You are on page 1of 100

Wingbox

Design
Designing, manufacturing, and testing of a
wingbox for Project GreenWing
Group E10E
Delft University of Technology
This page is intentionally left blank.
Wingbox

Design
Designing, manufacturing, and testing of a
wingbox for Project GreenWing
by

Group E10E

Aleksei Esser 5297656


Theophile Dubois 5223210
Mike Timmerman 5216796
Robin Bos 5260566
Didier van der Voort 5338158
Bart Meenderman 5328802
Ewan MacDonald 5219671
Lucas van der Klugt 5271177
Morris Jansen 5265207
Dragomir Nikolov 5216915
Project duration: April 19, 2021 – July 9, 2021
Mentor: Roderick Wassenaar TU Delft, Master Student
Contents
Summary iv
1 Introduction 1
2 Literature Study 2
2.1 The Wingbox . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2.2 Minimizing Environmental Impact of a Wing Box . . . . . . . . . . . . . . . . . . . . . . . . . 2
2.3 Operating Conditions of an Urban Air Mobility Aircraft and its Related Challenges . . . . . . . . 3
2.4 Safety Factor(s) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.5 Failure Modes and Locations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.6 Relevant Loads on an UAM Wing during various Phases of Flight . . . . . . . . . . . . . . . . . 4
2.7 The Influence of Wingbox Deformation on the Performance of a Wing . . . . . . . . . . . . . . 6
2.8 Manufacturability and Feasibility Assessment . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.9 Robot Motion Plan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3 The Wingbox Design 7
3.1 Wingbox Design Description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3.1.1 Function and Design Objectives of the Wingbox . . . . . . . . . . . . . . . . . . . . . . 7
3.1.2 Requirements and Constraints of the Wingbox. . . . . . . . . . . . . . . . . . . . . . . 7
3.1.3 Assumptions made for the own Wingbox Design. . . . . . . . . . . . . . . . . . . . . . 8
3.1.4 Free Design Parameters of the Wingbox . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.2 Description of the Wingbox Design Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
3.3 Description of Calculations for the Wingbox Design . . . . . . . . . . . . . . . . . . . . . . . 11
3.3.1 Cross-Section Properties of the Wingbox. . . . . . . . . . . . . . . . . . . . . . . . . . 11
3.3.2 Internal Force and Moment Diagrams of the Wingbox . . . . . . . . . . . . . . . . . . . 12
3.3.3 Bending Normal Stress Calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.3.4 Shear Stress Calculations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.3.5 Rivet Failure Calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.3.6 Final Stringer Configuration of the Wingbox . . . . . . . . . . . . . . . . . . . . . . . . 20
3.4 Design Justification for the Wingbox - Sustainability . . . . . . . . . . . . . . . . . . . . . . . 21
4 Manufacturing of the Wingbox 22
4.1 Discussion of Manufacturability of the Wingbox Design . . . . . . . . . . . . . . . . . . . . . 22
4.1.1 Product Complexity of the Wingbox . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
4.1.2 Product Variants & Production Friendly Design of the Wingbox . . . . . . . . . . . . . . 22
4.1.3 Wingbox Component Availability and Price . . . . . . . . . . . . . . . . . . . . . . . . 23
4.2 KUKA Robot Motion Plan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
4.2.1 Usage of KUKA Robot for Automated Manufacturing of the Wingbox . . . . . . . . . . . 23
4.2.2 Description of Robot Motion Plan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
4.3 Feasibility Assessment between Manual & Automated Manufacturing . . . . . . . . . . . . . . 26
4.3.1 Manual Manufacturability of the Wingbox . . . . . . . . . . . . . . . . . . . . . . . . . 26
4.3.2 Partly Automated Manufacturability of the Wingbox . . . . . . . . . . . . . . . . . . . . 27
4.3.3 Conclusion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
5 Testing of the Alfa Wing 29
5.1 Description of Test Objectives for the Alfa Wing. . . . . . . . . . . . . . . . . . . . . . . . . . 29
5.2 Description of Test Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
5.3 Calculation of Load Distribution of the Whiffletree . . . . . . . . . . . . . . . . . . . . . . . . 31
5.4 Description of Test Events for a Wingbox . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
6 Analysis of the Wingbox Testing 35
6.1 Discussion on Differences Between Own Design and Alfa Wing Design . . . . . . . . . . . . . . 35
6.1.1 Inboard Part of the Wingbox . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
6.1.2 Outboard Part of the Wingbox . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

i
Contents ii

6.2 Alfa Wing Failure based on own Calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . 36


6.2.1 Geometric Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
6.2.2 Sheet Buckling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
6.2.3 Shear Buckling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
6.2.4 Inter Rivet Buckling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
6.3 Explanation of Strain Gauges selected for Analysis . . . . . . . . . . . . . . . . . . . . . . . . 37
6.4 Analysis of the Strain Gauge Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
6.5 Description and Explanation of Failure Location . . . . . . . . . . . . . . . . . . . . . . . . . 38
6.6 Comparison of Test Data and Calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
6.7 Evaluation of Own Wing Design based on the Alfa Wing . . . . . . . . . . . . . . . . . . . . . 39
7 Conclusions and Recommendations 40
Bibliography 42
Appendices 42
A Task Distribution 43
A.1 Estimated Time Distribution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
A.2 Actual time distribution. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
A.3 Time spend on each part per person . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
B Material Properties, Availability and Dimensions 50
B.1 Material Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
B.2 Material Dimensions and Availability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
B.3 Rivet Properties and Dimensions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
C Manual Production Timeline 52
D Alfa Wing and E10E Wing Assembly Drawing 55
E Wingbox Production Plan 60
List of Symbols

ρ Density kg/m3

σ Normal stress Pa

σcr Critical buckling stress Pa

σi r Inter-rivet buckling stress Pa

A Cross-sectional area m2

b Stringer pitch m

E Young’s modulus Pa

F cr Critical buckling load N

F ul t Ultimate failure load N

I Second moment of area m4

Kc Compression buckling coefficient -

M Bending moment Nm

n r i vet Number of rivets -

n st r Number of stringers -

Q First moment of area m3

q Shear flow N/m

s Rivet spacing m

t Sheet thickness m

V Shear force N

ybar Distance from the neutral axis m


Summary
The wing is often considered the most important part of an aircraft. Without the wing it would not be able to
fly. A wing collapsing during flight would be disastrous, therefore it is very important that the wing is struc-
turally sound. To make sure of this, the loads in the wing are for a large part carried by a wingbox. The aim of
this report is to explain the function and document the design process of the wingbox for project GreenWing.
This is done while keeping in mind the sustainability aspect of the production and the environmental impact
of producing such a wingbox .

A literature study was conducted to make sure that sufficient background knowledge was known. Some im-
portant findings are:

• A wingbox is a closed structure inside a wing that efficiently resists different kinds of loads, especially
bending and torsion;

• A wingbox can fail in 5 different ways: buckling, inter-rivet buckling, shear buckling, bending and tor-
sion;

• To minimize the environmental impact of the wingbox it is important to improve flight efficiency and
to look into the environmental aspects of the material choice and the production methods.

To design the wingbox the following DELTA requirements had to be adhered to:

• Wingbox must buckle above 3 [kN ] ;

• Failure should occur at 3.3[kN ] ;

• The structure must be optimized for sustainable manufacturing, operations, and end-of-life processes.

Using the requirements for the iterative calculation process, the wingbox was designed with general dimen-
sions: 2.87 [m] x 0.4 [m] x 0.152 [m]. It includes 19 stringers, 4 panels, 4 ribs and a sleeve all made out of
recycled aluminium. (Figure: 1 shows a detailed image of the design).

To prevent any ambiguities in the manufacturing method, a production plan was created. In addition to
this manual production plan, a robot motion plan was produced which could potentially replace the drilling
steps in the manual production plan. The feasibility of each plan was compared and analyzed to determine
whether (partly) automating the production of the wingbox would be cost / time effective or not. The partly
automated production turned out to be the more effective method out of the two.

The tested wing is called the Alfa Wing, a whiffletree was used to uniformly apply bending stresses across the
wingbox (similarly to how lift would affect the wing). Numerous strain gauges and displacement sensors were
placed on the Alfa Wing during the test. The maximum force that the wingbox could carry was 3295.7 [N ].
The maximum displacement was around 150 [mm].

Even though the designed wingbox and the Alfa Wing are different, the test can be used to analyse the design
of this report. The calculations used for this design mostly focus on the first failure mode of buckling. They
do not take the ultimate failure into account. The Alfa wing buckles very early in the test but does fail at the
desired load. All this indicates that the wingbox in this report will buckle at the desired load but regarding its
ultimate failure load, not a lot can be said.

To summarise, the stringers on the top panel of the wingbox will prevent it from buckling and help it reach
the buckling design load, but they will probably also increase its ultimate load, exceeding the failure design
load. However, this cannot be said with certainty, as the calculations made did not focus on that aspect.

iv
List of Symbols v

(a) An exploded view of the inboard side of the wingbox.(own work) (b) An exploded view of the outboard side of the wingbox.(own work)

Figure 1: The design for the wingbox.


1
Introduction
The wing of an aircraft is often considered to be the most important structure of the aircraft. The most effi-
cient aircraft is often considered to be a ’simple’ flying wing, similar to TU Delft’s and KLM’s ’Flying-V’ design,
which is a blended wing body design and might be the future of aviation. In any case, the wing is a very im-
portant element of an aircraft and therefore it is extremely important that it does not fail. A wing experiences
many different loads due to its often complex shape. A very efficient way to withstand these loads is to use a
so called ’wingbox’, which is a closed box structure that mainly forms the internal wing structure. For DELTA’s
latest Project GreenWing, a 4-passenger Urban Air Mobility (UAM) vehicle flying at 150 knots, such a wingbox
needs to be developed to withstand a prescribed distributed load. As the name of Project GreenWing already
hints to, the project keeps the greater context of the design in mind with the tagline "Sustainable travel for all".

The purpose of this report is to describe and analyse the design of a wingbox whilst actively considering its
sustainability[1]. DELTA has provided three requirements that the wingbox needs to meet which includes a
minimum buckling and failure load while optimising for sustainability. The wingbox will be designed with
the following parts:

• The top, bottom, and side sheets;

• The stiffeners;

• The ribs;

• The rivets.

The exact design of the wingbox should include a specific number of parts with specified positioning and ma-
terial while keeping in mind the requirements and the sustainability of the design. The calculation process to
determine these variables must be done carefully to prevent buckling and failure before the described loads.
The sustainability of the design should be accounted for by considering the emissions and costs of material
production and aviation gasoline use. After the design of the wingbox is determined, the design should be
manufactured. For a design to be manufactured correctly, a detailed production plan should be made. Due
to the technology that is present today, it is possible to automate productions to a certain extent. For the
production of this wingbox, the feasibility of partly automating the production should be determined. Once
the wingbox is tested using a whiffletree setup, test data should be analyzed thoroughly, which should give a
valuable indication of the real-life performance of the designed wingbox structure.

This report will start with a literature study on the relevant topics in Chapter 2. With the acquired knowledge,
it will be followed by a description, calculation and justification of the wingbox design choices in Chapter 3.
After that, a description of the production and robot motion plan including a feasibility assessment between
the two production methods will be given in Chapter 4. Chapter 5 will describe the used test setup and the
events that occurred during the test on the Alfa Wing. This is then followed by an in depth analysis of the test
data and a comparison between the Alfa Wing and the own design in Chapter 6. Finally, a final conclusion
will be given in Chapter 7, including recommendations for future research.

1
2
Literature Study
To make sure that the wingbox is designed properly a lot of research needs to be done about the functionality
of the wingbox. To do this, a literature study is conducted. It deepens the understanding of the wingbox and
ensures that a proper design is made and that the test is analysed in a correct manner. Each section in this
chapter covers a part about the theory of the wingbox.

2.1. The Wingbox


Definition
A wingbox, also known as a torsion box, is a closed structure inside the wing that efficiently resists different
kinds of loads, especially bending and torsion. A wingbox can have any cross-sectional shape, as long as it
has a closed cross-section. A normal wingbox spans from the wing root to the wing tip.

Advantages of a box structure


While I-beams and other stiffened structures are great at resisting bending due to its large moment of inertia,
they are less suited for resisting torque. Since the aerodynamic forces do not create only bending moment
but also torque in the wing, it is important that the structure of the wing is well suited against torque. This
is especially important when there is also an engine placed on the wing, as the thrust of the engine will also
create a torque in the wing. Implementing a wingbox can reduce the thickness of the wing at given wingspan
due to its higher efficiency in resisting torsion compared to a two-spar wing concept. One of the biggest
advantages compared to the two-spar wing concept is that it generally allows for a more lightweight structure
[2].

2.2. Minimizing Environmental Impact of a Wing Box


The environmental impact of an aircraft design is mostly influenced by three main factors:

1. The type of fuel used;

2. The flight efficiency of the aircraft;

3. The manufacturing methods and materials used.

While designing the wingbox there is not much that can be done to influence the type of fuel used in the air-
craft. Because of that the focus should be on improving the flight efficiency and reducing the environmental
footprint of the manufacturing process.

The best way to improved flight efficiency is to reduce the weight of the aircraft. Even a small initial reduction
of the total weight of the aircraft body can lead to a so-called "snowball effect", which refers to the ultimately
large resultant weight reduction achieved after taking into account the way the design will change because of
the reduced load on other components (e.g. 1 [kg ] weight reduction in the wingbox means 1 [kg ] less that
has to be carried by the landing gear and ≈ 98 [N ] less lift force required during cruise, hence smaller and
lighter wings, meaning smaller fuel or battery pack size requirements for the same mission distance etc.).

Another area of focus when considering the environmental aspects of the design should be the ecological ef-
fects of the material choice. This includes minimizing the green house gasses used to produce the materials
needed as well as looking into the recyclability of those materials and various other aspects of their produc-
tion. A table of the environmental impact of the manufacturing of different aerospace materials is presented
below in Table 2.1.

2
2.3. Operating Conditions of an Urban Air Mobility Aircraft and its Related Challenges 3

Table 2.1: Embodied energy, carbon footprint, and water demand of some common aerospace materials.[3]

Energy [M J /kg ] Carbon [kg /kg ] Water [l i t r es/kg ]


Low and medium carbon steels 25-28 1.7-1.9 44-48
Aluminum alloys 200-220 12-13 1100-1300
Titanium alloys 650-720 44-49 190-210
CFRP 450-500 33-36 1300-1500
GFRP 150-170 9.5-11 150-170

2.3. Operating Conditions of an Urban Air Mobility Aircraft and its Related
Challenges
By definition the operating environment of an urban air mobility (UAM) aircraft are cities with all the conse-
quences that entails. Such aircraft will have to follow strict safety protocols to ensure that they will not present
a safety risk to pedestrians below and even to people indoors in nearby buildings. Unlike conventional air-
craft UAM ones are not allowed to experience even small instabilities during flight, as one such instability can
lead to a major crash and a possible death toll [4].

Another great concern is the noise created by such vehicles during flight. According to many UAM researchers
noise will probably be the greatest obstacle to wide adoption of UAM aircraft for commercial use.

UAM aircraft operations can also be influenced by many other factors such as weather, airspace busyness
and local community backlash.[5]. This has forced air regulation authorities to propose the term "practical
capacity" of an airspace. "Practical capacity is defined as the number of operations that can simultaneously
be accommodated within an airspace sector while accruing no more than a specified amount of average
delay." [5]. Figure 2.1 lists most factors that influence the practical capacity of an airspace sector .

Figure 2.1: Air control sector capacity influence diagram [5]

2.4. Safety Factor(s)


In engineering, safety factors are used to increase the strength of a structure to make a structure capable of
resisting loads higher than the actual loads. Safety factors are used to make the structure increase the relia-
bility and reduce the risk of failure. Usually, higher safety factors are used for structures for which failure can
be catastrophic [6].

The Federal Airworthiness Regulation specifies a 1.5 Ultimate Factor of Safety for external loads on the aircraft
structure during ground and flight time [7]. This also includes the loads on the wing structure.
2.5. Failure Modes and Locations 4

2.5. Failure Modes and Locations


In general, structural failure is defined as the instant when a structure is no longer capable of resisting the
stresses that are caused by the loads acting on the structure [8]. However, the definition of structural failure
can also differ between types of structures. For instance, for a safe-life structure, failure can also be defined
as "the development of a detectable crack" [9].

There are five different failure modes in the design of an aircraft wing box. These are buckling, inter-rivet
buckling, shear buckling, bending and torsion. The different failure modes result in different critical failure
locations:

• For buckling, the failure location is at the top skin of the wing box due to compression;

• For inter-rivet buckling, the failure location is between the rivets, also due to compression;

• For shear buckling, the failure location is at the spar webs due to shear forces;

• For bending, the failure location will be at the location furthest away from the neutral axis as the normal
stresses will be highest there. Since the magnitude of the bending moment differs along the length of
the wing box, the failure location will be at the location of the highest bending moment magnitude;

• For torsion, the failure location will be at the location on the cross-section with the smallest thickness.

2.6. Relevant Loads on an UAM Wing during various Phases of Flight


Assumptions
For this Section, some assumptions have been made on how the UAM vehicle will perform flight. Since it is
an urban vehicle, there will not be a lot of space to take off. This leads to the assumption that the vehicle will
take of vertically like a helicopter. When it has built up sufficient altitude it will accelerate horizontally and
start to fly using its wings.

Taxi
During taxi none of the motors on the wing will be engaged yet. Because of this, the only the remaining loads
acting on the wing will be the weight of the wing itself and the weight of engines mounted to it. The com-
bination of those loads creates bending moments on the wings and in case that the centers of masses of the
engines do not align with the wing centerline, a torsion load will also be present.

Takeoff
Takeoff will consist of two phases. The first being the vertical takeoff and the second being the horizontal ac-
celeration. During takeoff the engines will increasingly produce more thrust, lifting the vehicle of the ground.
Now the thrust and the weight of the wing act upon the wingbox together with their reaction force at the end,
equal to the weight of the vehicle.

When the UAM vehicle starts to accelerate horizontally lift and drag begin to act upon the wing while the
thrust shifts in a different direction. The weight of the wing remains unchanged. The reaction now gets a
component in the negative z direction due to the thrust being greater than the drag on the wing.

Cruise
During cruise the main loads on the aircraft will be the aerodynamic forces (lift and drag),the thrust created
by the propulsion system and the corresponding reaction forces that will act upon the wing. The asymmetry
of those loads induces bending and torsion moments in the wing and body of the aircraft.

Landing
It is assumed that the vehicle will land in the same manner as it takes off. It will decelerate horizontally before
landing vertically. The loads will be the same as during takeoff, except for the thrust that acts in the other
direction during the deceleration and with that the reaction force.

