22.06.2015 Views

SAWE Report - Cal Poly San Luis Obispo

SAWE Report - Cal Poly San Luis Obispo

SAWE Report - Cal Poly San Luis Obispo

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>SAWE</strong> Paper No. 3232<br />

Category No. 10<br />

The Vendetta<br />

Preliminary Design <strong>Report</strong><br />

by<br />

<strong>Cal</strong> <strong>Poly</strong>, <strong>San</strong> <strong>Luis</strong> <strong>Obispo</strong> Design Team<br />

consisting of<br />

Kolby Keiser Chris Droney (Team Leader) Nathan Schnaible<br />

Chris Atkinson Christopher Maglio Dan Salluce<br />

For Presentation at the<br />

61 st Annual Conference<br />

of<br />

Society of Allied Weight Engineers, Inc.<br />

Virginia Beach, Virginia 20-22 May, 2002<br />

Society of Allied Weight Engineers<br />

Serving the Aerospace-Shipbuilding-Land Vehicles and Allied Industries<br />

Permission to publish this paper, in full or in part, with credit to the author and the Society may be<br />

obtained, by request, to:<br />

Society of Allied Weight Engineers, Inc.<br />

P.O. Box 60024, Terminal Annex<br />

Los Angeles, CA 90060<br />

The Society is not responsible for statements or opinions in papers or discussions at the meeting.


Abstract<br />

<strong>Cal</strong> <strong>Poly</strong> proudly presents the Vendetta, a supersonic bomber designed to meet the criterion<br />

specified by the AIAA 2001/2002 Undergraduate Team Aircraft Design Request for Proposal.<br />

The mission to be flown by the Vendetta consists of a 1,750 nautical mile radius, all of which<br />

must be flown at Mach 1.6 at or above 50,000 feet. The aircraft must have low frontal radar<br />

cross-section and also be capable of dropping a 9,000-pound weapons payload. The Vendetta is<br />

being designed to replace the stealthy F-117 Nighthawk and B-2 Spirit as well as the supersonic<br />

F-15 Eagle and B-1 Lancer. The Vendetta’s current takeoff gross weight is 125,000 pounds and<br />

its empty weight is 57,000 pounds. Furthermore, an analysis from a low observables standpoint<br />

has been made possible by publicly accessible RCS code. The component weight buildup of the<br />

Vendetta has been developed using both class I and class II methodologies as well as by<br />

assigning mass properties to an actual solid model of the aircraft. Several challenges concerning<br />

the balance of the aircraft have been introduced by the immediate and abrupt shift in the<br />

aerodynamic center due to the acceleration from subsonic to supersonic Mach numbers.<br />

Solutions to this problem have been developed and will be presented in the report.<br />

i


Table of Contents<br />

Abstract______________________________________________________________________ i<br />

Table of Contents _____________________________________________________________ ii<br />

List of Figures________________________________________________________________ iv<br />

List of Tables _______________________________________________________________ vii<br />

Nomenclature _______________________________________________________________viii<br />

1 Introduction______________________________________________________________ 1<br />

2 Defining the Design Domain ________________________________________________ 7<br />

3 Configuration ___________________________________________________________ 10<br />

4 Stealth Considerations ____________________________________________________ 16<br />

5 Aerodynamics ___________________________________________________________ 22<br />

5.1 Wing Sizing ________________________________________________________ 22<br />

5.2 Wing Planform ______________________________________________________ 23<br />

5.3 Wing Thickness _____________________________________________________ 26<br />

5.4 Airfoil _____________________________________________________________ 28<br />

5.5 Lift Curve __________________________________________________________ 28<br />

5.6 Drag_______________________________________________________________ 31<br />

6 Propulsion ______________________________________________________________ 34<br />

6.1 Engine Selection _____________________________________________________ 34<br />

6.2 Inlets ______________________________________________________________ 41<br />

6.3 S-Duct _____________________________________________________________ 46<br />

6.4 Nozzle _____________________________________________________________ 47<br />

7 Materials and Structure____________________________________________________ 48<br />

8 Landing Gear ___________________________________________________________ 52<br />

9 Weight & Balance________________________________________________________ 55<br />

10 Stability and Control____________________________________________________ 60<br />

11 Performance __________________________________________________________ 71<br />

11.1 Performance Requirements_____________________________________________ 71<br />

11.2 Specific Excess Power Requirements_____________________________________ 74<br />

11.3 Turn Rate Requirement________________________________________________ 76<br />

11.4 Mission Requirements ________________________________________________ 78<br />

11.5 Takeoff & Landing ___________________________________________________ 79<br />

11.6 Performance Summary ________________________________________________ 81<br />

11.7 Alternate Missions ___________________________________________________ 82<br />

12 Payload ______________________________________________________________ 84<br />

13 Cockpit ______________________________________________________________ 86<br />

14 Systems ______________________________________________________________ 90<br />

14.1 Auxiliary Power Generation System _____________________________________ 90<br />

14.2 Vehicle Management System ___________________________________________ 91<br />

14.3 Fuel System_________________________________________________________ 92<br />

15 Manufacturing_________________________________________________________ 94<br />

16 Cost Analysis _________________________________________________________ 96<br />

Appendix___________________________________________________________________ 98<br />

Threats Chart______________________________________________________________ 98<br />

ii


Diffuser Efficiency _________________________________________________________ 99<br />

Literal Factor Forms ________________________________________________________ 99<br />

Foldout 1 ________________________________________________________________ 101<br />

Foldout 2 ________________________________________________________________ 103<br />

The Vendetta Design Team____________________________________________________ 105<br />

References_________________________________________________________________ 107<br />

iii


List of Figures<br />

Figure 1.1 - Design Mission Profile _______________________________________________ 1<br />

Figure 1.2 - F-111 Aardvark _____________________________________________________ 3<br />

Figure 1.3 - F-15 Strike Eagle____________________________________________________ 4<br />

Figure 1.4 - F-117 Night Hawk___________________________________________________ 4<br />

Figure 1.5 - B-1B Lancer _______________________________________________________ 5<br />

Figure 1.6 - B-2 Spirit__________________________________________________________ 5<br />

Figure 2.1 - Historical Weight Fractions ___________________________________________ 7<br />

Figure 2.2 - Constraint Plot______________________________________________________ 9<br />

Figure 3.1 - Nergal ___________________________________________________________ 10<br />

Figure 3.2 - Jackhammer ______________________________________________________ 10<br />

Figure 3.3 - Interdictor ________________________________________________________ 10<br />

Figure 3.4 - Big Paulie ________________________________________________________ 10<br />

Figure 3.5 - Initial Configuration ________________________________________________ 11<br />

Figure 3.6 - Radar Return of Initial Configuration___________________________________ 12<br />

Figure 3.7 - Second Configuration _______________________________________________ 13<br />

Figure 3.8 - Current Configuration _______________________________________________ 14<br />

Figure 3.9 - Inboard Layout ____________________________________________________ 14<br />

Figure 3.10 - Inboard Layout Continued __________________________________________ 15<br />

Figure 4.1 - Stealth Considerations_______________________________________________ 16<br />

Figure 4.2 - RCS Model Faceting________________________________________________ 17<br />

Figure 4.3 - Radar Cross Section at 0º Lookup Angle ________________________________ 19<br />

Figure 4.4 - Radar Cross Sections at 15º Lookup Angle ______________________________ 20<br />

Figure 4.5 - Radar Cross Sections for a Radial Sweep________________________________ 21<br />

Figure 5.1 - Optimization of Wing Area and Aspect Ratio ____________________________ 22<br />

Figure 5.2 - Effect of Wing Leading and Trailing Edge Sweep on Aircraft RCS ___________ 24<br />

Figure 5.3 - Wing Planform ____________________________________________________ 25<br />

Figure 5.4 - Effect of Root Chord Thickness on Wing Weight and Cross Sectional Area ____ 26<br />

Figure 5.5 - Effect of Root Chord Thickness on Fuel Burn ____________________________ 27<br />

Figure 5.6 - Airfoil Section at MAC______________________________________________ 28<br />

Figure 5.7 - Airfoil Section at Tip of Trailing Edge Flap______________________________ 28<br />

Figure 5.8 – Variation in Lift Curve Slope with Mach Number_________________________ 29<br />

Figure 5.9 - Lift Distribution of Wing with and without Twist _________________________ 30<br />

Figure 5.10 - Subsonic Wing Lift Curve __________________________________________ 30<br />

Figure 5.11 - Transonic Area Distribution _________________________________________ 31<br />

Figure 5.12 - Supersonic Area Distribution (Mach 1.6) _______________________________ 32<br />

Figure 5.13 - Drag Build-Up at 50,000 ft, Mach 1.6, Maneuver Weight, and 5% Static Margin 33<br />

Figure 6.1 - VAATE Goals_____________________________________________________ 38<br />

Figure 6.2 - Thrust Curves for Altitudes from Sea Level to 70,000 ft (21,300 m)___________ 39<br />

Figure 6.3 - Military TSFC Curves for Altitudes from Sea Level to 70,000 ft (21,300 m) ____ 40<br />

Figure 6.4 - Engine Sizing Plot__________________________________________________ 41<br />

Figure 6.5 - Optimum Deflection Angle for Mach 1.6 Flow ___________________________ 42<br />

Figure 6.6 - Pressure Recovery for a Two Shock versus Three Shock Inlet _______________ 42<br />

Figure 6.7 - Inlet Area Ratio____________________________________________________ 43<br />

iv


Figure 6.8 - Cost Association with Inlet Shocks_____________________________________ 44<br />

Figure 6.9 - Off Design Area Required for Engine Mass Flow _________________________ 45<br />

Figure 6.10 - Vendetta S-Duct __________________________________________________ 46<br />

Figure 6.11 - Diffuser Angle to the Engine Face ____________________________________ 46<br />

Figure 7.1 - Structure Buildup for Vendetta ________________________________________ 48<br />

Figure 7.2 - Wing Attachment Detail _____________________________________________ 49<br />

Figure 7.3 - Empennage Structural Layout_________________________________________ 50<br />

Figure 7.4 - V-n Diagram for Vendetta____________________________________________ 50<br />

Figure 8.1 - Main Gear Structural Attachment Point _________________________________ 52<br />

Figure 8.2 - Landing Gear Configuration Trade Study________________________________ 53<br />

Figure 8.3 - Main Gear Retraction Sequence _______________________________________ 53<br />

Figure 8.4 - Nose Gear and Main Gear Retraction Schemes ___________________________ 53<br />

Figure 8.5 - Completed Landing Gear ____________________________________________ 54<br />

Figure 8.6 - Vendetta with MJ-1 Lift Truck and 2000lb JDAM_________________________ 54<br />

Figure 9.1- Principle Axes _____________________________________________________ 56<br />

Figure 9.2 - Center of Gravity Excursion __________________________________________ 58<br />

Figure 10.1 - Longitudinal X-Plot at Mach 0.3 _____________________________________ 61<br />

Figure 10.2 - Horizontal Area Required for Static Stability with Cant Angle ______________ 62<br />

Figure 10.3 - Vertical Area Required for Static Stability with Cant Angle ________________ 63<br />

Figure 10.4 - Radar Cross Section Impact of 20° vs. 30° Vertical Cant Angle _____________ 64<br />

Figure 10.5 - Vendetta Empennage Configuration ___________________________________ 66<br />

Figure 10.6 - Mach Tuck Illustrated ______________________________________________ 66<br />

Figure 10.7 - Pitch Break Characteristics __________________________________________ 68<br />

Figure 10.8 - Pheagle Simulator _________________________________________________ 70<br />

Figure 10.9 - Flight Cab and Instruments __________________________________________ 70<br />

Figure 10.10 - Graphics and Environment _________________________________________ 70<br />

Figure 10.11 - Heads up Display ________________________________________________ 70<br />

Figure 11.1 - Fuel Consumption Envelope at Average Climb Weight____________________ 72<br />

Figure 11.2 - Drag on Aircraft in Loiter Conditions__________________________________ 73<br />

Figure 11.3 - 1g Military Specific Excess Power Envelope at Maneuver Weight ___________ 75<br />

Figure 11.4 - 1g Maximum Specific Excess Power Envelope at Maneuver Weight _________ 75<br />

Figure 11.5 - 2g Maximum Specific Excess Power Envelope at Maneuver Weight _________ 76<br />

Figure 11.6 - Maneuverability Diagram at 15,000 ft (4,572 m) and Maneuver Weight ______ 77<br />

Figure 11.7 - Fuel Consumption over Mission ______________________________________ 78<br />

Figure 11.8 - MPRL with 8 × 2,000 lb (907 kg) JDAM’s _____________________________ 82<br />

Figure 12.1 - L to R configurations 1, 2, 3 _________________________________________ 84<br />

Figure 12.2 - 180 inch MPRL___________________________________________________ 84<br />

Figure 12.3 - Ballute and Sabot _________________________________________________ 84<br />

Figure 12.4 - Bomb Bay Door Retraction Scheme___________________________________ 85<br />

Figure 12.5 - 30in (76.2cm) Ejector rack __________________________________________ 85<br />

Figure 12.6 - LAU-142A Ejection Sequence _______________________________________ 85<br />

Figure 12.7 - (8) 2000lb JDAM + MPRL__________________________________________ 85<br />

Figure 13.1 - Cockpit Width Trade Study _________________________________________ 86<br />

Figure 13.2 - Forward fuselage Comparisons_______________________________________ 86<br />

Figure 13.3 - Virtual Cockpit Model _____________________________________________ 87<br />

v


Figure 13.4 – Rectilinear Vision Plot of Forward Cockpit Position______________________ 87<br />

Figure 13.5 - Cockpit Display Arrangement________________________________________ 88<br />

Figure 13.6 - Helmet Mounted HUD _____________________________________________ 88<br />

Figure 13.7 - ACES II Ejection Seat______________________________________________ 89<br />

Figure 13.8 - AFCPS__________________________________________________________ 89<br />

Figure 14.1 - Sundstrand APS 3200 Location ______________________________________ 90<br />

Figure 14.2 - Fuel Tank Locations in Vendetta _____________________________________ 92<br />

Figure 14.3 - Fuel System Architecture ___________________________________________ 92<br />

Figure 14.4 - Retractable in-flight refueling boom ports, F22, F-117, B-2 ________________ 92<br />

Figure 15.1 - Routing Tunnel ___________________________________________________ 94<br />

Figure 15.2 - Manufacturing Breaks______________________________________________ 94<br />

Figure 15.3 - Assembly Line ___________________________________________________ 95<br />

Figure 16.1 - Cost Analysis ____________________________________________________ 96<br />

Figure 16.2 - Operating Cost ___________________________________________________ 97<br />

Figure 16.3 - Lifecycle Cost ____________________________________________________ 97<br />

vi


List of Tables<br />

Table 1.I - Required Weapons Loadout ____________________________________________ 1<br />

Table 1.II - Summary of Design Requirements ______________________________________ 2<br />

Table 1.III - Comparison of the F-111, F-117, B-2, B-1B, and F-15E_____________________ 6<br />

Table 2.I - Weight Fractions & Weights____________________________________________ 7<br />

Table 2.II - Weight Fraction Assumptions __________________________________________ 8<br />

Table 2.III - Constraint Assumptions ______________________________________________ 8<br />

Table 4.I - Common Ground Radars______________________________________________ 18<br />

Table 5.I - Wing Measurements _________________________________________________ 25<br />

Table 6.I - Engine Specifications of RFP Supplied Engine ____________________________ 34<br />

Table 6.II - RFP Dimensions Compared to the Snecma Olympus _______________________ 36<br />

Table 6.III - IHPTET Goals ____________________________________________________ 36<br />

Table 7.I - Materials Selection __________________________________________________ 51<br />

Table 9.I - Initial Component Weight Buildup______________________________________ 55<br />

Table 9.II - Final Component Weight Buildup______________________________________ 56<br />

Table 9.III - Inertia Estimation __________________________________________________ 56<br />

Table 10.I - Historical Aircraft Tail Volume Coefficients _____________________________ 60<br />

Table 10.II - Pitching Moment Coupling with Rudder Deflection for Various Vertical Cant<br />

Angles _________________________________________________________________ 64<br />

Table 10.III - Rudder Control Power Results for OEI Condition________________________ 65<br />

Table 10.IV - Longitudinal and Lateral Dynamic Mode Conformity with MIL-8785C ______ 69<br />

Table 11.I - RFP Performance Requirements_______________________________________ 71<br />

Table 11.II - RFP Design Mission _______________________________________________ 71<br />

Table 11.III - Detailed Design Mission ___________________________________________ 73<br />

Table 11.IV - Performance Measures of Merit______________________________________ 74<br />

Table 11.V - Mission Results ___________________________________________________ 78<br />

Table 11.VI - Fuel Consumption by Mission Segment _______________________________ 79<br />

Table 11.VII - Takeoff Flight Profile _____________________________________________ 79<br />

Table 11.VIII - Landing Flight Profile ____________________________________________ 80<br />

Table 11.IX - Takeoff Results __________________________________________________ 80<br />

Table 11.X - Landing Results ___________________________________________________ 81<br />

Table 11.XI - Performance Summary_____________________________________________ 81<br />

Table 11.XII - Alternate Mission Results__________________________________________ 83<br />

Table 13.I - Military Vision Specifications ________________________________________ 87<br />

Table 14.I - APU Selection Table________________________________________________ 90<br />

Table 14.II - Fuel System Sizing Requirements _____________________________________ 93<br />

vii


Nomenclature<br />

AIAA American Institute of Aeronautics and Astronautics<br />

A max Maximum Cross Sectional Area, ft 2<br />

AR Aspect Ratio<br />

C D Drag Coefficient<br />

C D parasite Parasite Drag Coefficient<br />

C D0 Zero Lift Drag Coefficient<br />

C Di Induced Drag Coefficient<br />

C Dwave Wave Drag Coefficient<br />

Cf Skin Friction Coefficient<br />

CG Center of Gravity, ft, in<br />

C L Lift Coefficient<br />

C l Section Lift Coefficient<br />

C Lmax Maximum Wing Lift Coefficient<br />

D Drag, lb<br />

F F Form Factor<br />

FS Fuselage Station<br />

L Length of Fuselage, ft<br />

L Lift, lb<br />

L/D Lift to Drag Ratio<br />

L HT Lever Arm of Horizontal Tail<br />

L VT Lever Arm of Vertical Tail<br />

M Mach Number<br />

M Mach Number<br />

MAC Mean Aerodynamic Chord, ft<br />

M cd0 max Mach Number for Maximum Wave Drag<br />

M cr Critical Mach Number<br />

NPF Net Propulsive Force, lb<br />

OEI One Engine Inoperable<br />

P Pressure, lb/ft 2<br />

P s Specific Excess Power, ft/s<br />

Q Interference Factor<br />

R Gas Constant, ft 2 /s 2 R<br />

RCS Radar Cross-Section<br />

Re L Reynolds Number Based on Length<br />

Re Lcutoff Cutoff Reynolds Number<br />

RFP Request for Proposal<br />

S Sutherland’s Constant, R<br />

S Wing Reference Area, ft 2<br />

SFC Thrust Specific Fuel Consumption, lb/lb hr, lb/lb s<br />

S HT Horizontal Tail Planform Area<br />

SL, Sea Level<br />

S ref Wing Reference Area, ft 2<br />

viii


S VT Vertical Tail Planform Area<br />

S w Wing Reference Area, ft 2<br />

S wet Wetted Area, ft 2<br />

T Thrust, lb<br />

T Temperature, R<br />

T/W Thrust to Weight Ratio<br />

T available Available Thrust, lb<br />

TOGW Takeoff Gross Weight, lb<br />

T required Required Thrust, lb<br />

TSFC Thrust Specific Fuel Consumption, lb/lb hr, lb/lb s<br />

V Velocity, knots, ft/sec<br />

V H Horizontal Tail Volume Coefficient<br />

V stall Stall Velocity, knots, ft/s<br />

V TD Touchdown Velocity, knots, ft/s<br />

V TO Takeoff Velocity, knots, ft/s<br />

V V Vertical Tail Volume Coefficient<br />

W Weight, lb<br />

W Fuel Flow Rate, slugs/s<br />

F<br />

W/S Wing Loading, lb/ft 2<br />

a Speed of Sound, ft/s<br />

a Temperature Lapse Rate in Troposphere, R/ft<br />

c Chord<br />

c Mean Aerodynamic Chord, ft<br />

c Mean Aerodynamic Chord, ft<br />

w<br />

c root Root Chord, ft<br />

c tip Tip Chord, ft<br />

e Span Efficiency Factor<br />

f Component Fineness Ratio<br />

g Acceleration Due to Gravity, ft/s 2<br />

h Altitude, ft<br />

k Surface Roughness, ft<br />

k 1 Induced Drag Coefficient<br />

l Length of Drag Component, ft<br />

n Load Factor<br />

⎛ g ⎞<br />

n Atmospheric Constant, ⎜n<br />

= = 5.2561⎟<br />

⎝ aR ⎠<br />

q Dynamic Pressure, lb/ft 2<br />

s Distance, n miles, ft<br />

t Time, s<br />

t Thickness of Wing to Chord Ratio<br />

y mac y Location of Mean Aerodynamic Chord, ft<br />

Λ LE Leading Edge Sweep, degrees, rad.<br />

Λ t max Sweep of Position of Maximim Thickness, degrees, rad.<br />

Trailing Edge Sweep, degrees, rad.<br />

Λ TE<br />

ix


α Angle of Attack, degrees, rad.<br />

α eff Effective Angle of Attack, degrees, rad.<br />

α i Angle of Attack Induced by Downwash, degrees, rad.<br />

δ T Engine Correction Factor<br />

γ Specific Heat Ratio<br />

γ Climb Angle, degrees, rad.<br />

λ Taper Ratio<br />

µ Dynamic Viscosity, lb s/ft 2<br />

µ brake Braking Coefficient of Friction<br />

µ roll Rolling Coefficient of Friction<br />

θ T Engine Correction Factor<br />

ρ Density, slugs/ft 3<br />

ψ Yaw Angle, degrees, rad.<br />

Subscripts<br />

0 Sea-Level Value<br />

∗ Value at Tropopause<br />

x


1 Introduction<br />

Every year, the American Institute of Aeronautics and Astronautics (AIAA) sponsors collegiate<br />

design competitions. The request for proposal (RFP) for the 2001-2002 team undergraduate<br />

aircraft design competition outlined the requirement for a stealth supersonic interdictor. The<br />

interdictor is introduced with a design mission as shown in Figure 1.1. The payload specified for<br />

this design mission is shown in Table 1.I. Because multiple weapon load outs are specified, it is<br />

clear that this, as with any modern aircraft, must be suited to more than one role.<br />

The RFP also gives many requirements for the aircraft. These include operating constraints as<br />

well as performance requirements. The design requirements are summarized in Table 1.II.<br />

External tanks may be used but must be retained for the duration of the flight. Bomb pylons may<br />

also be used suggesting the possibility of a non-stealth configuration. Another Important factor<br />

is that the aircraft must cost less than 150 million dollars. This is a very small price tag for an<br />

aircraft of this size and complexity.<br />

Figure 1.1 - Design Mission Profile<br />

Table 1.I - Required Weapons Loadout<br />

Loading # (Quantity) Weapon<br />

1 - Design (4) 2000 lb (907 kg) JDAM + (2) AIM-120<br />

2 (4) Mk-84 LDGP + (2) AIM-120<br />

3 (4) GBU-27 + (2) AIM-120<br />

4 (4) AGM-154 JSOW + (2) AIM-120<br />

5 (16) 250 lb (113 kg) Small Smart Bomb<br />

1


Table 1.II - Summary of Design Requirements<br />

Area Design Requirement Value (if applicable)<br />

Misc.<br />

Crew<br />

500 lb (227 kg), 2 pilots, single pilot operation<br />

Structure<br />

Fuel<br />

Stability<br />

Observables<br />

Operation<br />

Positive g’s<br />

7 (50% Internal Fuel)<br />

Negative g’s<br />

3 (50% Internal Fuel)<br />

Dynamic Pressure<br />

2,133 psf (102 kPa)<br />

Factor Of Safety 1.5<br />

JP-8<br />

Self Sealing<br />

Static Margin 10% to – 30%<br />

Active Flight Controls for Unstable Aircraft<br />

RCS (Front Aspect)<br />

0.05 m 2 , frequency range 1 – 10 GHz<br />

Balanced IR, Visual, Acoustical, RCS<br />

Internal Stores<br />

Runway Length 8,000 ft (2,438 m)<br />

Operate from NATO Airports<br />

All Weather Weapons Delivery<br />

Cost<br />

Max Cost<br />

Minimize Life Cycle Costs<br />

$150 Million, 2000 dollars<br />

Performance<br />

Supercruise Mission Radius<br />

Specific Excess Power<br />

1-g, Mach 1.6, 50,000 ft, Dry<br />

1-g, Mach 1.6, 50,000 ft, Wet<br />

2-g, Mach 1.6, 50,000 ft, Wet<br />

Instantaneous Turn Rate, Mach 0.9, 15,000 ft<br />

1,750 nm (3,240 km)<br />

0 ft/sec (0 m/s)<br />

200 ft/sec (61 m/s)<br />

0 ft/sec (0 ft/s)<br />

8 deg/sec<br />

2


The RFP also lists several aircraft that collectively fulfill the mission of the proposed interdictor.<br />

