AIAA 2019 2020 Undergraduate Team RFP Hi
AIAA 2019 2020 Undergraduate Team RFP Hi
Weizhuo 981656
Wang
Jack Wu 1109057
i
Contents XI.B Takeoff & Landing Performance 33
XI.C Other Performance Parameters 35
I The Team . . . . . . . . . . . . . . . . i XI.D Drag for All Flight Segments 35
XI.E Aircraft Performance Coeffi-
List of Figures . . . . . . . . . . . . . . . . . . iii cients . . . . . . . . . . . . 36
XI.F Payload-Range & Range
List of Tables . . . . . . . . . . . . . . . . . . iv
Mach Diagram . . . . . . . 38
XI.G Flight Envelope Diagram . . 38
II Nomenclature . . . . . . . . . . . . . . iv
XI.H Specific Excess Power Diagram 39
III Acronyms . . . . . . . . . . . . . . . . v XI.I Trade Studies . . . . . . . . 40
XI.I.1 Cruise condition trade study 40
IV Executive Summary . . . . . . . . . . . 1 XI.I.2 Step climb trade study . . . . 41
XI.J Fuel Requirement . . . . . . 42
V Introduction . . . . . . . . . . . . . . . 1 XI.K Performance Conclusion . . 43
II. Nomenclature
III. Acronyms
V. Introduction
The transition of global airline markets calls for an aircraft with a short-range high capacity mission profile. As
determined by the AIAA RFP, a 400 passenger aircraft with 3,500 nmi range (optimized for 700 nmi) is to be designed.
The RFP also stated that the design objectives are to minimize operating/ production cost and ensure high aircraft
reliability/ maintenance.
The target group of the JJJP is low-cost airliners in emerging markets. Primarily those that aim to maximize flights
across short-distance regional flights. To survive in this niche of the airline industry, there is an emphasis on minimizing
costs and maximizing passengers moved. As such, the JJJP has been designed with a focus on cost from the very
beginning. The JJJP contains design traits that give it an unparalleled efficiency at a competitive price point. The
Rolls-Royce UltraFan engines and large wingspan provide maximum fuel economy. The aluminum airframe allows for
rapid manufacturing and maintenance. Onboard auxillary systems and avionics ensure safety during all phases of flight.
The interior cabin is arranged to be volume efficient for a high passenger count. And finally, the folding wingtip systems
1
gives the JJJP the flexibility to access Group IV airports, maximizing it’s global and local access.
The JJJP is capable of carrying a 400 passenger payload a range of 3,500 nmi. Cruise speed and altitude are Mach
0.78 and 37,000 ft respectively. It has a 184 ft wingspan (171 ft folded) and 203 ft fuselage length. The MTOW and
empty weight are 443,000 lb and 241,500 lb respectively. It utilizes two Rolls-Royce UltraFan engines with a maximum
thrust of 62,500 lbf. The entry into service is expected to be 2029 at a unit cost of $ 180 million. A CAD model of the
are the largest influencers of the aircraft design. One further requirement that was chosen to be incorporated was sizing
the aircraft to fit into Group IV airports. This was to allow for increased access to regional airports.
flight attendants)
Takeoff Maximum 9,000 ft over a 35 ft ob- Dry pavement, sea level ISA + 15 °C
stacle at MTOW
flight altitude
autopilot
ICAO/EASA regulations
2
Optimal and divert mission profiles are shown in Figure 3. There are straightforward segments such as taxi/takeoff,
climb, cruise, descent, and landing/taxi. There is also the necessary capability to loiter for a certain range. This is
specified as a 200 nmi range to an alternate airport, a 30 minute hold, and a 5% contingency fuel.
A. Similarity Analysis
When searching for an aircraft to use as a seed for this design, the main parameters that were taken into consideration
were number of passengers and range. These two requirements have a significant impact on the overall size and
performance needed for the aircraft. Furthermore, modern aircraft were chosen over older ones to reflect the technology
that would be present in this design. From these criteria, the Boeing 777-200 and the Boeing 787-8 were used as the
two seed aircraft when designing the JJJP. A list of all other important parameters from the seed aircraft can be seen
For the aircraft that is being designed, a minimum of 400 passengers needs to be carried over a range of 3,500 nmi.
3
From Table 3 the Boeing 777-200 more closely aligns with these specifications. However, both seeds were used to do
the initial sizing process to verify that the dimensions would converge to similar values and that the sizing program is
working properly.
In order to quickly test out the effect of various parameters, a MATLAB sizing program was developed to conduct
an iterative sizing analysis, pragmatically. Specifically, the program first took in the requirements, seed parameters, and
starting guesses of the design parameters. It then applied the sizing equations from Raymer [6] and Roskam [7] to
The program repeated the iterations for different parameters such as cruise mach number, altitude, wing area, aspect
ratio, etc. Since these parameters were all varied across sizing iterations, it was critical to visualize the interactions
between the parameters. Therefore the program plotted a contour, allowing decisions to be made on the optimal
The constraints shown in Table 4 define the design space. These constraints were derived from the requirements
Parameter Value
Loiter 30 min
The constraints listed presented a valid design region in the constraint diagram in Fig. 4. The diagram is limited by
maximum cruise speed, takeoff distances, landing distances, and cruise limitations. The maximum cruise speed used is
mach 0.9. The blue cross denotes the designed point of JJJP. Specifically at maximum takeoff weight, the thrust to
weight ratio is 0.282, with wing loading at 114 psf. At this point, lowest thrust is required and penalties from takeoff
thrust and extra weight from large wing area are both minimized.
4
Fig. 4 Constraint diagram
C. Trade Studies
A set of trade studies were conducted to determine an optimal design. The qualifications for success in these studies
were (1) minimize fuel weight to reduce operating cost and (2) minimize empty weight to reduce manufacturing cost.
To determine fixed geometry parameters and flight conditions of the aircraft that optimize for the aforementioned
qualifications of success, the following set of values were varied: wing aspect ratio, wing area, wing sweep, wing taper,
By utilizing rough estimates, a lifetime cost model was incorporated into the trade studies. An empty weight to unit
cost estimation of $850 per pound and fuel cost to fuel weight of $1.12 per pound were added to the sizing code outputs.
After rerunning the sizing code, with this cost model implemented, the contour in Fig. 5 was produced. This contour
5
Fig. 5 Unit + fuel cost (USD millions) of AR vs. wing area
A key takeaway from this analysis was that rather than having a 171 ft fixed wingspan, it would be more economically
viable to have a 184 ft wingspan with wingtips that fold to 171 ft. This is assuming that the folding wingtip system
would weigh under approximately 2,000 lb. Based upon further trade studies for different parameters, final optimal
Parameter Value
6
The optimized parameters from Table 5 were input into the sizing program, and the aircraft model was then
lb
SFC 0.43 lb-hr Fuel Weight 110,000 lb
VIII. Configuration
A. Design Morphology
The aircraft configuration was originally designed in an effort to optimize for the design objectives specified in the
AIAA RFP [1]. Principally, the objective of optimizing cost for a 700 nmi range mission was closely considered. To
that end, considerations were made for FAA Advisory Circular AC 150/5300-13A, wherein specifications are listed for
geometric layout of runways, taxiways, aprons, and other facilities at civil airports [8]. The document specifies airport
design groups and their corresponding aircraft classes. Each design group is categorized by both the range of wingspans
and maximum tail height that the airport can accommodate. For instance, there are approximately one dozen airports
in the world that are built to accommodate aircraft within the specifications of “Group VI” [9]. A comprehensive
7
Table 7 Airport Classifications for per FAA and ICAO
Group Wing Span Tail Height Code Wing Span Outer Main Gear Span
Qualitative analysis was performed on alternative configurations, including a twin-fuselage design and a double-
decker design. A twin-fuselage design initially appeared enticing from a cost standpoint, as an existing fuselage
configuration (e.g. Boeing 737-800) could simply be duplicated, with a wing extending between the fuselages. Such a
design, similar to Scaled Composites’ Stratolaunch aircraft [10], could potentially save on both research and design
costs, as well as manufacturing costs via existing tooling and production facilities. The design was ultimately abandoned
due to the accelerations that would be experienced by passengers during turns, as a result of centripetal force. Additional
concerns were noted in the required structures for a wing spanning between the two fuselages.
The double-decker design was abandoned due to anticipated costs associated with the structural components required
to support the design and general manufacturing and transportation costs. In the same way that Airbus is required to
produce a380 components at multiple factories across Europe due to the size of the components, any double-decker
aircraft design may require similar manufacturing logistics. Similar to the Airbus a380, such a design would have large
sections (e.g. fuselage, wings, etc.) and traditional component transportation methods would not be feasible [11].
After the initial configuration, a trade study was performed to analyze potentially novel applications of technology
that would allow the aircraft to perform competitively in its class. The qualifications for the trade study were that any
configuration change would need to both further optimize cost for a 700 nmi mission and be realistically implementable
by the EIS date. Table 8 provides details on technologies that were studied.
8
Table 8 Innovative Technology Trade Study Concepts
While all of the studied technologies had the potential to improve performance, it was ultimately decided that the
Technology Readiness Level of the window-less configuration was too low to pursue at this time. Experts also believe
that passengers simply prefer window views to digital displays [13]. In contrast, despite having a very high Technology
Readiness Level, additive manufacturing was determined to be too expensive for this design, violating the qualifications
of this trade study. The folding wingtips concept was accepted as both technologically realistic and expected to provide
cost savings. This decision drove a wingspan configuration change from 171 ft to 184 ft. The consequence of this
decision was a new wing design with a larger wing span, larger aspect ratio, and larger wing area–improving in-flight
performance via lowering drag– while still allowing the aircraft to fly into Group IV airports. Analysis is presented in
Section XVI to verify that the added weight of the larger wing and folding system do not result in greater performance
1. Modeling
The original design, created in OpenVSP, consisted of a traditional low-wing single fuselage design, as shown in
Fig. 6. This design did not offer significant consideration to the nose cone or tail cone. Subsequently, they were left in
9
The design was later migrated into the Siemens NX environment and refined to incorporate a nose and tail cone that
more closely resembled those of existing high capacity aircraft, as shown in Fig. 7. This choice was made assuming that
such a configuration would likely have had previous analysis performed around it, acting as a good starting point for
additional refinement. Further consideration was paid to fitting internal structures in addition to simply allowing room
for the requisite seating. Higher fidelity details such as control surfaces and high lift devices (ailerons, flaps, rudder,
and elevators) were added to the model with the intention of undergoing further geometry changes as analysis was
performed. Details such as windows, doors, and landing gear were not yet included in the model.
As a result of in-depth analysis, presented throughout the rest of this report, in areas such as Stability and Control,
Aerodynamics, Structures and Loads, etc., the configuration was further refined to include additional details such as
a yehudi, landing gear, sized control surfaces and high lift devices, and other details (ref. Fig. 8). A 3-view and
engineering drawing of the current design can be found on page 11 and 12, respectively.
10
11
shown on page 11. Note: details on the landing gear can be found in Section XV.A. Key configuration parameters are
presented in Table 9.
Parameter Value
𝐶
4 Sweep 25 deg
Taper .15
Pilot viewing angles are a necessary part of the design and regulatory process for development of any aircraft
because of the need for the pilot to see outside to operate the aircraft. This has been taken into account in the design of
the JJJP. More specifically, this is proved by meeting the Advisory Circular 25.773-1 that clarifies 14 CFR §25.773 [14].
For simplification, the minimum angles required for viewing in the vertical and horizontal direction will be used. If the
JJJP meets the minimum required angles, then it should be certifiable in terms of the regulations. Another important
note is that although this model uses a human height of 5’8", with seat adjustments any height within the range specified
by the FAA could pass the viewing angle requirements. Table 10 shows the minimum viewing angles needed, and Fig
Pilot Above Below Left of Cen- Right of Left Unob- Right Un-
13
(a) Vertical viewing angles (b) Horizontal viewing angles
The interior design of this aircraft was driven by three main objectives: Meeting the RFP requirements for passenger
counts, meeting FAA safety and evacuation requirements, and making sure the passenger experience is positive.
The RFP had a few specific requirements that had to be complied with in the design of the interior. There were to be
at least 50 business class seats with a pitch of 36 in and a width of 21 in .There were also to be at least 350 economy seat
with a pitch of 32 in and a width of 18 in [1]. These were all complied with according to Fig. 12 . The decision was
made to stay at the minimum seat dimension values outlined by the RFP because these values are more in line with the
(a) Business class seat specifications (b) Economy class seat specifications
The next step was to fulfill the requirements from the FAA on the interior of the aircraft. Mainly, these requirements
are 14 CFR §25.815 for required aisle width, and 14 CFR §25.803 with clarifying regulations 14 CFR §25.807 and 14
CFR §121.291 for evacuation of the aircraft. Looking again at Fig. 12, it can be seen that the aisle width for business
class is 15 in on the ground and 27 in above the arm rest. For the economy class, the aisle is 16 in on the ground and 20
14
in. This meets the minimum requirements according to 14 CFR §25.815. For evacuation regulations, the focus was
mostly put on the type of exits needed in order to evacuate under the required 90 seconds with the consideration of a
possibility of fire on part of the aircraft. The JJJP employs three Class A exits per side and one Class B per side. Per 14
CFR §25.807, Class A exits can evacuate up to 110 passengers, and Class B exits can evacuate up to 75 passengers.
This means in total the airplane should be certifiable to evacuate in total 810 passengers or 405 passengers on one side
The layout of our interior seating is a 2-3-2 configuration for business class, and a 3-4-3 for the economy class.
This was done in order to balance passenger comfort, passenger count, and meeting requirements. Based on qualitative
research on seat configurations of different aircraft and airline companies, it was found that both configurations were one
of the most popular for their respective classes. This analysis, along with these classes being efficient at getting more
seats into a given area, makes this setup the best choice for the JJJP. An image of the overall seating configuration that
could be used for seat marketing is shown in Fig. 13. Please note that this image is not completely sized dimensionally,
All dimensions for the interior of the aircraft are shown below in Fig. 14 in inches. Not shown in the figure are two
doors, two bathrooms, two jump seats, and one galley space near the front of the aircraft. These will be contained in the
nose section of the aircraft behind the cockpit, so they were not included in the cabin drawing. Also not shown on Fig.
14 is the length of the baggage compartments. The forward baggage compartment has a length of 62 ft, and the aft
baggage compartment has a length of 64.5 ft with the last 12 ft tapering down in the tail section. For reference, this
corresponds to a total of 24 LD3 containers in the forward baggage compartment, and 20 LD3 containers in the aft
baggage compartment. LD3 containers are common containers used in the airline industry for baggage. These containers
have a maximum width of 79 in and a height of 64 in [15]. With the average width in the baggage compartment
being 170 in, there is more than enough room to house two rows of LD3 containers in the forward and aft baggage
compartments.
