development is becoming a reality
once again. Efforts are being made in both the development of a low-boom
design and a SST able to fly supersonically just overseas.
The design proposed in this project is a supersonic transport aircraft able to
carry 15 passengers and 4 crew members, having a maximum payload of 1900
kg. The aircraft is being able to fly at supersonic speeds just overseas, since it
has not be designed for reduced aerodynamic noise generation. The cruise flight
conditions are 1.7 Mach at an initial altitude of 14.55 km. The aircraft’s take off
mass is estimated as 24499 kg, while the wing area is about 63.75 m2 . The total
dry thrust at SL is 120 kN , provided by two low-bypass ratio turbofan EJ200
engines. In order to surpass the sound barrier, the aircraft has to incorporate
a carefully area-ruled design, with a parasite drag coefficient at the given cruise
conditions that should not exceed 0.021.
The aircraft, in order to exhibit satisfactory static longitudinal stability in
both subsonic and supersonic speeds, keeping the static margin to 0.1 for every
flight condition, uses fuel shifting for the aft movement of the aircraft center
of gravity during supersonic flight. The design incorporates a whole moving
horizontal tail is used instead of an elevator. During the supersonic cruise, the
aircraft is estimated to fly at about 3 deg angle of attack in trimmed condition.
Hence, a wing incidence of 3 deg would be chosen to minimize the fuselage drag
during cruise.
During the mission a total horizontal range of 6053 km can be covered, in
addition to a maximum loiter of 24 min. This range is rather smaller than
the desired requirement of 7200 km. However, it could be increased through
camber and wing twist optimization, so that the estimated drag-to-lift ratio of
5.764 reaches a more efficient value. The estimated time of a transatlantic flight
between London and New York is estimated at about 4 hrs.
The aircraft exhibits adequate flying and handling qualities in subsonic speeds,
like the dynamic stability modes evaluation has shown, from the obtained flight
test measurements. However, these qualities was not possible to be assessed for
the supersonic speeds as well, since the flight test failed to give reasonable results
using the simulation model created.
The two most important environmental concerns about the supersonic flight

107
are the noise generated by the shock waves, the so-called sonic boom, and the air
pollutants emissions. A discussion made about low-boom design shows that the
sonic boom is quite hard to be reduced in larger and heavier aircraft. Moreover,
the low-boom design and the aircraft performance optimization contradict each
other, so that this trade-off between them questions the economic viability of the
reduced sonic boom loudness designs. As regards the air pollution, the excessive
turn in supersonic flight could lead to two significant environmental impacts,
named the climate change and the ozone layer depletion. It has been shown that
the emissions of greenhouse gases, like the carbon dioxide, are bigger compared
to subsonic airliners, as a result of the higher fuel consumption. The higher
nitrogen oxide emissions, resulting from the supersonic flight, at the stratosphere,
especially in altitudes where the ozone concentration is maximum, constitute an
even more significant problem to be tackled.
During the conceptual design and the flight test, the aeroelastic effects have
been disregarded. Moreover, the aerodynamic coefficients referring to transonic
flights have been just interpolated following charts and theoretical guidelines
given from Raymer in reference [9], due to lack of valid empirical relations.
CFD simulation would be a more accurate way for obtaining the transonic
aerodynamic coefficients. Finally, CFD simulations could be used for the evaluation
and validation of the used subsonic and supersonic wing aerodynamic coefficients
as well.

108
109
Appendices

A. Airfoils Coordinates

NACA 64-006 coordinates [15].


110
NACA 64-009 coordinates [15].

111
B. Airfoils Aerodynamic Characteristics

NACA 64-006 lift and pitching moment coefficient [15].