These are each outlined below.<br />

F-111 - “Aardvark”<br />

The F-111 (Figure 1.2, Table 1.III) is specifically mentioned as the predecessor to the aircraft<br />

requested in the RFP. The F-111 officially entered service in 1967 and was retired in 1996; a<br />

replacement is badly needed. It has been partially replaced by several aircraft, each outlined in<br />

detail in the sections to follow. The F-111 is a very large aircraft capable of carrying a 31,000 lb<br />

(14,061 kg) payload over 2,000 nm (3,704 km). Both the payload and combat radius are large<br />

thus yielding a 91,000 lb (41,276 kg) aircraft. Though the F-111 is capable of Mach 2.2, but it<br />

does not cruise supersonically. The F-111 was designed for a very different mission than the one<br />

outlined in the RFP. The F-111 was actually designed to cruise subsonically to the target area,<br />

dash in supersonically at low level, drop its payload, and fly out of the threat area quickly. After<br />

retiring the aircraft, the air force decided a new aircraft was needed to drop precision weapons<br />

from remote airfields with minimal support.<br />

F-15E - “Strike Eagle”<br />

Figure 1.2 - F-111 Aardvark<br />

The F-15E Strike Eagle (Figure 1.3, Table 1.III) partially filled the role of the F-111 after it was<br />

retired. The F-15E was designed to be capable of both air superiority and ground attack<br />

missions. Superior maneuverability was achieved with the F-15E due to its high thrust-to-weight<br />

ratio and low wing loading.<br />

3


Figure 1.3 - F-15 Strike Eagle<br />

F-117 – “Night Hawk”<br />

The Night Hawk (Figure 1.4, Table 1.III) also aided in the replacement of the F-111. However,<br />

it has a vastly reduced payload capacity and a limited range. The F-111 is also not capable of<br />

supersonic speeds and is thus more vulnerable if it were detected. If a supersonic aircraft were<br />

detected, the window of opportunity for an attack is relatively small. Thus, faster aircraft have a<br />

tendency to be less venerable. Due to the small payload and high maintenance of the first<br />

generation stealth technology, the F-117 is a poor replacement for the F-111.<br />

Figure 1.4 - F-117 Night Hawk<br />

4


B-1B – “Lancer”<br />

The B-1 has never been a successful aircraft. Until the war against the Taliban, the B-1B was<br />

never used in combat. Originally it was designed to have a mission much like the one specified<br />

in the RFP. When the proposal for a high altitude supersonic bomber was withdrawn, politics<br />

wouldn’t allow the B-1 to vanish. The aircraft was converted into the B-1B. The B model was<br />

designated as a low level subsonic bomber much like the F-111 was. The B-1 is an expensive<br />

aircraft to operate due to its vast size. A picture and table of data are also provided for this<br />

aircraft (Figure 1.5, Table 1.III).<br />

B-2 – “Spirit”<br />

Figure 1.5 - B-1B Lancer<br />

The B-2 (Figure 1.6, Table 1.III) is a large stealth subsonic bomber. It is designed to carry vast<br />

amounts of payloads long distances undetected. The B-2 is a large aircraft that is very costly to<br />

operate.<br />

Figure 1.6 - B-2 Spirit<br />

5


Table 1.III - Comparison of the F-111, F-117, B-2, B-1B, and F-15E<br />

Manufacturer Lockheed<br />

General<br />

Dynamics<br />

Boeing Northrop Rockwell<br />

Type F-117 FB-111A F-15E B-2 B-1B<br />

b - ft<br />

(m)<br />

43.6<br />

(13.3)<br />

32.0<br />

(9.2)<br />

42.8<br />

(13.0)<br />

172.0<br />

(52.4)<br />

78.2<br />

(23.8)<br />

AR – – 3 – –<br />

L – ft<br />

(m)<br />

H – ft<br />

(m)<br />

Wing - ft 2<br />

(m 2 )<br />

W oe – lb<br />

(kg)<br />

W pl – lb<br />

(kg)<br />

W f – lb<br />

(kg)<br />

W to – lb<br />

(kg)<br />

66.6<br />

(20.3)<br />

12.5<br />

(3.8)<br />

913<br />

(84.8)<br />

29,500<br />

(133,801)<br />

5,000<br />

(2,267)<br />

73.5<br />

(22.4)<br />

17.1<br />

(5.21)<br />

–<br />

46,171<br />

(20,943)<br />

31,500<br />

(14,488)<br />

– –<br />

52,501<br />

(23,814)<br />

91,492<br />

(41,500)<br />

63.7<br />

(19.4)<br />

18.5<br />

(5.6)<br />

608<br />

(56.5)<br />

32,000<br />

(14,515)<br />

24,500<br />

(11,113)<br />

13,122<br />

(5,952)<br />

81,000<br />

(36,740)<br />

69.0<br />

(21)<br />

17.0<br />

(5.2)<br />

5274<br />

(490.0)<br />

153,700<br />

(69,717)<br />

40,001<br />

(18,144)<br />

200,003<br />

(90,720)<br />

375,998<br />

(170,550)<br />

147.0<br />

(44.8)<br />

34.0<br />

(10.4)<br />

1950<br />

(181.1)<br />

192,001<br />

(87,090)<br />

133,999<br />

(60,780)<br />

194,999<br />

(88,450)<br />

477,003<br />

(216,365)<br />

Max power loading – – 1.73 4.86 –<br />

Max level (Mach) 0.9 2.2 2.5 0 1.25<br />

Max combat radius<br />

nm<br />

Service Ceiling - ft<br />

(m)<br />

570 2,750 686 6,300 6,479<br />

–<br />

50,853<br />

(15,500)<br />

–<br />

50,000<br />

(15,240)<br />

The solution to the RFP is not a trivial one. The aircraft will have to be well area ruled and have<br />

a low frontal cross-section in order to minimize wave drag. The goal of this design is to meet or<br />

exceed RFP requirements while minimizing manufacturing and operating costs.<br />

–<br />

6


2 Defining the Design Domain<br />

The first estimation of aircraft weight used the iterative weight fraction method outlined in<br />

Roskam. This method calculates the weight fraction for each mission segment. The initial<br />

takeoff weight was guessed and the resulting empty weight was calculated. The resulting weight<br />

generally yields an unrealistic total weight fraction. In order to check the realism of the aircrafts<br />

weight fraction, a database of aircraft similar in mission was compiled. The aircraft gross<br />

takeoff weight was iterated until the total weight fraction landed on the historical trend. This<br />

trend is shown in Figure 2.1. The weights generated by this method include gross takeoff<br />

weight, fuel weight, and landing weight. The results of the weight fraction method are shown in<br />

Table 2.I. The assumptions made to generate the weight fractions are shown in Table 2.II.<br />

Vendetta Estimated Weight Fractions<br />

Cruise Back<br />

16%<br />

Reserve<br />

6%<br />

Misc.<br />

5%<br />

Warm -up &<br />

Takeoff<br />

6% Initial Climb<br />

11%<br />

Dash Back<br />

14%<br />

Dash Out<br />

17%<br />

Cruise Out<br />

25%<br />

Figure 2.1 - Historical Weight Fractions<br />

Table 2.I - Weight Fractions & Weights<br />

Mission Segment Weights<br />

Weights<br />

Start/Takeoff 6% Takeoff 108,400 lb (49,169 kg)<br />

Climb To Cruise 11% Empty 51,600 lb (23,405 kg)<br />

Cruise-Out 25% Fuel 47,600 lb (21,591 kg)<br />

Dash-out 17% Payload 9,054 lb (4,107 kg)<br />

Dash-Back 14% Fuel Weight Fraction 47.6%<br />

Cruise-Back 16%<br />

45 Minute Reserve 6%<br />

Misc. 5%<br />

Total 100%<br />

7


Many of the equations buried in the weight fraction method depend on assumed parameters. In<br />

many cases, values were assumed using figures and tables from Roskam, Nicolai, and Raymer.<br />

The assumptions used are listed in Table 2.2. The weight fraction method is by no means an<br />

accurate method for initial sizing. Inaccuracies of up to 10% are possible depending on the<br />

quality of the initial assumptions, and 20% is not uncommon for unusual missions such as the<br />

one outlined in the RFP. The weight fraction method may not be accurate for this type of aircraft<br />

due to the lack of similar aircraft in the database. There are really only three supercruise aircraft,<br />

the SR-71, YF-23, and the F-22. These aircraft all have a vastly different mission and may yield<br />

invalid sizes for the interdictor. Though the method may be flawed, it was used anyway due to<br />

the lack of a better method.<br />

Weight fractions provide a starting point for the weight of a proposed aircraft; however, the<br />

physical dimensions are not predicted. In order to determine the physical size, constraint plots<br />

were created. A constraint plot examines the relationship between two variables based on given<br />

requirements. Generally, the two variables used are wing loading and thrust to weight ratio.<br />

The RFP gives many constraints as shown earlier in Table 1.II. The majority of constraints can<br />

be written as functions of wing loading and thrust-to-weight ratio. This is the reason that it is a<br />

popular type of constraint plot. The equations for range, takeoff distance, and many others were<br />

found in Roskam, Nicolai, and Raymer. Many more assumptions were made to create the<br />

constraint plot; these are shown in Table 2.III.<br />

Table 2.II - Weight Fraction Assumptions<br />

SFC _Cruise 1.11<br />

SFC _Dash 1.11<br />

SFC _Turn 1.11<br />

SFC _Loiter 0.8<br />

L/D Cruise 10<br />

L/D Dash 10<br />

L/D Turn 10<br />

L/D Loiter 12<br />

Table 2.III - Constraint Assumptions<br />

C Lmax_TO 1.8<br />

C Lmax_CR 1.2<br />

C LCruise 0.2<br />

AR 3<br />

e 0.8<br />

The constraint equations show how thrust to weight ratio and wing loading relate to a given<br />

performance constraint. This allows engineers to determine which combinations of thrust to<br />

weight ratio and wing loading are acceptable. The constraint plot for the RFP is shown in Figure<br />

2.2. Note that any design point on the shaded side of a line would not meet the design<br />

requirements. This again depends on the accuracy of the aforementioned assumptions. The<br />

constraint plot clearly identifies a design domain, in which, any combination of thrust-to-weight<br />

ratio and wing loading will satisfy the design requirements.<br />

Combined with the weight fraction method, the constraint plot shows the physical size of the<br />

airplane. Because a preliminary weight was determined from the weight fraction method, the<br />

8


wing loading and thrust to weight ratio may be converted to wing area and thrust required. Once<br />

this information is found the designer may start the configuration layout and choose a power<br />

plant. In order to obtain the smallest aircraft, the design point should be at the highest wing<br />

loading and lowest thrust to weight as possible on the constraint plot. As the design point moves<br />

to the left on the constraint plot, the wing area increases. This yields a larger airframe. Large<br />

airframes are expensive and cost more to maintain.<br />

As the design point moves up on the constraint plot, the engines get bigger and heavier. This is<br />

not desired as the aircraft will burn more fuel and have to fly at a higher lift coefficient for a<br />

given Mach number. This would cause the fuel burn for a given mission and thus increase<br />

operating costs. If the design point deviates from the lower right, an explanation is required.<br />

The initial design point was chosen in the center of the design domain in order to allow for<br />

aircraft growth. The design point was chosen away from constraint lines to make the aircraft<br />

design performance less sensitive to changes in weight.<br />

10<br />

0.9<br />

0.8<br />

Current Design<br />

Point<br />

Takeoff<br />

Landing<br />

ψ<br />

Range<br />

Thrust to Weight Ratio<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

Initial Design<br />

Point<br />

Specific<br />

Excess<br />

Power<br />

Specific<br />

Excess<br />

Power<br />

40<br />

60<br />

80<br />

100 120 140<br />

Wing Loading, (psf)<br />

160<br />

180<br />

200<br />

Figure 2.2 - Constraint Plot<br />

9


3 Configuration<br />

The current configuration of the aircraft was developed through several iterations. The first<br />

iterations were individual designs developed by each of the team members. Four different<br />

designs were considered for the configuration the Vendetta would take and each was evaluated to<br />

a similar level of detail.<br />

The first of these designs was Nergal (Figure 3.1). This<br />

was a concept that utilized thrust vectoring for stability<br />

in the yaw axis. This configuration utilized a rotary<br />

bomb bay configuration and a side-by-side cockpit<br />

arrangement. This was the only design based on a<br />

single PW-F119 as its power plant.<br />

Figure 3.2 - Jackhammer<br />

Figure 3.1 - Nergal<br />

The Jackhammer shown in Figure 3.2 was another tailless<br />

design. This aircraft differed from the Nergal only in its canard<br />

arrangement and twin-engine approach. It included a rotary<br />

bomb bay as well as F119 engines as its primary power plants.<br />

The flight deck was a side-by-side layout that provided good<br />

visibility for the pilots.<br />

The Interdictor, shown in Figure 3.3, was based on the RFP engines<br />

and sported good propulsive efficiency due to the lack of an S-style<br />

duct. Like the previous aircraft, the Interdictor included a side-by-side<br />

cockpit.<br />

Figure 3.3 - Interdictor<br />

The Big Paulie (Figure 3.4) had ample fuel volume and<br />

utilized RFP engines. It was outfitted with a side by<br />

side cockpit and ACES II ejection seats.<br />

Figure 3.4 - Big Paulie<br />

10


The downselect between these aircraft was not difficult. This early work clearly showed that the<br />

engine provided by the RFP was far too large for the thrust it provided. The two aircraft<br />

designed for the F119 were both smaller and more space efficient. This narrowed the downselect<br />

to the Nergal and Jackhammer. Both of these aircraft were tailless and it was determined that<br />

the weight and drag benefits associated with the lack of a vertical tail would be outweighed by<br />

the costs associated with the thrust vectoring system. It was also determined that the aircraft<br />

were too unstable laterally to be controlled by an inexpensive, low bandwidth, thrust vectoring<br />

system. It was decided to begin with a new design incorporating the strong points of each<br />

aircraft.<br />

The first iteration of the aircraft is shown in Figure 3.5; it is a large aircraft that has many design<br />

flaws. The first and most obvious is the above-chine mounted inlet. This is easily seen in the<br />

aircraft’s front view. The chine causes a vortex roll-up that would be directly ingested by the<br />

inlet at high angles of attack. This is not desired, as it would cause poor and unpredictable<br />

performance at nearly any angle of attack. A low bypass ratio engine might tolerate these flow<br />

disruptions without problems however; the design utilizes a new engine with a bypass ratio of<br />

approximately 1.5. This type of engine will not tolerate the poor inlet location.<br />

• Span = 50 ft<br />

• m.a.c. = 23 ft<br />

• S ref = 965 sq. ft<br />

• TOGW = 121,600 lb<br />

• Empty Weight = 62,000 lb<br />

50’<br />

19’<br />

105’<br />

23’<br />

Figure 3.5 - Initial Configuration<br />

Another downfall to the initial configuration was the weight distribution. The fuel center of<br />

mass was not near the empty weight center of mass. This caused the aircraft to take off very<br />

stable and land very unstable. This could not be remedied due to the small volume available for<br />

fuel in the aft portion of the fuselage. The majority of the fuel volume in the aft portion of the<br />

aircraft was located around the engines. This is undesirable due to the possibility of a<br />

catastrophic failure of the engine fan disk or afterburner.<br />

11


Another problem arises from the 20° facet on the bottom of the fuselage. This created a large<br />

radar footprint underneath the aircraft, as shown in Figure 3.6. The vertical stabilizer also<br />

created a very radar visible configuration. The final flaw that drives the aircraft to the new<br />

configuration is the pitching moment characteristic of the fuselage. The side by side seating<br />

arrangement of the first iteration caused the fuselage to be excessively large in the areas forward<br />

of the aircraft’s neutral point. The pitch up tendencies of the aircraft grew very large with very<br />

small angles of attack. The control power of the horizontal surfaces was deemed unacceptable to<br />

combat this. The current configuration has a tandem cockpit arrangement, which will be detailed<br />

later in the report, to help solve some of these issues.<br />

Arc fuselage top creates<br />

steps caused by facets<br />

Less sensitive to<br />

lower frequency radar<br />

as wavelengths increase<br />

and can’t distinguish facets<br />

0<br />

1 GHz<br />

5 GHz<br />

10 GHz<br />

RCS in dBm 2<br />

Flat bottom<br />

Horizontal fuselage<br />

intersection creates a<br />

sharp cavity<br />

20° bottom facet<br />

Figure 3.6 - Radar Return of Initial Configuration<br />

The second configuration, shown in Figure 3.7, is very different than the first. The<br />

configuration features many changes that aide in solving the previously discussed problems. The<br />

cockpit was changed to a tandem arrangement as well as the addition of two canted tails in the<br />

place of the single vertical. The engines moved to the top of the fuselage to avoid detection from<br />

infra red sensors. The take off gross weight was decreased to 114,000 lb due to an improved<br />

engine deck and aerodynamics.<br />

12


• Span = 53 ft<br />

• m.a.c. = 32 ft<br />

• S ref = 1500 sq. ft<br />

• TOGW = 114,000 lb<br />

• Empty Weight = 55,000 lb<br />

35°<br />

53’<br />

19’<br />

98’<br />

18’<br />

13°<br />

Figure 3.7 - Second Configuration<br />

This configuration was generated with a different mentality than the previous airframe. The<br />

center of gravity was known before the first part was placed on the aircraft and every effort was<br />

utilized to keep it in the appropriate place. The weight and balance issues, though still present,<br />

were dramatically improved. The fuel load and payload compartment reside directly on the<br />

desired center of gravity, however, the empty weight was too far aft. The low mounted wing<br />

proved to be a structural challenge when incorporating a landing gear well. Another issue dealt<br />

with the cruise angle of attack. It was shown that the aircraft would cruise at approximately 4<br />

degrees. The forward chine on the fuselage would be shedding a vortex throughout the cruise<br />

portion of the mission resulting in higher drag. The chine angle should meet the onset flow<br />

angle. This prompted another revision to the aircraft.<br />

The design was further refined and the current iteration was created. The current iteration has no<br />

red flags and performs the mission well. The current iteration of the Vendetta is shown in Figure<br />

3.8 and detailed in Foldout 1 of the Appendix. It has a forward chine of 4 degrees and sound<br />

structural load paths, which will be discussed in detail later in this report. The Vendetta has<br />

grown a small amount and currently weighs 124,000 lb (56,364 kg).<br />

The aircraft has a tandem cockpit supported by a very long nose. The long nose offsets the mass<br />

of the large engines and the massive structure required for the full flying horizontal stabilizers.<br />

The APU is located in the engine compartment keeping the fuel and fire retardant systems as<br />

redundant as possible. The inlets are under wing mounted to keep them in clean flow throughout<br />

the flight envelope. The Vendetta has a 1500 ft 2 (139 m 2 ) wing area with a leading edge sweep<br />

of 40 degrees. The design drivers will be discussed in detail throughout following sections. The<br />

inboard layout can be seen in Figure 3.9 and 3.10.<br />

13


• Span = 54.7 ft<br />

• m.a.c. = 32 ft<br />

• Length = 103 ft<br />

• Root = 46.75 ft, 3% Thick<br />

• Tip = 3%<br />

• Λ c/4 = 27°<br />

• Λ LE = 40°<br />

• S ref = 1500 sq. ft<br />

• TOGW = 129,000 lb<br />

• Empty Weight = 60,000 lb<br />

55’<br />

103’<br />

51°<br />

20’<br />

13°<br />

Figure 3.8 - Current Configuration<br />

APU<br />

Weapons Bay<br />

Retracted<br />

Gear<br />

Engines<br />

Retracted<br />

Gear<br />

Figure 3.9 - Inboard Layout<br />

14


Wing Tank X 2<br />

6,366 lb each<br />

70% Volume Usage<br />

Forward Fuselage Tank<br />

23,069 lb<br />

80% Volume Usage<br />

Aft Fuselage Tank<br />

23,544 lb<br />

80% Volume Usage<br />

Tandem Cockpit<br />

Figure 3.10 – Inboard Layout Continued<br />

Full Flying<br />

Horizontal<br />

15


4 Stealth Considerations<br />

As mentioned by the RFP, the aircraft is required to meet a stealth requirement. This is a very<br />

important driver for the aircraft. The entire design is influenced by this consideration equally as<br />

much as aerodynamics. There are many low observable considerations to be taken into account.<br />

The first and most obvious is the radar cross section (RCS).<br />

From the aspect of RCS, there are many drivers for an aircraft. The majority of the radar return<br />

comes from the shaping of the aircraft. The fuselage is constructed from flat sides and constant<br />

radius curves. The sides are kept at a 60° angle from the horizontal and the bottom is kept flat<br />

(Figure 4.1). This is desired because, as later shown, the footprint of the aircraft remains small.<br />

Another feature is the canted tails. This keeps the surfaces in the empennage section from<br />

creating 90 degree angles. This is important because the 90 degree angle would radiate RF<br />

energy directly back in the direction of the source. The leading edge sweep is 40°. This creates<br />

spikes well off of the frontal aspect of the aircraft. All other leading edges are kept swept at this<br />

same angle in order to minimize the magnitude of the frontal spoke. The 15° look up angle was<br />

considered the most important aspect of the RCS. The majority of the threat encountered will be<br />

below the Vendetta. This implies that they will be looking up at the aircraft, not from the front.<br />

This is where the majority of the stealth considerations were taken into account. (See Figure 4.1)<br />

The Low observability requirements are not only for RCS. In fact, the RFP specifically specifies<br />

“Balanced Observables”. Aside form RCS, IR accounts for the next highest threat. Emissivity<br />

matching can reduce the IR signature of the aircraft. The Vendetta will be coated with a material<br />

with similar emissivity as the surroundings, aiding in the disappearance of the aircraft to an IR<br />

sensor. The actual odds of becoming invisible to the IR sensor are fairly unrealistic due to the<br />

cold surroundings. The aircraft is shadowed by what is essentially space at 50,000 feet. It is<br />

hard to hide a warm object when backlit by a cold space. The other stealth consideration in this<br />

area is the nozzles. These are axisymmetric nozzles that are proven to have lower IR signatures<br />

than the axis symmetric option. This, combined with the frontal look-up threat direction<br />

minimizes the impact on IR stealth.<br />

40° LE Sweep<br />

All other Surfaces<br />

Matched<br />

Hidden Canted<br />

Verticals<br />

60° Facet<br />

Figure 4.1 - Stealth Considerations<br />

16


Another area of concern is that of visual observability. This is not a very big problem as the<br />

aircraft cruises so high that visually detecting the Vendetta would be near impossible. The only<br />

threat here is the contrails left by the engines. The contrails can be minimized by contrail<br />

avoidance techniques and don’t pose a large problem.<br />

To quantitatively analyze the radar cross section of the Vendetta, Radbase2 software by Surface<br />

Optics was utilized. First, a faceted model was generated from the 3D model. Faceting was<br />

limited to only those necessary because of the demanding processing requirements. Facets were<br />

limited to 10 degree tolerances at roughly 0.017 feet minimums. The facetted model is presented<br />

as Figure 4.2.<br />

Figure 4.2 - RCS Model Faceting<br />

It can be seen that heavy facet optimization was needed to make sure that all facets followed<br />

tangency requirements to leave smoothly curved and splined surfaces. The spline arc on the top<br />

of the fuselage is modeled with facets every 10°. For the flat surfaces like the wings and<br />

empennage 10° is more than adequate.<br />

The Radbase2 RCS code calculates the radar returns based on Physical Optics and Chu-Stratton<br />

integral methods. These are highly computationally intensive. Because of this, bounces off of<br />

surfaces were limited to 2 after the initial bounce off the surface. This was deemed adequate for<br />

this level of analysis. The vertical-vertical polarization of the return and transmission was<br />

analyzed as it is the most relevant to how radar stations operate. Monostatic radars which both<br />

broadcast and receive were used in the analysis as there would be too many possibilities to<br />

calculate for bistatic radars.<br />

The code was allowed to iterate on the model with 1° azimuth increments and for 0° and 15°<br />

lookup angles. It was also run for 1, 5, 10, and 12 GHz radar frequencies. Most fast track and<br />

search radar runs at the higher frequencies while long range threat radars utilize the lower<br />

frequencies. The 1 to 10 GHz range covers most of the radars that are expected for the role of<br />

this aircraft. A table of common ground and surface radars with their respective frequencies is<br />

presented as Table 4.I.<br />

17


Table 4.I - Common Ground Radars<br />

Radar Manufacturer Frequency Image<br />

AN/TPS-43E Mobile<br />

Radar<br />

Westinghouse<br />

2.9 to 3.1 GHz<br />

AN/TPS-70<br />

Fixed Ground Radar<br />

Northrop<br />

Grumman<br />

2.9 to 3.1 GHz<br />

AN/SPS-49<br />

Typical Long Range<br />

Naval Radar<br />

Navy Research<br />

Labs<br />

850 to 942<br />

MHz<br />

AN/SPS-55<br />

Long Range Surface<br />

Search Radar<br />

ISC Cardion<br />

9.05 to 10.0<br />

GHz<br />

Data is not readily available for radars made by foreign manufacturers. However, these radars<br />

should be adequate for this level analysis because the properties for radar waves traveling<br />

through the air over long distances are similar.<br />

The 1 to 12 GHz range covers FM and XM radar bands which are most common threats. The<br />

