You are on page 1of 221

Humanitarian Emergency Relief Aircraft

Final Report of the International Design Synthesis Exercise 2005

Belfast / Delft, May 2005

Del-Viso Hopkins, P.S. Fick, P.W.


Gillen, M.T. Hiele, R.D.
Kennedy, S. De Hoon, M.D.J.
Matthews, K. Huertas Garcia, I.
Scott, K.M.L. De Jong, S.
Toal, D.J.J. Karatas, F.

Delft University of Technology


Humanitarian Emergency Relief Aircraft

Final report of the International Design Synthesis Exercise 2005

Principal Supervisors
Dr. R. Cooper Queens University Belfast
Ir. T.J. van Baten Delft University of Technology

Tutors
Dr. E. Benard Queens University Belfast
Dr. ir. O.K. Bergsma Delft University of Technology
Dr. R. Curran Queens University Belfast
Ir. M.M. Heiligers Delft University of Technology
Dr. M. Price Queens University Belfast

Copyright © 2005 by P.S. Del-Viso Hopkins, P.W. Fick, M.T. Gillen, R.D. Hiele, M.D.J. de Hoon, I.
Huertas Garca, S. de Jong, F. Karatas, S. Kennedy, K. Matthews, K.M.L. Scott, D.J.J. Toal, International
Design Synthesis Exercise 2005. All rights reserved.

No part of this report may be reproduced in any form by print, photo print, microfilm or any other means
without the written permission of the copyright holders.
Final Report of the International Design Synthesis Exercise 2005

Preface

This document has been written within the scope of the International Design Synthesis Exercise 2005.
The exercise is conducted by an international team, composed of six students from the Delft University of
Technology (NL) and six students from the Queen’s University Belfast (UK). The aim of the exercise was
to design a humanitarian relief and emergency aid cargo aircraft, capable of transporting cargo and
evacuees to and from almost every place in the world under the most severe conditions.

This report will present the final preliminary design of the aircraft called HERA (Humanitarian
Emergency Relief Aircraft). It builds further onto the Mid Term Report [8], which was submitted halfway
the project.

The design, and therefore also this report, would not have been realised without the continuous effort and
support of our supervisors, tutors and other experienced staff. First of all, the authors would like to thank
ir. G.N. Saunders-Smits for coordinating this international project. Without her tremendous effort, this
project would not have been organised. We would also like to thank our tutors – dr. E. Benard, dr. ir.
O.K. Bergsma, dr. R. Curran, ir. M.M. Heiligers and dr. M. Price – for their readiness to give technical
advice and feedback. In April the team members in Delft got the opportunity to see a C130J-30 in
maintenance at the Royal Dutch Airforce base in Eindhoven. We would like to thank professor BA.C.
Droste for arranging this visit as well as the Public Relations and the Maintenance Department from the
Eindhoven Airforce base for their warm welcome. Last, but certainly not least, the authors would like to
express their special gratitude to Dr. R. Cooper and Ir. T.J. van Baten for their guidance and helpfulness
throughout this stage of the project. We would like to thank them for the tremendous amount of time and
effort they spent into guiding, educating and helping us. They have stimulated us in all our activities and
made a great contribution to the final result.

Delft, May 4, 2005

David Toal Meta de Hoon


Ferhat Karatas Patrick Del-Viso Hopkins
Irene Huertas Garcia Peter Fick
Karen Scott Roeland Hiele
Kelly Matthews Sander de Jong
Martin Thomas Gillen Stefan Kennedy

Preface i
Final Report of the International Design Synthesis Exercise 2005

Abstract

Air transport has proved to be indispensable in humanitarian operations. These contain the air transport
of emergency aid, such as food supplies, vehicles, evacuees and medical equipment. However, aid
organisations around the world rely on the support from governments. Reductions in these budgets as
well as the increasing need to operate, while neutral, in emergency situations, resulted in the need for a
robust, low-cost, reliable cargo aircraft, independently operated by aid organisations for delivering
humanitarian relief and emergency aid cargo transport. This need was formulated in the following Mission
Need Statement:

“To provide low-cost, flexible and reliable air transport of aid, evacuees and medical equipment under difficult
circumstances whilst remaining impartial.”

Out of this statement the HERA project was born and this statement was used as a basis for every single
design activity. HERA stands for Humanitarian Emergency Relief Aircraft. Together with customers’
validation, requirements of the HERA were set up, which resulted in the preliminary concept. This was
presented later on to the customer. However, two requirements, which could not be met (a maximum
speed of 700 km/hr and 800 m for take off and landing on unprepared runways), were therefore lowered
after a review with the customer.

A more detailed design was then generated. At present, the HERA, almost 39 meters long and 49 meters
wide, is a high wing airplane with a high aspect ratio and large wing control surfaces. Its four six-bladed
Rolls Royce Ae1107c turbo propeller engines make it possible for the HERA to carry 35 tons of payload
in a cabin volume of 400 m3 to an objective over 5000 km away. The aircraft is capable of transporting 9
pallets, 70 stretchers or 130 evacuees and bring help to or evacuate people from disaster areas. The cargo
area is equipped such that the aircraft does not depend on ground facilities in case these are lacking. The
landing gear has been designed in such a way that the HERA is able to take off with its maximum take off
weight and land on unprepared airstrips, less than 1000 meters long.

The final configuration of the HERA complies with five of the customer requirements. These are the
cabin volume, the payload weight, the acquisition costs of $30 million and the maximum speed of 650
km/hr. It even surpasses the requirement for the maximum range with full payload, for which 4000 km
was required and almost 5500 km is achieved.

Abstract iii
Final Report of the International Design Synthesis Exercise 2005

However there is still one requirement set by the customers the HERA fails to meet. HERA’s first flight,
which was required to be in 2010, will be delayed with two years, caused by extensive testing and
certification of certain parts of the aircraft will, so the first flight is set to be in 2012.

This last drawback perhaps will underline the quality of the HERA to serve as a low-cost, flexible and
reliable air transport of aid, evacuees and medical equipment under difficult circumstances. Especially the
low costs are essential, so impartial, non-governmental organisations are able to purchase the HERA. It
shows that the HERA is a feasible concept and it is ready for full-scale development.

iv Abstract
Final Report of the International Design Synthesis Exercise 2005

Table of Contents

Preface................................................................................................................................................i

Abstract ........................................................................................................................................... iii

List of Symbols ...............................................................................................................................xiii

List of Figures ................................................................................................................................xxi

List of Tables................................................................................................................................xxiii

1 Introduction ............................................................................................................................... 1

2 Mission Analysis and Project Planning......................................................................................3

2.1 Mission specification...................................................................................................................................3


2.2 Top-level requirements...............................................................................................................................4
2.3 Functional Analysis .....................................................................................................................................5
2.3.1 Functional flow diagram ..............................................................................................................5
2.3.2 Functional breakdown .................................................................................................................6
2.3.3 Operations concept and flow diagram ......................................................................................6
2.4 Hardware-software block diagrams ..........................................................................................................7
2.5 Technical risk assessment...........................................................................................................................7
2.5.1 Risk identification .........................................................................................................................8
2.5.2 Risk analysis: the risk map ...........................................................................................................8
2.5.3 Risk handling .................................................................................................................................9
2.6 Project design and development logic......................................................................................................9

3 General Layout ..........................................................................................................................11

3.1 External configuration ..............................................................................................................................11


3.2 Cabin layout................................................................................................................................................12
3.2.1 Payload..........................................................................................................................................12
3.2.2 Non-pressurised ..........................................................................................................................13
3.2.3 Autonomous on- and offloading..............................................................................................14
3.3 Cockpit interior ..........................................................................................................................................14
3.4 Flight Control System ...............................................................................................................................14
3.5 Aircraft Specifications...............................................................................................................................15

Table of Contents v
Final Report of the International Design Synthesis Exercise 2005

4 Aerodynamics........................................................................................................................... 17

4.1 Wing.............................................................................................................................................................17
4.1.1 Wing geometry ............................................................................................................................17
4.1.2 Airfoil analysis .............................................................................................................................18
4.1.3 Wing lift........................................................................................................................................21
4.1.4 Wing twist ....................................................................................................................................22
4.1.5 Spanwise lift distribution ...........................................................................................................24
4.1.6 High-lift devices ..........................................................................................................................25
4.2 Aircraft drag breakdown...........................................................................................................................28
4.2.1 Wing ..............................................................................................................................................28
4.2.2 Fuselage ........................................................................................................................................28
4.2.3 Tailplane .......................................................................................................................................29
4.2.4 Engines .........................................................................................................................................29
4.2.5 Undercarriage ..............................................................................................................................29
4.2.6 Other factors................................................................................................................................29
4.2.7 Total drag .....................................................................................................................................30
4.2.8 Upsweep effects ..........................................................................................................................30
4.3 Aircraft aerodynamic centre.....................................................................................................................31
4.2.9 Aerodynamic centre of the wing ..............................................................................................31
4.2.10 Aerodynamic centre of wing and body ...................................................................................32
4.2.11 Pitching moment.........................................................................................................................33
4.4 Aircraft lift ..................................................................................................................................................34

5 Power Plant .............................................................................................................................. 39

5.1 Engine..........................................................................................................................................................39
5.2 Propeller ......................................................................................................................................................40
5.2.1 Assumptions Made in Preliminary Design .............................................................................40
5.2.2 Propeller Design Method ..........................................................................................................41
5.2.3 Results...........................................................................................................................................42
5.3 Engine locations.........................................................................................................................................45
5.4 Noise suppression .....................................................................................................................................45
5.5 Emissions....................................................................................................................................................46

6 Performance Analysis ............................................................................................................... 47

6.1 Constraints analysis ...................................................................................................................................47


6.1.1 Stall speed.....................................................................................................................................47
6.1.2 Takeoff Distance.........................................................................................................................48

vi Table of Contents
Final Report of the International Design Synthesis Exercise 2005

6.1.3 Cruise and maximum speed ......................................................................................................49


6.1.4 Climb.............................................................................................................................................51
6.1.5 Design Point ................................................................................................................................52
6.2 Climb performance....................................................................................................................................53
6.3 Load factor envelope ................................................................................................................................54
6.3.1 Construction of Manoeuvre Envelope....................................................................................54
6.3.2 Construction of Gust Envelope ...............................................................................................55
6.4 Flight envelope...........................................................................................................................................56
6.5 Payload-range diagram..............................................................................................................................58

7 Airframe Structural Design ...................................................................................................... 61

7.1 Materials selection .....................................................................................................................................61


7.2 Wing.............................................................................................................................................................62
7.2.1 Spanwise wing loading: shear and bending.............................................................................62
7.2.2 Spanwise wing loading: torsion.................................................................................................65
7.2.3 Wing structural analysis: assumptions and method...............................................................66
7.2.4 Axial stress ...................................................................................................................................67
7.2.5 Bending Shear Stress ..................................................................................................................68
7.2.6 Pure Torsion ................................................................................................................................69
7.2.7 Results...........................................................................................................................................70
7.3 Fuselage .......................................................................................................................................................73
7.3.1 Torsion of the fuselage ..............................................................................................................73
7.3.2 Bending of the fuselage..............................................................................................................78
7.3.3 Shear loading of the fuselage.....................................................................................................80
7.4 Empennage.................................................................................................................................................82
7.4.1 Horizontal tail: assumptions......................................................................................................82
7.4.2 Horizontal tail: results and discussion .....................................................................................82
7.4.3 Vertical tail: assumptions ...........................................................................................................84
7.4.4 Vertical tail: results and discussion...........................................................................................84
7.5 Cargo ramp .................................................................................................................................................86
7.5.1 Layout of the ramp .....................................................................................................................86
7.5.2 Structural analysis of the ramp..................................................................................................88

8 Undercarriage........................................................................................................................... 93

8.1 General layout ............................................................................................................................................93


8.2 Location on the HERA ............................................................................................................................93
8.2.1 Steering .........................................................................................................................................93
8.2.2 Longitudinal Tip-over ................................................................................................................94

Table of Contents vii


Final Report of the International Design Synthesis Exercise 2005

8.2.3 Lateral Tip-over...........................................................................................................................95


8.2.4 Ground clearance........................................................................................................................95
8.3 Tire-selection ..............................................................................................................................................95
8.4 Shock absorber...........................................................................................................................................96
8.5 Brakes ..........................................................................................................................................................97
8.6 Landing gear design...................................................................................................................................98
8.7 Moveable shelter ........................................................................................................................................98
8.8 Evaluation ...................................................................................................................................................99
8.9 Landing gear design...................................................................................................................................99
8.10 Shelter ..........................................................................................................................................................99
8.11 Landing gear-fuselage connection.........................................................................................................100
8.12 Weight estimation....................................................................................................................................100

9 Weight and Balance ................................................................................................................101

9.1 Class one component mass estimation ................................................................................................101


9.2 Detailed mass estimate............................................................................................................................102
9.2.1 Wing group mass estimate.......................................................................................................102
9.2.2 Fuselage mass estimate.............................................................................................................103
9.2.3 Tail mass estimate .....................................................................................................................103
9.2.4 Operational mass items and fixed equipment ......................................................................103
9.2.5 Total propulsion mass, landing gear mass, payload and fuel mass ...................................104
9.3 Centre of gravity analysis........................................................................................................................104
9.3.1 Component centre of gravity ..................................................................................................105
9.3.2 Loading Conditions and Corresponding Centre of Gravity Location .............................106
9.3.3 Centre of gravity range criteria and excursion......................................................................106
9.4 Mass moments of inertia ........................................................................................................................107
9.5 Conclusions ..............................................................................................................................................108

10 Control and Stability................................................................................................................109

10.1 Initial sizing of control surfaces ............................................................................................................109


10.2 Longitudinal stability and control derivatives .....................................................................................111
10.2.1 The Criteria for Longitudinal Static Stability........................................................................112
10.2.2 Tail Plane Sizing ........................................................................................................................113
10.2.3 Test for Longitudinal Static Stability .....................................................................................117
10.2.4 Longitudinal Dynamic Stability ..............................................................................................118
10.3 Lateral stability and control derivatives................................................................................................122
10.4 Initial sizing of the fin and rudder ........................................................................................................124
10.4.1 Heading after Engine Failure ..................................................................................................124

viii Table of Contents


Final Report of the International Design Synthesis Exercise 2005

10.4.2 Rudder angle to trim after engine failure ..............................................................................124


10.4.3 Inability of Rudder to Cause Fin Stall ...................................................................................125
10.4.4 Heading Hold in a Crosswind.................................................................................................125
10.4.5 Change of Heading Against a Failed Engine........................................................................125
10.4.6 Determination of Design Values............................................................................................125
10.4.7 Dihedral Angle ..........................................................................................................................127
10.4.8 Aileron Sizing ............................................................................................................................127
10.4.9 Crosswind Case .........................................................................................................................127
10.4.10 Roll performance ......................................................................................................................128
10.4.11 Dynamic Stability ......................................................................................................................129
10.5 Main results of the lateral stability analysis..........................................................................................130

11 Manufacturing, Assembly and Certification ...........................................................................131

11.1 Manufacturing ..........................................................................................................................................131


11.2 Joining techniques ...................................................................................................................................132
11.3 Aircraft assembly .....................................................................................................................................133
11.4 Verification/certification approach ......................................................................................................133

12 Cost Analysis ...........................................................................................................................135

12.1 Aircraft Estimated Price .........................................................................................................................135


12.1.1 Estimation of Research, Development, Test and Evaluation Cost (RDTE) ..................135
12.1.2 Method for Estimating Manufacturing and Acquisition Cost...........................................139
12.2 Reductions in Cost ..................................................................................................................................141
12.3 Return on Interest and Operational Cost ............................................................................................142
12.4 Operational Cost......................................................................................................................................143
12.4.1 Direct Operating Costs (DOC). .............................................................................................144
12.4.2 Indirect Operating Costs (IOC) .............................................................................................145
12.4.3 Total Operating Cost................................................................................................................146

13 Sustainable Development Strategy..........................................................................................147

13.1 Production ................................................................................................................................................147


13.2 Operational use ........................................................................................................................................148
13.3 Recycle plan ..............................................................................................................................................148

14 Conclusions and Recommendations.......................................................................................149

14.1 Conclusions ..............................................................................................................................................149


14.2 Recommendations ...................................................................................................................................150

Table of Contents ix
Final Report of the International Design Synthesis Exercise 2005

References ......................................................................................................................................151

Appendix 1 Mission Requirements ...........................................................................................155

Appendix 2 Comparison of Similar Aircraft ..............................................................................157

Appendix 3 Functional Analysis................................................................................................159

3.1 Functional flow diagram.........................................................................................................................159


3.2 Functional Breakdown............................................................................................................................160
3.3 Operational Flow Diagram ....................................................................................................................161

Appendix 4 Project Gantt Chart ................................................................................................162

Appendix 5 Three-View Drawing of the HERA .......................................................................165

Appendix 6 Payload Configurations..........................................................................................167

6.1 Payload configuration top view.............................................................................................................167


6.2 Payload configuration front view..........................................................................................................168

Appendix 7 CD-ROM................................................................................................................169

Appendix 8 Blade Element Theory ...........................................................................................173

Appendix 9 Rate of Climb Curves .............................................................................................175

Appendix 10 Landing Gear Design.............................................................................................177

10.1 Fixed landing gear design .......................................................................................................................177


10.2 Semi retractable landing gear design.....................................................................................................177
10.3 Shelter design............................................................................................................................................178
10.4 Connection strut-fuselage.......................................................................................................................178

Appendix 11 Weight and Balance Calculation Table .................................................................179

Appendix 12 Centre of Gravity Excursion Diagram ...................................................................181

Appendix 13 Aerodynamic Values For Horizontal Tail Sizing...................................................183

Appendix 14 Horizontal Tail Geometry......................................................................................185

Appendix 15 MATLAB Program Graphical User Interfaces ......................................................187

x Table of Contents
Final Report of the International Design Synthesis Exercise 2005

15.1 Graphical user interface P1....................................................................................................................187


15.2 MATLAB Graphical User Interface P2...............................................................................................188

Appendix 16 Cost Analysis..........................................................................................................189

Appendix 17 Exploded View of the HERA.................................................................................191

Table of Contents xi
Final Report of the International Design Synthesis Exercise 2005

List of Symbols

Symbols Definition Unit

− Mean Aerodynamic Chord m


c
A Aspect ratio -
An Area of cells for wing box m2
Aspar Area of spar m2
a Lift curve slope deg-1
acr Speed of sound at cruise m/s
a1 Horizontal Tail Lift Curve Slope rad-1
a2 Elevator Lift Curve Slope rad-1
Bp Number of propeller blades
Br Boom area of stringer and spars m2
b Wing span m
b Spacing in between stringers m
bf Fuselage width m
bf Fin span m
bfo Span outward flap section m
bfi Span inward flap section m
CACQ Acquisition Cost $
CD 3D Drag coefficient -
Cd 2D Drag coefficient -
CD0 Zero-lift drag coefficient -
(CDS)f Drag contribution from fuselage -
(CDS)n Drag contribution from engines -
(CDS)upsweep Drag contribution from upsweep -
(CDS)w Drag contribution from wing -
CL 3D Lift coefficient -
CLα Lift curve -
C Lα a Change in lift curve -

CLα)wf Lift curve fuselage and wing -


CLmax Maximum lift coefficient -
CLmaxL Maximum lift coefficient for landing -

List of Symbols xiii


Final Report of the International Design Synthesis Exercise 2005

CLmax TO Maximum lift coefficient for take-off -


CLwα Wing lift curve -
Cl 2D Lift coefficient -
CMAN Estimated manufacturing cost $
Cp Pressure coefficient -
CPRO Desired Profit $
CRDTE Cost of research, development, test and $
evaluation of HERA
c Mass coefficient -
c’ Extended chord length m
cf Flap chord m
cf root Fin chord at root m
cf tip Fin chord at tip m
cg Mean geometric chord m
cm Moment coefficient -
cMAC Mean aerodynamic chord m
clα Lift coefficient curve -
clb Basic lift distribution -
cr Wing root chord m
ct Wing tip chord m
D Drag N
D Propeller diameter m
D Strut diameter m
Dp Propeller diameter m
DHercules Tire diameter of tires Hercules “/m
DHERA Tire diameter of tires HERA “/m
dT Thrust from a blade element N
dQ Torque to rotate element Nm
E Jone’s edge velocity factor rad-1
E Modulus of elasticity Pa
e Oswald factor -
en Length of wing boxes sides m
F Overall Lift Correction Factor – Stick Fixed -
Fm Loading on main landing gear kg
Fn Loading on nose landing gear kg
G Shear Modulus Pa
g Gravitational acceleration m/s2

xiv List of Symbols


Final Report of the International Design Synthesis Exercise 2005

H True altitude m
H Wing box height m
h Trade-off weight factor -
h Step size -
hf Fuselage height m
I Moment of inertia m4
ih Root angle of incidence deg
J Advance ratio -
K Lift dependent drag factor -
KCG Chronological factor -
Kg Gust alleviation factor -
KN Static Margin -
KRF Rough field capability coefficient -
KSL Strut length coefficient -
KI Factor for calculating lift on wing plus body -
KII Factor for calculating lift on wing plus body -
KVS Variable sweep constant -
k Radius of gyration m
kb Pitch Radius of Gyration -
L Lift N
L Rib pitch m
La Additional lift distribution -
Lb Basic lift distribution -
Lp Rolling Moment due to Roll Rate -
Lv Rolling Moment due to Side slip -
lf Fuselage length m
lHT Horizontal tail arm m
lp Distance propeller to quarter chord line m
lT Tail Plane Setting Arm m
lVT Vertical tail arm m
M Mach number -
M Moment Nm
M0 Maximum take-off mass kg
M∞ Freestream Mach number -
MCR Crew mass kg
ME Empty mass kg
MF Fuel mass kg

List of Symbols xv
Final Report of the International Design Synthesis Exercise 2005

ML Landing mass kg
Mmaxop Maximum operating Mach number -
MPL Payload mass kg
MH Horizontal tail mass kg/m2
MT Tail mass kg/m2
Mtip Tip Mach number of propeller -
Pitching moment due to change in axial
MU -
velocity
Vertical tail mass kg/m2
MW Pitching moment due to normal velocity -
Mw Wing mass kg
Mw Wing mass kg/m2
Mq Pitching moment due to pitch rate -
M_B Body mass kg
M_CREW Crew mass kg
M_E Basic empty weight kg
M_F Fuel mass kg
M_FE Fixed equipment mass kg
M_N Nacelle mass kg
M_OE Operational empty weight kg
M_OP Operation items mass kg
M_PROP Propulsion group mass kg
M_SC Surface control mass kg
M_STR Total structural mass kg
M_T Tail group mass kg
M_TO Maximum take off weight kg
M_UC Landing gear mass kg
Pitching moment due to rate of change of
M_W -
normal velocity
M_ZF Maximum zero fuel weight kg
N Neutral Point -
NHercules Number of wheels on the Hercules -
NHERA Number of wheels on the HERA -
Nv Yawing Moment due to Sideslip -
n Number of criteria -
n Load factor -
n Manoeuvre factor -

xvi List of Symbols


Final Report of the International Design Synthesis Exercise 2005

nDP Propeller rotation speed m/s


nd Product of prop diameter and no. of m/s
revolutions per second
np No. of revolutions per second rev/s
P0 Static power kW
PTO Take-off power hp
Pbr Shaft power available kW
Ps Maximum unfactored vertical load per leg kg
p Pressure N/m2
p0 Static pressure N/m2
q Dynamic pressure N/s2
q Shear flow N/m
q Distributed load N/m
R Specific gas constant of air m2/s2K
R Range km
Rsegment Range of segment km
RC Rate of climb m/s
r Position from axis of rotation m
r Shape factor -
rf Shape factor fuselage -
rn Shape factor engine -
rRE Correction factor Reynolds number -
ruc Shape factor undercarriage -
rw Shape factor wing -
S Wing area m2
Sf Fin surface m2
SHT Horizontal tail area m2
SLFL Landing field length m
SNET Exposed wing area m2
ST Horizontal Tail Area m2
STOFL Take-off field length m
SVT Vertical tail area m2
Swf Wing area corrected for flaps m2
T Air temperature K
T Torsion Nm
T0 Engine static thrust N
t Thickness m

List of Symbols xvii


Final Report of the International Design Synthesis Exercise 2005

Ude Velocity of the gust m/s


V True airspeed m/s
V Shear force N
VA Manoeuvring speed m/s
VB Gust velocity m/s
VC Cruise speed (TAS) m/s
VD Design dive speed (TAS) m/s
VR Relative Velocity on blade element m/s
V0 Forward velocity of aircraft m/s
Va Approach speed m/s
Ve Equivalent airspeed m/s
Vfw Usable fuel tank capacity in the wing m3
Vs Stall speed m/s
Vy Shear Force N
W Aircraft weight N
W Tire width “/m
Wampr Fraction of take off weight Lbs
WL Aircraft design landing weight kg
w Wing loading Pa
X Design option score -
XU Axial force due to axial velocity -
XW Axial force due to normal velocity -
Xq Axial force due to pitch rate -
Axial force due to rate of change of normal -
X_W
velocity
xac Wing aerodynamic chord m
xAFT x-coordinate aft CoG position m
xFRONT x-coordinate front CoG position m
xm x-coordinate main landing gear m
xMTOW x-coordinate CoG position for MTOW m
xn x-coordinate nose landing gear m
yE Lateral distance of engine from CoG m
ZU Normal force due to axial velocity -
ZW Normal force due to normal velocity -
Zq Normal force due to pitch rate -
Normal force due to rate of change of normal
Z_W -
velocity

xviii List of Symbols


Final Report of the International Design Synthesis Exercise 2005

zAFT z-coordinate aft CoG position m


zFRONT z-coordinate front CoG position m
zMTOW z-coordinate CoG position for MTOW m

∆zCL Increment of lift due to fuselage interference m


Γ Dihedral deg
Л Wing sweep -
Λ1/4 Sweep quarter chord line -
Ω Angular velocity of blade rad/s
α Angle of attack deg
α Second segment climb angle deg
αf Fuselage incidence deg
αr Incidence of root section deg
άδ Theoretical flap lift factor -
α 0 root Zero lift angle of attack at wing root deg

α 0tip Zero lift angle of attack at wing tip deg

(αl0)r Zero lift angle of root section deg-1


α01 Zero lift angle of attack / unit of twist deg-1
β Sideslip angle deg
β Trim Tab Angle deg
γ Ratio of specific heat -
γ Second segment climb angle deg
δf Maximum flap deflection deg
εt Aerodynamic twist deg

ε gt Geometric twist deg

ζ Rudder angle rads


η Normalised spanwise coordinate -
η Elevator Angle deg
ηp Propeller efficiency -
ηp Propulsive efficiency -
ηT Tail Plane Setting Angle deg
ηδ Flap efficiency factor -

θ&& Pitch Acceleration -

λ Taper ratio -
Φ Roll rate deg/s

List of Symbols xix


Final Report of the International Design Synthesis Exercise 2005

φ Angle of relative velocity on blade element to -


plane of rotation
φ Power over frontal area hp/m2
ρ Air density kg/m3
σ Relative density
σ Stress Pa
σ Solidity of the annulus (ratio of total blade -
area to total area of annulus)
τ Shear Stress Pa
τ Thrust correction factor -

Acronyms and Abbreviations Definition


AEO All Engines Operating
AEP Estimated price per aircraft ($)
AEM Aircraft Empty Mass
AEW Aircraft Empty Weight
CGR Climb gradient
CoG Centre of Gravity
CS Certification Specifications
DSE Design Synthesis Exercise
EAS Equivalent Airspeed
FAR Federal Aviation Regulation
FBW Fly-by-Wire
FFD Functional Flow Diagram
JAR Joint Airworthiness Requirements
MTOW Maximum Take Off Weight
NGO Non Governmental Organisation
NZG Near Zero Growth
OEI One Engine Inoperative
SFC Specific Fuel Consumption
TAS True Airspeed
ZFM Zero Fuel Mass

xx List of Symbols
Final Report of the International Design Synthesis Exercise 2005

List of Figures

Figure Page

Figure 2.2.1 Top level of the requirements discovery tree. 4


Figure 2.4.1 Software block diagram for the FBW system. 7
Figure 2.5.1 Risk as function of probability of occurrence and severity of effect 7
Figure 2.5.2 Risk Map Design Possibilities 8
Figure 2.6.1 Complete work flow diagram for the HERA project. 10
Figure 3.1.1 The external configuration of the HERA 11
Figure 4.1.1 HERA wing planform 17

Figure 4.1.2 Cp distribution of the LS(1)-0417MOD airfoil at α = 2.65° and for Re = 25 ⋅ 10 6 and
M = 0.55 19

Figure 4.1.3 Cp distribution of the LS(1)-0413MOD airfoil at α = 1.43° and for Re = 25 ⋅ 10 6 and
M = 0.55 19
Figure 4.1.4 XFOIL results for the root and tip airfoil 19
Figure 4.1.5 Cp distribution over airfoil 20
Figure 4.1.6 Pressure distribution over airfoil Figure 4.1.7 Velocity distribution over airfoil 21
Figure 4.1.8 C L , α -curve of the wing 22
Figure 4.1.9 Zero-lift angle of attack per unit of twist for straight wings [37] (page 475) 23
Figure 4.1.10 Spanwise lift distribution 25
Figure 4.3.1 Location of the mean aerodynamic chord and aerodynamic centre of wing and wing / body
31
Figure 4.4.1 Fuselage / wing interference effects on the spanwise lift distribution [37] (page 478) 34
Figure 4.4.2 Geometric definitions of the various wing / body angles [37] (page 478) 35
Figure 5.1.1 Basic Dimensions of the Engine Nacelles 40
Figure 5.2.1 Change of Propeller Efficiency with Change in Pitch 44
Figure 5.2.2 Final Propeller Design 44
Figure 6.1.1 Constraints diagram 52
Figure 6.3.1 V-n diagram (manoeuvre envelope and gust envelope combined) 56
Figure 6.4.1 Flight envelope of the HERA 57
Figure 6.5.1 Payload-range characteristics of the HERA. 59
Figure 7.2.1 From top to bottom: a) Spanwise wing loading, b) Shear force diagram, 64
Figure 7.2.2 Spanwise variation of torsion about flexural centre 65
Figure 7.2.3 Idealised wing box 66

List of Tables xxi


Final Report of the International Design Synthesis Exercise 2005

Figure 7.2.4 Idealisation of structure for bending shear stress analysis 68


Figure 7.2.5 Notation and convention used for pure torsion loading 70
Figure 7.2.6 Internal view of root section 72
Figure 7.2.7 Close up of spar cap at root 72
Figure 7.2.8 Stringer close up 72
Figure 7.3.1 Torsion model for the centre part of the fuselage. 74
Figure 7.3.2 Torsion model for the front and rear part of the fuselage 75
Figure 7.3.3 Torsion model for the tail torsion box 75
Figure 7.3.4 Load case of the aircraft with its shear force and bending moment diagram 78
Figure 7.3.5 Structural modal of the bending analysis for the fuselage 79
Figure 7.3.6 Fuselage skin with Z-type stringers 80
Figure 7.3.7 Structural model of the shear force analysis for the fuselage 81
Figure 7.4.1 Overview of horizontal tail structure 83
Figure 7.4.2 Close up of horizontal tail spar cap 83
Figure 7.4.3 Stringer close up 84
Figure 7.4.4 Overview of fin structure 85
Figure 7.4.5 Close up of fin spar cap 85
Figure 7.4.6 Stringer close up 85
Figure 7.5.1 Configuration of backside of HERA 86
Figure 7.5.2 Ground track of wheels on ramp 87
Figure 7.5.3 Pallet positioned on ramp 87
Figure 7.5.4 Locker makes sure cargo floor and ramp stay parallel 88
Figure 7.5.5 Ramp cross section 88
Figure 7.5.6 Load case on I-beam 89
Figure 7.5.7 Free-body-diagram I-beam 89
Figure 7.5.8 Maximum loads to be distributed by the ramp 90
Figure 8.2.1 Taxi load cases 94
Figure 8.2.2 Lateral Tip-over 95
Figure 8.4.1 Layout of double acting strut 96
Figure 8.7.1 Moveable shelter design 98
Figure 10.1.1 From left to right: a) Geometry vertical tail plane, b) Geometry horizontal tail plane 110
Figure 10.2.1 Landing Configuration Constraints Diagram 115
Figure 10.2.2 Cruise Configuration Constraints Diagram 116
Figure 10.2.3 Pitching Moment Characteristic Graph for the HERA in Steady Level Flight 117
Figure 12.1.1 Reduction in RDTE Cost per Aircraft due to Increase in Number of Aircraft Produced
139
Figure 12.3.1 General Project Cash Flow 143

xxii List of Tables


Final Report of the International Design Synthesis Exercise 2005

List of Tables

Table Page

Table 3.5.1 HERA general specifications 15


Table 4.1.1 HERA wing geometry 18
Table 4.1.2 Main parameters from the cl , α -curve 19
Table 4.1.3 Parameters used and results obtained for flap analysis 27
Table 4.1.4 Maximum lift coefficients and stall speeds for cruise, takeoff and landing 28
Table 4.2.1 Increments for the drag polar for landing and takeoff configurations 31
Table 4.3.1 Required values for the use of Simpson’s rule 33
Table 5.1.1 Basic Information on the Ae1107c Engine 39
Table 5.2.1 Basic Outputs from Propeller Design 43
Table 5.5.1 Table of Emissions for the Ae3007 Engine 46
Table 6.1.1 Cruise correction factors 51
Table 6.1.2 CS-25 Climb Requirements 51
Table 6.1.3 Assumed aerodynamic values 52
Table 7.1.1 Material characteristics 61
Table 7.2.1 Expressions for q, V and M for the wing structural mass and the fuel mass. 63
Table 7.2.2 Summary of wing structure dimensions 71
Table 7.3.1 Results of the torsion calculations 77
Table 7.3.2 Cost comparison 80
Table 7.3.3 Minimum skin thickness due to maximum shear loading at the centre fuselage 81
Table 7.3.4 Total mass and material cost of the fuselage skin for combined shear and torsional loading
82
Table 7.4.1 Results for horizontal tail analysis 83
Table 7.4.2 Results of the fin structural analysis 85
Table 7.5.1 Optimisation results for cargo ramp 91
Table 8.2.1 C.o.G. locations 93
Table 8.2.2 Landing gear locations 94
Table 8.2.3 Loading ratios 94
Table 8.4.1 Strut loading 97
Table 8.4.2 Load-Stroke 97
Table 8.12.1 Landing gear weight 100
Table 9.1.1 Class One Mass Statement 102
Table 9.2.1 Aircraft Mass Statement – Class 1 104

List of Tables xxiii


Final Report of the International Design Synthesis Exercise 2005

Table 9.3.1 Centre of gravity assumptions 105


Table 9.3.2 Centre of Gravity Locations Relative to the Mean Aerodynamic Chord 106
Table 10.1.1 Tailplane dimensions 110
Table 10.2.1 Horizontal Tail Design 118
Table 10.2.2 Aero-Normalised Stability Derivatives 119
Table 10.2.3 Short Period Oscillation and Phugoid Time Signatures 121
Table 10.2.4 Short Period Oscillation and Phugoid Characteristic Coefficients 121
Table 10.3.1 Lv = Rolling Moment due to Sideslip 123
Table 10.4.1 Results of Nζ for a Range of Values of Sf/S and cf/c 126
Table 10.4.2 Results of the fin and rudder sizing 127
Table 10.4.3 Results for Dynamic Modes 130
Table 10.5.1 Summary of the main results 130
Table 12.4.1 DOC break down 145
Table 12.4.2 Total operational cost of aircraft 146
Table 14.1.1 Compliance matrix. 150

xxiv List of Tables


Final Report of the International Design Synthesis Exercise 2005

1 Introduction

In light of the recent tsunami disaster, air transport has been proven to be indispensable in humanitarian
operations. Countries from all over the world showed compassion by providing immediate live-saving aid
and long-term support for rebuilding the victim countries. This requires fast, reliable and robust air
transport of emergency aid, such as food supplies, vehicles, evacuees and medical equipment. For that
purpose aid organisations around the world have relied on the support from the governments. Cutbacks
in defence budgets as well as the increasing need of aid organisations to maintain impartiality in emergency
situations, have erected the need for a robust, low-cost, reliable cargo aircraft, which can be independently
operated by the aid organisations for delivering humanitarian relief and emergency aid cargo transport.

The aim of this report is to present the final preliminary design of the Humanitarian Emergency Relief
Aircraft (HERA), an aircraft capable of performing such a mission. The report will outline the tradeoffs
that were made during the preliminary design studies and will present preliminary results on for example
the aerodynamics, performance, structural design and stability of the HERA. It will show that the HERA
is a feasible concept and that it is ready for full-scale development.

The structure of the report is as follows. First, chapter 2 will discuss the mission specifications of the
HERA and will present a project life cycle planning. In chapter 3, the general layout and the top-level
specifications of the HERA will be introduced. Subsequently, chapter 4 will present an aerodynamic
analysis, including a drag breakdown and a determination of the position of the aerodynamic centre. The
engines of the HERA will be discussed in chapter 5. This is followed up by a performance analysis in
chapter 6, in which amongst others the climb performance and the payload-range performance of the
HERA will be discussed. The structural design of the wing, fuselage, tail and cargo ramp will be outlined
in chapter 7. Chapter 8 will bring up the design of the undercarriage. Next, chapter 9 will present a weight
breakdown and an analysis of the aircraft centre of gravity. The longitudinal and lateral control and
stability of the HERA will be studied in chapter 10. Subsequently, chapter 11 will give a manufacturing,
assembly and certification plan. In chapter 12 a thorough cost analysis will be conducted. A sustainable
development strategy will be presented in chapter 13. Finally, chapter 14 will sum up the conclusions and
will provide recommendations for further development.

Chapter 1 Introduction 1
Final Report of the International Design Synthesis Exercise 2005

2 Mission Analysis and Project Planning

The best design for the HERA is that design, which best satisfies all requirements. For that reason it is
very important to have a good insight into the aircraft mission and the related requirements. For that
purpose, the first part of this chapter will present a thorough mission analysis, including the mission
specification (section 2.1), an overview of the top-level requirements (section 2.2), a functional analysis
(section 2.3) and a risk analysis (section 2.4). Subsequently, the second part of this chapter (section 2.5)
will present a project design and development logic, outlining a planning for the complete project life
cycle. The current chapter will form the foundation for the remaining design process.

2.1 Mission specification

The aircraft’s mission is to deliver humanitarian aid to almost every place in the world under the most
severe conditions. The aircraft should be highly reliable and very cost-effective. It must be fast, safe and
easy to operate by non-governmental organisations (NGO’s).

The aircraft must be able to meet the NGO’s needs in terms of:
• Cargo transportation of emergency aid
- Bulk (food, water, fuel)
- Containers, pallets
- Equipment (jeeps, tents, portable medical centres)
• Emergency evacuation of people in case of natural or human disasters in lieu of normal civil
airliners
• Mobile operating theatre
• Easy and cheap maintenance under primitive conditions

The aircraft must be able to operate under the most difficult conditions in terms of:
• Weather
• Climate
• Terrain
• Hostilities (evasive action)

Currently most NGO’s rely heavily on support from governments around the world in meeting their
transport needs. However, cutbacks in defence spending as well as the increasing need to maintain their
impartiality in emergency conflict situations are reducing the government’s readiness to aid in

Chapter 2 Mission Analysis and Project Planning 3


Final Report of the International Design Synthesis Exercise 2005

humanitarian relief operations. Therefore, the need has arisen between NGO’s to have their own fleet of
aircraft, which are distinctively recognisable around the world as impartial humanitarian relief aircraft. The
mission requirements for this aircraft were specified by the client and have been included in appendix 1.
Based on the mission described above, the Mission Need Statement can be formulated. This is necessarily
limited to the bear essentials of what the aircraft is required to achieve and is given as follows:

“To provide low-cost, flexible and reliable air transport of aid, evacuees and medical equipment under difficult
circumstances whilst remaining impartial.”