112
NACA 64-006 drag polar [15].

113
NACA 64-009 lift and pitching moment coefficient [15].

114
NACA 64-009 drag polar [15].

115
Bibliography

[1] National Aeronautics and Space Administration, NASA Begins


Work to Build a Quieter Supersonic Passenger Jet, (2017, May
30), Retrieved from https://www.nasa.gov/press-release/
nasa-begins-work-to-build-a-quieter-supersonic-passenger-jet.
[2] Boom Technology, Official webpage, (2017, May 30), Retrieved from https:
//boomsupersonic.com/.
[3] B. Liebhardt, K. Lütjens, An Analysis of the Market Environment
for Supersonic Business Jets, German Aerospace Center (DLR) – Air
Transportation Systems, Hamburg, Germany, 2011.
[4] B. Liebhardt, K. Luetjens, V. Gollnick, Estimation of the Market Potential
for Supersonic Airliners via Analysis of the Global Premium Ticket Market,
American Institute of Aeronautics and Astronautics, German Aerospace
Center (DLR) – Air Transportation Systems, Hamburg, Germany, 2011.
[5] C. Carreras, M. Daouk, T. Downen, L. Jamonet, S. Lederle, D. Sharman,
R. Wertenberg, The case for a practical small supersonic transport,
International Council of the Aeronatuical Sciences, Massachusetts Institute
of Technology, Cambridge, USA, 2002.
[6] Airbus, Global Market Forecast Mapping Demand 2016/2035
[7] Boeing, Current Market Outlook 2015-2034
[8] Bombardier, Bombardier Business Aircraft Market Forecast 2016-2025
[9] Daniel P. Raymer, Aircraft Design : A Conceptual Approach, American
Institute of Aeronautics and Astronautics, AIAA, 5th edition, Virginia, USA,
2012.
[10] Jack D. Mattingly, William H. Heiser, David T. Pratt, Aircraft Engine
Design, American Institute of Aeronautics and Astronautics, AIAA, 2nd
edition, Virginia, USA, 2002.
[11] Thomas C. Corke, Design of Aircraft, Pearson Education Inc., New Jersey,
USA, 2003.
[12] E. Torenbeek, H. Wittenberg, Flight Physics. Essentials of Aeronautical
Disciplines and Technology, with Historical Notes, Springer, 2009.

116
[13] John D. Anderson, Jr, Aircraft Performance and Design, WCB/
McGraw-Hill, New York, USA, 1999.

[14] John D. Anderson, Jr, Funtamentals of Aerodynamics, WCB/ McGraw-Hill,


5th edition, New York, USA, 2011.

[15] Ira H. Abbot, Albert, E. von Doenhoff, Theory of Wing Sections, Dover
Publications Inc., New York, USA, 1959.

[16] Dennis Howe, Aircraft Conceptual Design Synthesis, Professional


Engineering Publishing, London, UK, 2000.

[17] Norman S. Currey, Aircraft Landing Gear Design : Principles and Practices,
American Institute of Aeronautics and Astronautics, AIAA, Washington,
DC, USA, 1988.

[18] National Advisory Committee for Aeronautics, Report 1135. Equations,


tables and charts for compressible flow, Washington, DC, USA, 1948.

[19] John D. Anderson, Jr, Modern Compressible Flow, with Historical


Perspective, WCB/ McGraw-Hill, 3rd edition, New York, USA, 2003.

[20] John D. Anderson, Jr, Introduction to Flight, WCB/ McGraw-Hil


International Edition, 8th edition, New York, USA, 2016.

[21] Snorri Gudmundsson, General aviation Aircraft Design : Applied Methods


and Procedures, Elsevier Inc., Waltham, USA, 2014.

[22] Klaus Hünecke, Jet Engines, Fundamentals of Theory, Design and


Operation, Motorbooks International Publishers & Wholesalers, Osceola,
USA, 1997.

[23] Pasquale M. Sforza, Theory of Aerospace Propulsion, Elsevier Inc.,


Waltham, USA, 2012.

[24] MTU Aero Engines, EJ 200 Turbofan Engine, Product Leaflet, (2017,
March 8), Retrieved from http://www.mtu.de/fileadmin/EN/2_
Engines/2_Military_Aircraft_Engines/1_Fighter_Aircraft/EJ200/
ProductLeaflet_EJ200.pdf.

[25] Mohammad H. Sadraey, Aircraft Design : A Systems Engineering Approach,


John Wiley & Sons Ltd, West Sussex, UK, 2013.

[26] Mark Drela, XFOIL : An Analysis and Design System for Low Reynolds
Number Airfoils, MIT Dept. Of Aeronautics and Astronautics, Cambridge,
Massachusetts, USA.

[27] John G. Lowry, Edward C. Polhamus, Technical Note 3911. A Method for
Predicting Lift Increments due to Flap Deflection at Low Angles of Attack in
Incompressible Flow, National Advisory Committee for Aeronautics, NACA,
Washington, USA, 1957.

117
[28] Bernard Etkin, Lloyd Duff Reid, Dynamics of Flight. Stability and Control,
John Wiley & Sons, Inc., 3rd edition, Toronto, Canada, 1996.

[29] Steven A. Brandt, Introduction to Aeronautics : A Design Perspective,


American Institute of Aeronautics and Astronautics, AIAA, 3rd edition,
Reston, Virginia, USA, 2015.

[30] Leland M. Nikolai, Grant E. Carichner, Funtamentals of Aircraft


and Airship Design. Volume I, American Institute of Aeronautics and
Astronautics, AIAA, Reston, Virginia, USA, 2010.