All force distributions on the wingbox per flight phase can be seen in Figure 2.2.
2.6. Relevant Loads on an UAM Wing during various Phases of Flight 5

Figure 2.2: FBD’s of the wingbox during several phases of flight (own work)
2.7. The Influence of Wingbox Deformation on the Performance of a Wing 6

2.7. The Influence of Wingbox Deformation on the Performance of a Wing


A wings performance is based on its shape. For example changing the sweep angle of a wing will influence its
C
critical mach number and changing the shape of the airfoil will influence the C l curve.
d

The wingbox is the skeleton of the wing. If the wingbox deforms, the wing deforms with it.[10] Another way
how the wing box can influence the wings performance is when there is a torque in the wingbox. The torque
will twist the end of the wingbox with respect to the start.[11] This will change the angle of attack at the end
of the wing, influencing the lift and drag coefficients. So it is up to the wingbox to make sure that the wings
shape does not change and the performance of the wing does not deviate from what was calculated in the
design stages, so there aren’t any nasty surprises during flight. For example running out of fuel before you
reach your destination because the drag was higher due to the deformations than was initially calculated. To
accomplish this the wingbox needs to be quite stiff.

2.8. Manufacturability and Feasibility Assessment


Manufacturability means the theoretical possibility to produce something with machines or by humans. This
means that if a certain shape is desired, then it is theoretically possible to achieve this shape manually or by
using machines. Feasibility, however, means the theoretical possibility to produce something with the given
resources and requirements. In other words, it assesses the achievability of the desired shape considering the
given resources and requirements like costs, available materials, and more.

For the wing box that needs to be designed, many shapes and dimensions are already specified. From the
available shapes and materials, design choices need to be made based the given requirements. Therefore,
the feasibility assessment will have a greater impact on the design choices than the manufacturability assess-
ment.

2.9. Robot Motion Plan


A robot motion plan is an algorithm that tells a computer how to move (part of) the robot from one place to
another, while also taking obstacles into consideration, like a GPS tells you how to move from point A to point
B [12].

For a robot motion plan design, there are certain factors that are important to take into consideration. These
factors are all about the obstacles that are encountered. Robot motion planning generally becomes harder
when the characteristics of the obstacles (like size and location) are unknown and when the obstacles are
moving. Therefore, the easiest situation would only include only static obstacles with known characteristics
[13].

When a manufacturing robot is operating, all parts that need to be manufactured must be secured in order
to prevent slippage. There are several methods to secure parts to the workbench. One of these methods is
by using clamps. These clamps can hold the part to the workbench when forces are used on them when, for
example, drilling holes. Another clamping method is by placing the part into a mould. If a part is fixed into a
mould, the mould prevents the part from slipping when a force is exerted on the part.
3
The Wingbox Design
Throughout this chapter the calculations used to obtain the final wingbox design are discussed. In 3.1 a
description of the design goals are discussed along with constraints and the free design parameters. The
key design assumptions are also discussed in 3.1. 3.2 includes a description of the design process, including
analysis of the applied load and the calculation of the geometric properties. Furthermore, 3.3 goes through
every calculation and provides the mathematical evidence to finalise the design configuration. Finally, 3.4
includes the justification of the design with respect to environmental concerns.

3.1. Wingbox Design Description


The wingbox design has to fulfill the requirements and constraints provided by DELTA. The wingbox is to be
used in their new project, Project GreenWing, as the internal wing structure. This means that the goal of this
design is to design a wingbox which is to function as the internal, load-carrying structure of the Urban Air
Mobility aircraft. The design of the wingbox should also align with the goal of Project GreenWing itself, which
is to create sustainable travel for all.

3.1.1. Function and Design Objectives of the Wingbox


The wingbox will function as the internal wing structure of the Urban Air Mobility (UAM) aircraft. Hence
it should be able to resist the loads imposed on the wing due to the lift it produces. The most significant
load which it will have to carry is the bending moment. Besides this load, it should also carry the torsional
and shear loads. Under these loads it should also resist against skin buckling, shear buckling, and inter-rivet
buckling.

Since the aim of DELTA’s new project is to improve the environmental impact and design sustainable travel
for all, another function the wingbox should fulfill is to reduce the environmental impact of the product over
its lifetime, hence the sustainability of the wingbox is important.

Taking the function and goal of the project stated earlier into consideration, the overall objective of the wing-
box can be stated as optimising the environmental impact. In concrete terms this means that manufacturing
and operating resources should be optimised. As a result, CO 2 emission during manufacturing and operation
of the aircraft should also be optimised. This means that minimizing weight is not the main design objective,
although it does indirectly impact the CO 2 emissions.

3.1.2. Requirements and Constraints of the Wingbox


DELTA set a number of requirements and constraints of the wingbox design. They are listed below.
The following performance requirements apply:

1. Buckling of the wingbox is allowed to occur above 3 [kN ];

2. Failure should occur at 3.3 [kN ];

3. The structure must be optimized for sustainable manufacturing, operations end end-of-life processes.

Besides, DELTA set the following (dimensional) constraints:

1. The overall dimensions of the two parts of the wingbox connected by the sleeve are 2.87 x 0.4 x 0.15 [m];

2. Only few materials can be used, these are new and recycled aluminium, also see Appendix B;

3. The rest of the properties of the materials that may be used can be found in Appendix B.

7
3.1. Wingbox Design Description 8

To conclude, there are also a few manufacturing constraints which must be taken into account during the
design of the wingbox. These are:

1. Stringers must be present in each corner of the wingbox (to attach the side plates to);

2. The side sheets must be flush with the edges of the top and bottom sheet of the wingbox;

3. When a stringer crosses a rib, a hole must be cut in the rib in order for the stringer to pass freely through
it;

4. Side stiffeners on the side sheet must stick out 1 [mm] below and above respectively the bottom and
top sheet;

5. The stiffeners on the webs should be placed on the outside of the wingbox;

6. Rivet holes are not allowed to coincide with the stringers’ vertical flanges.

3.1.3. Assumptions made for the own Wingbox Design


While designing the wingbox, a few assumptions are made in order to make calculations of internal stresses
possible and in order to make the design process less complex. The three major assumptions that are made
during the design are:

1. Thin walled assumption


The thin walled assumption is used to calculate the shear stresses due to the torsion and the shear force.
The thin walled assumption can be used if the thickness is less than a tenth of the smallest dimension
of the cross-section. In the case the thickness is 0.8 [mm] and the smallest dimension is the height; 150
[mm]. The ratio of ht > 10, hence the thin walled assumption can be used in this case.

2. Thin sheet assumption


Assuming the plates to be thin sheet allows using the buckling formulas, which help determining at
which critical stresses the plates will buckle, therefore making this assumption is necessary, but simi-
larly to the thin walled assumption, is also valid given the thickness to length of the plates.

3. Weight of wingbox
The weight of the wingbox is negligible compared to applied forces. An order of magnitude calculation
of the final weight of the wingbox can give insight in whether the weight is negligible compared to the
applied loads. The wingbox has a length of 2.9 [m], and has two plates with base 0.15 [m] and two
with base 0.4 [m], all plates have thickness 0.8 [mm]. Assuming the to be four corner stringers along
the entire length, with dimensions given in Table B.3, from Appendix B, given an estimated volume
of 3.248 × 10−3 [m 3 ], using the density of aluminium, Table B.1, gives an estimated weight of 88.6[N ].
Comparing this with an applied load of 3[kN ], it can be concluded that weight of the wingbox can be
neglected. This simplifies the free body diagrams, hence simplifies the analysis.

3.1.4. Free Design Parameters of the Wingbox


Taking into account all the previously set constraints and requirements, a few parameters are left that can be
varied. The determination of these parameters should ensure that the previous performance requirements
are met, hence that the wing can sustain the set maximum loads. However, a large number of combinations
of the values for these design parameters can be chosen such that the requirements are met, that’s why the
design should be optimized for the objective given in Subsection 3.1.1. The following design parameters are
still left to vary:

• Choice of material;

• Number, location and spacing of longitudinal stringers and vertical stiffeners;

• Number and location of ribs;

• Number of rivets and rivet spacing.

The determination of these design parameters such that requirements are met and such that the design is
optimised for the objective, will be done in the remainder of the chapter.
3.2. Description of the Wingbox Design Process 9

3.2. Description of the Wingbox Design Process


Designing a wingbox is a complex undertaking, in which many failure modes are present and many internal
stress calculations need to be performed. The wingbox is loaded by a whiffletree which distributes a total ap-
plied force over four distinct attachment points of the wingbox. It is clamped at one side, hence it will act as
a cantilever beam. The wingbox will thus experience a shear force and a bending moment due to the applied
force. Furthermore, since the whiffletree is slightly offset from the longitudinal center line, the wingbox will
also experience a torsional moment.

All of the aforementioned forces and moments will lead to internal forces and moments that will be different
throughout the length of the beam. Since there are two separate force requirements at which buckling and
failure can occur, there will be two free body diagrams made; one for the buckling requirement (Figure 3.3.A)
and one for the failure requirement (Figure 3.3.B). The entire wingbox can effectively be divided up into five
separate parts, each with their own internal shear force, bending moment and torsional moment. The five
parts are separated by a rib, hence the wingbox will contain four ribs, one between each section. This also
means that each rib is placed at an attachment point of the whiffletree. This is the optimal placing for the
ribs since they will provide a direct load path to introduce the applied loads into the entire wingbox structure.
The dimensions of each section and the positioning of the ribs is shown in Figure 3.1.

Figure 3.1: Wingbox section dimensions and rib positions (own work)

A note on the material: The option is given to choose between a new, stronger aluminium, and a recycled, but
weaker aluminium. Properties for both are given in Table B.1. All calculations are done using properties of
the recycled aluminium. Then, it is checked whether using the new aluminium will result in weight savings,
reducing fuel consumption over its lifetime, hence making it more attractive. This is done in section 3.4.

Normal Stress

The wingbox will experience normal stresses due to the internal bending moment. Since the internal bend-
ing moment is different in each section of the wingbox, each section will also experience a different normal
stress. Therefore, determining the amount of stringers needed and the spacing of them to meet the set re-
quirements, can be done separately for each section.

The normal stress experienced by a given section plays a role in the failure mode of panel yielding. Next to
panel yielding, it also plays a role in top panel buckling since the top panel will be in compression. In case of
the yielding, the maximum normal stress in the cross-section has to be considered, which will be the panel
furthest away from the horizontal neutral axis. Therefore, the top and bottom panel will play a key role when
considering the normal stress.

The key structural elements that allows for controlling the magnitude of the normal stress, are the longitu-
dinal stringers. They will increase the moments of inertia. If only panel yielding had to be considered, a
symmetrical configuration of the longitudinal stringers would be ideal. However, since only the top panel
is in compression, hence can potentially buckle, the normal stress in the top panel cannot exceed a certain
critical normal stress and a certain stringer pitch will have to be met in order to avoid buckling. Therefore the
longitudinal stringer configuration will be asymmetrical, with the top panel containing more stringers. This
will also cause the center line to shift upward, which in turn means that the bottom panel will experience the
highest normal stress. The bottom panel will thus be considered when checking for panel yielding.
3.2. Description of the Wingbox Design Process 10

Shear Stress

Next to a normal stress, the wingbox will also experience a shear stress. This shear stress is caused by the inter-
nal torsional moment and shear force. Just like the normal stress, the shear stress is also different within each
of the five sections previously considered. This, again, allows for optimally determining how many structural
elements, in this case stringers, each section needs. This will lead to the lightest option, and consequently
the most sustainable option.

Also the shear stresses experienced by each section dictate a number of failure modes. The most obvious one,
again, is the shear stress exceeding the maximum allowed shear stress. This time, however, the maximum al-
lowed shear stress is not as straightforwardly obtained as was the maximum allowed normal stress, which
was the yield strength. The maximum allowed shear stress will be derived from the Tresca Yield Criterion.
It will end up being depended on the yield strength and the normal stress experienced. This will be further
explained in more detail in Section 3.3.4. Next to panel shearing, the shear stress also plays a role in shear
buckling of the side panels.

When considering the shear stress, the leading edge and trailing edge panels will be the main consideration,
since in these panels the shear stress will be the highest. The dependence of the magnitude of the shear
stress is, however, not that straight forward. The biggest contribution to the shear stress is the shear force, the
contribution of the torsion will remain small comparatively. In this case, it could be reasoned that increasing
the amount of longitudinal stringers, hence increasing the moment of inertia, would decrease the shear stress
experienced in both panels. Although this is the case, the effect is negligible. To see why this is the case, the
following relation is given:
VQ
τ=
It
By increasing the amount of longitudinal stringers, only the first moment of area Q, and the second moment
of area I are changed. Taking a closer look at what these two properties resemble in their most basic form will
reveal why adding a stringer doesn’t affect the shear stress significantly:

I = Σ(y 0 )2 (A 0 ) (3.1) Q = Σ(y 0 )(A 0 ) (3.2)

where y’ is the moment arm from the neutral axis to the center line of that element and A’ is the area of an
element. By adding an extra stringer, only the area is increased, the moment arm between the neutral axis
of the cross-section and the center line of the stringer remains the same. This is showed in Figure 3.2. Since
both I and Q scale linearly with A’, the ratio of the new I and Q when adding an extra stringer, effectively just
increasing A’, will nearly be the same ratio as without the stringer added. Therefor one way to meet the shear
stress requirement, would be to change the dimensions of the wingbox cross-section, increasing the moment
Q
arm y’, and thereby effectively changing the ratio of I 1 .

Figure 3.2: Effect of adding a stringer on the first and second moment of area (own work)

1 By using an asymmetrical longitudinal stringer configuration, the ratio of Q could be slightly altered since having more stringers, e.g.
I
on the top panel, would shift the neutral axis, in this case upward. This way y 0 can actually be altered, however this would decrease
Q
y 0 , thus the ratio I would increase due to the relation of I and Q with y 0 (Equations 3.1 and 3.2). Hence an asymmetrical longitudinal
stringer configuration would have an adverse effect on the shear stress, making it an ineffective method to decrease the shear stress.
3.3. Description of Calculations for the Wingbox Design 11

Since the outer dimensions of the wingbox are constraint, increasing the moment arm y 0 cannot be achieved
for this design. The only other option is to increase the maximum allowed shear stress. As discussed earlier,
it depends on the yield strength of the material, as well as on the normal stress experienced in the wingbox.
Increasing yield strength or decreasing the normal stress would result in a higher maximum allowed shear
stress. A higher yield strength can simply be achieved by choosing a different material, and decreasing the
normal stress can in turn be achieved by increasing the number of stringers, as previously discussed. If the
shear stress requirement turns out to be limiting, increasing the number stringers could indirectly influence
the shear stress requirement.

In terms of shear buckling, the vertical stringers are the structural elements that play a role. Shear buckling
is mainly a concern for the side panels since the shear stress is highest in those. The shear stress should stay
below a certain critical shear stress to avoid buckling. This critical shear stress depends on the material, but
also on the unconstrained dimensions of the panel. By implementing ribs and vertical stringers along the
length of the wingbox, the constraint length can be decreased, increasing the critical shear buckling stress.
The exact relation is explained in more detail in section 3.3.4.

3.3. Description of Calculations for the Wingbox Design


3.3.1. Cross-Section Properties of the Wingbox
After all assumptions, constraints and requirements were stated, calculations began on the wing box design.
The goal of these calculations is to obtain values for the free variables as listed in Subsection 3.1.4. Further-
more, this description aims to provide the reader with an overview of the design process.

The first step in the design process is to establish the failure criteria of the shear stress in the wing box. There-
fore, it it necessary to determine the main properties of the different parts that will later form the wing box.
These main properties include the Area Moment of Inertia and the center line of each part. These were cal-
culated using the general Equation 3.3 and Equation 3.4.

bh 3 ỹ A 0
P
I xx = + Ad 2 . (3.3) ȳ = PA (3.4)
12 A A0
The following values were obtained after these calculations (Table 3.1):

Table 3.1: Calculated values using Equation 3.4 and Equation 3.3 w.r.t. the cross-section centerline of the wingbox (own data)

Part Name Amount Available Center line y [m] Moment of Inertia I xx [m 4 ] Cross-Sectional Area A [m 2 ]
Sheet A 2 0 2.2E-07 1.2E-04
Sheet B 2 0 2.2E-07 1.2E-04
Sheet C 4 0.074 1.7E-06 3.2E-04
Stringer A 6 0.0055 2.2E-09 6.0E-05
Stringer B 10 0.0055 2.2E-09 6.0E-05
Stringer C 6 0.0055 2.2E-09 6.0E-05
Stringer D 10 0.0055 2.2E-09 6.0E-05

The Area Moment of Inertia values in Table 3.1 of the plates added give a total Moment of Inertia of the plates
(the basic cross-sectional shape) of 4.00E-6 [m 4 ].

The next step in determining the basic design properties is to find Q (the first moment of the portion of the
area about the neutral axis). These are calculated using Equation 3.5

y¯0 A 0
X
Q= (3.5)

Using the values for the Area Moment of Inertia and the First Moment Q two important values can already be
calculated: the shear stress acting on the wing box and the normal stresses acting in the wing box.
3.3. Description of Calculations for the Wingbox Design 12

3.3.2. Internal Force and Moment Diagrams of the Wingbox


Now that all requirements and considerations have been mentioned, calculations on forces can start. Using
standard Force Equilibrium Equations and the buckle and failure requirements for the panel, the Free Body
Diagrams and subsequently the Internal Shear Force, Moment and Torsion Diagrams can be drawn. It is im-
portant to note that the weight of the wingbox itself was neglected during this calculation, as the exact weight
is not known yet and since the weight load of the panel acting downwards will be relatively small compared
to the large forces that the whiffletree produces that act upwards. The final Free Body Diagrams and Internal
Shear Force, Moment and Torsion Diagrams for both the buckling situation and the failure situation can be
found in Figure 3.3.

(a) Buckling requirement (b) Failure requirement

Figure 3.3: Free body diagram with internal force- and moment diagram (own work)
3.3. Description of Calculations for the Wingbox Design 13

3.3.3. Bending Normal Stress Calculations


The first step in calculating the bending normal stress is calculating the critical normal stress value (σcr i t ).
This value is limiting because panel buckling in the face sheets of the wing box will be present beyond this
σcr i t (better known as thin sheet buckling). The critical normal stress due to panel buckling can be calculated
using Equation 3.6
µ ¶2
t
σcr = K c E (3.6)
b
Here, K c is a stress concentration factor determined by the clamping method and the width over height ratio
of the panel side view. E is the E-modulus (in [P a] and bt is the thickness over height ratio of the panel.
The geometrical properties in Equation 3.6 are already defined, as those dimensions were already given (see
Table B.3). Using Equation 3.6 and applying it on the several sections yield five different σcr i t i c al -values.
These values can be found in Table 3.2.

Table 3.2: σcr i t i c al -values for wingbox sections & through 5 (own data)

Section 1 2 3 4 5
Critical Stress [M P a] 36.3 36.3 16.1 4.0 1.0

The actual normal stress for each section can be calculated by using Equation 3.7

My
σ= (3.7)
I
With M the moment applied by the whiffletree in [N m], y the coordinate of the location at which the stress is
to be determined in [m] and I the Area Moment of Inertia in [m 4 ].