From the Mission Need Statement the driving requirements, which drive the design to a large extent, can
immediately be identified. In the next section the Mission Need Statement will be expanded to a set of
top-level requirements.

2.2 Top-level requirements

Requirements are the foundation of the design process. It is therefore very important to have a good
overview of all relevant requirements. The purpose of this section is to provide an overview of the top-
level requirements. A complete overview can be found in reference [8].

The requirements have been derived from the mission specification and can be presented in a
requirements discovery tree. Figure 2.2.1 presents the top-level structure of the tree with a brief
explanation of each top-level requirement.

Figure 2.2.1 Top level of the requirements discovery tree.

4 Mission Analysis and Project Planning Chapter 2


Final Report of the International Design Synthesis Exercise 2005

The next step to defining requirements is to consider each category in turn and expand these into full sets
of quantifiable requirements. For that purpose the airworthiness requirements in the CS-25 should be
considered for example. This has been done for the mid term report, reference [8], but will not be
discussed here anymore.

From the top-level of the requirements discovery tree the acquisition cost of $30 million per unit can be
identified as a very important driving requirement. Other aircraft with a mission similar, such as the C130
Hercules and the Airbus A400M, are 2 to 3 times as expensive (a comparison of the HERA to similar
aircraft can be found in Appendix 2). Therefore, cost has been made one of the criteria for performing the
tradeoffs and ‘drives’ the design to a much greater extent than most other requirements.

Another design driver is the time requirement that the aircraft should be ready for service by 2010.
Nowadays most aircraft design projects of this size have a duration of 15 to 20 years. The requirement will
force the design towards proven concepts and allows little development.

The requirements analysis, which has been discussed in this section, does not stand on its own, but is very
much interrelated with the functional analysis. This will therefore be discussed in the next section and will
provide a deeper insight into the mission characteristics.

2.3 Functional Analysis

2.3.1 Functional flow diagram


Functional flow diagrams (FFD’s) help in discovering mission solutions and help to focus attention on
what the system must be able to do. This ensures that the system is built around the needed mission
requirements and functions instead of preconceptions.

The functional flow diagram can be found in Appendix 3.1. The initial level is the mission need statement
upon which the system must carry out. The functions below this must ensure that the initial mission
statement is fulfilled. The top-level functions are derived from the top-level requirements, which are
displayed in Figure 2.2.1. For example, the requirements discovery tree revealed the need for payload in
1.2; the associated function is therefore that the system must carry the payload.

Subsequently, the top-level functions are then divided into lower level functions, which describe the top-
level functions in more detail. To carry the payload includes carrying both cargo and evacuees. This is the
second level of functions and the functions are numbered 1.2.1 and 1.2.2 respectively.. The second level
can then be divided to give yet more detail; carrying cargo describing carrying pallets, vehicles, bulk and
equipment.

Chapter 2 Mission Analysis and Project Planning 5


Final Report of the International Design Synthesis Exercise 2005

2.3.2 Functional breakdown


The functions of section 2.3.1 can be grouped in a functional flow breakdown. This breakdown has been
included in Appendix 3.2 and can be divided in three categories: aircraft operation, maintenance and
support.

The aircraft must be operated to perform missions of air transport, the goal of these missions varying
between passenger operations, freighter operations and non-revenue operations, all performed under
various circumstances. The operation of the aircraft can be broken down in pre-flight operations, take-
off/landing operations, flight operations and post flight operations. This envelope must be executable
under various circumstances, anything from nominal to harsh, the latter more likely to happen in this type
of missions.

At the same time inspections are necessary to maintain the aircraft operational. These may be the expected
regular inspections or supplementary inspections when there is cause to believe that damage was inflicted
during operations. Also, to prolong the aircraft’s fife, upgrades might be necessary to adapt it to future
mission needs or simply to comply with new airworthiness regulations. For both these cases a production,
delivery and repair line of spare and new parts must be kept active in parallel during the aircraft’s life cycle.

Finally, the aircraft in itself cannot perform a mission without an organisation that is a resource of
logistics, organisation and education. Every type of aircraft demands a special training for the flight and
operations crew – such as flight crew and technicians. Therefore a system must be put in place to ‘educate’
the personnel working in and around the aircraft to ensure its safe maintenance.

2.3.3 Operations concept and flow diagram


HERA’s operation concept is to safely carry on an emergency evacuation of people in case of natural or
human disasters, to transport emergency aid cargo and/ or mobile operating theatres with medical staff
while maintaining high reliability and safety.

To show how the operations are carried on, an operations flow diagram has been developed that can be
found in Appendix 3.3. This diagram encompasses the three types of missions and describes the
procedure to follow to carry on the mission. It is essentially the operations done during the mission that
are also described in the functional flow diagram, see section 3.2.1 and or Appendix 3.1.
The top level chain of the operation is to first load cargo, then taxi and take off after a standard safety
check to head for the disaster zone. Once arrived, the aircraft is unloaded and/or departs again to carry
the evacuees to a safe zone. The second level gives a more detailed overview of each operation performed
per top-level box. The operations must be done sequentially before going to the next step to guarantee a
safe mission. This operations flow diagram may be used later on for the writing of the flight manual.

6 Mission Analysis and Project Planning Chapter 2


Final Report of the International Design Synthesis Exercise 2005

2.4 Hardware-software block diagrams

From the functional analysis (section 2.3) it is possible to indicate whether a certain function is performed
by a hardware of a software element. Since the HERA is an aircraft, it will mainly consist of hardware.
From the functional flow diagram, only 1.1.3 Manoeuvre, 1.1.5 Navigate and 1.6.2.4 Detect threat, use SW
elements. One such element is the Fly-by-Wire system (function 1.1.3). The flow diagram for that system
is given in Figure 2.4.1.

Figure 2.4.1 Software block diagram for the FBW system.

2.5 Technical risk assessment

A risk assessment is made to determine which features on an aircraft pose the highest risk of disrupting
the normal operation of the aircraft. In order to determine the risk one needs to know the probability that
a certain failure occurs and the possible consequences of that failure. The risk is the product of these two,
as can be seen in Figure 2.5.1.

Figure 2.5.1 Risk as function of probability of occurrence and severity of effect

Chapter 2 Mission Analysis and Project Planning 7


Final Report of the International Design Synthesis Exercise 2005

Although it is very difficult or even impossible to determine the exact possibility that a certain malfunction
occurs and what the possible consequences of this malfunction will be, it should be possible to say
whether this possibility is high or low and how severe the consequences of that failure would be.

2.5.1 Risk identification


Some factors that can produce risk are:

1. Engine failure 14. Interior environment system failure


2. Ramp failure 15. Cockpit oxygen supply failure
3. Undercarriage failure 16. Emergency systems failure
4. LG Shelter failure 17. Flaps failure
5. Tire Rupture 18. Flight control software error
6. Brakes failure 19. Autopilot failure
7. Loading system failure (Rollers or winch) 20. Communication system failure
8. Elevator failure 21. Flight management system failure
9. Rudder failure 22. Navigation system failure
10. Aileron failure 23. Pilot error
11. Wing structure failure 24. Extreme weather phenomena’s
12. Fuselage structure failure 25. Stability difficulties
13. Evasive measures failure 26. Short takeoff and landing (unprepared runways)

2.5.2 Risk analysis: the risk map


A very useful tool to give an overview of the risk factors is a risk map. A risk map is a table, which
categorises factors that can produce risk to their consequence on performance and their probability of
failure. The risk map for the factors that pose a certain risk mentioned above can be found in Figure 2.5.2.

High
High moderate 23 26
Moderate 4 1, 24
Failure probability

Low moderate 20 15 2, 5, 21, 22


Low 14 7, 19 25 6, 8, 9, 10, 11,
12, 13, 16, 17, 18
Negligible Marginal Critical Catastrophic

Performance Consequence

Figure 2.5.2 Risk Map Design Possibilities

8 Mission Analysis and Project Planning Chapter 2


Final Report of the International Design Synthesis Exercise 2005

2.5.3 Risk handling


In order to minimise risk, you can either try to minimise the change of occurrence of the failure, or try to
minimise the effects on performance. For some factors however, this can be difficult because high risk is
inherent to some factors. Short take-off and landing from unprepared runways (26) is always a risky
manoeuvre. There is not much that can be changed about this because it depends on the terrain, which
cannot easily be changed. Another factor that is a frequent cause of aircraft mishaps is pilot error (23).
Every time a human is part of a system, it is usually the systems weakest link. People will make an error at
some time; it is difficult to prevent. Probably the system that fails most on aircraft is the engine (1). Trying
to make an engine as reliable as possible can decrease the probability of this failure. Fortunately, due to
the number of engines, engine failure does not necessarily leads to an accident. The same goes for tire
rupture (5), something that can be common for aircraft operating from rough terrain. The tires used on
the HERA are designed to operate on unprepared runways, but tire rupture will happen at some time.
Thanks to the number of tires however, it does not automatically mean that the aircraft has become
useless. It can still be operated and finish the mission, albeit with some difficulty. Another thing that will
happen quite a number of times during a planes operational life, is experiencing extreme weather
phenomena. Although some of these phenomena can do quite some damage to the aircrafts exterior, the
plane should be designed in such a way that the aircraft is still able to operate. Ramp failure (2), Flight
management system failure (21) and Navigation system failure (22) are examples of things that can go
wrong, but will not immediately put the aircraft in danger. Brakes failure (6), Elevator failure (8), Rudder
failure (9), Aileron failure (10), Wing structure failure (11), Fuselage structure failure (12) and Evasive
measures failure (13) however will pose a great danger to the aircrafts performance and can very possibly
be the cause of an accident. Therefore the only way to minimise the risk is to keep the probability of
occurrence as low as possible. Fortunately, after over one hundred years of aviation, aircraft manufactures
know how to keep these failures from happening. The only thing that can change this is neglecting
maintenance.

2.6 Project design and development logic

The design and the development of an aircraft of the size of HERA may take years till market
introduction. The customer requirement is that the aircraft has to make its first flight in 2010. A decent
project planning is therefore indispensable.

The complete workflow from project initiation until aircraft disposal for the HERA project is outlined in
the project design and development logic, see Figure 2.6.1 below. The forecast is that the HERA will be
ready for market introduction by January 2012. The design philosophy is to use off-the-shelf equipment
and proven technology in order to limit the time required for development and certification.

Chapter 2 Mission Analysis and Project Planning 9


Final Report of the International Design Synthesis Exercise 2005

This report is indicated in Figure 2.6.1 by “DSE Final Report” and formally concludes phase 3 of the
project life cycle. The project design and development logic below can also be translated in the form of a
project Gantt chart. This Gantt chart can be found in Appendix 4.

Figure 2.6.1 Complete work flow diagram for the HERA project.

10 Mission Analysis and Project Planning Chapter 2


Final Report of the International Design Synthesis Exercise 2005

3 General Layout

This chapter will present the general layout of the HERA. First, in section 3.1 the external configuration
will be discussed. Subsequently, section 3.2 will outline the cabin layout. The cockpit interior will be
discussed in section 3.3. Section 3.4 briefly explains the choice of the flight control system. Finally, the
general specifications of the HERA will be listed in section 3.5.

3.1 External configuration

The external configuration of the HERA is not surprising when compared to other cargo aircraft, see
Figure 3.1.1. A three-view drawing of the HERA can be found in Appendix 5.

The high wing layout was selected to help prevent structural obstruction of the cargo bay floor. High wing
aircraft tend to be lower to the ground relative to low wing configurations and this assists in meeting the
mission requirements for loading the aircraft. The high wing is necessary to ensure propeller ground
clearance for the under wing nacelle configuration. There is one significant draw back of the high wing
configuration and that is the difficulty of inspecting and maintaining the structure.

Figure 3.1.1 The external configuration of the HERA

The wing design itself has several benefits for the aircraft. The high aspect ratio results in a very high lift-
to-drag ratio, and hence an improvement in aircraft range. There are problems inherent in a high aspect
ratio design. Firstly is the poor performance in turbulence. This will place greater demands on the avionics
systems to meet the mission requirement for operations in bad weather.

Chapter 3 General Layout 11


Final Report of the International Design Synthesis Exercise 2005

The fuselage design has been driven by the mission requirements for cargo capacity. This has specified the
minimum width and length of the cargo bay and the cockpit and rear have been designed around this.
Given the high drag due to the fuselage shape it is essential that the landing gear be stored in fairings and
the slope of the rear doors be controlled so as to limit pressure drag due to boundary layer separation.

The HERA is powered by four Rolls Royce AE1107c turboprop engines. Providing a shaft power of 4568
kW each, these engines are the most powerful turboprops currently available in the West and the only one
capable of allowing the aircraft to meet the mission requirements and service entry date of 2010. They are
used on the V-22 Osprey as well. The power plants will be further discussed in chapter 5.

The HERA has a conventional tail. The tail was designed initially around volume ratios taken from similar
aircraft configurations. This first estimate was later refined based on the aircraft stability analysis to ensure
adequate damping of the oscillatory and non-oscillatory modes of motion, adequate static stability in all
flight configurations, and the trim requirements.

The landing gear system has been designed to be semi-retractable and is based on existing military
technology found in the C-130. The main gear consists of four wheels in series attached to the main
fuselage. The gear is housed in fairings to reduce aerodynamic drag during flight. Due to the relatively far
forward centre of gravity position, it has been necessary to design stronger nose gear. The nose gear is
therefore a single bogie version of the main gear. This has reduced part count benefits and improves the
rough field handling qualities.

3.2 Cabin layout

The driving design requirement of the aircraft fuselage is primarily that of a cargo transporter with
secondary requirements including medical relief and civilian evacuation. Consequently, the cabin or more
specifically the cargo bay area has driven the overall design to a high degree.

3.2.1 Payload
The conclusive dimensions of the cabin in length, width and height are 20 m, 4.4 m and 3.5 m,
respectively. Below will be described what kind of payload can be carried and what payload configurations
could be attained. In Appendix 6 one can find a detailed drawing of the possible payload configurations.

12 General Layout Chapter 3


Final Report of the International Design Synthesis Exercise 2005

The cabin is designed to be able to carry:


• 2 light vehicles side by side
• 9 pallets of type 3610 (2.44m by 3.20m)
• 9 pallets of type 463L (2.15m by 2.65m)
• carry up to 130 evacuees
• transport a medical unit

The top edges of the cabin are not needed for carrying payload. The extra space that becomes available
will then be used for the wirings of the fly-by-wire system, air-conditioning system and other aircraft
systems. In Appendix 6.2 a front view drawing has been included of the cabin interior with different
configurations and dimensions to give an indication about its size.

It must be mentioned that there will be no problem to fit a medical unit inside the cabin. There are
companies (reference [50]) designing medical units on customer-based preferences. From references [42],
[49] and [58] information can be obtained of some currently used light vehicles.

As stated above, the cabin must allow 9 pallets of type 3610 or 463L (smaller ones) to be transported. Of
these pallets 8 will be carried on the fixed floor of the cabin, and one pallet will be carried on the ramp.
The pallets are all positioned in line and can be attached to the floor or ramp with lockers to secure them
during flight. In the first 4 meters of the fuselage cabin multidirectional rollers are positioned in the floor.
The task of these rollers is to move equipment to the sides of the cabin since the width of the ramp is
smaller than the cabin itself.

The cabin also allows for transport of 130 evacuees. The seats must be foldable and in the middle of cabin
removable in order to create flexibility in the cabin. When seats are not needed they do not have to be
transported as well. The removable seats are positioned in 4 rows of 34 seats and can be locked in the
floor as can be seen in Appendix 6.2. While transporting 130 evacuees there still is room for carrying 10
stretchers, 2 pallets of 3610 or 463L type or 2 light vehicles side by side. Furthermore, the cabin allows
carrying 70 stretchers in total or 8 light vehicles, which can be loaded and unloaded quickly.

3.2.2 Non-pressurised
Very characteristic for the HERA design is the non-pressurised cabin. This results in a very lightweight
and low-cost fuselage structure. The effect of that is enormous: the range of the HERA is around 40%
bigger than that of other cargo aircraft with a pressurised cabin. Also, because of the lower weight, less
powerful engines are required, which will result in a further weight and cost reduction.

Chapter 3 General Layout 13


Final Report of the International Design Synthesis Exercise 2005

In the case when the HERA needs to transport evacuees, the aircraft will be flying at a lower (emergency)
altitude of 2000 m, in order to have a convenient cabin climate for the passengers.

3.2.3 Autonomous on- and offloading


For fast on- and offloading the aircraft has been equipped with a big cargo ramp in the back. The cargo
ramp is 4 meters long and 3.5 meters wide and has been designed to withstand the maximum loads that
can occur during on- and offloading. As the opening must also be high enough, two split doors are
positioned above the cargo door. The ramp itself will be raised and lowered via a hydraulic piston with a
manual override in case of a pressure loss or other failure. The design of the cargo door will be further
discussed in section 7.5.

3.3 Cockpit interior

The crew of the HERA exists of four, namely two pilots, a loadmaster and a flight engineer/navigator.
During the flight, they are situated in the cockpit. Together they can operate the HERA to its full extend.

When flying with a non-pressurised aircraft, the crew needs oxygen supply above a certain height. During
the day this height is around 2440 m and during the night 1520 m since flying at night results in a higher
workload for the pilots. Therefore there are oxygen masks on board of the HERA. Together with the
masks, flow meters are used to control the oxygen flow. At an altitude of 7620 m a concentration of 50%
is needed. A pulse oxymeter is slided on a fingertip of the crew and uses red and infrared light to measure
the oxygen saturation of the tissue. Since there are no rules on the needed oxygen flow (it differs form
person to person), the pulse oxymeter provides the crew with valuable information when flying above
1520 m.

3.4 Flight Control System

The flight control system used for HERA is a fly-by-wire system. Unfortunately, at the moment the
acquisition cost are five to ten times as high as for a mechanical control system, however in the future this
incredible cost disadvantage will decrease since it has to be operational in 2010, although 2012 will be
more probable. A direct weight reduction of 300 kilograms can be expected by replacing the conventional
mechanical control elements by electrical components. A second indirect weight reduction can be
expected because structural parts will need less stiffening.

Due to this fly-by-wire system safety and aircraft handling are also improved because of stabilising
functions of computers and through automatic compensation for flight configuration changes. The fly-by-
wire system will improve the short takeoff and landing capabilities of the HERA.

14 General Layout Chapter 3


Final Report of the International Design Synthesis Exercise 2005

3.5 Aircraft Specifications

The following table gives an overview of the main characteristics of the HERA.

Table 3.5.1 HERA general specifications


Dimensions
Length 38.43 m
Wing span 48.5 m
Height 11.92 m
Fuselage diameter 4.88 m

Cabin
Maximum height 4.1 m
Maximum width 4.88 m
Length including ramp 23.5 m
Floor area 85.5 m
Max. Number of Pallets: 9 of type 463L or
9 of type 3610 or
1 mobile hospital or
130 seats or
70 stretchers
Cabin volume 400 m3

Weight
Maximum take-off weight 101 124 kg
Empty weight: 42 175 kg
Maximum allowable cabin 35 000 kg
load:

Engine
Engines Ae1107c
Number of propellers 6
Propeller diameter 4.4 m
Fuel tanks 5 on wings and
1 on fuselage

Performance

Maximum level speed 183 m/s


Maximum rate of climb 7 m/s
Service ceiling 7520 m
Take-off run 950 m
Landing run 950 m
Ferry range (with centre 14 600 km
section wing tank)
Range with full payload 5 460 km

Cost
Unit Cost US$ 29.7 M

Crew
Crew 2 pilots
1 navigator
1 loadmaster
Crew Complement medical 2 medics
missions
Date to be deployed 2012

Chapter 3 General Layout 15


Final Report of the International Design Synthesis Exercise 2005

4 Aerodynamics

Aerodynamics is very important for a good aircraft performance. In the present chapter the aerodynamics
of the HERA will be discussed. First, section 4.1 will investigate the aerodynamics of the wing. Next,
section 4.2 will present a drag breakdown of the aircraft yielding a value for C D 0 . The location of the
aerodynamic centre is very important for stability and control (chapter 10) and will be analysed in section
4.3. Finally, section 4.4 will give the lift of the complete aircraft including fuselage effects.

4.1 Wing

This section will discuss the aerodynamic design of the wing. The structural design will be discussed in
section 7.2. Main requirements posed on the aerodynamic design of the wing are:
• High C L max for short takeoff and landing: C L max TO ≥ 2.8 , C L max L ≥ 3.55

• High L/D in cruise (Mach 0.55): (L / D )cruise ≥ 16

• 17% thickness for structural reasons


• Favourable stall characteristics

4.1.1 Wing geometry


The wing planform chosen for the HERA is a straight tapered wing, as can be seen in figure 4.1. The
straight wing ensures a simple, lightweight and low-cost design. It is not necessary for the wing to be
swept, as the critical Mach number will not be reached while cruising at around Mach 0.55.

The large taper of λ = 0.36 is an attempt to approximate an elliptical lift distribution. Together with the
high aspect ratio of 11 it contributes to a good L/D during cruise and hence improves the performance of
the aircraft considerably.

Figure 4.1.1 HERA wing planform

Chapter 4 Aerodynamics 17
Final Report of the International Design Synthesis Exercise 2005

From the wing planform in figure 4.1 it can be seen that the flaps take up a large part of the total wing
area. The flaps are required to give the aircraft a high C L max for short takeoff and landing (STOL)
operations. They allow the aircraft to takeoff and land on unprepared runways within a very short distance
of 950 m.

The parameters related to the wing geometry are all summed up in table 4.1.

Table 4.1.1 HERA wing geometry


Parameter Symbol HERA Specification
Wing area S 214.1 m2
Wing span b 48.5 m
Aspect ratio A 11
Mean aerodynamic chord c 4.737 m
Mean geometric chord cg 4.112 m
Taper ratio λ 0.36
Sweep of quarter-chord line Λ1 / 4 0 deg
Root airfoil - LS(1)-0417MOD
Tip airfoil - LS(1)-0413
Geometric twist εt -1.22 deg
Root angle of incidence ih 2.0 deg
Dihedral angle Γ -2.0 deg

4.1.2 Airfoil analysis


The main selection criteria for the airfoil selection were:
• A high cl max for short takeoff and landing

• A high cl / c d for economy in cruise (Mach 0.55)

• 17% thickness for structural reasons


• Favourable stall characteristics

Several airfoils have been analysed using XFOIL [59]. In the end the NACA-LANGLEY LS(1)-
0417MOD was selected as the root airfoil for the HERA. For structural reasons a thinner airfoil was
selected for the tip, namely the NACA-LANGLEY LS(1)-0413MOD airfoil with a thickness of 13%c.

The airfoils were analysed for with XFOIL’s viscous solver for a Reynolds number of Re = 25 ⋅ 10 6 and a
Mach number of M = 0.55 . The Cp-distribution plots that were produced by XFOIL of the root and the
tip sections at their typical cruise angle of incidence (see section 4.4) are shown in Figures 4.1.2 and 4.1.3

18 Aerodynamics Chapter 4
Final Report of the International Design Synthesis Exercise 2005

Figure 4.1.2 Cp distribution of the LS(1)- Figure 4.1.3 Cp distribution of the LS(1)-
0417MOD airfoil at α = 2.65° and for 0413MOD airfoil at α = 1.43° and for
Re = 25 ⋅ 10 6 and M = 0.55 Re = 25 ⋅ 10 6 and M = 0.55

The airfoils have been analysed for a wide range of angle of attacks. The results are shown in Figure 4.1.4
below in the form of a cl , α -curve, a c d , α -curve and a c m , α -curve.

2,5
0,040
2,0
0,030
1,5
cl (-)

cd (-)

0,020
1,0

0,5 0,010

0,0 0,000
-5 0 5 10 15
-5 0 5 10 15
α (degrees)
α (degrees)

0
-5 0 5 10 15

-0,05
Root: LS(1)-0417MOD
cm (-)

-0,1 Tip: LS(1)-0413MOD

-0,15
α (degrees)

Figure 4.1.4 XFOIL results for the root and tip airfoil

From the cl , α -curve in Figure 4.1.4 it is possible to derive the lift curve slope, the maximum lift
coefficient and the zero-lift angle of attack. These parameters are listed in Table 4.1.2 below.

Table 4.1.2 Main parameters from the cl , α -curve

Chapter 4 Aerodynamics 19
Final Report of the International Design Synthesis Exercise 2005

Airfoil c l α (rad-1) c l max [-] α0


LS(1)-0417MOD (root) 9.00 2.35 -3.58
LS(1)-0413MOD (tip) 8.80 1.85 -4.02

In order to verify the XFOIL results a separate CFD analysis was carried out for the root section. This
method had many preparatory steps in order to be able to run a model for the airfoil under consideration.
The first step was the generation of the mesh file; a c mesh type was used with increased cell density
around the airfoil surface to increase the accuracy of the results.

The model analysed in Fluent used a Spalart-Allmaras turbulence model and was run for the
incompressible case as the Mach number is subsonic. This meant a segregated, steady solver was used.
The velocity inlet was specified with the various flow components in the x and y directions (at 2 degree
incidence x =179.89 m/s, y =6.282 m/s). Boundary conditions and operating conditions were taken from
various parameters during cruise and the solution run for 1000 iterations with a convergence criterion of
10-5.

The plot of pressure coefficient along the chord (Figure 4.1.5) produced a very similar plot to that
obtained using XFOIL (Figure 4.1.4). This plot has already been described and therefore the same
conclusions can be taken from it. The main difference was that the maximum pressure obtained using
XFOIL was higher than that of the CFD model, this could be due to the inaccuracy of the viscosity model
used in the XFOIL solver and differences due to the inputs available to the user. This therefore validates
the analysis of the airfoil using XFOIL. The contour plot of pressure coefficient over the airfoil shows a
favourable distribution without any pressure peaks.

Figure 4.1.5 Cp distribution over airfoil

20 Aerodynamics Chapter 4
Final Report of the International Design Synthesis Exercise 2005

Figure 4.1.6 Pressure distribution over airfoil Figure 4.1.7 Velocity distribution over airfoil

The velocity profile shows a favourable flow over the airfoil with little separation. The flow appears to be
quite well behaved over the surface with the flow remaining subsonic and hence no shocks are present as
would have been expected.

4.1.3 Wing lift


The results from XFOIL are 2D data and need to be corrected for 3D downwash effects. One of those
effects is the reduction in the lift curve slope. The three-dimensional lift curve slope for a straight wing in
incompressible flow can be determined using Torenbeek [37]:

cl α
C L wα = f (4.1.1)
E + cl α / πA

Here, f is a correction factor for wing taper, which can be assumed to be equal to 0.995 for 0.2 < λ < 1.0 .
Furthermore, E is Jone’s edge velocity factor, which is given by:


E =1+ (4.1.2)
A(1 + λ )

From this it follows that for the root section C L w α = 6.58 RAD-1.

Another 3D effect is the reduction in the maximum lift coefficient. A rough estimate is given in
Torenbeek [37]:

(cl max )root + (cl max )tip


C L max = 0.95 ⋅ (4.1.3)
2

Chapter 4 Aerodynamics 21
Final Report of the International Design Synthesis Exercise 2005

Using the data from Table 4.1.2 this gives C L max = 2.0 .

As the zero-lift angle of attack remains constant, the C L , α -curve for the wing can now be made and is

shown in Figure 4.1.8. A more accurate estimate for C L max will be determined in section 4.1.5.

CL,α-curve of the wing

2,5

1,5
CL (-)

0,5

0
-5 0 5 10 15 20 25
α (deg)

Figure 4.1.8 C L , α -curve of the wing

The C L required for cruise lies around 0.65. To achieve this the angle of incidence of the wing can thus
be set to 1.0 degrees. However this does not take into account any wing twist. This will be discussed in the
next two subsections.

4.1.4 Wing twist


From the airfoil analysis it was clear that for the root section to achieve a lift coefficient of 0.65, the airfoil
needed to be at an incidence of around 1.0 degrees. This value however does not take into account the 3D
effects associated with a wing. It is therefore obvious that the angle of incidence between the wing and the
fuselage will be greater than this value to produce the required lift coefficient. The reason for employing
twist along the wing is to prevent stalling on the outboard sections at higher angles of attack. Using the
following formula [37] (page 475) the angle of incidence at the root can be found as follows:

(
C L w = C L w α α r − α l 0 r − α 01ε t ) (4.1.4)

22 Aerodynamics Chapter 4
Final Report of the International Design Synthesis Exercise 2005

The lift curve slope for the wing is known to be 6.87 / rad and the required lift coefficient is 0.65. The
zero-lift angle of the root airfoil is equal to -3.725 degrees. α 0 is the zero – lift angle at angle of attack per
unit of twist and is found from Figure 4.1.9 to equal -0.25 for a wing with aspect ratio equal to 11 and
taper ratio equal to 0.36.

Figure 4.1.9 Zero-lift angle of attack per unit of twist for straight wings [37] (page 475)

This formula can be used iteratively to provide various combinations of root and tip incidence to give the
desired lift coefficient at cruise. The root incidence should not be any more than 3 degrees while there
should be no more than 5 degrees washout along the wing due to an increased induced drag penalty.
Using a root section incidence of 1 degree:

6.87π
CLw = CLwα ⎛⎜⎜⎝αr − (αl 0 )r − α 01εt ⎞⎟⎟⎠ = 0.65 = [1 − (− 3.725) − (− 0.25)εt ] (4.1.5)
180

This results in a tip incidence of εt = 2.78 0 which is positive which would not help the stalling properties

of the outer wing section. A root section of 2 degrees results in a tip incidence of εt = −1.22 0 . These
values seem acceptable and provide the required characteristics.

The same calculation for 3 degrees yields a value of εt = −5.22 which exceeds the 5 degrees of difference
requirement to prevent an added drag penalty. Therefore the incidence between the wing and fuselage was
selected to equal 2 degrees and the tip section placed at -1.22 degrees to help the stalling characteristics of
the outer wing.

Chapter 4 Aerodynamics 23
Final Report of the International Design Synthesis Exercise 2005

4.1.5 Spanwise lift distribution


Using the now available data it is possible to determine the spanwise lift distribution. For that purpose the
method outlined on page 473-474 in Torenbeek [37] will be used.

The spanwise lift can be divided into an additional lift distribution and a basic lift distribution:

cl = cl a + cl b (4.1.6)

where:

cg
cl a = C L La (4.1.7)
c

c g ε t cl a
cl b = Lb (4.1.8)
c E

Here, ε t is the aerodynamic twist of the wing tip, which is defined on page 440 of Torenbeek [37] as:

ε t = ε g t + α 0 root − α 0 tip (4.1.9)

Using the results from the airfoil analysis in section 4.1.2 this gives ε t = −1.22 − 3.58 + 4.02 = −0.78
degrees.

For a straight wing the non-dimensional parameters La and Lb are given by:

c 4
La = C1 + (C 2 + C 3 ) 1 −η 2 (4.1.10)
cg π

⎛ε ⎞
Lb = β ELa C 4 cos Λ β ⎜⎜ + α 01 ⎟⎟ (4.1.11)
⎝ εt ⎠

Here α 01 is related to La , and for a straight-tapered unswept wing with linear twist it is given by:

1 + 2λ 4
α 01 = −C1 − (C 2 + C 3 ) (4.1.12)
3(1 + λ ) 3π

The factors C1 , C 2 , C 3 and C 4 can be determined using figure E-5 on page 474 of Toorenbeek and
were found to be 0.5, 0.35, 0.35 and 0.5 respectively.

24 Aerodynamics Chapter 4
Final Report of the International Design Synthesis Exercise 2005

Finally the spanwise lift distribution can be determined by evaluating equations (4.1.8) and (4.1.9) using
the spanwise distributions of the wing chord, the twist and the wing thickness. Adding the two
contributions together results in the spanwise lift distribution as shown in Figure 4.1.10.

Spanwise lift distribution

CLcruise = 0.65 CLmax = 1.65 clmax local

2,5

1,5
q (N/m)

0,5

0
0 5 10 15 20 25 30
y (m)

Figure 4.1.10 Spanwise lift distribution

The blue line in Figure 4.1.10 represent the lift distribution during cruise. The orange line represents the
local maximum lift coefficient of the wing section. It decreases towards the tip because the airfoil used at
the tip has a smaller thickness. If the lift coefficient is increased to 1.65 (pink line) the local maximum lift
coefficient will be exceeded around y = 15 m and the wing will stall. The maximum lift coefficient is thus
equal to 1.65.

4.1.6 High-lift devices


High-lift devices are required for a high C L max during landing and takeoff in order to enable the aircraft
to operate from short unprepared runways. For that purpose the HERA makes use of flaps. Various flap
types were considered, including the single slotted flap, the double slotted flap with fixed vane, the double
slotted flap with variable geometry and the double slotted Fowler flap. A spreadsheet was made using the
method of Torenbeek [37]. The spreadsheet can be found on the CD in Appendix 7. The method will be
outlined below.

Chapter 4 Aerodynamics 25
Final Report of the International Design Synthesis Exercise 2005

In order to meet the takeoff and landing requirements, the flaps should give a high increment in C L max of
around 2.2. This maximum lift increment is given by the following equation:

S wf
∆ f CL max = 0.92∆ f C l max cos Λ 0.25 (4.1.13)
S

Here, the ratio of Swf/S is the ratio of the wing area affected by the flaps to the total wing area and is
calculated using:

Swf bfo − bfi ⎡ 1 − λ ⎛ bfo + bfi ⎞⎤


= 1+ ⎜1 − ⎟ (4.1.14)
S b ⎢⎣ 1 + λ ⎝ b ⎠⎥⎦

In this equation bfo and bfi relate to the flap span and are defined in figure G-13 of Torenbeek [37]. For

moderate angles of attack the two-dimensional lift increment ∆ f C l max can be estimated by the lift

increment at zero angle of attack:

c' ⎛ c' ⎞
∆fClo = ∆fC l 0 ' + C l 0 ⎜ − 1⎟ (4.1.15)
c ⎝c ⎠

Here, the ratio c' / c is the ratio of the extended chord length over the original chord length and can be
calculated by:

c' ∆c cf
= 1+ (4.1.16)
c cf c

Referring to equation (4.1.15) the two-dimensional lift coefficient at zero angle of attack, C l 0 , is taken

from the airfoil properties (see section 4.1.2) and is found to be 0.60. Furthermore, ∆ f C l 0 ' represents

the increment in the two-dimensional lift coefficient at zero angle of attack with the extended chord c'
and can be determined using:

∆ f C l 0 ' = 2πηδ α δ ' δ f (4.1.17)

Here, η δ is the flap effectiveness factor. This factor differs for each flap type and can be determined

using Torenbeek [37]. Furthermore, α δ ' is the theoretical flap lift factor and represents the rate of change

26 Aerodynamics Chapter 4
Final Report of the International Design Synthesis Exercise 2005

of the zero-lift angle with flap deflection. It is a function of the relative flap chord c f / c and can be

determined using figure G-2 of Torenbeek [37]. Finally, δ f represents the maximum flap deflection in

radians, which may differ for each airfoil.

In case of double-slotted flaps with variable geometry or double-slotted Fowler flaps, the lift increment
must be evaluated for each flap section separately using equation (4.1.15). The total lift increment is then
given by:

∆fCl 0 = ∆1Cl 0 + ∆ 2Cl 0 (4.1.18)

The analysis outlined above has been carried out for the four different flap types. The parameters used
and the results obtained are summed up in Table 4.1.3.

Table 4.1.3 Parameters used and results obtained for flap analysis
Double slotted Double slotted Fowler (double
Parameter Single slotted
(fixed) (variable) slotted)
0.75 (forward) 0.8 (forward)
ηδ 0.7 0.725
0.56 (aft) 0.56 (aft)
αδ ' 0.7 0.75 0.75 0.75
δf 30 (forward) 30 (forward)
30 50
30 (aft) 30 (aft)
∆c / c f 0.25 0.4 0.45 0.7
cf /c 0.35 0.4 0.4 0.4
c' / c 1.0875 1.16 1.18 1.28
Cl 0 0.65 0.65 0.65 0.65
∆ 1C l 0 - - 2.29 2.69
∆ 2 Cl 0 - - 1.74 1.94
∆fC l 0 1.81 3.55 4.03 4.63
S wf / S 0.64 0.64 0.64 0.64
∆ f CL max 1.06 2.08 2.36 2.71

Based on the results given in Table 4.1.3 the double slotted flaps with variable geometry were selected,
because of their high performance. Only the Fowler flap has a higher performance, but it also has some
major disadvantages due to its higher weight, complexity and cost. The double-slotted flap is not a
remarkable choice: it is used on many STOL cargo aircraft, such as the DHC-7 and the A400M.

The lift increment from the double slotted flaps is sufficiently high, so that leading edge high-lift devices,
such as slats and slots, are not needed. Using the value found in section 4.1.1 for the maximum lift
coefficient of the clean airfoil, the maximum lift coefficients and subsequently the stall speeds during

Chapter 4 Aerodynamics 27
Final Report of the International Design Synthesis Exercise 2005

takeoff and landing can obtained. These are listed in Table 4.1.4. For the landing a very flap angle of
20/20 degrees was assumed. This high flap angle could be made possible by to the fly-by-wire system.
Due to fly-by-wire it will generally be easier to perform short takeoff and landing operations.

Table 4.1.4 Maximum lift coefficients and stall speeds for cruise, takeoff and landing

Flap angle C L max Vstall


Cruise 0/0 deg 1.65 98.2
Takeoff 20/20 deg 3.2 44.8 *
Landing 30/30 deg 4.0 40.1 *
*) Stall speed at 85% gross weight

4.2 Aircraft drag breakdown

This section will use a method described by Torenbeek [37], page 148-152, to estimate the drag of the
aircraft. The drag of each component of the aircraft will be estimated by comparing it to the friction drag
of a flat plate having the same wetted area and length. It is expected that the CDO value for a large
turboprop aircraft will be in the region of 0.18 – 0.24.

4.2.1 Wing
The method used to estimate the drag produced by the wing is valid for t/c ratios below 0.2 and assumes
that the transition region occurs at around 10% of the chord. The drag is calculated using:

( CDS ) w = 0.0054rw⎛⎜⎝1+3(t / c)cos2 Λ.25 ⎞⎟⎠ S (4.2.1)

Where rw = 1 for cantilever type wings. This results in: ( CDS ) W = 1.746

4.2.2 Fuselage
The basic drag associated with the fuselage assumes a fully turbulent boundary layer for a Reynolds
number of 100 million based on the fuselage length. This fraction of the total drag neglects upsweep at
this stage, which will add a significant increment, this increment will be discussed later. The drag of the
fuselage is calculated using:

( CDS ) f = 0.0031rflf (bf + hf ) (4.2.2)

Where: r f is the shape factor and equals 1.1 for an aircraft with one side circular, the other rectangular
with some rounding, lf is the fuselage length which equals 40.22m, hf and bf are fuselage height and width

28 Aerodynamics Chapter 4
Final Report of the International Design Synthesis Exercise 2005

and equal 4.1 m and 4.88 m, respectively. This yields a value of 1.123 for the estimated drag of the
fuselage without upsweep effects.