[31] Richard T. Whitcomb, John R. Sevier, Jr., Technical Report R-72. A


Supersonic Area Rule and an Application to the Design of a Wing-Body
Combination with High Lift-Drag Ratios, NASA, Langley Research Center,
USA, 1960.

[32] Roelof Vos, Saeed Farokhi, Introduction to Transonic Aerodynamic,


Springer, Dodrecht, The Netherlands, 2015.

[33] Mohamad Sadraey, Spoiler design, Daniel Webster College, (2017,


March 26), Retrieved from http://faculty.dwc.edu/sadraey/Spoiler\
%20design.pdf.

[34] Courtland D. Perkins, Robert E. Hage, Airplane Performance, Stability and


Control, John Wiley & Sons, New York, USA, 1949.

[35] Maido Saarlas, Aircraft Performance, John Wiley & Sons, New Jersey, USA,
2007.

[36] Christopher Orlebar, The Concorde Story, Osprey Publishing, Oxford, UK,
2017.

[37] Merlin Flight Simulator Group, The Merlin MP521 Engineering Flight
Simulator, Merlin Products Ltd. and Merlin Flight Simulation Technologies,
(2017, May 16), Retrieved from http://www.merlinsim.com/mp521.htm.

[38] Michael V. Cook, Flight Dynamic Principles. A Linear Systems Approach


to Aircraft Stability and Control, Elsevier Ltd., 3rd edition, Waltham, USA,
2013.

[39] Military Specification MIL-F-8785C, Flying Qualities of Piloted Airplanes.,


Nov. 1980.

[40] Christine M. Darden, Progress in Sonic-Boom Understanding : Lessons


Learned and next Steps, NASA Langley Research Center, Hampton, Virginia,
USA, 1994.

[41] George T. Haglund, Low Sonic Boom Activities at Boeing, Boeing


Commercial Airplane Group, Seattle, WA, USA, 1994.

118
[42] Samon H. Cheung, Thomas A. Edwards, Supersonic Airplane Design
Optimization Method for Aerodynamic Performance and Low Sonic Boom,
NASA Ames Research Center, Moffett Field, USA, 1992.
[43] Sriram Rallabhandi, Sonic Boom Minimization through Vehicle Shape
Optimization and Probabilistic Acoustic Propagation. PhD dissertation ,
Georgia Institute of Technology, USA, 2005.
[44] Robert J. Mack, A Supersonic Business-Jet Concept Designed for Low Sonic
Boom, NASA Langley Research Center, Hampton, Virginia, USA, 2004.
[45] Percy J. Bobbitt, Application of Computational Fluid Dynamics and
Laminar Flow Technology for Improved Performance and Sonic Boom
Reduction, Eagle Engineering Inc., Hampton Division, USA, 1992.
[46] Kevin P. Shepherd, A Loudness Calculation Procedure Applied to Shaped
Sonic Booms, NASA Langley Research Center, Hampton, Virginia, USA,
1991.
[47] Harry W. Carlson, Raymond L. Barger, Robert J. Mack, Application of
Sonic Boom Minimization Concepts in Supersonic Transport Design, NASA
Langley Research Center, Hampton, Virginia, USA, 1973.
[48] G.J.J. Ruijgrok, D.M. van Paasen, Elements of Aircraft Pollution, VSSD,
Delft, The Netherlands, 2007.
[49] Christian N. Jardine, Calculating the Environmental Impact of Aviation
Emissions, Environmental Change Institute, Oxford University, UK, 2005.
[50] Saeed Farokhi, Aircraft Propulsion, John Wiley & Sons, Hoboken, USA,
2009.
[51] European Environment Agency, European Aviation Environmental
Report 2016, EASA, (2017, May 01), Retrieved from
https://ec.europa.eu/transport/sites/transport/files/
european-aviation-environmental-report-2016-72dpi.pdf.
[52] Lloyd R. Jenkinson, Paul Simpkin, Darren Rhodes, Civil Jet Aircraft
Design, American Institute of Aeronautics and Astronautics, AIAA, Reston,
Virginia, USA, 1999.
[53] Christian N. Jardine, Calculating the Carbon Dioxide Emissions of Flights,
Environmental Change Institute, Oxford University, UK, 2009.
[54] National Research Council, Commercial Supersonic Technology. The Way
Ahead, National Academy Press, Washington, DC, USA, 2001.
[55] F.W. Lipfert, Correlation of Gas Turbine Emissions Data, ASME, San
Francisco, California, USA, 1972.
[56] A. Wahner, M.A. Geller Scientific Assessment of Ozone Depletion:
1994. Subsonic and Supersonic Aircraft Emissions, National Oceanic and
Atmospheric Administration, USA, 1994.

119

You might also like