TAMU SAE Report

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 30

Texas A&M Aero Design Team

2016 SAE AERO DESIGN EAST

DESIGN REPORT

January 25, 2016

Texas A&M University – College Station, TX

Team 022
2
TABLE OF CONTENTS

1. Objective and Requirements ........................................................................................................ 4

2. Conceptual Design ....................................................................................................................... 5

2.1 Wing Configuration .................................................................................................................. 5

2.2 Fuselage Concept ...................................................................................................................... 6

2.3 Stability and Control ................................................................................................................. 6

2.4 Payload Support Structure ........................................................................................................ 7

3. Design and Analysis ..................................................................................................................... 7

3.1 Wing Design ............................................................................................................................. 7

3.2 Fuselage and Tail Design ....................................................................................................... 13

3.3 Payload Support Structure Design .......................................................................................... 16

3.4 Stability and Control ............................................................................................................... 16

3.5 Servo Selection ....................................................................................................................... 21

3.6 Propulsion System Design ...................................................................................................... 21

3.7 Takeoff Performance .............................................................................................................. 22

4. Experimental Data Collection .................................................................................................... 23

4.1 Motor and Propeller Testing ................................................................................................... 23

4.2 Material Properties ................................................................................................................. 24

5. Construction ............................................................................................................................... 25

6. Competition Flight Plan ............................................................................................................. 25

7. Conclusion.................................................................................................................................. 26

References ................................................................................................................................................. 28

3
1. OBJECTIVE AND REQUIREMENTS

The goal of the competition is to design, build, and fly a radio controlled aircraft capable of lifting

the maximum possible payload within the specified design constraints laid out by the competition. The

most critical design constraints include overall dimension specifications, allowable materials, takeoff

distance, and power plant limitations as listed in Table 1. In order to satisfy the requirements of the

competition, we concentrated (in order of priority) on maximum lift-to-drag ratio, maximum possible

lift, adequate controllability, and minimum aircraft empty weight. These considerations were used to

formulate trade-offs and rigorous optimization problems to manage the large number of parameters

dictated in aircraft design. This report carries the theme of optimization based on the aforementioned

considerations throughout the various discussions. Most of the analyses were performed in four steps:

definition of quantitative objectives and requirements based on a literature survey; compilation of the

equations and modeling required to determine the effect of parameters on meeting the requirements;

iterations to find the proper parameter values; and finally, testing to validate the entire process. This

report briefly addresses each of these issues on the topics of aerodynamics, structures, stability and

control, and propulsion.

Aircraft Requirements Description


Dimensions Maximum combined height, length, and width of 175 inches.
Materials No fiber-reinforced plastic components.
Takeoff Distance Maximum 200 feet.
Landing Distance Maximum 400 feet in the same direction of takeoff.
Maximum weight The maximum weight may not exceed 55 pounds with payload.
Must be able to securely load and unload payload within 1 minute each. Payload must
Payload Support
also be contained within a 4 in. x 4 in. x 10 in. closed payload bay.
Power Plant (motor) Engine must be a single, unmodified electric motor with a 1000 W limiter.
Landing Condition No parts may fall off the plane (except the propeller).
Battery 6 cell (22.2 volt) Lithium Polymer Battery.
Table 1: Design Requirements

4
2. CONCEPTUAL DESIGN

Wing Configuration

The conceptual design of the wing configuration consists of the wing’s shape and characteristics,

e.g., airfoil profile, winglets, taper, ailerons, and flaps. Each design parameter was optimized to ensure

considerable lift, reduced drag, minimal structural weight, and desired controllability of the aircraft.

Various wing configurations are illustrated in Figure 1, exhibiting their optimal lift distributions.

A rectangular wing offers the greatest lift while “creating trailing vortices which cause an additional drag

force on the wing” [1]. Theoretically, elliptical wings are the most efficient planform, since by the

reduction of induced drag, the kinetic energy of the trailing vortices is spread out along the span of the

wing, instead of being concentrated at the wing tips [1]. However, due to the elliptical wing’s undesirable

stall characteristics and the increased difficulty on the manufacturability of a wing that shape, the design

team selected to build a tapered wing instead. The tapered wing provides the same efficiency comparable

to the elliptical wing but without stalling simultaneously across the entire wing span. We chose to have a

moderate taper on both the rear and front wings to receive efficiency benefits while maintaining high lift.

One benefit of the selection of the tandem wing aircraft is that the design team was able to optimize

different airfoils for both wings. This variation ensures that the front wing stalls before the rear wing, thus

allowing the aircraft to keep control stability. We established that the rear wing was to be placed high up

on the vertical tail to guarantee that it is outside of the downwash of the front wing. To counterbalance the

stability effects of the high mounted rear wing, the team decided to have a low placement of the front wing

on the fuselage. To reduce induced drag and structural weight, we removed attachment hardware from the

airflow which was made possible by the use of a cantilevered wing.

Elliptical Rectangular Moderate Taper High Taper Pointed Tip Swept

Figure 1: Example of ideal aircraft lift distributions.

5
Fuselage Concept

The team designed the fuselage structure to contain the payload and abide by the weight,

dimensional, and aerodynamic constraints while incorporating a factor of safety of 3. In order to meet the

SAE Aero Design payload dimension requirements, a rectangular shape was chosen for the fuselage. In

order to increase aerodynamic efficiency while maintaining the structural integrity of the fuselage, drag-

causing frontal and overall surface area of the fuselage were minimized. To accommodate for motor

mounting and takeoff rotation angle requirements, different regions of the fuselage were tapered.