4.2.3 Tailplane
The tailplane drag is estimated in the same way as the wing drag, it is known that this accounts for
approximately 24% of the fuselage plus wing drag and therefore a factor of 1.24 was selected for rt .

4.2.4 Engines
The drag due to the engines is related to the power in order to take into account the variation of size and
shape. The drag is therefore calculated using:

Pto
( CDS ) n = 0.1rn (4.2.3)
φto

Where: rn is the shape factor of the engine which equals 1.0, Pto is the take – off power equal to 21,000hp
and φto is the power / frontal area and equals 3554.5 hp/m². This yields a value of 0.5908 for ( CDS ) n .

4.2.5 Undercarriage
The drag of the undercarriage is found by approximating its contribution to the total drag based upon
values obtained experimentally for various configurations of the undercarriage. Obviously for a fully
retractable undercarriage this factor would simply be 1. The landing gear is semi retractable and retracts
into fairings similar to that found on the C-130 and C-5A. The value recommended for the drag factor for
such a configuration, ruc = 1.08 .

4.2.6 Other factors


These various contributions neglect interference effects, surface imperfections and the likes of aerials,
slots etc… These should not have a huge effect on the total drag as this is a large aircraft and are relatively
small with respect to the size of the aircraft. Hence a correction factor must be employed for the effect of
Reynolds number on turbulent skin friction drag and other drag contributors. This correction factor is
found by using:

−0.2
Actual zero lift drag coefficient ⎛ Vcrlf ⎞
rRE = = 47 Re f −0.2 = 47⎜ ⎟ = 0.9976 (4.2.4)
Uncorrected drag coefficient ⎝ υcr ⎠

2
Where Vcr = 167 m / s, lf = 40.22m and υcr = 2.893e −5 m s −1 at 8000m.

Chapter 4 Aerodynamics 29
Final Report of the International Design Synthesis Exercise 2005

4.2.7 Total drag


Using the following equation yields the total drag minus upsweep effects:

( CD ) S = rREruc[rt {(CDS )W + ( CDS ) f } + ( CDS ) n] = 4.4696


0 (4.2.5)

Dividing through by the wing area gives a value of CD 0 of 0.02087 .

4.2.8 Upsweep effects


As mentioned earlier, the upsweep of the rear fuselage section will result in a significant increase in the
drag of the aircraft. According to the CAD model the upsweep angle is 10°, using figure 4.2.5, the frontal
drag increment due to upsweep at ~ 1° incidence is 0.02. This equates to a drag increment of:

⎛ ∆CDf (bf × hf ) ⎞
( CDS ) upsweep = ⎜ ⎟1.868e −3 (4.2.6)
⎝ S ⎠

Adding this value to the total drag gives a value of CD 0 = 0.02274 .

Figuur 4.2.1 Estimated fuselage drag increment due to rear upsweep and angle of attack [37] (page 505)

Carrying out the same procedure for the Hercules yielded a basic value of 0.02076 neglecting upsweep and
therefore verifies the value obtained for the HERA.

For take-off and landing, various increments must be added due to deployment of landing gear and flaps.
These increments have been estimated by Roskam [2] (127) and are shown in table 4.2.1.

30 Aerodynamics Chapter 4
Final Report of the International Design Synthesis Exercise 2005

Table 4.2.1 Increments for the drag polar for landing and takeoff configurations
Configuration ∆CD0 e
Take-0ff 0.02 0.8
Landing gear 0.025 0.83
Landing 0.075 0.75

1
This yields the following parabolic drag equations using the fact that K = :
πAe
For the clean aircraft: CD = 0.0227 + 0.0349CL2

During take-off: CD = 0.0677 + 0.0362CL2

During landing: CD = 0.1227 + 0.0386CL2

4.3 Aircraft aerodynamic centre

4.2.9 Aerodynamic centre of the wing


For a wing with zero sweep, the location of the aerodynamic centre is located at the mean quarter chord
point, that is the quarter chord point of the mean aerodynamic chord. This means the location of the
mean aerodynamic chord must be known in order to find the location of the aerodynamic centre.

For a straight tapered wing the lateral coordinate of the MAC can be found using:

b ⎛ 1 + 2λ ⎞
y= ⎜ ⎟ = 10.223m (4.3.1)
2 ⎝ 3(1 + λ) ⎠

Due to the geometry of the wing, it is obvious that the longitudinal coordinate of the aerodynamic centre
will have the same coordinate as the quarter chord point of the root quarter chord point. This is equal to
1.622 m and therefore the location of the wing aerodynamic centre is now known, as shown below.

Figure 4.3.1 Location of the mean aerodynamic chord and aerodynamic centre of wing and wing / body

Chapter 4 Aerodynamics 31
Final Report of the International Design Synthesis Exercise 2005

The location of the wing aerodynamic chord is defined as being the location of the aerodynamic centre
relative to the leading edge of the MAC. This is simply the quarter chord point of the MAC, which is
equal to 1.846 m and therefore: (xac / c )w = 0.25 .

4.2.10 Aerodynamic centre of wing and body


The aerodynamic centre of the wing and body is sensitive to the pressure distribution around it and not
the lift force and therefore it is difficult to predict an exact location. Correction factors will be included to
take into account body effect on the location of the aerodynamic centre. This method is outlined by
Torenbeek [37] (pages 476-481).

The first correction factor is due to the forward shift of the aerodynamic centre due to the fuselage
sections both aft and forward of the wing. The main contribution to this shift is the fuselage nose.

⎛ ∆f 1 xac ⎞ 1.8 bfhflfn


⎜ ⎟=− = −0.066 (4.3.2)
⎝ c ⎠ ( CLα ) wf S c

Hence it can be seen that the fuselage alone shifts the aerodynamic centre forward by 26.4%.

The second correction factor takes into account the additional lift created by the propeller, which causes a
shift in the location in the aerodynamic centre. This effect is significant as the props are mounted to the
wing and therefore they will move the aerodynamic centre further forward.

xac BpDp 2 lp
∆p = − 0 .5 ∑ = −0.00265 (4.3.3)
c S c( CLα ) wf

Where Bp is the number of blades, Dp is the propeller diameter and lp is the distance of the propeller from
the ¼ chord line. This is for one prop alone, so the overall effect is equal to:

xac
∆p = −0.00265 × 4 = −0.0106 (4.3.4)
c

Therefore the propellers have the effect of shifting the aerodynamic centre forward by a further 4.24%.
This results in the aerodynamic centre being located at:

⎛ xac ⎞ ⎛ xac ⎞ ⎛ ∆f 1 xac ⎞ ⎛ ∆pxac ⎞


⎜ ⎟wing / body = ⎜ ⎟w + ⎜ ⎟+⎜ ⎟ = 0.25 − 0.066 − 0.0106 = 0.1734 (4.3.5)
⎝ c ⎠ ⎝ c ⎠ ⎝ c ⎠ ⎝ c ⎠

32 Aerodynamics Chapter 4
Final Report of the International Design Synthesis Exercise 2005

4.2.11 Pitching moment


From the analysis on the airfoils, it was found that the pitching moment coefficient at cruise was equal to -
0.119 for the LS(1)-0417MOD and -0.114 for the LS(1)-0413MOD. The pitching moment coefficient of
the wing is given by:

2 b/2
( Cm ) wing =
ac
Sc
∫0
cmacc 2 dy (4.3.6)

This integral must be solved numerically using Simpson’s rule which states:

2 b/2 2 1
( Cm ) wing =
ac ∫ cmacc dy =
2
h[ y 0 + 4 y1 + yn] (4.3.7)
Sc 0 Sc 3

Where h is the step size and is equal to 12.1325.

Table 4.3.1 Required values for the use of Simpson’s rule


cr(y0) MGC (y1) ct (yn)
y location along span 0 12.1325 24.265
cmac -0.119 -0.117 -0.114
c² 42.09 19.466 5.476
cmacc 2 -5.008 -2.2775 -0.6242

Therefore the pitching moment of the wing can now be calculated.

2 1
( Cm ) wing =
ac h[ y 0 + 4 y1 + yn] = −0.1175 (4.3.8)
Sc 3

Although twist will also contribute towards the overall pitching moment, the significance is negligible in
the case of a straight wing as the ¼ chord sweep is equal to zero.

The contribution of the fuselage to the overall pitching moment coefficient is given by Munk’s theory [37]
(page 480), which states:

⎛ 2.5bf ⎞ π
bfhflf CL 0
∆fCmac = −1.8⎜1 − ⎟ = −0.0613 (4.3.9)
⎝ lf ⎠ 4 S c ( CLα ) wf

Therefore the pitching moment coefficient for the wing and fuselage is equal to:

Chapter 4 Aerodynamics 33
Final Report of the International Design Synthesis Exercise 2005

( Cm ) wf = ( Cm ) w + ∆fCm = −0.1805
ac ac ac (4.3.10)
This value is higher than the value obtained in chapter 10 as it does not take into account the contribution
to the overall pitching moment due to the engines as their contribution is very difficult to predict. The
analysis of the pitching moment with the tail is a stability issue and will be discussed in chapter 10.

4.4 Aircraft lift

Having already obtained the lift produced by the wing, the effect of the fuselage must now be taken into
account. This method for estimating the lift of the aircraft is that used by Torenbeek [1] (479 -482). There
are four main interference effects of the fuselage on the lift of the aircraft. The first is that the lift on the
fuselage nose is counteracted only partly by the download on the rear fuselage as shown in figure 4.4.1.
The second is the increased effective incidence of the wing sections near the wing / fuselage interface due
to the flow component normal to the fuselage (crossflow), this is identified in figure 4.4.1. Wing – lift
carry over is another factor, although the lift produced by the fuselage is less than the lift produced by an
equivalent wing section in place of the fuselage, there is considerable lift produced by this section as
shown in figure 4.4.1. The vertical displacement of the wing with respect to the fuselage has the effect of
altering the flow pattern and for a high wing design this leads to a reduction of lift.

Figure 4.4.1 Fuselage / wing interference effects on the spanwise lift distribution [37] (page 478)

These effects can be added together to give the total wing / fuselage lift coefficient as follows:

⎡ KII ⎤
CLwf = ( CLα )⎢(α
f − α 01) + { iw − ( αlor )}tε⎥ + ∆zCL (4.4.1)
⎣ KI ⎦

34 Aerodynamics Chapter 4
Final Report of the International Design Synthesis Exercise 2005

The various angles are illustrated below in figure 4.4.2. The factors KI and KII are factors for calculating
the lift on the wing plus the body.

Figure 4.4.2 Geometric definitions of the various wing / body angles [37] (page 478)

KI can be found using the following formula:

⎛ bf ⎞ Snet π bf 2
KI = ⎜1 + 2.15 ⎟ + (4.4.2)
⎝ b ⎠ S 2CLwα S

Here, Snet is the exposed wing area outside of the fuselage section and is found using simple geometry to
equal:

Snet = 181.45m 2 (4.4.3)

Hence KI is equal to:

⎛ bf ⎞ Snet π bf 2

KI = 1 + 2.15 ⎟ + = 1.056 (4.4.4)
⎝ b⎠ S 2CLwα S

( CLα ) wf is the lift curve slope of the fuselage and wing and is calculated using the factor KI as follows:

( CLα ) wf = KICL = 1.056 × 6.87 = 7.256rad −1


wα (4.4.5)

The factor KII is equal to:

⎛ bf ⎞ Snet
KII = ⎜1 + 0.7 ⎟ = 0.907 (4.4.6)
⎝ b⎠ S

The last parameter to be found is ∆zCL . For a high wing design: ∆zCLS / crbf = −0.1 , hence:

Chapter 4 Aerodynamics 35
Final Report of the International Design Synthesis Exercise 2005

crbf
∆zCL = −0.1 = −0.0148 (4.4.7)
S
Knowing these parameters allows the lift coefficient of the wing and fuselage to be found:

⎡ KII ⎤
CLwf = ( CLα )⎢(α
f − α 01) + {iw − ( αlor )}tε⎥ + ∆zCL = 0.569 (4.4.8)
⎣ KI ⎦

It is evident that this value is lower than the required lift coefficient of 0.65 for cruise and therefore the
aircraft will have to fly at incidence during cruise. An iterative process results in a value of 0.637 degrees
giving the required lift coefficient of 0.65 for cruise. However this neglects the impact of the tailplane,
which will further reduce the lift of the aircraft.

The overall lift coefficient of the aircraft can be calculated including tailplane effects using the following
equation which assumes that the aircraft is flying at 0.637 degrees incidence and therefore the wing /
fuselage coefficient is equal to 0.65.


⎜ xcg − xac ⎞⎟ ⎞⎟ c


CL = C 1+
Lwf ⎜ ⎜

⎟⎟ + Cmac = 0.599 (4.4.9)

⎝ ⎝ lh ⎟⎠ ⎟⎠ lh

It can be seen that the tailplane has a negative increment on the lift coefficient of the aircraft of
approximately -0.05.

The individual tailplane lift coefficient increment can be calculated using:

Sh qh c ⎛ ⎜ xcg − xac ⎟ ⎞
⎛ ⎞
∆ h CL = CL h = ⎜Cmac + CL⎜⎜ ⎟ ⎟ = −0.0502
S q lh ⎝ ⎝ c ⎟⎠ ⎠ (4.4.10)

From the stability section it is know that the tailplane has a lift coefficient of -0.3. This can be validated
using the above equation as follows:

Sh qh
∆ h CL = CL h = −0.0502 (4.4.11)
S q

This results in the airspeed over the tailplane being equal to 124.2 m/s, hence slowing from the 167 m/s
that occurred over the wing. This seems reasonable, as turbulence effects will reduce the energy of the
flow and therefore the airspeed. Hence the incremental lift form the tailplane has been validated by the
tailplane lift coefficient obtained in the stability section in chapter 10.

36 Aerodynamics Chapter 4
Final Report of the International Design Synthesis Exercise 2005

A lift coefficient of 0.599 was obtained for the aircraft flying at 0.637 degrees with a body minus tail lift
coefficient of 0.65. Using equation 4.4.8 and an iterative method the required angle at cruise can be found
to produce the required lift coefficient of 0.65. This was found to equal 1.05 degrees.

Earlier the incidence of the wing relative to the fuselage was selected based on an iterative method, which
included some basic requirements. Now that more information is available about the lift of the aircraft,
this value can be checked using the following formula at zero incidence during cruise:

CLwf − ∆zCL KI
iw = + α 01tε+ ( α
lo ) r = 2
0
(4.4.12)
KIICLWα KII

Therefore this validates the choice of the setting angle earlier of 2 degrees and the incidence of -1.22
degrees at the tip section.

Chapter 4 Aerodynamics 37
Final Report of the International Design Synthesis Exercise 2005

5 Power Plant

One of the particularities of the HERA design are the engines: they are the most powerful turboprops
currently available in the western world. This chapter will present the power plant analysis. First, section
5.1 will discuss the engine choice. Subsequently, section 5.2 will present an extensive propeller analysis. In
section 5.3 the location of the engine will be determined. Finally, the noise and emission characteristics of
the engine will be discussed in sections 5.4 and 5.5, respectively.

5.1 Engine

The engine configuration for this particular aircraft design is a simple one, consisting of four turboprop
engines mounted under the wings of the aircraft. A turboprop configuration was selected for a number of
reasons. The turboprop engines were seen to be more efficient than the turbofan engines, at the required
cruise speed and altitude. The turboprop engine configuration was also the cheapest of all of the
configurations considered; more information on the selection of the engine configuration can be found in
reference [8], page 70. With this configuration the Rolls Royce Ae1107c engine was selected as the most
suitable to power the aircraft, it was seen to have a number of advantages over the other turboprops
considered. The Ae1107c is currently in production for the V-22 Osprey, unlike the TP400, which is
currently in development for the Airbus A400M and therefore may not be ready in time for the proposed
date for the first flight. The Ae1107c also has many features which reduce the amount of maintenance
needed to be carried out on it while in service. The Ae1107c is based on the same core engine design as
the Ae2100 currently in service on the C-130J, which is a similar type of aircraft to this design, but much
smaller, see Appendix 2. This means that the technology behind the Ae1107c has been tried and tested in
all of the conditions that the HERA aircraft will be required to operate in. The actual cost of the Ae1107c
engine also compared very favourably to its rival, the TP400, costing approximately $2.3million to the
TP400’s $4million. Some basic information on the Ae1107c is shown in table 5.1.1 below.

Table 5.1.1 Basic Information on the Ae1107c Engine


Max Shaft Power 4586 kW Basic Weight 1158 kg Length (with gearbox) 3.29m
Pressure Ratio 16.7 Compressor 14HP Height (with gearbox) 0.91m
Length (core) 1.98 m Turbine 2HP, 2PT Width (with gearbox) 0.87m
Diameter (core) 0.87 m SFC (Cf) 72 µg/J

Due to the nature of the current application of the Ae1107c engine, on a high-tech military aircraft, very
little information can be obtained on both the dimensions and the performance of the engine. Due to the
nature of the V-22 the Ae1107c hasn’t got a normal gearbox, like that of the Ae2100. The power from a

Chapter 5 Power Plant 39


Final Report of the International Design Synthesis Exercise 2005

single engine is distributed across the aircraft to both of the sets of rotors using a unique system. It has
therefore been assumed that Rolls Royce will be capable of providing a standard gearbox for this engine.
The dimensions of the engine including the gearbox have been found by comparing the Ae1107c engine
to other turboprop engines with the gearbox present (such as the Ae2100 or the PW150). Upon such a
comparison it can be seen that the actual engine core accounts for approximately 60% of the length of
the engine and 95% of the overall height. It can be seen that generally the width is fairly constant. When
these ratios are used in conjunction with the information provided by the manufacturer [52], the
approximate dimensions including the gearbox, can be found.

The approximate dimensions for the engine including the gearbox can then be used to compare the
engine to the likes of the Ae2100. With the percentage differences in size between the two engines and
with the known dimensions of a nacelle for the Ae2100, an approximation of the size of nacelle needed to
house the Ae1107c can be made. For example the overall length of the Ae1107c is 8.95% larger than that
of the Ae2100, thus the nacelle for the Ae1107c can be approximated to also be 8.95% longer. The final
dimensions of the engine nacelle (shown in figure 5.1.1) can be seen to be similar to that of the nacelle on
the V-22, thus to some degree, validating the results.

Figure 5.1.1 Basic Dimensions of the Engine Nacelles

5.2 Propeller

5.2.1 Assumptions Made in Preliminary Design


During the preliminary design phase, see [8], a number of assumptions were made about the performance
of the propellers on the aircraft. These assumptions were integral to the results obtained from
performance constraints analysis using the Howe Spreadsheet. The main assumption made regarded the
advance ratio of the propellers. The advance ratio is a measure of the distance advanced by the propeller
in the direction of flight after one revolution, and is therefore a function of the forward velocity of the

40 Power Plant Chapter 5


Final Report of the International Design Synthesis Exercise 2005

aircraft, the number of revolutions per second and the diameter of the propeller. Equation 5.2.1 below
shows the calculation of advance ratio.

V0
J= (5.2.1)
np D

In the Howe spreadsheet method it was required that a value be entered for the denominator part of this
equation (nd). This value was taken as 85, which was based on information provided by Denis Howe in
his textbook, for a similarly sized transport aircraft. Thus one of the purposes of this portion of the
design is to validate that this assumption was correct and that a propeller designed to this criteria would
indeed be capable of providing the aircraft with sufficient thrust.

5.2.2 Propeller Design Method


The propeller will be designed to fulfil a number of different conditions. As mentioned previously the
multiple of the number of revolutions made by the propeller per second and the diameter of the propeller
must be approximately equal to the constant value used in the Howe spreadsheet during the preliminary
design stage. More importantly the propeller must be capable of providing an adequate amount of thrust
for cruise and take-off. The power that the propeller requires from the engine in order to provide this
thrust must not be greater than that which the engine is capable of providing in that particular segment of
the flight envelope.

The method used to design the propeller is based on the blade element theory as described on page 541 of
reference [21] and is briefly outlined in Appendix 8. Using this theory it was possible to construct a
MATLAB program to design a simple propeller. This program can be found on the CD in Appendix 7.

The main reason for using the blade element theory is that it provides a quick method for both the design
and indeed for the analysis of the propeller. Where the blade element theory used in the MATLAB code
can produce results in a few seconds, the more complex methods employed by industry would take many
hours for a single design cycle to complete. Using this method it was easy to run the program, check the
results against the requirements, then modify the input parameters as required and run the program again.
For example if too great a thrust was produced at cruise conditions, the diameter of the blades could be
reduced accordingly. In this manner a propeller blade could be designed to meet the criteria in a matter of
minutes. The same propeller blade design could then be entered into a second MATLAB program to carry
out the basic analysis of the blade at varying changes in pitch angle and at varying aircraft velocities. Once
again this would be carried out at a much faster rate than if computational fluid dynamics were to be used.

Chapter 5 Power Plant 41


Final Report of the International Design Synthesis Exercise 2005

The MATLAB program and the blade element theory itself do have a number of drawbacks which make
it less accurate compared to the more advanced CFD design techniques. Firstly the blade element theory
assumes that each of the blade elements considered in the analysis is not affected by the elements around
it, essentially each element is considered to be in a 2 dimensional flow. This will produce errors, as in a
real flow there will be components along the blade length. The second problem with the blade element
model is that no consideration is made of the affect of the propeller hub and spinner on the flow around
the blades, especially at the blade root. Using computational fluid dynamics the propeller blade can be
considered as a fully 3 dimensional body in a fully 3 dimensional flow, the propeller hub can also be
modelled and its affects on the blades can be taken into account when measuring the performance.

The airfoils used in the propeller design were from the HS1 series of airfoils, developed by Hamilton
Standard in 1981. Although this series of airfoils is slightly older, than for example the ARA-D series, it
was possible to obtain real experimental results from wind tunnel tests for the HS1-606 section. This
allowed a comparison between the experimental results and results for the same conditions obtained using
X-Foil. The X-foil results could therefore be scaled accordingly for the other airfoils in the series, and thus
databases of lift and drag coefficients could be obtained for all of the airfoils. It is these databases that the
MATLAB design and analysis programs use. However, this approach introduces a number of problems.
Firstly, because data is only present for a set airfoil section, it is impossible to gradually change the shape
of the airfoil sections along the length of the blade. This means that there are large “jumps” in the profile
of the blade. For example the blade may begin with an airfoil thickness of 20% and it will keep this
thickness for a certain distance along the blade and then suddenly jump to an airfoil of 12% thickness.
The second problem with this method is that the program is limited by the databases in so far as it is not
able to use a lift or drag coefficient which falls outside the limits of the database. If the database goes from
angles of attack of -18˚ to 26˚, the program does not have access to data for an angle of attack of 28˚. The
same is true for the Mach number of the flow. Also because of the nature of the databases, it is impossible
to include all variations of angle of attack and Mach number. So to find the values the program must
interpolate between those given in the database. This is yet another source for error. Once again the use of
CFD techniques would eliminate all of these errors: it would allow for a smooth transition of airfoil shape
along the blade and there would be no need for a database of lift and drag coefficients, as they would be
calculated directly. Due to the errors present in the design program, the final propeller design will be such
as to provide slightly more thrust than required for each stage of flight. Providing more thrust than is
required also gives the aircraft a certain amount of buffer if the weight of the aircraft was to increase
during further development.

5.2.3 Results
Using the propeller design program it was possible to design a propeller that would produce a thrust
slightly greater than that required at cruise. To do this, the conditions that were entered into the design

42 Power Plant Chapter 5


Final Report of the International Design Synthesis Exercise 2005

program were the same as the conditions that the aircraft would be expected to be under at cruise. The
forward velocity of the aircraft was assumed to be 181m/s, which is equal to the maximum cruise velocity
defined in the requirements. The altitude of the cruise was also used to define the atmospheric density and
temperature, of 234.2K and 0.507kg/m3 respectively. Using this data and by varying the number of blades
on the propeller, the diameter of the blades, the amount of sweep to the leading edge of the blades a
thrust was obtained. The amount of thrust generated was greater than that required to maintain cruise
conditions and also the shaft power that was required to turn the propeller was within the maximum shaft
power that the engine can produce at the cruise altitude. The final data output from the propeller design
program is shown below in table 5.2.1. Not shown in the table is the geometric data that was also an
output of the program that was later used to perform a more detailed analysis of the propeller.

Table 5.2.1 Basic Outputs from Propeller Design


Parameter Output Parameter Output Parameter Output
Thrust (kN) 15.7 Diameter (m) 4.4 Advance Ratio 2.1
Torque (kNm) 26.3 nd 86.2 Tip Mach No. 1.062
Power (kW) 3242.9 Velocity (m/s) 181 Propeller Efficiency 0.877
No. of Blades 6 Diameter of Spinner 0.79m

The geometric data that was outputted from the design program consists of a number of different things
such as the angle of each blade element to the plane of rotation, the chord length of each element and the
amount of sweep each element has. Sweep was only applied to the outer elements of the blade, where the
velocities approached the speed of sound. The angle of the blade elements was high close to the propeller
hub (about 80˚) and decreased along the blade length (to about 35˚). The chord of the propeller was
assumed to be constant for a large proportion of the blade with an elliptical tip (similar to that in figure 6-
11 on pg202 of reference [37]). The size of the spinner was taken to be approximately 0.16 times the blade
diameter; this figure came from looking at existing propellers. Using all of this geometric data, the
propeller analysis program can calculate the efficiency of the propeller at different aircraft velocities and
can also analyse the affect that changing the pitch of the propeller has on the efficiency. The main output
of running this program is the graph of propeller efficiencies against advance ratio, shown in figure 5.2.1
below. In this case the advance ratio only changes because of the forward velocity of the aircraft, the
diameter and the speed of rotation stay constant.

Each of the smaller curves represents a change in pitch of the propeller, it can be seen that by employing
the propeller designed previously as a variable-pitch propeller on this aircraft, a high propulsive efficiency
can be obtained over a fairly large range of advance ratios. The analysis can be repeated for differing
altitudes and new curves of propeller efficiency can be obtained. The overall curve of highest efficiency
for each pitch change (shown as the line joining the top of each of the smaller curves in figure 5.2.1) will
later be used in the performance analysis of the aircraft.

Chapter 5 Power Plant 43


Final Report of the International Design Synthesis Exercise 2005

Figure 5.2.1 Change of Propeller Efficiency with Change in Pitch

A third smaller program calculates the amount of thrust that the propeller blades can provide on takeoff.
This is achieved by changing the angle of the blade until the desired amount of thrust can be generated.
As with the analysis program it relies on the databases of lift and drag coefficients, and due to the very
large angles of attack of the blades the results will not be exact. Upon executing this program it can be
seen that the propeller will produce approximately 106.3kN of thrust per engine, with a shaft power of
3776.1kW required. A total thrust of 425.2kN can therefore be achieved at takeoff. This thrust is slightly
more than that required for the aircraft to takeoff, once again this is to prevent the complete redesign of
the propeller if the weight of the aircraft were to increase. The final propeller design which will meet all of
the necessary requirements is shown in figure 5.2.2 below, it must be noted that only the chord lines of
each of the blade elements making up the blades are drawn, the real blade will have airfoils going from
20% thickness at the root to 6% thickness at the blade tip.

Figure 5.2.2 Final Propeller Design

44 Power Plant Chapter 5


Final Report of the International Design Synthesis Exercise 2005

5.3 Engine locations

The position of the engines was determined by a combination of the basic principles set out on pg204 of
Torenbeek [37] (which uses the FAR regulations) and from inspection of other aircraft of a similar
configuration. The basic factors which define the location of the engines on the aircraft are the noise and
vibrations produced by the engines, the consideration of the safety of the passengers and crew if a
propeller blade was to break of and finally the distance from the propeller tip to the ground during
takeoff, landing and taxi. As mentioned previously the engines will be mounted, under the wings, with two
on either side. All of the distances will be taken from the centre line of the. The actual diameter of the
propeller will be taken as being 5m. This is slightly larger than the propeller designed previously, but by
allowing for a larger propeller it makes the aircraft more future proof if there is a need for a major refit, if
a different propeller needs to be used or a more powerful engine. To limit cabin noise and vibrations from
the engines it is necessary that the tip of the propeller be 10cm plus another 1.65cm for every 100hp the
engine can produce. This means a total of approximately 1m from the fuselage, and puts the inboard
engine approximately 6m from the centre line of the aircraft. The FAR regulations quote a minimum
distance of 0.23m between two props, however upon examining existing aircraft, such as the C-130 this
was seen to be approximately 2m. This places the outboard engine 13m from the aircraft centre line.

5.4 Noise suppression

The levels of noise both inside and outside a propeller driven aircraft can be quite high and in many cases
to provide the passengers and crew with an improved degree of comfort the level of noise will have to be
reduced. The main source of noise on this type of aircraft is the propeller, which can cause vibrations in
the airframe which travel into the cabin or, in the case of external noise, from the rotation of the
propeller. There are many different techniques which can be used to reduce the amount of noise from
simple passive noise reduction techniques to more active reduction techniques. The simplest and cheapest
way to reduce noise is through passive means, this can include the use of insulation in areas particularly
affected by noise, or through modifications to the engine exhaust and the actual propeller shape. Many
modern propellers have scimitar shaped blades to reduce the amount of noise they produce. Active noise
cancellation can also come in a variety of forms; however the methods which can be employed are more
complicated than those used in passive control. The simplest is the creation of anti-noise. This requires
microphones to be placed in the aircraft which pick up any noise produced, a computer analyses the noise
and produces sound waves which are equal and opposite, thus cancelling out the noise. This technique is
most affective when used in headsets. It can be used in large spaces, like a cabin, but requires large
speakers and is less effective. Another method which can be used is the “tuning” of an engine-propeller to
cancel out the noise from a neighbouring engine-propeller. This involves adjusting the propeller phasing,
which can be preset by the manufacturer or can be achieved actively during flight by a computer system.
A system similar to this is currently in development for the C-130 (reference [20]). Another active noise

Chapter 5 Power Plant 45


Final Report of the International Design Synthesis Exercise 2005

control technique is the use of a noise and vibration suppression system (NVS), similar to that used on the
Q400 (reference [44]). The noise inside the fuselage is caused by pulses of air hitting the side of the
fuselage from the turning propeller. The NVS system uses microphones coupled with active vibration
absorbers that produce counter vibrations and reduce internal noise. With the exception of the production
of anti-noise the active noise reduction techniques are very expensive. This means that their use in this
particular design will be very limited. Instead noise will be reduced through use of insulation and for the
crew the anti-noise producing headsets could be purchased. The reduction of noise for those on the
ground can be achieved through changes to the propeller design; the Saab 2000 (reference [54]) turboprop
aircraft only produces a maximum of 87.9 dB, by using scimitar propellers.

5.5 Emissions

It was previously mentioned that due to the nature of the current application of the Ae1107c engine, in
top secret military aircraft, little information is available on the emissions that the engine will produce. The
Ae1107c comes from a family of engines which share the same core design, the Ae3007 shares this engine
core and data on its emissions is available. This data can therefore be used to provide some indication of
the emissions of the Ae1107c. The available data on engine emissions is summarised in table 5.5.1 below
and was obtained from reference [48].

Table 5.5.1 Table of Emissions for the Ae3007 Engine


Parameter Result Parameter Result
Hydrocarbon emission index (T/O) 0.24 g/kg Carbon Monoxide emission (T/O) 0.87 g/kg
Hydrocarbon emission index (idle) 4.83 g/kg Carbon Monoxide emission (idle) 26.04 g/kg
Oxides of Nitrogen emissions (T/O) 17.9 g/kg Average Oxides of Nitrogen 10.95 g/kN
Oxides of Nitrogen emissions (idle) 3.37 g/kg Fuel Used Jet - A

Upon comparison of this engine with other similar sized engines it can be seen that the Ae3007 series of
engines holds up very well in terms of the amount of pollutants produced. In some cases the amounts of
carbon monoxide and NOx produced by the engine is much less than that which rival engines produce.
This means that the Ae3007 and thus the Ae1107c is a relatively efficient and environmentally friendly
engine.

46 Power Plant Chapter 5


Final Report of the International Design Synthesis Exercise 2005

6 Performance Analysis

The (economic) value of an aircraft is determined largely by its performance. This chapter will present a
performance analysis. First, in section 6.1 a constraints analysis will be carried out yielding the power
required to perform all flight operations. It will be shown that the AE1107c engines are capable of
delivering this power. Subsequently, in section 6.2 the climb performance will be analysed. The load factor
envelope will be presented in section 6.3. The flight envelope is the topic of section 6.4. Finally, section
6.5 will present the payload-range diagram of the HERA.

6.1 Constraints analysis

In the subsequent subsections the relation between the power loading P/W and wing loading W/S will be
discussed for stall, takeoff distance, cruise/maximum speed and climb. A special spreadsheet has been
made for this purpose, which can be found on the CD in Appendix 7.

6.1.1 Stall speed


In the requirements (see appendix 1) the landing field distance has been specified: s LFL = 800 m at 85%

gross weight. During a later requirements review the requirement was set to: s LFL = 950 m at 85% gross
weight. Roskam [29] suggests the following relation between the landing field distance and the approach
speed for FAR-25:

s LFL = 0.3Va2 (6.1.1)

Here s LFL is in ft and Va is in kts. In SI-units equation (6.1.1) becomes:

s LFL = 0.3456Va2 (6.1.2)

The approach speed Va is always defined in terms of the stall speed Vs :

Va = 1.3Vs (6.1.3)

Combining equations (6.1.2) and (6.1.3) yields:

s LFL = 0.584Vs2 (6.1.4)

Chapter 6 Performance Analysis 47


Final Report of the International Design Synthesis Exercise 2005

The maximum allowed stall speed Vs , max is thus given by:

s LFL
Vs ,max = (6.1.5)
0.584

For s LFL = 950 m this gives a maximum allowed stall speed of Vs , max = 40.1 m/s. From Table 4.1.4 it

follows that the actual stall speed during landing is exactly equal to Vs = 40.1 m/s. Therefore the
maximum stall speed requirement is just satisfied.

At this speed the lift L must still equal to landing weight WL. Hence:

1
WL = ρVs2,max C Lmax L S (6.1.6)
2

Combining this expression with equation (6.1.5), rearranging and translating to takeoff conditions finally
gives:

⎛W ⎞ 1 ⎛ s ⎞ ⎛W ⎞
⎜ ⎟ = ρ 0 ⎜ LFL ⎟C Lmax L ⎜⎜ MTOW ⎟⎟ (6.1.7)
⎝ S ⎠TO 2 ⎝ 0.584 ⎠ ⎝ WL ⎠

Here (WL WMTOW ) = 0.85 . Equation (6.1.7) will produce a vertical line in the parametric sizing chart.

6.1.2 Takeoff Distance


In the requirements (see appendix 1) the takeoff distance has been specified: sTOFL = 800 m at 85% gross

weight. During a later requirements review the requirement was reset to: sTOFL = 950 m at 85% gross
weight.

Using Roskam [29] the takeoff distance sTOFL is proportional to the takeoff wing loading (W S )TO and

the takeoff thrust-to-weight ratio (T W )TO :

⎧⎪ 0.85 ⋅ (W S ) ⎫⎪
sTOFL = 0.239⎨ TO
⎬ (6.1.8)
⎪⎩σC LmaxTO η p (T W )TO ⎪⎭

48 Performance Analysis Chapter 6


Final Report of the International Design Synthesis Exercise 2005

Here, σ denotes the density ratio, which for takeoff is simply equal to 1, and η p is the propulsive

efficiency. Rearranging gives:

⎧⎪ 0.85 ⋅ (W S ) ⎫⎪
(T W )TO = 0.239⎨ TO
⎬ (6.1.9)
σC η s
⎪⎩ LmaxTO p TOFL ⎪⎭

The thrust-to-weight applies to turbojet and turbofan airplanes. In case of a propeller airplane (turboprop)
the performance constraints are usually expressed in terms of the power-to-weight ratio: P/W. For that
purpose equation (6.1.9) must be multiplied with the takeoff safety speed V2, which for a four-engined
aircraft is specified as:

V2 = 1.15Vs (6.1.10)

Given from Table 4.1.4 that Vs TO = 44.8 m/s this gives V2 = 51.6 m/s. The power-weight-ratio
required for takeoff is thus given by:

⎛P⎞ 0.239V2 ⎧⎪ 0.85 ⋅ (W S ) ⎫⎪


⎜ ⎟ = ⎨
TO
⎬ (6.1.11)
⎝ W ⎠TO ηp ⎪⎩ σC LmaxTO sTOFL ⎪⎭

6.1.3 Cruise and maximum speed


In the requirements (see appendix 1) the cruise speed and maximum speed have been specified:
Vcr = 600 km/hr and Vmax = 700 km/hr. During a customer requirements review the requirement was

lowered to Vcr = 600 km/hr and Vmax = 700 km/hr. During cruise the following equations are
simultaneously satisfied:

1
L =W = ρV 2 SC L = qSC L (6.1.12)
2

1
T =D= ρV 2 SC D = qSC D (6.1.13)
2

Assuming a parabolic drag polar, the thrust-to-weight ratio can be derived, yielding:

⎛T ⎞ 1 ⎡ qC D0 (W / S )cr ⎤⎛⎜ WTO ⎞


⎟⎟
⎜ ⎟ = τ tf ⎢ + ⎥ (6.1.14)
⎝ W ⎠ TO η p ⎣ (W / S )cr qπAe ⎦⎜⎝ Wcruise ⎠

Chapter 6 Performance Analysis 49


Final Report of the International Design Synthesis Exercise 2005

This expression is valid for turbofan engines. For turboprop, equation (6.1.14) must be multiplied by the
cruise speed Vcruise to obtain the power-to-weight ratio:

⎛P⎞ V ⎡ qC D0 (W / S )cr ⎤⎛⎜ WTO ⎞


⎟⎟
⎜ ⎟ = cr τ tp ⎢ + ⎥ (6.1.15)
⎝ W ⎠ TO η p ⎣ (W / S )cr qπAe ⎦⎜⎝ Wcruise ⎠

Here (WTO Wcruise ) is the mission segment mass fraction for cruise and was found in [8] to be:

(WTO Wcruise ) = 0.820873 . The parameter τ is a correction factor, which corrects for two effects:
cruise altitude and Mach number.

The cruise altitude correction is needed, because the power/thrust produced by a turboprop/turbofan
engine decreases with altitude as a result of the lower atmospheric density. As such, the shaft power of a
turboprop engines can be corrected by relating it to the power at sea level according to the following
formula (see Ruijgrok [35]):

n
P ⎛ ρ ⎞
=⎜ ⎟ (6.1.16)
P0 ⎜⎝ ρ 0 ⎟⎠

Here, the exponent n has a value of 0.75 in the troposphere. Taking a cruise altitude of 7,500 km (25,000
ft) equation (6.1.16) produces a value of P P0 = 0.5534 .

The Mach number correction is needed in order to take into account variations in mass flow and turbine
inlet total temperature. If the Mach number increases the mass flow will increase, which will result in a
higher thrust. However, at the same time the inlet temperature will increase as well. This will limit the
maximum possible energy increase across the core of the engine, because of the maximum allowable
combustion chamber temperature. As a result, the generated thrust decreases. In case of a turboprop, the
variation of the available power as a function of the Mach number can be expressed as (see Shevell [36]):

γ
P ⎡ γ − 1 2 ⎤ γ −1 ⎡ T ⎤
= ⎢1 + M ∞ ⎥ ⎢1 − 0.25 cr M ∞2 ⎥ (6.1.17)
P0 ⎣ 2 ⎦ ⎣ T0 ⎦

This produces a value of P P0 = 1.1441 in cruise and P P0 = 1.1694 for maximum speed. So, the
Mach correction results in a positive contribution to the power available in cruise. This means that the
benefit gained from the increased mass flow is higher than the loss due to the increased turbine inlet total
temperature.