15
Fig. 14 2D view of seating arrangement and barrel section dimensioning
IX. Propulsion
A. Design Approach
The team’s design philosophy utilizes a rubber engine approach. The engine parameters were scaled from the seed
parameters according to varying thrust requirements. Specifically, parameters such as fan diameter, nacelle diameter,
integrated power plant system weight, engine length, and nacelle wetted area were scaled. According to the sizing result
discussed in section VII, the required total thrust was 125,000 lbf at sea level.
B. Trade Studies
At 125,000 lbf thrust requirement, turboprop technology would struggle to meet performance without great expense
in weight and subsequent cost. Even with the most powerful [16] turboprop in production (Europrop TP400), the design
still needs at least six engines. Electric propulsion requires carrying a very heavy battery that would not reduce weight
in flight. According to Raymer [6], jet fuel has a energy density 11000 Wh/kg , while the battery has at most 600 Wh/kg.
16
This means the equivalent battery would weigh over 2 million pounds. This factor alone determines that electric aircraft
is infeasible for this type of mission. This analysis narrows down the engine choice to turbofan or turbojet, provided the
The design of the propulsion system is mainly concerned with factors such as fuel economy, range requirements,
and cruise speed. The JJJP aims to maximize fuel economy while achieving the required range and capacity, as the
demand for short-haul routes can be highly sensitive to ticket prices [17]. Considering that the aircraft will be cruising
at transonic speeds, turbofan technology is a better choice than turbojet technology, since turbofan has a slower exhaust
velocity and higher mass flow rate. Specifically, the highest overall efficiency is achieved when exhaust velocity equals
two times the cruise velocity[6], which is closest to turbofan’s exhaust speed. The high mass flow rate of turbofan also
the engine.
Four engine configurations bring extra complexity and weight to the aircraft and do not provide any real gain in fuel
economy. Three engines configuration has inherent risks due to the mid-engine placement. When the failure occurs to
the mid engine, it is often catastrophic (AA191 O’Hare 1979 [18]). It was then decided to use two engines to power the
aircraft, each providing 62,500 pounds of thrust to meet the total thrust requirement.
C. Safety Considerations
In coordination with performance analyses, with one engine inoperative (OEI), the aircraft can meet the takeoff and
landing requirements set in the RFP (9000 ft) [1]. In such a condition, the pilot should immediately abort takeoff and
apply the brake if the speed is under V1 calculated in the performance section (Sec. XII). On the other hand, if the
speed is above V1, the pilot should continue to take off with only one engine. In both cases, the pilot is guaranteed to
To illustrate the engine operation over the designed mission, a plot of Throttle vs Time is provided below in Fig 15.
The orange line denotes cruise with step climb, while blue denotes same mission without step climb. This plot includes
the 3500 nmi primary mission and 200 nmi diversion. Specifically, primary mission ends at 28000 s as shown in Fig 15.
17
Fig. 15 Throttle used throughout the flight
From Fig 15, the JJJP is designed to cruise at 78%-90% throttle depending on the cruise condition. If step climb
is used, since the aircraft will be cruise at a higher altitude with thinner air, the throttle is maintained around 90%
throughout the cruise with short periods of 100% in order to climb to a higher altitude. A detailed analysis of the
E. Engine Selection
18
To summarize, an optimal engine should have low SFC, low weight, and at least 62,500 lb of thrust. With these
factors in mind, six current generation engine families and one in development engine were found to fulfill all the
requirements. Some key parameters for those engines are listed below in Table 11. GE90, GE9X, and Trent XWB
are slightly overpowered for this aircraft. However, they are worth mentioning in case more payload is desired, or the
need arises to cover high altitude airports. Also, it should be noted that each family of the engine comes with multiple
models. Therefore some parameters are shown as a range of values in table. For the next generation engine UltraFan,
the parameters are estimated based on the frozen design provided by Rolls-Royce. A CAD model was made according
to the parameters and pictures released by Rolls-Royce. Figure 16 is a rendering of the CAD model. The estimations of
UltraFan below are based on worst case parameters to provide a margin for JJJP design in case the supplier fails to meet
the target.
Engine Family Application Takeoff Thrust (dry) [lbf] SFC [lbf/lb-hr] Weight (dry) [lb]
The next-generation engine surpasses the current generation on almost all parameters, even with the most conservative
estimations. The engine is also scalable from 25,000 lb of thrust to 100,000 lb of thrust, meaning that there are much
more flexibility in the design process. RR UltraFan is chosen as it provides an unparalleled advantage on both fuel
F. Engine Characteristics
Since the UltraFan is not currently available, the performance characteristics of the engine need to be estimated.
Applying Raymer [6] equation of SFC vs. altitude vs. Mach on the known parameters of the UltraFan gives the following
characteristics (Fig 17). As altitude increases, specific fuel consumption decreases. And as the speed of the aircraft
increases, the SFC increases as less thrust is produced with same fuel flow rate.
19
Fig. 17 Raymer SFC estimation
Apart from mach number and altitude, SFC is also dependent on the throttle level. The engine characteristic model
took into account the throttle level too according to the method provided in Artur Bensel’s paper [25]. Figure 18 is
produced assuming cruise at M=0.8. As illustrated in the plot, thrust available decreases with altitude and the optimal
20
SFC occurs around 70% throttle.
The thrust available from each engine vs. altitude relation is estimated from the pressure ratio between the cruising
altitude and the sea level. The relation is shown in Fig. 19. From the plot, at cruise altitude each engine can produce
The inlet is designed to tilt downward 4 degrees to compensate for the cruising angle of attack and increase the inlet
capture area at cruise condition. At cruise condition, the inlet capture area is 90.5 ft2 , and the engine inlet area is 65.5
ft𝑡 2 . At optimal cruise condition M=0.78, this area ratio can decelerate the air to M=0.45 and convert 86.8% of the total
pressure to static pressure. According to Raymer equations [6], the inlet pressure recovery is estimated to be 99.15%.
Figure 20 shows the cross section of the engine. The nacelle cross-section is designed to be like an airfoil to
minimize drag. Located inside the nacelle are accessory gearbox, power management systems, and fuel delivery systems.
The nacelle can be opened at the end of the fan cowling to provide reverse thrust upon landing.
21
Fig. 20 Inlet, nacelle, and exhaust
The engine has two exhausts in total. One is bypass channel exhaust, and the other is combustion chamber exhaust.
Similar to current generation design, the hot core exhaust will be gradually mixed with cold bypass exhaust through a
chevron-shaped exhaust to reduce noise (Fig. 21). The routing of the exhaust is shown in the CAD cross-section (Fig
20).
Since typical thrust losses due to installation are 4% for low bypass and 8% for high bypass ratio engine ([26]) , the
installed thrust loss for UltraFan is estimated to be 10% since this engine is an ultra high bypass ratio engine (15:1).
With installed thrust requirement stands at 62,500 lbf, the uninstalled thrust required is estimated to be 69,500 lbf.
22
H. System Perspectives
1. Engine System
In general, engines are placed below the wing like majority of the commercial aircrafts, for both safety consideration
and structural reasons. Refer to Auxiliary Systems section (Sec. XVIII.B) for the details on FADEC setup and system
architecture.
2. Fuel System
The detail of the fuel system is further covered in the Auxiliary Systems section (Sec.XVII.C). The total amount of
fuel required for JJJP on a 3,500 nmi mission is 83,610 lb, according to the performance calculation (Sec.XII.J). The
X. Aerodynamics
A. Airfoil Selection
Airfoil selection was performed using the vortex-lattice software XFLR5. The software is useful as a means to
quickly compare characteristics of similar airfoils. However, it does not analyze airfoils in the transonic regime. When
cruising in the transonic regime, air flow on top of the airfoil will become supersonic and generate shock waves. This
is not desirable for the aircraft because upper-surface shock waves will cause loss of lift and increased drag. At the
same time, the shift in center of pressure will change pitching moment. For a swept wing, loss of lift will result in
sudden pitch down moment or "Mach tuck". Supercritical airfoils are design to minimize these effects by having a more
uniform thickness distribution. They move the camber location to the trailing edge to delay shock wave and turn strong
shock waves into weak shock waves [6]. SC-0412, SC-0714, RAE-5215, and DFVLR R-4 were chosen for potential
airfoil candidates [27]. They were analyzed at a fixed Reynolds number to get a sense of their performance.
23
(a) Lift slope comparison (b) Drag polar comparison
Figure 22 shows the aerodynamic performance of two high cambered airfoils (R-4 and SC-0714) and two relatively
low cambered airfoils (SC-0412 and RAE-5215). High cambered airfoils offer more lift and 𝐿/𝐷 compared to low
cambered airfoil but suffer from a high moment coefficient, as shown in Fig 22c. For example, the SC-0714 has a
𝐿/𝐷 = 66.3 at 𝛼 = 1.25, but the moment coefficient is 69.86% higher than the SC-0412. This contrast has a direct
impact on stability and control (e.g. increasing tail size to counter the pitching moment). As shown in the drag polars in
Fig 22b, performance of all four airfoils were similar in the non-stall region.
OpenVSP was used to do high speed trade studies for airfoil selection. Since OpenVSP cannot analyze 2D airfoils, a
24
(a) Lift slope comparison (b) Drag polar comparison
Figure 23 shows aerodynamics characteristics of the 3D wing using OpenVSP. Despite the general trend of the
aerodynamic performance curves matching the results from XFLR5, the lift slope is higher compared to XFLR5. The
Moment coefficient diagram and lift-to-drag ratios are consistent with XFLR5, where the low camber airfoils have
lower pitching moments and the lift-to-drag curves shift to the right. One of the limitation of OpenVSP however is the
Because of this analysis, the selection was narrowed down to the SC-0412 and RAE-5215. Historical trends provided
by Raymer in Chapter 4 of Aircraft Design : A Conceptual Approach suggest that a thickness to chord ratio (𝑡/𝑐)
of around 13% is appropriate for transonic aircraft [6], which makes the SC-0412 a more favourable pick than the
RAE-5215. Although a thinner airfoil will minimize drag and especially wave drag, it hinders other systems such as fuel
tank size, structural integrity, and high-lift system. Although the SC-0412 is lacking in terms of aerodynamic efficiency,
it was a well rounded airfoil. Thus it was chosen to be the airfoil that best met the necessary performance requirements
of the JJJP. The SC-0412 geometry is shown in Fig. 24. For tail airfoil selection, see section XII.B.
25
Fig. 24 NASA SC(2)-0412 Airfoil, max thickness of 0.12% at 37% chord, max camber of 0.13% at 83% chord
B. Wing Design
Wing parameters were determined by multiple trade studies in the initial sizing phase. Those studies examined
parameters such as wing area, aspect ratio, and taper. Other parameters were chosen based on historical data. Wing
twist is a solution for improved stall performance, so the outboard section of the wing doesn’t stall first. For geometric
twist, a typical -3 deg of twist will provide adequate stall characteristics suggested by Raymer [6]. For wing dihedral, a
low wing transonic aircraft will typically have three to seven degrees of dihedral [6]. The JJJP dihedral was determined
to be 6 deg based on roll stability calculations and engine ground clearance. Wing sweep reduced wave drag due to
reduction of chordwise flow and shock wave at an oblique angle on the wing.
Since the aircraft is designed for short haul high capacity missions, the cruise speed impact on travel time has less of
an effect than longer range missions. Initial sizing trade studies concluded that reducing the wing sweep from the seed
aircraft’s 31 degrees to 25 degrees would be a more optimal due to the lower cruise speed of Mach 0.78 [28]. A lower
wing sweep will also lower the weight of the wing. Wing sweep reduces the 𝑀𝑑𝑑 by a factor of inverse square root of
cosine of sweep angle according to Equation 4.15 in Roskam [29]. The Boeing 777-200 cruises at Mach 0.84 [27] with
a 31 degree sweep. A quick calculation shows, using Equation 4.15 with the same wing, a reduction of six degrees
of sweep will lower 𝑀𝑑𝑑 to 0.817. Reduction of Δ 𝑀𝑑𝑑 = 0.025 is an acceptable compromise for the benefit of less
A further investigation on this trade study was conducted using Delta method. Wing sweep angle for commercial
aircraft are typically around 31 deg. However, sweeping beyound 25 deg will see diminishing return due to its minimum
wave drag reduction. Using Delta method, 2 count of wave drag reduction was predicted going from 25 deg to 31 deg
wing sweep.
26
Table 12 Wing Parameters
Parameter Value
Wingspan 183.7 ft
𝐶
4 Sweep 25 deg
Taper .15
Dihedral 6 deg
For steady level flight at Mach 0.78 at 38,000 feet, a lift coefficient of 0.44 was calculated via the initial sizing code.
OpenVSPAERO, a fast vortex lattice solver, was used to do aerodynamic analysis on the wing for cruise condition. The
lift curve was determined by scaling the OpenVSP result at Mach 0 with Raymer’s prediction of transonic lift slope. To
obtain the scaling factor, simply use the typical swept wing high aspect ratio curve in Raymer Fig. 12.6 [6], and take the
ratio of lift slope at Mach 0.78 to lift slope at Mach 0. With the wing cruise lift curve, incidence angle was determined
to be 3 deg because the wing will satisfy the lift requirement at that angle of attack. 𝐶 𝐿 and 𝐿/𝐷 were plotted against 𝛼
in Fig 25a and Fig 25c respectively, and the mission points are indicated with black star. 𝐶 𝐿 = 0.44 at 𝛼 = 0° in Fig 25a
satisfied the cruise lift requirement. Additionally, 𝐿/𝐷 = 15.714 at 𝛼 = 0° is within a reasonable range when compared
to aircraft at this size. For example, the Boeing 747-400 has cruise 𝐿/𝐷 of 15.5 [30].
27
C. Aircraft Aerodynamic Characteristics
(a) Aircraft cruise Lift Curve Slope Graph (b) Aircraft cruise drag polar
For aircraft aerodynamics characteristics at cruise, three diagrams that showcases lift curve, drag polar, and L/D
curve are shown in Fig. 25. Cruise lift curve and L/D cruise were determined by shifting the wing’s polar to the left by
3 deg.
For takeoff/landing analysis, XFLR5 was used as a mean to construct aircraft polars. To use XFLR5, the airfoil
needs to be analyzed at a Reynold’s number from 8 million to 55 million (From tip chord to root chord) at Mach 0.202
for landing and Mach 0.23 for takeoff. Both analyses used an inviscid Vortex lattice method for flapped wings. Since
XFLR5 can only model plain flaps and slats, highlift device modeled for the JJJP closely resemble those of the Boeing
777 [31]. As shown in Fig. 27, landing flaps were set to 35 deg for leading edge flap and 50 deg for plain flap. 24 deg
and 34 deg was set respectively for takeoff. Key mission points are marked by the black star.