Stability and Control

The design process began by exploring the use of three different high lift configurations: the

tandem wing, the w-tail, and the flying wing. The team’s focus was to achieve desired stability, while

also gaining a competitive lifting capacity for the competition using hand calculations as well as XFLR5

for confirmation. After review of each configuration, it was decided to build a half-scale tandem

prototype, as seen in Figure 2, in order to show proof of concept. From literature on a general aviation

aircraft [6, 7], we determined the required ranges for the stability and control coefficients as shown in

Table 2. Further discussion about each coefficient shown can be found in Section 3.4. Eight successful

flight tests of the half-scale tandem prototype demonstrated our desired stability and control

characteristics.

Sign Minimum Maximum Mean


CLα + 4.5 5.73 5.2775
Cmα - -1.26 -0.683 -0.90767
Cmq - -35.6 -10 -22.12
Cyβ - -0.96 -0.362 -0.684
Clβ - -0.221 -0.074 -0.2965
Cnβ + 0.071 0.28 0.1505
Clp - -0.53 -0.41 -0.46333
Cnr - -0.3 -0.125 -0.20433
Figure 2: Half-Scale Tandem Prototype Clδa - -0.134 0.0461 -0.04395
Cnδr - -0.072 -0.109 -0.0905
Cmδe - -0.923 -1.34 -1.1315
Table 2: Ideal aircraft stability and control coefficient ranges.

6
Payload Support Structure

The philosophy dictating the placement and design of the internal payload support structure

emphasized minimizing weight, optimizing structural integrity, and improving accessibility to the internal

payload. It also allowed for strategic placement of the C.G. such that it is toward the forward part of the

fuselage for optimal stability and control. The majority of the payload box is supported by additional spars

that are attached to the fuselage while a small section is supported by the rear spars of the front wing as

seen in Figure 3.

Figure 3: a) Solidworks Model Depicting Payload Bay Placement in Fuselage


b) Solidworks Model Depicting Payload Bay Placement in Regards to the Front Wing

3. DESIGN AND ANALYSIS

After the team developed the conceptual design and identified the design parameters, we

determined the analysis methods for each subsystem of the aircraft. These analyses were performed in

an iterative fashion to identify the optimal parameter values for all features of the aircraft.

Wing Design

Airfoil Selection

The design team used the Xfoil plug-in within XFLR5 to test over 20 high lift, low Reynolds

number airfoils from the University of Illinois at Urbana Champaign’s airfoil database [16]. We tested

the airfoils on a Reynolds range from 25,000 to 300,000 at an angle of attack sweep from -5° to

15°. The Reynolds range was focused on atmospheric data of weather from Fort Worth during March,

with a large range to provide security in analysis. The results were compared and can be seen in the

Figure 4, which displays the data for the E423 and Figure 5, which displays the data for the Saint CYR

Bartel 24. As it can be seen in the figures, all of the airfoils provide high lift and efficiency numbers.

7
The team then collected the factors of efficiency, maximum lift coefficient, maximum drag coefficient,

manufacturability, and stalling characteristics for each airfoil. We formulated a basic optimization

process so that the parameters of lift, drag, and efficiency were normalized and then added as shown by

Equation 1.

𝐽 = −𝐶𝑑𝑟𝑎𝑔 (𝑡ℎ𝑟𝑢𝑠𝑡 − 𝑑𝑟𝑎𝑔)2 − 𝐶𝑙𝑖𝑓𝑡 (𝑤𝑒𝑖𝑔ℎ𝑡𝑚𝑎𝑥 − 𝑙𝑖𝑓𝑡)2 + 𝐶𝑒𝑓𝑓 (𝑒𝑓𝑓𝑖𝑐𝑖𝑒𝑛𝑐𝑦)2 (1)

where Cdrag, Clift, and Ceff are the normalizing coefficients. From these comparisons, as well as analyses

showing that it would be necessary for the front wing to produce 65% of the lift while the rear wing

supported the remaining 35%, we selected the E423 for the front wing and the Saint CYR Bartel 24 for

the rear wing. Aside from the lifting requirements, we chose the E423 because it would stall before the

back wing, which has a high stall angle. In doing so, we ensured that both of our wings would never stall

at the same time, protecting our pilot’s ability to retain control of the aircraft during every phase of

flight. In addition to the wings, the rudder fins were selected to have the NACA 0008 airfoils in order to

reduce drag and minimize flow separation.

Figure 4: E423: a) L/D vs. Alpha; b) Cl vs. Alpha

Figure 5: St. CYR 24 (Bartel 35-IIC): a) L/D vs. Alpha; b) Cl vs. Alpha

8
Wing Geometry

The same airfoil shape is sustained throughout both the front and rear wings, the E423 and the

Saint CYR Bartel 24, respectively. Using a combination of XFLR5 and Star CCM+, the designing team

was able to optimize the geometry of the wing through simulations.

Each wing has an incidence angle calculated to give high efficiency at cruise and provide

favorable stall characteristics. This optimization was done using XFLR5 and the Xfoil plug-in.

Efficiency versus angle of attack data was compared from which the front wing was determined to have

an incidence angle based on maximum efficiency at cruise. Likewise, the rear wing incidence angle was

calculated to allow for rear wing stall to occur after the front wing. After calculations, the incidence

angles were selected to be 6° and 2.1° for the front and rear wings, respectively. Because of the low

incidence angle on the rear wing, which decreases efficiency at cruise speeds, the team chose a high

aspect ratio for both front and rear wings. This decision subsequently improved both overall efficiency

and stall characteristics.

The design team then analyzed root-tip chord ratios from simulations in Star CCM+ to optimize

for efficiency as well as maximum lift. Due to dimensional constraints, we determined that the root-tip

chord for the front wing to be 16” at the root and 13.5” at the tip. In addition, we decided that the rear

wing be 11” at the root, tapering to 8” at the tip. To optimize the aerodynamic center, the front wing was

tapered linearly on the trailing edge while the back wing was tapered linearly on the leading edge. The

rear wing chord changes after 4” from the center of the plane to allow it to be attached to the vertical

control fins.