50 Performance Analysis Chapter 6


Final Report of the International Design Synthesis Exercise 2005

Finally the correction factor τ can be found by multiplying the altitude correction with the mach
correction. The results are summarised in table 6.1.1.

Table 6.1.1 Cruise correction factors


Altitude correction Mach correction τ
Cruise speed 0.5534 1.1441 0.6332
Maximum speed 0.5534 1.1694 0.6472

6.1.4 Climb
The climb requirements follow from the CS-25 airworthiness requirements [19]. These requirements are
based on several combinations, such as engine-out, landing gear position and flap position. In total there
are six climb requirements. These have been summarised in table 6.1.2.

Table 6.1.2 CS-25 Climb Requirements


Requirement Climb segment Engines CGR3 Flaps Gear Speed
(4 engines)
CS 25.121(a) Transition segment OEI1 >0.5% Takeoff Down V2
CS 25.111(c) Initial segment OEI1 >1.7% Takeoff Up V2
CS 25.121(b) 2nd segment OEI1 >3.0% Takeoff Up V2
CS 25.121(c) 4th segment OEI1 >1.7% Up Up 1.25 Vs
CS 25.119(a) Baulked landing AEO2 >3.2% Landing Down 1.3 Vs
CS 25.121(d) Discontinued approach OEI >2.7% Approach Up 1.5 Vs
1) One engine inoperative
2) All engines operating
3) Climb gradient required

In order to fulfil the climb requirements, the aircraft engines must be able to provide a certain excess
power. The required thrust-to-weight ratio with OEI then follows from:

⎛T ⎞ 1 N ⎛ 1 ⎞ 1
⎜ ⎟ = ⎜ + CGR ⎟ ⋅ (6.1.18)
⎝ W ⎠ TO η p N − 1 ⎝ L / D ⎠ 0.8

And with AEO:

⎛T ⎞ 1 ⎛ 1 ⎞ 1
⎜ ⎟ = ⎜ + CGR ⎟ ⋅ (6.1.19)
⎝ W ⎠ TO η p ⎝ L / D ⎠ 0.8

The corresponding power-to-weight ratios for turboprop can be found by multiplying equations (6.1.18)
and (6.1.19) with the climb speed, as specified in table 6.1.2. For that purpose the stall during takeoff from
Table 4.1.4 is used of Vs TO = 44.8 m/s.

Chapter 6 Performance Analysis 51


Final Report of the International Design Synthesis Exercise 2005

Based on Roskam [29] some assumptions have been made on the aerodynamic parameters. These
assumptions are listed in table 6.1.3.

Table 6.1.3 Assumed aerodynamic values


Configuration CL C D0 e
Clean aircraft 1.65 0.02 0.83
(∆C )
D0 flaps = 0.055
Landing Configuration 4.0
(∆C )
D0 gear = 0.015
0.77

(∆C )
D0 flaps = 0.015
Takeoff configuration 3.2
(∆C )
D0 gear = 0.015
0.79

6.1.5 Design Point


Having determined a series of relationships for the different performance requirements, it is now possible
to find the design point. In earlier stages this point was used as the starting point for further design; in the
present stage it can be used to verify whether the wing area and the power from the engines is still
sufficient to meet the performance requirements. For that purpose the expressions derived in the
preceding subsections will be plotted in a so-called constraints diagram. The resulting diagram is shown in
figure 4.3.1. For the climb requirements only the discontinued approach has been plotted, because it was
the most severe one.

Figure 6.1.1 Constraints diagram

52 Performance Analysis Chapter 6


Final Report of the International Design Synthesis Exercise 2005

From the diagram it follows that the design point is dictated by the maximum speed and the takeoff
distance. The design point is located at P / W = 18.4 W/N and W / S = 4650 N/m2. Using the latest
estimate for the maximum takeoff weight (see chapter 9) of 101124 kg, this gives that the required shaft
power is 4563 kW and the minimum required wing area is 213.3 m2. Given that the available shaft power
is 4586 kW and the wing area is 214.1 m2 this is just enough. However, this buffer is too small. It is
recommended that the analysis is repeated and extended in future project stages. Possibly, the
requirements have to be lowered again.

6.2 Climb performance

The climb performance describes the capacity to ascend or descend and its behaviour during the
manoeuvre. The important parameter in this behaviour is the rate of climb. It is described by formula
6.2.1:

Pa − Pr
RC = (6.2.1)
W

In order to find a worst case scenario and check whether the engines can perform then, the weight chosen
will be the MTOW.

As described in section 5.1 the power available can be computed for different altitudes and configurations.
Power available: using the propeller analysis program described in section 5.2 the efficiencies of the
propeller were found at the required altitudes.

To obtain the ‘power required’ curve at a specific altitude the power is calculated throughout the velocity
range marked by the lift coefficients determined by the wing. From the range of lift coefficients the range
of velocities are calculated with the following formula:

2W
V = (6.2.2)
C L ρS

where the weight is once again the MTOW.


Next, the power at a specific altitude level is determined by using the following equation:

C D2 2 W
Pr = W (6.2.3)
C L3 ρ S

Chapter 6 Performance Analysis 53


Final Report of the International Design Synthesis Exercise 2005

where:
CL
C D = C D0 + (6.2.4)
πAe
ρ = ρ (altitude) (6.2.5)

So for one value of CL the velocity and the power required is known if the air density is taken over form
the ISA tables. Some rate of climb curves are particularly interesting. The first one is off course the rate of
climb curve at sea level which shows the capacity to climb after take off or before landing. There is also
the rate of climb at cruise altitude and at the theoretical ceiling (the meaning of this will be explained in
the next section). These curves can be found in Appendix 9.

The spreadsheet used for the calculations in this section can be found on the CD in Appendix 7.

6.3 Load factor envelope

In steady level flight, wing lift supports the weight of the aircraft. However, during manoeuvres or when
flying through turbulent air, the structure of the aircraft is exposed to additional loads. Aircraft can be
designed to a maximum manoeuvring load factor depending on the operational role of the aircraft. It
follows that since the HERA is designed for relatively low speed, long range operations, its maximum load
factor will be lower relative to an aircraft such as a fighter jet. Typically, transport aircraft have maximum
load factors between 2.5 and 3.5 (page 21 of [27]). CS-25 regulations state that the maximum manoeuvre
factor is calculated by equation 6.3.1.

24000 (6.3.1)
n = 2.1 +
W + 10000

In this case, W is the aircraft mass in pounds, and n may not be less than 2.5. For the HERA, n = 2.2 so
the maximum manoeuvre factor selected is 2.5. Similarly, the regulations state that the negative limiting
load factor should not be less than -1. Again, due to the lower manoeuvrability of the HERA, this lower
limit value is selected.

6.3.1 Construction of Manoeuvre Envelope


The first step in the construction of the flight envelope is to determine the aircraft stalling speed for
various values of positive n. This can easily be achieved using equation 6.3.2.

2nw
VS = (6.3.2)
ρC L max

54 Performance Analysis Chapter 6


Final Report of the International Design Synthesis Exercise 2005

The maximum lift coefficients and wing loading parameters are taken from section 4.1. A similar analysis
can be performed using the minimum lift coefficient for the negative values of n, thus providing the lower
curve which can be seen in Figure 6.3.1. A line can then be drawn at VD, the design dive speed, which
was determined as 145 m/s EAS in the mid-term report [8], vertically up to an n value of 2.5. A line can
then be drawn horizontally towards the positive n stall curve until they intersect. In addition a vertical line
is drawn at the maximum cruise speed value with maximum and minimum n values of 2.5 and -1
respectively. Again, a horizontal line can be drawn, this time at n = -1 where it will intersect with the
negative n stall curve. The point for n = -1 and VC can then be linked to the n = 0 and VD point by a
simple line as shown in Figure 6.3.1. This completes the construction of the manoeuvre envelope.

6.3.2 Construction of Gust Envelope


A gust causes a load change due to the rapid change in angle of attack and consequently a change in wing
lift. The change in wing lift divided by aircraft weight gives the change in load factor n. Equation 6.3.3
shows the basic formula (in imperial units) for change in load factor n due to a gust where Ude is the
velocity of the gust, and C L α is the change in lift curve slope due to change in angle of attack.

C LαaVeU de
∆n = (6.3.3)
498w

However, equation 6.4.3 assumes a ‘sharp edged’ gust and relies only on the wing loading. This equation
also assumes that aircraft of a given wing loading will exhibit the same dynamic response. This
simplification does not include alleviation effects resulting from real unsteady aerodynamics therefore a
gust alleviation factor, Kg is implemented. This takes into account the gust gradient, airplane response and
lag in lift increment caused by the change in angle of attack (page 25 of [27]). Equation 6.3.3 then
becomes equation 6.3.4 (again in imperial units).

C LαaVeU de
∆n = 1 ± K g (6.3.4)
498w

The standard gust velocities of 60, 50 and 25 feet per second were applied to equation 6.3.3 to determine
the increase or decrease in loading caused by the gusts. To apply the results to the figure, gust lines are
drawn from n = 1 for stall speed, cruise speed and dive speed. The analysis assumes that the pilot will
decrease airspeed as turbulence is encountered (page 45 of [27]). Therefore a box can be drawn round
several points. For positive and negative values of ∆n , the points are as follows:
• Intersection of the 66fps line and stall curve
• Intersection of the 50fps line and cruise speed line
• Intersection of 25 fps line and design dive speed line

Chapter 6 Performance Analysis 55


Final Report of the International Design Synthesis Exercise 2005

The end result is the bold envelope seen in Figure 6.3.1. From this figure it can be seen that gust loads are
not a design problem for this aircraft, therefore the aircraft should be designed to the limit load factor of
2.5 and safety factor of 1.5 (i.e. ultimate load factor of 3.75).

V-n Diagram (Combined Flight Envelope)


3

+ve limit load factor


2.5
50 fps
2

66 fps 25 fps
Normal acceleration factor, n

1.5

1
VB VA VC VD

0.5

0
0 20 40 60 80 100 120 140 160

-0.5
VEAS

-1
-ve limit load factor
-1.5

Figure 6.3.1 V-n diagram (manoeuvre envelope and gust envelope combined)

6.4 Flight envelope

The flight envelope defines the speeds the aircraft can adopt at certain heights during its flight. These
limitations depend on altitude, engine properties, human crew and structural constraints.
The first limitation to the velocity range is the stall limit at low velocities. This is determined by the stall
properties of the wing and thus the lift coefficient range. Using maximum lift the aircraft can decelerate at
any one altitude up to a velocity given by the formula:

2W
VMIN = (6.4.1)
ρC LMAX S

Connecting the minimum speeds at different altitudes gives the minimum speed profile plotted in the
flight envelope, see figure 6.4.1.

56 Performance Analysis Chapter 6


Final Report of the International Design Synthesis Exercise 2005

Figure 6.4.1 Flight envelope of the HERA

With altitude air becomes thinner due to gravitational effects. Since the working principle of the propeller
is to push the air in order to obtain thrust, altitude has an important effect on the performance of the
engines. The thinner the air is, the lesser thrust the engines can provide for the propulsion of the aircraft.
This means that there will come an altitude where the air is to thin for the engine to provide thrust to
climb. This is called the theoretical ceiling. At this altitude the power that the engine can deliver is the
same as the power that is required for forward flight and is no longer able to climb. However the aircraft
will never be allowed to reach such a ceiling. The service ceiling, which has a specific margin of an excess
of thrust to provide a climb rate of 1 m/s, will be adopted as ceiling limitation coming from the engines.

Calculating the rate of climb for various altitudes it can be seen that it slowly lessens until it reaches the
minimum allowed of 1 m/s at a height of 15km, which becomes the theoretical ceiling of the HERA. The
resulting power and rate of climb curves can be found in Appendix 9.

However since the HERA does not make use of pressurisation the ceiling is actually restricted by the
crew’s physical limitations. The use of oxygen masks is mandatory after an altitude of 3000 m to prevent
diminished consciousness due to hypoxia. The masks allow the crew to rise up to an altitude of 7620 m.
This is thus the real ceiling for the HERA.

Every aircraft has a maximum operative Mach number. This limit is due to compressibility effects the
aircraft encounters when approaching the sonic range. The Mach number at which compressibility effects
commence is M=0.8. To be safely under that range the maximum operative Mach number of the HERA

Chapter 6 Performance Analysis 57


Final Report of the International Design Synthesis Exercise 2005

has been chosen to be M=0.75. To find the corresponding velocities at different altitudes the velocity can
be computed with the following formula (per altitude):

V = M γRT (6.4.2)

where the temperature is a function of the altitude, following the ISA tables. A collection of the velocities
at different altitudes gives the line ‘max. operative Mach number’ shown in Figure 6.4.1. Using a safety
margin of 10% gives the diving Mach number curve. There is a maximum operating speed that may not
be exceeded in flight. It must no be greater than the design cruising speed. The cruising speed is of 166.7
m/s at ground level. At different altitudes the airspeed becomes:

ρ0
V = Vh = 0 (6.4.3)
ρ

The collection of the data obtained at different altitudes gives the maximum operating speed curve. Using
in a safety margin of 10% gives the design diving speed curve. Collecting these curves in a graph shows a
range of speeds and altitudes in which the aircraft is operative. It is the flight envelope and is the surface
that is shaded in the graph of Figure 6.4.1.

The spreadsheet used for the calculations in this section can be found in Appendix 7.

6.5 Payload-range diagram

In chapter 5 the AE1107c was presented as the most suitable engine for the HERA. Using the
specifications of the AE1107c (Table 5.1.1), it is possible to construct a payload-range diagram. This
diagram may be regarded as the basis of the economic value of a transport airplane: an aircraft should be
capable of transporting large payloads as far as possible.

The payload-range diagram can be constructed using the Breguet-range equation:

η p CL ⎛W ⎞
R= ln⎜⎜ 1 ⎟⎟ (6.5.1)
SFC C D ⎝ W2 ⎠

This equation is valid, if the angle of attack is held constant during cruise and if the propulsive efficiency
η p and the specific fuel consumption SFC are assumed to be constant as well, see e.g. Ruijgrok [35].

58 Performance Analysis Chapter 6


Final Report of the International Design Synthesis Exercise 2005

The resulting payload-range diagram is shown in Figure 6.5.1 and compares the range performance of the
HERA to that of the competing A400M, the An-70 and the C130J Hercules. The range with full payload
of the HERA is estimated at 5490 km. Trading payload for fuel can further extend the range. The
maximum range at full fuel capacity and with a reduced payload of 10900 kg is 12000 km. Finally the
range of the HERA can be increased even more by further decreasing the payload. The ferry range (range
with zero payload) is estimated with equation (6.5.1) to be 14600 km.

One of the top-level requirements is that the aircraft should have a range with full payload of 4000 km.
From Figure 6.5.1 it can be seen that the HERA not only satisfies this requirement, but that it also
outscores all other competitors. This is due to the high aspect ratio of the HERA that is about 20%
higher. As a result the lift-to-drag ratio is about 10% higher. Another contribution comes from lower
operating empty weight of the HERA, which allows a higher weight fraction W1 W2 . The lower empty
weight of the HERA is achieved due to the non-pressurised cabin amongst others.

The spreadsheet used for constructing the payload-range diagram can be found on the CD in Appendix 7.

50000

40000 Maximum range at full payload


Payload Weight (kg)

30000
Trading payload for fuel HERA
Maximum range at full A400M
fuel capacity
20000 An-70
C130J

10000

Ferry Range

0
0 2000 4000 6000 8000 10000 12000 14000

Range (km)

Figure 6.5.1 Payload-range characteristics of the HERA.

Chapter 6 Performance Analysis 59


Final Report of the International Design Synthesis Exercise 2005

7 Airframe Structural Design

This section deals with the structural design of the airframe of the HERA. In section 7.1 the procedure
for material selection will be explained. This procedure will then be applied to several important aircraft
parts in the chapters such as the wing, fuselage, empennage and cargo ramp. The structural design of these
elements will be discussed in the sections 7.2 to 7.5.

7.1 Materials selection

The selection of an appropriate material and its subsequent conversion into a useful product with desired
shaped and properties can be a rather complex process. Several methods have been developed for
approaching a design and selection problem. The one used in this project is the case-history method.
Evaluating what has been done in the past, or what is currently being done by a competitor, can yield
important information. Using this as a starting base can reduce the total purchase cost (especially
manufacturing and qualification costs) of the aircraft drastically, which is the primary cause in a cost-
driven project. That’s why advanced composite materials can already be eliminated as much as possible
from the material selection since cost is a driving requirement. Besides that, HERA has to operate in
severe conditions. This means, if damage occurs to composite materials it is usually hard to determine the
exact location of the damage and to repair it. However, dents occurring in aluminium can easily be
recovered. The second major criterion is the weight, which can be achieved with a chosen material.
However, weight and cost are interchangeable since expensive material needs less of it to achieve the same
performance, which can mean a reduction in material cost. Other criteria are manufacturability (maximum
elongation, heat treatment of a specific material or alloy) and maintainability (fatigue properties). This
leads to the main alloys used in the aerospace industry for producing airframes: the aluminium 20xx and
70xx series. The properties of these two widely used materials are shown below in Table 7.1.1.

Table 7.1.1 Material characteristics


AL 2024-T3 AL 7075-T6
Yield strength (MPa) 345 503
Shear strength (MPa) 283 331
Max elongation (%) 18 11
Density (kg/m3) 2780 2810
Price ($/kg) 24 24

Using these properties, a selection can be made for the airframe structure. This will be discussed in the
next sections.

Chapter 7 Airframe Structural Design 61


Final Report of the International Design Synthesis Exercise 2005

7.2 Wing

7.2.1 Spanwise wing loading: shear and bending


In order to carry out a structural analysis of the spars and stringer in the wing, first the loads need to be
determined in terms of a shear force and a bending moment diagram. For that purpose the following
contributions will be considered:
- spanwise lift distribution
- wing structural mass distribution
- fuel mass distribution
- power plant mass
These contributions are schematically represented in Figure 7.2.1a.

The shear force and bending moment due to a distributed load are given by (see for example Meriam &
Kraige [26]):


V = q( y )dy (7.2.1)

M = q ( y ) ydy
∫ (7.2.2)

In order to simplify the integration procedures, the spanwise lift distribution from section 4.1.3 is not used
here. Instead an elliptical lift distribution will be assumed for now, as given on page 367 of Anderson [3]:

2
4L ⎛2y⎞
L ' ( y ) = q lift ( y ) = 1− ⎜ ⎟ (7.2.3)
bπ ⎝ b ⎠

This distributed load acts in the aerodynamic centre of the wing, which lies at 25%c (see section 4.3).
Substituting L = nW and integrating to obtain the shear force and bending moment gives:

2nW ⎡ϑ 1 ⎤
Vlift = − ⎢ − sin(2ϑ )⎥ (7.2.4)
π ⎣2 4 ⎦

nW sin 3 (ϑ )
M lift = (7.2.4)
π 3

b
Here, the transformation y = cos(ϑ ) was used.
2

62 Airframe Structural Design Chapter 7


Final Report of the International Design Synthesis Exercise 2005

The wing structural mass is linearly distributed over the wing planform, as is the fuel mass. The resulting
expressions for q, V and M are given in table Table 7.2.1. Here, M wing is determined in chapter 9 and

equals 8288.39 kg. The usable fuel tank capacity in the wing V f w can be determined using page 448-449

of Torenbeek [37] and is found to be 51.85 m3. With a fuel density of 790 kg/m3 this gives M fuel = 40960

kg.

Table 7.2.1 Expressions for q, V and M for the wing structural mass and the fuel mass.
Wing structural mass Fuel mass
Distributed
q wing =
2 M wing ng ⎡ (1 − λ )
1−

q fuel =
2 M fuel ng ⎡

(
1 − λ2 ⎤)
load b (1 + λ ) ⎢⎣ b/2
y⎥
⎦ ( )
b 1 + λ2 ⎣
⎢1
b/2 ⎦
y⎥

V wing =
2 M wing ng ⎡
ψ−
1 (1 − λ ) 2 ⎤
ψ ⎥ V fuel =
2 M fuel ng ⎡
ψ −
(
1 1 − λ2 2 ⎤
ψ ⎥
)
Shear force 1)
b (1 + λ ) ⎣⎢ 2 b/2 ⎦ ( )
b 1 + λ2 ⎣

2 b/2 ⎦
Bending 2 M wing ng ⎡ 1 2 1 (1 − λ ) 3 ⎤ (
2 M fuel ng ⎡ 1 2 1 1 − λ2 3 ⎤ )
M wing = ψ − ψ ⎥ M fuel = ⎢ ψ − ψ ⎥
moment 1) b (1 + λ ) ⎢⎣ 2 3 b/2 ⎦ (
b 1 + λ2 ⎣ 2) 3 b/2 ⎦
1) ψ = (b / 2 − y )

The contribution of the engines follows from the factored engine weight, which includes the engine, the
propeller and the nacelle, and was given in section 5.1: 2605.5 kg. If all contributions are now added , the
shear force and bending moment diagrams can be constructed. The maximum load factor the aircraft has
to be able to withstand, follows from the V,n-diagram (see section 6.4): n max = 2.5 . The shear forces and

bending moment diagrams for n max = 2.5 are given in Figure 7.2.1b and c, respectively. The diagrams
have been constructed using a safety factor of 1.5, so that they represent the limit load case.

The wing mass and fuel mass can be seen to have a relieving effect on the shear forces and bending
moments in the wing. The outboard fuel tanks have a larger relieving effect than the inboard fuel tanks
and are usually emptied last. The effect of that can be seen in figure Figure 7.2.1 as well. The maximum
loading occurs with empty wing fuel tanks. This load case will be used in the next subsection for
determining the structural arrangement of the wing and the wing fuselage connection.

The spreadsheet used to construct the shear force and bending moment diagrams can be found on the
CD in Appendix 7.

Chapter 7 Airframe Structural Design 63


Final Report of the International Design Synthesis Exercise 2005

250000

0
0 5 10 15 20 25 30
y (m )
V (N)

-250000

-500000

-750000

-1000000

15000000

10000000
M (Nm)

5000000

0
0 5 10 15 20 25 30
y (m )

Empty fuel tanks Fuel in outboard tanks Fuel in inboard and outboard tanks

Figure 7.2.1 From top to bottom: a) Spanwise wing loading, b) Shear force diagram,
c) Bending moment diagram.

64 Airframe Structural Design Chapter 7


Final Report of the International Design Synthesis Exercise 2005

7.2.2 Spanwise wing loading: torsion


In addition to the shear force and bending moment, which the wing will experience in flight, the
aerodynamic loads also cause torsion about the flexural centre of the wing. This torsion is caused by the
moment, which the airfoil produces and builds up from the tip to the wing root where it is a maximum.
For a straight tapered wing, the moment, and hence the torsion produced by the airfoil, can be found
using equation 7.2.5, where c(y) is the variation of the local chord with position along the span.

y
1
T= ρV 2 C m ∫ c 2 ( y )dy (7.2.5)
2 0

Since the local chord varies linearly with the span, an equation for c(y) was determined easily and is shown
as equation 7.2.6.

c( y ) = 0.1711y + 2.34 (7.2.6)

Filling into equation 7.2.5, an expression was produced for the variation of moment along the span and
can be found as equation 7.2.7.

T=
1
2
(
ρV 2 C m 0.00977y 3 + 0.400y 2 + 5.476y ) (7.2.7)

Using the value for Cm obtained in section 4.1.2, values for the moment about the flexural centre were
calculated and multiplied by the loading factor of 2.5 and safety factor of 1.5. The final torsion values
were plotted against various spanwise positions, and the resulting graph can be found in Figure 7.2.2.

Spanwise Variation of Torsion about Flexural Centre

0
0 5 10 15 20 25 30

-500000

-1000000
T (Nm)

-1500000

-2000000

-2500000
y (m)

Figure 7.2.2 Spanwise variation of torsion about flexural centre

Chapter 7 Airframe Structural Design 65


Final Report of the International Design Synthesis Exercise 2005

7.2.3 Wing structural analysis: assumptions and method


In flight, the wing is subjected to different forces: bending, shear, and torsion. Therefore, the structure
that is to be designed has to provide enough resistance to these combined forces and moments. To solve
this, the problem is simplified into three different cases that, combined, provide a good approximation to
the initial problem. The three cases are: (1) Axial stress, (2) Torsion, and (3) Bending Shear Stress.
Nevertheless, the wing is a complex element due to its particular geometrical section. Hence, assumptions
must be made to simplify and facilitate the calculations. The most important are stated below:

• The structure of the wing box is be idealised as a rectangular box, Figure 7.2.3.
• The front and rear spar caps are assumed to be of the same area and have locations at 15% and
55% of the local chord respectively. This allows enough space for the flaps (40% of the local chord)
and 5% for the flap mechanisms.
• It has been estimated that the wing structure contains 40 ribs; this was estimated by a comparison
of different cargo aircrafts and referring to page 22 of Roskam Part III [31].
• All stringers are assumed to have the same area, all parallel to the front spar and equally spaced.
• It is assumed that the lift distribution at each spanwise position would be of 75% at the front spar
and 25% at the rear spar. (Assumption used in the bending shear stress calculation).
• The axial stress is carried by the spars and by the stringers and it is therefore assumed that the skin
only carries shear flow.
• It is to be assumed that buckling stress of a component (e.g stringer) is greater than the yield stress
of the material since buckling of wing structures is not allowed. Of course, the actual stress in the
component should not exceed the yield stress of the material.
• The flexural centre and the centre of gravity coincide (35% of the local chord).
• Five different stations along the span will be considered to show the variation of component
dimensions along the span.
• For pure torsion analysis, a two cell structure (see Pure Torsion) is considered.
• As the wing aerofoil section decreases in thickness from 17% to 13%, to ease geometry
calculations, the NACA LS(1)-0417MOD (17%) is considered from root to halfway through the
semi span and the NACA LS(1)-0413MOD (13%) for the rest of the wing.

Figure 7.2.3 Idealised wing box

66 Airframe Structural Design Chapter 7


Final Report of the International Design Synthesis Exercise 2005

7.2.4 Axial stress


The axial load depends on the moment caused by aerodynamic forces on the wing. From structural
principles, the ETB equation can be applied to solve the axial stress problem (7.2.8). This expression can
be expanded and simplified for a case were all the axial stress is carried by the spars. Expanding the term
“I” (moment of inertia) and assuming that the maximum stress allowed will be that of the material yield
stress, the only unknown left is the area of the spar caps. This equation gives a first estimate of the
minimum spar cap area.

My M
σ= , where I = ΣAh 2 σ=
I ⎛H⎞ (7.2.8)
4 Aspar ⎜ ⎟
⎝2⎠

The next step to be taken is to apply the equation of skin buckling (7.2.9, page 177 [25]). From 7.2.8 the
maximum distance at which skin would not buckle can be found, and hence the number of stringers
obtained. A skin thickness is chosen at this stage, which is later verified in the torsion analysis.

2
⎛t⎞
σ skin = 3.62 E ⎜ ⎟ (7.2.9)
⎝b⎠

The stringers chosen are Z-shaped; a discussion on stringer selection follows later. On application of
stringers to prevent skin buckling, it should be noted that the stringers will now carry part of the axial
stress and so the spar cap area can be reduced. It has been assumed that the stringers carry 15 % of the
axial load, therefore the spar cap area found by equation 7.2.2.1 can be factored by 0.85 to find the new
spar cap area. As before, the yield stress is used to find the minimum areas, hence, the new equation is:

M
σ=
H
(4 Aspar + 2nAstringer ) (7.2.10)
2

From this area value, approximate stringer dimensions can be entered into the spreadsheet, which
calculates the area required by equation 7.2.9. Once the area is satisfied for chosen stringer thickness and
height, the stringer itself must be tested to ensure that the buckling stress of the chosen configuration is
greater than the yield stress of the material (equation 7.2.11 pp 176 [25]). The inertia in equation 7.2.11 is
that of the Z stringer added to that of the skin.

π 2 EI
σ= (7.2.11)
Astringer L2

Chapter 7 Airframe Structural Design 67


Final Report of the International Design Synthesis Exercise 2005

In addition, the actual maximum stress in the stringer for the loading described in a previous section is
computed to ensure that the maximum stress does not exceed that of the material.

Finally, once all stress requirements are satisfied, the preliminary values of skin thickness and areas of spar
caps and stringers are used in the next stage of the analysis process.

7.2.5 Bending Shear Stress


In real life, the lift acting from the aerodynamic centre causes bending shear across the wing cross-section
due to the fact that the centre of gravity is located aft this point. In this idealised case the bending is also
similar to the real case.

The loading assumption is divided into two cases, which have to be analysed. The first case being an
idealised box with 75% of lift acting on the front spar and the second case being 25% of the lift at the rear
spar. Calculating the shear flow over each surface and then, superimposing both results, gives the shear
flow around the box caused by bending. Figure 7.2.4 shows a graphical explanation.

Figure 7.2.4 Idealisation of structure for bending shear stress analysis

The calculation of the shear flow can provide an estimate of minimum skin thickness required based on
the material shear stress (equation 7.2.12). Since an initial value for skin thickness is already known from
the first stage of analysis, equation 7.2.12 is simply used as a check.

q = τ .t (7.2.12)

The shear flow around each box is found by dividing the upper and lower surface into skin panels in
between the stringers. Each of those smaller surfaces has a shear flow that can be calculated by equation
7.2.13 pp 339 [25], which considers only the “q” loads going in and out of each spar and stringer area
(note this equation assigns Br as boom area).

Vy s
q = q0 −
I
∑B
0
r yr (7.2.13)

68 Airframe Structural Design Chapter 7


Final Report of the International Design Synthesis Exercise 2005

This produces a solvable matrix. However, when all inputs are applied to equation 7.2.13, there are n-1
equations, i.e. one equation less than that needed. The additional equation is found by taking moments
around the node were the lift is applied.

0 = Hb(q1 + ... + q n − 2 ) + Hb(n − 1)q n −1 (7.2.14)

Finally, there are n-equations and n-unknowns, which are solved in matrix form as q = A-1.b giving the
shear flows for the two models. The shear flows for the 75% and 25% case are then added; hence the final
shear flow caused by bending is found. Moreover, it must not be forgotten that the shear flow caused by
torsion has to be superimposed to this shear flow.

7.2.6 Pure Torsion


In addition to the axial and bending stresses that affect the wing structure, the torque experienced by the
wing must also be considered. As mentioned previously, the various loads and moments can be
considered separately and their effects combined at the end of the analysis to determine a structure
suitable to withstand their combined effects. Therefore, this part of the analysis considers the case of
pure torsion. Since the section is to be loaded by a pure torque, the presence of the booms (spars and
stringers) has no effect. Even though the wing box is considered as a single cell, when the leading edge
section is attached to the front spar, the leading edge skin in fact carries some of the shear loads even if
this is not desired in design. As a result, for this case, the shear flow due to the torsion loads can be
assumed to be carried by both sections. According to Bredt-Batho torsion theory for a multi-cell box
structure, outlined in p383 of Megson [25], the torsion around the structure is given by equation 7.2.15
below.

N
T = ∑ 2 AR q R (7.2.15)
R =1

In addition, the rate of twist caused by the variable shear flow (pp 309 Megson [25]) is shown in equation
7.2.16 below.

dθ q ds
dz 2 A ∫ Gt
= (7.2.16)

Since the same material is being used throughout the wing box structure, G (shear modulus of elasticity) is
assumed to be constant and can therefore be removed from the integral. However the shear flow varies
and must be moved inside the integral. Another consideration is the application of the compatibility
theory that states that the rate of twist in one section must equal the other (equation 7.2.17).

Chapter 7 Airframe Structural Design 69


Final Report of the International Design Synthesis Exercise 2005

dθ dθ
= (7.2.17)
dz 1 dz 2

Figure 7.2.18 shows the notation and convention used in the analysis. Panel thickness is represented by t,
and panel length by e.

T
t1, e1 t2, e2

t3, e3 t5, e5
q1 q2

t4, e4

Figure 7.2.5 Notation and convention used for pure torsion loading

Applying equations 7.2.15 to 7.2.16 to Figure 7.2.5, expressions for shear flows q1 and q2 were found
(equations 7.2.18 and 7.2.19).

⎡ A ⎛e e e e ⎞ e ⎤
T ⎢ 1 ⎜⎜ 2 + 3 + 4 + 5 ⎟⎟ + 3 ⎥
q1 = ⎣ A2 ⎝ t 2 t 3 t 4 t 5 ⎠ t 3 ⎦
(7.2.18)
⎡e e A2 ⎛ e e e e ⎞⎤
2 A2 ⎢ 1 + 3 ⎛⎜1 + 2 A1 ⎞⎟ + 12 ⎜⎜ 2 + 3 + 4 + 5 ⎟⎟⎥
A2 ⎠ A t
⎣ t1 t 3 ⎝ 2 ⎝ 2 t 3 t 4 t 5 ⎠⎦
T − 2 A1 q1
q2 = (7.2.19)
2 A2

Once the shear flows for this case were determined, the shear flows from the bending shear analysis were
added giving a total shear flow in each skin panel. The shear stress in each panel was calculated using
7.2.12 for the skin thicknesses determined in the first step of this analysis and compared to the material
yield stress. For this aircraft, it is found that the torsion load was not the critical load for design (see
discussion of results) and thus all thicknesses produced from the previous analysis are found to be
adequate to accommodate the torsional load.

7.2.7 Results
From the above analyses the following table 7.2.2 summarises the results for 5 different positions along
the span.

70 Airframe Structural Design Chapter 7


Final Report of the International Design Synthesis Exercise 2005

Table 7.2.2 Summary of wing structure dimensions


Upper/Lower
Spars Stringers (spacing = 0.15m) Skin
Area F Web R Web Area Height Thickness Thickness
Case x (m) (m2) T (m) T (m) Number (m2) (m) (m) (m)
1.42E- 3.20E- 0.007
1 0 0.0055 0.006 14 0.042 0.004
02 04
9.58E- 2.47E- 0.005
2 6 0.0055 0.006 12 0.037 0.0035
03 04
4.53E- 1.54E- 0.003
3 13 0.005 0.0055 9 0.032 0.0025
03 04
1.70E- 9.38E- 0.002
4 18.25 0.0045 0.005 6 0.032 0.0015
03 05
2.41E- 3.08E- 0.001
5 23.25 0.003 0.002 4 0.031 0.0005
05 05

Table 7.2.2 shows the primary areas and dimensions out put from the analyses described. Since the
loading decreases along the span, it was expected that the geometry of the structure would also decrease.
This is clearly shown in Table 7.2.2. Although it is cheaper to manufacture sections of constant
dimensions, the effect would be extreme over-design for the sections of the wing closer to the tip. Excess
material increases weight and cost. The results above indicate that spar, stringer and skin dimensions
taper towards the tip. This may be the ideal solution but it is expensive to manufacture. Instead, the
dimensions will decrease in sections at different rib locations. What has not been determined from the
analysis are the actual spar dimensions. This discussion follows.

Spar Caps
When distributing the area of the spar caps, examples of various shapes must be considered. The most
common design is the double L structure where the spar web is placed between the L structures, which
provide stability for the web. When distributing the area over the double L structure, it can be observed
that a large increase in section occurs between the skin panels and the spar cap itself. Large changes in
section are prone to stress concentrations and thus should be avoided where possible. A solution to this
problem is to add an additional rectangular layer below the double L structure. This provides a more
gradual change in section since the dimensions of the double L are also reduced. In addition, the double
L structures should not have sharp corners. It is common for the L structures to be either bent or
extruded thus eliminating potential stress concentrations at the inner corner of an L. For example: for the
root case (1), the L shapes are found to be 15 cm length and height and of 1.2 cm thickness whilst the
second layer has a total length of 50 cm and thickness of 1.5 cm. The resulting section is shown below
(Figure 7.2.6), followed by a close up of the spar cap (Figure 7.2.7).

Chapter 7 Airframe Structural Design 71


Final Report of the International Design Synthesis Exercise 2005

Figure 7.2.6 Internal view of root section

Figure 7.2.7 Close up of spar cap at root

Stringers
Unlike the spar caps, the dimensions of the stringers are determined during the analysis stage based on the
required area and a buckling analysis. The choice of stringer has to be made at the first stage of analysis so
ensure the buckling calculations are correct. Several stringer shapes are available including Y shape, L
shape, Z shape and top hat configurations. Examples like the Y stringer are rarely used due to their
complexity and hence cost. The L shape is the simplest shape. However, for a given area, the L shape
must have greater thickness than say the Z shape. Again, this causes large changes in section between the
skin and the stringer causing undesirable stress concentrations. The top hat shape provides a solution to
this problem in that for a given area, the overall thickness can be smaller. However, the top hat is not
selected since it is known to have problems with moisture ingression, corrosion and is difficult to inspect.
Therefore, it follows that the logical choice for this design is the Z shape stringer. The Z shape is easily
manufactured by extrusion and is easily inspected as all sides are easily viewed. Again, for a given area, the
thickness can be smaller than the L shape thus alleviating potential problems of stress concentrations.

Figure 7.2.8 Stringer close up

72 Airframe Structural Design Chapter 7


Final Report of the International Design Synthesis Exercise 2005

Figure 7.2.8 shows a close up of the stringers for the wing root. The stringers are at 0.15 m spacing and
have filleted edges rather than sharp corners again to lessen the stress concentrations. Typically these
stringers would be riveted to the skin with a countersunk rivet on the outer skin panel to lessen drag.

Spar Web
The spar web thicknesses are quite straightforward outputs. During analysis it can be observed that to
withstand the load, very large skin thicknesses were required in the front spar web whilst relatively low
thicknesses were required in the rear spar web. Rather than to suffer the weight penalty of extra ribs,
which provide added strength not required at the rear spar, a stiffener is inserted in between the ribs on
the front spar. This allows the skin thicknesses to be decreased on the front web whilst having a lesser
weight penalty than additional ribs. Stiffeners are added up to position 4. Stiffeners are not required at
position 5 (see shading in Table 7.2.2).

7.3 Fuselage

In this chapter the structural analysis of the fuselage will be discussed. For most of the calculations the
book of Megson was used, reference [25].

The subjects that will be discussed are:


1. Torsional loading of the fuselage
2. Bending of the fuselage
3. Shear loading of the fuselage

To be able to treat these subjects separately, the assumption was made that the skin stiffeners carry the
entire bending load, while the skin itself carries all shear loads due to torsional moments and shear forces.

The goal of these calculations is to minimise the weight of the structure but still keep it strong enough to
cope with all load cases the aircraft might encounter during its operational life. To make sure this is the
case, the most extreme load cases are used for the calculations. To calculate these extreme cases either the
static case was multiplied with the max G-number and a safety factor, or a reasonable upper limit to the
extreme load was taken. Therefore the result of the torsion and shear calculations is a minimal thickness
needed to withstand the shear flows resulting from these loads and the result of the bending calculations is
a minimal stiffener area in order to be able to resist the bending moments.

7.3.1 Torsion of the fuselage


During flight an aircraft has to make all kinds of manoeuvres. Apart from bending and shear an aircraft
also experiences torsion, that is, it is loaded by a moment around its longitudinal axis.