28
(a) Aircraft takeoff/landing Lift Curve Slope Graph (b) Aircraft takeoff/landing drag polar
The wing was sized by maximum lift coefficient as determined by performance. 𝐶 𝐿𝑚𝑎𝑥 based upon the design
requirement is around 2.49, so the goal is to size the high lift devices to meet that. Leading edge slats and double slotted
Fowler flaps were used to achieve landing requirement. This was done mainly because double slotted Fowler flaps gave
two advantages. Extending the flap increase projected area to shift the lift curve to the left. On the other hand, slots
increase 𝐶 𝐿𝑚𝑎𝑥 by allowing high energy airflow from the bottom surface to the top. However, the wing will stall much
quicker. To counter that, leading edge slat is used to delay the stall, which shifts the lift slope to the right. At the same
time, increase 𝐶 𝐿𝑚𝑎𝑥 due to the increase in projected area. Using the combination of slat and double slotted flap will be
able to produce more lift maintaining the same stall angle of attack [6].
Sizing of the flap is done by using the similar configurations and dimensions of the high-lift system on the Boeing
777 and the SCW-2A airfoil [31][32]. It was then verified in XFLR5. However, XFLR5 model is limited in that it can
only model plain flaps. This can be adjusted for as XFLR5’s aerodynamics model uses the Vortex Lattice method. The
29
Vortex Lattice method does not account for flow separation, which can be used to simulate flow around slats and slotted
flaps. This is because slats and slotted flaps induce high energy air to flow from the bottom surface to the top, and thus
reduce adverse flow or flow separation. High-lift device dimensions were shown in Table 13.
Dimension value
The flap layout of the airfoil is going to be similar to SCW-2A [32]. This is possible as in comparison to the
SCW-2A, the SC-0412 has the same thickness to chord ratio at the same location and a minor difference in camber.
According to flap layout diagram in Fig. 29, the slat uses 14% of the chord, the vane uses 18%, and the flap uses 12%.
The cove should be properly sized to include volume for the flap extension hydraulics and actuators.
30
E. CAD Drawings
Fig. 29 Top and front drawing with major wing dimensions, along with flap layouts
F. Drag Buildup
Drag buildup is important because it impacts performance in cruise. Buildup was mainly done using Raymer’s
component buildup method. It yields similar cruise drag values when compared to Raymer’s equivalent skin friction
method, which verifies that cruise drag was accurately approximated. Wave drag was calculated using the Delta method
[33], with an estimated design lift coefficient of 0.50. Next, the design mach number was estimated using inputs of the
2D drag divergence Mach number, aspect ratio correction, and wing sweep correction. The design Mach number is 0.83,
which agrees with an earlier assumption. If the aircraft is cruising at 𝑀𝑑𝑑 or design Mach number, the aircraft will
experience 20 counts of drag. 10 counts of wave drag was estimated at Mach 0.78. Trim drag was determined using
Takeoff and landing drag buildup was estimated using a combination of Raymer for component drag and Roskam for
high-lift drag [6] [29]. Flap drag consisted of flap pressure drag, flap induced drag increment, and interference drag.
Flap induced drag increment and pressure drag contribute to the majority of the drag. The large induced drag increment
was due to flaps disturb the lift distribution, and thus it is no long being elliptical. Pressure drag on the other hand was
𝑆𝑤 𝑒𝑡
cause by large flap deflection angle. Drag buildup for landing and takeoff is shown in Fig. 16. For 𝑆𝑟 𝑒 𝑓 calculation, see
31
Table 14 Cruise Raymer component drag buildup with Delta wave drag prediction
C𝐷 𝑓 𝑙𝑎 𝑝 0.1438 0.0803
32
G. Trade study
For the wing sweep angle trade study, see Subsection X.B. For the airfoil selection trade study, see Subsection X.A.
XI. Performance
A. Performance Requirements
According to the RFP, the JJJP must be able to fly 3,500 nmi missions carrying 410 people on board [1]. This is
equivalent to a normal payload of 94,300 lb. It should also have a balanced field length of less than 9,000 ft and a
landing field length at the end of mission less than 9,000 ft when operating at ISA + 15°C. The approach speed at the
end of 3,500 nmi mission should also be less than 145 KCAS.
Additionally, the RFP states that an appropriate mission profile would be takeoff, 3,500 nmi cruise, descent, loiter 5
minutes, approach, divert climb, divert cruise 200 nmi, descent, loiter 30 minutes, approach, and then land, as shown in
Fig. 3.
It was decided by the group that a higher fidelity estimation of the performance, through out the stages of flight,
should be simulated. Specifically, a simulation was developed that took in the design parameters such as weight, thrust,
TSFC, maximum Cl, a drag model, etc. Within the simulator, the equations of motion were implemented to predict
the state of the aircraft. Then the simulation progressed in a time step fashion, simulating each stage of the mission,
including takeoff, climb, cruise, descent, loiter, approach, and landing. Additional capabilities were developed within
the performance simulator. For instance, the simulator can solve for the optimal amount of fuel to carry that would
fulfill the reserve requirement. It is also capable of performing a trade study to find the optimal operational schemes. It
To accurately account for the requirement stated in the RFP, all the takeoff and landing performance simulations
were simulated under a ISA+15°C condition. For takeoff distance, a 35 ft obstacle was assumed. In order to find the V1
33
speed, a range of speeds at which one engine fails was tested. Provided this data, the simulator then calculates the total
distance needed to continue with the take off or stop. From the simulated result, the V1 is found to be 165 KCAS at
maximum takeoff weight (MTOW), and balanced field length is found to be 7,850 ft, which easily fulfills the 9,000 ft
The comparison plot of the takeoff flight path is shown below in Fig 31. The plot shows a takeoff trajectory for AEO
Similarly, the landing performance can be determined with the time integration simulator. At ISA+15°C dry
condition, the landing distance of maximum landing weight (MLW) was 4,420 ft. Note that in all conditions, a thrust
Alternate conditions considered were wet and icy concrete/asphalt. Dirt and grass were not considered as the team
decided an aircraft of this size will not be used on grass and dirt strips. The BFL and LFL on all conditions were
34
Table 18 BFL / LFL at Sea Level
From the table above, JJJP can achieve the 9,000 ft requirement under both dry and wet conditions without the help
of thrust reversers. For icy conditions, further analysis was done with thrust reverser and the resulting landing field
length was found to be 6,960 ft. Therefore it is suggested that, when landing the JJJP on icy runways, thrust reverser
The maximum rate of climb at MTOW and takeoff condition is 3,500 ft/min. However, in normal operating
conditions, a less aggressive initial rate of climb of around 1,700 ft/min can be used. The JJJP is perfectly capable of
The service ceiling of the aircraft is dependent on the weight and speed that the aircraft is flying. The analysis
assumed a design cruise speed of mach 0.78. Therefore it was possible to calculate the heavy-weight (MTOW) ceiling,
medium-weight (50% fuel, normal payload) ceiling and light-weight (20% fuel, no payload) ceiling. Notice that the
ceiling calculated here was only the mechanical limit of the aircraft, purely dependent on the thrust and lift available.
Often time, the other systems would exert a lower limit on cruising altitude.
Examples of drag and drag coefficients for each flight segment are summarized in Table 20. At takeoff and landing,
drag are very low since the wing is not producing much lift at low speed. While for majority of the stages, JJJP is flying
with a similar range of drag values just above 20,000 lbf. The two exceptions are climb and approach. For climb, JJJP is
35
using more lift and consequently generating more induced drag. For approach, as flaps and slats are extended, more
drag are created to both descent and slow the aircraft down.
Segment C𝐷 Drag
Using the same drag model and aircraft parameters as the simulation, it was possible to calculate various performance
coefficients at different cruising speeds. Using automated tools, the following diagram was plotted at MTOW and
cruising altitude FL350 (Fig. 32). The coefficients are recorded in the Table 21.
36
Fig. 32 Performance coefficients
The three coefficients each represented the performance of the aircraft in different scenarios. Specifically, 𝐶𝐶𝐷𝐿
represented the endurance performance of the aircraft. At 𝐶𝐶𝐷𝐿 the drag was at minimum and therefore the thrust
𝑚𝑎𝑥
required is at minimum. The fuel consumption is minimized and the aircraft could stay airborne the longest.
! From the
1 3
𝐶𝐿2 𝐶𝐿2
plot, the best endurance occurred at Mach 0.73. The 𝐶𝐷 represented the range performance. At 𝐶𝐷 the aircraft
𝑚𝑎𝑥
3
𝐶𝐿2
could achieve its maximum range. From the plot, maximum range occurred at Mach 0.81. The third coefficient 𝐶𝐷
3
!
𝐶𝐿2
represented the loiter performance. The optimal loiter speed was Mach 0.6 according to the location of 𝐶𝐷 .
𝑚𝑎𝑥
𝐶𝐿
𝐶𝐷 𝑚𝑎𝑥 16.58 0.730
1
!
𝐶𝐿2
𝐶𝐷 22.65 0.813
𝑚𝑎𝑥
3
!
𝐶𝐿2
𝐶𝐷 14.68 0.571
𝑚𝑎𝑥
37
F. Payload-Range & Range Mach Diagram
The JJJP had a maximum payload cap higher than the requirements set in RFP[1]; this brought multiple advantages
when conducting short haul flights. Specifically the maximum payload of the aircraft was 150,000 lb, meaning it could
carry 50,000 lb of revenue payload on top of 400 passengers. It had plenty of capacity to either carry more people or fit
A plot of payload range diagram was provided below (Fig. 33). Each blue solid line represented the payload-range
performance at specific takeoff gross weight, while the outer most line represented the performance at MTOW. From the
plot, the JJJP can carry 103,550 lb of payload for 3500 nmi mission, 9,250 lb more than required. Otherwise, it can go
on mission up to 1455 nmi with maximum payload of 150,000 lbs. From the diagram, maximum range of the ferry
The flight envelope diagram at MTOW was shown below (Fig. 34). The absolute ceiling (heavy-weight) was 36,800
ft . The diagram considered the effect of the wave drag using the delta method [33]. Notice at low altitude, the structural
strength determined the maximum speed. The maximum equivalent airspeed of JJJP was 321 knots, which was shown
by the purple dashed line. The maximum speed at sea level was roughly Mach 0.49.
38
Fig. 34 Flight envelope @ MTOW
To study the climb performance of the aircraft, a specific excess power contour was plotted (Fig. 35). This plot
showed the power available to climb at certain speeds and altitudes. For example, at takeoff (Mach 0.25, altitude 0 ft),
the diagram suggested specific excess power was roughly 60 ft/s, which was equal to the 3,600 fpm maximum climb
rate. This agreed with our takeoff analysis that JJJP had a maximum initial climb rate of 3,500 fpm.
The red line denoted the maximum specific excess power at certain altitudes, which was ideally the best rate of
climb line. However in reality, flight envelope determined that the aircraft actually can not reach the optimal line when
flying at low altitude. Since the structural limit restricted the maximum speed JJJP can achieve at those altitudes.
39
Fig. 35 Specific excess power diagram
I. Trade Studies
The two most important factors in operation was the cruise speed and altitude, both of which determined the fuel
consumption in a significant way. If they were chosen poorly, the aircraft could be burning much more fuel than optimal
or even unable to reach the destination. It was possible to predict the fuel consumption performance with high fidelity
simulation available. Looping over different combinations of cruise speed and altitude, the contour map (Fig. 36) was
plotted to conduct the trade study. The contour in the plot showed the amount of fuel in pounds burnt in flying the 3,500
nmi mission, not counting the reserve, contingency or diverting fuel. The simulation accounted for the transonic wave
drag using the delta method [33], and accounted for the SFC change with respect to altitude, Mach number and throttle
level using the method provided in Raymer [6] and Artur’s paper ([25]).
40
Fig. 36 Cruising mach & altitude trade
The white regions on the contour plot were configurations that won’t reach the destination. Specifically, the red limit
on the top was set by the initial service ceiling of the JJJP, while the yellow limit on the right was set by the amount of
fuel carried to conduct the mission. In this case, the aircraft takes off with 90% full tank.
The optimal cruising Mach number and altitude were found to be M0.78 @ FL370, which was very close to the
designed cruise of M0.8 @ FL350. Also, notice that any speed too fast or too slow would cause the aircraft to burn more
fuel. This was expected since flying too slow would make the trip time unnecessarily long, while cruising at higher
speed would add wave drag. From the diagram, there was a trend that the higher the cruising altitude was, the lower the
fuel consumption. However, at MTOW, the aircraft reached the service ceiling at 37,000 ft, making it hard to climb to a
higher altitude. Step climb was a potential solution to this dilemma without requiring extra thrust or weight reduction.
As discussed in the previous trade study, it was obvious that the aircraft could achieve higher fuel efficiency at higher
altitude. It was necessary to cruise climb at a later stage of flight in order to maintain a high L/D. Therefore, a simulated
flight with cruise climb was plotted in Fig. 37 to compare the fuel consumption. The JJJP would try to climb to a higher
flight level (+1,000 ft) when the service ceiling allowed. The blue line in the plot showed the altitude of the aircraft with
respect to time. Four cruise climbs were conducted throughout the flight to maintain a high L/D. Over the 3,500 nmi
mission, step climb can save over 550 lb of fuel per flight.
41
Fig. 37 Simulated flight with step climb
J. Fuel Requirement
For the 3,500 nmi mission, the aircraft did not need to carry full fuel capacity. Therefore, to obtain an optimal
amount of fuel to carry, a script was used to solve iteratively an amount that fulfilled the regulation in 14 CFR §
121.639/645 and RFP, while also minimized the fuel burn. With the number obtained from trade study (M0.78 and
FL370), the optimal amount of fuel carried was 98,900 lb, in which 83,610 lb was burnt to fly the 3,500 nmi mission.
Upon the landing of the 3,500 nmi mission, there would be 15,290 lb of reserve fuel left in the tank. With step climb,
the total fuel carried was decreased to 98,400 lb, in which 86,060 lb was burnt and 15,340 lb remained upon landing.
The amount of fuel remaining during the 3,500 nmi mission was shown in Fig. 38. It was assumed that startup and
taxi burned 3,000 lb of fuel. Note that the takeoff and landing stage were so short and burned so little fuel that they were
hard to tell on the plot. Specifically, takeoff burned 468 lb and landing burned 25 lb of fuel (since for landing the engine
was basically in idle). Therefore on the plot the first sector is climb, the slope is steeper since engine operates at throttle
level higher than cruise. Then there is cruise stage which burn most of the mission fuel. The third stage is descent and
the curve is very flat, as the aircraft tries to glide to lower altitude. The last big stage is approach. As the flaps and slats
are extended, the drag created is significant and engine no longer operates in idle region. At the end of the mission,
roughly 15,000 lb of fuel remains in the tank. It consists of 200 nmi divert fuel, 35 min loiter fuel and 5% extra block
fuel.