After additional analysis, we concluded that neither wing would feature aerodynamic twist. This

decision came from the benefit of favorable stall characteristics created by the incidence angles and

aspect ratios of the wings, thus the aerodynamic twist would add manufacturing complications without

necessary returns in controllability and lift magnitudes.

Winglets were added to increase the effective span of the aircraft by reducing wing tip vortices

as seen in Figure 6. The team ran Star CCM+ simulations on half wing models and provided pressure

9
distributions for a full filleted winglet as seen in Figure 7. From this pressure distribution it was

determined that a linearly sloping winglet on the top portion of the wing would be ideal for preventing

vortices. The lower portion of the wing illustrated little change in pressure distribution.

Figure 6: Half Wing Model Wingtip Velocity Magnitudes

Figure 7: Half Wing Model of a) Upper and b) Lower Pressure Distributions

Structural Analysis

The engineering rationale behind all structural components in our design takes into account

structural theory, experience from past designs, and manufacturability. The wing structures were analyzed

using basic Euler-Bernoulli beam theory and aerodynamic data from prototype flight testing. Finite

Element Analysis was not used because of the inaccuracies of modeling anisotropic materials.

The main structure of both wings consist of a six spar system tapered in a trapezoidal array. Four

main load-bearing spars run along the wing laterally, while two smaller spars run along the trailing edges

of the wings to support control surfaces as seen in Figure 8. The top and bottom of the forward two-thirds

of the front wing is skinned with 1/16” balsa for increased flexural stiffness and airfoil definition. The rear

wing is fully skinned with 1/16” balsa for increased structural stiffness to reduce wing flutter which was

observed in the prototype testing. Moreover, additional webbing helped build torsion boxes to increase

10
torsional stiffness with minimal increase in weight. As previously stated, the front wing was designed to

lift 65% of the total lift and the rear wing the remaining 35%. A 2.5G turn was used in our analyses as the

worst case scenario because this was the maximum G loading seen in our prototype’s flight testing. A

factor of safety of 2 was used to account mainly for ground handling and rough landings. Thus, considering

a maximum takeoff weight of 55 pounds, as denoted by the competition rules, in a 2.5 G turn with a factor

of safety of 2, while assuming an elliptical lift distribution, the geometry of the basswood spars in the

front wing was then determined to support a total lift of 178.75 pounds. The geometry of the rear wing

spars was determined using the same method, except with a total lift of 96.25 pounds.

In order to analyze the second moment of area, the wing was modeled as a cantilevered beam. To

calculate the span-wise distributed lifting load, Equation 2 was utilized as shown below,

4𝑚𝑔 4𝑦 2 (2)
𝑤(𝑦) = √1− 2
𝑏𝜋 𝑏

where mg is the effective weight of the aircraft, b is the wing area, and y is the distance along the span

from the wing root [11]. Next, using the Euler-Bernoulli beam theory, the bending moment equation

reduces to,

1
I*  ymax M z (x) (3)
 max

where I* is the required second moment of inertia (value to be solved for), σmax is the maximum allowable

shear stress for basswood as dictated in Section 4.2, and ymax is the maximum distance in the vertical

direction, which in this case is half of the maximum chord thickness. Mz(x) is the bending moment at a

specific location along the span, which can be solved for at any distance along the span by using the

integral of Equation 2.

In order to mimic the required decaying second moment of area curve, as seen in Figure 9, and

minimize weight while taking into account manufacturability, we decreased the spars’ thickness at a single

location along each half-span. The spar thickness at every point exceeded the minimum required thickness

to maintain a FOS of 2 in a 2.5G maneuver.

11
To determine thickness, placement, and number of rib structures needed in the wing, we calculated

the distribution. From this, we design and placed the ribs so that each rib would support an equal pressure

caused by the lift [4]. The exact rib spacing for the front and back wing is shown in Figure 10.

Figure 8: Solidworks model depicting front wing and cross-sectional view of spar placements.

Figure 9: Plots depicting the required and actual second moment of area curves for the a) front and b) rear wings.

Figure 10: Rib Spacing [in.] for the a) Front and b) Rear Wing

Wing Design Results

After the final iteration, the front wing was determined to have a root chord of 16”, a tip chord of

13.5”, and a 98” wing span while the final iteration of the back wing determined a root chord of 11”, a tip

chord of 8”, and a 98” wing span. Both wings are tapered and do not have aerodynamic twist or dihedral.

12
Fuselage and Tail Design

Aerodynamic Considerations

The fuselage is the largest source of aerodynamic inefficiency because it does not act as an efficient

lifting surface but instead produces substantial amounts of viscous drag caused by its large surface area.

Because of the required dimensions of the payload bay, we focused on creating a fuselage with the smallest

possible form drag. One reduction of form drag we utilized was ensuring the nose of the aircraft is no

larger than the motor. In addition, it was decided that it was necessary to test a rounded edge and un-

rounded edge design. However, the drag reduction using a rounded body design was negligible and

ultimately not worth the construction difficulty involved.

Structural Analysis

As a result of the chosen tandem wing design, a relatively large amount of force will be transferred

through the fuselage body to the payload bay. In the past, wooden fuselages were found to be efficient for

single wing designs where the wing would be joined through a plywood plate to the payload. After

modeling, it was found that an aluminum fuselage could handle this increased loading while weighing

much less than wooden structures designed for the same conditions. Because of the fuselage being

aluminum, all fuselage components were able to be analyzed with Finite Element Analysis because they

could be considered isotropic. Analyzing a wooden fuselage this way would have been incorrect due to

the anisotropic nature of the material. Another major benefit was the consistent strength throughout the

fuselage. This allowed for a wider range of positions to mount the landing gear, the payload bay, servos,

and other hardware.

The fuselage structure is made up of an aluminum sheet that is bent to a shell, shown in Figure 11.