RFP specifically requires that the Vendetta has a frontal RCS of 0.05 m 2 . As the threat chart<br />

shown in Appendix A shows, most threats will be from below and at shallow angles of about 15°<br />

while at 50,000 during ingress. Because of this, the 0° and 15° lookup angles were analyzed. The<br />

results of the Radbase2 software are illustrated first in Figure 4.3, which depicts the radar cross<br />

section of the aircraft from a frontal, or 0° lookup angle.<br />

18


30<br />

20<br />

10<br />

0<br />

-10<br />

-20<br />

-30<br />

-40<br />

-50<br />

1 GHz<br />

5 GHz<br />

10 GHz<br />

12 GHz<br />

RFP Requirement<br />

Figure 4.3 - Radar Cross Section at 0º Lookup Angle<br />

Figure 18 shows that the vehicle does clearly meet the frontal RCS requirement of 0.05 m 2 (-12<br />

dB) set forth in the Request for Proposal. It also shows that the measures taken at shaping the<br />

aircraft are working. The leading edge and trailing edge of the wing come together closely. There<br />

is a large return directly from the side of the aircraft due to the wing tip and fuselage side. It can<br />

also be seen that although there are slight variations in the returns due to the different<br />

frequencies, they do not vary that much. This is due to the fact that the Vendetta is a rather large<br />

vehicle. None of the surfaces are small enough to interfere with the wavelengths of the radar.<br />

The weakest azimuth angle for the Vendetta is the 40° angle where the leading edge sends a large<br />

spike forward.<br />

Looking at the equally crucial 15° lookup angle cross section in Figure 4.4 reveals a slightly<br />

different picture.<br />

19


30<br />

20<br />

10<br />

0<br />

-10<br />

-20<br />

-30<br />

-40<br />

-50<br />

1 GHz<br />

5 GHz<br />

10 GHz<br />

12 GHz<br />

RFP Requirement<br />

Figure 4.4 - Radar Cross Sections at 15º Lookup Angle<br />

Figure 4.3 shows that the Vendetta meets and exceeds the 0° lookup angle returns. This is seen as<br />

highly advantageous. The shape of the bottom of the aircraft is effective in keeping spikes at a<br />

minimum. As mentioned earlier, this is a crucial area for the Vendetta. As most of its threats are<br />

from the ground, it is important that the aircraft has a limited return in this orientation.<br />

20


The software was also utilized to generate an RCS butterfly plot in a sweep around the vehicle to<br />

determine the footprint that it will leave as it flies above its threats. Figure 4.5 shows this sweep.<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

-10<br />

1 GHz<br />

5 GHz<br />

10 GHz<br />

12 GHz<br />

Figure 4.5 - Radar Cross Sections for a Radial Sweep<br />

It can be seen that the 30° facets on the bottom of the fuselage are deflecting radar away from the<br />

vulnerable lookup orientation. The aircraft is still producing a rather large return of almost 40 dB<br />

in this position, however. Once again there is little variation in the returns for various<br />

frequencies.<br />

It is important to note that the addition of radar absorbing material (RAM) would further reduce<br />

some of the returns on the aircraft. Also, currently the RCS software is treating the entire aircraft<br />

and all of its parts as purely reflective metal surfaces. This is a conservative approach. RAM<br />

could be applied in actuality to reduce some of the returns on the bottom and front of the aircraft.<br />

21


5 Aerodynamics<br />

The major aerodynamic aspects of the Vendetta are cruise lift-to-drag ratio and maximum<br />

subsonic lift coefficients for landing performance. Wing design is critical to both of these<br />

aerodynamic aspects as well as radar cross section. Area ruling was used to minimize the<br />

supersonic wave drag on the aircraft and minimize the aircraft’s fuel consumption over the<br />

mission.<br />

5.1 Wing Sizing<br />

The first aerodynamic aspects of the aircraft that were considered were the wing planform area<br />

and aspect ratio. To select the optimum wing planform area and aspect ratio, the effect of these<br />

two parameters on the specific excess power and fuel consumption over the design mission was<br />

studied. The 1g military specific excess power at an altitude of 50,000 ft (15,240 m) and Mach<br />

number of 1.6 was estimated using engine data and drag estimation based on component skin<br />

friction drag and area ruling. The fuel consumption of the aircraft was estimated by numerically<br />

integrating the engine fuel flow from over the design mission. The additional weight and<br />

maximum cross sectional area of larger wing areas was considered in calculations, however the<br />

mission profile was kept constant and the fuel weight at takeoff was kept constant at 59,250 lb<br />

(26,875 kg). The results shown in Figure 5.1 indicate that a wing planform area of<br />

approximately 1,500 ft 2 (139 m 2 ) and aspect ratio of 2 would maximize specific excess power<br />

and minimize fuel consumption.<br />

97<br />

96<br />

1,600 ft 2<br />

Design Point<br />

1,500 ft 2<br />

Fuel Onboard<br />

Specific Excess Power (ft/s)<br />

95<br />

94<br />

93<br />

92<br />

1,700 ft 2<br />

1,400 ft 2<br />

1,300 ft 2<br />

1,900 ft 2 1,200 ft 2<br />

1,100 ft 2<br />

Wing Area<br />

91<br />

1,800 ft 2 2,000 ft 2 Aspect Ratio 1.6<br />

2.2<br />

2.4<br />

2<br />

1.8<br />

1,000 ft 2<br />

90<br />

57,000 58,000 59,000 60,000 61,000 62,000 63,000 64,000 65,000 66,000<br />

Fuel Consumption over Mission (lb)<br />

Figure 5.1 - Optimization of Wing Area and Aspect Ratio<br />

22


5.2 Wing Planform<br />

The next aspect of the wing that was considered was the leading and trailing edge sweep angles.<br />

Because any edges on an aircraft reflect radar energy, the sweep angles of the edges of the wing<br />

were chosen to minimize radar energy reflected back to the source, especially in the frontal<br />

aspect of the aircraft where a specific RCS requirement is given by the RFP. To avoid reflecting<br />

radar toward the front of the aircraft, the leading and trailing edges of the wing had to be highly<br />

swept. In addition, the sweep angles could not be approximately 45º because a corner reflector<br />

would be created. These requirements led to a diamond shaped wing planform with leading and<br />

trailing edge wing sweeps of approximately 40º. Two initial designs were considered one having<br />

a 40º swept leading edge and a 30º forward swept trailing edge and the other having matched<br />

35.3º leading edge and trailing edge sweeps. A trade study was performed to select between<br />

these two wing configurations by studying the effect of the two configurations on RCS and<br />

aerodynamics. Figure 5.2 shows a comparison of radial sweeps of both configurations using<br />

RadBase2. The return from the 40º and 35.3º leading edge sweeps can be clearly seen in the<br />

plot. The leading edge spike on the matched leading and trailing edge configuration is<br />

approximately 15 dB lower than the other configuration; however it is 5º closer to the frontal<br />

aspect of the aircraft. The aerodynamic study of the two wing configurations indicated that<br />

approximately 1,000 lb (454 kg) of additional fuel would be required due to the additional wave<br />

drag of the lower leading edge sweep angle. Because of the aerodynamic benefits and because<br />

the RFP only gives frontal aspect RCS requirements, the 40º leading edge and 30º trailing edge<br />

configuration was chosen.<br />

23


1 GHz. 40º LE Sweep<br />

10 GHz. 40º LE Sweep<br />

1 GHz. 35.3º LE Sweep<br />

10 GHz. 35.3º LE Sweep<br />

RFP Requirement (-12 dB)<br />

50 dB<br />

40 dB<br />

30 dB<br />

20 dB<br />

10 dB<br />

35.3º LE Sweep<br />

40º LE Sweep<br />

0 dB<br />

-10 dB<br />

-20 dB<br />

-30 dB<br />

-40 dB<br />

-50 dB<br />

Figure 5.2 - Effect of Wing Leading and Trailing Edge Sweep on Aircraft RCS<br />

Once the wing area, aspect ratio, and sweep angles were chosen, the tip chord was kept at 8 ft<br />

(2.4 m) to avoid an overly small tip chord that could interact with radar wavelengths<br />

unpredictably. This resulted in the wing planform shown in Figure 5.3, with the measurements<br />

given in Table 5.I. Leading and trailing edge flaps, and ailerons were added to the wing. The<br />

chord high lift devices and control surfaces were kept at a constant percentage of the mean<br />

aerodynamic chord so that the hinge lines would parallel to the wing edges. The trailing edge<br />

flap chord is 20% of the mean aerodynamic chord and the leading edge flap and aileron are each<br />

10% of the mean aerodynamic chord. The trailing edge flap extends from the fuselage to 65% of<br />

the semi-span, the leading edge flap extends from the fuselage to 90% of the semi-span, and the<br />

aileron extends from the edge of the flap to 90% of the semi-span. No moveable surfaces were<br />

added to the last 10% of the semi-span so that radar obsorbing materials (RAM) could be added<br />

in the wing tip to minimize any returns from that edge.<br />

24


Figure 5.3 - Wing Planform<br />

Table 5.I - Wing Measurements<br />

Planform Area 1,500 ft 2 (139 m 2 )<br />

Span 54.8 ft (16.7m)<br />

Root Chord 46.8 ft (14.3 m)<br />

Tip Chord 8.0 ft (2.4 m)<br />

MAC 32.0 ft (9.8 m)<br />

y Location of MAC 10.5 ft (3.2 m)<br />

Aspect Ratio 2.0<br />

Leading Edge Sweep 40.0º<br />

Sweep at Quarter Chord 20.5º<br />

Sweep at Half Chord 4.7º<br />

Trailing Edge Sweep -30.0º<br />

Taper Ratio 0.17<br />

Leading Edge Flap Area 137 ft 2 (12.7 m 2 )<br />

Trailing Edge Flaperon Area 238 ft 2 (22.1 m 2 )<br />

Flapped Wing Area 935 ft 2 (86.9 m 2 )<br />

25


5.3 Wing Thickness<br />

The effect of wing thickness on the performance of the aircraft was studied so that the optimum<br />

thickness could be chosen. Initially a wing thickness of 3% of the chord was chosen based on<br />

existing supercruising aircraft. Increasing the root thickness of the wing was considered to<br />

reduce the weight of the wing. The effects of wing root thickness on wing weight, cross sectional<br />

area, and fuel consumption were studied. The weight of the wing was estimated using the<br />

method presented in Raymer, and the additional cross sectional area was calculated numerically.<br />

The resulting wing weights and cross sectional areas for wing root thicknesses from 3% to 6%<br />

are shown in Figure 5.4. The effect of the resulting weights and cross sectional areas on the fuel<br />

consumption during the mission were estimated using the same method used for the wing sizing.<br />

The results in Figure 5.5 show that the larger cross section of a thicker wing root adds more<br />

wave drag than the induced drag savings from the reduced wing weight. Based on this result, a<br />

constant wing thickness of 3% was chosen.<br />

8,500<br />

8,000<br />

t root = 3%<br />

Weight of Wing (lb)<br />

7,500<br />

7,000<br />

t root = 4%<br />

t root = 5%<br />

6,500<br />

t root = 6%<br />

6,000<br />

14 15 16 17 18 19 20 21 22<br />

Maximum Frontal Cross Sectional Area of Wing (ft 2 )<br />

Figure 5.4 - Effect of Root Chord Thickness on Wing Weight and Cross Sectional Area<br />

26


62,500<br />

Fuel Consumption over Mission (lb)<br />

62,000<br />

61,500<br />

61,000<br />

60,500<br />

60,000<br />

59,500<br />

59,000<br />

58,500<br />

58,000<br />

Fuel Onboard<br />

57,500<br />

3.0% 3.5% 4.0% 4.5% 5.0% 5.5% 6.0%<br />

Wing Root Thickness<br />

Figure 5.5 - Effect of Root Chord Thickness on Fuel Burn<br />

27


5.4 Airfoil<br />

The NACA 65A-003 airfoil section was chosen for the aircraft, because a symmetrical airfoil<br />

with the maximum thickness at 50% of the chord is optimum for supersonic flight. The airfoil<br />

ordinates given in Theory of Wing Sections for an NACA 65A-006 were scaled and interpolated<br />

using Lagrangian polynomials to define the geometry of the wing. The leading edge radius of<br />

the airfoil is 0.1% of the chord, which is approximately 3/8 inch at the mean aerodynamic chord<br />

and 1/10 inch at the tip. The airfoil sections at the mean aerodynamic chord and tip of the trailing<br />

edge flap are shown in Figure 5.6 and Figure 5.7 respectively. Because the chords of the flaps<br />

remain constant as the wing chord changes, each airfoil section has a different relative flap size.<br />

0.1<br />

0.05<br />

0<br />

-0.05<br />

-0.1<br />

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1<br />

Figure 5.6 - Airfoil Section at MAC<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

-0.05<br />

-0.1<br />

-0.15<br />

-0.2<br />

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1<br />

Figure 5.7 - Airfoil Section at Tip of Trailing Edge Flap<br />

5.5 Lift Curve<br />

To estimate the lift curve of the wing, first, the lift curve slope of the wing was estimated using<br />

standard subsonic theory, compressibility corrections, and linear supersonic theory. The<br />

resulting lift curve slopes are shown as a function of Mach number in Figure 5.8.<br />

28


5<br />

4.5<br />

4<br />

Lift Curve Slope (1/rad)<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

0 0.5 1 1.5 2 2.5 3<br />

Mach<br />

Figure 5.8 – Variation in Lift Curve Slope with Mach Number<br />

Next, the stall angle of the wing was estimated under subsonic conditions by calculating the lift<br />

distribution of the wing using LinAir. The section lift coefficient was calculated as a function of<br />

the span-wise location of the section for different wing angles of attack. The wing was assumed<br />

to stall when one of the section lift coefficients exceeded the maximum lift coefficient given in<br />

Theory of Wing Sections. The stall angle of attack of the wing was determined to approximately<br />

14º. Because the wing tip was shown to stall at a much lower angle of attack than the rest of the<br />

wing, adding a –3º angle of incidence to the wing tip was considered. The resulting twist<br />

extends the stall angle of attack to approximately 16º; however the twist decreased the lift<br />

coefficient at a given angle of attack and could impact RCS and supersonic aerodynamics.<br />

Ultimately, the non-twisted wing was chosen. The lift distributions of the wing with and without<br />

twist are shown in Figure 5.9.<br />

The effects of the trailing edge flap were estimated using the stall angle of attack and lift<br />

coefficient increments given in Nicolai. The effect of the leading edge flap was estimated by<br />

assuming that a 10º leading edge flap deflection would increase the stall angle of attack by<br />

approximately 10º, and the decrease in lift coefficient was estimated based on the change in<br />

effective angle of attack. The resulting subsonic lift curve at Mach 0.2 is shown in Figure 5.10.<br />

29


1<br />

Section Lift Coefficient<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

Max. Section Lift Coefficient<br />

0º Tip Incidence<br />

16º<br />

15º 16º<br />

15º<br />

14º<br />

13º<br />

14º<br />

12º<br />

13º<br />

12º<br />

- 3º Tip Incidence<br />

<strong>Cal</strong>culated Using LinAir<br />

0.2<br />

0.1<br />

0<br />

0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%<br />

Spanwise Distance (percent semi-span)<br />

Figure 5.9 - Lift Distribution of Wing with and without Twist<br />

2<br />

Lift Coefficient<br />

1.5<br />

1<br />

0.5<br />

0<br />

C L = 1.20<br />

C L = 1.15<br />

C L = 0.56<br />

C L α = 2.33 1/rad<br />

C L = 1.51<br />

30º TE Flap<br />

Deflection<br />

Clean<br />

LinAir<br />

10º LE Flap<br />

Deflection<br />

-0.5<br />

Tail Strike Angle (14º)<br />

-1<br />

-20 -15 -10 -5 0 5 10 15 20 25 30<br />

Angle-of-Attack (degrees)<br />

Figure 5.10 - Subsonic Wing Lift Curve<br />

30


5.6 Drag<br />

The drag of the aircraft was divided into four parts: parasite drag, wave drag, induced drag, and<br />

trim drag. The parasite drag was estimated using a component build method with form and<br />

interference factors. The wave drag was calculated using the formula presented in Brandt &<br />

Stiles. The wave drag efficiency factor was calculated from cross sectional area distributions<br />

using the de Kármán integral and the theoretical wave drag of a perfect Sears-Haack body. The<br />

cross sectional area distributions were measured at transonic and supersonic (Mach 1.6)<br />

conditions. The transonic case was measured by passing vertical planes through a solid model of<br />

the aircraft and measuring the intersecting area. The supersonic case was measured by passing<br />

Mach cones through the model, measuring the intersecting area, and projecting that area onto the<br />

vertical plane. For both cases, the engine capture area was subtracted from sections containing<br />

the inlet, engine, and nozzle. The resulting area distributions shown in Figure 5.11 and Figure<br />

5.12 match reasonably well with that of a perfect Sears-Haack body. Both distributions yield a<br />

wave drag efficiency factor of approximately 2.14 (based on 80 ft 2 (7.4 m 2 ) max. area and 100 ft<br />

(30.5 m) length).<br />

90<br />

80<br />

70<br />

Sears-Haack<br />

Wing<br />

Cross Sectional Area (ft 2 )<br />

60<br />

50<br />

40<br />

30<br />

Fuselage<br />

Vertical Tail<br />

Horizontal Tail<br />

20<br />

10<br />

0<br />

0 200 400 600 800 1,000 1,200<br />

Fuselage Station (inches aft datum)<br />

Figure 5.11 - Transonic Area Distribution<br />

31


90<br />

80<br />

Wing<br />

Cross Sectional Area (ft 2 )<br />

70<br />

60<br />

50<br />

40<br />

30<br />

Sears-Haack<br />

Fuselage<br />

Vertical Tail<br />

Horizontal Tail<br />

20<br />

10<br />

0<br />

0 200 400 600 800 1,000<br />

Fuselage Station (inches aft datum)<br />

Figure 5.12 - Supersonic Area Distribution (Mach 1.6)<br />

Induced drag was estimated using standard subsonic theory and the supersonic equation<br />

presented in Brandt & Stiles to calculate the induced drag term (k 1 ). Trim drag was calculated as<br />

induced drag generated by the horizontal tail at the lift coefficient required to trim the aircraft<br />

with a given static margin and moment coefficient. The resulting drag build-up for the aircraft at<br />

an altitude of 50,000 ft (15,240 m), Mach number of 1.6, maneuver weight of 94,735 lb (42,971<br />

kg), and 5% static margin is shown in Figure 5.13.<br />

32


0.06<br />

0.05<br />

Drag Coefficient<br />

0.04<br />

0.03<br />

0.02<br />

Induced Drag<br />

50,000 ft<br />

Maneuver Weight<br />

Trim Drag<br />

0.01<br />

Wave Drag<br />

Parasite Drag<br />

0<br />

0 0.5 1 1.5 2 2.5 3<br />

Mach<br />

Figure 5.13 - Drag Build-Up at 50,000 ft, Mach 1.6, Maneuver Weight, and 5% Static Margin<br />

33


6 Propulsion<br />

In developing the propulsion system for the Vendetta, the RFP specifications of supersonic cruise<br />

and stealth are the driving factors for the propulsions system. Due to the stealth criteria the fan<br />

blades of the engine must remain hidden which drives the engine placement inside the airplane.<br />

That combined with the supercruise criteria leads to four major aspects of the propulsion system:<br />

the Engine, Inlets, S-ducts and Nozzles.<br />

6.1 Engine Selection<br />

The RFP specifies that the airplane should perform the mission with adequate installed thrust. A<br />

Low-Bypass-Ratio Turbofan (LBR-TF) or a Turbojet (TJ) engine may be used to perform the<br />

mission. Equations are provided for creating an engine deck, they are as follows:<br />

LBR-TF:<br />

TJ:<br />

ρ<br />

Tdry<br />

= 0.9 TSL −dry<br />

(0.88 + 0.24 ABS( M − 0.6) )( )<br />

ρSL<br />

2 ρ 0.8<br />

Twet<br />

= TSL −wet<br />

(0.94 + 0.38 ABS( M − 0.4) )( )<br />

ρ<br />

1.4 0.8<br />

ρ<br />

Tdry<br />

= 0.9 TSL −dry<br />

(0.907 + 0.262 ABS( M − 0.5) )( )<br />

ρSL<br />

2 ρ 0.8<br />

Twet<br />

= TSL −wet<br />

(0.954 + 0.38 ABS( M − 0.4) )( )<br />

ρ<br />

SL<br />

1.5 0.8<br />

However an installed engine deck for a LBR-TF with an axisymmetric center body inlet and a<br />

mixed flow ejector nozzle was supplied with the RFP. Since it included physical dimensions and<br />

fuel flow values the RFP engine deck was used instead of the above equations. The RFP engine<br />

specifications are shown in Table 6.I.<br />

Table 6.I - Engine Specifications of RFP Supplied Engine<br />

Engine and Nozzle Length<br />

Propulsion System Length<br />

Fan Face Diameter<br />

Maximum Diameter<br />

Weight with Nozzle<br />

SL<br />

310 inches (787 cm)<br />

425 inches (1080 cm)<br />

50 inches (127 cm)<br />

65 inches (165 cm)<br />

7200 pounds (3266 kg)<br />

The engine supplied by the RFP includes fuel flow and thrust data for part power, idle power,<br />

and military power. All engine data supplied by the RFP is corrected to sea level and a Mach<br />

number of zero. Therefore, every value for thrust and fuel flow at each altitude and Mach<br />

34


number is given in corrected net propulsive force (NPF c ) and corrected fuel flow (WF c ). To find<br />

the actual thrust (NPF) and fuel flow (WF) the following equations were used:<br />

NPF = NPF ⋅ d<br />

c T<br />

0.6<br />

F<br />

=<br />

F<br />

⋅<br />

c T<br />

⋅<br />

T<br />

W W Q d<br />

2 3.5 P<br />

dT<br />

= (1 + .2 M ) ( )<br />

PSL<br />

2 T<br />

QT<br />

= 1+<br />

0.2 M ( ) T<br />

Once the data was uncorrected the military thrust was found. The RFP supplied equations that<br />

could be used to scale the engine based on a desired thrust. The scaling equations are as follows:<br />

NPF<br />

NewMeasurement = OldMeasurement( ) NPF<br />

base<br />

Axial length scaling exponent = 0.4<br />

Diameter scaling exponent = 0.5<br />

Weight scaling exponent = 1.0<br />

SL<br />

exponent<br />

The RFP engine produced a military thrust of 26,350 pounds (117,210 N) and had a cruise thrust<br />

specific fuel consumption (TSFC) of 1.19 1/hr (0.121 kg/h/N) for Mach 1.6 flow at 50,000 ft<br />

(15,240 m). TSFC is calculated using the following equation:<br />

TSFC =<br />

W F<br />

NPF<br />

The Vendetta would require two engines to perform the desired mission. The size, weight, and<br />

location of the engines have great effect on the size of the airplane. The larger the engines the<br />

wider the aft portion of the fuselage and the longer the airplane. For the size and weight of the<br />

RFP engine it produced merely too little thrust and burned too much fuel.<br />

Other engines were sought out and analyzed in an attempt to find a better performing engine that<br />

was smaller and lighter than that supplied. Through this research the Concorde’s Rolls-Royce<br />