42
Fig. 38 Fuel plot of simulated flight (3,500 nmi)
Thanks to the UltraFan engine, JJJP has a 20% reduction in fuel consumption compared to Boeing 787 on 3500 nmi
mission. Specifically, the fuel consumption per seat is only 1.83 L/100km, compared to 2.26 L/100km [35] for Boeing
787. Similarly, on short-haul mission (700 nmi), JJJP has 13% lower fuel consumption even compared to 737 Max.
Specifically, the fuel consumption per seat is only 1.98 L/100km, compared to 2.28 L/100km [36] for 737 Max.
K. Performance Conclusion
The sizing process was conducted with the aforementioned factors in Section XII.A. The 94,300 lb of the total
payload consists of 410 people on board, each weighing 200 lb and carrying 30 lb of baggage. The maximum payload
weight is set to 150,000 lb so that the aircraft can potentially carry more cargo when flying short-haul flights. In the
initial sizing code, the parameters were constrained so that the TOFL and LFL were shorter than 9,000 ft at ISA+15°C.
The target range is set to 4,000 nmi with 5% extra of reserve fuel so that there was enough fuel to divert to another
airport 200 nmi away and loiter for 35 minutes. In the higher fidelity simulation code, the actual BFL and LFL for the
mission were found to be 7,850 ft and 4,420 ft, respectively. Using the parameters and 𝑉approach = 1.3𝑉stall , the approach
speed at the end of the 3,500 nmi mission was 141.6 KCAS, fulfilling the requirement of maximum approach speed of
145 KCAS. According to the payload-range diagram 33, the JJJP can carry up to 103,550 lb of payload for 3,500 nmi
mission, surpassing the payload budget set at 94,300 lb. In conclusion, JJJP had met all of the performance requirements
43
XII. Stability and Control
A. Stabilizer Configuration
Several stabilizer configurations were compared against each other for this design, including conventional, T-tail,
and cruciform, each having their own pros and cons. Although both the T-tail and cruciform would help ensure the
horizontal tail is unaffected by the jet exhaust, the fact that a low wing twin engine configuration was chosen (already
putting the horizontal tail well out of proximity to jet exhaust) as well as the added structural weight to support the
horizontal tail on the vertical tail made the conventional tail the clear choice. Its proven reliability in stability for similar
aircraft and much lighter structural weight pushed this design to adopt it for its stabilizer configuration.
B. Stabilizer Sizing
In order to fully size the horizontal and vertical stabilizer, prior knowledge of some key geometric parameters,
namely the volume ratios and moment lever arm between the tail aerodynamic center and the aircraft center of gravity
was needed. Although there were historical values for these parameters, many of these had become outdated. Trade
studies were done in order to determine volume ratios and moment lever arms of the horizontal and vertical stabilizer.
The B777 and B787 were analyzed; the volume ratios and moment arms for these similar aircraft were presented in
Table 22 along with the values for the JJJP [37, 38].
The horizontal stabilizer was sized using the scissor plot shown in Fig. 39, constructed using methods outlined in
Torenbeek [39]. The vertical stabilizer was sized using the volume ratio method outlined in Raymer [6]. The remaining
geometric quantities were sized along historical ranges given in a lecture series by E.G. Tulapurkara [40]. The vertical
tail volume coefficient was chosen based off those for the B777 and B787 due to there being doubt that the historical
values presented by Raymer were still accurate. Table 23 outlined various geometric parameters of the tail, and a scaled
dimensional drawing was shown in Fig. 40. It was chosen to include a horizontal tail dihedral angle of Γ = 6 deg
(matching that of the wing) in order to ensure the stabilizer was completely out of the engine jetwash and disturbed air
from the wing, for greater lateral-directional stability, as well as for aesthetic purposes. Lastly, the NASA SC(2)-0010
44
airfoil, a thin airfoil with thickness ratio of 10% at 37% chord, was chosen for both the horizontal and vertical stabilizers
due to its symmetric geometry and transonic capabilities. The small thickness ratio reduced wave drag according to the
Delta method [33], and it had the ability to delay shock waves due to its uniform thickness distribution.
𝑏 72.62 ft 28.75 ft
AR 5 1.7
𝜆 0.3 0.31
Γ 6 deg –
45
(b) Vertical Tail
(a) Horizontal Tail
The JJJP utilized various control surfaces, namely an elevator, rudder, and ailerons, to maintain control of its attitude.
All three of these flight control surfaces were sized using historical values given in Raymer [6]. The geometric quantities
determined through this methodology were presented in Table 24 and scaled dimensionsal drawings of these devices is
presented in Fig. 40 and on page 12. The span ratio and span listed for the aileron is for a single device
D. Incidence Angles
It was decided that the JJJP’s horizontal tail will be able to adjust its incidence angle, allowing for a wide range of
incidence angles as shown in Table 23. This adjustable tail design was chosen for the purposes of efficient trimming,
allowing the elevator to be used solely for control. Further, Sadraey [41] points out that when flight cost was a major
design requirement, it was better to utilize an adjustable horizontal tail instead of a fixed or all-moving tail. Through
trim analysis, the necessary tail incidence angles for takeoff, cruise, and landing was determined in order to zero the
46
pitching moment of the JJJP.
E. Trim Analysis
Trim analysis was done using the methods outlined in and McCormick [42]. It was decided to have a variable
incidence horizontal tail for the JJJP; this system was to be used to trim the aircraft, with the elevator being used
mainly for control purposes. By varying the incidence of the horizontal tail, trim plots were made for takeoff, cruise,
and landing and are presented in Fig. 41. Take note that the convention of positive incidence angle being downward
(opposite of the wing) is used here. The goal is to achieve zero pitching moment about the center of gravity at cruise
with as small as possible tail incidence, making the aircraft passively stable and requiring little input from the pilot. The
wing incidence being 3 deg results in the need of 5.8 deg tail incidence at cruise. During takeoff and landing, there must
be a greater tail incidence in order to produce a moment that allows the JJJP to ascend/descend at the proper angle of
attack. Trim points were chosen to ensure the necessary lift at each mission segment was achieved. These incidence
angles were all within the deflection range stated in Table 23.
For the JJJP to have longitudinal static stability, it had to be able to resist changes in AoA by producing an opposing
pitching moment. In order words, 𝐶𝑚 𝛼 < 0. Various derivatives related to lift and moments were calculated using
methods outlined in Raymer [6] and McCormick [42] and were collected in Table 25.
47
Table 25 Longitudinal Stability Derivatives
Derivative 𝐶𝐿𝛼 𝐶𝑚 𝛼 𝐶𝑚 𝛿𝑒 𝜖𝛼
The negative value of 𝐶𝑚 𝛼 as well as the downward slopes of the trim plots shown in Fig. 41 indicate that the
JJJP possessed longitudinal static stability. An increase in the AoA resulted passively in the production of a restoring
pitching moment. Additionally, the negative value of 𝐶𝑚 𝛿𝑒 suggested that a positive elevator deflection (pitched up)
further resulted in a counterbalancing pitch moment. Further, using equations given in Raymer [6], the neutral point and
static margin were found using the forward and aft 𝐶𝑔 limits and were tabulated in Table 26. The three lengths were
measured from the nose of the JJJP whereas the static margin was given as a percentage of the MAC. Although slightly
higher than similar aircraft, this positive static margin nonetheless indicated that the 𝐶𝑔 of the JJJP was ahead of the
In order to get a full understanding of the JJJP’s stability, the lateral-directional static stability was also analyzed.
Using the method outlined in McCormick [42], the various derivatives used to determine lateral-directional stability
were found and were shown in Table 27. The negative value of 𝐶𝑙𝛽 implied that the JJJP naturally produced a restoring
roll moment given a change in sideslip angle. The positive value of 𝐶𝑛𝛽 indicated that a counteracting yaw moment
occurred in the event of a sideslip change. These, as well as the signs of the other derivatives in Table 27 proved that the
H. Dynamic Stability
Dynamic characteristics were also analyzed besides static ones. The six degrees of freedom equations of motion of
this aircraft, as given in McCormick [42], were solved for their roots, from which dynamic modes were determined as
48
well as other characteristics. These findings were tabulated in Tables 28 and 29. All but the spiral mode had negative
real part, indicating dynamic stability. However, the doubling time of 53.7 seconds allowed plenty of time for the pilot
Additionally, the rudder sizing was confirmed by determining the necessary deflection in the case of one engine
inoperative (OEI) for takeoff and landing and was presented in Table 30a. This rudder trim was determined using
equations outlined in the lecture notes of Dieter Scholz [43]. With the takeoff rudder trim being well within the deflection
range, the area of the rudder was confirmed to be large enough to counteract the most adverse condition of OEI.
Lastly, the FAA required transport aircraft be able to sufficiently perform normal lateral maneuvers with OEI (14
CFR §25.147). Due to the slight ambiguity in this statement, the FAA released in the Advisory Circular (AC) 25-7D
[44] that the test case for this would be for the aircraft to perform a roll maneuver through 60 deg (30 deg bank angle in
one direction to a 30 deg bank angle in the other direction). This maneuver was to be done within 11 seconds to meet
the requirements per AC 25-7D. The roll time calculated from equations in McCormick [42] was shown in Table 30b
and was found to be 8.5 seconds, well below the FAA’s test case guidelines.
49
XIII. Structures and Loads
1. V-n Diagram
In order to determine the aircraft’s flight envelope and to understand the structural limits of the designed aircraft, a
V-n diagram overlayed with a gust diagram was plotted using methods outlined in Roskam V and is shown in Fig. 42
[45]. This analysis helped find loads experienced during maneuvering and produced by gusts of wind. The gust lines
being fully within the boundaries of the solid V-n plot in Fig. 42 indicated gust loads had very little impact on the
JJJP. Unlike a much smaller airplane, the JJJP’s large weight and high cruise speed allowed it to not be considerably
affected by gusts. Further note, the limit load of 2.5 was chosen to comply with the minimum limit load given by 14
CFR §25.337(b).
A qualitative analysis of some of the load cases of interest experienced by the JJJP was shown in Fig. 43. The loads
on the wing included lift and the weight of the engine, idealized as a point mass. These forces were transferred from
the wing box to the fuselage via skin panels, ribs, and spars. Further, the JJJP’s fuselage experienced pressurization
loads. With the cabin pressure chosen to be equivalent to the air pressure at 8,000 ft altitude per the RFP [1], a pressure
differential of 7.78 psi was present at cruise, these values being tabulated in Table 31. This pressure differential was
used to define the minimum skin thickness of the fuselage. These pressurization forces traveled through the fuselage via
50
the skin, stringers, and longerons. Lastly, the landing loads experienced by the landing gear traveled through the main
Parameter Value
Additionally, the lift distribution was approximated using Schrenk’s approximation between a trapezoidal and
elliptical lift distribution over the wing and was presented in Fig. 44a. From this, moment and shear diagrams for the
wing were plotted and were presented in Figs. 44b and 44c. A safety factor of 1.5 was used for this methodology. The
moment and shear load were at a maximum at the wing root tapering off to 0 at the wing tip. The ultimate moment and
shear values, found at the spanwise root chord location, were found to be 8.084 × 107 lb-in and −1.998 × 105 lb. These
plots were used for material sizing and selection, specifically the spars and ribs.
51
(a) Lift distribution approximation
B. Material Selection
When considering what material should be used in the construction of the main structural components of an aircraft,
the two most common materials used are either aluminum alloys or composites.
Depending on the composite, they can have high strength at very low densities making them appealing to the
aerospace industry. The major problems that arise when dealing with a composite structure is the high cost of not only
the material, but also the complex manufacturing processes involved in a composite layup. Along with a high cost, it
is also extremely difficult to mend a composite fuselage and wing together as traditional fasteners will cause a stress
concentration that could result in failure at those locations [46]. If these issues are overcome, maintenance becomes
another key factor. Composite structures are much harder to conduct repairs on than metal ones and with this aircraft
focusing on short hauls, the structure will be prone to fatigue making fast, cheap and reliable repairs vital to keeping the
52
aircraft flight worthy [46].
With this knowledge, it was decided to go with a primarily aluminum alloy construction. Aluminum alloys have
been tested and used in industry for years, are relatively easy to manufacture and test, are readily available, and have
Because of this, four different aluminum alloys have been chosen to conduct a trade study including Al 2024, Al
7075-T73, Al 6061, and Al 7079. Depending on which structural component is being designed, different material
properties of the aluminum alloy should be focused on. For things like the wing spars, ribs, along with fuselage frames
a material with a high elastic modulus and yield strength are favored due to the high stress these components will see
during every flight. As for the skin of the fuselage and wings, a material with a high elastic modulus is desired as well
as high fracture toughness to avoid crack propagation in the case of damage to the surface. The physical properties if all
√
Material Yield Strength (MPa) Elastic Modulus (GPa) Fracture Toughness (MPa- 𝑚) Density (𝑔/𝑐𝑚 3 )
Al 7079 450 70 20 –
Based on this information, Aluminum 2024 is considered the best for the skin of the aircraft given its large elastic
modulus and fracture toughness. These properties mean that the surfaces can experience high loads and still maintain
its shape. For the fuselage this is important since there will be a pressure difference of 7.78 psi between the inside and
outside the cabin when cruising at 37,000 ft. Additionally, with a high elastic modulus the fuselage will not bulge as
much as a material with a lower elastic modulus. If there is extreme deformation of components such as the wing, the
aerodynamics will vary largely from what was predicted, so maintaining the desired shape is critical to maximizing the
performance of the aircraft. The fracture toughness determines how fast a crack will propagate in the material, with a
higher value corresponding to a slower propagation. Aluminum 2024 has one of the highest values of fracture toughness
which means if there is a crack in these surfaces, it will resist fracturing better than most of the other alloys investigated.
For structural components such as wing ribs and spars, Aluminum 7075 has been chosen for its high yield strength
and favorable elastic modulus, meaning it will be best suited for enduring the high shear and bending loads that these
Aside from the main structural components of the wing and fuselage, materials for other various components needed
53
to be determined. For the landing gear, an aluminum alloy was not a viable option due to the high loads and cycling
of this component making it prone to fatigue. Instead, the steel alloy, AISI 4340 was chosen for its high tensile yield
strength, allowing for repeated landings with minimal maintenance needed to keep to keep the landing gear in service
[48]. Another special consideration had to be given to the radome of the aircraft placed at the tip of the nose. Since all
radio signals will need to pass through, aluminum was not an option as it can block many of the signals from being
transmitted. To solve this issue the radome will be constructed of s-glass fiberglass which is non conductive and allows
the transmission of radio signals, while having a tensile yield strength of 4750 MPa and can operate at a maximum
C. Structural Arrangement
The main structural components of the the JJJP are analyzed in this section, including the fuselage, wing, empennage,
and horizontal and vertical stabilizers. Each of these components have structural configurations meant to handle and
distribute the loads expected through the full flight envelope to ensure failures will not occur. The complete structural
1. Wing Structure
The main wing structure consists of a series of ribs attached to two main spars that run the length of the wing and
connect at the center in the wing box as shown in Figure 46. Along with the main spars, there is also supplementary
spar placed near the root of the wing that acts as a support for the landing gear strut.