With all of the structure on the external surface, flexural stiffness and torsional rigidity are maximized

within the structure. The aluminum exterior is the primary structural element of the fuselage. Other

structures within the fuselage include aluminum ribs and cross-members that are used solely for the

payload support structure and for mounting avionics and servos. The geometry of this fuselage could not

be modeled as a beam because its width to length ratio is too high. Therefore, Euler-Bernoulli beam theory

13
could not be correctly applied to the design of the structure. Instead, the team analyzed the fuselage using

Finite Element Analysis in Abaqus.

Structural failure of our fuselage occurs in two instances. First, if the material reaches its yield

stress of 345 MPa and second, if the deflection of the fuselage structure causes flight failure. During flight

testing of previous aircrafts, G-loadings of 2.5 were encountered during turns. It was decided to build this

aircraft capable of handling 6 G’s of loading with a factor of safety of 3 to ensure structural integrity

during the flight, ground handling, and landing of the aircraft.

The fuselage analysis consisted of five main tests to model the loading situations the fuselage will

likely encounter. The structure was tested under: full rudder deflection, full aileron deflection, full elevator

deflection, steady level flight, and a landing situation. Using the results from these analyses, it was easily

observed where the structure would encounter the most stresses and what parts of the structure transferred

the most load. With this data, we were able to cut out minimal load bearing sections of the fuselage which

would reduce the overall empty weight of the aircraft, ultimately allowing for a higher maximum payload.

The results of these studies can be viewed in Figure 12.

Material was cut from the fuselage shell until analysis showed the presence of deflections that

would aerodynamically hinder the aircraft’s performance, as well as keeping with the factor of safety of

3 implemented for the fuselage. Using this process, a final fuselage design was constructed as shown in

Figure 11-b.

Figure 11: Solidworks model depicting a) original fuselage outline and b) final fuselage structure.

14
Figure 12: Abaqus model depicting Finite Element Analysis of load testing for the fuselage.

Top Left: Steady Flight; Top Right: Rudder Deflection Test; Bottom Left: Aileron Deflection Test; Bottom Right: Landing Loads.
All deformations scaled for viewing

Landing Gear

The team designed the main landing gear strut to absorb most of the landing loads, choosing a

tapered parabolic-like shape in order to evenly distribute stress propagations over the span of the landing

gear. Since the ideal parabolic-shaped landing gear design is difficult to fabricate using available

resources, we used small bends to imitate the ideal shape as closely as possible. A forward sweep was

added to the landing gear to provide a wider range for balance of center of mass during landing and takeoff

[6]. This sweep helped prevent tipping of the plane as a result of any moment caused by a shift in the

center of mass. The team then modeled the landing gear using a dynamic Finite Element Analysis drop

test with Solidworks as seen in Figure 13. Using these results and past experience we were able to design

a landing gear that would maximize weight savings while providing the needed structural strength to

sustain a 3G landing.

Figure 13: Solidworks model depicting a Finite Element Analysis drop test for the landing gear.

15
Payload Support Structure Design

The main factors the team considered when designing the internal payload bay, in order of priority,

were to minimize weight, ease payload loading, and decrease the number of major load bearing

components. The payload support structure was designed such that the wing lifting force is directly

transferred to the payload box that is strategically placed directly above the rear section of front wing. The

payload itself consists of a welded box made of steel plates. The box is then held together by bolts that

run vertically through the box. The payload box will be inserted into the fuselage through the top hatch

and will rest upon two thin aluminum ribs. A small fraction of the payload bay is cantilevered over the

rear spars of the front wing so as to allow for the required payload box positioning. Furthermore, the

payload will be secured to the aluminum fuselage by two additional bolts, which will in-turn secure the

landing gear as well.

Stability and Control

Static Stability

In the analysis of the stability and control of our aircraft, the priority was given to static stability

over dynamic stability, but both analyses were performed. The following stability models detail the

equations used during both prototype design and the full scale optimization of the tandem wing aircraft.

Longitudinal stability, Cmα, is crucial with regard to pilot control and desired maneuverability.

Well-documented stable aircrafts have a static margin (S.M.) within 5-40% using the equation below,

𝑆. 𝑀. = 𝑥̅𝑎𝑐 − 𝑥̅𝑐𝑔 (4)

The shift in aerodynamic center due to the fuselage contribution was found to be negligible using

Multhopp Method [5]. As the competition calls for high lift and minimal maneuverability, a static

margin of 30% was chosen for the prototype, which was later confirmed to be ideal by pilot feedback,

after flying the prototype. However, due to structural considerations regarding position of the payload

bay with respect to wing spars, the static margin was adjusted to 23%. Longitudinal stability was

calculated using the following derived equation:

16
−𝑥𝑎𝑐 𝑑𝜀
𝐶𝑚𝛼 = −𝐶𝐿𝛼𝑤 ( ) − 𝜂𝐻 𝑉𝐻 𝐶𝐿𝛼𝐻 (1 − 𝑑𝛼) (5)
𝑐̅

where xac is the x-coordinate of the aerodynamic center and VH is the horizontal tail volume ratio.

Downwash from the main wing was found to be negligible during wind tunnel testing of the prototype

due to ample spacing in height between the two wings.

Directional stability is critical due to the short coupled nature of dimensionally constrained

aircraft. In order to maximize lift and efficiency of the two wings, the rear wing was placed on top of

two vertical fins. This increase in vertical surface area led to optimal directional stability. Directional

stability, Cηβv, was calculated for the aircraft by the following derived equation:

𝑑𝜎 𝑆𝑣 𝑋𝑣1 (6)
𝐶𝑛𝛽𝑣 = 𝐶𝐿𝑎𝑣 (1 − 𝑑𝛽) 𝜂𝑣 𝑏𝑆

where ηv is the dynamic pressure ratio, Sv is the vertical tail area, xv1 is the distance from the A.C. of the

𝑑𝜎
vertical tail surface to the C.G., and 𝑑𝛽 is the side-wash effect which was assumed to be negligible

because of the ample spacing between the wingtips and vertical tails.