Snecma Olympus engine was found to be comparable to the RFP engine. However, the engine<br />

was first manufactured and flown in the Concorde in the mid 60’s through mid 70’s. Table 6.II<br />

compares the RFP engine to that of the Snecma Olympus. As can be seen the Snecma Olympus<br />

is very close in size and weight to that of the RFP, however it produces even more thrust than<br />

that of the RFP. Also the weight of the Snecma Olympus includes that of an afterburner whereas<br />

the RFP engine is without an afterburner.<br />

35


Table 6.II - RFP Dimensions Compared to the Snecma Olympus<br />

RFP<br />

Snecma Olympus<br />

Fan Face Diameter 50 in (127 cm) 47.5 in (120.65 cm)<br />

Length 310 in (787 cm) 280 in (711 cm)<br />

Weight 7200 lbs (3266 kg) 7000 lbs (3175 kg)<br />

Max Dry Thrust 26,356 lbs (117,237 N) 31,350 lbs (139,452 N)<br />

Based on this data the RFP engine was assumed to be an older engine; a more efficient and<br />

modern engine would be needed for the design of the Vendetta. The RFP engine deck was used<br />

as a baseline for designing a newer better engine, as it was the only full engine deck available. It<br />

was determined that an F119 engine would be the initial design engine for the airplane. This<br />

engine is currently used in the F-22 and a derivation of the engine (the F135) is to be used in the<br />

F-35.<br />

All information regarding the F119 is classified except that it is in a 35,000 lbs (155,700 N)<br />

weight class. Several methods were utilized to narrow in on the thrust produced by the F119.<br />

Through the use of The Integrated High Performance Turbine Engine Technology (IHPTET)<br />

program F119 characteristics were determined. IHPTET which began in 1988 and will culminate<br />

in 2005 consists of a 3 phase plan, utilizing the most current advancements in industry. “IHPTET<br />

is producing revolutionary advancements in turbine engine technologies due to the synergistic<br />

effect of combining advanced material developments, innovative structural designs and<br />

improved aerothermodynamics ” The 3 phases of the program are shown in Table 6.III.<br />

Phase III (2005)<br />

Phase II (1997)<br />

Phase I (Completed)<br />

Table 6.III - IHPTET Goals<br />

+100% Thrust/Weight<br />

-40% Fuel Burn<br />

+60% Thrust/Weight<br />

-30% Fuel Burn<br />

+30% Thrust/Weight<br />

-20% Fuel Burn<br />

The Air Force Research Laboratory states that Phase I of the program has been completed and<br />

that the technology has been applied to existing engines including the F100, F110, F414, and the<br />

F119. Based on this data Phase I was applied to the RFP engine deck to yield an F119 engine.<br />

This was done because both the RFP engine and the F119 are low bypass turbofan engines. The<br />

20% decrease in fuel burn was applied and then the weight was decreased by 21% and the thrust<br />

increased by 2% to account for the 30% change in thrust to weight. The resulting thrust produced<br />

by the F119 is 27,000 pounds (120,101 N), has a cruise TSFC of 0.94 1/hr (0.096 kg/hr/N) and a<br />

weight of 6,575 lbs (2,575 kg).<br />

Vendetta’s design is to be frozen in 2010 and future advancements and technologies are to be<br />

taken into account for its development. Phase II of IHPTET was to be completed by 1997 yet has<br />

not been achieved. However, Pratt and Whitney has proven a 40% increase in thrust to weight.<br />

36


The weight of the F119 was decreased down to 5,270 lbs (2,384 kg) to account for the additional<br />

10% increase in thrust to weight. However there is no evidence of any other advancements with<br />

the IHPTET program, and advancements in turbofan engines were sought out. The Versatile<br />

Affordable Advanced Turbofan Engine (VAATE) is an industry projection to 2020. Even though<br />

it builds upon IHPTET it uses the F119 as a base engine for its future goals. Figure 6.1 illustrates<br />

the goals for turbofan engines thru 2020 and Phase I goals of a 25% decrease in TSFC and a 45%<br />

decrease in cost by 2010. It is likely that this program will face similar problems in achieving its<br />

goal by 2010, in which case a decrease of 13% in TSFC was taken and an estimated 25%<br />

decrease in cost over the F119. The 13% change in TSFC was achieved by increasing the thrust<br />

by 11% and decrease the fuel flow 2%. The new VAATE technology engine has a sea level<br />

thrust of 30,000 lbs (133,500 N) and a cruise TSFC of 0.82 1/hr (0.084 kg/hr/N), however once<br />

inlet and ducting losses are accounted for the cruise TSFC is 0.92. 1/hr (0.094 kg/hr/N).<br />

37


38<br />

Figure 6.1 - VAATE Goals


The data for the idle power of the engine was based off of the original RFP. The idle data has not<br />

been changed since there is no distance credit or fuel credit for the descent portions of the<br />

mission and that is the only time idle settings would be used for this mission. The military thrust<br />

TSFC of the engine at various altitudes can be seen in Figure 6.2 and Figure 6.3 respectively.<br />

The afterburner model was created by multiplying the military thrust by 1.65 and the military<br />

fuel flow by 2. The equations provided by the RFP for calculating thrust from afterburner were<br />

not utilized as the maximum thrust produced by those equations yielded the same amount of<br />

thrust as the F119 at military power and altitude.<br />

Thrust (lbs/eng)<br />

35,000<br />

30,000<br />

25,000<br />

20,000<br />

15,000<br />

10,000<br />

Sea Level<br />

Alt 5,000 ft<br />

Alt 10,000 ft<br />

Alt 20,000 ft<br />

Alt 25,000 ft<br />

Alt 30,000 ft<br />

Alt 36,089 ft<br />

Alt 43,000 ft<br />

Alt 50,000 ft<br />

Alt 55,000 ft<br />

Alt 60,000 ft<br />

Alt 70,000 ft<br />

5,000<br />

0<br />

0 0.5 1 1.5 2 2.5<br />

Mach Number<br />

Figure 6.2 - Thrust Curves for Altitudes from Sea Level to 70,000 ft (21,300 m)<br />

39


TSFC<br />

3.0<br />

2.8<br />

2.6<br />

2.4<br />

2.2<br />

2.0<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

1.0<br />

0 0.5 1 1.5 2 2.5 3<br />

Mach Number<br />

Sea Level<br />

Alt 5,000 ft<br />

Alt 10,000 ft<br />

Alt 20,000 ft<br />

Alt 25,000 ft<br />

Alt 30,000 ft<br />

Alt 36,089 ft<br />

Alt 43,000 ft<br />

Alt 50,000 ft<br />

Alt 55,000 ft<br />

Alt 60,000 ft<br />

Alt 70,000 ft<br />

Figure 6.3 - Military TSFC Curves for Altitudes from Sea Level to 70,000 ft (21,300 m)<br />

Once the engine deck had been established the engine dimensions were once again considered.<br />

The fan face diameter of the new engine was assumed to be that of the RFP. Low bypass<br />

turbofans typically have smaller fan face diameters however as the bypass ratio increases the fan<br />

face diameter would increase as well. Since future technology is being taken into account it is<br />

likely that the engine that would produce this thrust would have a larger bypass ratio but smaller<br />

core keeping the fan face diameter comparable to that of the RFP. The length of the engine was<br />

estimated based off of previous trends in engines. Engines used to plot this data include the<br />

F100, F101, F110, F404 JT3D, JT8D and TF30. This data can be seen plotted in Figure 6.4<br />

below. Logarithmic trendlines were fitted to the data and then extrapolated out past 30,000 lbs<br />

(133,500 N). For a thrust of 30,000 lbs (133,500 N) the engine would have a maximum diameter<br />

of 60 inches (152 cm) and a maximum length of 220 inches (559 cm).<br />

40


Diameter (in)<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

Length (in)<br />

0<br />

0<br />

12,000 17,000 22,000 27,000 32,000<br />

Thrust (lbF)<br />

Figure 6.4 - Engine Sizing Plot<br />

6.2 Inlets<br />

Sizing the inlet for supercruise flight at 1.6 Mach posed an interesting problem. A pitot inlet is<br />

good up until about 1.6 Mach and it is by far the cheapest inlet possible. However the<br />

performance of the inlet above Mach 1.6 is very poor. The pressure recovery of a two shock inlet<br />

(one oblique and one normal shock) and a three shock inlet were analyzed. The optimum<br />

deflection angle for Mach 1.6 flow was found for a two shock inlet by finding the stagnation<br />

pressure loss across the oblique and normal shock for different deflection angles. The results<br />

were graphed in Figure 6.5 and the resulting deflection angle for the greatest pressure recovery<br />

was found to be 10.75 degrees yielding a pressure recovery of 97.65%. Finding the optimum<br />

deflection angle for a three shock inlet is more involved therefore a rough estimate of a six<br />

degree deflection angle followed by another 6 degree deflection angle was used to compare<br />

against the two shock inlet. The difference in on design pressure recovery is about 1% however<br />

the larger the deflection angles become the better the pressure recovery will become. The<br />

pressure recovery comparison can be seen in Figure 6.6. The military specification for inlets is<br />

given below and is represented in the graph.<br />

Mil Spec MIL-E-5008B<br />

η<br />

rSpec<br />

⎧ 1 M ≤ 1<br />

⎩1<br />

0.075( 1) 1 5<br />

0<br />

= ⎨ −<br />

1.35<br />

M0 − < M0<br />

<<br />

41


Deflection Angle Analysis for Mach 1.6<br />

0.99<br />

0.98<br />

0.97<br />

Pressure Recovery<br />

0.96<br />

0.95<br />

0.94<br />

0.93<br />

0.92<br />

Deflection Angle = 10.75<br />

Pressure Recovery = 0.9765<br />

0.91<br />

0.9<br />

0.89<br />

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15<br />

Flow Deflection Angle<br />

Figure 6.5 - Optimum Deflection Angle for Mach 1.6 Flow<br />

Pressure Recovery vs. Inlet Shocks<br />

0.99<br />

Total Pressure Recovery<br />

0.97<br />

0.95<br />

0.93<br />

0.91<br />

0.89<br />

0.87<br />

Design Point<br />

Mil-E-5008B<br />

0.85<br />

1 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 1.9 2<br />

Mach Number<br />

2 Shock 3 Shock angles 6 and 6 Mil Spec Mach 1.6<br />

Figure 6.6 - Pressure Recovery for a Two Shock versus Three Shock Inlet<br />

42


Other traits were looked at before choosing a two shock compression inlet. Figure 6.7 illustrates<br />

the area ratio compared to Mach number for the two and three shock inlets. The jumps in the<br />

curve, illustrate the Mach numbers at which the engine will work the hardest because they are<br />

the locations at which the shocks occur. Both of which will occur before the cruise Mach number<br />

of 1.6 therefore occurring during the climb portion of the mission.<br />

Area Ratio vs. Mach<br />

Area Ratio<br />

1.4<br />

1.35<br />

1.3<br />

1.25<br />

1.2<br />

1.15<br />

1.1<br />

1.05<br />

1<br />

0.95<br />

0.9<br />

0.8 1 1.2 1.4 1.6 1.8 2<br />

Mach Number<br />

2 Shock 3 Shock 6 deg and 6 deg<br />

Figure 6.7 - Inlet Area Ratio<br />

Figure 6.8 shows a rough cost relationship between different type inlets. As can be seen there is a<br />

lower cost associated with a two shock inlet. There are drawbacks to having more shocks, in that<br />

they drive the inlet to be larger, longer, and send multiple radar returns. The above traits do not<br />

show enough of a benefit to go with a three shock inlet therefore a two shock inlet was chosen.<br />

43


Figure 6.8 - Cost Association with Inlet Shocks<br />

Figure 6.9 below shows the off design inlet area ratio that is required for the Vendetta. The<br />

equations used to find the data are shown below. The actual capture inlet area is depicted by A 1 ,<br />

with the area at the shock being A s , and the actual flow area being captured by the inlet as A 0i .<br />

As the Vendetta increases in speed the engine requires a greater amount of inlet area for a<br />

constant mass flow rate.<br />

Mass Flow Ratio:<br />

A A A<br />

=<br />

A A A<br />

0i 0i s<br />

1 s 1<br />

Area Ratio:<br />

A<br />

A<br />

0i s s<br />

s<br />

ρ V<br />

=<br />

ρ V<br />

0 0<br />

44


Mass Flow Performance<br />

Inlet Area Ratio<br />

1.3<br />

1.2<br />

1.1<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

Design Point<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

0 0.5 1 1.5 2 2.5 3<br />

Mach Number<br />

Figure 6.9 - Off Design Area Required for Engine Mass Flow<br />

The inlet capture area was found by first estimating the mass flow rate required by the engine at<br />

the design point. The mass flow of the engine could be estimated using the following equation.<br />

Mass Flow Estimation:<br />

m<br />

e<br />

= 26( FrontFaceDiameter)<br />

2<br />

The front face diameter of 4 ft (122 cm) was used; this yielded a mass flow rate of approximately<br />

405 slugs/sec (5910 kg/sec). Now using the mass flow equation shown below, the area of the<br />

inlet could be found for the design mission.<br />

Mass Flow Equation:<br />

m<br />

= ρ AV<br />

Once this was done the mass flow equation was used to calculate the area at different altitudes<br />

based on conservation of energy. For the desired design point of 1.6 Mach and an altitude of<br />

50,000 ft (15,240 m) this was found to be slightly larger than 5 ft 2 (4645 cm 2 ); however at<br />

55,000 ft (16,764 m)t it was found to be about 6 ft 2 (5.6 m 2 ). Since different parts of the mission<br />

take place at several different altitudes above 50,000 ft (15,240 m), the inlet area was sized to 6.5<br />

ft 2 (6 m 2 ). By sizing the engine to 6 ft 2 (5574 cm 2 ) air could be bypassed from the inlet to cool<br />

the fuel.<br />

45


The inlet has a boundary layer diverter for high speeds and auxiliary doors for low speed flight,<br />

since the required inlet area at take off will be twice what it is at cruise. The final inlet sizing for<br />

Mach 1.6 is:<br />

• Inlet capture area = 6.5 ft 2 (6 m 2 )<br />

• Inlet compression angle 10.75 degrees<br />

• Inlet Pressure Recovery is 97.6%<br />

• Speed after Normal Shock, M=0.82<br />

The inlet is located on a boundary layer diverter on the lower side of the wing. This keeps any<br />

vortices produced off of the wing or side of fuselage from being ingested by the inlet, as well as<br />

aid in inlet capture at high angles of attack.<br />

6.3 S-Duct<br />

S-ducts were used to move<br />

the flow from the inlets to<br />

the engine faces so that the<br />

compressor face of the<br />

engine could not be seen.<br />

Stealth is a requirement for<br />

the mission and the<br />

Figure 6.10 - Vendetta S-Duct<br />

compressor face is a large<br />

contributor to radar return. The s-duct goes from a minimum area just after the inlet to a<br />

maximum area at the compressor face as can be seen in Figure 6.10. The s-duct shape<br />

progressively goes from a square at the inlet to an oval<br />

and then a circle at the engine face.<br />

3 o<br />

4 o<br />

The portion of the s-duct closest to the fan face is used<br />

to straighten and slow the flow before it hits the<br />

compressor. This is done by having that portion of the<br />

duct be fairly long and gradually diffuse up to the<br />

compressor face through an upper deflection angle of 3<br />

degrees and a lower angle of 4 degrees as shown in<br />

Figure 6.11.<br />

Figure 6.11 - Diffuser Angle to the Engine<br />

Face<br />

The s-duct is 28 feet (853 cm) in length with an average<br />

duct height of 2.7 feet (82 cm). The efficiency of the s-<br />

duct is found by calculating the average diameter of the<br />

duct and dividing that by the length of the duct. The<br />

figure presented in Appendix B was used to calculate<br />

the efficiency of the duct. This yielded a length over<br />

diameter of just over 10 and an engine area to inlet area of 2 which yielded a duct efficiency of<br />

91.5%.<br />

46


6.4 Nozzle<br />

The nozzle of the Vendetta will have an afterburner and thrust<br />

reversers incorporated. It is a converging-diverging nozzle which<br />

allows for backpressure control at off design Mach numbers<br />

when the afterburners are on. That leads into a low signature<br />

axisymmetric advanced nozzle similar to that seen in Figure<br />

6.12. The advanced nozzle is being used because it has<br />

comparable signature to that of a 2-D nozzle however it weights<br />

50% less, costs 60% less and requires 300 fewer parts.<br />

Figure 6.13 - Clam Style Thrust Reversers<br />

The nozzle will have<br />

thrust reversing<br />

capabilities to enable<br />

the aircraft to land on<br />

an icy runway and<br />

Figure 6.12 - Low-Signature<br />

Axisymmetric Advanced Nozzle<br />

stop within the required 8,000 ft (2483 m) as<br />

specified by the RFP. Clam style thrust reversers,<br />

similar to those seen in Figure 6.13 will be utilized to<br />

reverse 25% of the thrust. Thrust vectoring is will not<br />

be incorporated as the Vendetta is not required to<br />

maneuver like a fighter.<br />

47


7 Materials and Structure<br />

The overall layout of the Vendetta’s structural layout is shown in Figure 7.1. The wing structure<br />

is similar to that of an F-15 and the material selection is similar an F-22. The main load path is<br />

in the form of a central keel that runs from between the nozzles and engines to the nose gear<br />

attachment point. The weapons bay splits the keel in the center of the aircraft. The load is<br />

shifted from the keel to the aft weapons bay wall and back into the keel at the forward end of the<br />

weapons bay. A close up of the weapons bay is also shown in Figure 7.1.<br />

Main Wing Spars<br />

Weapons Bay<br />

Stiffeners<br />

Figure 7.1 - Structure Buildup for Vendetta<br />

48


The layout of Vendetta’s inlets and landing gear allow for a continuous structural member, in the<br />

form of a bulkhead, to carry the aerodynamic loads from each wing directly to the central keel.<br />

This ideology breaks down as the bulkheads move away from the main wing load paths. The<br />

weapons bay splits the forward wing attachment bulkheads. This occurs only well in front of the<br />

aerodynamic center of the wing. Just forward of the aerodynamic center is the main forward<br />

load path for the wing. The aft load paths are a ring structure around the engines and inlets. The<br />

Important thing to note is that where the primary loads are being distributed, between 25 to 50<br />

percent of the mean aerodynamic chord, the bulkheads are continuous.<br />

It is also important to note that the landing gear attaches to a bulkhead just forward of the aft<br />

closure to the weapons bay. This is important because it locates the airborne and ground laden<br />

load paths on top of each other. This allows for some redundancy in the structure and allows for<br />

a lighter aircraft. Another redundant feature is the aft main load path. This bulkhead acts as the<br />

main forward engine attachment point. Again this allows for a minimum of large structural<br />

bulkheads and thus creates a lighter aircraft. The wing attachment points are shown in Figure<br />

7.2.<br />

The empennage structure follows the same methodology as the wing attachment structure. The<br />

vertical tails attach to the aft primary carry through of the wing. The aft vertical attach point is<br />

the same as the primary load path for the horizontal tails. The horizontal tail is an area of<br />

concern for the Vendetta. The horizontal surfaces are capable of producing tremendous forces on<br />

the aircraft. The horizontal surfaces are full flying and thus must attach at a single point in the<br />

aircrafts structure. This bulkhead is a ring carry through type that distributes the load from the<br />

pivot point to the central keel. Two secondary bulkheads back up this main bulkhead. The<br />

empennage structure is shown in Figure 7.3.<br />

Aft Primary<br />

Bulkhead & Main<br />

Engine Attachment<br />

Forward Secondary<br />

Bulkheads<br />

Forward Primary<br />

Bulkhead<br />

Aft Secondary<br />

Bulkheads<br />

Main Gear<br />

Attachment<br />

Figure 7.2 - Wing Attachment Detail<br />

49


Vertical Attachment<br />

Points<br />

Horizontal Pivot<br />

10” Diameter Shaft<br />

Horizontal Structural Load<br />

Paths<br />

Figure 7.3 - Empennage Structural Layout<br />

The structure of Vendetta was created with the RFP load requirements in mind. A V-N Diagram<br />

shown in Figure 7.4 was created using the required maximum and minimum g’ limits, and<br />

knowing the maximum dynamic pressure the aircraft should withstand. This diagram shows the<br />

load envelope the aircraft can operate in. The diagram also shows the standard gust lines for 1-g’<br />

flight.<br />

8<br />

Max g Limit<br />

6<br />

Max q<br />

4<br />

Max Lift<br />

Gust Lines<br />

60 ft/sec<br />

g's<br />

2<br />

0<br />

0 ft/sec<br />

-2<br />

-4<br />

-60 ft/sec<br />

Min g Limit<br />

0 500 300 1000 600 1500 900 2000<br />

1200<br />

Equivlent Knots Equivalent Velocity (ft/sec) Airspeed<br />

Figure 7.4 - V-n Diagram for Vendetta<br />

50


The materials selection for Vendetta was difficult. Vendetta takes advantage of the benefits of<br />

modern composites while relying on the proven durability of more conventional metals. The<br />

materials selection for different components is shown in Table 7.I.<br />

Table 7.I - Materials Selection<br />

Forward fuselage: • Skins and Chine - composite<br />

• Bulkheads / Frames - resin transfer, molded composite, and<br />

aluminum<br />

• Fuel Tank Frame and Walls - resin transfer, molded composite<br />

• Side Array Doors and Avionics - formed thermoplastic<br />

Mid-fuselage: • Skins - composite and titanium<br />

• Bulkheads and Frames - titanium, aluminum, and composite<br />

• Fuel Doors - composite<br />

• Weapons Bay Doors:<br />

o Skins - thermoplastic<br />

Hat Stiffeners - resin transfer, molded composite<br />

o<br />

Aft fuselage • Forward Boom - welded titanium<br />

• Bulkheads and Frames - titanium<br />

• Keel Web - composite / HC Cocure<br />

• Upper Skins - titanium and composite / HC Cocure<br />

Empennage: • Skin and Closeouts - composite<br />

• Core - aluminum<br />

• Spars and Ribs - resin transfer, molded composite<br />

Wings:<br />

• Pivot Shaft - tow-placed composite<br />

The materials used in the construction of the main portion of the wing<br />

are titanium alloy (42%), composites (35%), aluminum alloy (24%),<br />

steel alloy, and some other materials. The following materials are used<br />

in the construction of the wing:<br />

• Skins - composite<br />

• Side of body fitting - titanium HIP cast<br />

• Spars<br />

o front - titanium<br />

o intermediate - resin transfer, molded composite and<br />

titanium<br />

o rear - composite and titanium<br />

Miscellaneous: • Duct Skins - composite<br />

• Landing Gear – Airmet 100 steel alloy<br />

51


8 Landing Gear<br />

Landing Gear design for the Vendetta has eight significant design drivers.<br />

1) Tire selection due to the high 150 knot takeoff and landing speed (277.8 km/hr)<br />

2) 120,000 lb gross weight (54,300 kg)<br />

3) Ease of loading and reloading weapons<br />

4) Tail Strike Angle<br />

5) Ground Handling Characteristics<br />

6) Structural Location<br />

7) Minimal Internal Volume Usage<br />

8) Low Weight<br />

Suitable structural attachment points dictated the main<br />

gear be positioned near the subsonic center of pressure on<br />

the main wing (near the main spar) shown in Figure 8.1.<br />

This placement, as well as limited internal volume, good<br />

ground handling characteristics, minimal frontal area, and<br />

ease of unloading and loading weapons led to the<br />

adoption of a tricycle landing gear configuration. The<br />

main gear configuration was then approximated as a 737<br />

type main gear, (near our GTOW) for volume purposes.<br />

Initial sizing began with tire selection. Following<br />

methodology outlined in Raymer the main gear of the<br />

Vendetta should carry 88% of the GTOW and the nose gear<br />

should carry 12%. Starting with a database of tires and<br />

wheels developed by the Aerospace and Ocean Engineering<br />

Department at Virginia <strong>Poly</strong>technic Institute and State<br />

Figure 8.1 - Main Gear Structural<br />

Attachment Point<br />

University and a listing of tires in Raymer’s Aircraft Design a Conceptual Approach the initial<br />

listing was narrowed to the choice of 36in x 11in (91.4cm x 27.9cm) tires for the main gear and<br />

24in x 7.7in (61cm x 19.6cm) for the nose gear. The tires selected allowed a 1.5 factor of safety<br />

over the dynamic landing load of the aircraft.<br />

Now knowing the approximate volume of the 737 landing gear configuration with usable tires a<br />

rough solid model of the fuselage and internal components was produced to determine the exact<br />

gear location. The main gear was then designed to fold into the allotted space. The initial design<br />

considered the smallest internal volume as well as smallest frontal area for a given load, as<br />

shown in Figure 8.2. After analyzing both the internal position the gear would have to fold into,<br />

behind and under the main inlet ducts, the tandem configuration was chosen.<br />

The next challenge presented was obtaining the necessary gear height for easy loading and<br />

unloading as well as a tip back angle which did not exceed the tail strike angle, and having that<br />

gear fit into the limited internal volume available. The gear retraction scheme adopted produced<br />