54
Fig. 46 Wing Structural Arrangement
This configuration is very common in the aerospace industry, especially for this class of transport aircraft, since it
provides a large amount of strength while maintaining a relatively low weight to maximize performance [7]. The front
and rear spar locations are typically constrained by the location of the the high lift devices on the wing, with a larger
torque box being advantageous in rib sizing as the thickness can be reduced since there is better distribution of the shear
torsional forces experienced during flight. For these reasons, the front spar was placed at 14% of the chord and the rear
spar located at 60% of the chord along the span of the wing. The exact size of the spars was determined by the ultimate
bending moment experienced at the interaction of the wing and body, while being constrained by the thickness of the
wing at that location [50]. A diagram of the front and rear spar dimensions is shown in Fig 47.
Once the spar size and locations is determined it is then possible to size and space the ribs along the span of the wing.
To do this, an idealized wing box analysis was used to calculate the shear flow through the rib at the intersection between
the wing and fuselage as this is the location of the ultimate shear force experienced. Again the exact methodology to
determine the rib thickness was done by following the methodology outlined in Niu [50]. From this the rib thickness
was determined to be 0.35 in. thick which was kept consistent along the wing for consistency and ease of manufacturing.
As for the spacing, an approximation was used from historical data of transport aircraft and was chose to be 24 in. apart
55
as presented in Roskam part 3 [7]. The orientation of the ribs was also considered as there are two primary option of
doing so. The first option would be to have the spars parallel to the freestream. The advantage of this configuration is
that the wing airfoil shape is held more accurately down the span of the wing, resulting in a slight improvement in the
performance of the aircraft. The major drawback of this configuration is the fact that there are large stress concentrations
at the intersection between the rib and the front spar resulting in reinforcements needed at these locations[50]. For this
reason, the ribs were placed perpendicular to the front spar resulting in a more structurally sound construction.
Once the preliminary sizing of the wing structure was completed, a linear elastic finite element analysis was done to
verify it could withstand the loads expected during the full flight envelope. To do this, a lift distribution taken from 1g
Shrenks approximation during cruise conditions was applied to the wing of the aircraft down the span. Since the skin
was not modeled in this simulation, the lift force was distributed along the top surface of the front and rear spar with the
front spar taking 56% of the load and the rear spar taking 44%. This was decided based on the location at which the lift
acts about at the root then doing a summation of the moments experienced by the front and rear spar. Along with lift, a
point force was placed at the location of the engine to represent its effects on the structure during flight. Once this was
done, the solution was obtained using the Nastran solver built into NX 12.0 , with the results shown in Figure 48.
56
From the simulation, it was determined that throughout most of the wing structure the stress levels were around
35,000 psi, along with a total displacement of 12.3 ft of the wingtip. Although it should be noted that the maximum
stress value was seen at a single node where the wing and fuselage intersect, and is likely an unrealistically high value
do to the way the mesh was constructed and connected at that location. Throughout the rest of the wing, The highest
values observed were around 60,000 psi, which is well beneath the yield strength of 73,000 psi for Aluminum 7075. In
terms of the deflection of the wing, a 12.3 ft deflection is consistent with aircraft of similar size, with the Boeing 777
having a maximum deflection of 15.5 ft [51], and the Airbus A350 seeing a maximum deflection of 17.6 ft [52]. Further
refinement of the FEA setup may be helpful to get more accurate results for the maximum stress, allowing for iterations
For the attachment method used, a wing box is constructed running the width of the fuselage. This consists of two
I-beams extended from the wing spar with longitudinal supports to add torsional rigidity. The wing box is connected to
the frames of the fuselage by a series of bolts. Both wings will be bolted onto the wing box where they meet at the
fuselage wing intersection with a set of lugs on the front and rear spar. This will allow for the aerodynamic forces to be
efficiently transferred from the wing to the spars and absorbed by the wing box.
2. Fuselage Structure
The fuselage construction followed similar methodologies to that of the wing. The main structural components of
the fuselage consist of the frames, stringers, bulkheads, and skin. Many of these parameters were determined following
approximations found in Roskam Part 3 [7]. The results of this can be seen in Table 33.
Since the JJJP will be flying at an altitude of 37,000 ft, it is essential that the cabin be pressurized to a maximum of
8,000 ft pressure altitude. This means the inside of the cabin will have to maintain a pressure of 10.92 psi throughout
the entire flight envelope, with a pressure difference of 7.78 psi between the cabin and atmosphere during cruise. The
pressurization was used in determining the skin thickness of the fuselage. To do this, a thin walled hoop stress analysis
was done to find the thickness required for safe operation of this aircraft [53]. Frame spacing and depth for this aircraft
was done using the approximations outlined in Roskam Part 3 [7], then confirmed with a similarity analysis using a
structural layout diagram of the Boeing 777-300. Using Siemens NX 12.0 to scale the image and measure the distance
between frames, it was found that the Boeing 777-300 has a frame spacing of 21.5 inches, which very similar to the
57
frame spacing used on the JJJP. These results were to be expected as the size of both aircraft are very similar. Another
design consideration taken into account of the fuselage came with the shape of the stringers to be used. Two of the most
common shapes used in this class of transport aircraft are J-stringers and Z-stringers. Z-stringers are very structurally
efficient at transferring loads and use less material to that of a J-stringer. J-stringers are less structurally efficient, but
have the advantage of dual fasteners due to its geometry [50]. Although this is a beneficial fail safe to have, lower weight
due to the smaller size and fewer fasteners needed for the Z-stringer proved to be more important than the extra fastening
Bulkhead and mainframe placement was also something to be considered when designing the structure of the
fuselage. A total of two bulkheads and three mainframes were decided to be used on the JJJP. The bulkheads were placed
in the nose and tail of the aircraft to maintain pressurization of the fuselage. The mainframes are placed right after the
cockpit as well as at the front and rear spar of the wing box allowing for the loads from the wing to be transferred to the
rest of the airframe structure. The full fuselage construction can be seen in Fig. 49
3. Empennage Structure
The philosophy behind the empennage construction was the same taken as the fuselage and wing shown earlier. The
tail cone section has identical frame and longeron dimensions as the fuselage tapering all the way down to the tip of
the tail. While the horizontal and vertical stabilizers follow the same construction as the wing with two main spars
and ribs perpendicular to the leading edge spar. The spacing of the ribs in the tail is set to 24 in. apart matching the
estimations shown in Roskam Part 3 [7]. The horizontal tail surface also has a wing box similar to that of the wing,
while the vertical tail will be mounted to the top of the fuselage along a reinforced longeron placed along the top surface
of the tailcone. The horizontal and vertical stabilizers can be seen in Fig 50.
58
(a) Horizontal Stabilizer Structure (b) Vertical Stabilizer Structure
In Part V of his book, Airplane Design, Roskam provides two methods of estimating airplane component weights:
Class I and Class II [45]. Class I consists of an empirical sampling estimation, wherein weights from other aircraft in
the same class are averaged and scaled based off of the designed-aircraft’s gross weight. Class II utilizes equations
based on the aircraft’s geometry and structure in order to estimate weights for the following components: fuselage,
empennage, fixed equipment, powerplant, landing gear, nacelle, and wing [45].
x y z
59
Table 35 Roskam Class I Weight Component Estimations
Using the Roskam Class I estimation method, a 2-dimensional analysis was performed to estimate the component
weights of the aircraft. The weights were estimated using the methodology and data provided on page 11 in Part V [45].
The resulting weight estimations are provided in Table 35, along with the location of their respective 𝐶𝑔 with respect to
the datum described in Table 34. The analysis was limited to 2 dimensions, as landing gear height, engine pylon length,
vertical tail span, and other parameters were in the process of being defined at the time of this analysis. The analysis was
further simplified to one dimension (along the length of the fuselage), due to a symmetrical configuration.
The component weight analysis was expanded by performing Roskam’s Class II method for determining component
weights [45]. In this method, Roskam provides equations from four sources: US Air Force, Cessna, General Dynamics,
and Torenbeek [45]. Roskam’s commentary suggests that the equations from General Dynamic and Torenbeek should
be representative for aircraft of this class. Both sets of equations were evaluated and the results were averaged for a final
Class II estimation.
Finally, the resulting weight estimations were validated against the weight fractions that were provided by Roskam
Class I analysis and the Boeing 787-8, for reference [54]. In some cases, the Class II equations specified weights for
sub-components of the weight groups. The resulting weight groups and their sub-component weights are provided in
60
Table 36 Roskam Class II Weight Fractions (Component/𝑊𝐸 )
61
Table 37 Roskam Class II Weight Estimations
Component Group Equations Used [45] Total Weight (Sub- 𝐶𝑔 Location (x,y,z)
(Sub-Components) Component Weight) (ft)
(lb)
Wing 5.7 62,000 (101.5,0,15.5)
Landing Gear 28,000
Nose Gear 5.41, 5.42 3,600 (25,0,4)
Main Gear 5.41, 5.42 24,400 (120,0,4)
Empennage 11,000
Horizontal Tail 5.19 8,000 (179,0,24)
Vertical Tail 5.20 3,000 (174,0,37)
Fuselage 5.26,5.27 48,500 (85,0,18.5)
Nacelle 5.35,5.37 5,000 (90,0,7.5)
Powerplant 30,500
Engine Estimation From Sim- 26,000 (90,0,6.5)
ilarity Analysis
Air Induction Sys- 6.9,6.10 500 (145,0,18.5)
tem, Engine Controls,
Starting System
Fuel System 6.20,6.23 4,000 (136,0,6.5)
Fixed Equipment 56,500
Flight Controls 7.5,7.6 4,400 (101.5,0,18.5)
Hydraulics & Pneu- Roskam Section 7.2 6,500 (145.5,0,18.5)
matics [45]
Electronic Systems 7.15,7.17 2,300 (145.5,0,18.5)
Instruments, 7.23,7.25 5,200 (20,0,18.5)
Avionics, Other
Electronics
Air-conditioning, 7.29,7.30 4,200 (101.5,0,18.5)
Anti-Ice System,
Pressurization
System
Oxygen System 7.35,7.38 1,200 (145.5,0,18.5)
APU 7.40 300 (174,0,23.75)
Furnishings 7.44,7.45 27,000 (85,0,18.5)
Baggage and Cargo 7.48,7.49 2,200 (145.5,0,18.5)
Equipment
Paint 7.51 3,200 (85,0,18.5)
𝑊𝐸 241,500 (102,0,15)
Provided the component weights found in Table 37, along with the parameters (i.e. fuel weight) found in the initial
sizing analysis, a comprehensive list of relevant operating weights can be found in Table 38.
62
Table 38 Operating Weights
B. Structural 𝐶𝑔
Using the data from the two Roskam weight analyses, a 𝐶𝑔 can be estimated for the aircraft structure. In addition
to the Roskam weight estimation method, the Mass Properties Engine in OpenVSP was used to calculate a 𝐶𝑔 of the
model. The calculation is performed by dividing the model into 150 segments and calculating a 𝐶𝑔 from using the
segment weights. Similar to the Roskam Class I analysis, the OpenVSP analysis was limited to a single dimension. The
OpenVSP model was later migrated to Siemens NX. The mass properties tool in NX also calculates a 𝐶𝑔 value. The
Due to lower fidelity in both the OpenVSP model and the NX model, in comparison to the constructed aircraft, the
structural 𝐶𝑔 values provided by the modeling methods are used only as validation for the values from the Roskam
calculations. Furthermore, as the Class II values incorporate aircraft-specific geometry, rather than exclusively empirical
data, the structural 𝐶𝑔 provided by Roskam Class II was used as the final value.
Table 39 𝐶𝑔 Estimations
C. 𝐶𝑔 Travel
As a result of attitude changes (e.g. pitching during take-off) and fuel burn, the 𝐶𝑔 will shift during a mission.
Furthermore, for any given flight, it unknown how full the passenger and cargo cabins will be, respectively. Thus, a
calculation must be performed for 𝐶𝑔 that accounts for the extremes (forward or aft) in possible loading distributions.
To account for this, Torenbeek suggest that a load and balance diagram be derived by calculating 𝐶𝑔 at a variety of
extremes. These extremes account for MTOW, MZFW, MLW, OEW, take-off pitch, landing pitch, and fuel burn [55].
63
Fig. 51 Loads and Balance Diagram
As shown in Fig. 51, the 𝐶𝑔 of the aircraft travels from 98.5 ft to 104.7 ft aft of the nose (17.5% MAC to 47.5%
MAC) – the shaded region of the plot. To capture this travel, the 𝐶𝑔 of the fuel was set to be entirely at the front, center,
and rear of the fuel tanks, respectively. The total aircraft 𝐶𝑔 was calculated at each of these increments, as well as across
Torenbeek describes the loading and distribution of passengers to be entirely arbitrary [55]. Thus, the 𝐶𝑔 travels
due to passenger distribution. To capture this phenomena, a Monte Carlo simulation was ran, where in the number of
passengers and crew were randomized in a uniform distribution from zero to 410, and their respective seating locations
were randomized, provided the seat layout described in Section VIII.C. Aisles, galleys, and restrooms were included in
the possible locations where weight my exists, accounting for the full interior of the cabin. The simulation was run 5000
times per operational loading condition (e.g. MTOW, MLW, varying fuel weight, etc.). The result of the Monte Carl
simulation was distinct clusters of passenger-driven 𝐶𝑔 travel. Torenbeek describes this behavior as the “loading potato”
[55].
Finally, Torenbeek notes that “a margin of a few percent M.A.C is usually assumed” [55]. Thus, a forward and rear
64
D. Cargo Hold Location Trade Study
Interestingly, it became quickly evident that the traveling 𝐶𝑔 , as a result of the randomized passenger distribution,
followed a nearly-quadratic trend with linear decreases of operational weight. This discovery yielded the question of
whether the location of the baggage hold (i.e. the location along the length of the fuselage allocated for baggage and
other cargo) could offset this quadratic trend, trying to keep the 𝐶𝑔 as close to the structural 𝐶𝑔 as possible during the
flight. A similar Monte Carlo simulation was run, repeating the process, but with the location of the cargo load being the
randomized variable. The qualification for success of this trade study would be a quantifiable increase in flight duration
that the 𝐶𝑔 remained close to the structural 𝐶𝑔 during flight. After completing the Monte Carlo simulation, it was found
that the only way the cargo loading would have sufficient offsetting effects was in lucky cases where the location of the
cargo grossly opposed the loading of the passengers. It was assessed that the randomization of the passenger loading
could easily vary flight-to-flight. Thus, the location of the cargo hold was determined exclusively by stability-driven
requirements, similar to the wing placement and other high weight components.