Lateral static stability derivative, Clβ, was difficult to determine due to both the lack of dihedral

as well as negligible sweep. Nonetheless, the equation below was used to calculate lateral static stability:

𝑧𝑣 𝑆𝑣
𝐶𝑙𝛽 = −𝐶𝐿 𝛼 𝑓(𝜆)𝛤 + 𝑔(𝐴𝑅, 𝜆)𝐶𝐿 𝛼 𝑡𝑎𝑛𝛬 − 𝜂𝑣 𝐶𝐿 𝛼 (7)
𝑣 𝑏𝑆

where λ is the tip-to-root taper ratio, ηv is the efficiency factor, and the last term is the contribution of the

vertical tail to Clβ. This shows that roll stability is proportional to dihedral angle and wing sweep.

However, this equation does not take into account the high wing placement of the rear wing, which is

the dominant contributor to roll stability for our aircraft. Combining this with large vertical tails above

the C.G. led to an excessive amount of roll stability as determined in XFLR5. A low front wing

placement was utilized by the design team to both balance roll stability as well as further reduce

downwash effects. By analyzing a value of Clβ = -.331 obtained from XFLR5, it was calculated that the

tandem had an “effective dihedral” angle of 6°.

17
In order to counteract the moments induced by the motor during cruise, the following equation

was derived from relations between lateral and directional stability, calculating the thrust line through

the C.G.:

𝜂𝛽 𝑃 (8)
𝜃 = sin−1
𝑇𝐿𝜔𝑙𝛽

Where ηβ and lβ are dimensionalized with respect to cruise conditions, ω is the rotation of the motor in

radians per second, and P and L are the power and moment arm of the motor, respectively. Using this

equation, the design team calculated that the motor be mounted at 7° to the right of the original thrust

line. This value was confirmed during flight testing of the full scale aircraft when no rolling moment

was observed during cruise.

Damping Analysis

Calculations for Cmq, pitch damping stability, were extracted from:

2 2 (9)
𝐶𝑚 𝑞 = ( ) (𝐶𝐿 𝛼 𝑙𝑤 𝑆𝑤 − 𝐶𝐿 𝛼 𝑙𝑡2 𝑆𝑡 )
𝑆𝑤 𝑐 2 𝑤 𝑡

The pitch lever arms, the distance from CG of entire aircraft to AC of each wing, for the wing

and lifting tail are the predominant contributors to pitch damping stability. The front wing is

destabilizing as seen in Equation 9, but this was easily overcome due to the large stabilizing lifting tail

and longer moment arm.

The equation for Cnr, directional damping, was derived as below:

1 2 (10)
𝐶𝑛 𝑟 = − ( ) ( ) (𝐶𝐿 𝛼 𝑙𝑣2 𝑆𝑣 )
𝑏𝑤 𝑆𝑤 𝑏𝑤

This equation shows that the directional damping was improved upon due to a smaller frontal wing area

as well as the increased vertical tail area.

The equation for Clp, lateral damping stability, was derived as:

1 1 + 3λ𝑤 1 1 + 3λ𝑡 (11)


𝐶𝑙 𝑝 = −𝐶𝐿 𝛼 ( ) ( ) − 𝐶𝐿 𝛼 ( ) ( )
𝑤 12 1 + λ𝑤 𝑡 12 1 + λ𝑡

18
Both the wing and the lifting tail were stabilizing for lateral damping stability, so the design team chose

to maximize the taper ratios of the wing, λw, and the lifting tail, λt, in order to make the lateral damping

stability stronger in opposition to the roll rate.

Primary Flight Control

The half-scale foam prototype’s known control derivatives were used as a basis for comparison.

The amount of increase or decrease for the values was determined from flight testing of the prototype

and pilot feedback. In addition, the team cross-referenced these coefficients with appropriate values

from a Cessna 172. This aircraft was used as a reference because it is a well-documented trainer aircraft,

and its main focus is stability with relaxed maneuverability.

The high placement of the rear wing caused significant roll stability, resulting in the need for

more roll authority than predicted. The pilot requested an increase in roll authority of roughly

35%. However, in order to maintain structural integrity of the wing, the chord ratio was decreased from

35% to an absolute maximum of 27%. Fortunately, CLα was increased drastically due to the selection of

the E423 airfoil, resulting in the need for only a small increase in aileron span. Clδα increased in

magnitude to 0.461, a 36% improvement over the prototype. The equation used for aileron control

power is defined as:

2𝐶𝐿 α𝑤 τ𝐶𝑟 𝑦 2 𝑦𝑜
2 λ−1
𝐶𝑙 𝛿 = [ 2 + 3( ) 𝑦3] (12)
λ 𝑆𝑏 𝑏 𝑦𝑖

where S, b, λ, 𝐶𝐿 α𝑤 , 𝐶𝑟 were defined from aerodynamic optimizations and the variables  , yo , and yi

were chosen to get a Cl value within the team’s ideal range.

After flight testing of the prototype, elevator control power was deemed sufficient by the pilot,

but the elevator span was increased from 50 to 60% of the rear wing in order to compensate for the

larger frontal wing area as seen in the volumetric ratio equation below.

𝑏
𝐶𝑚 𝛿𝐸 = −𝐶𝐿 𝛼ℎ ηℎ 𝑉ℎ 𝑏𝐸 τ𝑒 (13)

19
We assumed ηh to be unity, because of the low Reynolds number operating range and the high

placement of the rear wing.

The rudder control power was proven ideal during flight testing. This is due in large part by the

twin vertical fins as seen in the following derived equation:

𝑏
𝐶𝑛 𝛿𝑅 = −𝐶𝐿 𝛼𝑉 𝑉𝑣 η𝑉 τ𝑟 𝑏𝑅 (14)
𝑉

Due to necessary structural supports, chord ratio could not be increased any further than the set

value of 35%, which led to the decision to accept a penalty of 20% less control than the prototype in

order to maintain a larger wingspan.