52


a landing gear retraction scheme similar to an<br />

XB-70 Valkarie as shown in Figure 8.3. The<br />

complexity was necessary due to overall<br />

configuration drive of low supersonic maximum<br />

cross sectional area.<br />

Due to the Sears-Haack ideal of supersonic area<br />

distribution the nose gear could be easily placed<br />

in the forward fuselage. No need for complex<br />

folding arrangements led to the selection of a<br />

standard side-by-side tire configuration. The<br />

complete retraction schemes and nose wheel<br />

configuration can be seen in Figure 8.4. The<br />

completed landing gear configuration can be<br />

seen in Figure 8.5.<br />

The braking system for both the nose gear and<br />

main gear configuration would use a standard<br />

rotor disk braking mechanism. The heat-sink<br />

Figure 8.2 - Landing Gear Configuration Trade<br />

Study<br />

Figure 8.3 - Main Gear Retraction Sequence<br />

Figure 8.4 - Nose Gear and Main Gear Retraction Schemes<br />

will be made of carbon rather than steel because of the fact that carbon is superior to steel as it<br />

has a higher thermal conductivity and temperature limit, and the thermal expansion of carbon is<br />

lower. The only drawback to carbon brakes is the lower density of carbon compared to that of<br />

53


steel. This means that more<br />

braking material is required for<br />

a carbon braking system than<br />

what would be required for a<br />

steel system. The superceding<br />

benefit is that carbon offers a<br />

higher service life and has<br />

lower<br />

maintenance<br />

requirements than steel brakes.<br />

The sizing of the shock<br />

absorption system was<br />

designed around a hydraulic<br />

fluid pressure limit of 1,500 psi<br />

(10.3 GPa). The maximum<br />

load acting on each strut was<br />

then calculated and the<br />

corresponding piston area<br />

required to support this load<br />

was then calculated to be<br />

approximately 7 inches<br />

(17.8cm).<br />

Figure 8.5 - Completed Landing Gear<br />

The Vendetta’s landing gear, as seen<br />

in Figure 8.6 allows for drive up<br />

loading utilizing a MJ-1 or MHU-83<br />

lift truck. Landing gear sizing took<br />

account maximum lift truck reach to<br />

place weapons on the MPRL within<br />

the Vendetta’s main weapons bay.<br />

Figure 8.6 - Vendetta with MJ-1 Lift Truck and 2000lb JDAM<br />

54


9 Weight & Balance<br />

Weight and balance is one area often overlooked in aircraft design. The weights engineer<br />

develops a buildup that heavily influences aircraft performance engineer in terms of wing sizing,<br />

propulsion system selection, and in predicting aircraft performance. The stability and control<br />

engineer must rely on the weights engineer in order to size control surfaces, to develop flight<br />

control systems, and to develop stability derivatives. The configurator must work carefully with<br />

the weights engineer in order to develop a feasible inboard layout. In all aspects, the weights<br />

engineer plays a major role coordinating the design of any aircraft.<br />

After having sized the aircraft using the weight fractions technique, and after having developed<br />

an initial configuration, the next step is to develop a more accurate, class II weight buildup of the<br />

aircraft. The class II method used in the design of the Vendetta was developed from those<br />

methods found in the Nicolai, Raymer, and Roskam texts in order to obtain a collaborative and<br />

unbiased perspective. These methods involved defining several physical and geometric<br />

parameters of the aircraft. These parameters were inputs into a series of equations developed<br />

from historical weight trends. The weight estimations for various components as well as the<br />

level of agreement between authors are shown below in Table 9.I.<br />

Table 9.I - Initial Component Weight Buildup<br />

Roskam Nicolai Raymer Average Roskam Nicolai Raymer<br />

Component<br />

Weight<br />

(lbs)<br />

Weight<br />

(lbs)<br />

Weight<br />

(lbs)<br />

Weight<br />

(lbs)<br />

Accuracy<br />

(%)<br />

Accuracy<br />

(%)<br />

Accuracy<br />

(%)<br />

Structures<br />

Wing Group 9,687 11,466 7,870 9,674 0% -31% 26%<br />

Horizontal Tail 1,135 1,694 958 1,262 14% -62% 32%<br />

Vertical Tail 801 1,538 1,497 1,279 47% -34% -28%<br />

Fuselage 10,681 16,031 10,398 12,370 19% -52% 22%<br />

Main Landing Gear 2,742 2,969 1,156 2,289 -33% -52% 60%<br />

Nose Landing Gear 387 405 408 400 5% -2% -3%<br />

Propulsions 10,636 10,878 11,199 11,209 4% 0% -4%<br />

Systems 18,649 14,506 14,350 20,574 -29% 12% 13%<br />

Payload 9,500 9,500 9,500 9,500 0% 0% 0%<br />

Fuel 58,974 58,974 58,974 58,974 0% 0% 0%<br />

TOGW 123,653 128,435 128,435 127,531 4% -2% -2%<br />

The detailed weight buildup of the structures, control surfaces, systems, payload, and fuel groups<br />

has been compacted in order to save space and can be viewed in its entirety in Foldout 1. The<br />

table indicates that all three authors tend to disagree to some extent in their weight estimates of<br />

certain components, and for other components, one author may have no way of estimating that<br />

components weight at all. The most accurate way to develop the most detailed component weight<br />

buildup was to consider all three methods and take the average shared between them. An<br />

author’s estimation was discarded if it did not agree to within ±30% of the average of the other<br />

authors’ estimations. Once the estimations were in agreement to within ±30%, they were<br />

averaged in order to develop a weight buildup for the entire aircraft. The class II weight buildup<br />

for the Vendetta after the downgrading process is shown below in Table 9.II.<br />

55


Table 9.II - Final Component Weight Buildup<br />

Roskam Nicolai Raymer Average<br />

Component<br />

Weight<br />

(lbs)<br />

Weight<br />

(lbs)<br />

Weight<br />

(lbs)<br />

Weight<br />

(lbs)<br />

Structures<br />

Wing Group 9,687 XXXXXX 7,870 8,779<br />

Horizontal Tail 1,135 1,694 958 1,262<br />

Vertical Tail 801 1,538 1,497 1,279<br />

Fuselage 10,681 XXXXXX 10,398 10,540<br />

Main Landing Gear 2,742 2,969 1,156 2,289<br />

Nose Landing Gear 387 405 408 400<br />

Propulsions 10,636 10,878 11,199 11,209<br />

Systems 18,649 14,506 14,350 20,574<br />

Payload 9,500 9,500 9,500 9,500<br />

Fuel 58,974 58,974 58,974 58,974<br />

TOGW 124,800<br />

Inertias were calculated using guidelines outlined by the Society of Allied Weight Engineers<br />

(<strong>SAWE</strong>). Each components mass and location in reference to the aircraft center of gravity was<br />

used to calculate that components inertia. The sums of these inertias were then used in<br />

calculating the total moments of inertia about the<br />

Vendetta’s principal axes shown in Figure 9.1. In order to<br />

determine whether or not these values were accurate, the<br />

moments of inertia were transformed into non-dimensional<br />

radii of gyration coefficients. These coefficients were then<br />

compared to typical values for a jet bomber provided by<br />

<strong>SAWE</strong>. These comparisons are shown below in Table<br />

9.III.<br />

Table 9.III - Inertia Estimation<br />

Figure 9.1- Principle Axes<br />

Inertias (slug*ft 2 )<br />

Ix Iy Iz<br />

Vendetta 35,248 807,819 838,575<br />

Nondimensional Radii of Gyration<br />

Rx Ry Rz<br />

Vendetta 0.11 0.29 0.38<br />

<strong>SAWE</strong> 0.31 0.33 0.47<br />

Accuracy 63% 13% 18%<br />

The table indicates that the inertias are well<br />

within the typical values for a jet bomber<br />

except about the roll axis. This is because the<br />

Vendetta is similar to a typical jet bomber in<br />

length; however, it has a much shorter<br />

wingspan. This would constitute a smaller<br />

moment of inertia about the roll axis.<br />

After having developed an initial configuration and a more detailed class II weight buildup, the<br />

next step was to balance the aircraft. This was done for two types of payload, the first being<br />

fixed equipment and the second being variable payload. The variable payload aboard the<br />

Vendetta consists of both fuel and weapons because the weight and center of gravity of these<br />

56


items varies throughout the mission. The fixed equipment aboard the Vendetta is considered in<br />

this case to be everything other than the variable payload.<br />

Vendetta was first balanced it with the fixed equipment and then with the additional variable<br />

payload. This was done by first allowing the configurator to develop an inboard configuration.<br />

The weights engineer then calculated the center of gravity location resulting from this inboard<br />

arrangement. This process was iterative in that the weights engineer and configurator had to<br />

continuously modify the inboard arrangement until the center of gravity location was at the<br />

desired location.<br />

In order to minimize the trim drag on the aircraft, it was opted that the aircraft’s center of gravity<br />

location stay as close to the aerodynamic center as possible. This was a difficult task because of<br />

the dramatic shift, 12% MAC, in the location of the aerodynamic center when transitioning from<br />

subsonic to supersonic flight conditions. A trim tank was considered in order to allow the center<br />

of gravity to follow the aerodynamic center during this dramatic shift in order to maintain a<br />

neutrally stable condition at both subsonic and supersonic flight conditions; however, this idea<br />

was discarded because the trim tank would only require additional fuel volume in an already<br />

congested aircraft. Without a trim tank, in order to minimize drag by keeping the center of<br />

gravity as close as possible to the aerodynamic center the aircraft would have to fly with an<br />

unstable static margin, subsonically, and with a stable static margin, supersonically.<br />

The current arrangement of the fixed payload is such that it provides for a 5% unstable static<br />

margin at subsonic flight conditions. A center of gravity monitor makes use of fuel burn control<br />

in order to keep the aircraft as close as possible to the neutrally stable flight condition.<br />

Furthermore, with full fuel tanks, full weapons load, and subsonic flight conditions, i.e. takeoff,<br />

the aircraft is balanced such that it provides for a 5% unstable static margin. With the empty<br />

weight and takeoff gross weight balanced to provide a 5% unstable static margin, and with an<br />

aerodynamic shift of 12%, the aircraft immediately goes to a 7% stable static margin upon<br />

transitioning to supersonic flight. The center of gravity monitor then controls the fuel burn in<br />

such a way that the center of gravity follows the aerodynamic center and the Vendetta maintains<br />

neutral stability.<br />

Because the fuel load is constantly changing throughout the mission, balancing the fuel load<br />

throughout the mission can be a challenging task. Furthermore, the deployment of various<br />

weapons at any point during the mission makes this balancing process even more difficult.<br />

Because of the complexity involved in developing a center of gravity monitor, a computer code<br />

was developed in order to simulate the center of gravity monitor. The first step in developing<br />

this code was to obtain the best solution to balance the fuel payload throughout the mission. The<br />

code required four inputs including; the weight and location of the fixed equipment, the location<br />

and weight of the fuel at any given time, the amount of fuel burned at intervals throughout the<br />

mission profile, and the desired center of gravity location at that interval. With these inputs, the<br />

code can then determine which tank to burn fuel from in order to obtain the center of gravity<br />

location closest to that corresponding to the desired static margin. The code then outputs the<br />

center of gravity location and the remaining fuel payload. This is done at 10-second intervals<br />

57


throughout the 5-hour mission. Using this data, the center of gravity path can then be plotted as<br />

it tracks that path corresponding to the desired static margin.<br />

The next step was to balance the weapons payload. Because the weapons payload was placed in<br />

a rotary launcher, the center of gravity of the payload was concentrated in one location. If it had<br />

been placed in a more conventional arrangement spread across the belly of the aircraft, the center<br />

of gravity of the weapons would have also been spread across the belly of the aircraft. By<br />

concentrating the center of gravity of the weapons payload in one location and placing the<br />

weapons payload on top of the aircraft’s empty weight center of gravity location, deployment of<br />

the weapons payload did not generate any problems in balancing the aircraft or in disturbing the<br />

static margin. The center of gravity is shown tracking along the path of the desired static margin<br />

by means of fuel monitoring and pumping in Figure 9.2.<br />

Figure 9.2 - Center of Gravity Excursion<br />

The figure indicates that the center of gravity location at takeoff gross weight is slightly aft of<br />

the neutral point; however, the center of gravity tracks the desired static margin shortly after<br />

the aircraft has transitioned to supercruise. Notice the path of the aerodynamic center as it<br />

shifts during the transition from subsonic to supersonic flight. It is clear that the aircraft flies<br />

supersonically shortly after takeoff, or when the aircraft’s gross weight is just below takeoff<br />

gross weight. Furthermore, near the zero fuel weight, the aircraft flies subsonic for the<br />

remainder of the flight. The figure also indicates that with the current fuel tank arrangement,<br />

the desired static margin cannot be tracked during the final portion of the supercruise because<br />

there is not enough fuel available to properly trim the aircraft. At this point, the center of<br />

58


gravity is influenced by only the fixed weight of the aircraft and again the aircraft remains at<br />

a 5% unstable static margin during landing. This plot indicates that the center of gravity<br />

monitor works together with the control system in order to minimize trim drag while at the<br />

same time maintaining the aircraft’s controllability.<br />

59


10 Stability and Control<br />

To initially size the horizontal tail, tail volume coefficients from historical aircraft were<br />

analyzed. This was done in an attempt to determine the rough size of the horizontal and vertical<br />

tail surfaces prior to addressing stability and control issues. The tail volume coefficients are<br />

unitless parameters defined by geometric values relating the size of the empennage surface to the<br />

aircraft. The horizontal and vertical tail volume coefficients are defined in the following<br />

equations.<br />

V<br />

H<br />

=<br />

SHTL<br />

c S<br />

W<br />

HT<br />

W<br />

V<br />

V<br />

=<br />

SVT<br />

L<br />

b S<br />

W<br />

VT<br />

W<br />

Because the demands for most supersonic cruising aircraft are considered similar to a certain<br />

extent, the historical values of tail volume coefficients are used to back out the planform areas<br />

for the horizontal and vertical surfaces. Similar aircraft and their tail volume coefficients are<br />

presented in Table 10.I.<br />

AIRCRAFT<br />

Table 10.I - Historical Aircraft Tail Volume Coefficients<br />

TAIL VOLUME COEFFICIENTS<br />

V H<br />

V V<br />

Boeing SST (2707-300) 0.36 0.049<br />

Concorde n/a 0.080<br />

GD F-111A 1.28 0.064<br />

Rockwell B1B 0.80 0.039<br />

TU-22M 1.11 0.087<br />

TU-144 n/a 0.081<br />

AVERAGE 0.58 0.067<br />

Using the average tail volume coefficient for these similar aircraft yielded a horizontal stabilizer<br />

area of 386 ft 2 . This is rather large and may be attributed to the fact that these vehicles require<br />

large robustness in CG travel without the use of a flight control augmentation system (CAS).<br />

Likewise, the vertical tail would require 196 ft 2 of area. This number is driven slightly larger due<br />

to the fact that some of the larger historical tail volumes are inflated because these aircraft’s<br />

verticals are mounted on booms which extend aft. These booms allow for greater moment arms<br />

and make the vertical more effective.<br />

The effects of horizontal tail area on longitudinal static stability were looked at in an attempt to<br />

determine what the driving factors for horizontal tail area are. A Roskam class II method was<br />

60


used to see how the increased weight of a bigger horizontal affects the longitudinal static margin.<br />

This “X-plot”, as it is commonly known, is shown as Figure 10.1.<br />

Figure 10.1 - Longitudinal X-Plot at Mach 0.3<br />

It is notable that only about 110 ft 2 of horizontal area is required to keep the Vendetta neutrally<br />

stable at Mach 0.3. It can be seen that as the tail grows, the CG of the entire configuration shifts<br />

aft. This also shifts the effective neutral point (center of pressure) of the aircraft aft at a faster<br />

rate than the CG shifts aft. A horizontal that is bigger than 100 ft 2 yields a stable aircraft but will<br />

pay the price in trim drag if the aircraft is too stable.<br />

A stable static margin is necessary in flight without the use of a digital flight control margin. The<br />

RFP mandates this as well as adherence to MIL-8785C, the military specification for handling<br />

qualities of aircraft. A statically unstable aircraft would have a tendency to pitch up in a static<br />

level condition. The purpose of the horizontal tail is to apply a force which counteracts this<br />

offending moment. This comes at the price of trim drag, however. As the elevator is deflected,<br />

drag is created and this hurts the overall aircraft performance in cruise. It is because of this that a<br />

neutrally stable or marginally stable (1-3%) aircraft is desired in cruise.<br />

To complicate matters, it can be seen from Figure 6.1 that areas above 110 ft 2 are required for a<br />

Mach number of 0.3. The aerodynamic center (center of pressure) on the wing and most surfaces<br />

propagates aft as the Mach number passes the transonic regime. This shift effectively leaves the<br />

61


difference in neutral point and center of gravity greater. This difference means the aircraft is<br />

actually more stable in a supersonic cruise. The fact that the center of gravity is so far forward in<br />

relation to the neutral point causes the aircraft to pitch down. More trim is required which causes<br />

drag. This phenomenon is known as Mach tuck.<br />

It is because of this that the weight and balance of the aircraft must be closely in synch with the<br />

control system. Trim drag will be minimized and controllability will be enhanced with<br />

completely integrated systems.<br />

Canting the horizontals in a v-tail configuration was investigated in an attempt to shape the<br />

empennage in a stealthy manner. The effective area of the vertical and horizontal are functions of<br />

the square of the cosine of the cant angle. These effects are reflected in Figure 10.2.<br />

Figure 10.2 - Horizontal Area Required for Static Stability with Cant Angle<br />

It can be seen from the plot that as the cant angle increases the total planform area of the<br />

horizontal must increase to maintain the nominally desired static stability of 5%. Five percent<br />

was chosen because at this stage in the sizing it was uncertain what the dynamic characteristics<br />

of the aircraft would be. Attempting to maintain a minimally statically stable aircraft would ease<br />

the job of control system design if future needs warrant one. Angles up to 30° were looked at<br />

because it would be unwise from an RCS point of view to approach a 90° angle created by larger<br />

cants near 45°. Beyond 45° the trend would be the same; however the horizontal would drive the<br />

area instead of the vertical.<br />

This plot shows that only 118 ft 2 of horizontal area is required to maintain the desired static<br />

margin. This is far off from the historical class I method and by initial inspection appears small.<br />

62


This leads us to believe that the area required maintaining static stability is not the driving factor<br />

in the size of the horizontal. Control power required to rotate the aircraft, dynamic<br />

considerations, and high angle of attack recovery will most likely drive this size.<br />

A similar study was conducted on the vertical stabilizer to see what area would required for<br />

varying cant angles to maintain 0.001 (1/degree) lateral weathercock stability. This is illustrated<br />

in Figure 10.3.<br />

Figure 10.3 - Vertical Area Required for Static Stability with Cant Angle<br />

From Figure 6.3 it can be seen that at 30°, 165 ft 2 of vertical area is required to maintain 0.001<br />

(1/degree) of lateral weathercock stability. Although the 30° cant angle on the verticals was<br />

initially selected to match the bottom fuselage facets for RCS considerations, lowering that angle<br />

to 20° would allow other advantages. Shallower cant angles are easier to manufacture, require<br />

less structure, weigh less, and have less coupling with pitch modes. For these reasons, the impact<br />

on RCS was investigated for the 20° cant angle as well as the pitch coupling term for rudder<br />

deflection, C<br />

δ<br />

.<br />

m r<br />

The RCS code was run on two aircraft configurations. The same wing, fuselage, and horizontal<br />

were modeled with the vertical planforms mounted at both 20° and 30°. The results of that study<br />

are shown as Figure 10.4.<br />

63


40<br />

30<br />

20<br />

10<br />

0<br />

-10<br />

-20<br />

-30<br />

-40<br />

-50<br />

20° Canted Vertical<br />

30° Canted Vertical<br />

RFP Requirement<br />

Figure 10.4 - Radar Cross Section Impact of 20° vs. 30° Vertical Cant Angle<br />

Figure 69 clearly shows that there is an impact on the RCS for changing the cant angle. The RFP<br />

required -12 dBm 2 return is shown in red. As mentioned in the RCS section, this requirement is<br />

only mandated for the frontal 0° azimuth angle. Going to a 20° cant does not violate this<br />

requirement and yields the aforementioned benefits.<br />

The effective area of a rudder sized to 27% mean aerodynamic chord of the vertical was<br />

calculated in the horizontal plane of the aircraft. In normal non-canted configurations, Cm<br />

δ<br />

is<br />

r<br />

nonexistent. Table 10.II shows the values for this coupling term and various cants.<br />

Table 10.II - Pitching Moment Coupling with Rudder Deflection for Various Vertical Cant Angles<br />

C<br />

δ<br />

Vertical Cant Angle<br />

(165 ft 2 27% m.a.c. Rudder) m r<br />

0° 0.0000<br />

10° 0.0004<br />

20° 0.0009<br />

30° 0.0021<br />

64


The extra 10° cant resulted in a substantially larger pitch coupling term. In addition to the<br />

complications of canting more, a 30° angle would mean that a more complex mixer and control<br />

system would be required. This would add to the cost and is avoided.<br />

It is important to note that the previous static methods do not take into account the dynamic<br />

characteristics or modes of this aircraft. With such a large amount of the fuselage in front of the<br />

center of pressure, the Vendetta may require a complex yaw damper or larger vertical to<br />

compensate.<br />

The size of the vertical could potentially be driven by the one engine inoperative (OEI) control<br />

power requirements. Because the engine nozzle centerlines are mounted considerably offset from<br />

the centerline at 3 feet (0.9144 m), a large yawing moment will be created if the Vendetta loses<br />

an engine during takeoff. The engines produce roughly 45,000 pounds of thrust and would<br />

generate a 135,000 foot-pound moment. Table 10.III shows the results of the rudder control<br />

power analysis for this critical OEI condition.<br />

Table 10.III - Rudder Control Power Results for OEI Condition<br />

PARAMETER NOTATION VALUE<br />

Side Force due to Rudder C yδr 0.0105<br />

Rolling Moment due to Rudder C lδr 0.0072<br />

Rudder Effectiveness C nδr -0.0070<br />

OEI Critical Yawing Moment<br />

135,000 ft-lbs<br />

Rudder Deflection Required in OEI Condition at Takeoff 13.6°<br />

With a rudder effectiveness of -0.0070 (1/deg), a 13.6° rudder deflection is required to keep the<br />

aircraft flying straight in the OEI condition on takeoff. This is not too large, and would suffice by<br />

allowing approximately another 10° of rudder deflection for the pilot to yaw the aircraft beyond<br />

the straight condition for controllability. In this condition, the aircraft would be susceptible to<br />

large amounts of sideslip, β.<br />

This rudder deflection would be substantially higher if a higher cant angle were used. In these<br />

critical situations where the aircraft is in danger, the added drag created by the mixing is desired<br />

to be as little as possible.<br />

A separate 4 surface empennage was now made necessary because going to a purely V-tail was<br />

shown to be ill-advised at this stage because of the aforementioned studies. If a pure v-tail was<br />

chosen, it would have to be full-flying due to the demand placed on the surface and hinge lines in<br />

supersonic flight. This would require a large actuator and large structural members in the aft<br />

portion of the aircraft. This would considerably drive the configuration away from initial RCSfriendly<br />

layouts as well as increasing complexity and cost.<br />

65


The Vendetta configuration utilizes a 20° cant on the<br />

verticals and a separate full-flying horizontal as seen<br />

in Figure 10.5. It was mentioned earlier that one of<br />

the reasons the horizontal tail volume coefficient<br />

was larger in the historical aircraft was because<br />

those aircraft did not utilize control augmentation<br />

systems or digital fly-by-wire control systems. Not<br />

only did they have to account for wide shifts in CG,<br />

they also had to combat the muck tuck problem<br />

associated with breaking the sound barrier.<br />

Figure 10.5 - Vendetta Empennage Configuration<br />

Figure 10.6 shows that as the aircraft<br />

Mach trim<br />

exceeds the critical Mach number,<br />

the center of pressure of the wing<br />

and other control surfaces travels aft.<br />

In the case of the Vendetta, this<br />

leaves the CG an extra 12% m.a.c. in<br />

front of the neutral point; this makes<br />

it 12% more stable. This 12% shift<br />

was calculated with the Air Force’s<br />

Data Compendium (DATCOM)<br />

mg<br />

12% m.a.c.<br />

methods.<br />

Figure 10.6 - Mach Tuck Illustrated<br />

The shift in the neutral point of the<br />

wing means that the horizontal<br />

would have to deflect to keep the Vendetta from “tucking” under. The trim drag created could be<br />

avoided by shifting the CG, by altering the neutral point, or designing the aircraft to be unstable<br />

subsonic and stable supersonic.<br />

The use of a trim tank was investigated to pump fuel aft and shift the CG closer to the neutral<br />

point in supersonic cruise. This notion was dismissed because the tank would be a vacant waste<br />

of space and would complicate ground procedures where refueling would have to leave the tank<br />

vacant.<br />

The use of an extra flying surface such as a canard could be used as well. The canard would<br />

destabilize the aircraft by moving the neutral point forward and closer to the CG but it would<br />

make the Vendetta even more uncontrollable in the subsonic landing and takeoff conditions. This<br />

extra control surface would add to the cost and complexity.<br />

A fuel management system could be used to burn fuel from certain tanks to keep the CG travel in<br />

check. After analyzing the abrupt shift in the neutral point when the Vendetta climbs to its cruise<br />

condition, it was decided that the fuel management system could not pump fuel fast enough to<br />