A. Trade Study
The first step in designing a landing gear system is deciding what type of configuration will be optimal for the
application at hand.The first decision to be made is weather the landing gear system will be fixed or retractable. The
factors when deciding are based on drag, weight, cost, and maintenance. Although cost and weight are important, for
transonic flight the drag produced will be the major factor in deciding what system to pursue. For this reason, fixed
landing gear will not be realistic option for this type of aircraft as the drag being produced during cruise will have
detrimental effects on the performance expected. The next decision needed to be made is the configuration to be used on
this aircraft. The three most common types of landing gear are tricycle, bicycle, and tail draggers. When choosing a
configuration for a transport aircraft, the factors deemed most important are stability of the system, line of sight, ease of
use, and practicality [56]. A figure of merit can be seen in Table 40 , with the weighting system of 1-3. For this study a
3 was most favorable while a 1 is unfavorable. The basis for which stability was determined came from the likelihood
of a lateral tip over and loss of control when steering on the runway. Line of sight is taken into account for the pilots
visibility of the runway on takeoff and landing based off of the landing gear configuration. Ease of use took into account
the takeoff and landing procedure with each type of landing gear configuration. Practicality was based off of the space
and placement available for the landing gear on this particular plane being designed.
65
Table 40 Landing Gear configuration Figure of Merit
Stability 3 1 2
Line of sight 3 2 1
Ease of use 3 2 1
Practicality 3 1 2
Total 12 6 6
Based on this assessment, the tricycle landing gear configuration is the best choice for this situation, as is of no
surprise as this is almost exclusively the configuration chosen of other aircraft in the same class. This configuration
results in favorable ground control, a level floor during taxi for comfort of the passengers, as well as easy takeoff and
B. Configuration
Once the configuration is selected it is then possible to determine the location and height of the main and nose gear.
The most aft cg location was used to ensure stable takeoff and landing procedures. The dimensions were calculated
following the methods detailed in Roskam Part II meeting the criteria for lateral tip-over, longitudinal tip-over, and
ground clearances [57]. With the dimensions of the landing gear known as well as the MTOW, the next step would be to
determine the loads experienced by the nose and main gear. This was done again following the approximation in Roskam
Part II [57]. These equations give an estimate for the loads experience by the landing gear and with this information it is
possible to use Table 9.2 in Roskam Part II to estimate the size of the tires needed as well as the number of wheels per
strut [57]. The results of the calculations for the landing gear to be used on the designed aircraft are shown in Table 41.
Distance Height Wheels Tire Tire Load Shock Shock Ab- Shock Absorber
from per size (in Pres- per Absorber sorber Di- Stroke Length
Nose strut x in) sure Strut Length ameter
Nose 21 ft 10.5 ft 2 40 × 190 psi 2.91 × 17.4 in. 6.2 in. 12 in.
Gear 15.5 104 lb
Main 108 ft 11.4 ft 6 52 × 200 psi 2.13 × 17.4 in. 15.7 in. 10 in.
Gear 20.5 105 lb
With the current landing gear configuration, the tailstrike angle was calculated to be 15 deg and a lateral tipover
angle of 12 deg. These values are both sufficient in preventing tail strikes from happening on takeoff or landing as well
66
as providing clearance for the nacelles while taxiing. After the loads, location, and height of the landing gear system
were determined, it is then necessary to go a step further and obtain information on the shock absorbers required for a
safe landing. Oleo shock absobers were chosen to be used due to their high efficiency at transferring the loads to the
airframe structure upon landing, and the sizing methodology used follows what is shown in Roskam Part 4 and takes
into account descent rate as well as the maximum landing weight [56]. The results of these calculations can also e seen
in Table 41. Knowing the exact dimensions along with the expected load, exact tires were then selected for this aircraft.
For the nose, Goodyear 405K89-2 Flight Leader tires were chosen and are rated for 39,500 lbs which is well above what
is expected. For the main gear, Goodyear 520K09-7 Flight Leader was chosen which are rated at 63,700 lb [58]. This
is well above the 35,500 lb of force expected by each tire on the main gear, allowing for a large factor of safety and
ensuring a safe landing is possible even in the event of a tire blowout. Due to the high loads of seen by the landing gear
during every flight, a material stronger and more durable than Aluminum must be used to maximize lifespan of this
system. For this reason AISI 4340 will be used for its favorable strength, toughness, and fatigue strength. Although
there are stronger titanium alloys, the cost of these materials is much greater and the physical properties of the steel
For the actuation system of the landing gear, the nose gear will retract forward into the nose of the aircraft as this is
the safest configuration since it reduces the chances of the gear shearing off during landing. A forward retraction also
allows for gravity to assist in the deployment of the gear in the case of a failure in the retraction system. The main
gear will fold inboard to the wing and fuselage structure as this is optimal for avoiding systems such as fuel tanks and
high lift devices located in the wing structure. The retraction system will be hydraulically powered as discussed in the
Systems subsection. A visualization of the stowed landing gear configuration can be seen in Fig 52.
67
XVI. Folding Wingtip System
A. Structures
The wing frame includes additional structure at the folding wingtip hinge and surrounding location to ensure
integrity during flight. The main wing section and wingtip have two sets of 0.175 in ribs at the hinge boundary. The
hinge itself is composed of two sets of 0.5 in titanium brackets and shafts. The maximum loads on the hinge are during
flight, and are determined by the shear and moment calculations based upon a schrenk’s lift distribution. The resulting
shear and moments acting on the hinge are 4,800 lb and 186,000 lb-in respectively, which includes a 1.5 safety factor.
Utilizing a FEM analysis, the maximum equivalent (Von-Mises) stress of the brackets was determined to be at the shaft
holes. The bracket thickness was then sized to be meet a 1.25 safety factor, at 0.5 in thick. Four sets of shafts are utilized
for rotation and locking of the hinge. The shear and moment forces result in a maximum transverse load of 44,000 lb per
shaft. The resulting shear stress is low enough to ensure a safety factor of 5.
68
Fig. 54 Folding wingtip structural hinge CAD model
The structural weight breakdown can be seen in Table 42. The total weight is approximately 100 lb.
Ribs 21.3 lb
B. Systems
The systems within the folding wingtip are a significant part of the logistics and weight of the overall wingtip design.
There are two major actuation systems being used that consist of four total actuators. Two electric actuators are being
used to lock the wingtip in place during flight mode. They only require a small amount of power to slide the pins in
and out of place, so electric actuation is a viable option in this case. The other two actuators will be hydraulic rotary
actuators that actuate the wingtip up and down. This system was chosen to be hydraulic because it contrasts with the
electric locking actuator which in turn makes the wingtip systems less likely to fail all at once. The hydraulic actuators
were also chosen because the design will use the close-by hydraulic power from the ailerons when the plane is on the
ground. The system was designed this way because aileron actuation is not needed during taxi, so some hydraulic power
can easy be rerouted to the wingtip system with very little extra hydraulic line run.
69
The wingtip folding system was designed with rotary actuation because after an analysis of linear actuation, it was
decided that with the hinge design, the moment arm and stroke of the actuator would make it difficult to actuate the
wingtip properly. This was because of the very small moment arm the actuator would have, along with the fact that the
arm would have to stick out of the bottom of the wing. This would make a gap in the wing skin that would need to be
filled, but fixing this problem would add more complexity and weight to the design. The rotary actuator has a simpler
The wingtip rotates on a hinge inside the end of the wing. The rotary actuator will rotate the hinge pin and the
wingtip with it that will be fixed to the hinge pin. To calculate the weight added by these systems, a simple analysis was
done. The moments from both weight of the wingtip along with potential gusts were calculated as a function of the
angle that the wingtip is folded to determine the maximum torque that the actuators would have to overcome at any
one point in time. It was determined that the maximum torque from weight comes at 0 deg (extended position) and
maximum torque from gusts comes from 90 deg (folded position). Overall, adding them together does not result in a
significant increase at torque for any one point, as they are biggest at opposite angles. Through this calculation it was
found that the actuators needed to overcome a torque of 8,600 lb-in. Adding an extra factor of 10 percent to allow the
actuation to occur in a timely manner makes the torque needed 9,500 lb-in.
After determining a torque needed, an estimate could be made on the weight of a system. Using existing aerospace
grade hydraulic rotary actuators from companies such as Parker, an estimate of weight of the system was found using
linear interpolation based on the ranges of torque and weight of the existing actuators [59]. The actuators for the locking
mechanism only move the pin into place, so they were estimated off the size of small linear actuators from Moog [60].
Table 43 shows the weight added by the wingtip systems. The total systems weight added for both sides is 85 lb.
Locking Actuator 5
The primary motivation for the folding wingtip system was to gain the performance benefits of a 184 ft wingspan
while maintaining access to Group IV airports that require a 171 ft wingspan. A larger wingspan results in lower fuel
required but a heavier wing and resulting aircraft. For this design decision to be valid, the decrease in fuel cost must be
greater than the increase in unit cost. As such, a trade study was performed between two cases - an aircraft with 184 ft
70
wingspan and folding wingtips (the JJJP) and an equivalent aircraft with a 171 ft wingspan and no folding wingtips.
The initial respective aircraft were based upon the optimal aspect ratio and wing area as determined by the sizing
analysis in section VII. The empty weight calculation is based on the Roskam Class II weight estimation methodology
[45]. The fuel burn is based on the timestep integration specified in section XI. The fuel burn calculation is based on a
700 nmi mission with a 400 passenger payload and the minimum fuel necessary to meet the range.
Utilizing the CERs specified in section XVIII, a cost analysis was performed based upon the parameters specified in
Table 44. The unit cost was based off the Roskam unit cost analysis, assuming 1000 units manufactured [61]. The
operational cost was based off the Raymer operating cost analysis, assuming 4,500 flight-hours per year [6]. For the
purposes of this trade study, a 0% profit was implemented. The total lifetime cost is simply the sum of the unit cost and
operational cost over a 25 year timespan. The results can be seen in Table 45.
The trade study concludes a savings of approximately $ 9,300,000 or 0.36% of the overall lifetime cost. The
monetary savings may seem relatively small. However in a competitive airline industry in which profit margins are thin
and one in which aircraft are bulk ordered, the overall savings are significant.
71
XVII. Auxiliary Systems
A. Flight Controls
The flight controls are an integral part of the systems of the aircraft. In order to do any basic function with the
aircraft, the flight controls are required. Flight controls include the input to the control system, the control surfaces, and
This aircraft will utilize a fly-by-wire (FBW) system of flight controls. A FBW system uses a primary flight computer
that sends electronic signals relayed from the pilot controls to the control surfaces instead of the conventional method of
using pulleys and long lengths of cable to actuate the control surfaces. There are a few reasons why the FBW system
is a better choice than the conventional flight control systems. In fact, with their decision for making the 777 a FBW
system, Boeing gives the following six reasons : Overall weight reduction, integration of several systems into one, better
handling of the aircraft, ease of maintenance, ease of manufacturing, and greater flexibility with changes and integration
of new systems into the aircraft [62]. On the downside, the cost of the system is significantly increased, although savings
from reduced weight are gained when the aircraft is in service. This is a trade off between cost now and cost later, but in
the end the positive traits of the FBW system outweigh the extra cost. One of the biggest advantages for this aircraft will
be the weight savings that the FBW system provides. Since this aircraft is designed for short haul missions, reducing
weight is one of the biggest ways to improve efficiency. FBW also has some added safety benefits that come along
with it. Because the system is running all of the controls through the primary flight computer, it is easy to enact safety
protocols that aid the pilot in emergency situations as well as keeping the aircraft out of emergency situations. The use
of the primary flight computer also allows for an easy integration and use of autopilot equipment, as it can easily receive
The FBW system, in order to create safety in redundancy, will employ three primary flight computers (PFC) with
three lanes that each control the surfaces of the aircraft. One PFC will take input from the autopilot, while the other take
input from the pilot controls. They will also be configured in a way that allows for another PFC to help out in the case of
a failure or overload of another PFC. This means the system has a significant amount of redundancy that translates to a
safer aircraft. Shown in Fig 55 is a basic diagram showing how the FBW system works and interacts with the aircraft to
get input from the cockpit to the actuator control electronics. These electronics communicate the inputs to the actuators
in order to correctly actuate at the right time. In the system, the flaps, ailerons,leading edge slats, and each side of the
elevator will each have 2 actuators. The rudder will employ 3 actuators. The spoilers and trim are each slated to have
one actuator.
72
Fig. 55 Fly-by-wire flight control setup diagram
B. Engine Controls
Engine controls in the aircraft are what converts pilot input into the corresponding action within the engine. Some
of the main aspects that the engine controls deal with are fuel connections, throttle input, cockpit alerts and warnings,
air data requirements, fire detection and protection, engine start and restart, and engine health monitoring [63]. The
JJJP, along with most other modern aircraft, will utilize a Full Authority Digital Electronic Control (FADEC) unit to
control the engine and all of it’s functions [63]. The FADEC will allow the integration of the engine into the flight
controls with the rest of the aircraft. This integration will allow all the information that is going to and from the engine
to be consolidated. The pilot can then have warning lights or buzzers from the instruments to allow proper warning if
something is wrong. Doing this makes the aircraft safer without causing a data overload for the pilot. An example of the
layout of the FADEC system and how it provides feedback is shown in Fig 56 [63]. The FADEC system for this aircraft
will come from Rolls-Royce because they are the developers of the engine and have the technical capabilities to create a
FADEC system.
73
C. Fuel System
The fuel system stores fuel in the aircraft and transports it when needed to the engines, APU, or out of the aircraft.
The fuel system of the JJJP can be seen below in Fig 57. This system uses two wing tanks and one center tank. Each
engine can be fed from the wing tank on its respective side, or the center tank. Fuel can also be transported from side to
side, either through the center tank and into the other wing, or past the center tank altogether and into the other side.
Another line that can be fed by any of the three tanks goes from the center back to the APU.
Within the fuel systems, there are a number of safety precautions that will be put into place. Check valves will be
placed throughout the system to keep problems in the system isolated. Surge tanks will also be implemented in case of a
decrease in fuel density while the tanks are full. The size of the surge tanks were calculated from a change in density
after a temperature change of 15 °C. Finally, there will also be vent tanks that hold fuel that is about to be jettisoned in
case of an emergency to decrease the weight of the aircraft. The vent tanks were sized off of historical data.
The location of the tanks inside the aircraft are shown in Fig. 58. Going from inboard to outboard, there is the
center tank in the fuselage, the main wing tanks that are split up into two sections, the surge tanks, and the vent tanks.
The two sections of the wing tank function as one tank, but are separated to help prevent major fuel sloshing. Each tank
uses a wet wing approach, because it allows the wing to keep its internal structure, creates baffles from the ribs to help
74
Because the maximum range required of 3,500 nmi is significantly less than any other commercial aircraft of this size,
achieving the necessary fuel volume requirement was not an issue. The fuel volumes of each tank are shown in Table 46.
The tank volumes were calculated using the maximum fuel volume needed at standard temperature and pressure along
with the available space within the aircraft structure. The required fuel volume calculated from performance data is
19,450 gal.