Dynamic Analysis

Dynamic analysis is crucial when predicting an aircraft’s response to perturbations in flight since

static stability does not inherently mean dynamic stability. XFLR5 was used to extract eigenvalues

evaluated at the moments of inertia, provided through Solidworks, for our maximum payload. The

aircraft exhibits exceptional short period, phugoid, spiral, and dutch roll qualities as referenced to MIL-

STD-1797A, but slightly unstable in roll mode; all of which can be seen in Table 3.

Longitudinal Modes Lateral Modes


Phugoid Short Period Spiral Roll Dutch Roll
Natural Damping Natural Damping Time Time Natural Damping
Frequency Ratio Frequency Ratio Constant Constant Frequency Ratio
0.731 0.0256 8.61 0.708 62.4 3.85 4.67 0.28
Table 3: Longitudinal and Lateral Mode Qualities

Final Configuration

The outcome for our 2016 competition aircraft is based upon the ultimate goal of improvement

from last year’s v-tail design and optimization of the prototype. It was first necessary to develop and fly

a prototype based on the confirmed stability and control derivatives for our 2015 aircraft. This was done

in order to create a baseline for comparison due to a lack of reference material on tandem wing aircrafts.

The stability of the tandem wing was then cross referenced with the Cessna 172 for verification. After

full scale flight testing, the risk of designing a tandem wing aircraft was justified when it proved to be

sufficiently stable and had a much greater lifting capacity for payload than our previous year’s aircraft,

all of which can be seen in Table 4.

20
Stability Derivatives V-Tail 2015 Prototype Tandem Wing Cessna 172
Cma -0.945 -3.151 -1.339 -0.613
Clb -0.108 -0.363 -0.331 -0.092
Cnb 0.159 0.18 0.154 0.059
Cmq -4.559 -22.576 -11.37 -12.4
Clp -0.64 -1.26 -1.238 -0.484
Cnr -0.109 -0.097 -0.073 -0.094
Control Derivatives
Cmδe -0.724 -1.535 -1.558 -1.122
Clδa 0.355 0.339 0.461 0.229
Cnδr -0.069 -0.148 -0.122 -0.065
Misc.
Max Theory Lift (lbs.) 35 N/A 52 N/A
L/D 8.5 13.5 12.4 9
Span (in.) 95.5 N/A 98 N/A
Table 4: Stability Analysis Results

Servo Selection

The design team calculated the maximum required torque, shown in Table 5, using control

surface derivatives evaluated at cruise conditions with a maximum deflection of 35°. The moment arm

of the servo is equal to the moment arm of the control horn, thus the maximum required torque of each

surface is directly compared to the servo torque. Both rudder surfaces are driven by the same servo.

Control Surface Maximum Required Servo Torque [oz-in] Servo Model Factor of Safety
Torque [oz-in]
Aileron 50.8 76.37 Hitec HS-625MG 1.5
Elevator 35.2 76.37 Hitec HS-625MG 2.17
Rudder (2) 45.0 76.37 Hitec HS-625MG 1.7
Flap 21.8 76.7 Hitec HS-625MG 3.6
Table 5: Control Surface Servo Torque Data

Propulsion System Design

The primary design constraint for the propulsion system was the 1000W power limit. The team

utilized eCalc Propeller Calculator to predict the static performance of different motors and propellers.

The team created a spreadsheet to compare 14 different motors in thrust performance for the same power

draw of 1000W. In the previous years, the limiter cut the throttle during takeoff. Therefore, this year’s

initial design was to pick a motor that used less than 1000W static and in addition to testing the

properties of the power limiter using a Pixhawk. After analyzing the spreadsheets, the data trends were

used to determine the optimal propeller diameter and motor kv: 21in and 200-270 kv, respectively, in

which we then compared 3 motors with these calculated specifications.

21
After the Pixhawk testing proved the eCalc trends were accurate, eCalc was then used to select

the final motor. The static power required and static thrust of each combination were compared, with

most desirable combinations being further analyzed using thrust versus airspeed data provided by eCalc.

The Scorpion SII-5525-210kv motor and APC 21x13E propeller were found to provide the most thrust

per unit power at takeoff speeds. In a static condition, this motor draws 930W for a propeller of 13in of

pitch at the predetermined optimal 21in. diameter. To account for pitch loss at forward velocity, we

initially chose a 20x11 propeller. After flight and wind tunnel testing, we selected a 21x13 APC

propeller. APC propellers boast high efficiency, durability, and availability.

Takeoff Performance

In order to calculate the theoretical maximum payload the plane can lift, it is necessary to

combine aerodynamic simulation, wind tunnel, and experimental data from thrust tests as a function of

velocity. Using motor data of thrust versus velocity, it was possible to model the thrust of the motor by

the simple polynomial equation below:

𝑇 = 0.0026𝑉 2 + 0.0139𝑉 + 0.009 (15)

Equation 15 outputs the thrust, T, of our motor in pounds-force with respect to velocity. From here, it is

possible to take three different approaches to determine takeoff velocity and the takeoff distance. The

three methods used to validate the calculations were an iterative method of equating force to the product

of mass and acceleration, Anderson’s takeoff equation, and Advanced Aircraft Analysis.