To analyze this normal stress a look must be taken at the location where the normal stresses are the largest.
This means a few assumptions with regard to the cross-section and calculations must be made. These are:

1. The part of the cross-section where the normal stress is the largest i.e. limiting must not contain one of
the ribs, as these help distributing the normal stresses;

2. It can be derived from Equation 3.7 that the normal stress is maximum at the largest y-coordinate. This
location is at either the top or bottom sheet of the wing box. In this case, it is maximum at the bottom
plate, due to the center line of the cross section being slightly above the neutral axis.

Using these calculations, the actual normal stress values can be obtained. These values are once again indi-
cated per section and can be found in Table 3.3.

Table 3.3: Normal stresses due to bending (own data)

Section 1 2 3 4 5
Normal Stress (top) [M P a] 36 21.5 10.2 2.66 0
Normal Stress (bottom) [M P a] 53.2 31.8 13.2 2.92 0

The values given in Table 3.3 can be compared with the limiting values in Table 3.2. Doing this gives that
only the normal stress on the bottom plate in section 1 (53.2 [M P a]) is exceeding the thin sheet buckling
criterion. However, the bottom part of the wingbox can be neglected in this comparison as it is in tension
and not compression and will therefore not buckle at all. All this results in the conclusion that the proposed
wingbox design will withstand the bending stresses during testing.
3.3. Description of Calculations for the Wingbox Design 14

3.3.4. Shear Stress Calculations


Obtaining the shear stress for a given section
The internal forces and moments that contribute to the shear stress are the torsion and the shear force. The
shear stress can be calculated for each of the five sections. The method to obtain the shear stress for each
section is the same. The shear stress induced in the cross-section by the torsion is given in Figure 3.4 and for
the shear force it is given in Figure 3.5. The thickness of the panels is not to scale, and the stringers were also
left out to simplify the diagram.

Figure 3.4: Internal shear diagram due to internal torsion (own work)

Figure 3.5: Internal shear diagram due to internal shear force (own work)

This means that in one of the side panels, the shear stress should be added, while in the other one it should
be subtracted. The panel in which it should be added is the leading edge panel, and it should be subtracted
for the trailing edge panel. The shear stress is maximum at the neutral axis of the cross-section.

Firstly, the shear stress due to the torsion can be calculated using the following formula, assuming the cross-
section is thin-walled, which was proved to be a valid assumption in section 3.1.3:

T
τt or si on =
2A m t

where T is the internal torsional moment, A m is the enclosed area of the cross-section and t is the wall thick-
ness. The thickness of the panel is t = 0.8[mm] and the enclosed area is calculated as A m = hb = 0.0596[m 2 ].
3.3. Description of Calculations for the Wingbox Design 15

Secondly, the shear stress due to the shear force is calculated using the following formula:

VQ
τshear f or ce =
2t I
where Q is the first moment of area, I is the second moment of area, t is the panel thickness at the neutral axis
(N.A.), and V is the internal shear force. Calculating the cross-section properties, Q and I, was explained in
section 3.3.1.

Then the shear stress in the leading and trailing edge become:

VQ T
τt r ai l i ng = −
2t I 2A m t

VQ T
τl ead i ng = +
2t I 2A m t
A stringer configuration was obtained in section 3.3.3, such that the panel buckling and normal stress require-
ments are met. This configuration can be used to check the shear stress and shear buckling requirement. If
these requirements are not met, the configuration should be adjusted in a next iteration to make sure the
requirements are met.

Maximum allowed shear stress requirement


The maximum allowed shear stress requirement requires the shear stress in the panels to remain below a
certain value when a total load of 3.3[kN ] is applied. Therefore, the free body diagram that can be used is
shown in Figure 3.3.B. The amount of longitudinal stringers for the top- and bottom panel is shown in Table
3.4. Additionally, the center line, the second, and first moment of area, the internal shear force and torsion,
and all relevant shear stresses are also given.

Table 3.4: Internal Shear Stresses (own data)

Section 1 2 3 4 5
n st r i ng er (Top Panel) 7 7 5 3 2
n st r i ng er (Bottom Panel) 2 2 2 2 2
Section Properties
Center line y [m] 0.0895 0.0895 0.0845 0.0785 0.075
Second moment of area I [m 4 ] 6.22E-06 6.22E-06 5.85E-06 5.41E-06 5.14E-06
First moment of area Q (at neutral axis) [m 3 ] 4.48E-05 4.48E-05 4.19E-05 3.85E-05 3.65E-05
Internal forces and moments
Torsion T [kNm] 0.0825 0.0540 0.0254 0.0085 0
Shear force V [kN] 3.30 2.16 1.02 0.34 0
Internal Stresses
Shear due to torsion [MPa] 0.866 0.566 0.267 0.089 0
Shear due to shear force [MPa] 14.9 9.72 4.55 1.51 0
Shear stress (leading edge) [MPa] 15.7 10.3 4.82 1.60 0
Shear stress (trailing edge) [MPa] 14.0 9.16 4.29 1.42 0

The maximum allowed shear stress per section can be derived from, the Tresca Criterion. The criterion says
that the maximum shear stress on an element should be less than half the yield strength of the material:
σY
τmax ≤
2
The maximum shear stress on an element depends on both the shear stress on it, as well as the normal stress
on it. Therefore, a Mohr’s circle should be constructed to derive how the yield stress, shear stress, and normal
stress are related. The Mohr’s circle, and the accompanying stress element are shown in Figure 3.6.
3.3. Description of Calculations for the Wingbox Design 16

Figure 3.6: Mohr’s circle with stress element of the wingbox (own work)

From Figure 3.6, it can be seen that:

σY σnor mal 2
r
τmax = = τ2al l owed + ( )
2 2

Rearranging this relation to the maximum allowed shear stress in a given panel yields:

1q 2
τal l owed = σY − σ2nor mal
2
From this relation, it becomes evident that lowering the normal stress increases the maximum allowed shear
stress. One thing should be noted, which is that the maximum yield stress and maximum normal stress are
in reality not present at the same point; the maximum yield stress is at the neutral axis, while the maximum
normal stress is in the bottom panel. However, since there are a lot of intermediate cases in which both a
normal stress and a shear stress act, assuming both the maximum shear and normal stress to act on the same
stress element serves as a limiting case. So, filling in the normal stresses per section obtained in section 3.3.3,
yields the maximum allowed shear stresses per section, given in Table 3.5

Table 3.5: Maximum allowable shear stress per section for the wingbox (own data)

Section 1 2 3 4 5
Yield Strength [M P a] 88.0 88.0 88.0 88.0 88.0
Maximum Normal stress [M P a] 53.2 31.8 13.2 29.2 0
Maximum Shear Stress [M P a] 15.7 10.3 4.82 1.60 0
Maximum Allowed Shear Stress [M P a] 35.0 41.0 43.5 44.0 44.0

Additionally, the maximum shear stresses are also shown in Table 3.5, which in this case is the shear stresses
at the neutral axis at the side of the leading edge. As can be seen, the shear stresses stay far below the maxi-
mum allowed shear stress, hence the configuration is sufficient to meet the shear stress requirement.

Shear buckling requirement


Since the side panels experience high shear stresses, they are vulnerable to shear buckling. A critical shear
stress can be determined which puts a constraint on the maximum shear stress that can be experienced by
the panels. The panels are not allowed to buckle up to a load of 3[kN ], hence the shear stresses experienced
under this load should remain below the critical shear stress. The value for the critical shear stress can be
determined through following equation:
µ ¶2
t
τcr = K s E (3.8)
b
where K s is the shear buckling coefficient, E is the elastic modulus, t is the thickness of the plate and b is the
unrestricted width of the panel. Since the shear stresses are highest in the leading edge panel, this will be the
limiting case. In that case, the thickness is 0.8 [mm], the unrestricted width is the height of the panel, which
is 0.1484 [mm] and the elastic modulus of the material is 75 [GP a]. That only leaves the K s as a free variable
which can be influenced. Its value is determined from Figure 3.7, obtained from [1].
3.3. Description of Calculations for the Wingbox Design 17

Figure 3.7: Shear buckling coefficients for skin-stiffener panel [1]

For the side panels of the wingbox, case 3 was picked, with top and bottom being hinged and the sides being
clamped. Since the side panels are divided by ribs on the inside, each section can be considered to have a
unrestricted length a equal to their respective lengths and unrestricted width b equal to their height. Case 3
was picked because each of these sections are part of one complete panel, except for the third section which
is split up by the sleeve. The top and bottom side of each section is only connected by a stringer to the top
panel, hence is closest approximated by the hinged case.

For each of these sections, a required K s value can be calculated using equation 3.8. This K s value corresponds
to a certain ba , from which the unrestricted length a can be determined. If this calculated length a is shorter
than the length of the section, then a vertical stringer is needed to break up the section into smaller ones such
that the unrestricted length a yields a satisfactory value for the K s . Table 3.6 shows this determination of the
amount of vertical stringers per section. Since the leading and trailing edge each experience a different shear
stress, a distinction has been made between them.

Table 3.6: Vertical Stringers for Leading and Trailing Edge Panel (own data)

Section 1 2 3 4 5
Shear Buckling Coefficient K s (LE Panel) 7.73 5.06 2.69 0.79 0
Shear Buckling Coefficient K s (TE Panel) 6.88 4.50 2.11 0.70 0
Maximum ba (LE Panel) 1.5 4.2 >6 >6 >6
Maximum ba (TE Panel) 2 >6 >6 >6 >6
Maximum length a [m] (LE Panel) 0.223 0.623 >0.890 >0.890 >0.890
Maximum length a [m] (TE Panel) 0.2968 >0.890 >0.890 >0.890 >0.890
Section length [m] 0.45 0.60 0.70 0.60 0.55
Required Vertical Stringers (LE Panel) 2 0 0 0 0
Required Vertical Stringers (TE Panel) 1 0 0 0 0

From Table 3.6, it can be seen that only the first section needs vertical stringers in order for the side panels not
to buckle. More specifically, the leading edge panel needs two vertical stringers and the trailing edge panel
needs one vertical stringer.
3.3. Description of Calculations for the Wingbox Design 18

3.3.5. Rivet Failure Calculations


Rivet spacing to resist buckling
Buckling between rivets can occur on the top plate as it is compression. Therefore to ensure the structure
buckles above 3[kN ], Equation 3.9 can be used [14]:
µ ¶2
t
σi r = 0.9cE (3.9)
s
"σi r " being the stress applied to the thin plate, and "c" being a constant associated with the rivet used, in this
case 2.1 [14]. "E " is the Young’s modulus of the material, "t " is the thickness of the thin sheet and "s" is the
spacing between rivets.
σi r is obtained from the maximum applied stress due to bending moment, in each section. 3.3
Rearranging Equation 3.9 for "s" allows the relevant values to be inputted:

t
s=q (3.10)
σi r
0.9cE

This gives the results given is Table 3.7

Table 3.7: Maximum rivet Spacings to resist buckling (own data)

Section 1 2 3 4
Spacing [mm] 48.53 62.74 90.96 178.45

Rivet spacing to resist shear


The rivets will also be loaded in shear. For calculating the required rivet spacing the shear flow is required,
which can be calculated from the values in Table 3.4 using equation 3.11. The shear flow in the rivets con-
necting the top plate is calculated from the equation 3.12

q = τ·t (3.11)
V ·Q
q= (3.12)
I
To use equation 3.12, Q must be known.
Q = ȳ · A (3.13)
2
In equation 3.13 A is 320 [mm ] and ȳ can be obtained from table 3.8. Using equation 3.13 the values in Table
3.8

Table 3.8: First moment of area Q values for shear in sheet (own data)

Section 1 2 3 4 5
ȳ to top plate [mm] 60.1 60.1 65.1 71.1 74.6
ȳ to bottom plate [mm] 88.3 88.3 83.3 77.3 73.8
Q at top plate [mm 3 ] 1.92E-05 1.92E-05 2.08E-05 2.28E-05 2.39E-05
Q at bottom plate [mm 3 ] 2.83E-05 2.83E-05 2.67E-05 2.47E-05 2.36E-05

Using the data from Table 3.8 and the equation 3.12 with the shear force of 3.3[kN ], the q values in Table 3.9
are obtained.

Table 3.9: Shear flow in each section of the wingbox (own data)

Section Shear flow [kN/m]


Section 1 2 3 4
Top 10.2 6.68 3.62 1.43
Side-leading edge 25.2 16.5 7.71 2.56
Side-Trailing edge 22.4 14.7 6.86 2.27
Bottom 15.0 9.82 4.63 1.55
3.3. Description of Calculations for the Wingbox Design 19

To obtain the relevant spacing to resist the calculated shear flows the equation 3.14 is used :
F r i vet
s=³ (3.14)
q
´
n st r i ng er s

With F r i vet being the maximum force a singular rivet can withstand, in this case 1060[N ], And n st r i ng er s the
amount of stringers to be riveted to the plate loaded in shear, found in Table 3.4. The results of this calculation
is found in Table 3.10.

Table 3.10: Maximum rivet Spacing to resist shear (own data)

Section Spacing [mm]


Spacing Location 1 2 3 4
Top 727.10 1111.50 1464.39 2227.36
Bottom 141.29 215.98 457.48 1365.88
Side Front 168.47 128.76 274.95 829.16
Side Back 141.98 144.69 309.20 933.06

Final Maximum Spacing


Through comparing the required spacings for buckling and shear it becomes apparent that buckling is far
more critical for the rivet spacing in the top plate. Therefore we take the smaller spacing to ensure the buck-
ling does not occur.

Table 3.11: Maximum rivet Spacings (own data)

Section Spacing [mm]


Spacing Location 1 2 3 4
Top 48.53 62.74 90.96 178.45
Bottom 141.29 215.98 457.48 1365.88
Side Front 168.47 128.76 274.95 829.16
Side Back 141.98 144.69 309.20 933.06

This results in the rivet spacing in Table 3.11. A key point to note is that the calculated values are the maximum
rivet spacing therefore a Smaller spacing can be used to ensure there is little risk of rivet failure. Furthermore
only a whole number of rivets can be used so a better rivet spacing is found with the following steps. To do
this, Equation 3.15 is used:

L
n r i vet = +1 (3.15)
s max
"L" being the available length, in this case the length of each section ,as seen in figure 3.1, and "s max " being
the maximum rivet spacing. The results of this equation are rounded up to ensure there is more rivets than
required. The results of these calculations are compiled in Table 3.12

Table 3.12: Minimum Rivets in each section (own data)

Section Number of rivets


Rivet Location 1 2 3 4
Top 11 9 6 4
Bottom 5 4 2 2
Side Front 2 3 2 2
Side Back 3 3 2 2

To calculate the new rivet spacing, Equation 3.16 is used:

L
s= (3.16)
n r i vet − 1
3.3. Description of Calculations for the Wingbox Design 20

Table 3.13: Adjusted spacings for whole number of rivets (own work).

Section Spacing [mm]


Spacing location 1 2 3 4
Top 45 75 140 200
Bottom 113 200 700 600
Side Front 150 75 150 150
Side Back 75 75 150 150

Table 3.13 Shows the maximum spacing if all rivets are to have equal spacing.

3.3.6. Final Stringer Configuration of the Wingbox


Having determined the design parameters per section, they still need to be compiled into one complete and
coherent design configuration of the wingbox. Thereby ensuring that there is structural continuity between
sections such that the applied loads have a path through which they are introduced into the support and
stress concentrations are prevented. Up till now, it was assumed that the longitudinal stringers would be
evenly distributed across the width of the panel, however this would lead to discontinuities between sections,
therefore the configuration shown in Figure 3.8 was derived.

Figure 3.8: Wingbox stringer configuration (own data)

In order to ensure this structural continuity, the spacing of the stringers in the third section is not even any-
more. There is a bigger gap between the two outer stringers at the top and bottom. The gap between them is
roughly 130 [mm]. Calculating the maximum allowed distance to prevent buckling can be done with equa-
tion 3.6. This gives a maximum unrestricted width b of 130 [mm]. Hence spacing the stringers as shown in
Figure 3.8 is still sufficient to prevent buckling up to 3 [kN ].
3.4. Design Justification for the Wingbox - Sustainability 21

3.4. Design Justification for the Wingbox - Sustainability


The production of the materials used in the wingbox design comes with a price with regard to environment
as well as cost. As some of the design objectives of this wingbox are to minimize weight and minimize emis-
sions, these factors were also taken into account in the design. First, it will be explained why cost is taken as
measurement for the environmental impact of the design. Afterwards, major design choices with regard to
environment will be explained and justified by calculations.

The total CO 2 -emissions of the wingbox production and design can be divided in different parts:
1. Emissions from manufacturing;
2. Emissions from operations of the UAV;
An overview with all details of the emissions involved can be seen in Table B.2. As can be seen in this table, the
emissions are related to the panel weight, as more aluminium means more emissions. Also, from the table
one can see that the emissions can be expressed in a certain cost per tonne emission. This means that we
can express the saved weight and the emissions both in cost, making cost the perfectly suitable parameter to
minimize in the design.

For the proposed design previously discussed, recycled aluminium was used. Using a calculated total volume
of the aluminium used in the wingbox gives an required aluminium mass of 11.7 [kg ]. This mass of recycled
aluminium will have CO 2 -emissions of 3.1 [kg ] , this costs 0.13 euro cents. If new aluminium was used, the
same calculation would result in an emission of 12 [kg ] CO 2 with a cost of 0.54 euro cents. In this calculation
for the new aluminium, it is assumed that the mass of aluminium used stays the same. In reality however, this
does not have to be the case, as the new aluminium is slightly stronger and thus there is less new aluminium
needed for the wingbox production. The proposed design is more environmentally friendly, as it has lower
CO 2 -emissions.

Now that the emissions of new and recycled aluminium (and their costs) are known, a more detailed look can
be taken at the total cost of the design. Now, also the total mass of the wingbox is considered. The UAV uses
0.2 liters of fuel for every kilogram it has to fly around. Using the total mass of the wingbox, it is obtained that
the UAV will need 2.4 liters of fuel to carry the wingbox design. This gives an emission of 5.17 [kg ] which costs
0.22 euro cents.

A general overview of all emissions per wingbox part and their costs can be seen in Tables 3.14 and 3.15. It
can be seen that the wingbox design made of recycled aluminium, as discussed in this chapter, is the most
environmental-friendly option. That is also why this material choice was made.

Table 3.14: CO 2 -emission properties of the Wingbox design made of new aluminium (own data)

New Aluminium Mass [kg] CO 2 -emission [kg] Emission Costs [EUR]


Panels 10.92 9.78 0.42
Stringers 0.30 1.08 0.05
Ribs 0.52 1.87 0.08
Operational N/A 5.17 0.22
Totals 11.75 17.90 0.76

Table 3.15: CO 2 -emission properties of the Wingbox design made of recycled aluminium (own data).

Recycled Aluminium Mass [kg] CO 2 -emission [kg] Emission Costs [EUR]


Panels 10.92 2.89 0.12
Stringers 0.30 0.082 0.00
Ribs 0.52 0.14 0.01
Operational N/A 5.17 0.22
Totals 11.75 8.28 0.35
4
Manufacturing of the Wingbox
Throughout this chapter some manufacturing considerations are made like manufacturability of the design
and automation of the manufacturing process. In 4.1 the manufacturability of the wingbox design is dis-
cussed. Then, 4.2 gives a description of a robot motion plan for a KUKA robot which could be used to au-
tomate the drilling of holes. Finally, 4.3 a feasibility assessment is made between manual and automated
manufacturing to see which degree of automation would be most cost effective. The complete manual pro-
duction plan including time estimations for each step can be found in Appendix E.