Wing 7,100
Center 5,250
Surge 300
Vent 150
Total 20,350
D. Hydraulics System
The hydraulic system of this aircraft will be used for wheel braking, nose wheel steering, flight control actuation,
and high lift device actuation. Figure 59 shows a basic layout of the hydraulic system layout of the JJJP.
75
The two hydraulic reservoirs on the left and right side of the aircraft are the main reservoirs, and the one in the center
is the standby system.The standby system will only actively be supplying pressure to the rudder of the aircraft along with
thrust reversers when activated. The left and right side systems will be run off power from the engine, whereas the center
standby system will be driven electrically with an option to run it using a ram air turbine in emergencies. The systems
are connected so if one system fails, another can supply emergency power to continue the operations of the aircraft.
Each control surface will also have at least one electric backup hydraulic actuators, which is a self contained system that
would still work if all three of the hydraulic systems were compromised. This creates more redundancy in the system
that leads to a safer hydraulic system overall. It should also be noted that the two main reservoirs do not exclusively
operate the systems on that given side. Some control surfaces are staggered between systems so that all the control
surfaces do not automatically fail when there is a problem. Instead, the other system that is still running gives partial
control to each side of the wing, keeping things balanced until emergency power can be obtained. Also, fail-safes are
put into place that will allow the standby reservoir or the main reservoirs to reroute power to other parts of the system in
the case that one of the other reservoirs goes down. This is managed by shutoff valves that are shown in Fig. 59, along
with other check valves that will be implemented in various places in the hydraulic design. As the design progresses and
more specific needs of the hydraulic system are established, more specifics of the system will be put into place.
E. Electric System
The electrical system generates and provides power to all the electrical systems on the aircraft. These systems
mainly consist of avionics, actuation, flight controls, and sensors. The electrical system is modeled after the Boeing 777
electrical system, because the electrical power requirements are similar based on the design of the other systems in the
aircraft [64]. The power for these systems is taken from one generator on each engine as well as well as a generator on
the APU. These generators provide 120 kVA of power each. Power for the electrical system is stored in one lithium-ion
battery that is kept near the front of the aircraft. This battery deviates from the 777 design, but with increasing battery
technology, a lithium-ion battery should not be an issue for this aircraft by 2029. From historical data, the battery was
estimated to need to weigh 70 lb in order to be big enough to store the necessary power [65]. The battery requires DC
power, so some other equipment must be used to convert from AC to DC and back to AC. To convert from AC to DC, a
Transformer Rectifier Unit can be used. Going back to AC, a power inverter is needed. The power required for the
systems in the aircraft will either run on 28 VDC or 115 VAC [64]. For systems with lower amperage requirements, 28
VDC will be used. Much of the avionics of the aircraft along with the wingtip locking actuators will run on DC. For
higher amperage systems 115 VAC will be used because it can achieve a higher amperage. The power follows a path
from the engines/APU to the front of the aircraft were the electrical equipment bay and battery is. The power is then
distributed throughout the aircraft to all the systems that need it.
76
F. Pneumatic System
The pneumatic system provides air pressure to various parts of the aircraft. It does this by bleeding air out of the
engine or the APU and supplying it to the necessary systems. Those systems include the environmental control system,
deicing system, pressurization, engine starting, and windscreen ice/rain prevention. Figure 60 shows a simple diagram
The environmental control system seeks to keep the passengers comfortable and the rest of the cargo and equipment
at a safe temperature. The environmental control system mainly takes bleed air from the pneumatic system and uses it
take care of things within the aircraft. Its main function is to keep the cabin pressurized and at a comfortable temperature
for the passengers on the aircraft. A basic diagram of what the temperature control system in the aircraft will look like is
shown in Fig 61 [63]. Using this system, the cabin will be pressurized to 8,000 ft in accordance with 14 CFR Part
25.841.
77
Fig. 61 Temperature control diagram [63]
Another large part of the environmental control system is keeping ice off of the aircraft. This aircraft, in order to
meet RFP requirements, must have a deicing system [1]. As previously stated, the deicing power will be provided by
the pneumatic system. Hot air from the engine will be brought to the leading edges of the wings, tail, engines, and
windscreen. This method provides a simple yet effective way of keeping ice from building on critical areas of the
aircraft. Shown in Fig 62 is a diagram of how the system will utilize the hot engine air to prevent ice on the leading edge
78
H. Emergency Systems
The aircraft will be equipped with emergency equipment for passengers, flight attendants, as well as the aircraft
overall. Supplemental oxygen will be built into the aircraft as well as portable oxygen masks with tanks will be put into
the aircraft to fulfill 14 CFR §121.333 in the FAR. Five hand fire extinguisher will be conveniently located throughout
the aircraft in case of a fire. One crash axe and two megaphones will also be installed in the aircraft, all in accordance
with 14 CFR §121.309. In the case of an evacuation of the aircraft, the aircraft will employ approved emergency exit
slides at each exit in accordance with 14 CFR §25.810. Finally, if there is an evacuation, all of this equipment will
help the passengers in order to exit in the allotted 90 seconds as stated by the FAA in 14 CFR §121.291. The aircraft
will also have life preservers, life rafts with appropriately equipped survival kits, and approved locator transmitters
can be installed for over-water operations in accordance with 14 CFR §121.339. For the aircraft itself, fire detectors,
fire suppressants, angle of attack sensors, and other emergency equipment to report the health of the aircraft will be
I. Avionics
Avionics in an aircraft manage communication, navigation, as well as integrate many other systems together in order
to display relevant information to the pilot for apt decision making. It is important to have a comprehensive yet not
overwhelming avionics system so that the pilot has all information and flight data necessary. However they should not
bogged down with unnecessary information. The avionics should also contain redundancy in order to achieve acceptable
safety standards. When looking at suppliers for more modern Boeing and Airbus aircraft, it is clear that they are trying
to achieve this. They never order all their equipment from the same company and have multiple companies supplying
For example, the Boeing 777 is using both Honeywell and Collins Aerospace for the majority of their flight and data
management systems [67]. Honeywell Aerospace supplies Boeing with AIMS, which integrates the multiple functions
of avionics into one single system [67]. Collins Aerospace also provides the B777 with flight control systems that are
used on the aircraft [67]. If they were to only use one avionics manufacturer, they run the risk of the company having
the same problem in different parts of the avionics because the problem was not found. Because these are different
systems created by different companies, they are much less likely to fail at the same time or exhibit the same problem.
The JJJP will be using an Integrated Modular Avionics(IMA) for its Flight Management System(FMS). IMA is
the opposite of the traditional federated architectures, and its concept is very similar to a desktop computer, where
everything gets processed by the centralized processing units. That is, the IMA is essentially a "barebones computer"
that allows for increased customization. The caveat being that all the modules need to be compatible. By having IMA,
the number of line replaceable units(LRU) is significantly lower. For example, Boeing utilized the IMA approach to off
shave 2,000 pounds off the avionics suite of the new Boeing 787 [68]. IMA is an economic solution to avionics because
79
of its flexibility, weight savings, and maintenance savings. Figure 63 demonstrated that Federated system require every
LRU to have its own processor. In contrast, IMA uses a centralized processor, and it will reduce the communication
needed for its systems. Thus, IMA reduces weight and complexity [69]. The following IMA architecture commonly
used by major commercial aircraft were put into comparison: Rockwell Collins Pro Line Fusion, Honeywell Aerospace
Airplane Information Management System (AIMS), and General Electric Common Core System(CCS). The Pro Line
Fusion was commonly seen in smaller aircraft such as the Airbus A220 and the Bombardier Global 5000, and Honeywell
AIMS was used on older Boeing 777. GE CCS was determined to be the JJJP’s avionics integrator, because it was used
on the Boeing 787 Dreamliner as well as the new Boeing 777X. Ultimately, the GE Common Core System was chosen
to be the JJJP’s avionics integrator, because the other two option was simply eliminated. The Pro Line Fusion only
exists in small commercial aircraft and business jets. On the other hand, the Honeywell AIMS was an IMA used in
older Boeing 777. CCS is a promising solution to avionics. According to GE, CCS saves hundreds of pounds of weight
80
For avionics integration, systems of the Boeing 787 Dreamliner was used, which is listed in Table 47 [71]. One
important note here is systems will come with packages. For example, Honeywell navigation package comes with
inertial reference system, air data system, and multi-mode receiver [71]. As well as pitot tube, instrument landing
system (ILS), GPS, GPS landing system (GLS), and VHF omni-directional radio range (VOR) from the Honeywell
Multi-mode receiver. On other hand, the Collins Aeropsace ISS-2100 package offers weather detection, traffic alert, and
collision avoidance. Additionally it has terrain awareness and warning function integrated into a single system [72].
The system was capable of VFR and IFR flight with the navigation package, along with autopilot from Honeywell. In
extreme weather condition, Honeywell flight control offers CAT II approach, windshear guidance, and integrated thrust
management [73].
Communication Communication Antennas Cobham SATCOM HGA-7001 high gain satcom antenna system
Flight and Data Management Flight Management Systems GE Aviation Systems Common Core System (CCS)
Flight and Data Management Operating System Green Hills Software Integrity-178B operating software
Flight and Data Management Avionics Management Systems Collins Aerospace Avionics Management Systems package
Flight and Data Management Flight Control System Honeywell Aerospace Flight Controls and Autopilots
Indicators and Instruments Electronic Flight Instrument Systems Collins Aerospace LCD Displays & Fuel Quantity Indicators
Indicators and Instruments Electronic Flight Instrument Systems Thales Avionics Electronic Flight Instrument Systems
Warning Systems Configurable Integrated Surveillance System Collins Aerospace ISS-2100 package
81
XVIII. Cost Analysis
Two methods were utilized for estimating the research, design, testing, and evaluation (RDTE) cost and flyaway
cost. The first was the RAND DAPCA IV model as specified in Raymer [6]. DAPCA IV is a cost model developed by
the RAND corporation which estimates the amount of hours necessary for RDTE and flyaway (production) in terms
of engineering, tooling, manufacturing, and quality control. The hours are then multiplied by their respective labor
rates, from which the total RDTE and flyaway costs can be determined. The second method is laid out in Roskam VIII
[61]. It follows a similar process of estimating manhours and process costs, but is higher fidelity and uses more aircraft
parameters as inputs.
Certain economic factors were assumed. The DAPCA IV cost model is calibrated to 2012 and the Roskam model is
calibrated to 1989. As such, the respective historical inflation rates were utilized between that time and 2020 based off
the U.S. Bureau of Labor Statistics (BLS) data [74]. From there, a steady 2% inflation rate was assumed between 2020
and 2029. In calculating unit cost, a 15% profit margin was also applied. The final cost estimations are presented in
Table 48.
One result of this estimation is that RDTE cost is fairly constant across various production runs. This is because the
majority of these costs are incurred well before production begins. Flyaway cost increases as expected considering that
the main component is manufacturing. The result of this is that unit cost decreases with higher production runs due to
82
B. Direct Operating Cost
Direct operating costs were estimated utilizing approximations from Raymer [6]. This includes fuel, crew,
maintenance, depreciation, and insurance costs. Fuel costs were based on the fuel burn and mission time for a 700 nmi
flight. For the JJJP, a two-man cockpit crew was assumed. Regarding maintenance, 10 manufacturing manhours per
flight-hour was implemented. The depreciation model was based on a 90% value loss over a 12 year lifespan. The final
C. Unit Cost
Unit cost was estimated utilizing two separate methodologies. The first was using both the DAPCA IV and Roskam
CER. RDTE + Flyaway costs were simply divided by the production quantity to determine a unit cost. The second was
by reviewing historical data. A simple analysis was done of aircraft empty weight vs. unit cost. The data included
aircraft from the Boeing commercial aircraft family (B737, B747, B767, B777) [75]. A linear line of best fit was
calculated, and then a unit cost was determined based off the JJJP empty weight. The slope of the line of best fit is
83
about $914 per pound. This matches a rough rule-of-thumb quoted in Raymer of about $800 per pound [6]. The data is
A summary of all results are presented in Table 50. The unit costs from the DAPCA IV and Roskam method are
lower than that of the historical data analysis. Sources of error for this estimate are discussed in the model uncertainties
section.
84
D. Cost Reduction Methods
The main design objectives of this aircraft are to minimize manufacturing and operating cost. As such, a number of
cost reduction methods have been considered. One design decision was the choice of material for the airframe. The 3
main options were a fully aluminum structure, a fully composite structure, and a partial aluminum-composite structure
(B777X). Composite frames provide a higher fuel efficiency which is beneficial for long range missions, however there
are diminishing returns as range decreases. Simply from a manufacturing and maintenance standpoint, it was decided to
Another set of methods are centered around the passenger experience. Recently, many low-cost carriers have
been able to reduce operating costs while maintaining a consumer base. One example is installing Wi-Fi for in-flight
entertainment as opposed to in-seat entertainment. This would benefit the customer and result in a lower systems
weight/cost, however the antenna and radome necessary for this increases the fuselage skin friction drag. Another
method is choosing lightweight seats with lower padding compared to normal seats. The shorter 700 nmi mission time
One cost reduction method that was explored was a windowless airframe. This concept is fairly simple. The
removal of all cabin windows results in significant loss in structural weight. To maintain passenger comfort, high quality
television panels would be setup throughout the cabin to visualize the exterior and replicate windows. A preliminary
sizing paper by the University of Bologna estimated about 3,200 lb for a Boeing 777 type aircraft [76]. This accounts
for structural weight savings and electronics weight gains. The implementation of a windowless airframe would result
in reduced fuel burn and manufacturing costs but slightly higher operating and maintenance costs. The overall fuel cost
and emissions savings for an airline would be significant. However, the ticket price savings for a passenger would be
approximately $ 6 for a 700 nmi mission. It was then assessed that the average consumer would likely not find this
tradeoff worth the potential discomfort of flight, and the concept was not further pursued.
E. Model Uncertainties
The models used in this section are fairly low fidelity and are subject to many sources of error propagation. One is
lack of knowledge in the pricing models of third-party suppliers. Pricing of specific onboard auxiliary systems and
avionics modules are not directly known. Avionics cost was calculated using a weight based estimation of $ 4,500 per
pound as suggested by Raymer [6]. However, the JJJP does not use systems or avionics that stray significantly from the
This cost model is also unable to account for possible technical delays during the development phase which would
result in a higher RDTE cost than predicted. The majority of the aircraft uses design features with high technology
readiness levels (aluminum airframe, folding wingtips, FBW system). However, the UltraFan engine utilized will push
technological boundaries in terms of blade design and efficiency. As such, potential time delays and budget overruns
85
could come about.
Finally, this model is not able to account for pure economic factors that are present at large firms. For example,
while a unit cost can be estimated, heavy discounts are often given to airlines that purchase bulk orders. Also, financial
factors such as interest rates, profit margins, and inflation are realistically more complex than modeled in this report.