For the first of the three methods, the plane is simulated using a MATLAB code that calculates

position, speed, and acceleration at half-second intervals. To begin, initial environmental conditions

such as density and gravity are entered into the code. Plane data, e.g., lift, drag, and motor thrust values,

are then inputted. Starting at a velocity of 0 m/s, the program calculates the current thrust based on the

current velocity and dynamic pressure. With this thrust value and dynamic pressure it is possible to use

the data values to interpolate the corresponding lift and drag values. Acceleration can then be calculated

and integrated for velocity and distance. During takeoff, the pilot will accelerate the plane until the back

wing naturally lifts up and then move the tail down to increase angle of attack and lift. To simulate this

22
takeoff process, the program uses the lift and drag values at 5° angle of attack until the lift of the back

wing equals the weight of the back wing. At this point, the code begins to use the lift and drag values at

0° angle of attack. After an arbitrary period of time, the program switches to using 5° angle of attack data

to simulate the pilot elevator input to lift off.

The second method is using Anderson’s takeoff equation, listed below, from his Introduction to

Flight textbook. Using this equation, we were able to calculate an approximate takeoff distance.

𝑑 = 1.44𝑊 ∗ 2𝑔𝜌𝑆𝐶𝐿𝑚𝑎𝑥 (𝑇 − 𝐷) (16)

The last method used Advanced Aircraft Analysis (AAA). AAA is an analytical program used to

develop the flight and performance characteristics of an aircraft after inputting needed aircraft

specifications. From this, we were able to determine an approximate takeoff distance.

Figure 14, illustrates all of the calculated takeoff distances from each method.

Figure 14: Takeoff distances for various calculations methods.

4. EXPERIMENTAL DATA COLLECTION

Motor and Propeller Testing

Through a literature search and the software used to predict the power plant performance, we

found that propeller thrust varies linearly with airspeed. In order to accurately model a motor-propeller

system, we elected to experimentally test the propulsion system at multiple velocities for use in the

takeoff prediction model which determines the maximum possible payload. We constructed the test

stand, shown in Figure 15, for use in Texas A&M University’s 3ft. x 4ft. subsonic wind tunnel. The

overall dimensions roughly represent the forward section of the fuselage and houses all of the

23
electronics necessary to run the motor. We collected the thrust and power usage at airspeeds ranging

from 0 to 50 miles per hour. The resulting thrust versus airspeed trend is modeled by:

𝑇 = 11.97 − 0.056𝑉 (17)

According to the trend line, the maximum thrust is 11.97lbs. when the aircraft is static. Additionally, as

desired, the motor never reached the 1000W limit. The experimental data shows suboptimal

performance when compared to the ideal static thrust of 12.5lbs. provided by the motor analysis

software; however, the resulting 11.97lbs. of thrust proves capable of lifting our desired maximum

payload.

Figure 15: a) Test stand design. b) Test stand in the 4’x3’ wind tunnel operating at 30 miles per hour.

Material Properties

To determine the material properties of the two most populous woods used in our wings, we

conducted tensile testing. Our primary wood building materials, basswood and balsa, were tested in two

different thicknesses; 0.063 and 0.125in. Dog bone-shaped specimens were laser cut to ensure consistent

sizing for the cross sectional areas based on ASTM D143 standards [5]. For each material and thickness,

five tests were performed to ensure a consistent average of results. From the stress-strain curve, Young’s

Modulus was subsequently calculated [8]. The results are shown in Table 6 which exhibit, as expected,

that the thinner specimens have a higher density and lower porosity, making a higher Young’s Modulus

and vise-versa.

Balsa
Thickness (in) Average Tensile (MPa) Average Young's Modulus (MPa)
0.063 N/A 272.58
0.125 N/A 122.29
Basswood
Thickness (in) Average Tensile (MPa) Average Young's Modulus (GPa)
0.063 81.03 2.52
0.125 67.24 2.14
Table 6: Material Testing Results of Basswood and Balsa Wood

24
5. CONSTRUCTION

Construction of the plane consisted of two phases; construction of the fuselage and construction

of the two wings. To remove human error, a CNC laser cutter was utilized to cut all flat wood pieces and

a CNC water jet table was used for the metal components. We also custom designed jigs for our wings to

provide exactness and repeatability during construction.

To construct the fuselage, following strict drawings, the metal components were bent using large

sheet metal benders. The resulting pieces were assembled using rivets, with piano hinges used on the

rudders and hatches. The flexible nature of the thin aluminum allowed for constant corrections to be made

to the assembly to ensure the final product was to spec.

To construct the wing, the wing ribs were slid onto straight rods running through a plywood jig.

After gluing the payload plate between the inside ribs, we epoxied the leading and trailing spars made of

overlapping basswood 1/8” inch spars directly onto the ribs. A block of wood was added to the leading

edge of the ribs and the jig was cut off. We used laser cut handheld guides to sand the leading edge to the

correct shape. Spar web and skin was then added to the wing. Additionally, balsa filler was applied to any

imperfections to create a smooth surface. To build the control surfaces, we used the laser cutter to create

the leading edge shape and etch lines onto it to serve as guides for the placement of the control surface

ribs. The team then epoxied hinges to the trailing edge spar with the opposing side of the hinge epoxied

to the control surfaces.

Throughout construction, we used medium thickness cyanoacrylate, CA, glue for general

attachments, e.g., spars to ribs, epoxy for plywood attachments, and thin CA for cracks and other small

openings. All metal pieces are held together using standard metal rivets. Standard milled steel plates were

hand cut and welded together to form the payload box. To create the landing gear, we used a metal bending

brake to bend the landing gear to precise angles.

6. COMPETITION FLIGHT PLAN

In order to maximize the team’s flight score, we prepared a detailed flight plan and payload

prediction based upon the scoring rubric. This flight plan is a carefully considered balance between

25
maximum round scores and a feasible and, most importantly, repeatable payload prediction. We plan to

fly every round at a consistently heavy payload while retaining confidence in the repeatability in our

flights.

During optimization of the payload prediction score, our team came to an important realization

that the maximum payload prediction score does not occur precisely at the payload prediction. Instead,

the maximum final flight score (FFS) occurs at exactly one half pound above the predicted payload. This

comes from the result of the combination of the parabolic prediction bonus with the superimposed linear

raw payload score which can be seen in Figure 16.