4.1. Discussion of Manufacturability of the Wingbox Design


The manufacturability of any design depends on the following factors:

• Product Complexity

• Product Variants

• Component Availability and Prize

• Production Friendly Design

4.1.1. Product Complexity of the Wingbox


The product complexity affects the amount of time needed to complete the product. By making a product
less complex, the production time might be reduced massively.

The design of our wingbox would not be considered to be complex. This is because the type of parts needed
are all basic parts (stringers, aluminium plates, ribs and rivets). There are no special systems or unique parts
needed for this design. The only parts that are somewhat unique for our wingbox are the ribs delivered by the
IAC.

4.1.2. Product Variants & Production Friendly Design of the Wingbox


The product variants describes the level of variance there is in all the parts that are from the same category.
This variance might be caused by a difference in the hole spacing or hole layout or the amount of cutouts a
part has.

Almost all parts involved in the wingbox assembly have unique hole and cut-out spacings. This design deci-
sion results in a far from ideal manufacturability, since having parts with the same kind of spacing will result
in a way less time-consuming and error-sensitive production process. The only parts that have some sort of
repetition are:

• Five top L-stringers A are similar in hole spacing.

• Three vertical side stringers are similar in hole spacing.

• Of the four ribs, the hole spacing on the side flanges and on the bottom flange are the same.

22
4.2. KUKA Robot Motion Plan 23

4.1.3. Wingbox Component Availability and Price


Whether a component is widely available or not has a use effect on the manufacturability of a part, since very
rare parts often cost way more and are harder to obtain.

For our wingbox design all available materials are delivered by IAC, as seen in Appendix B. The prices of the
parts used are relatively cheap, as seen in Table 4.1:

Table 4.1: fill something in person who knows what they are doing

Average Prices
Stringers 17.5 [m] => 41 [EUR]
Sheets 2.4 [m 2 ] => 72 [EUR]
Ribs 0.8 [m 2 ] => 25 [EUR]
Total Approx. Cost 138 [EU R] (not including rivets or any machinery)

These parts are relatively cheap, since they are widely available, and they are easily produced, as they can be
simply molded or folded. They are also all made out of aluminum which is a common material, and common
means cheap.

4.2. KUKA Robot Motion Plan


4.2.1. Usage of KUKA Robot for Automated Manufacturing of the Wingbox
DELTA wants to assess the feasibility of incorporating some automation into the manufacturing process of
the wingbox. Specifically, a robot should be programmed to efficiently drill all of the holes in the parts of the
wingbox. In order to have to robot accurately drill the holes, the robot should also swap between parts and
place them onto a sacrificial plate. Parts should also be clamped as to not have it move while being drilled. In
order to visualise, and estimate processing time, this drilling process should be animated in a Robot Motion
Plan. This is a virtual robot which runs on a robot program, as it would be a real robot running on the pro-
gram. This allows to visualise the robot program.

Since some parts of the wingbox have more than a hundred holes, it can be very beneficial to have them
be drilled automatically using a robot. Having a robot drill the holes will also result in a more accurate and
reproducible product than doing it manually. Therefore, a motion plan is made which shows the automated
drilling by the robot for some of the wingbox parts.

4.2.2. Description of Robot Motion Plan


The goal of this robot motion plan is to be able to estimate how much time it takes to manufacture a given
number of wingboxes using the KUKA robot as part of the manufacturing process. Using this estimate, it can
then be decided whether utilizing the KUKA robot would indeed be more beneficial than doing the complete
manufacturing manually. This is done in section 4.3.

In order to display the automated drilling of wingbox parts by the robot, and eventually estimate the time
the robot would take to drill all holes in all parts of the wingbox, only a carefully selected number of parts was
chosen to animate the drilling process for. This is done to efficiently make an estimate, and if it turns out that
using the KUKA in the manufacturing process is beneficial, programs can be written to drill all holes, ready
for assembly.

The parts that were chosen to be animated are:

• Corner Stringer (Part: AE1222-I-WB-ST-24);

• Longitudinal Stringer (Part: AE1222-I-WB-ST-02);

• Rib (Part: AE1222-I-WB-RIB-15);

• Top Inner Panel (Part: AE1222-I-WB-Plate-01);

• Leading Edge Side Panel (Part: AE1222-I-WB-Plate-12).


4.2. KUKA Robot Motion Plan 24

A distinction is made between corner stringers and other kinds of stringers, like the longitudinal ones, be-
cause the corner stringers needs to be drilled on both flanges, hence they use a different drilling sequence.
Since there are stringers with different amount of holes, the time to drill one hole can be measured, and the
time to swap the stringer with another part, and then a linear relation can be established to estimate the
drilling time as a function of the amount of holes in the form of:

T t ot al = t onehol e n hol es + t sw appi ng −t i me

The same kind of relation can also be established to estimate the drilling time of the remaining ribs, top- and
bottom panels and side panels, based on the corresponding parts that were animated.

The manufacturing set-up and process using the KUKA robot is explained next:

Swapping of Parts
The parts that need to be drilled are initially positioned on a separate table(not the drilling one). To transfer
the parts to the drilling table one by one, a swapping process needs to be performed by the KUKA robot. The
tools and process used differs between the different parts.

For the plates and the ribs, a suction tool is used to transfer from the initial table to the drilling table and
back. The plates are transferred to a specified position on the drilling table that allows the clamps to clamp
the plates to the drilling table.

For the stringers, a gripper tool is used to transfer from the initial table to the drilling table and back. The
stringers are squeezed between the fingers of the gripper to allow the robot to transfer it from the initial table
to the drilling table and back. Due to the excellent maneuverability of the KUKA robot, it is possible to orient
the stringer correctly and place them in the correct position on the drilling table to allow the clamps to clamp
the stringers to the drilling table.

It is important to note that an extra maneuver is required for the corner stringers since they have to be drilled
on both flanges. After one flange is drilled, the corner stringer is grabbed by the gripper, flipped using the
robot motion, and then placed onto the table again to allow for the drilling of the other flange.

The swapping sequence is displayed for one part in Figure 4.1.

(a) Picking up part (b) Transferring part (c) Placing down part

Figure 4.1: Swapping sequence of a wingbox part by the KUKA Robot (own work)

Swapping between Robot Tools


The KUKA robot is required to swap between robot tools to perform different steps of the program. There are
three tools that the robot uses: the suction tool, the gripper tool, and the drill tool. In the animation, the tools
are simply swapped in an instant by only showing the tool that is required to perform the current task. In
reality, however, the tool changing is more advanced since the tools need to be attached and detached. This
can be done by adding a tool stand that allows the robot to temporarily store a tool when it is not being used.
This additional tool changing time will be accounted for when determining the feasibility of the automatic
production in a later stage.
4.2. KUKA Robot Motion Plan 25

Clamping Set-Up
To make sure the parts do not move when being drilled by the KUKA robot, they should be fixed in a spe-
cific position. This is done using two different clamping systems: one for the plates and ribs and one for the
stringers. Both clamping systems are designed in CATIA V5.

The clamping system for the plates consists of three clamp screws on the side of the drilling table. In the
animation, the clamp screws are only moving inwards and outwards to allow for the plates to be placed in
the correct position without tightening the clamp screws. In reality, the clamp screws should be tightened to
clamp the plate to the sacrificial plate and the drilling table. The tightening of the clamps will be accounted
for when determining the feasibility of the automatic production in a later stage.

The clamping system for the stringers consists of multiple pneumatic clamps in the middle of the drilling
table, oriented towards the sacrificial plate. When a stringer is placed on the edge of the sacrificial plate, the
shafts of the pneumatic clamps push against the stringer and subsequently against the sacrificial plate. Mo-
tion of the sacrificial plate is resisted by multiple fixed blocks on the other side of the sacrificial plate. The
pneumatic clamps are automated and do therefore not need any manual adjustment.

(a) Top view (b) Isometric view

Figure 4.2: Kuka Robot setup (own work)

Drilling of Ribs
The ribs are pieces of sheet metal which are cut to shape and then goes through a forming process to make the
flanges of the ribs. However, drilling the holes after the forming process is quite impractical since the force
applied to the flanges to drill the holes of the flanges might deform them. The forming process should be
done after drilling the holes in order to prevent this. Therefore, the holes of the ribs are drilled in the unfolded
piece of sheet metal of a rib. The drilling sequence is displayed for the rib in Figure 4.3. Note however, that
the drilling sequence is similar for each part.
4.3. Feasibility Assessment between Manual & Automated Manufacturing 26

Figure 4.3: The rib being drilled with a drill tool in RoboDK (own work)

Visualization Errors
In the robot manufacturing animation, there are two important things to note that need to be taken into
account when using the KUKA robot in real life:

• When picking up the top plate, side plate and corner stringer, the orientation of these parts flips for
very short amount of time. This is a small bug in the program, but should not influence the robot when
operating in real life;

• When the KUKA robot places the corner stringer on the drilling table to drill the second flange of the
stringer, the end of the robot is positioned slightly inside the table. When operating in real life, this can
be avoided by using a smaller table, since the right side of the table will not be used for anything.

4.3. Feasibility Assessment between Manual & Automated Manufacturing


Now that both the manual production plan and robot motion plan are described, the feasibility of each will be
compared to determine an ideal manufacturing plan for large-scale production of the wingbox. This can be
either fully manual, or partly automated (i.e. hybrid). It is not possible to fully automate the manufacturing
process since the KUKA robot is only capable of performing the drilling steps.

To compare the feasibility of the manual production and the robot motion plan, the following costs (per hour)
will be taken into account:
Table 4.2: Manual and machine manufacturing costs. From [1]

Category Rate [€]


Manual labour 90/hr/person
Overhead - Manual labour (Use of facilities and tools) 35/hr
KUKA run time 1000/hr
KUKA setup mould/clamping system 250 per setup

4.3.1. Manual Manufacturability of the Wingbox


Pro’s:

• External programming isn’t needed for anything as humans can take simple oral commands or human
written commands. this is a pro, as making a good program can be time consuming.

• If something ends up going wrong or the manufacturing plan is a bit confusing, the workers can easily
adjust and improvise to make sure the wingbox is still manufactured, while an error in the program of
KUKA would simply stop the whole process.

• Working fully manually, allows for task distributions and shorter working periods.
4.3. Feasibility Assessment between Manual & Automated Manufacturing 27

Cons:

• Human workers are prone to errors, which can hinder the quality of the final product. This is especially
true after long work hours. This then leads to necessary breaks, breaks take time and in this case time
costs money as the workers are paid hourly.

• Managing 20 people can end up being difficult; requires a more complex planning and the efficiency of
big groups can sometimes be lower than expected if coordination and planning is not executed prop-
erly.

Using, mainly past experience on team production processes and basic understanding of the tools at hand,
a production timeline was created, by assigning estimated duration for each individual task of the produc-
tion process, assigning different numbers of workers depending on the task and then overlapping the most
amounts of tasks possible to make the theoretical timeline of the wingbox production as efficient as possible.

Production timeline:

The timeline can be seen in the Figure C.1 of Appendix C.

From this timeline, and based on the prices mentioned in Table 4.2 the final cost for the manual production
of the wingbox is thought to be: 3448.75 [€].

4.3.2. Partly Automated Manufacturability of the Wingbox


Pro’s:

• KUKA robot is fast and efficient for repetitive manufacturing steps, that have some sort of pattern, like
drilling the panels of the wingbox, if the programming is done well, enter the spacing, and distances
from the edge and the drilling should be done very fast.

• The robot can simply make sure the drilling is done perpendicularly and straight down through the
panels and stringers. While for a human, this takes time and is still prone to error, aligning the drill
properly is meticulous and a human worker will never be perfect while KUKA can be.

• KUKA can work without supervision once the pieces are set up and clamped properly, it could if neces-
sary work while human workers are taking a break, making the manufacturing process efficient as no
time is wasted.

Con’s:

• The KUKA is expensive to use and setup; the hourly wage is 11 time more expensive than a human
worker. This means, that for it to be economically beneficial, the KUKA would have to execute a given
task 12 times faster than the human. Taking into account the setup time for each part to be drilled with
the KUKA, the total time to complete the task using the KUKA might actually end up being slower than
12 times faster the time for the human to complete the task. This would mean the robot can possibly
not be so beneficial; budget wise.

• Setting up the program to KUKA requires skills (adds work-hours which are not included in this budget);
if not operated and programmed properly, holes could be drilled at the wrong locations. This could
prevent the wingbox to be properly manufactured and assembled.

Production Timeline:

The timeline can be seen in the Figure C.2 in Appendix C.

From this timeline, and based on the prices mentioned in table 4.2 the final cost for the partially automated
production of the wingbox is thought to be: 2422.08 [€].
4.3. Feasibility Assessment between Manual & Automated Manufacturing 28

4.3.3. Conclusion
Overall, the KUKA manufacturing option is deemed to be the best due to its fast, efficient and cheap pro-
duction. The KUKA production only take 145 [mi n] compared to 255 [mi n] for the manual production. It is
also estimated to only costs 2422.08 [€] compared 3448.75 [€] for the manual production. Clearly, the KUKA
production option is superior.

This is all determined from the created timetables in appendix C. However, these timetables are not based on
any experimental production times; therefore, general margins were added based on basic understanding of
the handling of the tools at hand. These margins were added to compensate for the fact that these duration
are simple estimates. This does reduce the accuracy and reliability of the timelines; however, they are good
tools to use for general planning, and budgeting.
5
Testing of the Alfa Wing
This chapter discusses why and how the wingbox will be tested, and what happened during the actual test.
This chapter starts with Section 5.1 which discusses what is wanted of this test and why it is performed. The
Section 5.2 subsequently describes how this test is actually going to be performed. Section 5.3 then goes more
in dept on what the test set up actually does to the test subject and Section 5.4 describes what was seen and
measured during the test.

5.1. Description of Test Objectives for the Alfa Wing


With every design, of course, must come testing. This is done to ensure that the final product functions safely
under the design loads. Due to time constraints, not every wingbox can be tested individually. Instead, a so-
called “Alfa Wing” will be tested. This wing is designed by the IAC Manufacturing department, and the data
from the test will be used to analyze the personal design calculations while keeping the differences between
the two designs in mind.

The Alfa Wing is not entirely representative of the design, but the test data will nevertheless be useful in
evaluating what can be improved. The goal of the test is to answer the following question: do the design
calculations accurately represent the test outcome? The purpose of answering this question is to determine
whether the wingbox designed is actually capable of fulfilling the requirements set. This kind of thinking will
allow for the evaluation of the wingbox design proposed in this report.

5.2. Description of Test Systems


The wingbox will be tested using a whiffletree. A whiffletree is a machine that can distribute a single load over
multiple fastening points. Essentially turning a point load into a distributed load. The more fastening points
there are, the more accurate the distributed load will be.

The whiffletree used for this test will be a whiffletree that uses 4 fastening points. In reality the wingbox
will experience a distributed load of course, so it is not an entirely accurate manner to test the wingbox. How-
ever a nice approximation can be made.

Below in Figure 5.1 is a schematic drawing of the whiffletree. As can be seen the fastening points are about
equally spaced with distances of 600-700 [mm] between them. Point S is not in the middle of the whiffletree
but more to the left. This will cause the forces at the first and second points to be greater than at the third
and fourth points, just like the lift on a wing is greater close to the fuselage than far from it. More on this in
the section where the loads of the whiffletree are calculated. Another picture of the final set-up can be seen
in Figure 5.2[15].

29
5.2. Description of Test Systems 30

Figure 5.1: Schematic drawing on how the whiffletree will be fastened to the wingbox[1]

Figure 5.2: The set-up of the Alfa wing test just before testing[15]

Before the test the wingbox will be mounted to a steel frame and clamped at the root side to the frame. During
the test a load will be introduced in point S. The whiffletree distributes this load over the wingbox causing it
to deform. While measuring the load in point S, the deformation will be measured by 22 strain gauges located
all over the wingbox. The precise locations of each sensor can be seen in Figure 5.3.

Figure 5.3: Locations of the strain gauges on the wingbox[1]


5.3. Calculation of Load Distribution of the Whiffletree 31

5.3. Calculation of Load Distribution of the Whiffletree


To calculate the load distribution on the whiffletree the method of sections was used. A reference image used
for the calculations can be found in Figure 5.4 A load will be applied at point S. This load will translate to both
ends of the bar. However these ends do not have the same distance to point S. The end closest to the root of
the wingbox is 370 [mm] away. The furthest end is 830 [mm] away. This gives the formulas:
X
= FS − F A − FB = 0 (5.1)
y
X
= 0.37 ∗ F A − 0.83 ∗ F B = 0 (5.2)
MS

Figure 5.4: Whiffletree points[1]

If we take F S as the variable this will result in the equations

0.83
FA = FB (5.3)
0.37
and thus

37
FB = FS (5.4)
120
making

83
FA = FS (5.5)
120
Now F A is further divided into FC and F D . As both FC and f D are 300 mm away from where F A connects to
the lower left bar, these forces will be equal in magnitude.
X
= F A − FC − F D = 0 →
− F A = 2 ∗ FC = 2 ∗ F D (5.6)
y

Thus

FA 83
FC = F D = = FS (5.7)
2 240
On the right side of the whiffletree things get slightly more difficult as The distance from E to B is not the same
as from F to B. This calls for a moment equation.
X
= 0.2 ∗ F E − 0.4 ∗ F F = 0 →
− FE = 2 ∗ FF (5.8)
MB

X 3
= FB − FE − FF = 0 →
− FB = 3 ∗ FF = ∗ FE (5.9)
y 2
37
FE = FS (5.10)
180
37
FF = FS (5.11)
360
All these equations were set up under the assumption that the wingbox is static and does not deform too
much. As F S increases and the wingbox starts to deform, these equations will get less and less accurate. A
complete overview of the calculations results can be found in the Free Body Diagram in Figure 5.5.
5.4. Description of Test Events for a Wingbox 32

Figure 5.5: FBD of the wingbox (own work).

5.4. Description of Test Events for a Wingbox


As the force on the whiffletree gets bigger the wingbox can be seen bending upwards. When the wingbox
is bent a certain amount the top part of the wingbox starts to buckle near the basis at which the wingbox
is clamped into the set up as can be seen in Figure 5.6 [16]. The wingbox can also be seen twisting slightly,
because the force of the whiffletree is not applied through the centre line causing a torsional load.

Figure 5.7 shows the displacement of the leading and trailing edge of the wingbox at 1750 [mm] from the
clamp edge vs the force on the main whiffletree attachment. From this can be seen that there is is a slight
rotation of the wingbox as these displacements are not equal. Also the maximum force on the wingbox seems
to be 3295.7 [N ], and the maximum displacement seems to be around the 150 [mm].

The strain on the sleeve is depicted in figure 5.8. These values were obtained from strain gauges oriented
along the sleeve as depicted in Figure 5.3, at a 30°angle with the rest of the set up. Gauge A the strain on the
gauge on the bottom of the sleeve, and gauge B the strain on the top of the sleeve.