Labor in the form of engineering, manufacturing, and support is not subject to constant hourly rates as specified in
DAPCA IV. For example, airframe manufacturing in the United States is heavily influenced by labor union agreements.
A. Acoustics
When considering the acoustic profile of the aircraft there are three locations at which sound levels are measured with
limits set by the size and year of certification. The three measurements taken are lateral, flyover, and approach. Lateral
is the full power reference noise measurement point at a distance of 1,476 ft from the center line of the runway. The
flyover measurement is taken at a distance of 21,325 ft from the start of takeoff roll, while approach is the measurement
taken at a distance of 6,561 ft from the centerline of the runway threshold [77]. For the JJJP, the noise levels allowed
depend on the maximum takeoff weight as well as time of certification. The exact levels of each stage were found by
From this the noise levels that need to be met at each limit were determined to be 97 EPNdB for flyover, 101 EPNdB
for lateral, and 103 EPNdB for approach. To know whether or not these levels can be met, acoustic data was taken from
the European Aviation Safety Association (EASA) database[79]. Information regarding the noise measured at each
86
location along with the exact engine model used was gathered. Following the trends observed from this information
would then allow for an estimation to be made of the acoustic profile of the JJJP, with the result shown in 66.
(a) Cumulative noise level estimate (b) Thrust effects on lateral noise levels
(c) Thrust effects on flyover noise levels (d) Thrust effects on approach noise levels
Fig. 66 Acoustics data showing effects of thrust on the noised produced at each stage
From this, the effects of thrust can be seen in the lateral section, with a trend of increased noise with an increase in
thrust as this is the position where max thrust must be used. When using this data to see where the JJJP will be in terms
of noise produced, the plane is expected to use 62,500 lb of thrust per engine during takeoff and following the trend that
puts it near 86-88 EPNdB. This range is well under the 101 EPNdB allowed for a stage 5 aircraft. For flyover, the results
are somewhat less clear, as it is very common for aircraft to cut back on power when approaching this location in order
87
to meet the noise requirement, thus thrust has less of an effect on the noise seen at this location [77]. If assuming worst
case scenario base on this data, the maximum noise produced by the engine to be used on this aircraft would range
between 81-86 EPNdB which is well below the limit of 97 EPNdB allowed. At approach it can be seen that thrust has
almost no effect on the noise produced as during this stage the engines are most often near idle producing very little
thrust and noise . The outliers in this data may be due to airframe noises that con sometimes be unavoided due to the
construction of the aircraft or aborted landings requiring an increasing in thrust and thus noise produced . By following
the trends of other aircraft, the noise produced for this aircraft can be estimated at 94-95 EPNdB that is still under the
There are some uncertainties when creating a model like this based purely on historical data of engines. This is even
more the case in this situation as the JJJP is planning on using the Rolls Royce UltraFan engine which currently has not
yet entered service, so there is no acoustics data available for this exact model. In order to estimate the noise being
produced, different variations of the Rolls Royce Trent 1000 were investigated as this is the closest engine on the market
at this time to the UltraFan. Although, since the Ultrafan does have a bypass ratio much larger than most engines on the
market today, these estimates for the acoustic profile are very conservative. With a low bypass ratio engine, the exit
velocity of the air through the engine is much lower resulting in a more quiet and less turbulent exhaust overall. This
reduces the noise produced when compared to an engine with similar thrust but lower bypass ratio. Another uncertainty
comes with construction of the aircraft and noises that it may make when flying that are very hard to model and predict
B. Emissions
In terms of environmental concerns, the emissions profile of this aircraft is of upmost importance, especially with
trends in political and social movements demanding for more consideration of the effects all industries have on the
environment. One of the added benefits of having such an efficient aircraft and engine such as the JJJP and Rolls Royce
UltraFan is the increased performance allows for less fuel to be burned per mile than some older planes on the market
with outdated technology. To get an accurate estimation of the total emissions to be expected from this aircraft, the
total flight hours per year ranged from 2500 to 4500 hours and an average life span of 25 years was predicted from this
aircraft to get an idea of the total emissions produced from flight of this aircraft. The results of this can be seen below in
Table 51.
88
The results in Table 51 not only include the emissions from the total flight hours, but also a fixed amount associated
with the emissions of manufacturing. The production of metal is a very high energy process, thus there are usually high
amounts of emissions generated from it. Since this aircraft is largely comprised of aluminum, it was estimated that
for every pound of aluminum, 9.2 pounds of 𝐶𝑂 2 are produced [80]. This estimate can be further refined if specific
manufacturing processes are investigated deeper as well as looking into using recycled materials to further reduce the
XX. Conclusion
A new type of high capacity short range aircraft is necessary to adjust for a developing global airline market.
The AIAA RFP outlines the requirements for a potential aircraft with complementary design objectives. The major
constraints are a 400 passenger capacity and 3,500 nmi range. Design objectives are to minimize operational and
production cost. Given this problem statement, the JJJP was designed.
The JJJP features a conventional tube-and-wing design. It can carry a nominal payload of 93,400 lb to a maximum
range of 3,500 nmi at a cruise speed of Mach 0.78. A high aspect ratio wing and UltraFan engines ensure high efficiency
at an affordable cost. The 7 ft folding wingtips allow it to operate out of Group IV airports. The MTOW and empty
weight are 443,000 lb and 241,500 lb respectively. Entry into service is expected in 2029 at a unit cost of $ 180 million.
JJJP is able to meet all RFP requirements and all major FAA requirements.
Based upon all assessments in this report, the JJJP design does not contain any flaws that would require revision. As
89
XXI. References
[1] “Request for Proposal High Capacity Short Range Transport Aircraft,” , 2020. URL https://www.aiaa.
org/docs/default-source/uploadedfiles/education-and-careers/university-students/design-
competitions/undergraduate-team-aircraft-design-competition/undergraduate-aircraft-high-
capacity-short-range-transport-aircraft.pdf?sfvrsn=b6081273_0.
[2] “Boeing 777 Specs, what makes this giant twin work?” , 2020. URL http://www.modernairliners.com/boeing-
777/boeing-777-specs/.
787-dreamliner-specs/.
[6] Raymer, D. P., Aircraft design: a conceptual approach, American Institute of Aeronautics and Astronautics, 1989.
[7] Roskam, J., Layout Design of Cockpit, Fuselage, Wing, and Empennage: Cutaways and Inboard Profiles, DARcorporation,
1997.
13A-chg1-interactive-201907.pdf.
[9] Dix-Colony, K., “Operating the 747-8 at Existing Airports,” AERO, 2010. URL https://www.boeing.com/commercial/
aeromagazine/articles/2010_q3/3/.
[10] Wall, M., “Stratolaunch to launch hypersonic vehicles from world’s biggest airplane,” 2020. URL https://www.space.com/
stratolaunch-hypersonic-vehicles-worlds-biggest-airplane.html.
[11] Morales, J., “The A380 Transport Project and Logistics,” , 2006.
[13] Street, F., “Are windowless planes the future of travel?” 2019.
1.pdf.
[16] “TP400-D6 ENGINE – THE MOST POWERFUL TURBOPROP ENGINE IN PRODUCTION,” , 2020. URL http:
//www.europrop-int.com/the-tp400-d6/.
90
[17] Mark, “Elasticity of Demand for Air Travel,” 2015. URL https://econfix.wordpress.com/2015/05/07/elasticity-
of-demand-for-air-travel/.
[18] NTSB, “Aircraft Accident Report AA191 Chicago O’Hare May 25 1979,” , 1979.
TCDS%20IM%20E%20102_issue10_20191213.pdf.
[20] “EASA type-certificate data sheet Trent 1000/7000,” , 2020. URL https://www.easa.europa.eu/sites/default/
files/dfu/TCDS%20E%20036%20issue%2016_20191105.pdf.
[21] Kjelgaard, C., “Rolls-Royce UltraFan Design Frozen,” 2018. URL https://www.ainonline.com/aviation-news/air-
transport/2018-07-16/rolls-royce-ultrafan-design-frozen.
[22] O’Connor, K., “Rolls-Royce Begins Building UltraFan Demonstrator,” 2020. URL https://www.avweb.com/ownership/
engines/rolls-royce-begins-building-ultrafan-demonstrator/.
[23] “EASA type-certificate data sheet Trent XWB,” , 2020. URL https://www.easa.europa.eu/sites/default/files/
dfu/EASA%20E111%20TCDS%20issue%2012_Trent%20XWB.pdf.
EASA%20TCDS%20IM%20E%20002_GE90%20series_Issue4_Final_18Dec2019.pdf.
[25] Bensel, A., “Characteristics of the Specific Fuel Consumption for Jet Engines,” 2018. URL https://slack-
files.com/files-pri-safe/TST914E8Z-FTU9VALSF/characteristics_of_the_specific_fuel_consumption.
pdf?c=1583735532-9a0b3b403a89e214.
[26] Dimitriadis, G., and Léonard, O., “Aircraft Design: Aircraft Propulsion,” 2018. URL http://www.ltas-cm3.ulg.ac.be/
AERO0023-1/ConceptionAeroTurbomachine.pdf.
[28] “Boeing 777 Specs, what makes this giant twin work?” , 2017. URL http://www.modernairliners.com/boeing-
777/boeing-777-specs/.
[29] Roskam, J., Airplane Design Airplane Design Part VI: Preliminary Calculation of Aerodynamic Thrust and Power Characteristics,
DARcorporation, 1997.
[30] Rodrigo Martinez-Val, E. P., and Palacin, J., “Historical Perspective of Air Transport Productivity and Efficiency,” 2012.
[31] Rudolph, P. K. C., High-Lift Systems on Commercial Subsonic Airliners, National Aeronautics and Space Administration,
September 1996.
[32] Harry L. Morgan, J., Experimental Test Results of Energy Efficient Transport (EET) High-Lift Airfoil in Langley Low-Turbulence
[34] Nicolai, L. M., and Carichner, G. E., Fundamentals of Aircraft and Airship Design, Volume 1 – Aircraft Design, AIAA, 2010.
[35] Hamilton, S., and Fehrm, B., “737 MAX 8 could be enabler for some LCC Long Haul,” 2014. URL https://leehamnews.
com/2014/12/08/737-max-8-could-be-enabler-for-some-lcc-long-haul/.
[36] Bhaskara, V., “ANALYSIS: A320neo vs. 737 MAX: Airbus is Leading (Slightly) – Part II,” 2016. URL https://web.archive.
org/web/20160206082857/http://airwaysnews.com/blog/2016/02/05/a320neo-vs-737-max-pt-ii/.
[37] “Data A: Aircraft Data File Table 4: Boeing Aircraft,” , 2001. URL https://booksite.elsevier.com/9780340741528/
appendices/data-a/table-4/table.htm.
[38] “Boeing 787 -8 (Dreamliner) sample analysis. (2005),” , 2006. URL http://www.lissys.demon.co.uk/samp1/index.
html.
[39] Torenbeek, E., Advanced Aircraft Design: Conceptual Design, Analysis and Optimization of Subsonic Civil Airplanes, John
[41] Sadraey, M. H., Aircraft Design A Systems Engineering Approach, John Wiley Sons, Ltd, 2013.
[42] McCormick, B. W., Aerodynamics, Aeronautics, and Flight Mechanics, John Wiley & Sons, Inc, 1995.
[45] Roskam, J., Airplane Design Part V: Component Weight Estimation, DARcorporation, 1997.
[46] Jordan, K. B., Care and Repair of Advanced Composites, SAE International, 1998.
GroupID=178.
[48] “AISI 4340 Steel, normalized, 100 mm (4 in.) round,” , 1996. URL http://asm.matweb.com/search/SpecificMaterial.
asp?bassnum=M434AE.
[52] “Pushing the A350 XWB to the brink,” , 2013. URL https://www.youtube.com/watch?v=B74_w3Ar9nI.
92
[53] “Thin Walled Vessel,” , 2006. URL http://www.eng.fsu.edu/~kalu/ema4225/lec_notes/Web%20Class_6_final.
[54] “Lissys Ltd opens up its Boeing 787 analytic performance tool,” 2008. URL http://www.lissys.uk/boeing787.html.
[55] Torenbeek, E., Synthesis of Subsonic Airplane Design, Kluwer Academic Publishers, 1982.
[56] Roskam, J., Airplane Design Airplane Design Part VI: Layout Design of Landing Gear Systems, DARcorporation, 1997.
[57] Roskam, J., Preliminary Configuration Design and Integration of of the propulsion system, DARcorporation, 1997.
[59] “Hydraulic Rotary Actuator - LTR Series (light duty),” , 2017. URL https://ph.parker.com/us/en/hydraulic-rotary-
actuator-ltr-series-light-duty.
[60] “Actuation and Motion Systems Product Guide,” , 2010. URL https://www.moog.com/content/dam/moog/literature/
MCG/actprodguide.pdf.
[61] Roskam, J., Airplane Cost Estimation: Design, Development, Manufacturing and Operating, DARcorporation, 1997.
[62] Bartley, G. F., “Boeing B-777: Fly-By-Wire Flight Controls,” 2001. URL https://www.davi.ws/avionics/
TheAvionicsHandbook_Cap_11.pdf.
[63] Moir, I., and Seabridge, A., Aircraft Systems: Mechanical, Electrical, and Avionics Subsystems Integration, John Wiley & Sons
Ltd, 2008.
[64] Andrade, L., and Tenning, C., “Design of Boeing 777 electric system,” Aerospace and Electronic Systems Magazine, IEEE,
Batteries-and-Advanced-Airplanes.
[68] Ramsey, J. W., “Integrated Modular Avionics: Less is More,” 2007. URL https://www.aviationtoday.com/2007/02/
01/integrated-modular-avionics-less-is-more/.
[69] Watkins, C., Integrated Modular Avionics: Managing the Allocation of Shared Intersystem Resources, AIAA, 2006.
[70] Wagenen, J. V., “GE Aviation to Bring Common Core System to Boeing 777X,” 2014. URL https://www.aviationtoday.
com/2014/12/22/ge-aviation-to-bring-common-core-system-to-boeing-777x/.
93
[72] “ISS-2100 Configurable Integrated Surveillance System,” , 2020. URL https://www.collinsaerospace.com/
en/what-we-do/Commercial-Aviation/Flight-Deck/Surveillance/Integrated-Surveillance/Iss-2100-
Configurable-Integrated-Surveillance-System.
systems-and-displays/flight-controls-and-autopilots.
bca/#/prices.
[76] Bagassi, S., PRELIMINARY DESIGN OF A LONG RANGE WINDOWLESS AIRCRAFT CONCEPT, *Department of Industrial
Certification-Workshop-2006/Boettcher_3.pdf.
21092/stage-5-airplane-noise-standards.
certification-noise-levels.
[80] “Amounts of CO2 Released when Making Using Products,” , 2012. URL http://www.co2list.org/files/carbon.htm#
RANGE!A83.
94