Chapter 7 Airframe Structural Design 73


Final Report of the International Design Synthesis Exercise 2005

There are three big factors that produce torsion in the fuselage during flight:
1. Deflection of the ailerons during flight produces a torsion entering the fuselage at its centre. Its
magnitude depends mainly on the angle of deflection and the current airspeed.
2. Apart from a moment along the planes vertical axis deflection of the rudder also produces a
moment along the planes longitudinal axis. This is due to the fact that the point on the vertical tail
where the force is experienced is above the planes longitudinal axis.
3. Gust loads will produce all kinds of loads on wings, horizontal- and vertical stabilisers. These gust
loads will create torsion in a similar way as deflection of a control surface such as the aileron or
rudder does.

When calculating the required skin thicknesses and stiffener areas the assumption is made that the
stiffeners carry all the bending stresses and that the skin carries all the shear loads and the torsion.

Of the three load cases described above the first one has proven to be the critical one. So the skin
thicknesses will be calculated using the torsion that is produced when the aircraft fully deflects its aileron
very abruptly during maximum flight speed multiplied with a safety factor of one and a halve. The torsion
found in this way is approximately 12 . 106 Nm.

The model representing the fuselage consists of three different parts:

1. The centre part of the fuselage is modelled by a section consisting of three cells; namely the cabin,
the wings torsion box and a small compartment under the cabin floor. See Figure 7.3.1.

Figure 7.3.1 Torsion model for the centre part of the fuselage.

74 Airframe Structural Design Chapter 7


Final Report of the International Design Synthesis Exercise 2005

2. The front and rear fuselage sections connected to the centre fuselage. These are modelled by a
two cell section, one of which is the cabin and the other one the compartment under the cabin
floor. See Figure 7.3.2.

Figure 7.3.2 Torsion model for the front and rear part of the fuselage

3. Normally the rear part of an aircraft is a closed section just like the rest of the aircraft. But since
the emergency cargo airplane is also used to perform airdrops, it should be able to fly with its
ramp down, so the rear fuselage is effectively an open section in this situation. Therefore the rear
aircraft structure should be able to transfer the torsion both with and without the ramp lowered.
Off course the ramp lowered situation is the critical one. In order to withstand the torsion in this
part of the fuselage, there is a torsion box present in the upper part of the rear fuselage where the
ramp is located. The skin thicknesses in this tail torsion box will have to be dimensioned in such a
way, that it can withstand the torsion by itself without the ramp transferring any torsional loads.
See Figure 7.3.3.

Figure 7.3.3 Torsion model for the tail torsion box

Chapter 7 Airframe Structural Design 75


Final Report of the International Design Synthesis Exercise 2005

First a certain thickness is assumed for each skin panel in the cross-section. This thickness will be
minimised using Excel’s solver in such a way that in each skin panel the shear stress is lower than the
shear yield stress for the type of aluminium we are using.

According to the book of Megson the shear flows due to torsion in a multicell structure can be calculated
using the following formulas:

N
T = ∑ 2 ⋅ AR ⋅ q R (7.3.1)
R =1

and for each cell

dθ 1 ds
=
dz 2 ⋅ AR ⋅ G REF ∫ q⋅ t
R * (7.3.2)

where

G
t* = t (7.3.3)
G REF

When the change of rotation is set equal for each cell

⎛ dθ ⎞ ⎛ dθ ⎞ ⎛ dθ ⎞
⎜ ⎟ =⎜ ⎟ = ... = ⎜ ⎟ (7.3.4)
⎝ dz ⎠1 ⎝ dz ⎠ 2 ⎝ dz ⎠ N

N+1 equations are obtained with N+1 unknowns, namely each qR and one dυ/dz, which can be solved
using Excels solver for instance.

After finding the constant shear flows in each cell, the shear flows in the walls of each cell are easy to
calculate. Now the thicknesses can be decreased until the shear stress in the skin panels approaches the
shear yield stress. This process is an iterative one, so for each time the thicknesses are decreased the solver
has to calculate the new shear flows, after which it is possible to determine whether the thicknesses can be
decreased or not.

After repeating this process a number of times, a minimal required skin thickness can be found for each
skin panel. For production reasons all thickness for a certain section are set to the same value, so what it
all comes down to is finding that one minimum skin thickness for which all the shear stresses in all skin
panels are lower than the shear yield stress. The results can be found in table 7.3.1.

76 Airframe Structural Design Chapter 7


Final Report of the International Design Synthesis Exercise 2005

Table 7.3.1 Results of the torsion calculations


Minimum required Skin mass of the Material cost of the
skin thickness (mm) entire fuselage part entire fuselage part ($)
(kg)
Centre fuselage
Al 2024-T3 0.9 413 9915
Al 7075-T6 0.7 325 7795

Front and rear


fuselage
Al 2024-T3 1.3 1062 25496
Al 7075-T6 1.1 909 21806

Tail torsion box


Al 2024-T3 10.0 2623 62959
Al 7075-T6 9.0 2386 57275

From this table it follows that choosing Aluminium 7075-T6 instead of Aluminium 2024-T3 gives both a
weight and a cost advantage. But the penalty of choosing this material is a less fatigue resistance.

As can be seen from the table, the thicknesses are somewhat large, especially for the tail torsion box,
however the calculations are for an extreme case. Under normal operation such forces will not be
encountered and therefore the skin could have been thinner in general. Besides that, the method used to
calculate the shear flows assumes that the object is clamped on one side, so that reaction forces occur. In
this case however, there can be no reaction forces since the fuselage is free to rotate. Therefore the shear
flows calculated are the shear flows that would occur when the fuselage would be clamped. These shear
flows can be seen as the shear flows that occur due to inertia of the fuselage at the instant the ailerons are
deflected when the wings start to rotate, but the fuselage remains still. Also, for the tail torsion box, the
assumption was made that the tail torsion box will carry all the torsion loads, while in reality, the sidewalls
of the fuselage at the ramp opening will also carry torsion loads. Last, the assumption was made that the
skin carries all the torsion loads, but in reality, the stiffeners will also carry some of the loads. Nevertheless
for safety and fatigue reasons the thickness is not reduced any further.

However, the thicknesses found during these calculations are not yet the minimal needed skin thickness to
keep the aircraft from failing under extreme operation. The minimal thicknesses generated this way are the
minimum needed thickness to resist torsion. These should be added to the minimum needed skin
thickness under shear loads (both loads could happen at the same time, so the shear flows should be
superimposed on each other), and the skin should be supported by the stiffeners for which also a minimal
area will be calculated to resist bending. These will be calculated in subsequent chapters.

Chapter 7 Airframe Structural Design 77


Final Report of the International Design Synthesis Exercise 2005

7.3.2 Bending of the fuselage


In this part the sizing of the stringers in the fuselage will be estimated. The task of the stringers is to take
up bending stresses resulted from bending moments. In reality the skin of the fuselage takes up bending
stresses as well, but to reduce complexity and save time for a first estimate, an assumption is made here.
The idealisation is made by assuming that all bending stresses are carried by the stringers (represented as
boom areas) while the skin is only effective in shear.

The most extreme load case for bending and shear occurs when the pilot makes an extreme manoeuvre
which produces the maximum allowable load factor of 2.5 g. Assuming that half of the MTOW is in the
wing and the other half is distributed along the fuselage length, a shear force diagram and a moment
bending moment diagram can be constructed. These can be found in Figure 7.3.4. From the shear force
diagram and the bending diagram the maximum shear force and the maximum bending moment can be
found which, after multiplying with the maximum load factor of 2.5 g and a safety factor of 1.5, gives a
reasonable estimate of the maximum shear force and bending moment which might be encountered
during operational life.

Figure 7.3.4 Load case of the aircraft with its shear force and bending moment diagram

78 Airframe Structural Design Chapter 7


Final Report of the International Design Synthesis Exercise 2005

The fuselage consists of 36 stringers distributed along the contours of the skin. The boom areas represent
the stringers, and from the engineering bending theory (simplified by symmetric crossed sections) the
bending stresses can be calculated with:

Mx
σz = y (7.3.5)
I xx

with

36
I xx = ∑ Ai y i (7.3.6)
i =1

By choosing sufficiently large boom areas, the bending moment of inertia increases and results in a
decreased bending stress, which must be below the yield stress of the material. This results in boom areas
as shown in Figure 7.3.5 below for AL7075.

Figure 7.3.5 Structural modal of the bending analysis for the fuselage

The type of stringers used in the fuselage is called Z-stringers. These are commonly used in aircraft. Hat-
type stringers are better in carrying loads but void can concentrate at the edges. The Z-type stringers are
positioned in a way that moisture will not concentrate at the edges. This is illustrated in Figure 7.3.6.

Chapter 7 Airframe Structural Design 79


Final Report of the International Design Synthesis Exercise 2005

Figure 7.3.6 Fuselage skin with Z-type stringers

The material used for the production of these stringers is AL 7075-T6. This alloy is often used for
stringers since its tensile strength is rather high. However, 7xxx alloys exhibit reduced stress-corrosion
resistance, which means frequent maintenance is required. Under normal conditions the stress on the
stringers will be about 35% of the yield stress (taking gust loading into account during flight). With this
alloy the total mass amounts 1850 kg and material cost $44,250. See Table 7.3.2 below where a
comparison is made with AL2024-T3.

Table 7.3.2 Cost comparison


AL2024-T3 AL7075-T76
Mass (kg) 2700 1850
Material cost ($) 65,000. - 44,250. -

To prevent buckling of the stringers, frames will be used. These frames are located with a distance of 60
cm from each other in the front and rear fuselage (this is a traditional distance of frames in commercial
aircraft). In the centre fuselage the frame distance will become 50 cm since the landing gear will be
suspended to these frames. This means that frames in the centre fuselage will get thicker (and thus
heavier) since higher loads must be carried here.

7.3.3 Shear loading of the fuselage


In this section the minimum thickness of the fuselage skin will be estimated due to the maximum shear
loads acting on the centre fuselage. The maximum shear load can be found from the shear force diagram
in Figure 7.3.4 above. To make a first estimate it is again assumed that the skin only carries shear (and not
some of the bending stresses). It is also assumed that the shear load acts in the middle of the fuselage.

The fuselage is modelled as a thin walled single cell structure with same boom areas as found for the
bending stresses. The floor is removed from the model to reduce the amount of calculations. This reduces
the model from a multi cell to a single cell structure. For this model the engineering bending theory for
shear loads is used to calculate the shear flow acting on the skin. The equation for this is shown below.

Sy 36
qs = −
I xx
∑A y
i =1
i i + q s ,0 (7.3.7)

80 Airframe Structural Design Chapter 7


Final Report of the International Design Synthesis Exercise 2005

The closed sectioned fuselage first needs to be opened with a ‘cut’. Then the shear flows are calculated as
for an open sectioned structure. After calculating these shear flows the structure will then be closed again
by introducing a circular shear flow. This circular shear flow can be found with a moment equation at the
centre of the fuselage. To get the final shear flows around the fuselage, the circular shear flow must be
added to the shear flows calculated for the open section. The result of this can be seen in Figure 7.3.7.

Figure 7.3.7 Structural model of the shear force analysis for the fuselage

By taking the maximum shear flow (near the middle), the minimum thickness of the skin can be
determined at the centre fuselage. This is done for two aluminium alloys as can be seen from the table
below. It must be mentioned that the maximum shear load only occurs at the centre of the fuselage. This
means that skin thickness (and thus weight) can be reduced at front and rear fuselage.

Table 7.3.3 Minimum skin thickness due to maximum shear loading at the centre fuselage
AL2024-T3 AL7075-T76
Thickness (mm) 0.48 0.41
Mass (kg) 820 710
Material cost ($) 19,650. - 17,000. -

From cost perspective it does not matter which material to choose since they are close to each other.
However the strain rate of AL2024 is 18% with respect to 11% of AL7075. This means that production of
parts can be done easier (thus cheaper) with AL2024. But a more important advantage is the crack
corrosion resistance of AL2024, which is higher than AL7075. However extreme torsion (as calculated

Chapter 7 Airframe Structural Design 81


Final Report of the International Design Synthesis Exercise 2005

above) can occur in combination with shear. This means that the shear flows of both load cases (by super
imposing) must be added to determine the minimum thickness of the skin. Thickness results from
torsional loading of the fuselage show much larger differences between the both alloys as can be seen
from Table 7.3.3 in the torsion section. So using AL7075 will give reduction in material cost ($15,000), but
more important is the reduction in weight (600kg). And since the fuselage is non-pressurised, fatigue does
not play a major role as it does in commercial jets. That’s why AL7075 will be selected above AL2024 for
the skin of the fuselage. The weight and cost of both alloys for the total fuselage skin can be seen below in
Table 7.3.4.

Table 7.3.4 Total mass and material cost of the fuselage skin for combined shear and torsional loading
AL2024-T3 AL7075-T76
Mass (kg) 4920 4330
Material cost ($) 120,000. - 105,000. -

7.4 Empennage

This section presents the structural layout and design of the empennage. The methods used for analysis
are almost identical to those laid out in section 7.2 with a few minor adjustments. Some major
assumptions have been made which means that the results from the analysis are likely to be quite
conservative. However, the results do provide a useful starting point for the next stage in design.

7.4.1 Horizontal tail: assumptions


For the horizontal tail structural design, the same approach as for the wing was taken. As both structures
are very similar and have the similar function this assumption is acceptable. The same wing box approach
with stringers, spars and skin tapering along the span was taken. Therefore, practically all the assumptions
from section 7.2.3 are valid. The loading (in comparison with the wing) was now negative because the
horizontal tail plane has to produce negative lift to trim the aircraft. The loading was idealised as elliptical
were the CLmax and the maximum CM of the aerofoil were considered. Loading diagrams were produced
for the horizontal tail from the spreadsheet used for the wing loading analysis. However, this time, the
effect of the engines was removed, as was the effect of fuel in the tanks since this does not apply to the
horizontal tail. In addition, the location of the spars was changed to 25% and 65%. The stringer cross-
section chosen was also changed to a simple L shape (see later). Finally, 18 ribs were used on the
structure as specified by the skin buckling conditions (see previous section).

7.4.2 Horizontal tail: results and discussion


The analysis was performed for three different positions along the span of the horizontal tail. The output
from this analysis can be found in Table 7.4.1.

82 Airframe Structural Design Chapter 7


Final Report of the International Design Synthesis Exercise 2005

Table 7.4.1 Results for horizontal tail analysis


Spars Stringers U/L Skin
Area F Web R Web T Area Height Thickness Thickness
Case x Number
(m2) T (m) (m) (m2) (m) (m) (m)
1 0 2.04E-03 0.004 0.0045 16 1.41E-04 0.025 0.003 0.005
2 3.26 8.06E-04 0.0035 0.004 13 9.60E-05 0.025 0.002 0.003
3 7.26 2.39E-05 0.003 0.0025 8 3.94E-05 0.025 0.0008 0.001

Spar Caps
When distributing the area of the spar caps, the double L structure used in the wing design was considered
suitable for this design. Unlike for the case of the wing, the double L structure does not need an extra
layer to prevent large changes in section and thus stress concentrations. Again the L structures follow the
bend pattern identified for the wing analysis. For the root case, the height of the spar cap is 25 cm with a
thickness of 1 cm. Figure 7.4.1 shows an overall view of the profile of the horizontal tail at the root,
followed by Figure 7.4.2, which shows a close up of the spar cap.

Figure 7.4.1 Overview of horizontal tail structure

Figure 7.4.2 Close up of horizontal tail spar cap

Stringers
In the analysis for the wing, Z shape stringers were chosen as they provided the required area at a reduced
section thickness. In this case, the Z shape stringer provides an extremely small thickness for a certain
area. Therefore, the stringer selection was changed (as indicated at the start of this section) to a simple L
profile. The simple shape provides adequate area without the change in section problems which would
have been encountered had this section been used for the wing. From Table 7.4.1.1, it can be seen that at
the root, the stringers have a thickness of 3 mm and a height of 2.5 cm.

Chapter 7 Airframe Structural Design 83


Final Report of the International Design Synthesis Exercise 2005

Figure 7.4.3 Stringer close up

Figure 7.4.3 shows the L stringers at a spacing of 0.11 m, again with filleted edges rather than sharp
corners again to lessen the stress concentrations. Like the Z shape stringers, these would typically be
riveted to the skin.

Spar Web
Again, in a similar fashion to the wing, the front spar web required a stiffener to be placed between the
ribs for the root and mid-span sections but not for the outer position.

The spreadsheets for the horizontal tail structural analysis can be found on the CD in Appendix 7.

7.4.3 Vertical tail: assumptions


From section 10 it can be seen that the vertical tail (fin) has a symmetrical aerofoil section. This implies
that during steady level flight the fin has only a small load on its structure. Whilst this is true in part,
during disturbances or manoeuvres the fin washes through the air. Disturbances and manoeuvres can
cause extreme loads on this structure. For the purposes of this analysis, the loading on the fin will be
computed assuming the HERA is flying at maximum cruise speed with an instantaneous rudder deflection
of 30 degrees. Following the example of the previous analyses, these loads were multiplied by the ultimate
load factor of 3.75. One difference between this analysis and that followed for the horizontal tail, is that
the weight of the fin itself does not act in such a way as to relieve the torsional loads on the structure,
therefore this effect is removed from the spreadsheet (see CD-ROM in Appendix 7, “VtailSolution”).

Due to the large rudder (50% chord), the spars were placed at 0.15 and 0.45 of the local chord (5% margin
allows for control linkages etc). As for the horizontal tail, L-shaped stringers were again used for the fin
design (reasons in previous section apply).

7.4.4 Vertical tail: results and discussion


Table 7.4.2 below presents the results for the fin analysis

84 Airframe Structural Design Chapter 7


Final Report of the International Design Synthesis Exercise 2005

Table 7.4.2 Results of the fin structural analysis


Spars Stringers U/L Skin
Area F Web R Web Height Thickness Thickness
Case x Number Area (m2)
(m2) T (m) T (m) (m) (m) (m)
1 0 4.70E-03 0.005 0.0055 13 2.19E-04 0.033 0.0035 0.0045
2 4.67 1.59E-04 0.004 0.0045 9 9.68E-05 0.033 0.0015 0.0020
3 9.67 2.84E-05 0.003 0.0025 6 2.98E-05 0.030 0.0005 0.0010

Spar Caps
Following the example of the horizontal tail, the spar caps for the fin will consist of ‘bent’ double L
shapes. For the case at the root, the spar caps have a height of 12 cm with a thickness of 1 cm. Figure
7.4.4 shows the torsion box formed by the spar caps in relation to the complete root chord. The effect of
the large rudder can clearly be easily identified.

Figure 7.4.4 Overview of fin structure

Figure 7.4.5 Close up of fin spar cap

Stringers
As mentioned previously, the fin also adopts the L shaped stringers. In the case of the root section, the
stringers have a thickness of 3.5 mm and a height of 3.3 cm. Figure 7.4.6 shows these stringers spaced at
0.10 m intervals and with filleted edges.

Figure 7.4.6 Stringer close up

Chapter 7 Airframe Structural Design 85


Final Report of the International Design Synthesis Exercise 2005

Spar Web
Again, in a similar fashion to the horizontal tail, the front spar web required a stiffener to be placed
between the ribs for the root and mid-span sections but not for the outer position. The spreadsheets for
the horizontal tail structural analysis can be found on the CD in Appendix 7.

7.5 Cargo ramp

The general layout of the backside of the aircraft is mostly determined by the requirements of the loading
of the light vehicles, the medical unit and the pallets. A small loading angle will ease the loading but it will
increase the mass needed for structural stiffness. The opening must be high enough to load the medical
unit. That is why a splitting of the rear door in an upper and a lower half is chosen. The upper half
consists of two parts opening along the longitudinal axis of the airplane. The requirement for a minimum
cargo box height of 3.9 m is lowered to 3.5 m, since the requirement came from an ability to load a truck,
which is not the case for the HERA. The manufacturer is able to produce a medical unit that fits in the
airplane, so the height of the cargo compartment is never a problem. The height of the medical unit is
assumed to be 2.5 meters. The undercarriage determines the height of the floor. The floor height together
with the ramp length determines the loading angle. A sketch of the backside is shown in figure 7.5.1.

Figure 7.5.1 Configuration of backside of HERA

7.5.1 Layout of the ramp


The main objective of the cargo ramp is to ease the on- and off loading of cargo in and out of its cargo
bay. The three most important requirements that the ramp must fulfil are:
• Light vehicles
• Medical unit
• Pallets

86 Airframe Structural Design Chapter 7


Final Report of the International Design Synthesis Exercise 2005

The first two are able to drive in and out of the aircraft, while the last must be loaded using rollers (and a
winch when being on unprepared runways). The ground track of the light vehicles and the medical unit
are approximately 1.5 m and 2.5 m, respectively. This means that from the centreline in both directions at
a distance of 0.75 and 1.25 meters, the ramp should be able to distribute the loads of the light vehicles and
the medical unit, see figure 7.5.2.

Figure 7.5.2 Ground track of wheels on ramp

To easily load the pallets, 2 directional rollers on the ramp are used. To simplify the structural design of
the ramp these loads (of pallets with a maximum width of approximately 3 m) should be distributed at the
same distances as for the tires of the vehicles and the medical unit, so at distances 0.75 and 1.25 meters
from the centreline. To decrease the bending moment in the pallets generated by the large distance
between the inner rollers a fifth roller system is introduced at the centreline. This will divide the ramp (of
a total width of 3.5 meters) into six parts, two inner parts of 0.75 m and four of 0.5 m, see figure 7.5.3.
Knowing the rollers will be in the way when loading the light vehicles or the medical unit, these can be
removed and reattached by a locking system.

Figure 7.5.3 Pallet positioned on ramp

Another aspect of the ramp is that when loaded in flight, the effective loading area of the cargo bay is
increased. This means that the fuselage length will be smaller to transport the same cargo. This loading
area will be used to put pallets of approximately 5000 kg. This means the ramp has to be locked along its

Chapter 7 Airframe Structural Design 87


Final Report of the International Design Synthesis Exercise 2005

side to distribute the loads to the surrounding structure. When locking the ramp from movement in the
lateral axis of the plane, the ramp will be part of the structural stiffness of the tail. Important is that the
pneumatic suspension system to move the ramp, which is subjected ‘groundside’, must be able to move
when loaded by a pallet of 5000 kg. Knowing the pallet’s centre of gravity is approximately halfway of the
ramp and the suspension at the end, it means the suspension system has to move a load of 2500 kg. The
ramp must also possess rings and lockers to lock the cargo to the floor.

The last aspect is that the ramp can be used to perform airdrops during flight. The part around the ramp
that connects the tail to the fuselage is designed in such a way that airdrops during flight of medical and
food packages can be performed. To do this in a safe way, a locker is introduced which disables the ramp
to open further than parallel to the cargo floor, see figure 7.4. This locker is also used when loading the
aircraft on prepared runways, where external lifters are used.

Figure 7.5.4 Locker makes sure cargo floor and ramp stay parallel

7.5.2 Structural analysis of the ramp


The easiest way to calculate the ramp is to look at every wheel track or roller system along the ramp
separately. To distribute the loads in these paths the assumption is made to use I-beams. The
configuration of the ramp is given in figure 7.5.5.

Figure 7.5.5 Ramp cross section

88 Airframe Structural Design Chapter 7


Final Report of the International Design Synthesis Exercise 2005

When considering the I-beams, the assumption for calculations is made that the horizontal flanges will
distribute the bending stresses and the vertical web will deal with the shear stresses.

The way to solve this is to assume that a force P loads every I-beam in the middle, see figure 7.5.6.

Figure 7.5.6 Load case on I-beam

Now the free-body-diagram for the system can be produced together with a cut halfway of the beam and
introducing a vertical force, a horizontal force and a moment, see figure 7.7.

Figure 7.5.7 Free-body-diagram I-beam

From figure 7.5.7 and the theory used in the book ‘Mechanics of materials’ by Gere and Timoshenko, the
bending moment M and the shear force S can be calculated. Now the normal stresses in the flanges of the
I-beams can be calculated by using equation 7.5.1 in which If represents the moments of inertia of the
flanges. The shear stresses in the web of the I-beams can be derived using equation 7.5.2, in which AW
represents the Area of the web.

My
σ= (7.5.1)
If

S
τ= (7.5.2)
AW

Chapter 7 Airframe Structural Design 89


Final Report of the International Design Synthesis Exercise 2005

With these formulas the normal and shear stresses can be optimised in such a way that the normal stresses
will not exceed the normal yield stress σyield. The same will go for the shear stresses, which should not
exceed the shear yield stress τyield.

Only thing left is the maximum loading that is subjected to every beam. Now during on- and off loading
on the ground, according to JAR, a safety factor of 1.5 should be applied. In flight the loading on the
ramp can experience a maximum acceleration of 2.5 g. When multiplying this, it gives a safety factor of
3.75. In flight the only loading is one 3610 pallet. The assumption is made that the load is evenly
distributed over the five beams. So 1000 kg each, times the safety factor of 3.75, will be the highest load
case during flight. When loading on the ground the highest forces generated at the outer beams will be by
the medical unit. This will be ¼*30000 kg times a safety factor of 1.5. Assumed is ¼, since it is only one
wheel that will be on the beam. The maximum load created by the pallet will be a pallet placed in length
direction, so spread over three roller systems. This will give a loading of 1667 kg times a safety factor of
1.5. The light vehicle will give a maximum loading of 4000, which gives a maximum loading per two
wheels of 2000 kg, this again times a safety factor of 1.5. When comparing the loads in flight and on the
ground the maximum loads can be determined for every I-beam. The results are given in figure 7.5.8.

With all this data, the moment of inertia of the flanges and the area of the web can be optimised and the
total mass can be minimised.

Figure 7.5.8 Maximum loads to be distributed by the ramp

The results of this optimisation are given AL 7075 in table 7.5.1. This material has been chosen above AL
2024 because cost is one of the most important aspects. The amount of material needed for the ramp will
be less for AL 7075 due to its material properties and since the costs for both materials are merely the
same, the costs for the ramp will be lower than for AL 2024. The ramp is highly exposed to cyclic loading.
Because AL 7075 is more sensitive to fatigue (than AL 2024), it needs more inspection. This requests
more maintenance.

90 Airframe Structural Design Chapter 7


Final Report of the International Design Synthesis Exercise 2005

Table 7.5.1 Optimisation results for cargo ramp


Two outer beams Three inner beams
Thickness of flange 1.7 cm 1.1 cm
Thickness of the web 1.7 cm 1.1 cm
Width of the flange 10 cm 5 cm
Deflection 2.26 cm 1.87 cm
Weight 76.4 kg 37.1 kg

The total weight of the five I-beams will come down to 264.1 kg. The areas of the flanges can be spread
out in horizontal direction to carry the loads. These thicknesses are calculated by assuming a worst-case
scenario. Also the point loads are, in reality, distributed loads and the surrounding structure will help
carrying the load. The ramp will be a closed box, so when the extra stiffened sheet material is taken into
account, the total weight of the ramp will be approximately 360 kg.

Since the ramp had a thickness of 0.2 m, there is a gap between the topside of the ramp and the ground
when lowering the ramp. To bridge this distance, a so-called ‘ramplet’ is used. An extra option is to use a
‘support device’ to put under the groundside of the ramp. This will lower the loading angle. Only thing is,
that longer ramplets are needed to bridge the ramp-floor distance, see figure 7.5.8.

Chapter 7 Airframe Structural Design 91


Final Report of the International Design Synthesis Exercise 2005

8 Undercarriage

The undercarriage is an important part of the HERA, due to the requirement to operate from both
prepared and unprepared runways with primitive facilities. First the general layout resulting from the Mid
Term Report is given. Next, the location on the HERA is calculated and the tires, shock absorbers and
brakes are selected. Then, the design is presented and evaluated.

8.1 General layout

In the Mid Term Report it is stated that a tricycle layout would suit the HERA best. Further, the chosen
configuration is the fixed undercarriage with a moveable shelter.

8.2 Location on the HERA

From Chapter 9 the Centre of Gravity analysis:

Table 8.2.1 C.o.G. locations


Aircraft Empty Mass xAFT = 19.34 m zAFT = 3.53 m
Zero Fuel Weight xFRONT = 17.43 m zFRONT = 3.16 m
Maximum Take-off Weight xMTOW = 17.57 m zMTOW = 3.60 m

8.2.1 Steering
From Ref. [32], pg. 14, for adequate steering a minimum normal force must act on the nose gear so the
appropriate levels of friction forces needed for steering can be generated.

Fm 0.92 Fm
= ∠11.5 (8.2.1)
Fn 0.08 Fn

where:
Fm = Loading on main landing gear

Fn = Loading on nose landing gear

From Chapter 9 the Centre of Gravity analysis, the following first estimates were generated:

Chapter 8 Undercarriage 93
Final Report of the International Design Synthesis Exercise 2005

Table 8.2.2 Landing gear locations


xn 3.84 m
xm 21.28 m

Though these are merely first estimates, for now, these are used. Check on steering:

Fn (17.57 − 3.84) = Fm (21.28 − 17.57 )


(8.2.2)
Fn + Fm = 101124

Fn (17.43 − 3.84) = Fm (21.28 − 17.43)


(8.2.3)
Fn + Fm = 76929

Fn (19.34 − 3.84) = Fm (21.28 − 19.43)


(8.2.4)
Fn + Fm = 40602

Figure 8.2.1 Taxi load cases

Table 8.2.3 Loading ratios


Fm
Aircraft Empty Mass = 7.99 ∠11.5
Fn
Fm
Zero Fuel Weight = 3.60 ∠11.5
Fn
Fm
Maximum Take-off Weight = 3.70 ∠11.5
Fn

This leads to the conclusion that steering is possible, using the current location of the nose and main
landing gear.

8.2.2 Longitudinal Tip-over


From ref. [30], pg. 218: For a tricycle configuration, the main landing gear should be positioned behind
the most aft position of the centre of gravity location.

x AFT ∠xm (8.2.5)

94 Undercarriage Chapter 8
Final Report of the International Design Synthesis Exercise 2005

8.2.3 Lateral Tip-over


From ref. [30], pg. 219: Given the most forward position of the centre of gravity, the angle ψ determines
lateral stability while on the ground, see Figure 8.2.2.

Figure 8.2.2 Lateral Tip-over

From stability requirement ψ ≤ 55°, take ψ = 55° to determine minimum ground tracks. Calculations
result in a minimum ground track of 5.76 m.

8.2.4 Ground clearance


From ref. [30], pg. 221: Both longitudinal and lateral ground clearance oppose no problem, since the
HERA has a high tail and a high wing configuration.

8.3 Tire-selection

From ref. [28], pg. 273, a four-wheel bogey would be the logical choice for the HERA. In order to reduce
our frontal area, since the landing gear is fixed, put the wheels behind one another. This gave a total of 10
wheels. Use formula (8.3.1) where the HERA and Hercules are compared to determine the size of the tires
of the main landing gear. The Hercules is used since this aircraft also has the requirement to be able to
land and take-off from unprepared runways.

2
⎛ MTOWHERA ⎞⎛ DHercules ⎞
N HERA = N Hercules ⋅ ⎜⎜ ⎟⎟⎜⎜ ⎟⎟ (8.3.1)
⎝ MTOW Hercules ⎠⎝ D HERA ⎠

2
Here: 10 = 6 ⋅ ⎛⎜ 101124 ⎞⎟⎛⎜ 56 ⎞⎟ results in a diameter for the HERA-tires of 52”~ 1.32 m.
⎜ ⎟
⎝ 70380 ⎠⎝ DHERA ⎠

For the HERA, radial-ply tires are preferred. Compared to bias-ply tires, these are less stiff, have lower
cornering force, higher footprint, higher durability, run cooler, lower mass, lower wheel stresses, less

Chapter 8 Undercarriage 95
Final Report of the International Design Synthesis Exercise 2005

rolling resistance and higher cut resistance. Using ref. [51] to check whether there are radial-ply tires of
this size already in production, in turns out that tires with 52x21 R22 are produced for the B777. Since a
few years, the combination of radial technology and new materials led to a new type of tires: the Near
Zero Growth. These are more resistant to punctures and loss of pressure and are therefore more
sustainable. This new type of tire is used for the HERA.

Since the HERA must be able to land on both prepared and unprepared runways, the tire inflation
pressure is chosen to be around 3.2 kg/cm2. This enables the HERA to land on grass, see ref. [37], pg.
344. Normal, grooved NZG tires, provided with ribs, provide good adhesion on wet runways and
minimise the effect of cutting action of stones and flints and thus account for stones on unprepared
runways. The requirements on the shock absorber are high to enable the aircraft to also land on prepared,
hard, runways.

8.4 Shock absorber

There are different options for the shock absorber. These are:
• Steel coil springs and ring springs
• Steel leaf spring
• Rubber spring
• Air
• Oil
• Internally Sprung wheels
• Oleo-pneumatic

The oleo-pneumatic shock absorber is chosen, since they have the highest efficiency of all types and the
best energy dissipation, see ref. [5], pg. 75. The use of double acting struts will improve the shock-
absorption characteristics during taxi-conditions over rough or unpaved fields.

Figure 8.4.1 Layout of double acting strut

96 Undercarriage Chapter 8
Final Report of the International Design Synthesis Exercise 2005

A first assumption for the compression ratios for large cargo aircraft can be:
• Static to extended: 4/1
• Compressed to static: 3/1

The static loading has to be determined:

Table 8.4.1 Strut loading


Main gear Nose gear Per main strut Per nose strut
MTOW 79048 kg 21364 kg 9881 kg 21364 kg
ZFM 60205 kg 16723 kg 7526 kg 16723 kg
AEW 36084 kg 4516 kg 4511 kg 4516 kg

The maximum loading per strut, results in the following:


Load static = 21364 kg
Load extended = ¼ . 21364 kg = 5341 kg
Load compressed = 3. 21364 kg = 64092 kg

Since HERA’s MTOW is comparable to B720B, the static to compressed distance of the B720B is used:
total stroke: 20” ~ 0.508 m and static to compressed: 15 %.

Table 8.4.2 Load-Stroke


Load Stroke
5341 0
21364 0.4318
64092 0.508

From the strut loading the diameter of the strut can be determined using ref. [37], pg. 361.

D = 1.3 + 0.11 Ps (8.4.1)

where Ps is the maximum unfactored vertical load per leg. Here: 1.3 + 0.11 21364 = 0.18 cm.

8.5 Brakes

The use of carbon as heat sink material is chosen since carbon has a very low thermal expansion resulting
in faster and more uniform heat transfer. Other options were: steel, beryllium and Teflon. Beryllium oxide
is toxic, so handling beryllium needs special precautions. This is not an option, since HERA must be able

Chapter 8 Undercarriage 97
Final Report of the International Design Synthesis Exercise 2005

to operate from airfields with primitive conditions. Teflon has a low wear resistance at high temperatures,
which makes it unattractive for the HERA. Steel brakes have a high structural weight, due to difference in
density. Steel: 7833 kg/m3 and carbon 1688 kg/m3. Though carbon brakes are more expensive, it seems
the best option considering the low weight and performance at high temperatures.

8.6 Landing gear design

Main characteristics: A symmetric piece on where the wheels are attached to the struts, see Appendix 10.1,
part A. Though usually the nose gear is smaller than the main gear, HERA’s nose is relatively highly
loaded. Therefore using the symmetric connection, making the nose and main interchangeable, only
results in a small mass-penalty. This increases the maintainability since there are less different parts and
therefore also decreases production-costs. Part B provides rigidity.

8.7 Moveable shelter

First determine the maximum total deflection:


Maximum tire deflection: a 49% deflection is desirable; see ref. [5], pg. 127. The tire deflects to rolling
radius, see ref. [28], pg. 286. So, given the radius of 0.66 m, the deflection is: 0.66 –(0.49*0.66)/3=0.11 m.
The rolling radius is thus: 0.55 m. Maximum shock absorber deflection: 0.51 m. A total ground clearance
of at least 0.62 m is required. For the shelter, the following idea was generated:

Figure 8.7.1 Moveable shelter design

At A: a roller with a piston serving as an opening mechanism.


At B: (magnetic) locks to decrease the vibration of the doors, when opened.
At C: jettison locks.

98 Undercarriage Chapter 8
Final Report of the International Design Synthesis Exercise 2005

Both sides of the shelter had to move outward because the minimum ground clearance left little room
under the belly of the HERA, only 0.38 m. The control lies in the cockpit, but if this fails, there is hand-
control in the cargo bay. In case this fails as well, the shelter can be ejected using the jettison locks. For
the sizing, clearance allowances are needed between the tire and adjacent parts of the aircraft e.g. strut and
the shelter. These clearances are based on: the maximum dimension of the inflated tire, the effect of
centrifugal forces at high-speed rolling increasing the diameter and a growth allowance due to service. The
latter is small since HERA uses NZG tires. For the calculation of the clearance, ref. [37], pg. 361 is used:
• Clearance in radial direction: 0.12 D = 0.12x1.32 = 0.16 m.
• Clearance in normal direction: 0.18 W = 0.18x0.53 = 0.10 m.
• In total this design gave a frontal area of 3.66 m2 and a total length of 6.90 m.

8.8 Evaluation

As stated above, there is only 0.38 m height left between the belly and the ground to store part of the
shelter so it is now fully deflected outward. This resulted in a design not yet existing. This implies a lot of
effort has to be put in the design, development and testing, meaning high costs and more development
time. The maintainability is also worse than first expected since there are several moving parts after all.
There is now a shift in the trade-off and the option of having a semi-retractable landing gear might now
be the best option.

8.9 Landing gear design

The landing gear design had the simplicity of the Hercules undercarriage serving as example. So for the
retraction system, look at the retraction system of the Hercules. A layout sketch has been provided for in
Appendix 10.2. The location on the aircraft and loading are comparable. The range of applicability of the
HERA would not propose a problem to a Hercules-landing gear, since the Hercules also has a high range
of applicability. According to the Maintenance Department of the Netherlands Royal Airforce Base
Eindhoven, the undercarriage of the Hercules does not require much maintenance. Since this is proven
flight design, the development time is rather short and the testing required is minimal. The landing gear is
controlled from the cockpit but if necessary there is manual control in the cargo bay, which unlocks the
mechanism and deploys the landing gear. If this fails, there is manual control to help the deployment.
Both these manual controls are located at part C.

8.10 Shelter

Since a landing gear similar to the Hercules landing gear is used, first look at Hercules’ shelter. Compared
to the fixed landing gear, the shelter can now be located higher. According to the Maintenance

Chapter 8 Undercarriage 99
Final Report of the International Design Synthesis Exercise 2005

Department of the Netherlands Royal Airforce Base Eindhoven, maintenance of the shelter was easy. The
opening-close system is also simplified compared to the first design. In total this design gave a frontal area
of 1.90m2 and a length of 6.70 m. The design can be found in Appendix 10.3.

This semi-retractable system seems the best choice. However, if taken into account the amount of time
needed for development and testing, it will probably not be before 2012 that this is ready for flight.

8.11 Landing gear-fuselage connection

The loads applied to the landing gear are guided upward through the struts. The struts are connected to
frames. These frames are able to carry the loading coming from the landing gear. These ‘landing gear
frames’ have a spacing of 1.65 m. For the complete design of the fuselage frames, see chapter 7.3. For the
connection part, see Appendix 10.4.

8.12 Weight estimation

From ref. [5], pg. 259, method I, the weight of the landing gear can be determined using formula (8.12.1):

Landing gear weight = 0.046x(KSL+KRF)xWL (8.12.1)

Where: KSL = strut length coefficient = 0.85 (short gears)


KRF = rough field capability coefficient = 0.15
WL = aircraft design landing weight = 0.85xMTOW = 0.85x101124 kg

From ref. [5], pg. 261, method II, determine the weight of the landing gear, using formula (8.12.2):

1.17
Landing gear weight = 20.45 ⋅ K CG ⎛⎜ WL ⎞⎟ (8.12.2)
⎝ 1000 ⎠

Where: KCG = chronological factor depending on runway, mount type and year of design = 0.81

The results of the two methods are given in Table 8.12.1 below.