For the gauges that will be analyzed gauges 1,2,7,8 were chosen both on the inboard and outboard side. In
Figure 5.9 the strain vs force is given for the strain gauges on the inboard side of the wingbox. Figure 5.10
depicts the strain versus force on the outboard gauges.

Figure 5.6: The buckling that can be seen at the basis of the wingbox.[16]
5.4. Description of Test Events for a Wingbox 33

Figure 5.7: Graph showing the displacement of the two displacement sensors vs the force on the whiffletree.(own work)

Figure 5.8: The strain measured at position A and B.(own work)


5.4. Description of Test Events for a Wingbox 34

Figure 5.9: The strain on the selected gauges on the inboard side.(own work)

Figure 5.10: The strain on the selected gauges on the outboard side.(own work)
6
Analysis of the Wingbox Testing
This Chapter deals with processing the information obtained from the tests as described in Chapter 5. In this
Chapter the differences between the tested Alfa wingbox and the designed wingbox are discussed, and the
test on the tested Alfa wingbox is analyzed.

In Section 6.1 the differences between the Alfa wingbox that was tested and the wingbox described in this
report’s Design Chapter (Chapter 3) are discussed and it is discussed how these differences influenced the
Alfa Wing test. Section 6.2 describes the calculations used to design the wingbox from Chapter 3 are used to
estimate what to expect from the Alfa Wing tests with regard to buckling and failure et cetera. In Section 6.3
the choice of gauges to use to obtain the Alfa Wing testing data and the justification for this is presented and
in 6.4 the data from these strain gauges is discussed. Section 6.5 then explains how the wingbox failed in the
test. Then, in Section 6.6, the data from the test and the data from the calculations are compared. Section 6.7
then puts all of this together to use the values obtained from the Alfa test to draw a conclusion on how the
wingbox designed in this report would perform in a similar test.

6.1. Discussion on Differences Between Own Design and Alfa Wing Design
Now that the Alfa wing has been tested, a comparison must be made between the design tested and the
design created based off aforementioned calculations. There are quite a few differences in the Alfa wing’s
design and the design proposed in this report, let’s call this design the "E10E Wing". Keep in mind that this
is a description of the differences between the two wingboxes and can be quite unclear. The differences can
be best understood using a quick visual analysis. For reference, look at the technical drawings of the two
wingboxes in Appendix D.

6.1.1. Inboard Part of the Wingbox


The most striking difference between the two wingboxes is that the Alfa wing places the stringers on the bot-
tom panel while the E10E wing places them against the top panel. Furthermore in the E10E wing, 5 stringers
start at the root of the wing and end after 1 [m]. Here the two outermost stringers end completely and the 3
innermost are continued with 3 other stringers that continue until the end of the inboard. The alpha wing
uses 3 long continuous stringers from the root until the end of the inboard.

There is also a difference in riveting. Where the rivets in the alpha wing leave some space at the root and at
the sleeve end, the rivets in the E10E wing are continuous over the whole stringer from start until the end.

As for the side stringers, the Alfa wing uses evenly spaced side stringers along the leading and trailing edge.
There are 5 on one side and 4 on the other, all equally spaced. The E10E wing only uses a total of 3 side
stringers. 2 on one side and 1 on the other. The rib placement on both wings also differs greatly.

6.1.2. Outboard Part of the Wingbox


For the outboard part of the wingbox, the stringers are again fastened to different plates like for the inboard
part. The E10E wing starts at the sleeve end with a continuation of the 3 middle stringers mentioned in the
previous subsection. All 3 stringers end after a short while. The middle one however, is continued by adding
another 60 [cm] long stringer after it. The Alfa wing uses two stringers in this section. In stead of starting
the stringers at the sleeve end of the wingbox, the stringers start at the tip end of the wingbox. One is slightly
longer than the other.

35
6.2. Alfa Wing Failure based on own Calculations 36

The number of ribs is also different. The E10E Wing uses 2 ribs, while the Alfa Wing uses only one rib. Re-
garding the side stringers, the E10E Wing uses none and the Alfa Wing uses evenly spaced stringers over the
whole length again. Just like in the previous subsection 5 on one side and 4 on the other.

6.2. Alfa Wing Failure based on own Calculations


6.2.1. Geometric Properties
First the geometric properties of the Alfa wing are calculated using the method described in Chapter 3, see
Table 6.1.

Table 6.1: Geometric properties of the alfa wingbox (own data).

Section
Section Property 1 2 3 4 5
y bar 65.47658 65.47658 65.47658 68.34321 75
I 6.58E-06 6.58E-06 6.33E-06 5.78E-06 5.14E-06
Q_centre 4.48E-05 4.48E-05 4.19E-05 3.85E-05 3.65E-05

The properties of section 5 are calculated however it is not loaded under the whiffletree so no failure is ex-
pected in this section therefore it is omitted from further calculation.

6.2.2. Sheet Buckling


Using methods described in Chapter 3, to calculate failure load of the Alfa Wing, a critical weakness is high-
lighted with failure due to sheet buckling. This expected load that will cause failure due to buckling is 70 [N ].
To obtain this value, First the critical stress was calculated using Equation 3.6. The results of this calculation
are compiled in Table 6.2

Table 6.2: Critical buckling stress in alfa wing (own work).

Section
Value 1 2 3 4
Critical Stress [Pa] 1.01E+06 1.01E+06 1.01E+06 1.01E+06

Using the rearrangement of Equation 3.7 for M.

σI
M= (6.1)
y

From Equation 6.1 the internal moments that cause failure are found as compiled in Table 6.3.

Table 6.3: Internal moments (own work).

Section
Value 1 2 3 4
Internal Moment [Nm] 78.45 78.45 75.43 71.32

Using the static equilibrium of the whiffletree as discussed in Chapter 3, The values in Table 6.4 are found

Table 6.4: Failure load in each section (own work).

Section
Value 1 2 3 4
Overall load for failure in each section [N] 70.05 138.31 271.83 1156.50
6.3. Explanation of Strain Gauges selected for Analysis 37

6.2.3. Shear Buckling


Also from the methods in Chapter 3, Failure due to shear buckling is expected at 3691.56[N ]. This is obtained
using Equation 6.2
τ
V= Q
(6.2)
0.025
2t I + 2t Am

First τcr i t i c al is obtained from Equation 3.8, This gives a result of 1.67 · 107 [P a]. Inputting this value into
Equation 6.2 along with the geometric data for section 1 found in Table 6.1. This calculates the critical load to
be 3691.56[N ]

6.2.4. Inter Rivet Buckling


Also the expected failure stress due to the rivet spacings can be calculated using Equation 3.9. This yields the
maximum stresses, which can be found in Table 6.5.

Table 6.5: Maximum stress to avoid inter rivet buckling (own work).

Section
Value 1 2 3 4
Max stress [Pa] 1.01E+08 1.01E+08 1.01E+08 1.01E+08

Using the same method as used to calculate the maximum load for sheet buckling, the maximum load to
resist inter rivet buckling is 7 [kN ]. Which is far above the desired load, therefore not a concern

6.3. Explanation of Strain Gauges selected for Analysis


The strain gauges that were selected for analysis on the Alpha Wing are A & B, 8, 7, 2, and 1 on the inner
wingbox and gauges 8, 7, 2, and 1 on the outer wingbox. Their locations have great significance in the overall
design of our wingbox. A & B are quite self-explanatory, at this point in the wingbox we have a discontinuity
in the design properties, so these locations will be analyzed thoroughly. Then, strain gauges will be analysed
to compare how the top and bottom of the wingbox will deform. For example, locations 8 and 7 are opposite
each other on the horizontal axis of the wingbox. Therefore, if the strain gauges for both locations are given,
one can see if the wingbox will experience pulling apart of the two sides on both the outer and the inner
wingbox. This effect is important to analyze as it could result in the failure of the wingbox. Strain gauges 8
and 7 were specifically chosen on both the inner and outer wingbox due to their great difference in distance
from the clamped edge. The other strain gauges, 2 and 1, have been chosen because of how close they are to
the sleeve connecting the inner and outer wingbox.

6.4. Analysis of the Strain Gauge Data


The gauges placed on the wingbox sleeve show some interesting results in the data in Figure 5.8. Mainly that
the gauge on the bottom of the wingbox seems to be in compression and that the top side is in tension. This
is opposite of what was expected, and the strain is also very low, this is probably because the sleeve is a lot
thicker than the rest of the wingbox. It is also possible that in the data the data for gauge A and B are switched.

The gauges on the inboard side behave somewhat more as expected as shown in Figure 5.9. The top is in
compression, and the bottom is in tension. The only gauge that shows somewhat weird behaviour is gauge 8
for which the strain moves left to right, and shows permanent deflection. This is because the gauge is placed
over the area where the wingbox buckles.

The gauges on the outboard side are also somewhat different as all gauges are in tension even the ones on the
top as shown in Figure 5.10. The ones on the bottom are still more in tension, which is logical. It also seems
that the gauges are not completely placed correctly as they start and end at a strain of 0.0004.
6.5. Description and Explanation of Failure Location 38

6.5. Description and Explanation of Failure Location


First and foremost a limitation of this analysis needs to be established. This analysis of the failure mode is
based on the videos of the test of the Alfa wing provided by the IAC testing facility [17]. The angles in the
videos are quite limiting and thus the accuracy of this analysis is also limited, since certain events could be
not visible because of the video and photo angles.

In the video it can be seen that the Alfa Wing starts to bend upwards. This is of course caused by the load
pulling upwards that is applied by the whiffletree. This upward force causes bending stresses that causes
compression stress on the top side of the wingbox and tension on the bottom side. The compression causes
the first visible deformation and failure. It causes the panel buckling in the top panels. The buckling is greatest
in the first section of the wing, close to the root as this is where the largest moment is applied as can be
seen in figure 3.3. As the formula for the bending stress suggests (Equation 3.7), for large applied moments
the (compressive) bending stress is also higher. This explains why panel buckling failure happens first at the
wingbox root. Close to the root, one can also see that the heavy panel compression causes inter-rivet buckling
between the rivets in the first section.

6.6. Comparison of Test Data and Calculations


The Test Data and Calculations on the Alfa Wing design will be compared to each other with regard to two
main loading modes: shear stresses and normal stresses. As for the shear stresses, it was calculated that the
shear stress-carrying leading edge of the Alfa Wing would fail if the shear stress exceeded 10.2 [M P a] in sec-
tion 3 and that the trailing edge would fail at a shear stress of 1.32 [M P a] measured in section 4. The Alfa
Wing met these shear failure criterion, as shear buckling did not play a major part in the eventual failure of
the wingbox. Therefore, shear stresses are not limiting in this case and the Alfa Wing meets the shear buckling
requirement for the required 3.3[kN ] load from Subsection 3.1.2.

Then, panel buckling is considered. According to the calculations made on the Alfa Wing, the Alfa Wing
would not meet the panel buckling requirement. The calculated limiting normal stress on the top panel is
2.49 [M P a] (compression) in section 4 and 1.32 [M P a] (tension) in section 4 on the bottom panel. The critical
normal stress in section 4 for both top and bottom side is 1.0 [M P a], which indeed suggests the panel would
fail due to panel buckling. These failure modes and expected normal stresses were calculated at a force of 3.3
[kN ] meaning that at this point the theoretical Alfa Wing would already have failed. When a look is taken at
the actual Alfa Wing test data, one can see that the maximum load the Alfa Wing can carry is approximately
3296 [N ] before it fails. This means that the real Alfa Wing is a bit better in carrying normal stresses than the
theoretical wing, meaning it definitely meets the normal stress requirement for a force of 3.3 [kN ].

There is however one requirement the Alfa Wing does not meet: the buckling requirement which states that
buckling should only occur above 3 [kN ] (see Subsection 3.1.2). However from the displacement graph (see
Figure 5.7) it can be seen that the structure already starts permanently deforming from 200 [N ] onward and
the curve only becomes steeper from there on out. In the video it is also just about shown that the base of the
wingbox starts buckling very early around the 55 [sec] in the test it can already be seen buckling a slight bit
even if it is hard to see because of the camera angle, and there is no way of correlating the applied force to the
time in the video. This is most likely explained by the lack of stringers on the top of the wingbox, even though
the load case puts a large compressive force on that area. The combination of the high compressive load and
total lack of reinforcement causes the wingbox to buckle early on.
6.7. Evaluation of Own Wing Design based on the Alfa Wing 39

6.7. Evaluation of Own Wing Design based on the Alfa Wing


Although there are very big differences between the Alfa Wing and the wing designed in this report, the data
from the Alfa Wing can be used to evaluate our own design. Both by taking the results from the Alfa Wing test
and approximating how these would change for the other wingbox design, and by using these results to test
the calculations used to design the other wingbox.

Firstly the calculations seem to be quite useful for determining when the wingbox will initially buckle, this is
only the smallest permanent deformation, but it is present. The calculations do not state anything however
about when ultimate failure will occur and as the Alfa Wing proves the first buckling and ultimate failure can
be quite far removed from one another. This all would suggest that the wingbox designed in this report would
buckle at the desired point, but that not much can be said about its ultimate failure.

Secondly, from the data it can be concluded that the Alfa Wing buckled very early but eventually did fail at the
right time. Comparing this with the wingbox from this report it would seem that that wingbox would buckle
much later due to the large difference in stringers preventing the buckling in the upper plate. This might
also suggest that the wingbox from this report is over designed because of this difference, but that might
ultimately come down to the role of the bottom stringers. The Alfa Wing has much more bottom stringers as
the wingbox from this report and if its strength comes from that than the difference in ultimate failure might
not be so large.
7
Conclusions and Recommendations
The purpose of this report was to describe and analyse the design of a wingbox for use in project GreenWing.
On top of this, the design also had to be optimized for sustainability, and different manufacturing methods
had to be considered.

The design that was eventually settled on was one made of recycled aluminium for better sustainability. The
design is a standard wingbox made up of two parts with a sleeve in between. For reinforcement there are
two side stringers on the leading edge and 1 on the trailing edge, corner stringers in each corner throughout
the entire wingbox, stringers running along the top of the wingbox to prevent buckling, ribs in between the
sections of the wingbox, and many rivets to ensure rigidity. More details can be seen in Figure 7.1. The design
is expected to perform as desired regarding buckling, but regarding ultimate failure, not much can be said.

(a) An exploded view of the inboard side of the wingbox.(own work) (b) An exploded view of the outboard side of the wingbox.(own work)

Figure 7.1: The design for the wingbox (own work).

The design of the wingbox had to meet three requirements which included minimum buckling load, failure
load, and sustainability. The design also had three (dimensional) constraints: overall dimensions of the wing-
box, available materials, and properties of the materials that could be used. After doing many calculations
for different types of loads, the final configuration of the wingbox was determined, which can be seen in Fig-
ure 7.1. This final configuration includes a specific amount of top stringers, corner stringers, side stringers,
ribs and rivets, which will in theory meet the requirements that were mentioned before. However, the wing-
box was not manufactured and hence not tested. Instead, an Alfa wingbox was manufactured, to which the
theoretical wingbox was compared.

Still, a production plan was written to describe the manufacturing procedure for the wingbox. In addition, a
robot motion plan was made that could potentially replace the drilling steps of the manual production plan.
The feasibility of each production method was compared and the robot production option turned out to be

40
41

the superior option due to its fast, efficient and cheap production.

As mentioned before, testing data on the Alfa wing was used to analyse the wingbox design. A comparison
between the two designs suggested that the wingbox will buckle at the desired load. However, due to the am-
biguity of the Alfa wing test data regarding the ultimate failure, the expected ultimate failure performance of
the wingbox turned out to be rather inconclusive.

Recommendations

Using the lessons learned from the testing of the Alfa Wing and the engineering design process of the own
Wingbox, a few recommendations can be made. These recommendations aim at optimizing future design
iterations of the Alfa Wing and/or the own Design. First, recommendations on the own Wingbox design will
be done. Then, the recommendations for the Alfa Wing will be made.

Making recommendations on the own Wingbox design is a bit of a challenge, as no real testing was done on
the design and therefore there is no testing data available. In Chapter 6, an attempt was made to apply the
testing data of the Alfa Wing to the own Wingbox design. Based on this comparison, a few recommendations
can be made. In general, it is highly advised to iterate the design process described in this report in order to
optimize the E10E Wingbox design. In this report, during the calculations, already a few basic iterations were
made in order to meet all requirements and design objectives. Nevertheless, additional iteration of the design
will definitely yield a better wingbox. For example, it can be investigated if the wingbox could be made from
another material that has a better performance when it comes to the design requirements and objectives of
the E10E Wingbox. This could possibly also entail using materials with a lower carbon footprint (a more en-
vironmentally friendly design).

Second, a lot of emphasis was put on meeting the buckling requirements during the design process of the
E10E Wingbox. Therefore, a future recommendation would be to put the focus more on optimizing for the
ultimate failure requirements. It of course is beneficial to focus a bit more on meeting the buckling require-
ment, as buckling often is followed by failure, however in this design process, the ultimate failure requirement
could have had more priority. This would further optimize the E10E Wingbox design (in future iterations).

Finally, with regard to the manufacturing of the E10E Wingbox, it is recommended to use the KUKA manufac-
turing option, because it is fast, efficient and cheap as explained in Chapter 4. The more accurate production
using the KUKA would also mean a better performing wingbox relative to a manual made one, because there
are fewer production errors.

As for the Alfa Wing, a few recommendations can be made too. First of all, the Alfa Wing (as explained in
Chapter 6) buckled quite rapidly after applying load to it. A look could (and should) be taken at the load paths
in the structure of the wingbox. Mainly a lack of stringers and a lack of proper placement of the stringers was
the probable cause of the rapid buckling. Therefore, it is advised to iterate this part of the design.

Second, the IAC Testing Department’s video ([17]) could have been recorded more clearly. The camera angles
do not show all details necessary to perform more accurate analysis on the Alfa Wing. A recommendation, or
more a kind of a guideline for future tests on the Alfa Wing would be to capture and record tests very clearly,
so that all test data and test videos can be easily used as a reference.

Finally, during the Alfa Wing Test, inter-rivet buckling occurred. A closer look should be taken in future itera-
tions at how this inter-rivet buckling can be prevented, by for example decreasing the rivet spacing, or adding
additional stringers to prevent the whole sheet at the root from buckling.
Bibliography
[1] Important Aerospace Company, AE1222-I Wingbox for an Urban Air Mobility Vehicle Reader, 2021. Aca-
demic year 2020/2021.

[2] Alderliesten, R. C., Introduction to Aerospace Engineering - Structures and Materials, TU Delft, 2011. URL
https://textbooks.open.tudelft.nl/textbooks/catalog/view/15/22/80-1.
[3] Ashby, M., Shercliff, H., and Cebon, D., Materials: engineering, science, processing and design,
Butterworth-Heinemann, 2018.