To determine the theoretical maximum payload that our aircraft can lift, we used a combination of

aerodynamic simulations of the entire body, experimental data of the force produced by the motor as a

function of airspeed, and a dynamic simulation of the aircraft rolling down the runway. For a given mass

of the aircraft, we numerically integrated the Fx = max equation to output the velocity and position of the

aircraft on the runway. At every time step iteration, the velocity determined the amount of lift possible at

5° angle of attack. When lift exceeded the weight of the aircraft, takeoff was achieved. This process was

repeated for a range of payloads to find the maximum payload capable of takeoff utilizing a maximum of

95% of the allowable runway length. Based off these analyses, we estimate the maximum payload to be

29lbs. The payload prediction equation thus reflects a value one half pound less than our maximum

payload value whose graph can be viewed at the end of this report.

40
Score Curve
Final Flight Score

30 Max Prediction
Max Score
20

10

0
0 2 4 6 8 10 12 14 16 18 20
Payload [lbs]

Figure 16: Plot of the final flight score for a single round as a function of payload lifted.
For an example, the graph assumes a max predicted payload of 16.5 lbs. and the max score occurs at 17 lbs.

7. CONCLUSION

Throughout the design process, we performed optimization and trade-off studies of the four

major components of aircraft design, aerodynamics, structures, stability and control, and propulsion.

26
From these we were able to accomplish the objectives and satisfy the requirements of the competition.

Our main design focal points were adequate controllability, maximum lift-to-drag ratio, maximum

possible lift, and minimum aircraft empty weight, each of these were targeted while continuously

considering the critical design constraints.

For the topic of structures, trade studies were performed between manufacturability, total aircraft

empty weight, and structural integrity. Near the payload structure, more weight was needed in order to

provide a more reliable structure. Where in the vertical tails, large sections of the solid aluminum

sheeting could be cut out and the integrity of the structure still held. In the topic of aerodynamics, the

airfoils selected and chord lengths were optimized considering the weighted importance of lift, drag,

maximum weight of the aircraft, and overall efficiency. Computational fluid dynamics (CFD)

simulations also determined the lift, drag, and efficiency of the aircraft as a whole. Within stability and

control, the team additionally used CFD analyses to determine the coefficients of lift for both the front

and rear wings in order to calculate the aircraft’s static margin. Subsequently, trade-off studies were

conducted between controllability and lifting capabilities. For propulsion, the design team analyzed

motors using eCalc then verified these findings with motor testing in the wind tunnel, which led us to

determine our maximum payload prediction.

Based on literature survey, detailed modeling analyses, trade study, optimization, and

experimental testing, we can undoubtedly declare that we developed an aircraft capable of lifting as

much payload as possible within the design constraints. Integrating a design process similar to that

found in industry, our engineering team worked in specialized topic groups surrounding aerodynamics,

structure, stability and control, and propulsion. Through multiple design iterations and optimization, we

were able to produce a thorough, imperforated aircraft with tremendous potential to perform well at the

competition.

27
References

[1] "Wing Planform," Dauntless Aviation, [Online]. Available: http://www.dauntless-soft.com. [Accessed


2014].

[2] R. Nelson, Flight Stability and Automatic Control, Second ed, McGraw-Hill 1997, p.456.
[3] A. Lennon, R/C Model Aircraft Design, M.A. Markowski, Ed., Air Age Media, Inc, 1996.
[4] A. Bedford and K. Liechti, Mechanics of Materials, Upper Saddle River, NJ: Prentice Hall, 2000.
[5] J. Roskam, Airplane Design Part II: Preliminary Configuration Design and Integration of the
Propulsion System, Lawrence, Kansas: Design, Analysis, and Research Corporation, 2004.
[6] “Standard Test Methods for Small Clear Specimens of Timber,” ASTM International Standards,
2009.
[7] W.D. Callister, Materials Science and Engineering: an Introduction, 7th ed., 2007.
[8] I.H. Abbot and A.E. Von Doneho, Theory of Wing Section, Mineola, NY: Dover, 1949.
[9] D.H. Allen and W.E. Haisler, Introduction to Aerospace Structural Analsys, Wiley, 1985.
[10] T.C. Corke, Design of Aircraft, Upper Saddle River, NJ: Pearson Education, Inc., 2003.
[11] D.P. Raymer, Aircraft Design: A Conceptual Approach, 3rd ed., AIAA, 1999.
[12] B.H. Tongue and S.D. Sheppard, Dynamics: Analysis and Design of Systems in Motion, Wiley,
2005.
[13] K. Wood, Aerospace Vehicle Design, vol. 1: Aircraft Design, Boulder, Colorado: Johnson
Publishing Company. 1968.
[14] E. Seckel, Stability and Control of Airplanes and Helicopters, Academic, 1964, p.506.
[15] R. Woodward, “Aluminum and Aluminum Alloys and Designations,” 4 September 2011.
[Online].
[16] “UIUC Airfoil Data Site,” 2010. [Online]. [Accessed 2012-2013].
[17] M.H. Sadraey, Aircraft Design: A Systems Engineering Approach John Wiley & Sons, 2013.
[18] J.D. Anderson, Fundamentals of Aerodynamics, 5th ed., McGraw-Hill, 2010.
[19] Katzoff, Mutterperl. The End-Plate Effect of a Horizontal-Tail Surface on Vertical Tail Surface.
UNT Digital Library.
[20] Murray, Harry. Wind Tunnel Investigation of End-Plate Effects of Horizontal Tails on a Vertical
Tail Compared with Available Theory. National Advisory Committee for Aeronautics. [Accessed
January 11, 2006].

28
Texas A&M University's Aero Design Team #022
Payload Prediction Curve
Density Altitude
35

Payload = -0.00049x+29
30

25

20
Payload (lbs.)

15

10

0
0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000
Density Altitude (ft.)

You might also like