66


trim the aircraft. The likewise was true when decelerating. The aircraft would suddenly go<br />

unstable. Because of this, use of a digital flight control system (DFCS) which is provided as GFE<br />

would allow the aircraft to fly unstable subsonic. The DFCS could easily allow a 0% - 7%<br />

unstable aircraft takeoff and land.<br />

The wing was placed and the empennage sized for the Vendetta to be 5% unstable in the<br />

subsonic regime and 7% stable in the supersonic regime without CG modification due to the<br />

12% shift. The fuel management system could then be used to enhance cruise performance by<br />

pumping fuel in a way which results in neutral or marginal static stability.<br />

A DFCS will not impact the design too much because complex navigation and autopilot systems<br />

will already have to be incorporated into the design. In addition to this, the DCFS will be used to<br />

enhance the dynamic modes of the aircraft. These may require it due to the large fore body and<br />

unstable pitch break exhibited by the Vendetta. Also, the 2010 delivery date will mean that next<br />

generation control laws and hardware could be implemented. All modern fighters being designed<br />

today utilize such systems already. The DFCS along with the fuel management system would<br />

maintain the static and dynamic stability.<br />

DATCOM and the compiled Digital DATCOM Fortran code proved to be useful tools in<br />

calculating many of the aerodynamic stability and control derivatives for the Vendetta. This was<br />

done in an attempt to identify problematic behaviors and to adhere to MIL-8785C.<br />

It was calculated that the Vendetta’s fuselage fore body will destabilize the aircraft an additional<br />

3.1% in subsonic cruise and 5.0% in supersonic cruise. The wing was placed to account for this.<br />

This is much improved over previous configurations where the fuselage destabilized the aircraft<br />

up to 16%. This is due to the fact that so fuselage with a large mean width was in front of the CG<br />

and NP.<br />

Figure 10.7 shows the Vendetta’s pitch break characteristics in the subsonic low speed and<br />

supercruise regimes given a CG location that would yield a statically stable aircraft.<br />

67


-0.6<br />

-0.4<br />

UNSTABLE<br />

NEUTRAL<br />

-0.2<br />

Lift Coefficient (C L )<br />

-0.25 -0.2 -0.15 -0.1 -0.05 0 0.05 0.1 0.15 0.2 0.25<br />

0<br />

0.2<br />

0.4<br />

M = 0.2<br />

M = 1.6<br />

0.6<br />

0.8<br />

Moment Coefficient (C m )<br />

Figure 10.7 - Pitch Break Characteristics<br />

This figure shows that as the Vendetta rotates and has some angle of attack in the low speed<br />

subsonic (M=0.2) regime, it will want to continue to rotate and break away. In the supercruise,<br />

the aircraft behaves mush more linearly. The subsonic characteristics are of some concern, but<br />

even simple feedback schemes in the DFCS could solve this problem. The supersonic<br />

characteristics are actually more desirable because the maneuvering required is very light and the<br />

control system will not be oscillating the control surfaces, which creates unnecessary drag, to<br />

keep the aircraft flying straight.<br />

A full state-space based model for the aircraft driven by a Taylor expansion and fit into equations<br />

of motion was developed for flight simulator validation. These forms are too complex for simple<br />

dynamic analysis, so the literal factor forms of the dynamics modes were used to determine<br />

flying qualities and conformity with MIL-8785C.<br />

The literal factors are nothing more than simplifications of the transfer function forms for<br />

longitudinal and lateral modes of interest. These forms omit insensitive stability derivatives. The<br />

literal factor forms for the modes of interest are provided in the Appendix. The conformity with<br />

the military specifications for handling quality is shown in Table 10.IV.<br />

68


Mode<br />

Table 10.IV - Longitudinal and Lateral Dynamic Mode Conformity with MIL-8785C<br />

Damping ratio (ζ) Natural Frequency (ω n )<br />

Vendetta MIL-8785C Vendetta MIL-8785C<br />

MIL-8785C<br />

Level<br />

Phugoid 0.094 > 0.04 0.091 - I<br />

Short Period 0.921 0.35 – 1.3 4.721 - I<br />

Dutch Roll 0.103 > 0.08 1.960 > 0.4 I<br />

Table 10.IV shows that the Vendetta satisfies all of the military specifications for these three<br />

important modes while in a subsonic cruise with the CG monitor. The only thing of concern<br />

regarding these results is high value for undamped natural frequency in the Dutch Roll mode. It<br />

is not uncommon for aircraft of this size and type to incorporate fairly simple yaw dampers<br />

operating on the yaw rate. With the use of the DFCS, the Vendetta would have no problem<br />

keeping that mode in control. Because there is a large amount of robustness available with CG<br />

excursion and the DFCS, the longitudinal modes are well within the Type I military<br />

specifications and would remain there in the supercruise.<br />

69


Validation of an aircraft design such as this is fairly hard to accomplish with traditional models.<br />

Traditional wind tunnel testing is limited to the availability of supersonic wind tunnels. Because<br />

of this, the <strong>Cal</strong> <strong>Poly</strong> Pheagle Flight Simulator was used to evaluate handling qualities, ground<br />

handling due to landing gear placement, up-and-away tasks, and low speed performance.<br />

A non-linear table lookup scheme was incorporated in the computer simulation to lookup the<br />

DATCOM derived stability and control derivatives. These derivatives were used to develop<br />

tables for the major derivatives as functions of Mach number and angle of attack. The simulator<br />

and flight cab are shown as Figure 10.8 and Figure 10.9, respectively.<br />

Figure 10.8 - Pheagle Simulator<br />

Figure 10.9 - Flight Cab and Instruments<br />

The static flight cab allows the pilot to see the a full panel of operational instruments as well as<br />

feel up to 50 pounds of force on the stick. The stick used was that from an F-15 Eagle. Graphics<br />

models were incorporated for pilot cues and situational awareness as seen in Figure 10.10. A full<br />

HUD was also incorporated and used as illustrated in Figure 10.11. Stick forces and dynamics<br />

were approximated based on existing information from current fighter aircraft.<br />

Figure 10.10 - Graphics and Environment<br />

Figure 10.11 - Heads up Display<br />

Results of validation proved that the DFCS control laws would need to be developed for the<br />

aircraft to be flyable in all flight regimes.<br />

70


11 Performance<br />

11.1 Performance Requirements<br />

The performance requirements listed in the RFP, shown in Table 11.I, consist of specific excess<br />

power requirements, a turn rate requirement, mission requirements, and takeoff and landing<br />

requirements. The specific excess power and turn rate requirements are measured at maneuver<br />

weight, which is defined as 50% internal fuel, and a payload of 2 × AIM-120 AMRAAM’s and 4<br />

× 2,000 lb (907 kg) JDAM’s. The mission requirements are defined as the ability to perform the<br />

design mission listed in Table 11.II.<br />

Table 11.I - RFP Performance Requirements<br />

Supercruise Mission Radius<br />

1,750 nm (3,240 nm)<br />

Supercruise Mach Number 1.6<br />

1g Military Specific Excess Power at 50,000 ft & Mach 1.6 0 ft/s (0 m/s)<br />

1g Maximum Specific Excess Power at 50,000 ft & Mach 1.6 200 ft/s (61.0 m/s)<br />

2g Maximum Specific Excess Power at 50,000 ft & Mach 1.6 0 ft/s (0 m/s)<br />

Maximum Instantaneous Turn Rate at 15,000 ft & Mach 0.9 8.0 deg/s<br />

Takeoff Field Length 8,000 ft (2,438 m)<br />

Landing Field Length (Icy) 8,000 ft (2,438 m)<br />

Table 11.II - RFP Design Mission<br />

1. Takeoff and acceleration allowance<br />

a. Fuel allowance for warm-up<br />

b. Fuel to accelerate to climb speed at maximum thrust (no distance credit)<br />

2. Climb from sea-level to optimum supercruise altitude<br />

3. Supercruise 1,000 nm (1,852 km) at Mach 1.6 and optimum altitude<br />

4. Climb above 50,000 ft (15,240 m)<br />

5. Dash 750 nm (1,389 km) at Mach 1.6 above 50,000 ft (15,240 m)<br />

6. Weapons delivery<br />

a. 180º turn at 50,000 ft (15,240 m) and Mach 1.6<br />

b. Drop air-to-surface weapons<br />

7. Dash 750 nm (1,389 km) at Mach 1.6 above 50,000 ft (15,240 m)<br />

8. Supercruise 1,000 nm (1,852 km) at Mach 1.6 and optimum altitude<br />

9. Descend to sea-level (no distance credit or fuel used)<br />

10. Reserve for 30 min at sea-level and speed for maximum endurance<br />

The RFP design mission explicitly defines some aspects of the required mission, while other<br />

aspects of the mission such as cruise altitudes and loiter speed are arbitrary. Within the<br />

constraints of the design mission, a detailed mission was created and optimized to minimize fuel<br />

71


consumption. The main aspects of the mission that were optimized were the initial climb<br />

sequence, the cruise and dash altitudes (dash altitude must be greater than 50,000 ft (15,240 m)),<br />

and the loiter speed. The optimum climb sequence was found by creating a flight envelope with<br />

lines of constant climb rate to fuel flow ratio (dh/dW F ) at the average climb weight of the<br />

aircraft. The climb profile that minimizes the fuel required to climb the aircraft to a given cruise<br />

condition is then found by drawing a flight path between the initial and cruise conditions that<br />

follow the maximum climb rate to fuel flow ratio. The resulting flight path and fuel<br />

consumption envelope are shown in Figure 11.1.<br />

50,000<br />

dh /dW F = 0 ft/lb<br />

100 ft/lb<br />

40,000<br />

Stall Limit<br />

50 ft/lb<br />

Altitude (ft)<br />

30,000<br />

20,000<br />

100 ft/lb<br />

150 ft/lb<br />

200 ft/lb<br />

250 ft/lb<br />

10,000<br />

300 ft/lb<br />

q Limit<br />

0<br />

0 0.5 1 1.5 2<br />

Mach<br />

Figure 11.1 - Fuel Consumption Envelope at Average Climb Weight<br />

The optimum cruise and dash altitudes were found by running a series of missions at different<br />

altitudes and finding the mission with the lowest fuel consumption. Because the aircraft’s<br />

weight decreases as fuel is burned, the optimum cruise altitude increases over the mission<br />

profile. It was found that the optimum sequence of cruise altitudes began at 52,000 ft (15,850 m)<br />

for the initial cruise and increased by 2,000 ft (610 m) for each successive cruise or dash segment<br />

resulting in a final cruise altitude of 58,000 ft (17,678 m). The optimum loiter speed was found<br />

as the speed at which the minimum drag occurred under loiter conditions (sea level and 61,000 lb<br />

(27,669 kg) weight). The drag on the aircraft under these conditions is plotted as a function of<br />

Mach number in Figure 11.2 showing that the minimum drag occurs at Mach 0.35 or 391 ft/s<br />

(119.2 m/s). The resulting detailed mission is listed in Table 11.III.<br />

72


30,000<br />

25,000<br />

Wave Drag<br />

Drag (lb)<br />

20,000<br />

15,000<br />

Sea-Level<br />

Loiter Weight<br />

10,000<br />

5,000<br />

Induced Drag<br />

Parasite Drag<br />

0<br />

0 0.2 0.350.4 0.6 0.8 1 1.2<br />

Mach<br />

Figure 11.2 - Drag on Aircraft in Loiter Conditions<br />

Table 11.III - Detailed Design Mission<br />

1. Warm-up 2 min at idle thrust<br />

2. Takeoff – Accelerate to takeoff speed 266 ft/s (81.1 m/s)<br />

3. Accelerate to Mach 0.86 at maximum military thrust<br />

4. Climb to 15,000 ft (4,572 m) and accelerate to Mach 0.9<br />

5. Climb to 17,500 ft (5,334 m) and accelerate to Mach 1.25<br />

6. Climb to 33,000 ft (10,058 m) and accelerate to Mach 1.72<br />

7. Climb to 45,000 ft (13,716 m) at Mach 1.72<br />

8. Climb to 52,000 ft (15,850 m) and decelerate to Mach 1.6<br />

9. Cruise 1,000 nm (1,852 km) at 52,000 ft (15,850 m) and Mach 1.6<br />

10. Climb to 54,000 ft (16,459 m) at Mach 1.6<br />

11. Dash 750 nm (1,389 km) at 54,000 ft (16,459 m) and Mach 1.6<br />

12. Descend to 50,000 ft (15,240 m) at Mach 1.6<br />

13. Turn 180º at n = 1.55<br />

14. Drop 4 × 2,000 lb (907 kg) JDAM’s<br />

15. Climb to 56,000 ft (17,069 m) at Mach 1.6<br />

16. Dash 750 nm (1,389 km) at 56,000 ft (17,069 m) and Mach 1.6<br />

17. Climb to 58,000 ft (17,678 m) at Mach 1.6<br />

18. Cruise 1,000 nm (1,852 km) at 58,000 ft (17,678 m) and Mach 1.6<br />

19. Descend to Sea-Level and decelerate to loiter speed 391 ft/s (119.2 m/s)<br />

20. Loiter 30 min at Sea-level and 397 ft/s (121.0 m/s)<br />

21. Decelerate to landing speed 266 ft/s (81.1 m/s)<br />

22. Land – decelerate to zero speed<br />

23. Unload non-fixed equipment (2 × AMRAAM’s and crew 1,280lb (5,806 kg))<br />

73


In addition to the performance requirements, the RFP includes performance measures of merit<br />

listed in Table 11.IV. The measures of merit are not specific requirements, but are measures by<br />

which competing aircraft will be judged.<br />

Table 11.IV - Performance Measures of Merit<br />

Mission duration, radius, and fuel burn by mission segment for design mission<br />

Takeoff and landing distance for design mission on dry and icy runways at sea-level<br />

Maximum Mach number at 36,000 ft (10,973 m)<br />

1g Maximum thrust specific excess power envelope<br />

2g Maximum thrust specific excess power envelope<br />

5g Maximum thrust specific excess power envelope<br />

Maximum thrust sustained load factor envelope<br />

Maximum thrust maneuvering performance diagrams at sea-level and 15,000 ft (4,572 km)<br />

11.2 Specific Excess Power Requirements<br />

Compliance with the specific excess power requirements is best shown using specific excess<br />

power envelopes. Figure 11.3 shows the 1g military specific excess power envelope. The RFP<br />

requirement of 0 ft/s (0 m/s) at Mach 1.6 and 50,000 ft (15,240 m) is met with 95.7 ft/s (29.2<br />

m/s) specific excess power. The 1g maximum (afterburner) specific excess power envelope,<br />

shown in Figure 11.4, shows that the RFP requirement of 200 ft/s (61.0 m/s) at Mach 1.6 and<br />

50,000 ft (15,240 m) is met with a value of 316 ft/s (96.3 m/s). This envelope also shows that<br />

the maximum Mach number at 36,000 ft (10,973 m) is 2.1. The 2g maximum specific excess<br />

power envelope, shown in Figure 11.5, shows that the RFP requirement of 0 ft/s (0 m/s) at Mach<br />

1.6 and 50,000 ft (15,240 m) is met with a value of 108 ft/s (32.9 m/s).<br />

74


70,000<br />

60,000<br />

50,000<br />

Stall Limit<br />

P s = 0 ft/s<br />

RFP Requirement 0 ft/s<br />

P s = 200 ft/s<br />

Altitude (ft)<br />

40,000<br />

30,000<br />

P s = 100 ft/s<br />

Flaps<br />

P s = 200 ft/s<br />

20,000<br />

P s = 300 ft/s<br />

q Limit<br />

10,000<br />

P s = 400 ft/s<br />

0<br />

0 0.5 1 1.5 2 2.5<br />

Mach<br />

Figure 11.3 - 1g Military Specific Excess Power Envelope at Maneuver Weight<br />

70,000<br />

P s = 0 ft/s<br />

60,000<br />

Stall Limit<br />

50,000<br />

RFP Requirement 200 ft/s<br />

Altitude (ft)<br />

40,000<br />

30,000<br />

20,000<br />

Flap<br />

P s = 200 ft/s<br />

P s = 400 ft/s<br />

P s = 600 ft/s<br />

P s = 600 ft/s<br />

q Limit<br />

10,000<br />

P s = 800 ft/s<br />

0<br />

0 0.5 1 1.5 2 2.5<br />

Mach<br />

Figure 11.4 - 1g Maximum Specific Excess Power Envelope at Maneuver Weight<br />

75


70,000<br />

Altitude (ft)<br />

60,000<br />

50,000<br />

40,000<br />

30,000<br />

20,000<br />

10,000<br />

Stall Limit<br />

P s = 200 ft/s<br />

P s = 400 ft/s<br />

P s = 600 ft/s<br />

P s = 0 ft/s<br />

RFP Requirement 0 ft/s<br />

P s = 600 ft/s<br />

q Limit<br />

0<br />

0 0.5 1 1.5 2 2.5<br />

Mach<br />

Figure 11.5 - 2g Maximum Specific Excess Power Envelope at Maneuver Weight<br />

11.3 Turn Rate Requirement<br />

The maximum instantaneous turn rate requirement of 8.0 deg/s at 15,000 ft (4,572 m) and Mach<br />

0.9 is shown in the maneuverability diagram at 15,000 ft (4,572 m) in Figure 11.6. The<br />

maneuverability diagram shows that the required turn can be sustained with military power. The<br />

maximum sustainable turn rate under maximum power is 13.1 deg/s.<br />

76


30<br />

n<br />

2<br />

3<br />

4 5 6 7<br />

r = 2,000 ft<br />

25<br />

r = 4,000 ft<br />

Turn Rate (deg/s)<br />

20<br />

15<br />

10<br />

5<br />

Stall Limit<br />

Max. P s = 0<br />

Mil. P s = 0<br />

RFP Requirement<br />

8 deg/s<br />

r = 6,000 ft<br />

r = 8,000 ft<br />

r = 10,000 ft<br />

q Limit<br />

0<br />

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2<br />

Mach<br />

Figure 11.6 - Maneuverability Diagram at 15,000 ft (4,572 m) and Maneuver Weight<br />

77


11.4 Mission Requirements<br />

By completing the design mission, the requirements for a supercruise Mach number of 1.6 and<br />

mission radius of 1,750 nm (3,240 km) are met. To determine the fuel capacity required to<br />

perform the design mission, the mission was simulated by numerically integrating the fuel burn<br />

rates over the mission profile. The mission simulation also allowed certain aspects of the<br />

mission such as cruise altitudes to be optimized. Table 11.V lists the results of the mission<br />

simulation, Figure 11.7 shows a breakdown of fuel consumption by mission segment, and Table<br />

11.VI lists the fuel consumption by mission segment.<br />

Reserve<br />

4%<br />

Misc.<br />

5%<br />

Accelerate &<br />

Climb<br />

12%<br />

Cruise Back<br />

18%<br />

Cruise Out<br />

28%<br />

Dash Back<br />

15%<br />

Dash Out<br />

18%<br />

Figure 11.7 - Fuel Consumption over Mission<br />

Table 11.V - Mission Results<br />

Total fuel consumption<br />

57,940 lb (26,280 kg)<br />

Mission radius<br />

1,750 nm (3,240 km)<br />

Total distance traveled over mission 3,980 nm (7,370 km)<br />

Total mission duration<br />

4 hr. 57 min.<br />

Takeoff weight<br />

124,800 lb (56,610 kg)<br />

Empty weight<br />

56,270 lb (25,520 kg)<br />

Fuel weight (total fuel onboard) 59,250 lb (26,880 kg)<br />

Maneuver weight<br />

95,830 lb (43,470 kg)<br />

Landing weight<br />

57,550 lb (26,100 kg)<br />

Average cruise lift to drag ratio 6.6<br />

78


Table 11.VI - Fuel Consumption by Mission Segment<br />

1. Warm-up 2 min at idle thrust 83 lb (37.6 kg)<br />

2. Takeoff (Accelerate to takeoff speed 266 ft/s) 174 lb (78.9 kg)<br />

3. Accelerate to Mach 0.86 at maximum military thrust 741 lb (336.1 kg)<br />

4. Climb to 15,000 ft and accelerate to Mach 0.9 724 lb (328.4 kg)<br />

5. Climb to 17,500 ft and accelerate to Mach 1.25 935 lb (424.1 kg)<br />

6. Climb to 33,000 ft and accelerate to Mach 1.72 3,528 lb (1,600.2 kg)<br />

7. Climb to 45,000 ft at Mach 1.72 805 lb (365.1 kg)<br />

8. Climb to 52,000 ft and decelerate to Mach 1.6 214 lb (97.1 kg)<br />

9. Cruise 1,000 nm at 52,000 ft and Mach 1.6 15,893 lb (7,208.9 kg)<br />

10. Climb to 54,000 ft at Mach 1.6 262 lb (118.8 kg)<br />

11. Dash 750 nm at 54,000 ft and Mach 1.6 10,589 lb (4,803.1 kg)<br />

12. Descend to 50,000 ft at Mach 1.6 53 lb (24.0 kg)<br />

13. Turn 180º at n = 1.55 658 lb (298.5 kg)<br />

14. Drop 4 × 2,000 lb JDAM’s 0 lb (0 kg)<br />

15. Climb to 56,000 ft at Mach 1.6 381 lb (172.8 kg)<br />

16. Dash 750 nm at 55,000 ft and Mach 1.6 8,879 lb (4,027 kg)<br />

17. Climb to 58,000 ft at Mach 1.6 139 lb (63.0 kg)<br />

18. Cruise 1,000 nm at 58,000 ft and Mach 1.6 10,532 lb (4,777.2 kg)<br />

19. Descend to Sea-Level and decelerate to loiter speed (391 737 lb (334.3 kg)<br />

ft/s)<br />

20. Loiter 30 min at Sea-level and 397 ft/s 2,453 lb (1112.6 kg)<br />

21. Decelerate to landing speed 266 ft/s 132 lb (59.9 kg)<br />

22. Land (decelerate to zero speed) 25 lb (11.3 kg)<br />

23. Unload non-fixed equipment (2 × AMRAAM’s and crew<br />

1,280lb)<br />

0 lb (0 kg)<br />

11.5 Takeoff & Landing<br />

The RFP requires that the aircraft be able to takeoff and land on an icy standard NATO runway<br />

8,000 ft (2,438 m) long. Takeoff and landing calculations were done according to MIL-C5011A.<br />

Takeoff and landing were simulated by numerically integrating velocity and rate of climb to<br />

determine distances and altitudes over a standard flight profile. Additional drag due to flaps and<br />

landing gear was taken into account for takeoff and landing as well as -25% military thrust from<br />

a thrust reverser during landing. The takeoff and landing profiles used in the simulation are<br />

listed in Table 11.VII and Table 11.VIII.<br />

Table 11.VII - Takeoff Flight Profile<br />

1. Ground roll under max. power up to 1.2 V stall<br />

2. Rotate for 3 s<br />

3. 1.15g pull-up to transition to climb<br />

4. Climb out at a 10º climb angle over 50 ft obstacle<br />

79


Table 11.VIII - Landing Flight Profile<br />

1. Approach from 50 ft obstacle on 3º glide slope<br />

2. 1.15g pull-up for flare – touchdown at 1.2 V stall<br />

3. Roll for 3 s<br />

4. Brake and apply thrust reversers until stopped<br />

MIL-C5011A defines field length to be the distance required to takeoff and clear a 50 ft (15.2 m)<br />

obstacle or the distance to land from a 50 ft (15.2 m) obstacle. Takeoff and touchdown speed are<br />

defined as 1.2 times the aircraft’s stall speed, and the speed over the 50 ft (15.2 m) obstacle must<br />

be greater or equal to 1.3 times the stall speed for both takeoff and landing. The takeoff gross<br />

weight for the design mission of 124,800 lb (56,610 kg) was used for the aircraft weight for both<br />

takeoff and landing calculations. This allows the aircraft to land immediately after takeoff<br />

without the need to jettison fuel or weapons. Takeoff and touchdown speeds were always greater<br />

than the required 1.2 stall speed because of the acceleration during the 3 second rotation and roll<br />

periods. During takeoff, due to the high speeds of the aircraft, the 50 ft (15.2 m) obstacle was<br />

cleared before the climb angle was reached, so the climb segment of the profile was ignored.<br />

Also, to simplify the calculations, the landing simulation was run backward so that the<br />

touchdown point could be found without having to calculate the altitude and speed at the<br />

beginning of the flare necessary to have the touchdown occur at the correct altitude and speed.<br />