[4] Thipphavong, D., Apaza, R., Barmore, B., Battiste, V., Belcastro, C., Burian, B., Dao, Q., Feary,
M., Go, S., Goodrich, K., Homola, J., Idris, H., Kopardekar, P., Lachter, J., Neogi, N., Ng, H.,
Oseguera-Lohr, R., Patterson, M., and Verma, S., “Urban air mobility airspace integration con-
cepts and considerations,” 2018. doi:10.2514/6.2018-3676, URL https://www.scopus.com/inward/
record.uri?eid=2-s2.0-85051668637&doi=10.2514%2f6.2018-3676&partnerID=40&md5=
b970fbe7388dd9a7d64c3e6a8e781c2e, cited By 68.
[5] Vascik, P., and Hansman, R., “Scaling constraints for urban air mobility operations: Air traf-
fic control, ground infrastructure, and noise,” 2018. doi:10.2514/6.2018-3849, URL https:
//www.scopus.com/inward/record.uri?eid=2-s2.0-85051643710&doi=10.2514%2f6.
2018-3849&partnerID=40&md5=042e93575e33a2104203f2c2fa3865d5, cited By 18.
[6] Wilhite, L., “What Is The Factor Of Safety?” Onsite Safety, 2018. URL https://www.onsitesafety.
com/safety-articles/what-is-the-factor-of-safety/.

[7] Modlin, C. T., and Zipay, J. J., “The 1.5 & 1.4 Ultimate Factors of Safety for Aircraft & Spacecraft - History,
Definition and Applications,” NASA, February 2014. URL https://ntrs.nasa.gov/api/citations/
20140011147/downloads/20140011147.pdf, accessed 26-04-2021.
[8] Findlay, S. J., and Harrison, N. D., “Why aircraft fail,” Materials Today, Vol. 5, 2002, pp. 18–25. URL
https://www.sciencedirect.com/science/article/pii/S1369702102011380.

[9] Showers, D. R., “Fatigue, Fail-Safe and Damage Tolerance Evaluation of Metallic Structure for Normal,
Utility, Acrobatic, and Commuter Category Airplanes,” FAA, September 2005. URL https://www.faa.
gov/documentLibrary/media/Advisory_Circular/ac23-13A.pdf, accessed 25-04-2021.
[10] Anderson, J., Introduction to Flight, 8th ed., Vol. 12, McGraw-Hill Education, 2016. URL https://books.
google.nl/books?id=uMV5cgAACAAJ.
[11] Hibbeler, R. C., Mechanics of Materials, 10th ed., Pearson education, 2018.

[12] Owen-Hill, A., “Back to Basics: Robot Motion Planning Made Easy,” RoboDK, 2019. URL https:
//robodk.com/blog/robot-motion-planning-made-easy/.

[13] Roberts, E., “Motion Planning in Robotics,” Stanford University, 1999. URL https://cs.stanford.
edu/people/eroberts/courses/soco/projects/1998-99/robotics/basicmotion.html, ac-
cessed 25-04-2021.

[14] Important Aerospace Company, AE1222-I Design and Construction Launcher Compression Panel Design,
2021. Academic year 2020/2021.

[15] Company, I. A., “Alfa Wing Alfa pre test (8) picture,” , 2021.

[16] Company, I. A., “Alfa Wing Alfa after test (5) picture,” , 2021.

[17] Company, I. A., “Alfa Wing Test Video,” , 2021.

[18] Department, I. T., “Alpha Inboard - Assembly Drawing,” , 2021.

[19] Department, I. T., “Alpha Outboard - Assembly Drawing,” , 2021.

42
1 2 3 4 5 6 7 8 9 10 11 12 13 14

Making task distribution

Literature study

Manufacturing S7

Introduction manufacturing plan

Tools and materials list

Manufacturing steps
A.1. Estimated Time Distribution

43
Design

Calculations failure loads

Description design goals and assumptions

Explanation iteration

Design choice and justification technical

Design choice and justification sustainable

Catia drawings
Task Distribution
A
1 2 3 4 5 6 7 8 9 10 11 12 13 14

Handing in production plan

Manufacturing S14

Description of the production plan


A.1. Estimated Time Distribution

Discussion of manufacturability

Description of robot motion plan

Feasibility assessment between manual and robot manufacturing

Testing

Description of test objective

Description of test system

Calculation of load distribution over wiffletree setup

Description of test events


44
1 2 3 4 5 6 7 8 9 10 11 12 13 14

Analysis

Discussion of differences between designs

Calculation of alfa wing failure mode


A.1. Estimated Time Distribution

Explanation and justification of strain gauges

Description and explanation of failure area

Comparison of alfa wing test data with failure load calculations

Cover/lay-out

Reference list

Appendix

Summary

Table of contents

Conclusion and recommendations

Handing in report
45
1 2 3 4 5 6 7 8 9 10 11 12 13 14

Making task distribution

Literature study

Manufacturing S7
A.2. Actual time distribution.

Introduction manufacturing plan

Tools and materials list


A.2. Actual time distribution.

Manufacturing steps

Design

Calculations failure loads

Description design goals and assumptions

Explanation iteration

Design choice and justification technical

Design choice and justification sustainable

Catia drawings
46
1 2 3 4 5 6 7 8 9 10 11 12 13 14

Handing in production plan and technical drawings

Manufacturing S14
A.2. Actual time distribution.

Description of the production plan

Discussion of manufacturability

Description of robot motion plan

Feasibility assessment between manual and robot manufacturing

Testing

Description of test objective

Description of test system

Calculation of load distribution over wiffletree setup

Description of test events


47
1 2 3 4 5 6 7 8 9 10 11 12 13 14

Analysis

Discussion of differences between designs


A.2. Actual time distribution.

Calculation of alfa wing failure mode

Explanation and justification of strain gauges

Description and explanation of failure area

Comparison of alfa wing test data with failure load calculations

Cover/lay-out

Reference list

Appendix

Summary

Table of contents

Conclusion and recommendations

Handing in report and robot motion plan


48
A.3. Time spend on each part per person 49

A.3. Time spend on each part per person


Table A.1: The task division and time spent on each task.

Time
Activity People spent
(min)
Making task distribution Robin, Aleksei 420
Literature study Didier, Lucas, Dragomir 450
Introduction manufacturing plan Didier 60
Tools and materials list Theophile, Aleksei 90
Manufacturing steps Robin, Didier, Theophile, Aleksei 840
Calculations failure loads Mike, Bart, Ewan 750
Description design goals and assumptions Mike, Bart, Ewan 180
Explanation iteration Mike, Ewan 250
Design choice and justification technical Mike, Dragomir, Ewan 570
Design choice and justification sustainable Bart 120
Catia drawings Mike, Dragomir, Ewan, Bart 900
Catia 3D model Dragomir 1500
Handing in production plan and technical drawings Dragomir, Bart 15
Description of the production plan Mike 15
Discussion of manufacturability Theophile, Aleksei 330
Description of robot motion plan Mike, Didier, Dragomir 480
Feasibility assessment between manual and robot manu-
Theophile, Aleksei 960
facturing
Description of test objective Lucas, Morris 50
Description of test system Lucas, Morris 40
Calculation of load distribution over whiffletree setup Robin, Lucas, Morris 120
Describing test events Robin, Lucas 240
Discussion of differences between designs Lucas 60
Calculation of Alfa wing failure mode Bart, Ewan 120
Explanation and justification of strain gauges Robin, Morris 120
Description and explanation of failure area Robin, Lucas 60
Comparison of alfa wing test data with failure load calcula-
Robin 90
tions
Cover/lay-out Dragomir, Robin 60
Reference list everyone -
Appendix Mike 30
Summary Lucas, Aleksei 240
Table of contents Robin 30
Conclusion and recommendations Robin, Bart, Didier, Aleksei 120
Handing in report and robot motion plan Bart, Didier 15
Material Properties, Availability and
B
Dimensions
B.1. Material Properties
Table B.1: Material properties of supplied wing box components [1]

Item Material ρ[g /cm 3 ] E [M P a] σul t [M P a] σ y [M P a]


Sheets (A-C) and ribs New Aluminium 2.78 75000 130 110
Sheets (A-C) and ribs Recycled Aluminium 2.78 70000 104 88
Stringers (A-D) New Aluminium 2.78 75000 300 200
Stringers (A-D) Recycled Aluminium 2.78 70000 240 160

Table B.2: Average emissions and price of aluminium production and aviation gasoline use [1]

Sector Material Carbon Dioxide Emissions Price (as of 2021-03-22)


Manufacturing New Aluminium 3.55 kg/kg Al €42.72/tonne Carbon Dioxide
Manufacturing Recycled Aluminium 0.27 kg/kg Al €42.72/tonne Carbon Dioxide
Operations Aviation Gasoline 2.20 kg/L fuel €42.72/tonne Carbon Dioxide

B.2. Material Dimensions and Availability


Table B.3: Material availability list [1]

Item Quantity L [mm] W [mm] t [mm] Notes


Sheet A 2 1300 148.4 0.8 Wingbox- short side
Sheet B 2 1500 148.4 0.8 Wingbox- long side
Sheet C 4 1300/1500 400.0 0.8 Wingbox- top, bottom Incl. 30◦ angle
Rib 4 398.4 148.4 0.8 Flanges included
L-stringer A 6 1000 20x20 1.5 Can be cut, if required
L-stringer B 10 1250 20x20 1.5 Can be cut, if required
L-stringer C 6 1500 20x20 1.5 Can be cut, if required
L-stringer D 10 152 20x20 1.5 Side stiffeners
Rivets 1500 - - - Total number of rivets

50
B.3. Rivet Properties and Dimensions 51

B.3. Rivet Properties and Dimensions


Table B.4: Specifications Fabory omni Dome head blind rivet 3.2x6mm [1]

Parameter Value
Diameter [mm] 3.2
Length [mm] 6
Grip Range [mm] 1.5-3.0
Head Shape Mushroom head
Material body/Mandrel Steel/Steel
Shear load [N ] 1,060
Tensile load [N ] 1,285
C
Manual Production Timeline

52
53

Figure C.1: Production timeline for a fully manual production process.(own work)
54

Figure C.2: Production timeline for a partly automated production process.(own work)
D
Alfa Wing and E10E Wing Assembly
Drawing

55
56

A
16

16
15

15

Isometric view
Scale: 1:10
14

14

Section view C-C


Scale: 1:4
13

13

Section view B-B


12

Scale: 1:4 12

C C
11

11

Rear view
Scale: 1:4
10

10

B B
9
9

Top view
Scale: 1:4
8
8

A A

7
7

Left view Front view Right view


Scale: 1:4 Scale: 1:4 Scale: 1:4

6
6

5
5

Bottom view
Scale: 1:4

4
4

3
3

2
2

Section view A-A


Scale: 1:4

This drawing is our property.


It can't be reproduced
or communicated without
DASSAULT SYSTEMES
our written agreement.
DRAWING TITLE
DRAWN BY DATE
Alpha Inboard - Assembly Drawing
XXX 5/24/2020
CHECKED BY DATE SIZE DRAWING NUMBER REV
1
1

XXX XXX A0 X
DESIGNED BY DATE
XXX SCALE 1:4 WEIGHT(kg) 1.95 SHEET 1/1
XXX

P O N M L K F E D C B A

Figure D.1: Inboard Alpha Wing Assembly Drawing [18]


57

A
16

16
15

15

Isometric view
Scale: 1:10
14

14

Section view C-C


Scale: 1:4
13

13
12

Section view B-B 12


Scale: 1:4
11

11
C C

Rear view
Scale: 1:4
10

10

9
9

B B

8
8

Top view
Scale: 1:4

A A 7
7

Left view Front view Right view


Scale: 1:4 Scale: 1:4 Scale: 1:4

6
6

5
5

Bottom view
4
4

Scale: 1:4

3
3

2
2

Section view A-A


Scale: 1:4
This drawing is our property.
It can't be reproduced
or communicated without
DASSAULT SYSTEMES
our written agreement.
DRAWING TITLE
DRAWN BY DATE
Alpha Outboard - Assembly Drawing
XXX 5/24/2020
CHECKED BY DATE SIZE DRAWING NUMBER REV
1
1

XXX XXX A0 X
DESIGNED BY DATE
XXX SCALE 1:4 WEIGHT(kg) 1.76 SHEET 1/1
XXX

P O N M L K F E D C B A

Figure D.2: Outboard Alpha Wing Assembly Drawing [19]


P
O
N
M
L
K
J
I
H
G
F
E
D
C
B
A

8
8

4
10

Isometric view[2]
Scale: 1:10

7
7
Isometric view
5 Scale: 1:10

3
13

12

6
6
11

6
Isometric view
Scale: 1:10
16
Bill of Material: AE1222-I-WB-Subassembly-A

5
5
Number Quantity Part Number Type Nomenclature Revision
1 1 AE1222-I-WB- Assembly Inner Top Panel A4
Subassembly-A.1
15 1 AE1222-I-WB- Assembly Inner Bottom Panel A4
Subassembly-A.2
14 1 1 AE1222-I-WB-Plate-11 Part Sheet B Inner TE A4
Side Plate
2 1 AE1222-I-WB-Plate-12 Part Sheet A Inner LE A5
Side Plate
3 1 AE1222-I-WB-ST-13 Part TE Side Panel L A4
Stringer

4
4
4 2 AE1222-I-WB-ST-14 Part LE Side Panel L A4
290
stringer
140 5 1 AE1222-I-WB-RIB-15 Part Inner Rib 1 A4
6 1 AE1222-I-WB-RIB-16 Part Inner Rib 2 A4

Bill of Material: AE1222-I-WB-Subassembly-A.1


Number Quantity Part Number Type Nomenclature Revision
7 1 AE1222-I-WB-Plate-01 Part Sheet C Inner Top A4

130.2
1.8
1
Plate
8 1 AE1222-I-WB-ST-02 Part Inner TE Top Corner A4
Left view Front view Right view Rear view Stringer
Scale: 1:10 Scale: 1:10 Scale: 1:10 Scale: 1:10

3
9 1 AE1222-I-WB-ST-03 Part Inner LE Top Corner A4 3
Stringer

Figure D.3: Inboard E10E Wing Assembly Drawing (own work)


10 5 AE1222-I-WB-ST-04 Part L Stringer A4
Recapitulation of: AE1222-I- 11 1 AE1222-I-WB-ST-05 Part L Stringer A4
WB-Subassembly-A 12 1 AE1222-I-WB-ST-06 Part L Stringer A4
Different parts: 16 13 1 AE1222-I-WB-ST-07 Part L Stringer A4
Total parts: 21
Quantity Part Number
Bill of Material: AE1222-I-WB-Subassembly-A.2
1 AE1222-I-WB-Plate-01
Number Quantity Part Number Type Nomenclature Revision
1 AE1222-I-WB-ST-02
14 1 AE1222-I-WB-Plate-08 Part Sheet C Inner A7
1 AE1222-I-WB-ST-03 Bottom Plate

2
5 AE1222-I-WB-ST-04 15 1 AE1222-I-WB-ST-09 Part Inner TE Bottom A5 2
1 AE1222-I-WB-ST-05 Corner Stringer
Bottom view 1 AE1222-I-WB-ST-06 16 1 AE1222-I-WB-ST-10 Part Inner LE Bottom A4
Scale: 1:10 Corner Stringer
1 AE1222-I-WB-ST-07
215
1 AE1222-I-WB-Plate-08
1 AE1222-I-WB-ST-09 This drawing is our property.
It can't be reproduced
1 AE1222-I-WB-ST-10 or communicated without
DASSAULT SYSTEMES
1 AE1222-I-WB-Plate-11 our written agreement.
DRAWING TITLE
1 AE1222-I-WB-Plate-12 DRAWN BY DATE
1 AE1222-I-WB-ST-13 dnikolov 12-May-21 Inner wingbox Assembly
2 AE1222-I-WB-ST-14 CHECKED BY DATE SIZE DRAWING NUMBER REV

1
XXX 1
1 AE1222-I-WB-RIB-15 XXX A1 AE1222-I-WB-Subassembly-A A5
DESIGNED BY DATE
1 AE1222-I-WB-RIB-16 SCALE 1:10 WEIGHT(kg) 2.03 SHEET 1/1
Group E10E XXX
58

P O N M L K F E D C B A
P
O
N
M
L
K
J
I
H
G
F
E
D
C
B
A

8
8

9
8

10

7
7
2
1

14

6
3 6

11

12 Isometric view
Scale: 1:10
13 Exploded view
Scale: 1:10
6

5 7

5
5

Bill of Material: AE1222-I-WB-Subassembly-B


Number Quantity Part Number Type Nomenclature Revision
1 AE1222-I-WB- Assembly Outer Bottom Panel A4
Subassembly-B.1
1 AE1222-I-WB- Assembly Outer Top Panel A4
Subassembly-B.2
1 1 AE1222-I-WB-Plate-26 Part Sheet A Outer TE A4
Side Plate
2 1 AE1222-I-WB-Plate-27 Part Sheet B Outer LE A4

4
4
Side Plate
Top view 3 1 AE1222-I-WB-RIB-28 Part Outer Rib 1 A4
Scale: 1:10 4 1 AE1222-I-WB-RIB-29 Part Outer Rib 2 A5

Bill of Material: AE1222-I-WB-Subassembly-B.1


Number Quantity Part Number Type Nomenclature Revision
5 1 AE1222-I-WB-Plate-17 Part Sheet C Outer A9
Bottom Plate
6 1 AE1222-I-WB-ST-32 Part Outer TE Bottom A4
Corner L Stringer

3
7 1 AE1222-I-WB-ST-18 Part Outer LE Bottom A4 3
Left view Front view Right view Rear view Corner L Stringer

Figure D.4: Outboard E10E Wing Assembly Drawing (own work)


Scale: 1:10 Scale: 1:10 Scale: 1:10 Scale: 1:10
Bill of Material: AE1222-I-WB-Subassembly-B.2
Number Quantity Part Number Type Nomenclature Revision
Recapitulation of: AE1222-I-
WB-Subassembly-B 8 1 AE1222-I-WB-Plate-19 Part Sheet C Outer Top A4
Plate
Different parts: 14
9 1 AE1222-I-WB-ST-20 Part Outer TE Top Corner A4
Total parts: 14
L Stringer
Quantity Part Number
10 1 AE1222-I-WB-ST-21 Part Outer LE Top Corner A4
1 AE1222-I-WB-Plate-17 L Stringer
1 AE1222-I-WB-ST-32 11 1 AE1222-I-WB-ST-22 Part L Stringer A4

2
2
1 AE1222-I-WB-ST-18 12 1 AE1222-I-WB-ST-23 Part L Stringer A4
1 AE1222-I-WB-Plate-19 13 1 AE1222-I-WB-ST-24 Part L Stringer A4
1 AE1222-I-WB-ST-20 14 1 AE1222-I-WB-ST-25 Part L Stringer A4
1 AE1222-I-WB-ST-21
1 AE1222-I-WB-ST-22
Bottom view 1 AE1222-I-WB-ST-23 This drawing is our property.
Scale: 1:10 It can't be reproduced DASSAULT SYSTEMES
1 AE1222-I-WB-ST-24 or communicated without
our written agreement.
1 AE1222-I-WB-ST-25 DRAWING TITLE
1 AE1222-I-WB-Plate-26 DRAWN BY DATE

1 AE1222-I-WB-Plate-27 dnikolov 12-May-21 Outer Wingbox Assembly


CHECKED BY DATE SIZE DRAWING NUMBER REV

1
1 AE1222-I-WB-RIB-28 XXX 1
XXX A1 AE1222-I-WB-Subassembly-B A5
1 AE1222-I-WB-RIB-29 DESIGNED BY DATE
Group E10E SCALE 1:10 WEIGHT(kg) 1.73 SHEET 1/1
XXX
59

P O N M L K F E D C B A
E
Wingbox Production Plan

60
"This is the Production Plan in Appendix E"

Contents
1 Introduction 1
2 Tools and Materials 2
2.1 Safety . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2.2 Measuring & Marking . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2.3 Punching, Drilling & Finishing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2.4 Clamping & Riveting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2.5 Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
3 Manufacturing 4
3.1 Sizing and Marking Stringers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
3.1.1 Top Stringers (A) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
3.1.2 Top Stringers (B) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
3.1.3 Top Corner Stringer #1 (C) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3.1.4 Top Corner Stringer #2 (C) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3.1.5 Top Corner Stringer #3 (B) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3.1.6 Top Corner Stringer #4 (B) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3.1.7 Bottom Corner Stringers #5/6 (C) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3.1.8 Bottom Corner Stringer #7/8 (B) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3.1.9 Side Stringers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3.2 Marking Panels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.2.1 Top Panels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.2.2 Bottom Panels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
3.2.3 Side Panels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
3.3 Marking and Cutting Ribs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
3.4 Attaching Stringers to Panels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.4.1 Top Panels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.4.2 Side Panels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.5 Assembly . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.5.1 Attaching Ribs to Bottom Panel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.5.2 Attaching Side Panels to Assembly . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.5.3 Attaching Top Panel to Assembly. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.5.4 Connection of both parts of the wingbox. . . . . . . . . . . . . . . . . . . . . . . . . . 20
Bibliography 22
A Drawings 23

i
"This is the Production Plan in Appendix E"

1
Introduction
For DELTA’s new GreenWing project, a wing-box will be manufactured for their new Urban Air Mobility Ve-
hicle (UAM). The wing-box is the skeleton of the wing, making sure it can carry its own weight, all the aero-
dynamic loads when flying as well as the applied loads from the engine thrust. A wing-box is essentially a
hollow cantilever beam, reinforced by stringers on the inside, to reduce weight and increase the strength if
the wing-box.
When working on an engineering project like the Green Wing project, the product has to be manufactured
using a production plan. Without a production plan, it is impossible for an outsider to produce the right parts
and assemble everything properly without compromising the final result. This is why a detailed, concise, and
clear production plan is crucial to the success of this project.