Table 8.12.1 Landing gear weight


Method I Method II
Nose 791 kg 607 kg
Main 3163 kg 2429 kg
Total 3954 kg 3036 kg

100 Undercarriage Chapter 8


Final Report of the International Design Synthesis Exercise 2005

9 Weight and Balance

In the process of developing a conceptual design for the midterm report, it was observed early on that the
accurate determination of aircraft mass and the control of that mass, would be a critical design driver. This
was due to the severe mission requirements and availability of a suitable engine. Perhaps the second
critical consequence of the aircraft mass is the effect on unit purchase cost and operating costs. The
position of the aircraft in the market place will typically be determined by sale cost per unit payload and a
favourable ratio over competitors can make or break a design. An accurate determination of the aircraft
mass is extremely important.

The present chapter sets out the method and final results of class one mass estimate, a detailed mass
estimate and the location of the aircraft centres of gravity for the pertinent loading conditions. The class
one mass estimate takes provided component mass estimates from similarly configured aircraft and
averages the values. These average values are then compared with the design mass estimate and any
difference in the sum is allocated to the components on a basis of the ratio of the component mass to the
basic empty weight. Based on the component geometry, component centres of gravity are estimated along
with their relative positions. From this data an aircraft mass statement is generated along with a centre of
gravity excursion diagram. The Excel Spread Sheets presenting the actual calculations can be found in the
accompanying CD in Appendix 7 in the file marked “Weight and Centre of Gravity”.

9.1 Class one component mass estimation

This method uses what is essentially empirical data from similar aircraft to provide a first estimation of the
component masses. This information was provided courtesy of Stanford University and can be found in
the CD file for this chapter. One draw back with this information was that it was supplied in US units,
however as the mass data is being non-dimensionalised as part of the estimation process, it is of very little
consequence.

The component masses are then averaged and allocated to the corresponding components on the HERA.
This provides the values of the first estimate. From this an empty weight is calculated and compared with
the empty weight from the parametric sizing exercise. The difference is then re-allocated on a basis of the
ratio of the component mass to the empty weight of the aircraft. This results in the second mass estimate.
See table 9.1.1 for clarification. The second mass estimate forms the benchmark for the detailed mass
estimate.

Chapter 9 Weight and Balance 101


Final Report of the International Design Synthesis Exercise 2005

Table 9.1.1 Class One Mass Statement


Component 1st Estimate (kg) Adjustment (kg) 2nd Estimate (kg)
Wing 9785.97 442.35 10228.32
Tail 2293.12 103.65 2396.77
Body 10277.51 464.57 10742.08
Landing Gear 3878.94 175.34 4054.28
Nacelle 1741.62 78.73 1820.35
Engine Systems 4200.68 189.88 4390.56
Flight Controls 898.48 40.61 939.10
Aux. Power System 314.45 14.21 328.66
Instruments 313.36 14.16 327.52
Hydraulics 659.76 29.82 689.58
Electrics 1200.38 54.26 1254.64
Avionics 1346.23 60.85 1407.09
Equipment 2009.27 90.82 2100.10
Air Con 647.13 29.25 676.38
Anti-Icing 469.54 21.22 490.77
Load Handling 34.73 1.57 36.30
Engines 3437.50 155.38 3592.88
Component Fraction Initial Mass Estimates
Empty Weight 43509 Empty Weight 45475
Difference 1967
Fuel 24195 Fuel 24195
Crew + Payload 35300 Crew + Payload 35300
Max Takeoff
Max Takeoff Weight 103004 Weight 104970

9.2 Detailed mass estimate

As more data becomes available it becomes possible to apply more sophisticated methods for aircraft
mass estimation. This data requires accurate aircraft geometries and the major construction materials to
calculate a mass per unit volume. With accurate masses provided for the propulsion group and the landing
gear, the job is reduced in scope; however it is still high in detail and provides the values needed for target
structural mass and data for the inertias needed in the stability analysis. The following section provides an
outline for the wing group, fuselage, and the equipment and tail mass estimates.

9.2.1 Wing group mass estimate


The wing group is defined as the basic structure of the wing plus the control surfaces. The following
equation (9.3.1) will provide an estimate of the wing group mass based on the volume. It is applicable to
aluminium structures and modern design and manufacturing processes.

0.4
⎛ R ⎞
⎜⎜1 − ⎟⎟
M (9.2.1)
M W = 0.021265(M TOM × N) × S 0.7819 × A(1 + λ ) × ⎝ TOM ⎠
0.4843 0.4

Λ×t /c

102 Weight and Balance Chapter 9


Final Report of the International Design Synthesis Exercise 2005

Values for the mass of the aircraft flaps are again based on a mass per unit dimension, but his time the
dimension is area. Assuming composite flap surfaces are affordable, the factor for mass can be reduced
from the value offered in Jenkinson [24] from 70 kg/m2 for complex all metal flaps to 15 kg/m2 for
simple composite flaps. An intermediate 25 kg/m2 was chosen for the mass estimation.

9.2.2 Fuselage mass estimate


The factors influencing fuselage mass depend on the configuration and role of the aircraft. As this is a
transport aircraft with a large rear door and strengthened floor, it would be expected that the fuselage
mass be high relative to other aircraft. However, as this aircraft is un-pressurised a large mass saving can
be made by not having to design to constrain a large pressure differential. The following equation is
provided by Howe [23].

(
M B = 0.039 2 L F D F V D )
1.5
(9.2.2)

The result of this calculation has to be augmented to account for the nature of the aircraft design. It needs
to be increased by 17% to account for the mass and structural discontinuity of the rear door and ramp. It
also contributes to the increased structural strengthening due to the fuselage-mounted landing gear. This
apparent mass spiral is capped by a reduction of 8% due to the un-pressurised fuselage (which is a
conservative estimate).

9.2.3 Tail mass estimate


An accurate evaluation of the tail structural mass is essential, not because it represents a significant
proportion of the aircrafts mass, but because it is located at the very back of the fuselage at the furthest
distance from the centre of gravity. The total tail mass is considered to be the structural mass of the
horizontal and vertical stabilisers and the respective control surfaces. The mass is again calculated using a
mass per unit area set based on values suggested in Jenkinson [24]. The values chosen were; kH = 25
kg/m2 and kV = 28 kg/m2.

M T = M H + M V = S H k H + SV kV (9.2.3)

9.2.4 Operational mass items and fixed equipment


As with so many of the other mass estimates, the operational mass items calculation is highly dependent
on the configuration and the role of the aircraft. The data provided in Jenkinson [24] is not entirely
suitable for the HERA, but there are CS requirements for carrying passengers and the safety equipment
needed to do so. As such the HERA is required to carry; approximately 2.27 kg per passenger of basic

Chapter 9 Weight and Balance 103


Final Report of the International Design Synthesis Exercise 2005

food and water; 1.36 kg per passenger of toileting provisions and 3.4 kg of safety equipment (active such
as inflation vests and passive such as hand rails). The total fixed equipment mass is dependant on aircraft
type and a fairly high percentage of maximum take off mass can be assumed. This value was set at 8%.

9.2.5 Total propulsion mass, landing gear mass, payload and fuel mass
The values for total propulsion mass and landing gear mass were at first estimated. Propulsion mass was
determined from a statistical analysis of the power available from an engine and its total mass, including
associated equipment and the nacelles. Landing gear was calculated in a similar manner with 86% of the
total figure being assigned to the main undercarriage. In the latter stages of the design analysis it was
possible to get exact figures (see section 8.12) and these were substituted into the aircraft mass statement.
The fuel requirement was calculated by the propulsion group for the midterm report [8] and the payload is
stipulated in the mission requirements.

These mass estimates provided a mass breakdown and allows for the first aircraft mass statement to be
created, see Table 9.2.1. Basic Empty Weight, Operational Empty Mass, Maximum Zero Fuel Weight and
Maximum All Up Weight are the sum of the preceding results.

Table 9.2.1 Aircraft Mass Statement – Class 1


Wing Mass Estimate MW kg 8146
Tail Group Mass Estimate M_T kg 2898
Body Mass Estimate M_B kg 12303
Nacelle Mass Estimate M_N kg 1537
Landing Gear Mass Estimate M_UC kg 3977
Surface Control Mass
Estimate M_SC kg 1048
Total Structural Mass M_STR kg 28300
Propulsion Group M_PROP kg 10422
Fixed Equipment M_FE kg 4970
Basic Empty Weight M_E kg 43765
Op. Items Mass M_OP kg 1048
Crew Mass M_CREW kg 279
Op. Empty Mass M_OE kg 45093
Max. Zero Fuel Weight M_ZF kg 80093
Fuel Mass M_F kg 24195
Maximum All Up Weight M_TO kg 104288

9.3 Centre of gravity analysis

The analysis for the centre of gravity of the aircraft must provide results for all feasible loading conditions.
These results will have a direct impact on the final layout of the aircraft, the undercarriage design, and the

104 Weight and Balance Chapter 9


Final Report of the International Design Synthesis Exercise 2005

aircraft handling qualities linked with stability during flight. The method begins by estimating the centre of
gravity locations for the components listed in the aircraft mass statement. This is followed by the
moments being taken about the nose of the aircraft to locate the overall centre of gravity. Based on these
results, it is then possible to determine the range of centre of gravity locations in three orthogonal axes (x,
y, z). The range of travel with respect to the mean geometric chord is checked against statistical data and if
it is found that the configuration lies out side that range, the aircraft has to be re-configured. A centre of
gravity excursion diagram is then plotted.

9.3.1 Component centre of gravity


The methods for estimating the component centres of gravity vary depending on the component and the
axis being analysed. Beginning with the x-axis, several assumptions have to be made based on the
configuration of the aircraft to determine component centre of gravity relative to the fuselage length.
These assumptions for the major mass items are listed in table 9.3.1.

Table 9.3.1 Centre of gravity assumptions


Component x- Axis Centre of Gravity Location
Wing Group
0.1c aft of c 14 set to MTO C.o.G
Total Power Plant 0.45l f CosΛ 1 / 4
Fuselage and Fixed Equipment 0.45lf
Payload 0.47lf

Using this data the moments can be taken about the nose of the aircraft and the equation derived so as to
make the centre of gravity the subject.

⎡ ⎛ 0.45M PP ⎞⎤
c ⎢0.1⎜⎜ M SURF + M WING + M FUEL − ⎟⎥ +
⎢⎣ ⎝ CosΛ 1 / 4 ⎟⎠⎥⎦
x cg = (9.3.1)
M TO − M W − 1.1M MG − M SURF − M FUEL
l f [0.55(M FUSE + M OPIT + M PAY + M FIXED ) + 0.9M TAIL + 0.1(M NG + M CR )]
M TO − M W − 1.1M MG − M SURF − M FUEL

The aircraft was assumed to be symmetrical about the y-axis. By assuming the components are basic
shapes, the centres of gravity in the z-axis can be determined, see Roskam [30].

Chapter 9 Weight and Balance 105


Final Report of the International Design Synthesis Exercise 2005

9.3.2 Loading Conditions and Corresponding Centre of Gravity Location


Based on the result given by equation 9.3.1, it is possible to determine the centre of gravity locations for
all feasible loading scenarios. These conditions are listed in Appendix 11. The full range of data used for
these calculations can be found in the CD file in Appendix 7; however the basic method can be
determined from the table. Each component mass was allocated a number and the moments taken about
the nose of the aircraft to determine the centre of gravity of that configuration.

Working on an Excel sheet, it is possible to enter in a second estimate value of the maximum take off
centre of gravity. Looking back at table 9.3.1 is can be seen that the wings are placed relative to the aircraft
centre of gravity. By setting up a two cell iteration, it is possible to let the values of MTI centre of gravity
converge and more accurately locate the wings and other component masses.

9.3.3 Centre of gravity range criteria and excursion


The next step in this exercise is to ensure the centre of gravity travel for the different loading conditions is
within a specified range. This range is provided in Roskam [30] and is based on aircraft of a similar
configuration and role. The range given for large transport aircraft is anywhere between 60 cm and 205
cm. As can be seen from table 9.3.2, the HERA lies just within that range. The maximum and minimum
span is approximately 1.4 m. This implies that while the nose gear will have to be strengthened to cope
with the most forward condition, the aircraft constitutes a viable design when considering the likely
consequences for ground handling and stability during flight. These values are plotted in a Centre of
Gravity Excursion graph presented in Appendix 12. Also of importance is the vertical travel of the centre
of gravity. It can be seen in the graph that the maximum height of the centre of gravity is 3.98 m. This is
at a point nearly equal to the height of the fuselage and may present a tip over problem for fuselage
mounted landing gear. However as it was decided in the midterm report to place the landing gear inside
fairings, extending the effective fuselage width, the tip over problem should be minimal. This is discussed
in greater detail in the landing gear section of the report.

Table 9.3.2 Centre of Gravity Locations Relative to the Mean Aerodynamic Chord

MAC 18.00 m 22.74 m


Mass MAC - Location
kg -
Aircraft Empty Mass 43765 0.36
Op. Empty Mass 45093 0.32
Zero Fuel Mass 80093 0.07
Zero Payload Mass 69288 0.30
Max T-O Mass 104288 0.15

106 Weight and Balance Chapter 9


Final Report of the International Design Synthesis Exercise 2005

9.4 Mass moments of inertia

Once the mass of the parts are known and geometrically located on the aircraft, the inertia moments can
be calculated. These inertias will determine part of the stability properties of the HERA.

The inertia moments required for the stability are: Ixx, Iyy, Izz, Ixy, Iyz, Ixz. They can be computed with the
following formulae:

( )
I XX = ∫ y 2 + z 2 dm (9.4.1)
m

( )
I YY = ∫ x 2 + z 2 dm (9.4.2)
m

( )
I ZZ = ∫ x 2 + y 2 dm (9.4.3)
m

I XY = ∫ xy dm (9.4.4)
m

I XZ = ∫ xz dm (9.4.5)
m

I YZ = ∫ yz dm (9.4.6)
m

While some of the parts can be considered concentrated point masses, such as the engines, other masses,
such as that of fuel, are distributed over a certain area. For these masses a new term must be introduced,
the radius of gyration. This term takes into account the fact that the mass is distributed over a certain area
and so it replaces one of the coordinate values of the mass, which coordinate value depends on which
moment of inertia is being calculated. Using reference [2] the radius of gyration is determined as:

S
k= (9.4.7)
3

where S represents the length over which the mass is distributed.

In the case of the rolling moment of inertia S is the semi-span of the wing. The radius of gyration comes
thus to replace the y-coordinate in the inertia equation. The same method can be applied to calculate the
other inertias for distributed masses. Adding the inertias of the different masses gives the total inertias of
the whole aircraft with respect to its axes.

The results can be found in the spreadsheet “Weight and Balance” on the CD in Appendix 7. The aircraft
is symmetrical, so the Iyz and Ixy moments of inertia are zero.

Chapter 9 Weight and Balance 107


Final Report of the International Design Synthesis Exercise 2005

9.5 Conclusions

From the final mass and centre of gravity data it is possible to examine the feasibility of the design so far
and attempt to understand the implications for other sections of the report. Considering first the aircraft
mass statements it is possible to validate the mass estimated made in the midterm report. In the initial
sizing exercise a maximum take off weight was calculated to be 100,400 kg before a margin of error was
applied. The class one method and the detailed mass estimate have provided more realistic estimates of
104288 kg. This value may be considered as more accurate as more data has been available at this stage on
which to base the calculations. This value, while higher than the original, does validate the margin of error
applied in the earlier stage and as such does not require a re-design of the wings or re-selection of the
engines based on the parametric sizing charts.

The centre of gravity calculations have been calculated and will provide the first data needed for the
design of the undercarriage and the stability analysis. The forward most location may require structural re-
enforcement of the nose gear and the aft most location may be too far to the rear to provide for minimum
static stability and control response. With some re-working it was possible to bring the values within
tolerance, and this is the data supplied as the final iteration. The relevance of this is discussed in the
undercarriage and stability sections of the report.

It is also necessary to comment on the feasibility of the design with respect to the data provided by the
analysis. The more accurate method of mass prediction has resulted in a higher mass estimate. This will
directly influence the operational capabilities of the aircraft and its marketability. This is alleviated by the
fact that the centre of gravity locations do not provide significant problems for the other aspects of the
design. Consequently it is possible to say that the design remains feasible, given the mission requirements.

108 Weight and Balance Chapter 9


Final Report of the International Design Synthesis Exercise 2005

10 Control and Stability

Stability is a very important property for an aircraft, especially for a transport aircraft, which handles a
large amount of mass and thus large moments around its axes. The aircraft should be inherently stable to
avoid stall condition. This physical condition can be modelled into a mathematical model that further
defines the geometry of the HERA, in particular the control surfaces of the tail section. That will be the
topic of the current chapter.

There are two types of stability: lateral and longitudinal stability. Longitudinal stability describes the
behaviour of the aircraft along the lateral (Y-) axis and is controlled by the elevator. The
upward/downward gusts, the wing and elevator forces are of major influence on longitudinal stability.
Lateral stability describes the behaviour of the aircraft along the X- and Z-axes and is controlled by the
rudder control surfaces. The side gusts and lateral forces have a major influence on the lateral stability.

To be able to calculate the stability of the tail surfaces, a first estimate must be produced of which the
control derivatives will be derived and the model refined or adjusted if necessary. This will be done in
section 10.1

10.1 Initial sizing of control surfaces

From the various configuration possibilities of a tail design the conventional ‘inverted T’ configuration is
chosen to form the tail of the HERA. This choice is based on a number of requirements and criteria
which play a role on the design of the whole airplane as well as the local parts: cost, weight, technical
performance, weight, reliability/robustness, feasibility and development risk, and maintainability cost.

The classical method of producing a first estimate of the tail control surfaces is starting off from a classical
surface volume coefficient and calculate the areas according to the following formulae:

cVT bW SW
SVT = for the vertical tail (10.1.1)
LVT

c HT C w SW
S HT = for the horizontal tail (10.1.2)
LHT

This calculation is obtained amongst others from Raymer [28]. The surface areas become then:
SVT= 39.9 m2 SHT= 61.6 m2

Chapter 10 Control and Stability 109


Final Report of the International Design Synthesis Exercise 2005

Once the area of the surfaces has been estimated the wingspan and other data can be determined by
geometry from Figure 10.1.1:

b = AS (10.1.3)

(a) (b)

Figure 10.1.1 From left to right: a) Geometry vertical tail plane, b) Geometry horizontal tail plane

The mean aerodynamic chord can be computed with the following formula from Raymer [28]:

2 1 + λ + λ2
c MAC = c root (10.1.4)
3 1+ λ

This results in a first estimate of the vertical and horizontal tailplanes. The dimensions can be found in
table (10.1.1).

Table 10.1.1 Tailplane dimensions

Vertical tail Horizontal tail


c root = 6.64 m c root = 7.04 m
c top = 3.32 m c top = 3.52 m
MAC = 5.16 m MAC = 5.48 m

Traditionally the horizontal tailplane has a sweep of about 5° due to aerodynamic considerations. The
HERA has adopted this angle. The vertical tailplane usually has an angle between 0° and 20° to prevent a
too low critical Mach number. The HERA has adopted a 20° angle.

110 Control and Stability Chapter 10


Final Report of the International Design Synthesis Exercise 2005

A comparison was made between several aircraft for airfoil types. Roskam [33] was used as a reference for
this. Classical airfoil types are used for both surfaces: the horizontal tailplane will use airfoil profile NACA
0010 and the vertical tailplane will use NACA 0015. Once those are selected, the aerodynamic properties
can be calculated making use of the program XFOIL, available as freeware from Mark Drela [59]. This
program calculates in 2D the properties of the airfoils, i.e. CL CD0 CM. This data must then be converted
to 3D for the whole tail section.

First the derivative ∂CLα must be obtained from the XFOIL data. This is the slope of the lift coefficients
versus the angles of attack. Then this slope can be used to calculate the 3D ∂CLα curve by using the
equation:

a0
∂C Lα =
a (10.1.5)
1+ 0
πAe

where the aspect ratios and Oswald factor are as determined in the tailplane spreadsheet, which can be
found in Appendix 7. Since both airfoils are symmetrical, the curves for both airfoils have a value of zero
for a zero angle of attack. The graphs can thus be determined by:

C L = ∂C Lα α (10.1.6)

These graphs can be found in tailplane spreadsheet on the CD in Appendix 7. They will be used to
calculate the longitudinal and lateral stability.

10.2 Longitudinal stability and control derivatives

Static stability of an aircraft can be readily defined as the tendency of that aircraft disturbed from steady
level flight, to return to its pre-disturbed state. A dynamically stable aircraft may be defined as one where
the energy contained within a disturbance is dampened. The analysis of the static and dynamic stability
characteristics of the HERA is the first step in the process of understanding how the aircraft will respond
when disturbed in steady level flight. It will also supply data to begin the future analysis of the control
force inputs and control column travel required to change the flight speed and state.
Longitudinal modes of motion include the linear motion about the aircraft x and z-axis and the angular
rotation about the y-axis consequently it is these modes that will form the central focus of the
investigation.

Chapter 10 Control and Stability 111


Final Report of the International Design Synthesis Exercise 2005

This report examines the longitudinal static and dynamic stability of the HERA. Using the link between
longitudinal stability and tail volume ratio, it sizes the horizontal stabiliser based on the trim requirements
in all configurations. This results in the determination of minimum static margin, horizontal tail area, tail
setting angle and minimum elevator deflection. This analysis is carried out using an Excel sheet developed
as part of the Aircraft Design 3 module at The Queen’s University of Belfast, School of Aeronautical
Engineering.

10.2.1 The Criteria for Longitudinal Static Stability


Before the analysis can begin, it is important to define the terms to be used and those by which the
properties of the aircraft will be measured. An aircraft is said to be trimmed in steady level flight if the
resultant forces about its centre of gravity are zero i.e. lift is equal to weight, thrust is equal to drag and all
axial moments are zero. For the aircraft to be stable it should display the following pitching moment
characteristics during wind tunnel testing and indeed during flight, see figure 10.2.1.

Figure 10.2.1 Pitching Moment Characteristic

For this configuration, an aircraft disturbed from condition A to B by an accidental control input or up
gust would experience a negative or “nose down” pitching moment, induced so as to provide a tendency
for the aircraft to return to its original state. Disturbance from A to D will produce a positive or “nose
down” restoring tendency.

The final terms, which need to be defined are the Neutral Point –N and the Static Margin –KN. The neutral
point is the aerodynamic centre of the aircraft including the tail plane and the static margin is defined as
the distance between the neutral point and the centre of gravity non-dimensionalised with respect to the
mean aerodynamic chord, see equation 10.2.1. The static margin is also the gradient of the pitching
moment characteristic graph, provided the sign of the gradient is reversed. From these the two criteria for
static stability in trimmed flight can be defined.

112 Control and Stability Chapter 10


Final Report of the International Design Synthesis Exercise 2005

dCMG V ' a1 ⎛ dε ⎞
KN = − = −(h − ho ) + ⎜1 − ⎟ (10.2.1)
dCL 1 + F a ⎝ dα ⎠
ST lT
V '= (10.2.2)
Sc
ST a1 ⎛ dε ⎞
F= ⎜1 − ⎟ (10.2.3)
S c ⎝ dα ⎠

The first is a positive static margin that is, the centre of gravity must lie forward of the neutral point. As
the static margin is the reverse sign of the CMG versus CL slope, the gradient of this curve must be negative
at the trim lift coefficient. The second is that the pitching moment coefficient at zero lift coefficient must
be positive. For the HERA, which is a conventional aircraft with a positively cambered wing, a tail plane
must be used to achieve a stable configuration and correct the inherent nose down pitching moment.
Equation 10.2.1 also demonstrates that if the centre of gravity is aft of the mean aerodynamic chord,
stability of the aircraft must be provided by the tail plane by increasing the area ST or by increasing the tail
plane setting arm, lT.

10.2.2 Tail Plane Sizing


The method out lined in section 10.1 gives an estimation of the horizontal tail based on empirical data
gathered from similarly configured aircraft. As a direct consequence, it does not take into consideration
stability or control needs. This section provides a description of the method used to estimate the surface
area, geometry and location required for a specified amount of longitudinal stability and the range of
elevator deflections required for rotation authority and trimmed flight. The method is based on that
provided by Denis Howe in “Aircraft Conceptual Design Synthesis” [1]. The method assumes a
symmetrical aerofoil for the horizontal stabiliser. The aerodynamic data used in this method is listed in
Appendix 13.

The Cruise Trim Requirement


The cruise configuration for the aircraft can be analysed using the total pitching moment coefficient about
the centre of gravity, stick fixed, equation 10.2.4.

zT V' ⎡ a1 ⎛ dε ⎞ ⎤
CMG = CMo + CL (h − ho ) + CT − ⎢CL a ⎜1 − dα ⎟ + a1ηT + a2η + a3 β ⎥ = 0 (10.2.4)
c 1+ F ⎣ ⎝ ⎠ ⎦

This equation can be rearranged to make ηT, the tail plane setting angle, the subject equation 10.2.5. By
assuming zero elevator and trim tab angle, ηT becomes the setting angle to trim the aircraft. As the area
ratio ST/S is contained in the above equation, the tail plane setting angle can be calculated for a range of

Chapter 10 Control and Stability 113


Final Report of the International Design Synthesis Exercise 2005

area ratios and centre of gravity locations. For the range of centre of gravity locations, a mean setting
angle can be determined. This estimate will result in the minimum elevator required to trim the aircraft in
any configuration.

zT
CMo + CL (h − ho ) + CT
ηT = c − CL ⎛⎜1 − dε ⎞⎟ = 0 (10.2.5)
a1
1
ST lT a ⎝ dα ⎠
Sc 1+ F

There is a considerable amount of aerodynamic data required to make use of these equations. Methods of
analysis based on empirical data can be found in the ESDU engineering database [57]. It is important to
note that for accuracy and to provide validation; several of the more important values were calculated
separately from the aerodynamics section. While this in some circumstances constituted a doubling up of
the work load, it does provide a means of co-validation. The values of most importance to accurately
estimate were CMo and ho. For these values wing, fuselage and nacelle effects had to be estimated and
these calculations can be found in the Excel sheet in the CD file for this chapter.

Elevator Angle Required for Landing


The elevators will be required to compensate for the increased nose down pitching moment caused by the
deployment of the flaps. For this configuration, a range of elevator angles is calculated using the same area
ratios and centre of gravity ranges as the cruise condition carpet plot, figure 10.2.2. Taking the average tail
plane setting angle, the re-arranged pitching moment coefficient equation becomes:

zT
CM O + C L (h − ho ) + CT
η= c − a1 ⎡ C L ⎛⎜1 − dε ⎞⎟ + η ⎤ = 0 (10.2.6)
a2 ⎢⎣ a ⎝ dα ⎠
T⎥
S l′ 1 ⎦
a2 T T
S c 1+ F

To solve this equation it is necessary to re-evaluate the aircrafts aerodynamic characteristics to account for
the effect of flap deployment. However there is significant difficulty in accurately predicting this data.
Abbott [1] states “….no adequate theory has been developed to predict the aerodynamic characteristics. Consequently the
information required... (for flap effects)…is obtained entirely by empirical methods” Applying the methods in ESDU
provided the increment in pitching moment coefficient, the rest of the data was provided by the
aerodynamics specialist group. Looking at figure 10.2.2 it can be seen that there is a limit set on the
elevator angle to trim of 270. This represents 67% of the available elevator travel for the C-130 and is set
so as to provide sufficient available authority to manoeuvre in the event of emergency. Given the lack of
this highly detailed data for the HERA, it is a reasonable assumption at this stage of the design process. It
can be seen that the elevator has sufficient authority for rotation in the most forward centre of gravity

114 Control and Stability Chapter 10


Final Report of the International Design Synthesis Exercise 2005

condition with out exceeding the 67% travel limit. This demonstrates there will be no necessity for an all
moving tail plane.

0
ηT = -6.65 degrees

0.15
-5
0.25

h
Elevator angle to trim (degrees)

-10

-15 0.40

0.35
0.30 Chosen
-20
0.25 S T /S=0.3
Limit on η S T /S
0.15
-25
Tak e-off
rotation

-30

-35

Figure 10.2.1 Landing Configuration Constraints Diagram

Longitudinal Static Stability in Cruise


Equation 10.2.1 provides a means of estimating the static margin for a range of tail plane area ratios and
centre of gravity locations. This is done by making ST/S the subject.

ST lT 1 a1 ⎛ dε ⎞
K N = −(h − ho ) + ⎜1 + ⎟ (10.2.7)
S c 1 + F a ⎝ dα ⎠
ST K N + (h − ho )
=
S lT′ 1 a1 ⎛ dε ⎞ (10.2.8)
⎜1 − ⎟
c 1 + F a ⎝ dα ⎠

To make use of this method, it is necessary to provide a value for minimum static margin. Typically a
value of 0.1 can be assumed. This is entered into the Excel sheet and a static stability line is presented on
the carpet plot. Typically a viable configuration will lie to the right of this line.

Chapter 10 Control and Stability 115


Final Report of the International Design Synthesis Exercise 2005

Rotation Authority in Forward Centre of Gravity Position


The ability of the aircraft to rotate during the take-off ground run is limited by the most forward centre of
gravity position due to its larger moment arm about the undercarriage. The minimum value of h was
determined to be 0.15MAC and this was entered into the excel spreadsheet along with the average tail
plane setting angle. The rotation condition can be modelled by using the equation for conservation of
angular momentum, equation 10.2.9. This used the main undercarriage as the fulcrum.

MkB2 &&
θ = CMO + CL (hG − hO ) −
Mg
(hG − h) − CT zG −VT ⎡⎢CL a1 ⎛⎜1− dε ⎞⎟ + a1ηT + a2η⎤⎥ (10.2.9)
ρS c qS c ⎣ a ⎝ dα ⎠ ⎦

In this equation, kb is the pitch radius of gyration for the aircraft and θ&& is the pitch acceleration. This
equation is re-arranged to make the area ratio the subject (equation 10.2.10) and a range of values plotted
in the cruise landing carpet plot.

2
Mg
CM O + CL (hG − hO ) − (hG − h ) − CT zG − MkB θ&&
ST qS c ρS c
= (10.2.10)
S ⎡ a1 ⎛ dε ⎞ ⎤ lT
⎢CL a ⎜1 − dα ⎟ + a1ηT + a2η ⎥
⎣ ⎝ ⎠ ⎦c

With this last data range, the carpet plot can be presented and discussed, figure 10.2.3

-3

-4 Chosen S T /S = 0.3
0.40

-5
0.2
Tailplane setting angle to trim (degrees)

h
0.15
-6

-7 0.05 Chosen ηT
Static
stability in 0.40
cruise 0.35
-8
Kn =0.1 S T /S
-9
0.20
Take-off

-10
rotation

0.2
-11

-12

Figure 10.2.2 Cruise Configuration Constraints Diagram

116 Control and Stability Chapter 10


Final Report of the International Design Synthesis Exercise 2005

From this constraints diagram it is possible to get a value for the minimum area ratio which satisfies the
requirements for static stability and elevator authority. The value of ST/S=0.3 has been selected as it lies to
the right of both the take-off rotation constraint and the static stability constraint. Based on a wing area of
214.1m2 this gives a horizontal tail area of 65m2 It does appear that this area ratio does not provide
sufficient static stability at an aft centre of gravity location of 0.4MAC, however from chapter 9, it has
been determined that the most aft centre of gravity location in any feasible flight configuration is 0.3MAC.

10.2.3 Test for Longitudinal Static Stability


Based on the criteria outlined in the introductory section, it is now possible to evaluate the stick fixed
static stability of the HERA.

Pitching Moment Characteristic

0.2

0.15
Pitching Moment Coefficient

0.1

0.05

0
0 0.2 0.4 0.6 0.8 1 1.2
-0.05

-0.1
Lift Coe fficient

Maximum Payload Kn=0.261

Figure 10.2.3 Pitching Moment Characteristic Graph for the HERA in Steady Level Flight

The static margin in cruise at maximum payload was calculated to be KN=0.261 satisfying the first
requirement for static stability. At the most aft feasible centre of gravity position in flight, the static margin
was calculated to be KN=0.116 demonstrating that all feasible centre of gravity locations, the centre of
gravity lies ahead of the Neutral Point.

The second requirement for static stability can be seen in figure 10.2.3. For flight at maximum payload
CMo is positive at CL=0. Also a trimmed condition exists for a value of CL=0.65. This is the optimum CL
specified by the aerodynamics specialist group and would imply that the tail provided by the Howe
analysis is the optimum at this stage of the design.

Chapter 10 Control and Stability 117


Final Report of the International Design Synthesis Exercise 2005

The analysis, working under the assumption of a symmetrical aerofoil has provided a tail plane setting
angle of -6.650 however, typical values for cargo aircraft are between 00 and -20. This can be achieved for
the HERA by selecting an aerofoil with the necessary aerodynamic properties when used within this range
of setting angles. One such aerofoil is the NACA 23012. If it is inverted and set at an angle of -1.750 it
provides the required negative lift to trim the aircraft at zero elevator angle. While this aerofoil was
selected from many others due to its favourable aerodynamic characteristics, it was later found that this is
the aerofoil in use on the C-130 Hercules.

An attempt was made to verify this assumption by calculating the lift coefficient of the tail plane during
trimmed steady level flight. The analysis in the “Calculated Data” page of the excel sheet for this chapter
under predicts the tail plane lift coefficient required trimmed flight. It gives a value of CLT=-0.252. Hand
calculation to validate this value has shown that a value of CLT=-0.289. The reason for this has been
determined to be that the excel calculation neglects the effects due to the thrust provided by the engines.
Based on this analysis, the following horizontal tail design was determined. A CAD schematic is provided
in Appendix 14.

Table 10.2.1 Horizontal Tail Design


Horizontal Tail Area - ST 65 m2
Root Chord - cr 5.25 m
Tip Chord - ct 2.62 m
Span - bT 16.5 m
Tail Plane Setting arm - lT 16.06 m
Aspect Ratio - AT 4.2 -
Leading Edge Sweep – ЛT 12 Degrees
Setting Angle – ηT -1.75 Degrees
Maximum Required Deflection – ± 25 Degrees
ηMAX
Static Margin in Cruise - KN-CRUISE 0.261 -
Tail Plane Lift Coefficient - CLT -0.289 -
NACA Aerofoil 23012 -

10.2.4 Longitudinal Dynamic Stability


An analysis of an aircrafts flight dynamics is an analysis of the motion during disturbance from a trimmed
condition. This analysis determines the longitudinal aerodynamic stability derivatives and examines the
behaviour of the short period oscillation and the phugoid modes of motion. To aid in the process, a
MATLAB program constructed as a design tool and learning aid has been used. This tool has been

118 Control and Stability Chapter 10


Final Report of the International Design Synthesis Exercise 2005

validated through several years use in the Aircraft Design 3 and Aircraft Dynamics 3 [39] module at The
Queen’s University of Belfast, School of Aeronautical Engineering.

Longitudinal Aerodynamic Stability Derivatives

Table 10.2.2 Aero-Normalised Stability Derivatives


Aero-Normalized Stability Derivatives
XU -0.13 ZU -1.3 MU 0
XW 0.44 ZW -7.24 MW -1.94
Xq 0 Zq -3.08 Mq -10.4
X_W 0 Z_W 0 M_W -4.12

Short Period Oscillation Approximation from Equations of Motion


The short period motion of an aircraft is dependant on the static margin. Typically, a larger static margin
causes the mode to become oscillatory. This analysis will examine the characteristic equation of the mode
and give the time to half amplitude, the period and the damping ratio.
Assuming forward speed does not change during the motion of the aircraft, the longitudinal equations of
motion simplify to:

⎡⎛ ZW& ⎞ ⎤ ⎛ Zq ⎞
⎢⎜⎜1 − ⎟⎟λ − ZW ⎥ − ⎜⎜1 + ⎟λ
⎣⎝ µ1 ⎠ ⎦ ⎝ µ1 ⎟⎠
=0 (10.2.11)
⎛ M W& ⎞ ⎛ iy Mq ⎞
− ⎜⎜ λ + M W ⎟⎟ ⎜⎜ λ − ⎟λ
⎝ µ1 ⎠ ⎝ µ1 µ1 ⎟⎠

⎛ Mq MW& ⎞ ⎛ ZW M q µ1MW ⎞
λ2 − λ⎜⎜ ZW + + ⎟+⎜ − ⎟=0 (10.2.12)
⎝ iy i y ⎟⎠ ⎜⎝ i y iy ⎟

A'= 1 (10.2.13)

⎛ M q MW& ⎞
B' = −⎜ ZW + + ⎟ (10.2.14)
⎜ i i ⎟
⎝ y y ⎠
⎛ ZW M q µ1 M W ⎞
C'= ⎜ − ⎟ (10.2.15)
⎜ i iy ⎟
⎝ y ⎠

Chapter 10 Control and Stability 119


Final Report of the International Design Synthesis Exercise 2005

From this the following can be calculated:

Natural Period –

A' (10.2.16)
Tn = 2πσ
C'

Time to half amplitude –


A'
t1 / 2 = −(2σ ln 2) (10.2.17)
B'

Damping Ratio –

⎛ ln 2 ⎞⎛ Tn ⎞
ζ =⎜ ⎟⎜⎜ ⎟⎟ (10.2.18)
⎝ 2π ⎠⎝ t1/ 2 ⎠

Period –
Tn
T= (10.2.19)
1−ζ 2

Phugoid Approximation from Equations of Motion


The phugoid mode can be described as a lightly damped oscillatory motion where kinetic and potential
energy is exchanged. This takes the form of true airspeed and altitude. Assuming there is no significant
change in pitching moment about the centre of gravity the longitudinal equations of motion can be
simplified to:

⎛ Xq ⎞
(λ − X U ) − ⎜⎜ λ − C L ⎟⎟ (10.2.20)
⎝ µ1 ⎠ =0
− ZU −λ
λ + λ 3C D + 2C L2 = 0
2 (10.2.21)

λ = −C D ± C D2 − 2C L2 (10.2.22)

The Damping Ratio –


CD
ζ = (10.2.23)
2C L

120 Control and Stability Chapter 10


Final Report of the International Design Synthesis Exercise 2005

Period of Oscillation –

T= (10.2.24)
2C L

The MATLAB program provided the following results, table 10.2.3

Table 10.2.3 Short Period Oscillation and Phugoid Time Signatures


Short Period Oscillation Phugoid
T (s) 5.66 T(s) 80
t1/2 1.48 t1/2 161
ζ 0.389 ζ 0.042

The coefficients of the characteristic equations are as follows.

Table 10.2.4 Short Period Oscillation and Phugoid Characteristic Coefficients


A’=1 B’=10.6 C’=187 D’=26.9 E’=151.5

From the results listed in table 10.2.3 the aircraft displays some problems with its dynamic behaviour. The
short period oscillation damping ratio is of insufficient magnitude to suggest proper damping of the
mode. Acceptable values lie above 0.5. The dampening of the SPO is virtually independent of the static
margin, so while the aircraft is statically stable in all flight modes it does not necessarily mean it has the
appropriate handling qualities expected by a pilot. However, the SPO damping is provided by the tail
plane, which has been designed with this in mind. Therefore, while the tail may not provide all of the
necessary damping, it will be well with in the abilities or a relatively simple avionics system to make up the
difference.