The results of the takeoff and landing simulations that are listed in Table 11.IX and Table 11.X<br />

show that the RFP requirements for takeoff and landing on an icy 8,000 ft (2,438 m) runway are<br />

met.<br />

Table 11.IX - Takeoff Results<br />

Weight<br />

124,800 lb (56,610 kg)<br />

Maximum lift coefficient 1.2<br />

Stall speed 242 ft/s (73.7 m/s)<br />

Takeoff speed 266 ft/s (81.1 m/s)<br />

50 ft obstacle speed ≥ 290 ft/s (88.4 m/s) 326 ft/s (99.4 m/s)<br />

Rolling friction coefficient 0.025<br />

Runway length 3,842 ft (1,171 m)<br />

Field length over 50 ft obstacle 5,280 ft (1,609 m)<br />

80


Table 11.X - Landing Results<br />

Weight<br />

124,800 lb (56,610 kg)<br />

Maximum lift coefficient 1.2<br />

Stall speed 242 ft/s (73.7 m/s)<br />

Takeoff speed 266 ft/s (81.1 m/s)<br />

50 ft obstacle speed ≥ 290 ft/s (88.4 m/s) 294 ft/s (89.6 m/s)<br />

Dry braking friction coefficient 0.3<br />

Icy braking friction coefficient 0.1<br />

Thrust reverser effectiveness<br />

25% Mil.<br />

Dry runway length 3,857 ft (1,176 m)<br />

Dry field length over 50 ft obstacle 5,256 ft (1,602 m)<br />

Icy runway length 5,707 ft (1,739 m)<br />

Icy field length over 50 ft obstacle 7,107 ft (2,166 m)<br />

11.6 Performance Summary<br />

As shown above, the Vendetta meets all performance requirements in the RFP. A summary of<br />

the performance requirements is shown in Table 11.XI.<br />

Table 11.XI - Performance Summary<br />

Requirement<br />

Value<br />

Supercruise Mach Number 1.6 1.6<br />

Supercruise Mission Radius 1,750 nm (3,240 km) 1,750 nm (3,240 km)<br />

Dash Altitude > 50,000 ft (15,240 km) > 54,000 ft (16,460 km)<br />

1g Mil. Ps M = 1.6, 50,000 ft 0 ft/s (0 m/s) 95.7 ft/s (29.2 m/s)<br />

1g Max. Ps M = 1.6, 50,000 ft 200 ft/s (61.0 m/s) 316 ft/s (96.3 m/s)<br />

2g Max. Ps M = 1.6, 50,000 ft 0 ft/s (0 m/s) 108 ft/s (32.9 m/s)<br />

Max. Turn Rate M = 0.9, 15,000 ft 8 deg/s 13.1 deg/s<br />

Takeoff Field Length 8,000 ft (2,438 m) 5,280 ft (1,609 m)<br />

Landing Field Length 8,000 ft (2,438 m) 7,107 ft (2,166 m)<br />

Fuel Burn Over Mission ─ 57,940 lb (26,280 kg)<br />

Maximum Mach Number at 36,000 ft ─ 2.1<br />

81


11.7 Alternate Missions<br />

In addition to the design mission, the Vendetta shows great performance on alternate missions.<br />

The multi purpose rotary launcher (MPRL), shown in Figure 11.8, carried by the Vendetta allows<br />

it to carry a total of 8 × 2,000 lb (907 kg) bombs compared to the 4 required for the design<br />

mission (no AMRAAM’s can be carried in this configuration.) Compared to an earlier custom<br />

designed rotary launcher, the MPRL has only a 4 in (10 cm) greater diameter. The extra cross<br />

sectional area due to the MPRL only adds approximately 1,300 lb (590 kg) of fuel consumption<br />

for the design mission. The performance of the Vendetta over three alternate missions was<br />

calculated. Fully loaded missions and subsonic missions flown at Mach 0.8 and an altitude of<br />

approximately 30,000 ft (9,144 m) were considered. The results shown in Table 11.XII indicate<br />

that only a small loss of range occurs due to the additional weight of 8 × 2,000 lb (907 kg)<br />

bombs, and the range of the aircraft can be greatly extended by flying subsonic (although it<br />

extends the mission to 11 hours long.)<br />

Figure 11.8 - MPRL with 8 × 2,000 lb (907 kg) JDAM’s<br />

82


Table 11.XII - Alternate Mission Results<br />

Design Mission<br />

Mission Radius 1,750 nm (3,240 km)<br />

Takeoff Weight<br />

124,800 lb (56,610 kg)<br />

Mission Time<br />

4 hr. 57 min.<br />

8 × 2,000 lb (907 kg) bombs – Supersonic<br />

Mission Radius 1,690 nm (3,130 km)<br />

Takeoff Weight<br />

132,800 lb (60,240 kg)<br />

Mission Time<br />

4 hr. 51 min.<br />

4 × 2,000 lb (907 kg) bombs – Subsonic<br />

Mission Radius 2,380 nm (4,410 km)<br />

Takeoff Weight<br />

124,800 lb (56,610 kg)<br />

Mission Time<br />

11 hr. 16 min.<br />

8 × 2,000 lb (907 kg) bombs – Subsonic<br />

Mission Radius 2,300 nm (4,260 km)<br />

Takeoff Weight<br />

132,800 lb (60,240 kg)<br />

Mission Time<br />

10 hr. 56 min.<br />

83


12 Payload<br />

Weapon internal layout drove the Vendetta’s size and layout. For small C.G. excursion due to<br />

weapons deployment all stores<br />

were initially positioned as close to<br />

the C.G as possible. As shown in<br />

Figure 12.1, three configurations<br />

were produced. Configuration one<br />

utilizes a standard weapons bay<br />

configuration. The large weapons<br />

bay drove the configuration to over<br />

120 ft (36.576m) in length when<br />

room for landing gear and<br />

weapons targeting systems were<br />

integrated. In an effort to decrease<br />

overall size a small rotary launcher<br />

was designed and integrated into a<br />

second configuration. This step<br />

increased the maximum cross<br />

sectional area by 5ft 2 (0.464m 2 Figure 12.1 - L to R configurations 1, 2, 3<br />

)<br />

and shortened the length of the<br />

aircraft to 95ft (28.9m).<br />

Figure 12.2 - 180 inch MPRL<br />

The next iteration of the design utilized the<br />

existing 180in (4.57m) MPRL out of the B-1B<br />

and shown in Figure 12.2. This caused the final<br />

configuration to grow to 105ft (32m) in length<br />

and maximum cross sectional area of 55 ft 2<br />

(5.11m 2 ). Utilizing the MPRL allowed for a wide<br />

selection of alternate missions and configurations<br />

to be discussed later.<br />

In an effort to ascertain the feasibility of RFP delineated weapons as supersonic deployment<br />

candidates research was done for each proposed system. As can be seen<br />

in the foldout only one of all of the weapons has even been wind tunnel<br />

tested for supersonic deployment. This led to the proposed systems for<br />

possible supersonic deployment.<br />

Retrofitting many of the weapon systems with a ballute and sabot,<br />

shown in Figure 12.3, would aide in supersonic stability. The use of a<br />

bomb bay supersonic flow deflector and acoustical resonance damping<br />

system as well as flow modification system would aide in deployment.<br />

Figure 12.3 - Ballute<br />

and Sabot<br />

84


Standard ten degree fall clearance is maintained for all weapons<br />

stowed and weapons bay doors were designed to rotate into the<br />

bomb bay, shown in Figure 12.4, and not into the free stream.<br />

Rotating the bomb bay doors into the fuselage has no detrimental<br />

effects on lateral stability, allows for the usage of lighter bomb bay<br />

doors, and lowers the radar cross section of the aircraft when the<br />

bomb bay is open versus a door which opens into the free stream<br />

such as those found on the F-117.<br />

Figure 12.5 - 30in (76.2cm)<br />

Ejector rack<br />

Alternative<br />

uses for the Vendetta is the main<br />

reasoning behind the choice of the<br />

MPRL. According to MIL-A-8861B a<br />

fighter/attack aircraft must be capable of<br />

withstanding, under “basic flight design<br />

gross weight” +7.5, -3 g’s. under<br />

“maximum design gross weight” +5.5,<br />

-2 g’s. According to the RFP we must<br />

Figure 12.4 - Bomb Bay<br />

Door Retraction Scheme<br />

In an effort to minimize undesirable underbody flow the entire<br />

underside of the aircraft was kept as flat as possible. The MPRL<br />

chosen allows the use of 30in (76.2cm) ejector racks shown in Figure<br />

12.5. The rack has electrically fired impulse cartridges, a gas operated<br />

mechanism, and is designed to forcibly eject conventional or nuclear<br />

stores up to 4000 lb (1810 kg) weight class. The LAU-142A ejector<br />

was used with the AIM-120C shown in Figure 12.6.<br />

be able, with the design load, and 50% of fuel volume structurally be able to withstand +7 -3 g’s.<br />

Thus it is conceivable to design the aircraft to structurally withstand the RFP’s mandated +7 -3 g<br />

loading with the RFP design load of (4) 2000lb (905 kg) JDAMs and (2) AIM-120’s.<br />

Alternatively the aircraft could then be loaded with (8) 2000 lb (905 kg) JDAM’s shown in<br />

Figure 12.7, and still meet the military specification of +5.5 -2 g’s at “maximum design gross<br />

weight”. This fully satisfies the RFP and military specifications.<br />

This doubling of weapons load shortens mission radius by<br />

150 miles (241 km) (see performance section). Alternative<br />

mission configurations might also include eight 2000lb (905<br />

kg) fuel tanks for increased ferry range or 8 JASSM next<br />

generation cruise missiles specifically designed for the B-1’s<br />

MPRL.<br />

Weapons guidance is accomplished with the on board<br />

INS/GPS system (all weapons), a AN/APG-77 radar system<br />

(AIM-120 guidance) as well as an on-board second<br />

generation tessa LANTIRN system (laser designated<br />

weapons). More detailed information on weapons can be found in Foldout 2 of the Appendix.<br />

Figure 12.7 - (8) 2000lb JDAM + MPRL<br />

Figure 12.6 - LAU-142A Ejection Sequence<br />

85


13 Cockpit<br />

Cockpit Design began with the RFP requirement of a crew of two. Comparison trades of tandem<br />

versus side-by-side seating configurations were performed as shown in Figure 13.1. Analysis of<br />

the two configurations was based upon a cockpit solid model where instrumentation, controls,<br />

circuit breakers and military aft pilot vision requirements (MIL-STD-850B) were used to<br />

construct the tandem and side by side arrangements.<br />

Figure 13.1 - Cockpit Width Trade Study<br />

Tandem vs. side by side configuration has very little effect on frontal area of the cockpit<br />

configuration in military aircraft. It was found due to instrumentation and control placement,<br />

which were the major cockpit width contributors, that other factors must be taken into account<br />

before a final decision could be made on pilot placements. Weapons configuration, the use of the<br />

180 in (457 cm) MPRL, favored side by side seating placement. The rational was the width of<br />

the rotary launching system allowed for the use of a side-by-side seating arrangement. This<br />

arrangement allowed greater pilot communication as well as<br />

the elimination of many redundant circuit breakers as well as<br />

instrumentation. However preliminary stability and control<br />

analysis revealed a need to narrow the forward fuselage, as<br />

shown in Figure 13.2, due to adverse C mα characteristics of a<br />

wide nose section (see stability and control section).<br />

Figure 13.2 - Forward fuselage<br />

Comparisons<br />

Therefore the decision was made to utilize a tandem seating<br />

configuration. This configuration offered a smaller frontal<br />

area, a much more ideal supersonic (M=1.6) area plot, and a<br />

better field of vision for the primary pilot.<br />

86


Table 13.I - Military Vision Specifications<br />

Forward Pilot<br />

5.1.1 Vision<br />

azimuth<br />

(°)<br />

up<br />

(°)<br />

down<br />

(°)<br />

0 10 11<br />

20 20<br />

30 25<br />

90 40<br />

135 20<br />

11°<br />

5°<br />

5.1.2 Aft Pilot Position<br />

0 5<br />

Figure 13.3 - Virtual Cockpit Model<br />

Utilizing this information the virtual cockpit model shown in Figure 13.3 was generated. The<br />

solid model also took into account the use of an ejection seat, room for instrumentation, controls,<br />

switch placement as well as the above military vision specifications.<br />

Further vision refinement produced rectilinear vision plots as shown in Figure 13.4.<br />

Canopy reinforcing structure was removed from the areas between 25° and 40° up to aide in inflight<br />

refueling vision. Runway vision areas are defined and every effort was made to increase<br />

downward vision to aide in ground handling as well as takeoff and landing.<br />

Takeoff and Landing vision is inherently limited in supersonic aircraft. In an effort to reduce<br />

pilot workload and increase flight ability, multifunction displays (MFD) in modern aircraft can<br />

double as vision aid tools. MFD 1 incorporated into the glare shield and upper instrument panel<br />

Figure 13.4 – Rectilinear Vision Plot of Forward Cockpit Position<br />

87


that could be used in landing and takeoff to increase downward<br />

vision, meshing seamlessly with the actual cockpit over nose<br />

view shown in Figure 13.5.<br />

MFD’s 2, 3, and 4 display moving map imagery, flight critical<br />

data, and mission critical information. The moving map<br />

display could also double in landings as another artifical vision<br />

aide. Perhaps utilizing infra-red or other electromagnetic<br />

spectrums for poor weather penetration and increased all<br />

weather capabilities. The standard dash mounted HUD was<br />

dropped in favor of a current helmet mounted HUD systems<br />

under development shown in Figure 13.6. The HUD allows far<br />

superior situational awareness as well as more aerodynamic<br />

canopy configuration.<br />

Figure 13.5 - Cockpit Display<br />

Arrangement<br />

Figure 13.6 - Helmet<br />

Mounted HUD<br />

Aircrew safety was a primary concern in the design of Vendetta. Due to<br />

RFP requirements the majority of the Vendetta’s mission will occur<br />

above the military specified ceiling for flight without a full pressure suit<br />

(50,000ft). Further research revealed the reasoning behind the<br />

specification. The NASA Bioastronautics study SP-3006 shows that<br />

animal and human life functions become critically affected by the lower<br />

oxygen content and lower pressure of the upper atmosphere. The study<br />

outlines how physiological effects such as the bends and hypoxia as<br />

well as the extremely low temperatures of high altitude within seconds<br />

render a human unconscious and dead in a mater of minutes. Also<br />

outlined is the Armstrong Line (63,000 ft), or the altitude at which<br />

water, at room temperature, will freely boil. In the study it shows how<br />

animals survived momentary exposure to altitudes higher than 63,000ft<br />

due to intravenous pressure keeping the blood within their veins liquid.<br />

Balancing this information against the economics and long prep and turn around time associated<br />

with full pressure suits the decision was made to opt for a partial pressure suit configuration. The<br />

advanced fighter crew protection system is shown in Figure 13.8. A partial pressure suit system<br />

was developed specifically for this altitude mission. It represents the next step beyond current<br />

systems and offers low unit cost in comparison to full pressure suits as well as low turn around<br />

time due to no necessity for a suiting procedure which involves lowering of blood nitrogen levels<br />

such as those used in the U-2.<br />

88


Ejection seat trade<br />

studies will need to be<br />

performed for the<br />

final configuration.<br />

The ACES II ejection<br />

seat shown in Figure<br />

13.7 was initially<br />

chosen as the premier<br />

American seat<br />

available. However<br />

due to global politics<br />

the availability of<br />

many superior<br />

Russian seats would<br />

produce a necessary<br />

Figure 13.7 - ACES II<br />

Ejection Seat<br />

trade between increased cost of bi-lingual<br />

design teams, language retrofit on new<br />

equipment, product per system cost and the<br />

unknown safety superiority of the Russian<br />

seat(s).<br />

Figure 13.8 - AFCPS<br />

Further pilot safety issues were found in the<br />

canopy design. The use of an ejectable canopy<br />

was thrown out over the choice of a shape<br />

charge cutting system due to the possibility of<br />

aircrew and canopy striking each other after<br />

ejection.<br />

89


14 Systems<br />

The systems of the Vendetta will closely follow the design architecture of the F-22. Technology<br />

advances by 2020 will render most of the electronics aboard the F-22 antequated, thus the next<br />

generation of this system should be implemented. Design trades on communications, processor<br />

I/O (such as unified vs. federated), system redundancy, and actuation will have to be performed<br />

on the new system.<br />

14.1 Auxiliary Power Generation System<br />

The auxiliary power generation<br />

system consists of two components:<br />

an auxiliary power unit (APU) aand<br />

a self-contained energy storage<br />

system (SES). APU selection<br />

involved examining a number of<br />

mid-sized, gas turbine, generators<br />

with output exceeding the minimum<br />

350 Hp (261.1 kW) estimated as<br />

needed for the Vendetta. A<br />

shortened list appears in Table 14.I.<br />

The RFP lists an APU but research<br />

Company<br />

Table 14.I - APU Selection Table<br />

points to the cost, volume and weight of the APU specified as being a Ram Air Turbine. A Ram<br />

Air Turbine (RAT) was eliminated due to the need for ground power and previous service<br />

experience. Modern designs utilize a SES for power backup due to its higher reliability,<br />

invulnerability to aircraft maneuver position, and airspeed.<br />

Model<br />

Honeywell 131-9A<br />

Honeywell 36-300<br />

Honeywell 331-200<br />

Pratt and<br />

Whitney PW901<br />

Sundstrand<br />

Sundstrand<br />

APS2100<br />

APS3200<br />

Startup<br />

Ceiling<br />

Dry<br />

Weight Rating Power/Weight<br />

ft (m) lb (kg) kW Hp Hp/lb (kW/kg)<br />

41000 350<br />

(12497) (158) 343 460 1.31 (2.17)<br />

35000 300<br />

(10668) (136) 291 390 1.3 (2.14)<br />

43000 500<br />

(13106) (226) 432 579 1.16 (1.91)<br />

25000 835<br />

(7620) (378) 1145 1535 1.84 (3.03)<br />

37000 270<br />

(11278) (122) 376 504 1.87 (3.08)<br />

39000 305<br />

(11887) (138) 450 603 1.98 (3.26)<br />

Due to common unreliability of in-flight APU startup, startup ceiling<br />

was not seen as a major driver in APU selection. Overall high power<br />

to weight ratio as well as a rating above 350 Hp (261.1 kW) and small<br />

size was seen as the main APU selection criteria. With this in mind the<br />

Sundstrand APS 3200 was selected. Placement of the APU can be<br />

seen in Figure 14.1.<br />

Figure 14.1 - Sundstrand<br />

APS 3200 Location<br />

The SES is a design point to be focused on in the next level of design.<br />

Current hypergolically fuelled systems offer high power to weigh ratio<br />

but fuels are hazardous and expensive. Next generation fuel cell<br />

systems offer reasonably high power to weight ratios with standard<br />

JP-8. Trade studies considering cost of development, service life cost<br />

savings, and internal volume usage will have to be performed.<br />

90


14.2 Vehicle Management System<br />

The vehicle management system (VMS) of the aircraft provides flight and propulsion control and<br />

includes the following systems: Control stick, Throttle controls, Rudder pedals, and actuators,<br />

Air data probes, Accelerometers Leading-edge, flap drive actuators, Primary flight control<br />

actuators .<br />

The choice of primary actuator type was the first trade examined. Four choices are commonly<br />

available for these systems:<br />

1) Electrohydrostatic<br />

2) Electric<br />

3) Pneumatic<br />

4) Hydraulic<br />

Pneumatic systems were eliminated due to low power to weight, large comparative size, and low<br />

power transmission efficiencies. Electrohydrostatic actuators, although offering many benefits<br />

such as a “line replaceable unit”, optimized per service dynamic impedance shaping, and<br />

optimized K factor suffered from low observability and weight problems as a consequence of<br />

electric power transmission throughout the aircraft( EM emissions in the IR (actuator heating)<br />

and Radio (cables) spectra). This low observability problem to an even greater degree affected<br />

the all electric system which utilizes a comparatively large and heavy actuator.<br />

Thus the decision was made to utilize an all hydraulic system with digital fly-by-wire control.<br />

The F-119 utilized in the F-22 has a PTO driving two 72 gpm (273 lpm) pumps (main line), four<br />

pumps total to supply two independent 4000 psi (27.6 GPa) systems. The choice of the high<br />

pressure systems was due to weight and volume considerations. Peak hydraulic system demand<br />

will be satisfied via an air pressurized hydraulic reservoir allowing for a constant energy bleed<br />

from the engines.<br />

Further detail was then examined into the electric system of the aircraft. Borrowing the electrical<br />

system from the F-22, a Smiths Industries 270 volt, direct current (DC) electrical system was<br />

chosen. It uses two PTO driven 65 kilowatt generators.<br />

Next level design trade studies need to be completed on the redundancy level of the electrical<br />

and flight control systems. Flight control system design should be aimed at Operation-<br />

Operational-Operational Fail (OOOF) or better, safety design. INS/GPS and other navigational<br />

systems should be examined and a proper redundancy level chosen.<br />

91


14.3 Fuel System<br />

Initial fuel system design began with<br />

configuration placement of fuel tanks<br />

symmetrically about the CG laterally and<br />

longitudinally. The two fuselage tanks (Figure<br />

14.2) serve to trim the aircraft in flight,<br />

necessary for the shift in neutral point location<br />

due to supersonic flight.<br />

Aircraft sizing began with fuel load<br />

internal volume necessary to complete the<br />

mission. The final configuration provides for<br />

Figure 14.2 - Fuel Tank Locations in Vendetta 68,000 lb (30,770 kg) of fuel to be carried<br />

within 80 %(fuselage) and 75 %(wings) volume<br />

usage tanks. All tanks in the aircraft are<br />

pressurized with nitrogen gas from the on-board inert gas generating system (OBIGGS).<br />

Pressurizing is minimal due to structural constraints and JP-8’s low vapor pressure of 0.029 psia<br />

@ 100 °F (200 Pa @ 42.3 °C). Nitrogen reduces fuel fumes and thus the chance of an accidental<br />

explosion. All tanks on the aircraft are self sealing and feature flame resistant overflow and<br />

exhaust venting.<br />

Single point fueling and de-fueling can be<br />

performed from the starboard side of the<br />

forward fuselage Figure 14.3. This fueling<br />

point shares common lines with the Air Force<br />

retractable fueling boom port (Figure 14.4)<br />

located on the upper portion of the same<br />

segment of forward fuselage. Both ports offer<br />

fueling rates as high as 1100 gpm (4164 lpm),<br />

the maximum KC-10 Fuel Probe refueling rate<br />

(Table 10.II).<br />

Fuel is then power transferred to tank three<br />

and four. From tank four it is distributed to<br />

wing tanks one and two.<br />

Figure 14.3 - Fuel System Architecture<br />

Figure 14.4 - Retractable in-flight<br />

refueling boom ports, F22, F-117, B-2<br />

92


Gravity ground refueling is provided via ports on the upper surface of the wing into tanks one<br />

and two. From there fuel can gravity feed into tanks three and four. This same gravity feed<br />

system can be utilized in flight given power feed system loss.<br />

All major thermal transfer within the aircraft is performed by the fuel system. The air cooled fuel<br />

cooler utilizing inlet duct boundary layer and flow control diversion air, placed ahead of the main<br />

gear bays dissipates all kinetic heating experienced by the airframe as well as systems heat<br />

generated. Dual heat exchangers are utilized for combat survivability. Fuel flow rate system<br />

requirements (Table 14.II) will be used to size fuel lines and pumps. All fuel lines are redundant<br />

to provide for fuel circulation and system combat survivability. The Vendetta is fitted with an<br />

onboard fire suppression system, utilizing a halon suppressant in the engine bay and fuel cutoff<br />

valves at all fuel tanks. The system is designed so it can be retrofitted with a more<br />

environmentally friendly suppression chemical in the future.<br />

Table 14.II - Fuel System Sizing Requirements<br />

Fuel System Sizing Requirement<br />

Fuel Flow Rate<br />

GPM (LPM)<br />

CG Shift Requirement Between #3 and #4 Fuel Tanks 100 (379)<br />

KC-10 Probe Maximum Refueling Rate 1100 (4164)<br />

(2) F-119 Turbofan @ Max Thrust With Reheat 294 (1113)<br />

Air Cooled Fuel Cooler (ACFC) 400 (1514)<br />

93


15 Manufacturing<br />

Several manufacturing considerations have been considered and planned for in the design of the<br />