The aim of this report is to describe the production plan of a wingbox for project GreenWing. The tools and
materials which will be necessary and a detailed step by step manufacturing plan will be described to give
the manufacturers producing the wingbox all the knowledge needed to do this. The production plan should
contain all the content to correctly manufacture the wingbox with the intent that it will perfectly match the
designed wingbox.

This report will be structured as follows. Chapter 2 will provide the lists of the tools and materials that will be
used during the manufacturing process of the wingbox. The Manufacturing steps are described in chapter 3.
And in appendix A, all the relevant technical drawings will be present.

1
"This is the Production Plan in Appendix E"

2
Tools and Materials
2.1. Safety
• Safety Goggles; must be worn at all times

• COVID Regulations; must be followed at all times

• Hearing Protection; must be worn when using the band saw

2.2. Measuring & Marking


• Pencil

• Permanent Marker

• Try Square

• Steel Ruler

• Tape Measure

2.3. Punching, Drilling & Finishing


• Centre Punch

• Hammer

• Drill Bit ø = 2.5[mm] (Pre-drilling pilot hole)

• Drill Bit ø = 3.2[mm] (final drilling finish)

• Battery Powered Drill

• Hole saw bit ø = 44[mm]

• Hand Reamer (For plate deburring)

• Handheld Counter Sink (For hole deburring)

2.4. Clamping & Riveting


• Clamp screw used to temporarily clamp a plate to a stringer/stiffener

• Rivet clamp ø = 2.5[mm], used to temporarily fasten a hole after pre-drilling

• Rivet clamp ø = 3.2[mm], used to temporarily fasten a hole after final drilling

• Hand Riveter

2
"This is the Production Plan in Appendix E"
2.5. Materials 3

2.5. Materials
To manufacture this wing-box, quite a few parts will be used:

• Short rivets 3.2 [mm] diameter, 1.5 - 3.0 [mm] grip length

• M6 bolts

• 2x 1300 [mm] long 148.8[mm] high aluminium side panels (sheet A)

• 2x 1500 [mm] long 148.8[mm] high aluminium side panels (sheet B)

• 2x 1300/1500[mm] long 400 [mm] wide with a 30°incline aluminium top panels and 2 aluminium bot-
tom panels of the same dimensions (sheet C)

• 5 x 1000 [mm] long 20x20 [mm] wide aluminium L-stingers (stringer A)

• 6 x 1250 [mm] long 20x20 [mm] wide aluminium L-stingers (stringer B)

• 4 x 1500 [mm] long 20x20 [mm] wide aluminium L-stingers (stringer C)

• 3x 152 [mm] long 20x20 [mm] wide aluminium L-stringers (stringer D)

• 4x 398.4 [mm] long 148.4 [mm] wide aluminium ribs

• 1 connecting metal sleeve


"This is the Production Plan in Appendix E"

3
Manufacturing
Disclaimer: Abide to all the safety and covid rules, from start to finish during the manufacturing process.

3.1. Sizing and Marking Stringers


This section discusses where to put the markings for the rivets on the stringers and how to cut the stringers
to length.

3.1.1. Top Stringers (A)


(Time Estimate: 10 min.)

1. Grab a stringer of type A, and lay it down on the workbench;

2. Next, indicate the midpoints of one of the flanges of the stringer at the ends of the stringer; connect the
midpoints with a dashed line;

3. Starting 45 [mm] from the left side of the stringer, mark the first rivet;

4. After the first rivet marking, draw 9 more rivets with 45 [mm] in between each of them;

5. Draw a line 62 [mm] after the last rivet marking, and mark another rivet place;

6. Mark 8 more rivets along the dashed line, keeping a spacing of 62 [mm] now;

7. Repeat this whole process for each of the 5 type A stringers, make sure to choose the same flange for
each stringer, so that in the final product they will be oriented in the same way.

8. Indent every marking with the center punch and the hammer.

3.1.2. Top Stringers (B)


(Time Estimate: 15 min.)

1. Grab a stringer of type B, and lay it down on the workbench;

2. Next, indicate the midpoints of one of the flanges of the stringer at the ends of the stringer; connect the
midpoints with a dashed line;

3. Now using the band saw and following the dimensions in table 3.1, cut one long stringer into 4 new
stringers;

4. Pair up the 207 [mm] stringer and the 371 [mm] stringer and refer to them as ’pair 1’;

5. Pair up the 307 [mm] stringer and the 271 [mm] stringer and refer to them as ’pair 2’;

6. For pair 1, put them down end to end with the 207 [mm] stringer on the left. Starting from the left end
of the 207 [mm] stringer, mark as many rivet placings as possible along the dashed line, while keeping
a 90 [mm] spacing in between them (6 rivets).

7. Following the same idea mark rivet placings for pair 2. With the 307 [mm] stringer on the left and start
from the 307 [mm] stringer’s left side.

4
"This is the Production Plan in Appendix E"
3.1. Sizing and Marking Stringers 5

8. Now, grab another B stringer and cut it into the two lengths as displayed in table 3.1;

9. Place them end to end with the 339 [mm] stringer on the left and start from the left end of the 339 [mm]
stringer, mark as many rivet placings as you can fit, keeping a 90 [mm] spacing between each rivet.

10. Indent every marking with the center punch and the hammer.

Table 3.1: Lengths to which B stringers need to be cut

ORIGINAL STRINGER (quantity) CUT PARTS QUANTITY


B (1250 [mm]) (1) 371 [mm] 1
271 [mm] 1
207 [mm] 1
307 [mm] 1

B (1250 [mm]) (1) 339 [mm] 1


839 [mm] 1

3.1.3. Top Corner Stringer #1 (C)


(Time Estimate: 5 min.)

1. Grab a type C stringer and lay it down on a workbench;

2. Next, indicate the midpoints of one of the flanges of the stringer at the ends of the stringer; connect the
midpoints with dashed lines;

3. 45 [mm] away from the left most edge of the stringer along both dashed lines, mark the location of the
first rivet;

4. Mark another 9 rivet places 45[mm] apart from each other;

5. 62 [mm] away from the last marked rivet place, mark a new rivet place;

6. Mark another 8 rivet places 62[mm] apart from each other, still following the dashed line;

7. 90 [mm] away from the last marked rivet place, mark a new rivet place;

8. Mark as many other rivet places 90[mm] apart from each other as you can fit, still following the dashed
line;

9. Mark a line 11 [mm] away from the right side of the stringer, and using the band saw (making sure to
follow safety procedures) cut the stringer along the newly drawn line;

10. Indent every marking with the center punch and the hammer.

3.1.4. Top Corner Stringer #2 (C)


(Time Estimate: 5 min.)

1. Grab a type C stringer and lay it down on a workbench;

2. Next, indicate the midpoints of one of the flanges of the stringer at the ends of the stringer; connect the
midpoints with dashed lines;

3. Mark a line 11 [mm] from the left edge of the stringer and using the band saw, cut off this 11 [mm] long
piece of stringer;

4. Starting from the newly cut edge, mark as many other rivet places 100 [mm] apart from each other as
you can fit on the dashed lines;

5. Indent every marking with the center punch and the hammer.
"This is the Production Plan in Appendix E"
3.1. Sizing and Marking Stringers 6

3.1.5. Top Corner Stringer #3 (B)


(Time Estimate: 5 min.)

1. Grab a type B stringer and lay it down on a workbench;

2. Next, indicate the midpoints of one of the flanges of the stringer at the ends of the stringer; connect the
midpoints with dashed lines;

3. 45 [mm] away from the left most edge of the stringer along the dashed line, mark the location of the
first rivet;

4. Mark another 9 rivet places 45[mm] apart from each other;

5. 62 [mm] away from the last marked rivet place, mark a new rivet place;

6. Mark another 8 rivet places 62[mm] apart from each other, still following the dashed line;

7. 90 [mm] away from the last marked rivet place, mark a new rivet place;

8. Mark as many other rivet places 90[mm] apart from each other as you can fit, still following the dashed
line;

9. Indent every marking with the center punch and the hammer.

3.1.6. Top Corner Stringer #4 (B)


(Time Estimate: 5 min.)

1. Grab a type B stringer and lay it down on a workbench;

2. Next, indicate the midpoints of one of the flanges of the stringer at the ends of the stringer; connect the
midpoints with dashed lines;

3. Starting from the left most edge of the stringer, mark as many other rivet places 100 [mm] apart from
each other as you can fit, still following the dashed line;

4. Indent every marking with the center punch and the hammer.

3.1.7. Bottom Corner Stringers #5/6 (C)


(Time Estimate: 5 min.)

1. Grab 2 type C stringers and lay them down on a workbench;

2. Next, indicate the midpoints on each flange of both stringers at the ends of the stringer; connect the
midpoints with dashed lines;

3. Mark a line 11 [mm] from the left edge of the stringer and using the band saw, cut off this 11 [mm] long
piece of stringer;

4. Starting from the newly cut edge, mark as many other rivet places 45 [mm] apart from each other as
you can fit, still following the dashed line;

5. Indent every marking with the center punch and the hammer.
"This is the Production Plan in Appendix E"
3.1. Sizing and Marking Stringers 7

3.1.8. Bottom Corner Stringer #7/8 (B)


(Time Estimate: 5 min.)

1. Grab 2 type B stringers and lay them down on a workbench;

2. Next, indicate the midpoints of each flange of the stringer at the ends of the stringer; connect the mid-
points with dashed lines;

3. Starting from the left most edge of the stringer, mark as many other rivet places 45 [mm] apart from
each other as you can fit, still following the dashed line;

4. Indent every marking with the center punch and the hammer.

3.1.9. Side Stringers


(Time Estimate: 2-3 min.)

1. Grab a type D stringer;

2. Next, indicate the midpoints of one of the flanges of a stringer at the ends of the stringer; connect the
midpoints with a dashed line;

3. Mark 2 points each 10 [mm] away from either edge of the stringer, along the dashed line;

4. In between these two points, mark another point

5. Using the center punch and the hammer, indent the drawn markings;

6. Repeat the same process for the other 2 type D stringers.


"This is the Production Plan in Appendix E"
3.2. Marking Panels 8

3.2. Marking Panels


This section discusses where to put the markings for rivets or stringers on the panels.
For all panels also mark holes as shown in Figure 3.1.

Figure 3.1: Holes for attaching sleeve. (own work)

3.2.1. Top Panels


(Time Estimate: 20 min.)

For these steps, use Figure 3.9 as a reference. In this figure the red lines represent the centre lines of the
stringers and the panels are oriented with the insides pointing upwards.

1. Grab 2 panels of type C to start with the top panels;

2. Put one panel flat on a workbench like the upper panel shown in Figure 3.9;

3. Draw 2 vertical dashed lines 450 [mm] and 1050 [mm] from the non inclined short edge of the panel;

4. Draw a dashed line 66 [mm] from and parallel to the upper long edge of the panel starting at the non
inclined short edge until the vertical 1050 [mm] line;

5. After that, draw 4 more dashed lines parallel to the upper long edge of the panel with 67 [mm] between
the lines;

6. For the 3 middle lines, continue the dashed line until the end of the panel is reached;

7. Now flip the panel, keeping it landscape with the short edge on the left;

8. Draw 2 dashed lines 10.8 [mm] from and parallel to either of the long edges of the panel;

9. Starting 45 [mm] from the left short edge of the panel along both dashed lines, mark the location of the
first rivet;

10. Mark another 9 rivet places 45 [mm] apart from each other on each dashed line;

11. 62 [mm] away from the last marked rivet place, mark a new rivet place on each dashed line;

12. Mark another 8 rivet places 62 [mm] apart from each other, still following each dashed line;

13. 90 [mm] away from the last marked rivet place, mark a new rivet place on each dashed line;
"This is the Production Plan in Appendix E"
3.2. Marking Panels 9

14. Mark as many other rivet places 90 [mm] apart from each other as you can fit, still following each dashed
line;

15. Draw 2 vertical dashed lines 450 [mm] and 1050 [mm] from the non inclined short edge of the panel;

16. At these 2 vertical dashed lines, mark 5 rivet places starting from the bottom long edge of the panel. For
the rivet spacing, use the E, D, C, B and A X-locations consecutively from Figure 3.3 to mark each rivet
place, starting 38 [mm] from the bottom long edge of the panel;

17. Now take the other inclined C panel and put it flat on a workbench like the lower panel shown in Figure
3.9;

18. Draw 2 vertical dashed lines 550 [mm] and 1150 [mm] from the non inclined short edge of the panel;

19. Draw a dashed line 133 [mm] from and parallel to the upper long edge of the panel starting from the
inclined short edge until the 1150 [mm] line;

20. After that, draw 2 more dashed lines parallel to the upper long edge of the panel with 67 [mm] between
the lines;

21. For the middle line, continue the dashed line until the 550 [mm] line is reached.

22. Now flip the panel, keeping it landscape with the short edge on the right;

23. Draw 2 dashed lines 10.8 [mm] from an parallel to either of the long edges of the panel;

24. Starting from the right short edge of the panel, mark as many other rivet places 100 [mm] apart from
each other as you can fit on each dashed line;

25. Draw 2 vertical dashed lines 550 [mm] and 1150 [mm] from the non inclined short edge of the panel;

26. At the vertical dashed line at 550 [mm], mark 5 rivet places starting from the bottom long edge of the
panel. For the rivet spacing, use the E, D, C, B and A X-locations consecutively from Figure 3.4 to mark
each rivet place, starting 39 [mm] from the bottom long edge of the panel;

27. At the vertical dashed line at 1150 [mm], mark 5 rivet places starting from the bottom long edge of the
panel. For the rivet spacing, use the D, C, B, A and E X-locations consecutively from Figure 3.5 to mark
each rivet place, starting 39 [mm] from the bottom long edge of the panel;

28. Keep both panels landscape and on the workbench showing the side containing the most amounts of
rivet markings;

29. Horizontally measuring 67.3 [mm] away from the slanted edge draw a dashed line, parallel to the
slanted edge, on each of the panels;

30. Mark the location of the stringers with horizontal dashed lines (making sure they intersect the slanted
dashed lines) indicating the center line of one of the stringers’ flanges;

31. Highlight the intersection point and using the center punch, indent the intersection points of the lines.

3.2.2. Bottom Panels


(Time Estimate: 10 min.)

1. Grab 2 non-modified type C panels and put them landscape on a workbench;

2. Align the inclined short edges of the panels so it will become one long rectangular panel. Make sure
that the diagonal incline goes from bottom left to top right;

3. Draw 4 vertical dashed lines 450 [mm], 1050 [mm], 1650 [mm] and 2250 [mm] from the left short edge
of the rectangular panel;

4. Flip the panels but keep them landscape;


"This is the Production Plan in Appendix E"
3.3. Marking and Cutting Ribs 10

5. Horizontally measuring 67.3 [mm] away from the slanted edge draw a dashed line, parallel to the
slanted edge, on each of the panels;

6. Mark the location of the corner stringers with horizontal dashed lines (making sure they intersects the
slanted dashed lines) indicating the center line of the stringers’ flanges that will be attached to the
bottom panel;

7. Highlight the intersection point and using the center punch, indent the intersection points of the lines.

3.2.3. Side Panels


(Time Estimate: 10-15 min.)

1. Grab 2 panels of type A and 2 panels of type B;

2. Put one type B panel flat and landscape on a workbench;

3. Draw one vertical dashed line 315 [mm] from one of the short edges.

4. Along this vertical dashed line, mark 3 dots: 2 dots each 9 [mm] away from either long edge of the panel
and one in the middle of these points;

5. Using the center punch and the hammer, indent the marked dots. After that, put the panel aside;

6. Put one type A panel flat and landscape on a workbench;

7. Draw 2 vertical dashed lines 225 [mm] and 405 [mm] from one of the short edges;

8. Along this vertical dashed line, mark 3 dots: 2 dots each 9 [mm] away from either long edge of the panel
and one in the middle of these points;

9. Using the center punch and the hammer, indent the marked dots;

10. Flip both previously marked panels, indented side down, keeping them landscape. Also put the other
type A and type B panels flat and landscape on the workbench;

11. On each panel, draw 2 dashed lines both 10.8 [mm] from and parallel to either of the long edges of the
panels.

3.3. Marking and Cutting Ribs


(Time Estimate: 20-25 min.)