The tail plane does however play virtually no role in the dampening of the phugoid; consequently the
damping ratio is an order of magnitude smaller than that for the SPO. This is because the dampening
provided by the tail plane is proportional to the rate of change of pitching moment. It can be assumed
that in phugoid, pitching moment is more or less constant through out the oscillation. The period of the
phugoid, of about 80 seconds means that the pilot will feel obliged to get involved in correcting the
aircraft back to steady level flight. Typically this results in the pilot adding to the energy of the system,
worsening the situation. This will require the avionics system to continually dampen the phugoid by means
of a feed back loop, improving the handling qualities of the aircraft and removing the problem.
The next step in the design process will be to estimate the control forces required to make the aircraft
change its flight condition. This will require more detailed information about the avionics system and the
acceptable control force limits.

Chapter 10 Control and Stability 121


Final Report of the International Design Synthesis Exercise 2005

10.3 Lateral stability and control derivatives

This section deals with the lateral stability and control of the HERA. The main objectives are to estimate
the following:
• the size, shape and location of the fin necessary to provide a specific degree of directional stability;
• the size of the rudder and the range of deflection necessary to control the aircraft throughout its
speed envelope;
• the wing dihedral angle;
• the size and location of the ailerons.

As for the longitudinal stability and control, the ESDU website was used to obtain certain data needed to
complete this section and a Matlab program was used as guide through the ESDU process for estimating
the aerodynamic stability derivatives. The data found using this website and the Matlab program was used
along with data already calculated from other sections such as wing area, wing span etc. to find the values
outlined above.

The following aerodynamic stability derivatives were required to complete this section:
• Lv = Rolling moment due to sideslip
• Nv = Yawing moment due to sideslip
• Lp = Rolling moment due to roll rate.

These derivatives have been determined using ESDU website along with a Matlab program. Lv was found
by estimating the contributions from the wing planform [11], wing dihedral [17], body [11], wing position
[9], nacelles [9], fin [14] and flaps [12]. The values were estimated using the equations and graphs provided
by each document and the resulting values were then added together to give the value for Lv. Below is an
example of how one of the contributions to the final value of Lv was calculated.

(Lv)w = wing planform contribution

A = 11, Λ1/4 = 0°, λ = 0.36, Λ1/2 = -2.75°, CL = 0.65, M = 0.55

1 ⎡ (1 − λ ) ⎤
× = 0.043 (10.3.1)
A ⎢⎣ (1 − λ ) ⎥⎦

Using the calculated value above and the values of sweep angle at ¼ chord and ½ chord along with a
provided graph the following value was found:

122 Control and Stability Chapter 10


Final Report of the International Design Synthesis Exercise 2005

⎡ (Lv )w ⎤
⎢ ⎥ = 0.05 (10.3.2)
⎣ CL ⎦

(Lv) )w = 0.05 × CL = 0.0325 (10.3.3)

M cos Λ1 / 2 = 0.55 × cos(− 2.75) = 0.55 (10.3.4)

Using another graph the following value was found: Km = 1.14 Using these values the final contribution
to Lv by the wing planform was found to be:

(Lv )w = 1.14 × 0.0325 = 0.03705 (10.3.5)

The other contributions to Lv were found using similar methods and the values calculated for each are
shown below:

Table 10.3.1 Lv = Rolling Moment due to Sideslip

Lv = Total of Contributions
(Lv)w = 0.03705
(Lv)d = 0
(Lv)B = -0.00246
(Lv)h = -0.0249
(Lv)ninner = -0.00929
(Lv)nouter = -0.00646
(Lv)F = -0.0414
(Lv)f = 0.0299

These values were added together to give a total value for Lv = -0.0176

Nv was found by estimating the wing and body contribution [10], nacelle contribution [10], fin
contribution [14] and flaps contribution [13] to the total value. These values were then added together to
give the final value for Nv = 0.1593.

Lp was found by estimating the contributions of the wing [18], fin [15], dihedral [16] and tailplane [18] [15]
and adding the resulting values together. The final value was found to be Lp = -0.2797.

Chapter 10 Control and Stability 123


Final Report of the International Design Synthesis Exercise 2005

These values were then used a number of times to determine sizes and positions of the fin, rudder and
ailerons. This will be the topic of the next section.

The spreadsheet used for the determination of the lateral control derivatives can be found on the CD in
Appendix 7.

10.4 Initial sizing of the fin and rudder

10.4.1 Heading after Engine Failure


This section shows that our aircraft will still be able to maintain a straight heading with no more than 5°
of bank angle. Ne is a non-dimensional aerodynamic coefficient expressing the yawing moment generated
by the out of balance thrust from the engines and it can be calculated using the following equation:

TyE
Ne = (10.4.1)
qSb

Where T is the thrust of one engine (T = 14174.44N), q is the dynamic pressure (q = 7294.74Pa), S is the
wing area (S = 214.1m2), b is the wing span (b = 48.5m) and yE is the lateral distance from the engine to
the centre of gravity (yE = 12.25m). The results is: Ne = 0.00229

10.4.2 Rudder angle to trim after engine failure


This section deals with obtaining a minimum value for the rudder derivative. After an engine fails the
speed of the aircraft is reduced to 1.2Vs. The rudder angle required to maintain the heading at this speed
must not exceed 0.75times the maximum rudder deflection. The maximum rudder deflection was assumed
to be 20° or 0.35 rads. A yawing moment was generated by the engine and this was countered by an equal
and opposite yawing moment from the fin. The yawing moment is Nζζ where Nζ is a non-dimensional
aerodynamic derivative. Therefore the yaw trim condition can be written as:

Nζ ζ = Ne (10.4.2)
Ne
ζ = (10.4.3)

But the rudder deflection cannot exceed 0.75 times the maximum deflection so:

1.33 Ne
Nζ ≥ ≥ 0.0087 (10.4.4)
ζ max

124 Control and Stability Chapter 10


Final Report of the International Design Synthesis Exercise 2005

10.4.3 Inability of Rudder to Cause Fin Stall


This section deals with obtaining a maximum value for the rudder derivative. Full rudder deflection must
never lead to a dynamic sideslip angle that would cause fin stall. The directional static stability of the
aircraft is expressed by the non-dimensional aerodynamic derivative Nv. If the sideslip angle is β then:

Nζ = − Nvβ (10.4.5)

Dynamic fin stall may be delayed to about β = 0.5 radians so it may be expressed as:

0.5 Nv
Nζ ≤ ≤ 0.2277 (10.4.6)
ζ max

10.4.4 Heading Hold in a Crosswind


This sections deals with the fact that it must be possible to hold a steady sideslip angle, and to change
heading, against a crosswind. To effect a heading change of 0.2 rads against a crosswind requires that:

0.2 Nv (10.4.7)
Nζ ≥ ≥ 0.091
ζ max

10.4.5 Change of Heading Against a Failed Engine


It may be necessary to hold a steady sideslip angle when landing in a crosswind with a failed engine. If the
maximum sideslip angle is 0.262 rads then:

Nζ ≥
(Ne + 0.262 Nv ) ≥ 0.1258 (10.4.8)
ζ max

10.4.6 Determination of Design Values


This section deals with finding the area of the fin and chord sizes of the fin and rudder which satisfy all of
the above inequalities. Using the following formula:

⎡⎛ cf ⎞0.47 ⎤
Nζ = F ⎢⎜ ⎟ + 0.08⎥VF (10.4.9)
⎣⎢⎝ c ⎠ ⎦⎥

where (cf/c) is the mean ratio of the rudder chord to the fin chord and VF is the fin volume ratio which is
(Sflf/Sb). The factor F is 2.3 for a body mounted tail. Nζ was calculated for a range of values of (cf/c) and

Chapter 10 Control and Stability 125


Final Report of the International Design Synthesis Exercise 2005

(Sf/S). The range of values of (cf/c) was from 0.2-0.5 and for (Sf/S) was 0.1-0.3. The results are shown in
table 10.4.1 below.

Table 10.4.1 Results of Nζ for a Range of Values of Sf/S and cf/c


Sf/S
0.1 0.15 0.2 0.25 0.3
0.2 0.04291 0.06436 0.08581 0.10727 0.12872
cf/c 0.3 0.05060 0.07590 0.10120 0.12650 0.15181
0.4 0.05702 0.08553 0.11405 0.14256 0.17107
0.5 0.06264 0.09396 0.12528 0.15659 0.18791

From the table above the lowest value of (Sf/S) must be chosen which satisfies all of the following:
|Nζ| ≤ 0.2276
|Nζ| ≥ 0.0087
|Nζ| ≥ 0.091
|Nζ| ≥ 0.125
The lowest value of (Sf/S) which satisfies all these requirements is 0.2 at (cf/c) = 0.5. For S = 214.1m2 this

gives S F = 42.82m 2 .

The fin geometry can now be determined using:

⎧ ⎡ (TR × crF ) ⎤ ⎫
S F = bF × ⎨crF + ⎢ ⎥⎬ (10.4.10)
⎩ ⎣ 2 ⎦⎭

For TR = 0.5 and bF = 6.4 m this gives:

⎛ cr + 0.5crF ⎞
42.82 = 6.4 × ⎜ F ⎟ (10.4.11)
⎝ 2 ⎠

This gives crF = 5.35m and crT = crF × TR = 2.67 m .

The mean aerodynamic chord of the fin is given by:

MAC =
2
crF × ⎢
( )
⎡ 1 + TR + TR 2 ⎤
(10.4.12)

3 ⎣ (1 + TR ) ⎦

126 Control and Stability Chapter 10


Final Report of the International Design Synthesis Exercise 2005

The results of the initial sizing of the fin and the rudder are summed up in Table 10.4.2 below.

Table 10.4.2 Results of the fin and rudder sizing


Parameter Symbol Value
Fin surface SF 42.82 m2
Fin chord at root c F root 5.35 m

Fin chord at tip c F tip 2.67 m

Fin mean aerodynamic chord MAC Rudder 2.08 m


Fin span bF 6.4 m
Rudder surface S Rudder 19.98 m2
Rudder relative chord crudder / c fin 0.5

10.4.7 Dihedral Angle


As the HERA is a high wing aircraft there is no need for wing dihedral for ground clearance of the
engines. Dihedral can also be employed to ensure the stability of an aircraft in different dynamic modes.
However the HERA has no dihedral, even though it is unstable in the slow spiral and Dutch roll modes,
as this means it is a simpler design and easier to manufacture etc. The fact that this will leave the aircraft
dynamically unstable is discussed in more detail in section 10.4.11.

10.4.8 Aileron Sizing


It is assumed that the aileron span is limited by the trailing edge flap requirements. The span of the
ailerons is therefore 7.35 m. Hence only the MAC of the aileron has to be calculated and it is found from
the ratio of aileron chord to wing chord.

10.4.9 Crosswind Case


Aileron authority is expressed by the non-dimensional derivative Lζ. In a sideslip of magnitude β, the
lateral stability generates a rolling moment Lvβ, and this must be balanced by a rolling moment generated
by the ailerons if the aircraft is to remain wings-level. If the maximum sideslip angle is 0.2rads and the
aileron angle used is to be no more than 50% of the maximum available which is 0.28rads then:

0.4 Lv
Lv ≥ ≥ −0.0251 (10.4.13)
ζ max

This sets the lower limit on the aileron authority and the aileron chord.

Chapter 10 Control and Stability 127


Final Report of the International Design Synthesis Exercise 2005

10.4.10 Roll performance


This section deals with the aircrafts ability to roll from 30° left to 30° right in under 7 seconds in normal
conditions or in under 11 seconds with only one engine operating, but only 50% of the maximum aileron
deflection is allowed in this case. The aileron authority may then be calculated from:

bΦLp
Lξ ξ max ≥
⎡ ⎤ (10.4.14)
V ⎢t1 +
K1
e (
1 − K1t1
−1 ⎥ )
⎣ ⎦

Where t1 is the time taken for the manoeuvre, V is the TAS, Φ is the change in bank angle and K1 is a
constant given by:

1
ρSV 2 Lp (10.4.15)
K1 = − 2
Ix

Where Ix is the rolling moment of inertia of the aircraft. It was estimated from the mass distribution and
geometry of the aircraft in section 9.4. Using those values the rolling moment of inertia of each major
section of the aircraft was estimated and they were added together to give the overall rolling moment of
inertia of the aircraft. See section 9.4 for a more in depth description.

The total rolling moment of inertia of the whole aircraft was found to be 8757538 kgm. When this value
was inserted into the formula for K1 along with ρ = 0.5252 kg/m3 and V = 166.67m/s then K1 was found
to be 0.807. This value was then inserted into equation (10.4.14), yielding Lξ ξ max ≥ −0.01746 and thus

Lξ ≥ −0.06238 .

This value was then used, as the maximum value of Lζ is chosen as the basis for the aileron sizing, in the
following formula to obtain the ratio (cf/c)A:

η ⎡⎛ cF ⎞ 0.16 ⎤
0.47

Lξ = − a ⎢⎜ ⎟ + ⎥ (Fξi − Fξo ) (10.4.16)


2 ⎣⎢⎝ c ⎠ A ⎦⎥

Here, a is the wing lift curve slope, η is the location of the aileron mid-point non-dimensionalised with
respect to the wing semi-span, A is the wing aspect ratio, and Fζi and Fζo are inner and outer correction
factors for the spanwise load distribution on the aileron. They are given by:

128 Control and Stability Chapter 10


Final Report of the International Design Synthesis Exercise 2005

⎛ ⎡ 0.8 ⎤ ⎞
Fξ (η ) = ⎜⎜1 − ⎢ ⎥ ⎟⎟(1 − η ) − 1.34 × 10 (12 + A tan Λ1 / 2 − 8λ ) (cos[π (η − 0.5)]) (10.4.17)
1.25 −4 2.5

⎝ ⎣ βm A⎦ ⎠

Fζinner was found to be 0.2832 while Fζouter was found to be 0.0538. These values were used along with that
for Lζ in the formula:

η ⎡⎛ cF ⎞ 0.16 ⎤
0.47

Lξ = − a ⎢⎜ ⎟ + ⎥ (Fξi − Fξo ) (10.4.18)


2 ⎣⎢⎝ c ⎠ A ⎦⎥

From this it was found that c F / c = 0.00672 . As the MAC of the wing is 4.737m then the MAC of the

aileron is, cF = 0.03186 m. It was found that the aileron began at 16.5 m along the span of the wing and
ended at 23.85 m along the span, therefore it is 7.35m long, see Figure 4.1.1 for the wing planform of the
HERA. Therefore the area can easily be calculated as: 1.28 × 1.74 = 0.23415m 2

This area is obviously extremely small and only indicates the minimum aileron area to control the aircraft.
For that reason the area of the aileron on the HERA was chosen to be 5.39m2 as this would mean the
HERA has very good aileron control and could perform turns easily. This value was chosen using some
basic assumptions about the general areas of ailerons on aircraft. As the MAC is 4.737m then it was
assumed that the chord of the wing at the aileron was approximately 3m and that the chord of the aileron
was 25% of this, so it was 0.75m. This was used along with the span of the aileron to find the area of
5.39m2.

10.4.11 Dynamic Stability


The aerodynamic stability derivatives were also used in a Matlab program along with a number of
dimensions, operating conditions, aerodynamic parameters, inertias and longitudinal derivatives to find the
characteristics of the dynamic stability of the HERA. The program produced results for the roll
subsidence, slow spiral and dutch roll for the aircraft and included values such as real and imaginary roots,
time to half amplitude and the period. It can be seen from the results obtained that the HERA is
dynamically unstable in both the slow spiral and dutch roll modes. The slow spiral can be stabilised with
the introduction of dihedral to the wing, however this would further destabilise the dutch roll which is a
far more important mode. Normally to stabilise the dutch roll mode the area of the fin would be
increased and this is possible to do for the HERA. At present the area of the fin is 42.82m2 and this could
be increased by 5%-20% and the results iterated to find the most suitable area for the fin of the aircraft,
which will have stability in dutch roll. However the dutch roll is not very unstable and the HERA employs
fly-by-wire control systems. Hence, it is sufficient to leave the area of the fin at its current value as the

Chapter 10 Control and Stability 129


Final Report of the International Design Synthesis Exercise 2005

instability of dutch roll and the slow spiral can be corrected easily with the aid of computer control. Table
10.4.3 lists the results for the lateral stability modes.

Table 10.4.3 Results for Dynamic Modes


Roll Subsidence Slow Spiral Dutch Roll
Real root -11 Real root 0.33566 Real root 0
Imaginary root NA Imaginary root NA Imaginary root 31.9875
T(s) NA T(s) NA T(s) 3.1199
t1/2(s) 0.9931 t1/2(s) -32 t1/2(s) -65

10.5 Main results of the lateral stability analysis

The following is a summary of the main results from sections 10.4 and 10.5.

Table 10.5.1 Summary of the main results

Parameter Symbol Value Unit


Rolling moment due to sideslip Lv -0.017608662
Yawing moment due to sideslip Nv 0.159379722
Rolling moment due to roll rate Lp -0.279757771
Area of Fin Sfin 42.82 m2
Area of Rudder Srudder 19.98 m2
Area of Aileron Saileron 5.39 m2
MAC of Fin MACfin 4.163 m
MAC of Rudder MACrudder 2.08 m
MAC of aileron MACaileron 0.75 m
Dihedral Γ 0 degrees

130 Control and Stability Chapter 10


Final Report of the International Design Synthesis Exercise 2005

11 Manufacturing, Assembly and Certification

One of the requirements is that the aircraft must be ready for service by 2010. In order to be able to reach
this milestone, potential problems should be identified as early as possible in the project life cycle. It is
therefore very important to pay continuous attention to manufacturing, assembly and certification issues.
This will be the topic of the present chapter. Section 11.1 will treat the manufacturing processes used to
produce the parts. The joining techniques will be discussed in section 11.2. Subsequently, section 11.3 will
discuss the assembly method. Finally, section 11.4 will elaborate on the approach to the verification and
certification of the HERA design.

11.1 Manufacturing

For an overview of the different parts of the HERA, an exploded view has been made, which can be
found in Appendix 17. The manufacturing techniques that will most probably be used to produce some of
these parts will be discussed below.

The parts of the fuselage that have to transfer very high loads are made in a casting process. Parts like
these include door substructures, wing ribs and spars, wing-fuselage connection, hinges, brackets, cabin
floor-fuselage connection, ramp beams, window frames, turbine blades, etc. The best casting process that
can be used to produce these parts is sand casting, because of its low equipment, mould, labour and raw
material cost and its series size.

Some components that can be cast successfully may also be manufactured economically by other
methods, such as forging. A typical characteristic of forging is that it produces components with locally
varying grain structures, this adding good strength and toughness to forged parts. Typical forged products
in HERA are heavily loaded parts such as the undercarriage, the fuselage frames and engine mountings.

The stringers used to stiffen the fuselage skin, cabin floor, and wing panels are produced using an
extrusion process. The stringers could also be made using a bending process. However, for this type of
process, the length of the stiffener would be limited to the size of the press, while the length of the
stringer made using extrusion could be virtually infinite. Another process that can be used to produce
stringers is roll forming, but due to the high cost of this process, extrusion is the better option for the
HERA.

Chapter 11 Manufacturing, Assembly and Certification 131


Final Report of the International Design Synthesis Exercise 2005

For the parts of the skin with large curvature, such as the skin panels used on the fuselage and the wing,
roll bending is used. Roll bending of sheet can also be used for the manufacture of other single-curved
parts with large curvatures used on the airplane.
Stretching or stretch forming is a process for the manufacture of large and slightly double curved shells.
These shells are e.g. present in the front and aft sections of the aircraft fuselage, in particular in the
cockpit area and the transition area close to the tail section. The tail section itself and the section between
the cockpit and tail are single curved skin sheets made by roll bending. Also wing leading edges are
produced using this process.

Rubber forming is the most important process in the aircraft industry for the forming of metal parts:
rubber forming makes about 50% of all sheet metal parts. The main reason for this high volume of parts
is the low production costs which are related to the tooling costs. Rubber forming is primarily a process
for the fabrication of flanged parts, like small frames and stiffeners, or parts with certain details.

11.2 Joining techniques

When connecting plates like skin panels to each other, or connecting stiffeners to plates, adhesive bonding
can be used. Adhesive bonding has some characteristics that distinguish it from riveted joints. An adhesive
bond has a long lifetime because there are no holes in the material; stress concentrations are absent,
making them very insensitive to metal fatigue. The life span of a bonded joint can be ten times as much as
of a riveted joint. The joint can be stronger than the surrounding material. Bonded joints are smooth and
airtight (fuel tight). There compulsory for the production of sandwich structures. Unfortunately, adhesives
usually require high temperatures and pressures using large and expensive bond presses or making
autoclaves necessary and bonded joints are not suitable for disassembly, difficult to repair and can only be
loaded in shear. However, bonding is less labour intensive than riveting, and can be corrosion resistant.
Adhesive bonded joints with fewer fasteners benefit aircraft by improving aerodynamic and signature
performance, eliminating fuel leak paths and simplifying manufacturing assembly, all resulting in a more
robust and affordable aircraft. Because of this, adhesive bonding should be applied as much as possible.

For parts that have a higher change of being damaged or parts that need regular inspection and repair,
riveted joints should be used. Riveted joints are very reliable due to their widespread and well-documented
use. Apart from this they are easy to inspect, simple to repair and can often be applied using simple tools.
On the downside of riveting are: fatigue- and corrosion-sensitive, uneven surface, poor efficiency (joint is
weak in comparison to surrounding material), high labour costs and the fact that they can only be loaded
in shear.

132 Manufacturing, Assembly and Certification Chapter 11


Final Report of the International Design Synthesis Exercise 2005

For joints that must be able to be unfastened, bolts and nuts should be used. Also in situations where
there should be pretension in the joint for an increased fatigue life or in situations where thick plates must
be connected and adhesive bonding is not possible due to sheets that are too thick. The amount of control
over the joint is larger with bolts and nuts than for rivets, but they are also more expensive. Bolts and nuts
are very strong and can be loaded both in shear and normal stresses. They are used in places where large
concentrated forces have to be carried, like the complicated fittings used to connect wings and fuselage.
Bolted connections are usually heavy and are subject to high tolerances, so they must be fitted with care.

11.3 Aircraft assembly

The production of the HERA will be done in serial, which means that the HERA will be put together in
the so-called assembly line. This line consists of a number of stations, each with its own crew working on
every aircraft the same kind of work. The product goes from station to station and remains in each station
an amount of time equal to the delivery interval. But before the assembly can be done the parts have to be
produced first. This means that the HERA will be built up from different sections. The synthesis of the
these sections has been displayed in an exploded view of the HERA, which can be found in Appendix 17

With such a structural breakdown, work can be divided among workstations, making sure the time needed
corresponds to the delivery period. Waiting periods of parts are minimised in this way. The breakdown of
the HERA consists of cockpit section, wings, front- and rear fuselage, centre fuselage and empennage. To
assemble these sections jigs and fixtures are needed. The task of the jigs is to support and position those
sections before joining them together. This way, time for delivery and cost requirement can be met.

11.4 Verification/certification approach

As with any project a certain amount of verification is necessary during and at the end of the development
phase. It must be verified that the product being developed complies with the requirements set by both
the customer and the law. In the aeronautical world this testing is more extensive than in other industries
and can cost a substantial amount of money if care is not taken to choose the appropriate verification and
certification methods.

The first subject to be verified is whether the customer’s wishes are satisfied. This must be done during
the whole process, especially at the milestones. The later a deviation is found in the design, the more
costly a rectification becomes.

Next it must be verified that the product itself satisfied the technical requirements set up by both the
customer and the authorities. To be able to fly over a country, the aircraft must not satisfy the national

Chapter 11 Manufacturing, Assembly and Certification 133


Final Report of the International Design Synthesis Exercise 2005

laws, but also the international regulations and agreements signed by that country. In the United States
and many other countries these regulations are the FAR (Federal Aviation Regulations). In Europe, the
JAA are applicable at present. However these regulations are about to change since EASA has been
established through Regulation of the European Parliament establishing a European Aviation Safety
Agency. By the time the HERA has to be certified the CS will have replaced the JAA. Also agreements of
the ICAO are applicable in some countries. It can be concluded that it is best to certify the HERA
according to the FAR and CS, keeping in mind the ICAO agreements.

To certify the HERA two verification methods are possible: analysis (FEM, thermal, electrical, RAMS)
and testing. Testing is more expensive and can only take place once the model is more or less completed.
Also, faults found during test will be expensive to correct. However some parts, especially crucial
mechanical parts will need (destructive) testing to be certified. This applies mostly to the structural part of
the aircraft, especially the tail section. Analysis of the design with software is a lot cheaper than testing and
can be done at earlier stages of the design, resulting in relatively less expensive corrections. Making
different types of analyses and corroborating them with some tests will reduce the testing needed, and
thus the cost of the certification phase.

The testing will be integrated in phase 4 of the design (see the design and development logic in section
2.6). It will start after the technical documentation is completed around. The prototype models will be
used for the physical tests.

134 Manufacturing, Assembly and Certification Chapter 11


Final Report of the International Design Synthesis Exercise 2005

12 Cost Analysis

This chapter gives a cost analysis for the HERA and will consist of four main sections; an estimation of
the price per aircraft, ways of reducing the estimated price per aircraft, a discussion on return on
investment and an estimation of the operating costs of the HERA.

12.1 Aircraft Estimated Price

This section deals with estimating the overall price of the aircraft that the customer will have to pay. It is
based on Roskam’s method and has two main stages. The first deals with estimating the cost of research,
development, testing and evaluation of the HERA and the second finds the manufacturing and acquisition
cost. These two stages are then combined to give the overall estimated price per aircraft.

12.1.1 Estimation of Research, Development, Test and Evaluation Cost (RDTE)


The estimation of the research, development, testing and evaluation cost is broken down into sections.
The resulting costs of each of these sections are then added together to give the overall RDTE cost. Each
of these sections is discussed below.

12.1.1.1 Airframe Engineering and Design Cost (Caedr)


The airframe engineering and design cost associated with the conceptual, preliminary and detailed design
phases can be computed from:

Caed r = (MHRaed r ) × (Re r ) (12.1.1)

Where MHRaedr is the total engineering man-hours required to complete the conceptual, preliminary and
detailed design and is calculated from:

MHRaed r = 0.0396(Wampr ) (V max )1.526 (Nrdte )0.183 (Fdiff )(FCAD )


0.791
(12.1.2)

Where:

Wampr = inv log[0.1936 + 0.8645(WTO )] (12.1.3)

Chapter 12 Cost Analysis 135


Final Report of the International Design Synthesis Exercise 2005

Vmax is the maximum speed of the HERA and is equal to 350.96 knots. Nrdte is the number of aircrafts
produced for the RDTE phase. There are two aircraft for flight tests and one aircraft for airframe static
tests. Fdiff is a factor, which accounts for the difficulty of the design of the HERA. As the HERA is a
relatively simple design incorporating proven technology a small factor was used. FCAD is a judgement
factor, which takes into account the experience the manufacturers have with CAD; as the more experience
they have the more cost efficient they are when using it. Most manufacturers are very familiar with CAD
nowadays so a small factor was chosen. Rer is the engineering dollar rate per hour for the airframe
engineering. It includes costs such as direct engineering labour, overheads, administrative costs etc.

12.1.1.2 Development Support and Testing Cost (Cdstr)


This section deals with the costs incurred from wind tunnel, structural, systems, propulsion testing etc.
The total cost of these activities is calculated from:

Cdst r = 0.008325(Wampr ) (V max )1.89 (Nrdte )0.346 (CEF )(Fdiff )


0.873
(12.1.4)

Where CEF is a cost escalation factor, which takes into account the fact that Roskam’s book uses data
from 1989 and updates it to values applicable to today’s industry.

12.1.1.3 Flight Test Airplanes Cost (Cftar)


This section deals with the cost of producing the flight test airplanes using the following equation:

Cfta r = C (e + a )r + C manr + C matr + C toolr + C qcr (12.1.5)

Where C(e+a)r is the cost of the engines and avionics for the RDTE phase and is found from:

( )
C (e + a )r = C er N e + C pr N pr + C avionicsr ( Nrdte − N st ) (12.1.6)

Where Cer is the cost per engine, Ne is the number of engines per aircraft, Cpr is the cost per propeller, Np
is the number of propellers per aircraft, Cavionicsr is the cost of avionics equipment per aircraft and Nst is
the number of aircraft used for static tests.

Cmanr is the manufacturing cost of the flight test airplanes and is calculated from:

(
C manr = MHRmanr Rmr )( ) (12.1.7)

136 Cost Analysis Chapter 12


Final Report of the International Design Synthesis Exercise 2005

Where Rmr is the manufacturing labour rate in dollars per man-hour and MHRmanr is the number of
manufacturing man-hours required in the conceptual, preliminary and detailed design phases. It can be
calculated from:

= 28 . 984 (Wampr )0.74 (V max ) (Nrdte )0.524 (Fdiff )


0 . 543
MHRman r
(12.1.8)

Cmatr is the cost of materials used in the manufacture of the flight test aircraft and is calculated from:

C matr = 37.632(Fmat )(Wampr ) (V max )0.624 (Nrdte )0.792 (CEF )


0.689
(12.1.9)

Where Fmat is a correction factor, which takes into account the type of materials used in the aircraft. As the
HERA is made primarily of proven aluminium alloys, a low factor was chosen.

Ctoolr is the tooling cost associated with the manufacturing of the flight test aircraft and it is calculated
from:

(
C toolr = MHRtoolr Rtr )( ) (12.1.10)

Where MHRtoolr is the tooling man-hours required in the conceptual, preliminary and detailed design
phases. It is calculated from:

MHR tool r = 4 . 0127 (Wampr )0.764 (V max )


0 . 899
(Nrdte )0..178 (N r r
) (Fdiff )
0 . 066 (12.1.11)

Where Nrr is the RDTE production rate in aircraft per month. Rtr is the tooling labour rate per man-hour
in dollars. Cqcr is the cost associated with the quality control of the manufacturing of the flight test aircraft
and it is found using:

(
C qcr = 0.13 C manr ) (12.1.12)

Where Cmanr is found from equation 12.1.7.

12.1.1.4 Flight Test Operations Cost (Cftor)


This section deals with the costs incurred during flight and simulation tests and is found from:

C fto r = 0 . 001244 (Wampr )1.16 (V max ) (Nrdte − N st ) (CEF )(Fdiff )(Fobs )


1 . 371 1 . 281
(12.1.13)

Chapter 12 Cost Analysis 137


Final Report of the International Design Synthesis Exercise 2005

Where Fobs is a factor that depends on the importance of having good stealth qualities. Obviously this
does not apply to the HERA so a value of 1 is used.

12.1.1.5 RDTE Profit (Cpror)


This section deals with the profit made from RDTE activities. This is included as a cost in the RDTE
phase. It is found using:

( )
C pror = F pror (C RDTE ) (12.1.14)

Where Fpror is the desired percentage profit of the total RDTE cost. It was taken to be 0.1. CRDTE is found
from equation 12.1.16.

12.1.1.6 Cost to Finance the RDTE Phases (Cfinr)


This section deals with the cost to finance the RDTE phase. If money is borrowed to finance the RDTE
phase then interest must be paid on the loan and this is where this cost comes from. Even if internal
funds were used to finance the RDTE phase then money would be lost due to those funds not earning
interest and this would be the cost. Due to the lack of better information the following formula was used
to find this cost:

( )
C finr = F finr (C RDTE ) (12.1.15)

Where Ffinr is taken to be 0.1 and CRDTE is found from equation 12.1.16. When all of these costs were
found it was possible to calculate the total cost of the RDTE phases for the HERA using the following
formula:

C RDTE = C aed r + C dstr + C ftar + C ftor + C tsf r + C pror + C finr (12.1.16)

The total cost obtained using this formula is for the entire RDTE phase and this is a fixed cost that will
not change throughout the manufacturing process. Therefore a part of this cost must be applied to each
aircraft produced i.e. the cost must be divided equally over the number of aircraft being produced.
Therefore the more aircraft that are produced, the less percentage of the total RDTE cost applied to each
one. Hence, if more aircrafts are produced, the overall price of the aircraft will decrease. It is clear from
Figure 12.1.1 that it would only make financial sense if at least 300 aircraft are sold, otherwise the RDTE
cost is too high per aircraft. Figure 12.1.1 illustrates this.

138 Cost Analysis Chapter 12


Final Report of the International Design Synthesis Exercise 2005

RDTE COst Per Aircraft versus Number of Aircraft Produced

25000000

20000000
RDTE Cost Per Aircraft

15000000

10000000

5000000

0
0 100 200 300 400 500 600 700 800 900 1000 1100
Number of Aircraft Produced

Figure 12.1.1 Reduction in RDTE Cost per Aircraft due to Increase in Number of Aircraft Produced

12.1.2 Method for Estimating Manufacturing and Acquisition Cost


This section deals with estimating the manufacturing cost and acquisition costs during the manufacturing
phase. The difference between these costs is the profit made by the manufacturer:

C acq = C man + C pro (12.1.17)

An estimation of the unit price per aircraft can be obtained from:

AEP = {(C man + C pro + C RDTE ) / N m } (12.1.18)

Where Nm is the number of aircraft produced (this does not include the aircraft produced as test aircraft).
Having stated this formula the rest of section 12.2.2 deals with estimating the manufacturing cost Cman and
manufacturer’s profit Cpro.

12.1.2.1 Airframe Engineering and Design Cost (Caedm)


Airframe engineering and design consists of a number of activities, such as design studies and integration
studies to meet with individual customer requirements, analysis of reliability, maintainability and
accessibility etc. The total cost of these activities is found from:

( )( )
C aed m = MHRaed program Rer − C aed r (12.1.19)

Chapter 12 Cost Analysis 139


Final Report of the International Design Synthesis Exercise 2005

Where MHRaedprogram is the total amount of engineering man-hours required for the entire aircraft program
and it can be found using:

MHR aed program


= 0 . 0396 (Wampr )0.791 (V max )
1 . 526
(N program )
0 .. 183
(Fdiff )(FCAD ) (12.1.20)

Caedr is found from equation 12.1.16 and Rem is the engineering man-hour rate in dollars for the entire
aircraft program. It is assumed to be equal to Rer. Nprogram is the total number of aircraft produced
(including test aircraft).

12.1.2.2 Airplane Program Production Cost (Capcm)


The cost of the program production consists of a number of components and can be found using:

C apc m = C (e + a )m + C int + C man m + C mat m + C tool m + C qc m (12.1.21)

Where C(e+a)m is the cost of the engines and avionics equipment, Cint is the cost of the airplane interior,
Cmanm is the cost of labour during manufacturing, Cmatm is the cost of the materials used to manufacture
the aircraft, Ctoolm is the cost of tooling during manufacturing of the aircraft and Cqcm is the cost of quality
control associated with producing the aircraft.

12.1.2.3 Cost to Finance the Manufacturing Phase (Cfinm)


This section is basically the same as section 12.2.6 and follows the same rules except that the cost found is
for the manufacturing phase. It is found using the formula:

C fin m ( )
= F pro m (C MAN ) (12.1.22)

where Ffinm is taken to be 0.1 and CMAN is found from equation 12.1.7.

12.1.2.4 Manufacturing Cost (CMAN)


The total manufacturing cost is the total of the costs outlined in the sections above and is found using:

C MAN = C aed m + C apcm + C ftom + C finm (12.1.23)

140 Cost Analysis Chapter 12


Final Report of the International Design Synthesis Exercise 2005

12.1.2.5 Profit (CPRO)


This section deals with the profit made on manufacturing activities and is taken to be 10% of the total
cost of manufacturing:

( )
C PRO m = F pro m (C MAN ) (12.1.24)

Where Fprom is taken to be 0.1 and CMAN can be found from equation 12.1.23.

The acquisition cost can now be calculated using equation 12.1.17. Using the total manufacturing cost, the
profit desired and the RDTE cost calculated in equation 12.1.16 the estimated price per aircraft can be
found using equation 12.1.18. The price will depend on the number of aircrafts produced and, as stated in
section 12.1.1, it does not make financial sense to produce less than 300 aircraft. The AEP if 150 aircraft
are produced was found to be approximately $49.08 million. Appendix 16 shows the detailed results to
estimate the price per aircraft. The price obtained was checked for realism by comparison to a value
obtained based solely on the take off weight of the aircraft. This value was found from:

AMP = (CEF )[inv log{3.3191 + 0.8043(log WTO )}] (12.1.25)

The value obtained for the HERA using this formula was approximately $45.3 million, which means the
estimated cost of the HERA is in fair agreement with this [29].

12.2 Reductions in Cost

The AEP found in section 12.1 does not meet the customer requirement of $30 million so some ways of
reducing the cost needed to be found. As the method used was based on the take off weight of the aircraft
any weight savings, such as reduced landing gear weight due to the fact it is semi-retractable, are included
in the figure for take-off weight that the cost estimation is based on. So there were reductions in cost due
to lighter landing gear, fly-by-wire control systems and the fact that the aircraft was not pressurised among
others. These cost savings are included in the reduced take off weight of the aircraft. The effect of take off
weight on aircraft price is important when the number of aircrafts produced is low, because of the
‘learning curve’ effect i.e. as more experience is gained in manufacturing the aircraft it becomes cheaper to
do so. Therefore the more aircrafts produced, the cheaper the aircraft becomes. The aircraft also be
comes cheaper as more are produced as the fixed RDTE cost is then spread over a greater number of
aircraft. Reducing the aircraft production rate will also reduce the overall cost of the aircraft. Cheaper
manufacturing and tooling labour rates could also be used to reduce the overall cost.

Chapter 12 Cost Analysis 141


Final Report of the International Design Synthesis Exercise 2005

The main way in which the cost of the HERA can be reduced is by producing more aircrafts. However
customers have to be found for these aircraft. Nowadays there are a huge number of NGO’s e.g. Red
Cross, Oxfam, AirServ etc. Each of the NGOs is involved in a huge number of humanitarian aid projects
throughout the world that require air transport. The market is definitely there for an aircraft like the
HERA, as the main competitors it would face, are African and Eastern European. These operators do not
perform to the highest safety standards so that operating costs can be kept to minimum. The HERA will
satisfy the highest safety standards while keeping operating costs to minimum. Reports on humanitarian
missions involving aircrafts, show that NGO’s would be prepared to pay slightly higher operating costs if
the aircraft is of a higher standard, as this will decrease costs incurred for spare parts, maintenance,
accidents etc. Competitor’s fleets of aircraft are also aging and the HERA would be a very attractive
option for the NGO’s. It is also larger than many of the aircraft used now, with the exception of perhaps
the Antonovs, and would be more cost efficient than flying more than one smaller aircraft to the same
place [4][47]. There is also the possibility of entering the cargo transport aircraft industry, as many
privately owned companies would be interested in a cost effective aircraft, which is capable of carrying a
payload of 35000 kg. Therefore it is reasonable to suggest that on top of the 150 aircraft ordered by the
customer, at least a further 150 aircraft could be ordered by a combination of NGO’s and privately owned
cargo transport companies.

The original requirement was that the first aircraft be delivered by 2010 but as seen in section 8.10 the
landing gear will not be certified until 2012. Therefore a reduction in production rate can be employed
since the first aircraft will be delivered in 2012.

If the aircraft are manufactured in Russia there will be huge savings in labour costs due to the cost of
living and economy there. There will be a 60% percent saving in manufacturing labour rates and a 65%
saving in tooling labour rates that will considerably reduce the cost of each aircraft.

The original estimated price per aircraft using the original parameters, labour rates, number of aircraft to
be produced and delivery date was $49,080,907 and this has been reduced to $29,657,761 using the
reductions described above. This meets the customer requirement of under $30 million and the AEP
could be further reduced with orders from more customers. The detailed results for the reduced estimated
cost of the HERA can be seen in Appendix 16.