Vendetta. One of these considerations is part commonality, which reduces the total number of<br />

separate parts that must be manufactured; thereby lowering manufacturing costs. The Vendetta<br />

is symmetrical in that both left and right wings,<br />

landing gear, horizontal and vertical stabilizers,<br />

etc. will be manufactured almost exactly the<br />

same. Furthermore, all flying surfaces are<br />

symmetrical and would provide an additional<br />

improvement in manufacturing costs.<br />

The design has also taken into consideration the<br />

routing of electrical lines, hydraulic lines, etc.<br />

These systems would be interconnected through a<br />

routing tunnel through the fuselage of the aircraft.<br />

This would reduce installation complexity and<br />

therefore reduce the amount of labor involved in<br />

the installation process. The routing tunnel as it<br />

passes beside the forward fuel tank, and aft past<br />

the weapons bay is shown below in Figure 15.1.<br />

Figure 15.1 - Routing Tunnel<br />

Manufacturing breaks include the wings, empennage, forward, center, and aft portions of the<br />

fuselage, as well as the landing gear itself. These breaks are shown in Figure 15.2.<br />

Figure 15.2 - Manufacturing Breaks<br />

94


The entire propulsion system will be capable of being dropped out the bottom of the aircraft<br />

which will provide for an easy installation during the manufacturing process as well as allowing<br />

for easy access during routine maintenance. Computer-aided manufacturing will enable more<br />

complex parts to be machined by computer-numerically-controlled (CNC) machining tools.<br />

Large items such as bulkheads can be easily machined from a single piece of metal. This would<br />

be required in order to meet the stringent structural load limits mandated by the AIAA RFP.<br />

Inspection and maintenance panels will be placed wherever possible throughout the aircraft<br />

without compromising the low-observability requirements. Furthermore, access panels will be<br />

built as “structural doors” able to carry through the skin loads that will also be required to meet<br />

the stringent structural load limits. These access panels will ease maintenance and reduce<br />

maintenance hours required per flight hour.<br />

The assembly line would allow for major components, such as the wing, fuselage, and<br />

empennage to be pre-fabricated at possibly other site locations and brought in to a central<br />

assembly line as shown below in Figure 15.3.<br />

Figure 15.3 - Assembly Line<br />

95


16 Cost Analysis<br />

The final and perhaps most important issue in the purposed development of the Vendetta is the<br />

cost analysis. The methodologies used in developing this analysis were found in the Raymer and<br />

Nicolai texts. Despite the fact that the Nicolai text was written in 1974, when adjusted for<br />

inflation, the method was accurate to within 5% of that method found in the 1999 Raymer text.<br />

Both of these analyses are adjusted for inflation to 2000 dollars. The methods used in the cost<br />

analysis were based on the DAPCA IV model developed by the RAND Corporation. This model<br />

provided a means of calculating the operating cost, life cycle cost, flyaway cost, and the cost<br />

required for research, development, test, and evaluation, or RDT&E.<br />

The RDT&E cost was predicted to be approximately $6.5 billion; whereas, the flyaway cost for a<br />

200 unit buy was calculated to be $128.5 million. This cost approximately 15% under that cost<br />

required by the AIAA RFP set at $150 million dollars per 200 unit buy. The cost per aircraft<br />

based on the number of aircraft purchased is shown below in Figure 16.1.<br />

Figure 16.1 - Cost Analysis<br />

The figure indicates that the cost per aircraft at a 600 unit buy is significantly less at $80.5<br />

million. Note the cost of engineering, development, manufacturing, and materials in the cost<br />

breakdown per unit at a 600 unit buy in comparison to the cost breakdown per unit at a 200 unit<br />

buy; the percentages associated with development and engineering decreases while the<br />

manufacturing and materials percentages increase. This is due to the fact that at a 600 unit buy,<br />

96


there are more aircraft available to help pay the $6.5 billion cost associated with RDT&E.<br />

Furthermore, there is a learning curve associated with the development of a large quantity of<br />

aircraft and the airplane become even less costly to produce.<br />

Four factors were considered when determining the operating cost of the Vendetta. These factors<br />

included the cost of the fuel and pilots, as well as the cost of parts and maintenance personal.<br />

Raymer estimates that a bomber flies approximately 400 hours per year and requires 40<br />

maintenance man hours per flight hour. In addition, because the Vendetta is designed to fly very<br />

fast at high altitudes, the fuel cost is a large percentage of the total operating cost. The operating<br />

cost of the Vendetta is calculated to be $13,000 per flight hour. This cost breakdown is shown<br />

below in Figure 16.2.<br />

Figure 16.2 - Operating Cost<br />

One final cost that must be considered beyond the cost of RDT&E, flyaway, and operations is<br />

the lifecycle cost. This cost considers the cost of RDT&E, flyaway, and operations over a 30<br />

year period at 400 flight hours per year, as well as the cost of disposal. This cost is totaled at<br />

$293 million per aircraft at a 200 unit buy. A breakdown of the lifecycle cost is shown below in<br />

Figure 16.3.<br />

Figure 16.3 - Lifecycle Cost<br />

97


Appendix<br />

Threats Chart<br />

Common Surface to Air Missiles and Threats to Airborne Aircraft<br />

98


Diffuser Efficiency<br />

Literal Factor Forms<br />

Mode Damping ratio (ζ) Natural Frequency (ω n )<br />

Phugoid<br />

ζ<br />

P<br />

−X<br />

u<br />

=<br />

2ω<br />

Zα<br />

Mq<br />

+ M α<br />

+<br />

Short Period u<br />

0<br />

ζ =<br />

SP<br />

2ω<br />

Dutch Roll<br />

ζ<br />

DR<br />

1<br />

β 0 r<br />

=− ⎜ ⎟<br />

2ωn<br />

u0<br />

DR<br />

n<br />

n<br />

⎛Y<br />

⎝<br />

p<br />

SP<br />

+ u N<br />

⎞<br />

⎠<br />

ω<br />

n<br />

DR<br />

ω<br />

n<br />

P<br />

=<br />

Z M<br />

−Zg<br />

u<br />

u<br />

α q<br />

ω<br />

n<br />

= −<br />

SP<br />

u0<br />

=<br />

0<br />

M<br />

α<br />

YN − NY+<br />

uN<br />

β β β r 0 β<br />

u<br />

0<br />

99


Wing Break Point<br />

Fuel<br />

Aft Fuselage<br />

Break Point<br />

56°<br />

14°<br />

13°<br />

Forward Fuselage<br />

Break Point<br />

54.7' [16669.5]<br />

20.7'<br />

[6294.8]<br />

5.3'<br />

[1621.3]<br />

Fuel<br />

Fuel<br />

102.8' [31337.8]<br />

5°<br />

11°


(4) Mk-84 LDGP + (2) AIM-120<br />

Weapon Weight<br />

1,967 lb (890 kg)<br />

Configuration Weight 10,222 lb (4,625 kg)<br />

Weapon Length<br />

12.6 ft (3.84m)<br />

Weapon Diameter<br />

18 in (45.7cm)<br />

Tail Span 2 ft (0.61 m)<br />

Max Drop Height<br />

Unlimited<br />

Max Tested Drop Velocity M=1.3<br />

Guidance<br />

Ballistic<br />

Weapon Information:<br />

Development of the Mk 84 Low Drag General<br />

Purpose Bomb for use by the United States armed<br />

forces began in the 1950’s. The Mk 84 bomb, which is<br />

fitted with 30 in (0.762m) spaced suspension lugs, is<br />

packed with 942 lb (426 kg) of Tritonal or H-6. The<br />

known inventory of Mk 81, 82, and 84 bombs is 1.13<br />

million.<br />

(4) GBU-27 + (2) AIM-120<br />

Weapon Weight<br />

2,165 lb (980 kg)<br />

Configuration Weight 11,014 lb (4,984 kg)<br />

Weapon Length 13.9 ft (4.24 m)<br />

Weapon Diameter<br />

14.6 in (37 cm)<br />

Tail Span 2 ft (0.61 m)<br />

Max Drop Height<br />

Unlimited<br />

Max Tested Drop Velocity Unknown<br />

Guidance<br />

Semi-Active Laser<br />

Weapon Information:<br />

The GBU-27 is a modified GBU-24 Paveway III<br />

designed for internal carriage in the F-117A. This LGB<br />

carries the designation GBU-27 /B and uses a BLU-<br />

109 /B penetrator bomb for its warhead. The main<br />

modifications made to the GBU-24 were to have<br />

shorter adaptor rings and to use the GBU-10’s rear<br />

wing unit to decrease the bomb’s length, and to clip the<br />

canards in order to make the weapon fit into the small<br />

F-117A Bomb Bay. The other major difference was the<br />

use of radar absorbing materials in order to prevent the<br />

bombs from being picked up by enemy radar once the<br />

aircraft’s bomb doors were opened. As a result of these<br />

modifications, the GBU-27 has a shorter range than the<br />

GBU-24, which can also be launched at lower<br />

altitudes.<br />

Guidance is by semi-active laser, the scanning<br />

detector assembly and laser energy receiver being<br />

mounted in the front of the canister behind the glass<br />

dome. After the bomb is released the laser error<br />

detector measures the angle between the bomb’s<br />

velocity vector and the line between the bomb and<br />

target. Steering corrections are made by moving the<br />

nose mounted canard control fins to adjust the bomb’s<br />

trajectory to line up with the target. The tail fins/wings<br />

are for stabilization purposes only. Target illumination<br />

for the system may be either by an aircraft-mounted<br />

laser marker (not necessarily the parent aircraft) or a<br />

ground-based laser transmitter.<br />

(4) 2000lb JDAM +(2) AIM-120<br />

Weapon Weight 2,100 lb (950 kg)<br />

Configuration Weight 10,754 lb (4,866 kg)<br />

Weapon Length 13.2 ft (4.02 m)<br />

Weapon Diameter 18 in (46 cm)<br />

Tail Span 2 ft (0.61 m)<br />

Max Drop Height Unlimited<br />

Max Drop Velocity M=1.3 tested<br />

Guidance<br />

GPS / INS<br />

Weapon Information:<br />

A parallel program to the AGM-154 JSOW the<br />

GBU-31 JDAM program began in the late 1980’s. The<br />

goal of the program was to produce a low cost guided<br />

munition. Interesting to note is the GBU-31 is soon to<br />

be replaced by the GBU-32/35. This new weapon, will<br />

utilize an I-1000 1000 lb (452.5 kg) penetrator warhead<br />

and is intended for future use in the F-22 raptor. This<br />

weapon, the GBU-32/35 is being used to size the<br />

raptor’s weapon bays.<br />

The GBU-31 utilizes both the Mk 84 and BLU-109<br />

warheads. Due to the Mk 84’s low cost, and<br />

commonality, it was chosen for the solid model seen<br />

above. The GBU-31 consists of three major<br />

subassemblies. The warhead (Mk 84), Saddleback stub<br />

wing assembly (attaches at hardpoints, three<br />

components), and a bolt on tail cone guidance kit.<br />

The guidance kit, contained within the replacement<br />

bolt-on tail cone consists of the following key<br />

elements: combined inertial measuring unit and GPS<br />

receiver; flight control computer; battery and power<br />

distribution unit; tail actuators and four movable<br />

clipped delta fins in a cruciform configuration. In<br />

keeping with other GPS guided weapons, the unit is<br />

believed to be fitted with two GPS antennas, one on<br />

top of the unit for initial flight and one in the tail for<br />

good reception during terminal maneuvering.<br />

Prior to bomb release the guidance unit will be fed<br />

with aircraft position, velocity and target coordinates<br />

through the aircraft to bomb interface. After release the<br />

bomb will guide itself to the target by means of rear fin<br />

deflection, which are driven by commands from an<br />

onboard computer that is constantly being updated by<br />

the GPS. The combination of the INS/GPS is expected<br />

to allow the bombs to hit within 32.8 ft (10 m) to 49.2<br />

ft ( 15 m) of their targets. Wind tunnel tests in 1996 are<br />

reported to have cleared JDAM for release at up to<br />

Mach 1.3.<br />

(4) AGM-154 JSOW + (2) AIM-120<br />

Weapon Weight 1,064 lb (481 kg)<br />

Configuration Weight 6,610 lb (2,991 kg)<br />

Weapon Length 14 ft (4.26 m)<br />

Weapon Diameter 21 in (53 cm)<br />

Tail Span 24 in (0.61 m)<br />

Max Drop Height Unlimited<br />

Max Drop Velocity Subsonic<br />

Guidance<br />

GPS / INS<br />

Weapon Information:<br />

In the late 1980’s the US Navy began a review of<br />

conventional weapons with the intention of reducing<br />

the number of weapon types. New systems were<br />

selected for future development: JDAM, TSSAM,<br />

JASSM, and the advanced interdiction weapon system<br />

to be later named Joint Standoff Weapon (JSOW).<br />

The JSOW program is intended to replace six<br />

existing weapons: the AGM-65 Maverick, AGM-123<br />

Skipper, AGM-62A Walleye, Rockeye and APAM<br />

(Anti-Personnel/Anti-Material) submunition<br />

dispensers, and laser- and TV- guided bombs.<br />

Of particular attention on the previous list is:<br />

1) All weapons are air to ground.<br />

2) This weapon is designed to replace the<br />

GBU-27, one of the weapons on the RFP<br />

attachment 3 list.<br />

The JSOW is an aerodynamically shaped,<br />

unpowered glide dispenser with a rectangular crosssection<br />

body shape. It is made up of three major<br />

sections: a streamlined nose fairing that houses the<br />

guidance and control system, a rectangular center<br />

section payload container for holding the bomblets<br />

(this is fitted with two folding high aspect ratio wings<br />

on its upper surface, and two standard 30 in (0.762 m)<br />

spaced suspension lugs); and the tail section which has<br />

six fixed, sweptback rectangular fins positioned<br />

radially on the boat tail and contains the flight control<br />

system.<br />

(16) 250 lb Small Smart Bomb<br />

Weapon Weight 250 lb (113 kg)<br />

Configuration Weight 5,500 lb (2,489 kg)<br />

Weapon Length 8.2 ft (2.5 m)<br />

Weapon Diameter 6 in (0.15 m)<br />

Max Drop Height Unlimited<br />

Max Drop Velocity Unknown<br />

Guidance<br />

GPS / INS<br />

Weapon Information:<br />

The Small Smart Bomb is a 250 lb (113 kg) weapon<br />

that has the same penetration capabilities as a 2000lb<br />

(905 kg) BLU-109, but with only 50 lbs (22.6 kg) of<br />

explosive. With the INS/GPS guidance in conjunction<br />

with differential GPS (using all 12 channel receivers,<br />

instead of only 5) corrections provided by GPS SPO<br />

Accuracy Improvement Initiative (AII) and improved<br />

Target Location Error (TLE), it can achieve a 5-8m<br />

(16.4 to 26.3 ft) CEP. The submunition, with a smart<br />

fuze, has been extensively tested against multi-layered<br />

targets by Wright Laboratory under the Hard Target<br />

Ordnance Program and Miniature Munitions<br />

Technology Program. The length to diameter ratio and<br />

nose shape are designed to optimize penetration for a<br />

50lb (22.6 kg) charge. This weapon is also a potential<br />

payload for standoff carrier vehicles such as<br />

Tomahawk, JSOW, JASSM, Conventional ICBM, etc.<br />

The Swing Wing Adapter Kit (SWAK) is added to give<br />

the SSB standoff of greater than 25 nm (48.6 km) from<br />

high altitude release. The wing kit is jettisoned at a<br />

midcourse way point if penetration is required so that<br />

velocity can be increased after wing release. For soft<br />

targets the wing kit continues to extend the glide range<br />

until small arms threat altitude is reached. At this point<br />

the wings are released. With INS/GPS guidance,<br />

coupled with AII, a 6-8 m (19.7 to 26.3 ft) CEP can be<br />

achieved. This wing kit allows the SSB to be directly<br />

attached to the aircraft at any 300 lb (135.75 kg) store<br />

station. The major advantage to the 250 lb (113.125<br />

kg) small smart bomb is an improved number of targets<br />

per pass capability.<br />

AIM-120 C AMRAAM<br />

Weapon Weight 327 lb (148 kg)<br />

Configuration Weight 5,500 lb (2,489 kg)<br />

Weapon Length 12 ft (3.657 m)<br />

Weapon Diameter 7 in (0.1778 m)<br />

Fin Span 1 ft 6 in (0.457 m)<br />

Max Drop Height Unlimited<br />

Max Drop Velocity Supersonic<br />

Guidance<br />

Command from<br />

Launch Aircraft<br />

INS<br />

Monopulse Radar<br />

Seeker<br />

Weapon Information:<br />

The Advanced Medium-Range Air to Air Missile<br />

(AMRAAM) AIM-120 development program was<br />

started in 1975. It was designed to follow on and better<br />

the performance of the Aim-7 Sparrow and be carried<br />

on the F-14, F-15, F-16 and F/A-18 aircraft. In the late<br />

90’s a modified(smaller) version of the missile, the<br />

AIM-120C was developed to be fitted to the F-22<br />

Raptor. This newer version also incorporates a dual<br />

mode active and passive radar seeker. The AIM-120C<br />

is deigned to be rail, ejector or trapeze launched. On<br />

the F-22 the AIM-120C is launched using an EDO<br />

corp. LAU-142/A hydraulic / pneumatic ejector.<br />

In a typical engagement the missile is launched and<br />

first guided by on-missile inertial navigation, with<br />

command guidance updates from the launch aircraft.<br />

The missile then goes into the mid-course autonomous<br />

mode and continues to guide by inertial navigation<br />

only. Finally, the terminal mode is automatically<br />

initiated by the missile itself when the target is within<br />

rage of the missile’s active monopulse radar seeker,<br />

which then guides the missile onto the target aircraft.<br />

Foldout 2 – Weapons Information


The Vendetta Design Team<br />

Chris Atkinson is a bachelor’s candidate in aerospace engineering<br />

with a computer science minor. His responsibilities for the Vendetta<br />

included aerodynamic and performance analysis as well as validation<br />

modeling. He participates in the <strong>Cal</strong> <strong>Poly</strong> Flight Simulation Group and<br />

enjoys fencing and hiking.<br />

Chris Droney is a blended program master’s candidate in aerospace<br />

engineering. Chris was the team leader and lead configurator for the<br />

Vendetta. He designs, builds, and flies R/C planes and gliders. Chris also<br />

has his private pilot’s license and enjoys flying when he gets the chance.<br />

Kolby Keiser is a 23 year old master’s candidate in aerospace<br />

engineering. Her responsibility in the Vendetta design group was aircraft<br />

propulsion. In her spare time, Kolby enjoys hiking, exercising, and other<br />

outdoor activities, in addition to swing dancing. She also participates in<br />

diabetes education in the community.<br />

105


Chris Maglio is a bachelor’s candidate in aerospace engineering. His<br />

responsibilities in the Vendetta design group encompassed configuration,<br />

weapons, and systems. Chris is an active member of Central Coast Polo<br />

and the USPA. Besides pursuing the hobby of mountain biking<br />

throughout the academic year, Chris, working as a consultant, has<br />

designed and manufactured numerous nautical, aerospace, civil, and<br />

automotive products.<br />

Dan Salluce is a 23 year old blended program master’s candidate in<br />

aerospace engineering. His responsibilities in the Vendetta design<br />

included stability and control analysis, radar cross section determination,<br />

and flight simulator validation. Dan is interested in control system<br />

design. He participates in the <strong>Cal</strong> <strong>Poly</strong> Flight Simulation Group, teaches<br />

Aerospace classes at <strong>Cal</strong> <strong>Poly</strong>, and enjoys R/C gliders and the <strong>San</strong> <strong>Luis</strong><br />

<strong>Obispo</strong> nightlife and nearby beaches.<br />

Nathan Schnaible is a 5 th year aerospace engineering student who<br />

will graduate with a bachelor’s degree in June of 2002. Nathan worked<br />

on the weights & balances, manufacturing, and cost analysis areas of the<br />

Vendetta. His background in the aerospace industry includes<br />

maintenance duties on refurbished Albatross’, ground service on both<br />

corporate and commercial aircraft, as well as piloting experience.<br />

Nathan will be attending the United States Navy Flight School in the<br />

spring of 2003.<br />

106


References<br />

1. “Ejection Systems,” www.bfg-aerospace.com, BF Goodrich Corporation, 2001.<br />

2. “Military Standard – Aircrew Station Controls and Dispalys: Location, Arrangement, and<br />

Actuaion of, for Fixed Wing Aircraft,” United States Department of Defense, 1991.<br />

3. “Military Standard – Aircrew Station Geometry for Military Aircraft,” United States<br />

Department of Defense, 1976.<br />

4. Abbott, I. H., Von Doenhoff, A. E. Theory of Wing Sections, Dover Publications, INC.<br />

New York 1959.<br />

5. Cummings D. Boeing Long Beach, Advanced Aircraft Design<br />

6. Currey, N. S. Aircraft Landing Gear Design: Principles and Practices, AIAA,<br />

Washington DC, 1988.<br />

7. Dillenius, M. F. E. Perkins, S. C., Nixon, D., “Pylon Carriage and Separation of Stores,”<br />

Tactical Missile Aerodynamics: General Topics, Edited by Michael J. Hemsch, Vol. 141,<br />

Progress in Aeronautics and Astronautics, AIAA, New York, 1992, pp. 575-666.<br />

8. Goodall, J. C. Americas Stealth Fighters and Bombers, Motorbooks International<br />

Publishers and Wholesalers, Osceola, WI, 1992<br />

9. Jane’s All The World’s Aircraft 2000-2001, Janes Information Group Inc. Alexandria,<br />

Virginia, 2000<br />

10. Jane’s Avionics 2001-2002. Janes Information Group Inc. Alexandria, Virginia, 2001<br />

11. Lennox D. Jane’s Air-Launched Weapons Issue 35. Janes Information Group Inc.<br />

Alexandria, Virginia, 2000<br />

12. Mattingly, Jack D., Elements of Gas Turbine Propulsion, McGraw-Hill, Inc. New York,<br />

NY, 1996.<br />

13. MIL-A-8860B<br />

14. MIL-A-8861B<br />

15. MIL-STD-850B<br />

16. MIL-E-5008B<br />

107


17. NACA-TN-3182, “Manual of the ICAO Standard Atmosphere <strong>Cal</strong>culations by the<br />

NACA”, NASA, 1976<br />

18. Nicolai, L. M. Fundamentals of Aircraft Design, METS Inc., <strong>Cal</strong>ifornia, 1984.<br />

19. Oates, G. C. Aircraft Propulsion Systems Technology and Design, AIAA, Washington<br />

DC, 1989.<br />

20. Raymer, D. P. Aircraft Design: A Conceptual Approach – Third Edition, AIAA,<br />

Washington DC, 1999.<br />

21. Roskam, J. Airplane Design, Part I: Preliminary Sizing of Airplanes, DARcorporation,<br />

Kansas, 1997.<br />

22. Roskam, J. Airplane Design, Part II: Preliminary Preliminary Configuration Design and<br />

Integration of the Propulsion System, DARcorporation, Kansas, 1997.<br />

23. Roskam, J. Airplane Design: Part III, Roskam Aviation And Engineering Corporation,<br />

Ottawa, KS, 1989, pp 1-34.<br />

24. Roskam, J. Airplane Flight Dynamics and Automatic Flight Controls, DARcorporation,<br />

Kansas, 1979.<br />

25. Wilcox, F. J., Baysal, O., Stallings, R. L., “Tangential, Semisubmerged, and Internal<br />

Store Carriage and Separation,” Tactical Missile Aerodynamics: General Topics, Edited<br />

by Michael J. Hemsch, Vol. 141, Progress in Aeronautics and Astronautics, AIAA, New<br />

York, 1992, pp. 667-721.<br />

26. www.aeronautics.ru/nws002/f22/diagram05.jpg<br />

27. www.aeronautics.ru/nws002/f22/diagram06.jpg<br />

28. www.aeronautics.ru/nws002/f22/systems.htm<br />

29. www.af.mil/news/efreedom/bombs.html<br />

30. www.af.mil/news/factsheets/KC_10A_Extender.html<br />

31. www.af.mil/news/factsheets/KC_135_stratotanker.html<br />

32. www.arfl.afr.mil<br />

33. www.aoe.vt.edu/aoe/faculty/Mason_f/M96SC.html<br />

34. www.batnet.com/mfwright/spacesuit.html<br />

108


35. www.dfrc.nasa.gov/PAO/PAIS/HTML/FS-061-DFRC.html<br />

36. www.eureka.findlay.co.uk/archive_features/Arch_Automotive/n-push/n-push.html<br />

37. www.fas.org/man/dod-101/sys/ac/equip/lau-142.htm<br />

38. www.fas.org/man/dod-101/sys/ac/equip/lau-142.htm<br />

39. www.fas.org/man/dod-101/sys/missle/amraam-5.jpg<br />

40. www.fas.org/man/dod-101/sys/smart/agm-154.htm<br />

41. www.fas.org/man/dod-101/usaf/docs/mast/annex_f/part06.htm<br />

42. www.globalsecurity.org/military/systems/aircraft/f-22-fcas.htm<br />

43. www.sff.net/people/geoffrey.landis/vacuum.html<br />

44. www.skf-linear.co.il<br />

109

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!