12.3 Return on Interest and Operational Cost

The main financial criterion on which the aircraft design should be judged is the return on investment
(ROI) to the company. The variability in the manufacture and operation, make the use of this parameter
difficult. During the first years of manufacture there is little cash inflow prior to the sales of the aircrafts.

142 Cost Analysis Chapter 12


Final Report of the International Design Synthesis Exercise 2005

There is only a small gain due to small payments of customers. Figure 12.3.1 is a good example of cash
flow during the design and launch of a new aircraft: There are heavy losses due to design and
manufacturing and cash flow picks up when the product is commercialised. So, it can be said that the
‘return’ of the ROI will depend on the terms and conditions of the investment loans and not solely on the
aircraft’s technical and operational performance. [38]

Figure 12.3.1 General Project Cash Flow

Therefore, it is hard to say what the actual ROI will be for the HERA, as this project only specifies for a
specific task required for one “imaginary” client with a possible order of 150 aircraft. As the figures earlier
presented, more investors will be needed to break even.

12.4 Operational Cost

The operational cost, as will be explained in this section, is a very complex quantity to calculate. There are
many factors influencing cost that are often outside the control of designers and predictions which means
that estimations and methods employed to estimate future costs, can only by partially realistic in their
results. Additionally, operational cost is divided into two costs; the DOC and IOC. The following section
will provide a reference for potential customers and estimates what the operational cost would be in the
case of an assumed frequent flight, which will be most probable to be frequently done. Assumptions:
• It will be assumed that the great majority of flights undertaken by the HERA will be at a MTOW of
101 tonnes, a range of 4000 km and a cruise speed of 324 knts (167 m/s). This is due to the fact
that the aircraft is designed for this specific mission and will probably fly the same route on
repeated occasions.

Chapter 12 Cost Analysis 143


Final Report of the International Design Synthesis Exercise 2005

• A block speed of 273 knts, a block annual utilisation of 1300 hours and were estimated from the
available performance data.

The method here presented is adopted from Roskam [34]. Although ref [34] calculated DOC and IOC for
a commercial airliner, the method and costs can be adapted to actual cargo aircraft. All formulas, graphs
and trends can be found in [34] or on the CD in Appendix 7. Additionally, at this stage of design it is
difficult to decide and predict accurately some costs.

The results obtained are intended to give a guideline and a magnitude of cost to possible customers of
what the cost per nautical mile (nm) is for flying and maintaining the HERA.

12.4.1 Direct Operating Costs (DOC).


DOC is usually expressed in US Dollars per nautical mile. This type of cost usually deals with the costs
produced by,
(1) DOC of Flying, which deals with costs related to cost of crew and fuel and oil used.
(2) DOC of Maintenance, which deals with: maintenance labour cost for engines, airframe and
systems, and maintenance material costs.
(3) DOC of Depreciation, which deals with depreciation costs of the aircraft, engine, propellers,
avionics and spare parts.
(4) DOC of Landing Fees. Landing fees are a matter of uncertainty. As this is aircraft designed to
transport refugees and humanitarian aid, in some cases, landing fees will not be applicable. Hence,
conservative values were used for the analysis.
(5) DOC of financing. According to [34], the cost of finance is around 7% of the total DOC.

The following table shows a summary and a break down structure of the DOC. The CD in Appendix 7
contains all the details of each DOC cost.

144 Cost Analysis Chapter 12


Final Report of the International Design Synthesis Exercise 2005

Table 12.4.1 DOC break down


1. DOC of flying
Crew cost 0,78 $/nm
Fuel and oil cost; C_pol 3,38282736 $/nm
2. DOC of maintenance 0,47545245 $/nm
Maintenance labour cost for airframe and systems 0,15300542 $/nm
Maintenance labour cost for engines 0,677016 $/nm
Maintenance materials cost for airplane and systems 2,59253686 $/nm
Maintenance materials cost for engines 1,69833356 $/nm
3. DOC of depreciation
Aircraft Depreciation Cost 2,16757158 $/nm
Engine depreciation Cost 3,12249291 $/nm
Airplane+Engines Spare parts Depreciation Cost 2,82138555 $/nm
4. DOC of landing fees, navigation fees and registry taxes
Cost of Landing Fees 0,61927913 $/nm
Cost of Navigation Fees 0,00462963 $/nm
5. DOC of financing 1,75 $/nm

The total DOC for the HERA was estimated to be $25 per nautical mile.

12.4.2 Indirect Operating Costs (IOC)


IOC methods deal with those costs not directly attributable to a particular aircraft type or the flying costs
of a particular operation. The expenses are usually classified under the following headings
1. Aircraft and traffic servicing
2. General and administrative overhead
3. Ground property and equipment maintenance and depreciation expenses
Management and operational aspects are predominant factors in the indirect costs and these are outside
the control of the aircraft designer. Additionally, they are very difficult to predict because of their
uncertainty factor, which is really dependent of the usage of the aircraft and the missions it is assigned.
Indirect operating costs are not insignificant; they can account for between 25 – 60% of the total
operating expenses. For the purpose of the HERA, IOC has been estimated to a value of 50% of the
DOC costs, therefore the IOC is $12.5/nm

Chapter 12 Cost Analysis 145


Final Report of the International Design Synthesis Exercise 2005

12.4.3 Total Operating Cost


Their result from the addition of DOC and IOC is the following table. As no other data is available for
comparison it is hard to say if these results are acceptable for a humanitarian cargo aircraft.

Table 12.4.2 Total operational cost of aircraft


Each Nautical Mile costs (DOC+IOC) 37,5 USD/nm
Seat mile 0,28846154 USD
Total stage cost 81000 USD

146 Cost Analysis Chapter 12


Final Report of the International Design Synthesis Exercise 2005

13 Sustainable Development Strategy

When designing a new transport aircraft, special consideration must be given to the sustainable
development. Bearing in mind the continuously growing (air) traffic density, world population and
pollution levels, a newly developed aircraft should be designed with the well being of future generations in
mind. Every system that is now built should bear this in mind. Together with the pressure of public
opinion and the threat of resource depletion, an aircraft must be designed limiting aerial pollution, noise
and the amount of waste material, not only when being in service but also when being produced and
when being disposed at the end of life.

Since the target of the HERA is to be in flight at 2010 and will be flying for approximately another 30 to
40 years, a long-term view considering present and future needs is required.

13.1 Production

Sustainable development concerns more than the aircraft being in use. To be sustainable, the production
techniques and facilities should to be looked at. It should be the aim to integrate environmental protection
into the production process itself, through measures designed to avoid the use of harmful substances and
prevent pollutant emissions being released in the first place, or at least reduce them to a minimum.

Environmental training activities for staff and for external partners along the life-cycle chain are also of
important influence on being as sustainable as possible when producing, using and disassembling the
HERA.

Things like the reduction of noise emissions from HERA production, the introduction of advanced
mechanical engineering techniques, such as dry machining and low-volume lubrication, the use of highly
developed paint techniques and reducing paint waste should be taken into account. At the same time the
handling of potentially water-polluting substances should be optimised. Closed-loop water cycles should
be implemented in the production sector to help conserve resources. Even things like eco-friendly
refrigerants in air conditioning systems and preventive measures to protect groundwater and soil should
be taken into account.

A simple production aspect for example is the adhesive-coated sand used in the manufacturing process.
The sand serves as a mould for aluminium-cast parts. Normally the sand used is sent to a cement kiln for
recycling. It is possible to recycle and re-use this in-house, thus preserving valuable resources.

Chapter 13 Sustainable Development Strategy 147


Final Report of the International Design Synthesis Exercise 2005

13.2 Operational use

For the HERA turboprops are used instead of turbofans. This will decrease the fuel consumption
drastically with about 15-20%. Since the AE1107c is an engine used on military aircraft, there is not much
information available about noise and emissions.

Also, the HERA does not have a pressurised cockpit and cabin. Compared to a pressurised, this will lower
the fuel consumption and extend the range. The non-pressurisation also means that less power is needed
from the engines for environmental control, which also lowers the fuel consumption.

13.3 Recycle plan

The main parts of the HERA, like fuselage, tail, wings, cockpit and ramp are manufactured using
aluminium, which is easy to recycle. Systems like fly-by-wire can be reused. Engines are disassembled and
potential elements are reused. Interior like seats, if still in a good condition, can be reused in similar
aircraft.

148 References
Final Report of the International Design Synthesis Exercise 2005

14 Conclusions and Recommendations

As closing chapter of this report, conclusions on the design of the HERA and its fulfilment of the
customer’s desires can be drawn, as well as a comment on the research and design yet to be done and
recommendations for a future team that will go on with the development of the HERA project. The
conclusions on the present design are discussed in the following section 14.1. A series of
recommendations for the future can be found in section 14.2.

14.1 Conclusions

The mission HERA was born from the necessity of the customer to acquire an aircraft that satisfies a
number of conditions that have lead to the mission need statement:

“To provide low-cost, flexible and reliable air transport of aid, evacuees and medical equipment under difficult
circumstances whilst remaining impartial.”

From this mission need statement the HERA project emerged. A mission analysis set up a series of
requirements that were subsequently validated by the customer and that ended up in a preliminary
concept. After the preliminary concept a further analysis of the systems was done.

This analysis has resulted in the present configuration. The HERA, almost 39 meters long and 49 meters
wide, is a high wing airplane with a high aspect ratio and large wing control surfaces. Its four six-bladed
Rolls Royce Ae1107c turbo propeller engines make it possible for the HERA to carry 35 tons of payload
in a cabin volume of 400 m3 to an objective over 5000 km away: up to 9 pallets, or up to 70 stretchers or
130 evacuees to bring help to necessitated zones or evacuate people from disaster areas. The cargo area is
equipped in such a way that the aircraft is independent from ground facilities in case these are lacking. The
landing gear has been designed so that the HERA is able to take off with its maximum take off weight and
land on unprepared airstrips under 1000 meters long.

This configuration satisfies six requirements set up by the customer by complying with them or even
surpassing them in the case of the range. However there is still one customer wish the HERA does not
comply with. The customer will have to wait an additional two years before seeing the HERA operative in
missions due to development and certification of certain systems. These two non-technical drawbacks
perhaps even underline the quality of the aircraft to be delivered, which will guarantee a safe and reliable
aircraft. A further analysis of the project by a future team should be able to reduce the development time
and the costs in order to satisfy the customer’s wishes. The requirements compliance is shown in the
requirements matrix in Table 14.1.1.

Chapter 14 Conclusions and Recommendations 149


Final Report of the International Design Synthesis Exercise 2005

Table 14.1.1 Compliance matrix.


Compliance Requirement HERA Specification Comments
3 Maximum payload of 35,000 kg 35,000 kg
3 Cabin volume 400 m3
Range with full payload of 4000
3 5490 km
km
Requirement was lowered to
650 km/hr after
3 Maximum speed of 700 km/hr 650 km/hr
requirements review with
the customer
Requirement was lowered to
STOL within 800 m from
3 950 m 950 m after requirements
unprepared runways
review with the customer
Acquisition cost no more than
3 $29.7 million
$30 million
Some items like the
undercarriage need
2 First flight in 2010 2012
extensive testing before
certification.

14.2 Recommendations

To aid future development of the HERA a few recommendations can be made. These are:

• The weight of the aircraft must be watched closely during further development. Efforts should be
made to maintain it or even reduce it in further structural analysis, by either looking at the
structure itself or the choice of materials. Should an increase in weight become inevitable, more
power should then be directed from the engines for the propulsion, cutting off from the systems
budget. This should be possible since the lack of pressurisation allows for a lower need for power
for the aircrafts systems.
• The undercarriage must still be further developed for a more detailed and realistic concept, whilst
trying to maintain the cost low.
• The HERA has an unstable behaviour in the Dutch roll condition, which is an undesired
property. Further attempts should be made by a stability team to modify the empennage surfaces
or surfaces’ configuration to obtain a more stable behaviour in that flight manoeuvre.
• It would also be very desirable to obtain support of Rolls Royce in the future for more detailed
design of the engines and the associated components.

If these recommendations are taken into account in future stages of the HERA design it should be
possible to achieve the goals set up by the HERA project within the limits imposed by the clients.

150 References
Final Report of the International Design Synthesis Exercise 2005

References

Literature

1 Abbott, I.H, Von Doenhoff, A. E., Theory of Wing Sections, 1st Editon, Great Britain: Dover
Publications, 1980, ISBN 0 48660 5868

2 AIAA, AIAA Aerospace Design Engineers Guide, 5th Edition, 2003, ISBN: 1-86058-424-1

3 Anderson, J.D, Fundamentals of Aerodynamics, McGraw Hill International Edition, 2001, ISBN 0-07-
118146-6

4 Coniglio, S., Military Technology-A Comparable Report of Military Transport Aircraft Programmes, 2003,
page 51-60.

5 Currey, N.S., Aircraft Landing Gear Design: Principles and Practices, first edition, Washington: American
Institute of Aeronautics and Astronautics, Inc., 1988, ISBN: 0930403-41-X

6 Del-Viso Hopkins, P.S., P.W. Fick, M.T. Gillen, R.D. Hiele, M.D.J. de Hoon, I. Huertas Garcia, S.
de Jong, F. Karatas, S. Kennedy, K. Matthews, K.M.L. Scott, D.J.J. Toal, DSE Project Plan,
International Design Synthesis Exercise 2005, Delft/Belfast, February 2005

7 Del-Viso Hopkins, P.S., P.W. Fick, M.T. Gillen, R.D. Hiele, M.D.J. de Hoon, I. Huertas Garcia, S.
de Jong, F. Karatas, S. Kennedy, K. Matthews, K.M.L. Scott, D.J.J. Toal, DSE Requirements Analysis
Report, International Design Synthesis Exercise 2005, Delft/Belfast, February 2005

8 Del-Viso Hopkins, P.S., P.W. Fick, M.T. Gillen, R.D. Hiele, M.D.J. de Hoon, I. Huertas Garcia, S.
de Jong, F. Karatas, S. Kennedy, K. Matthews, K.M.L. Scott, D.J.J. Toal, Mid Term Report –
Conceptual Design of a Humanitarian Relief and Emergency Aid Cargo Aircraft, International Design
Synthesis Exercise 2005, Delft/Belfast, March 2005

9 ESDU Engineering Data Sheet 73006, Effects of Isolated Body and Wing-Body Interference on Rolling
Moment due to Sideslip (with addendum A for nacelle effects

10 ESDU Engineering Data Sheet 79006, Wing-Body Yawing Moment and Sideforce Derivatives due to Sideslip
(with addendum A for nacelle effects)

11 ESDU Engineering Data Sheet 80033, Contribution of Wing Planform to Rolling Moment Derivative due to
Sideslip, at subsonic speeds

12 ESDU Engineering Data Sheet 80034, Effect of Trailing-edge Flaps on Rolling Moment Derivative due to
Sideslip

13 ESDU Engineering Data Sheet 81013, Effect of Trailing-edge Flaps on Sideforce and Yawing Moment
Derivatives due to Sideslip

References 151
Final Report of the International Design Synthesis Exercise 2005

14 ESDU Engineering Data Sheet 82010, Contribution of Fin to Sideforce, Yawing Moment and Rolling
Moment Derivatives due to Sideslip, in the Presence of Body, Wing and Tailplane

15 ESDU Engineering Data Sheet 83006, Contribution of Fin to Sideforce, Yawing Moment and Rolling
Moment Derivatives due to Rate of Roll in the Presence of Body, Wing and Tailplane

16 ESDU Engineering Data Sheet 85006, Contribution of Wing Dihedral to Sideforce, Yawing Moment and
Rolling Moment Derivatives due to Rate of Roll at Subsonic Speeds

17 ESDU Engineering Data Sheet A.06.01.03, Stability Derivative Contribution of Full-span Dihedral to
Rolling Moment due to Sideslip

18 ESDU Engineering Data Sheet A.06.06.01, Stability Derivative, Rolling Moment due to Rolling for Swept
and Tapered Wings

19 European Aviation Safety Agency (EASA), Certification Specifications for Large Airplanes – CS-25,
Brussels: EASA, 2003

20 Hammond, D., Noise Reduction Techniques for Special Operations C-130 Using Active Synchrophaser Control,
Air Force Research Laboratory, 1999

21 Houghton, E.L., P.W. Carpenter, Aerodynamics for Engineering Students, 5th Edition, Great Britain:
Butterwoth-Heinemann, 2003, ISBN:0-7506-5111-3

22 Howe, D., Aircraft Conceptual Design Synthesis, first edition, Burry St Edmunds: Professional
Engineering Publishing, 2000 ISBN: 1-86058-301-6

23 Howe, D., Aircraft Weight Prediction – Part 1, CofA Design Note DES126/1, Cranfield University,
UK, 1970

24 Jenkinson, L.R. Simpkin, P. Rhodes, D., Civil Jet Aircraft Design, 1st Edition, Great Britain: Antony
Rowe Ltd, 1999, ISBN: 0 340 74152 X

25 Megson, T.H.G., Aircraft Structurees for Engineering Students, 3rd Edition, Oxford: Butterworth-Heinemann,
2001, ISBN: 0-340-70588-4

26 Meriam, J.L., L.G. Kraige, Engineering Mechanics – Statics, 4th Edition, SI version, New York: John
Wiley & Sons, Inc., 1998, ISBN: 0-471-24164-4

27 Niu, M.C.Y., Airframe Structural Design-Practical design information and data on aircraft structures, 9th
printing, Hong Kong: Conmilit Press Ltd., 1988, ISBN: 962-7218-04-X

28 Raymer, D.P., Aircraft Design: A Conceptual Approach, 3rd Edition, Reston, Virginia: AIAA Education
Series, 1999, ISBN: 1-56347-281-0

29 Roskam, J., Airplane Design – Part I: Preliminary Sizing of Airplanes, 2nd Edition, Kansas:
DARCorporation, 1985, ISBN: 1-884885-42-X

152 References
Final Report of the International Design Synthesis Exercise 2005

30 Roskam, J., Airplane Design – Part II: Preliminary Configuration Design and Integration of the Propulsion
System, second edition, Kansas: DARcorporation, 1985, ISBN: 1-884885-43-8

31 Roskam, J., Airplane Design – Part III: Layout Design of Cockpit, Fuselage, Wing and Empennage: Cutaways
and inboard profiles, DARCorporation, 2002, ISBN: 1-884885-56-X

32 Roskam, J., Airplane Design – Part IV; Layout Design of Landing Gear and Systems, second edition,
Kansas: DARcorporation, 1985, ISBN: 1-884885-43-8

33 Roskam, J., Airplane Design – Part VII: Determination of stability, control and performance characteristics:
FAR and military requirements, 2nd Edition, Kansas: DARcorporation, 1988

34 Roskam, J., Airplane Design – Part VIII: Airplane Cost Estimation: Design, Development, Manufacturing and
Operating, Kansas, DARcorporation, 2002, ISBN 1-884885-55-1.

35 Ruijgrok, G.J.J., Elements of Airplane Performance, 2nd Edition, Delft: Delft University Press, 1996,
ISBN: 90-6275-608-5

36 Shevell, R.S., Fundamentals of Flight, 2nd Edition, Englewood Cliffs: Prentice-Hall, 1989, ISBN: 0-13-
339060-8

37 Torenbeek, E, Synthesis of Subsonic Airplane Design, Delft University Press, Martinus Nijhoff
Publishers, 1982, ISBN 90-247-2724-3

38 Wang, J., J.K. Watterson, Design 2: 205AER214 – Lecture notes, Belfast: The Queen’s University of
Belfast

39 Watterson, J.K. Aircraft Dynamics 3, 1st Edition, UK: The Queen’s University of Belfast, School of
Aeronautical Engineering. 2005

40 Watterson, J.K. Flight Mechanics 1&2, 1st Edition, UK: The Queen’s University of Belfast, School
of Aeronautical Engineering, 2002

References 153
Final Report of the International Design Synthesis Exercise 2005

Internet sources

41 Airbus Military, http://www.airbusmilitary.com

42 American Suzuki Motor Corporation, http://www.suzuki.com

43 An-70 Specifications, http://www.theaviationzone.com/factsheets/an70_specs.asp

44 Bombardier Q400, http://www.q400.com/q400/en/quiet.jsp

45 Federation of American Scientists – Rest of World Military Aircraft,


http://www.fas.org/man/dod-101/sys/ac/row/index.html

46 Federation of American Scientists – US Military Aircraft,


http://www.fas.org/man/dod-101/sys/ac/row/index.html

47 Humanitarian Missions which Require Air Transport


http://amos.qub.ac.uk/redirect/http%3a%2f%2fwww.magisterludi.com%2fsitomagister%2fpublic
_html%2faviation%2fpublications%2fpdf%2fhumanitarian_mission.pdf

48 ICAO Engine Emissions Database,


http://www.caa.co.uk/default.aspx?categoryid=702&pagetype=90

49 Land Rover, http://www.landrover.com

50 Medcoach, http://www.medcoach.com/products/clinic/index.html

51 Michelin Aircraft Tires, http://www.airmichelin.com

52 Rolls-Royce Defence Products,


http://www.rolls-royce.com/defence_aerospace/products/transport/ae1107c/default.jsp

53 Russian Engineering Salaries, www.us.design-reuse.com/news/news8350.html

54 Saab Aerospace, Saab 2000, http://www.saab.se/node5507.asp

55 Stanford University Aircraft Aerodynamics and Design Group, Department of Aeronautics and
Astronautics, http:/adg.stanford.edu/aa241/structures/weightstatements.html

56 The Aviation Zone, http://www.theaviationzone.com

57 The ESDU Engineering Data Base, http://www.athens.ac.uk

58 Toyota, http://www.toyota.com

59 XFOIL, http://raphael.mit.edu/xfoil

154 References
Final Report of the International Design Synthesis Exercise 2005

Appendix 1 Mission Requirements

Performance Maximum speed 700 km/h


Cruise speed 600 km/h
Takeoff distance 800 m (unprepared runways)
Range with full payload 4000 km (single trip)
Maximum payload 35000 kg
Cabin Minimum height 3.9 m
Minimum width 4m
Minimum volume 350 m3
Freight hold capacity 2 light vehicles side-by-side
9 pallets 88x108 ins 463L
9 96x125 ins 3610 types
Passengers Up to 130 (evacuees)
Loading Autonomous on- and offloading and cross loading capability, independent of
ground equipment
Landing capabilities Runway length 1000 m on hard field airstrip
Must be able to operate from runways of 800 on soft field airstrip at 85%
gross, i.e. ½ fuel but with evacuees (the airstrip would have no facilities)
Hospital capability The aircraft must be able to serve as a mobile operating theatre
Sustainability Provide a decommissioning plan for this aircraft for responsible disposal after
use
Cost $40 million (based on a production series of 150 aircraft)
Timeframe First flight should take place in 2010
Safety Evasive capabilities
Safety measures for transporting fuel

Appendix 1 Mission Requirements 155


HERA A400M An-70 C141B Starlifter C-130J Hercules C-160 Transall

Appendix 2
Appendix 2

Prime contractor - Airbus Military Antonov Lockheed Martin Lockheed Martin Transall
4 Allison AE1107c 4 TP400-D6 4 ZKMB Progress D-27 4 P&W TF33-P-7 4 Allison AE2100D3 2 Rolls-Royce Tyne
Power plant
turboprops turboprops propfans turbofan engines turboprops 22 turboprops
Power/thrust 4586 kW (6150 shp) 8200 kW (11000 shp) 10,300 kW (14,000 shp) 94 kN 3425 kW (4591 shp) 4552 kW (6100 shp)
Wingspan 48.5 m 42.4 m 44 m 48.8 m 40.4 m 40 m
Length 38.43 m 43.8 m 40.2 m 51.3 m 29.7 m 32.4 m
Height 11.92 m 14.6 m 16.1 m 11.9 m 11.6 m 11.7 m

Comparison of similar aircraft


Max. payload 35,000 kg 37,000 kg 47,000 kg 41,220 kg 18,956 kg 16,000 kg
Empty weight 42175 kg 70,000 kg 73,000 kg 67,185 kg 34,274 kg 28,758 kg
MTOW 101124 kg 130,000 kg 130,000 kg 155,585 kg 79,379 kg 49,100 kg
Cruise speed 600 km/hr Mach 0.68 – 0.72 750 km/hr 805 km/hr 630 km/hr 495 km/hr
Max. speed 650 km/hr 780 km/hr 800 km/hr 885 km/hr 648 km/hr 520 km/hr
Final Report of the International Design Synthesis Exercise 2005

Range 1) 5460 km 3400 km 1350 km 4750 km 5250 km 1800 km


Max. range 2) 12,000 km 6950 km 7400 km - 8300 km -
Service ceiling 7520 m 11,800 m (37,000 ft) 12,000 m (39,370 ft) 12,680 m (41,600 ft) 10,000 m (33,000 ft) 7924 m (26,000 ft)
Comparison of Similar Aircraft

Takeoff distance 950 m 940 m 900 m - 594 m 1100 m


Landing distance 950 m 625 m 900 m - 594 m 360 m
Cost per unit 3) $29.7 million ~ $80 million ~ $50 million ~ $48 million ~ $60 million -
1) at maximum payload
2) maximum range at reduced payload
3) 2002 US dollars

References used: [41], [43], [45], [46], [56]

157
Final Report of the International Design Synthesis Exercise 2005

Appendix 3 Functional Analysis

3.1 Functional flow diagram

Appendix 3 Functional Analysis 159


Final Report of the International Design Synthesis Exercise 2005

3.2 Functional Breakdown

160 Functional Analysis Appendix 3


Final Report of the International Design Synthesis Exercise 2005

3.3 Operational Flow Diagram

Appendix 3 Functional Analysis 161


Final Report of the International Design Synthesis Exercise 2005

Appendix 4 Project Gantt Chart

162 Project Gantt Chart Appendix 4


Final Report of the International Design Synthesis Exercise 2005

Appendix 4 Project Gantt Chart 163


Final Report of the International Design Synthesis Exercise 2005

Appendix 5 Three-View Drawing of the HERA

Will be printed separately!!!

Appendix 4 Project Gantt Chart 165


Final Report of the International Design Synthesis Exercise 2005

Appendix 6 Payload Configurations

6.1 Payload configuration top view

Appendix 6 Payload Configurations 167


Final Report of the International Design Synthesis Exercise 2005

6.2 Payload configuration front view

168 Payload Configurations Appendix 6


Final Report of the International Design Synthesis Exercise 2005

Appendix 7 CD-ROM

This appendix contains a CD-ROM with all the spreadsheets and programs that were used for the
analysis. It includes the following:

File name File type Description


Constraints Analysis Turboprop MS Excel spreadsheet This spreadsheet was constructed for the
constraints analysis that was discussed in
section 6.1.
Functional Flow Diagram Visio file and image file This is the functional flow diagram for a
better detailed view
N2 Visio file and image file This is the N2 chart for a better detailed
view
Airfoil analysis MS Excel spreadsheet Spreadsheet in which the 2D airfoil data
from XFOIL are analysed
Spanwise lift distribution MS Excel spreadsheet This spreadsheet calculates the basic and
additional lift distribution from which the
spanwise lift distribution can be obtained.
Using this sheet the CLMAX can be
obtained.
Analysis of high-lift devices MS Excel spreadsheet Contains data of the various flap types and
their effect
Wing Planform AutoCAD drawing Drawing showing the geometry of the
wing
Airfoil selection Folder Folder with all the data for the selection
and calculations on the wing’s airfoil.
Flap increments MS Excel spreadsheet Sheet describing the expected increments
from different flap types
Wing concept final MS Excel spreadsheet Sheet containing data of the wing
geometry and flap data
Wing MAC AutoCAD drawing Drawing of the wing locating the
aerodynamic centre of the wing and the
aircraft
Twist sheet MS Excel spreadsheet Sheet which calculates the lift coefficient
for the aircraft with various setting angles
and cruise angles
ICAO engine missions databank MS Excel spreadsheet Database of engine emissions for various
issue turbofan engines
Ae1107c Acrobat PDF file Engine specifications from Rolls Royce
T56 Acrobat PDF file Engine specifications from Rolls Royce
Ae2100 Acrobat PDF file Engine specifications from Rolls Royce
Tp400 Acrobat PDF file Engine specifications from Rolls Royce
C130 Noise reduction Acrobat PDF file This Acrobat file explains the efforts made
to reduce the noise emission of the
Hercules C-130
Solid nacelle Image file Picture of the Nacelle
Flight Performance MS Excel spreadsheet This spreadsheet contains the calculations
of the performance of the engines,
including propeller efficiencies, rates of
climb, power curves and flight envelope.

Appendix 7 CD-ROM 169


Final Report of the International Design Synthesis Exercise 2005

Payload Range MS Excel spreadsheet This spreadsheet was created to make the
payload-range diagram in section 6.6.
Tailplane Estimation MS Excel spreadsheet This spreadsheet has the data of the phase
2 estimation of the tailplane and the
aerodynamic characteristics of the airfoils
used.
H tail loading MS Excel spreadsheet Modification of Wing Loading for H tail
HT shear flow MS Excel spreadsheet Modification of Shear flow for H tail
H Tail solution MS Excel spreadsheet Modification of Wing Solution for H tail
V tail loading MS Excel spreadsheet Modification of Wing Loading for V tail
V tail solution MS Excel spreadsheet Modification of Wing Solution for V tail
VT shear flow MS Excel spreadsheet Modification of Shear flow for V tail
Ramp (bending and shear) MS Excel spreadsheet Calculations on shear and bending in I-
beams of ramp
Car specifications w.r.t. the ramp MS Excel spreadsheet Used as a basis for dimensioning the cargo
design box
Shear flow MS Excel spreadsheet This sheet was created to calculate shear
flows around an idealised torsion box for
the wing.
Case1_16Stringers Matlab file This is an example of the MATLAB files
used to calculate the shear flow caused by
bending shear stress.
Wing solution MS Excel spreadsheet This sheet was created to calculate spar
areas, stringer areas etc for various
spanwise wing positions
Fuselage shear and bending MS Excel spreadsheet This spreadsheet calculates the shear and
calculation bending of the fuselage.
Weight and balance analysis – MS Excel spreadsheet This sheet was created to provide an
final data aircraft mass breakdown and centre of
gravity
Tailplane estimation MS Excel spreadsheet This spreadsheet calculates the first tail
plane estimation as described in section
10.1, including the output from XFOIL
HERA stability analysis – final MS Excel spreadsheet This sheet was created to size the
data horizontal tail for required longitudinal
stability.
Lateral design MS Excel spreadsheet This sheet was created to calculate stability
derivatives and final fin and aileron
geometry
Detail cost estimation MS Excel spreadsheet This sheet was created to estimate the
price per aircraft without cost reduction
Final cost MS Excel spreadsheet This sheet was created to estimate hte
price per aircraft after cost reductions
Operational cost MS Excel spreadsheet This sheet was created to estimate the
operational cost of the aircraft.
XFOIL Folder with program Program to determine the 2D aerodynamic
files properties of airfoils
Propeller Design Program Matlab file Designs propeller for given conditions
Propeller Analysis Program Matlab file Analyses propeller design to produce
efficiencies for changes in pitch
Propeller Take-off Program Matlab file Program to calculate take-off thrusts
produced by propeller for different pitch
angles
Propeller Data Folder with various data Data on performance and geometric data
of each aerofoil section used

170 CD-ROM Appendix 7


Final Report of the International Design Synthesis Exercise 2005

Propeller Cad Folder with drawings Propeller models


Propeller Prices Acrobat PDF file Prices for propellers and related systems

Appendix 7 CD-ROM 171


Final Report of the International Design Synthesis Exercise 2005

Appendix 8 Blade Element Theory

Both the propeller design and the propeller analysis programs use the blade element theory. Essentially
this splits the entire length of a propeller blade up into a predetermined number of elements. Each
element is then considered separately as if it were in a simple 2 dimensional flow. The main difference
between the design and analysis programs is that the propeller design program defines the geometry of the
blade. This is achieved by considering each of the blade elements in turn. The relative velocity of the
airflow striking the element is determined and depends on the forward speed of the aircraft and the speed
of rotation of the element. The element is then initially defined as meeting this relative velocity at an angle
of 0˚. An iteration is necessary to find the relative velocity. The variables a and b are changed, which then
changes the angle that the relative velocity is to the plane of rotation ( φ ), as in equation 1.

V0 (1 + a )
tan φ = (1)
Ωr (1 − b )

With the value of φ calculated the thrust and torque on the blade element can be found, from which new
values for both a and b can be determined, which are then used to once again, find new values for a and b.
This cycle continues until an appropriate degree of accuracy is reached. The new values of a and b are
calculated using equations 2 and 3.

a 1
= .σ .t. cos ec 2φ (2)
1+ a 4
b 1
= .σ .q. cos ec 2φ (3)
1− b 2

The thrust and torque for each element is calculated from the lift and drag coefficient of that element,
which comes from the database and depends on the Mach number of the flow and the angle of attack.

Cl
t = Cl [cos φ − tan γ sin φ ] where tan γ = (4)
Cd
q = Cl [sin φ − tan γ cos φ ] (5)

dT dQ
= πrσVR .t and = πr 2σVR .q
2 2
(6)
dr dr

Appendix 8 Blade Element Theory 173


Final Report of the International Design Synthesis Exercise 2005

The total thrust and torque can be found by simply summing up the values for each element. With these
values, the power required to turn the blades and the overall efficiency of the propeller can be found. The
design program will alter the angle of attack of each element to gain the optimum efficiency. The airfoil
section will change along the length of the blade going from 20% thickness to 6%. Using thicker airfoil
sections at the blade root, increases the strength of the blade. The analysis program keeps all of the
geometry constant but uses the same techniques described to calculate the efficiencies at various speeds
and at various changes in the pitch of the propeller blade.

174 Blade Element Theory Appendix 8


Final Report of the International Design Synthesis Exercise 2005

Appendix 9 Rate of Climb Curves

Power at 15000m

14000

12000
Power (kW)

10000

8000

6000

4000
0 50 100 150 200
Velocity

Power Available Power Required

ROC at 15000 m

1.2

1
Rate of Climb (m/s)

0.8

0.6

0.4

0.2

0
0 50 100 150 200 250
Velocity (m/s)

Appendix 9 Rate of Climb Curves 175


Final Report of the International Design Synthesis Exercise 2005

Pr Vrs Velocity for Cruise

3 2
20000 y = 0.0021x - 1.0976x + 179.58x + 5821.2
18000
16000
14000
Power (kW)

12000
10000
8000
6000
4000
2000
0
50 100 150 200 250
Velocity (m/s)

Power Required V min power


Power Available Poly. (Power Available)

ROC at sea level

12

10

8
Rate of Climb (m/s)

0
0 50 100 150 200
-2

-4
Velocity (m/s)

176 Rate of Climb Curves Appendix 9


Final Report of the International Design Synthesis Exercise 2005

Appendix 10 Landing Gear Design

10.1 Fixed landing gear design

10.2 Semi retractable landing gear design

Appendix 10 Landing Gear Design 177


Final Report of the International Design Synthesis Exercise 2005

10.3 Shelter design

10.4 Connection strut-fuselage

178 Landing Gear Design Appendix 10


Final Report of the International Design Synthesis Exercise 2005

Appendix 11 Weight and Balance Calculation Table

Mass xicg xicgMi yicg yicgMi zicg zicgMi


i kg m kgm m kgm m kgm
1 Wing Mass Estimate 8146 19.2 156180 0 0 5 40729
2 Surface Control 1048 23.0 24172 0 0 5 5240
3 Tail Group Mass Estimate 2898 34.6 100239 0 0 8.5 24676
4 Body Mass Estimate 12303 19.2 236403 0 0 2.1 26205
5 Total Propulsion Mass 10422 16.6 172647 0 0 4.1 42730
6 Nose Gear 537 3.8 2063 0 0 0.5 268
7 Main Gear 3440 21.7 74510 0 0 0.5 1720
8 Total Fixed Equipment 4971 19.2 95524 0 0 2.1 10589
Empty Mass 353534 kg
⎛ 8 ⎞
⎜ ∑ M i xi ⎟
xcgM E = ⎝ i =1 ⎠
ME
Empty Mass CoG x 19.7 m y0m z 3.5 m
9 Crew 279 3.8 1072 0 0 2.13 594.3
10 Operational Items 1048 17.3 18131 0 0 2.13 2233.2
Operational Empty Mass 450923 kg
⎛ 10 ⎞
⎜ ∑ M i xi ⎟
⎝ i =1 ⎠
x cgM OE =
M OE
Op. Empty Mass CoG x 19.5 m y0m z 3.4 m
11 Fuel 24195 19.2 463887 0 0 5 120975
Un-laden Take-off Mass 69288 kg
⎛ 11 ⎞
⎜ ∑ M i xi ⎟
⎝ i =1 ⎠
x cgM U −TO =
M U _ TO
Un-laden T-O Mass CoG x 19.4 m y0m z 3.98 m
12 Payload 35000 17.3 605273 0 0 2.13 74550
Maximum Take-off Mass 104288 kg
⎛ 12 ⎞
⎜ ∑ M i xi ⎟
=⎝ ⎠
i =1
x cgM OE
M TO
Max. T-O Mass CoG x 18.7m y0m z 3.4 m

Appendix 11 Weight and Balance Calculation Table 179


Final Report of the International Design Synthesis Exercise 2005

Appendix 12 Centre of Gravity Excursion Diagram

Appendix 12 Centre of Gravity Excursion Diagram 181


Final Report of the International Design Synthesis Exercise 2005

Appendix 13 Aerodynamic Values For Horizontal Tail Sizing

Maximum Take Off Weight Maximum Landing Weight


Mass 104280 kg Mass 88200 kg
Wing 4811 Pa Wing Pa
Loading Loading 4041.298
S 214 m2 S 214.1 m2
L/D 17.98 - L/D 7 -
Thrust 57290 N Thrust 123606 N
Mach No. 0.55 - Mach No. 0.16 -
Static 35651 Pa Static Pa
Pressure Pressure 101000
CL 0.64 - CL 2.23 -
CT 0.035 - CT 0.32 -
a 7.2 rads-1 a 5.8 rads-1
a1 3 rads-1 a1 3 rads-1
a2 2.5 rads-1 a2 2.5 rads-1
dε 0.396 - dε -
dα dα 0.396
CMo -0.1798 - CMo -0.299 -
4.737 m m
c c 4.737
lT 16.06 m lT 16.06 m
zT -0.7 - zT -0.7 -
xUC 0.98 - xUC 0.98 -
h0 0.177 - h0 0.177716 -
h 0.15 - h -0.11606

Appendix 13 Aerodynamic Values for Horizontal Tail Sizing 183


Final Report of the International Design Synthesis Exercise 2005

Appendix 14 Horizontal Tail Geometry

Appendix 14 Horizontal Tail Geometry 185


Final Report of the International Design Synthesis Exercise 2005

Appendix 15 MATLAB Program Graphical User Interfaces

15.1 Graphical user interface P1

Appendix 15 Matlab Program Graphical User Interfaces 187


Final Report of the International Design Synthesis Exercise 2005

15.2 MATLAB Graphical User Interface P2

188 Matlab Program Graphical User Interfaces Appendix 15


Final Report of the International Design Synthesis Exercise 2005

Appendix 16 Cost Analysis

Appendix 16 Cost Analysis 189


Final Report of the International Design Synthesis Exercise 2005

190 Cost Analysis Appendix 16


Final Report of the International Design Synthesis Exercise 2005

Appendix 17 Exploded View of the HERA

Will be printed separately!!!

Appendix 11 Weight and Balance Calculation Table 191

You might also like