This section discusses where to mark the rivet-holes on the ribs and how to cut holes for the stringers in the
ribs. There are 3 types of ribs, for which the rivet holes are all the same except for the holes on the top flanges,
which are spaced differently . Also, the amount of cut-outs for the stringers to go through is different. First the
marking and drilling of the holes in the bottom and side flanges of all ribs will be explained. Next the drilling
of the top flanges will be described for each type of rib, since these are not the same. After that the cut-out
dimensioning and cutting will be explained for the relevant ribs.
Side and bottom holes
"This is the Production Plan in Appendix E"
3.3. Marking and Cutting Ribs 11

Figure 3.2: side and bottom hole dimension on all ribs

1. Perform the following steps, which describe the marking of the holes on the top and side flanges on all
ribs and use figure 3.2 as a reference.

Table 3.2: Type of ribs and the amount needed to be manufactured

Type of Rib Number of cut-outs amount to be manufactured


Type A 5 2
Type B 3 1
Type C 1 1

2. On all the flanges of the rib, mark the center line of the flange (approx. 7.2 [mm] from the outer edge)
on both outer ends of the flange and use a ruler to draw the center line with a pencil;

3. On the bottom flange, indicate the first hole location at 44.2 [mm] away from the left edge. Mark four
more co-linear dots each spaced 65 [mm] away from each other;

4. On both side flanges, indicate the first hole 24.6 [mm] away from the outermost edge and mark two
more co-linear dots each also spaced 24.6 [mm] away from each other;

5. When marking is done, use a center punch at all the marked holes. Drill with a 3.2 [mm] bit through
all the marked holes. Use a sacrificial plate underneath all holes while drilling to prevent damaging the
workbench;

Top flange holes Type A

1. On the top flange of the Type A Rib, indicate the locations of the rivet holes, visible in figure3.3;

Figure 3.3: Hole dimensions of the Type A rib

2. Make an indent with the center punch at all the marked points and drill with a 3.2 [mm] bit through all
the marked holes. Use a sacrificial plate when drilling to prevent damage to the workbench.

Top flange holes Type B


"This is the Production Plan in Appendix E"
3.3. Marking and Cutting Ribs 12

1. On the top flange of the Type B Rib, indicate the locations of the rivet holes, visible in figure3.4;

Figure 3.4: Hole dimensions of the Type B rib

2. Make an indent with the center punch at all the marked points and drill with a 3.2 [mm] bit through all
the marked holes. Use a sacrificial plate when drilling to prevent damage to the workbench.

Top flange holes Type C

1. On the top flange of the Type C Rib, indicate the locations of the rivet holes, visible in figure3.5;

Figure 3.5: Hole dimensions of the Type C rib

2. Make an indent with the center punch at all the marked points and drill with a 3.2 [mm] bit through all
the marked holes. Use a sacrificial plate when drilling to prevent damage to the workbench.

Table 3.3: Type of ribs and the amount needed to be manufactured

Type of Rib Number of cut-outs amount to be manufactured


Type A 5 2
Type B 3 1
Type C 1 1

Cut-outs for type A, B and C

For the cutouts on the type A (5 cut-outs), type B (3 cut-outs), type C (1-cutouts )s ribs, the following steps
need to be preformed. Use the figures 3.6, 3.7 and 3.8 as references in these steps for making the cut-outs on
all the ribs indicated in table 3.2.

1. Mark the cut-out lines with the appropriate dimensions as indicated in figures 3.6, 3.7 and 3.8.

2. Use the band-saw to cut out the straight marked ’cut-out’ lines on the top flange.

3. Use a 44 [mm] Hole saw to saw off the semi-circles on each cut outs.
"This is the Production Plan in Appendix E"
3.3. Marking and Cutting Ribs 13

4. Use a hand reamer to refine the cut-out edges.

Figure 3.6: Cut-out dimensions of a Type A rib

Figure 3.7: Cut-out dimensions of a Type B rib


"This is the Production Plan in Appendix E"
3.4. Attaching Stringers to Panels 14

Figure 3.8: Cut-out dimensions of a Type C rib

3.4. Attaching Stringers to Panels


This section discusses how each type of stringer is attached to the appropriate panels. For all the steps in-
volving drilling, make sure to use a sacrificial plate.

3.4.1. Top Panels


(Time Estimate: 30-35 min.)

For this subsection use Figure 3.9 as a reference as to where to put the stringers. The red lines in the figure
represent the centre lines of the stringers and the figure should be read as if the inside of the panel is pointing
up. In Table 3.4 information can be found on which stringer should be used where. Use the following steps
to rivet everything together. Note that all stringers need to be attached to the inside of the top panels and
should have the exact same orientation. Figure 3.10 can be used as a reference for the first panel of this step
the second panel should follow a similar idea just with a different layout.

1. Grab the top panel marked like the upper panel in Figure 3.9 and grab the 5 type A stringers;

2. Align the dashed centre line of one type A stringer with one of the dashed lines on the top panel that
represent the red a-lines from Figure 3.9;

3. Clamp the stringer to the panel on both ends using 2 clamp screws;

4. First drill pilot holes through the stringer and the top panel using the battery powered drill with the 2.5
[mm] drill bit in the same locations as the indents. Rivet clamps can be used to secure drilled holes;

5. Next, use the battery powered drill with the 3.2 [mm] drill bit to widen the 2.5 [mm] holes through the
stringer and panel. Again, rivet clamps can be used to secure the drilled holes;

6. Use a counter sink to debur all the drilled holes by putting the sharp end into a hole on the burred side
and turning it clockwise around its own axis until the burr is removed;

7. Use the rivet gun to rivet the stringer and the top panel together from the inside out;

8. Repeat steps 2 - 7 for the other 4 type A stringers;

9. Grab the 307 [mm], 339 [mm] and 371 [mm] type B stringers and align their centre lines with the dashed
lines on the top panel that represent the red b-, c- and d-lines, respectively, from Figure 3.9;

10. For these 3 stringers, repeat steps 2 - 7;

11. Now grab the other top panel marked like the lower panel in Figure 3.9:

12. Grab the 271 [mm], 207 [mm] and 839 [mm] type B stringers and align their centre lines with the dashed
lines on the top panel that represent the red e-, f- and g-lines, respectively, from Figure 3.9.

13. For these 3 stringers, repeat steps 2 - 7;


"This is the Production Plan in Appendix E"
3.4. Attaching Stringers to Panels 15

Table 3.4: Type of stringer to length, see Figure 3.9

Letter Length [mm]


a 1000
b 307
c 339
d 371
e 271
f 207
g 839

Figure 3.9: Layout of the stringers (red) on the top panels.


"This is the Production Plan in Appendix E"
3.4. Attaching Stringers to Panels 16

Figure 3.10: One of the finished top panels (own work)


"This is the Production Plan in Appendix E"
3.4. Attaching Stringers to Panels 17

3.4.2. Side Panels


(Time Estimate: 20-25 min.)

Use the following steps to rivet the corner stringers and the side stringers to the side panels. Note that the
side stringers should have the exact same orientation. The orientation of the corner stringers is specified in
Figure 3.12.

Figure 3.11: All 4 marked side panels with named dashed lines. Note that in this figure, ’marked’ means that the panel has at least one
marked vertical dashed line on the back of the panel, displayed as the dashed grey lines.

1. Grab the 4 side panels, and place them landscape on a workbench, showing the dashed lines parallel to
the long edges of the panels;

2. Grab the 8 corner stringers and align each of the dashed centre lines of the stringers with the dashed
lines on the panels. In Table 3.5 it is specified which corner stringer should align with which dashed
line on which panel. Make sure to align the corner stringers as specified in Figure 3.12;

3. Clamp one of the stringers to the side panel on both ends with 2 clamp screws;

4. First drill pilot holes through the stringer and the side panel using the battery powered drill with the 2.5
[mm] drill bit in the same locations as the indents. Rivet clamps can be used to secure drilled holes;

5. Next, use the battery powered drill with the 3.2 [mm] drill bit to widen the 2.5 mm holes through the
stringer and panel. Again, rivet clamps can be used to secure the drilled holes;

6. Use a counter sink to debur all the drilled holes by putting the sharp end into a hole on the burred side
and turning it clockwise around its own axis until the burr is removed;

7. Use the rivet gun to rivet the stringer and the side panel together from the inside out through the drilled
holes;
"This is the Production Plan in Appendix E"
3.5. Assembly 18

Table 3.5: Corner stringer number with the line it has to align with on the side panel. The corner stringer numbers refer to the numbers
from Section 3.1, and the names of the lines the stringers need to align with and their corresponding panels can be found in Figure 3.11.

CORNER STRINGER LINE TO ALIGN WITH SIDE PANEL


#1 e Marked Type B
#2 g Unmarked Type B
#3 a Marked Type A
#4 c Unmarked Type A
#5 f Marked Type B
#6 h Unmarked Type B
#7 b Marked Type A
#8 d Unmarked Type A

8. Repeat steps 3 - 7 for all the other corner stringers;

9. Now flip the marked type A panel and the marked type B panel, corner stringer side down, keeping
them landscape;

10. Align 2 type D stringers with the vertical dashed lines on the marked type A panel and one type D
stringer with the vertical dashed line on the marked type B panel;

11. Clamp one of the stringers to the side panel on both ends using 2 clamp screws. The stringer should
stick out 1 [mm] on each long edge of the panel;

12. First drill pilot holes through the stringer and the side panel using the battery powered drill with the 2.5
[mm] drill bit in the same locations as the indents. Rivet clamps can be used to secure drilled holes;

13. Next, use the battery powered drill with the 3.2 [mm] drill bit to widen the 2.5 [mm] holes through the
stringers and panel. Again, rivet clamps can be used to secure the drilled holes;

14. Use a counter sink to debur all the drilled holes by putting the sharp end into a hole on the burred side
and turning it clockwise around its own axis until the burr is removed;

15. Use the rivet gun to rivet the stringer and the side panel together from the outside in;

16. Repeat steps 11 - 15 for the other 2 type D stringers.

3.5. Assembly
This section discusses how to attach all the ribs, bottom/top panels, and the side panels. After the first three
sections the two halves of the wingbox should look like the drawings in the appendix A figure A.8 and figure
A.5

3.5.1. Attaching Ribs to Bottom Panel


(Time Estimate: 15 min.)

For all the steps involving drilling make sure to use a sacrificial plate.

1. Put the bottom panel with the dashed lines at 450 [mm] and 1050 [mm] on a table with the 1300 [mm]
side on your right and the 1500 [mm] side on your left;

2. Place the type A ribs parallel to the short non inclined side of the panel at the dashed lines 450 [mm]
and 1050 [mm] from this side in a way that they are symmetrical along the panel’s centre line. Make
sure the cut outs for the stringers are facing upwards;

3. Use a hammer and centre punch to make small indents in the places where the rivets will connect the
ribs to the panel;

4. First drill pilot holes through the ribs and the bottom panel using a 2.5 [mm] drill bit in the same loca-
tions as the indents. Rivet clamps can be used to secure drilled holes;
"This is the Production Plan in Appendix E"
3.5. Assembly 19

5. Use a 3.2 [mm] drill bit to widen the 2.5 [mm] holes through the ribs and panel. Again rivet clamps can
be used to secure the drilled holes;

6. Use a counter sink to debur the drilled holes;

7. Use the rivet gun to rivet the ribs and bottom panel together from the inside out;

8. Repeat this on the other bottom panel but with a type C rib at 550 [mm] and a type B rib at 1150 [mm]
away from the short non inclined side. Again make sure the cut outs for the stringers are facing upwards.

3.5.2. Attaching Side Panels to Assembly


(Time Estimate: 20 min.)

For all the steps involving drilling make sure to use a sacrificial plate.

1. Place the type B panel with 1 side stringer on the long side of the first bottom panel in such a way as
depicted in picture 3.12 [1];

2. Now use a hammer and a centre punch to make small indents in each of the locations marked for the
rivets along the stringer that now lies in the corner of the bottom panel and the side panel;

3. First drill pilot holes through the stringer and bottom panel using a 2.5 [mm] drill bit in the same loca-
tions as the indents. Rivet clamps can be used to secure drilled holes;

4. Use a 3.2 [mm] drill bit to widen the 2.5 [mm] holes through the stringer and panel. Again rivet clamps
can be used to secure the drilled holes;

5. Use a counter sink to debur the drilled holes;

6. Use the rivet gun to rivet the stringer and panel together from the inside out;

7. Do the same for the type A panel with 2 side stringers, on the short side of the bottom panel;

8. Rivet the side panels and the ribs together using the previously made markings on the side panels,
together with the steps used to rivet the corner stringer and bottom panel together;

9. Repeat this on the other bottom panel with the remaining side panels. For this bottom panel, the type
B panel without side stringers should be on the long side of the panel and the type A panel without
stringers should be on the short side of the panel.
"This is the Production Plan in Appendix E"
3.5. Assembly 20

Figure 3.12: How the top-/ bottom and side plates should be aligned. [1]

3.5.3. Attaching Top Panel to Assembly


(Time Estimate: 30-35 min.)

For all the steps involving drilling make sure to use a sacrificial plate.

1. Place the top panel With the most stringers on top of the side panels with side stringers aligning them
in the same way as the bottom- and side panel, and aligning the long and short sides of the top and
bottom panels making sure the stringers are facing downwards;

2. Now use a hammer and a centre punch to make small indents in each of the locations marked on the
top panel for the rivets of the corner stringer;

3. First drill pilot holes through the top panel and stringer using a 2.5 [mm] drill bit in the same locations
as the indents. Rivet clamps can be used to secure drilled holes;

4. Use a 3.2 [mm] drill bit to widen the 2.5 [mm] holes through the panel and stringer. Again rivet clamps
can be used to secure the drilled holes;

5. Use a counter sink to debur the drilled holes;

6. Use the rivet gun to rivet the top panel and corner stringer together from the outside in;

7. Do this process on either side of the top panel;

8. Rivet the top panel and the ribs together using the markings on the outside of the top panel and the
same steps for riveting as described in the previous steps;

9. Repeat this with the other bottom panel and top panel.

3.5.4. Connection of both parts of the wingbox


(Time Estimate: 10 min.)
"This is the Production Plan in Appendix E"
3.5. Assembly 21

1. Gather both wingbox parts, as well as the sleeve;

2. Drill though the indented holes on the non-stringer sides of the top and bottom panels;

3. Debur, the previously drilled holes; and place each wing box part in both sides of the sleeve, keeping 70
[mm] in between each part of the wing box;

4. Now, grab a few bolts and bolt the sleeve and the wing box parts together to finalize the design.
"This is the Production Plan in Appendix E"

Bibliography
[1] TUDelft, “Wingbox for an Urban Air Mobility Vehicle Reader,” , 2021. URL https://brightspace.
tudelft.nl/d2l/le/content/292963/viewContent/2119267/View.

22
"This is the Production Plan in Appendix E"

A
Drawings

23
P
O
N
M
L
K
J
I
H
G
F
E
D
C
B
A

16
16

15
15

14
14

Bill of Material: AE1222-I-WB-Assembly


Number Quantity Part Number Type Nomenclature Revision
1 AE1222-I-WB- Assembly Inner wingbox A5
Subassembly-A
1 AE1222-I-WB- Assembly Outer wingbox A5

13
13
Subassembly-B
Isometric view Isometric view 1 AE1222-I-WB- Assembly Connecting Sleeve A4
Scale: 1:10 Scale: 1:10 Subassembly-C

Bill of Material: AE1222-I-WB-Subassembly-A


Number Quantity Part Number Type Nomenclature Revision
1 AE1222-I-WB- Assembly Inner Top Panel A4
Subassembly-A.1

12
12
1 AE1222-I-WB- Assembly Inner Bottom Panel A4
Subassembly-A.2
1 1 AE1222-I-WB-Plate-11 Part Sheet B Inner TE A4

400
Side Plate
2 1 AE1222-I-WB-Plate-12 Part Sheet A Inner LE A5
Side Plate
3 1 AE1222-I-WB-ST-13 Part TE Side Panel L A4
Top view Stringer

11
11
Scale: 1:10 4 2 AE1222-I-WB-ST-14 Part LE Side Panel L A4
1161.75 stringer
2870 5 1 AE1222-I-WB-RIB-15 Part Inner Rib 1 A4
6 1 AE1222-I-WB-RIB-16 Part Inner Rib 2 A4

Bill of Material: AE1222-I-WB-Subassembly-A.1


Number Quantity Part Number Type Nomenclature Revision
7 1 AE1222-I-WB-Plate-01 Part Sheet C Inner Top A4

10
10
Plate
8 1 AE1222-I-WB-ST-02 Part Inner TE Top Corner A4
Stringer

150
9 1 AE1222-I-WB-ST-03 Part Inner LE Top Corner A5
Front view Stringer
Left view Right view Rear view
Scale: 1:10 10 5 AE1222-I-WB-ST-04 Part L Stringer A4
Scale: 1:10 Scale: 1:10 Scale: 1:10
11 1 AE1222-I-WB-ST-05 Part L Stringer A4
12 1 AE1222-I-WB-ST-06 Part L Stringer A4

9
9
13 1 AE1222-I-WB-ST-07 Part L Stringer A4

Bill of Material: AE1222-I-WB-Subassembly-A.2


Number Quantity Part Number Type Nomenclature Revision
14 1 AE1222-I-WB-Plate-08 Part Sheet C Inner A7
Bottom Plate
15 1 AE1222-I-WB-ST-09 Part Inner TE Bottom A5

8
Corner Stringer 8
Bottom view
Scale: 1:10 16 1 AE1222-I-WB-ST-10 Part Inner LE Bottom A4
Corner Stringer

25
17 Bill of Material: AE1222-I-WB-Subassembly-B
Number Quantity Part Number Type Nomenclature Revision
22 1 AE1222-I-WB- Assembly Outer Bottom Panel A4
Subassembly-B.1

7
1 AE1222-I-WB- Assembly Outer Top Panel A4 7
20 Subassembly-B.2

Figure A.1: Subassembly (own work)


17 1 AE1222-I-WB-Plate-26 Part Sheet A Outer TE A4
27 Side Plate
30
18 1 AE1222-I-WB-Plate-27 Part Sheet B Outer LE A4
Recapitulation of: AE1222-I- Side Plate
28 WB-Assembly 19 1 AE1222-I-WB-RIB-28 Part Outer Rib 1 A4
Different parts: 33 20 1 AE1222-I-WB-RIB-29 Part Outer Rib 2 A4

6
24 29 6
Total parts: 39
21 19 Quantity Part Number
Bill of Material: AE1222-I-WB-Subassembly-B.1
33 1 AE1222-I-WB-Plate-01
Number Quantity Part Number Type Nomenclature Revision
1 AE1222-I-WB-ST-02
21 1 AE1222-I-WB-Plate-17 Part Sheet C Outer A9
1 AE1222-I-WB-ST-03 Bottom Plate
5 AE1222-I-WB-ST-04 22 1 AE1222-I-WB-ST-32 Part Outer TE Bottom A4
1 AE1222-I-WB-ST-05 Corner L Stringer

5
1 AE1222-I-WB-ST-06 23 1 AE1222-I-WB-ST-18 Part Outer LE Bottom A4 5
1 AE1222-I-WB-ST-07 Corner L Stringer
18 1 AE1222-I-WB-Plate-08
15 1 AE1222-I-WB-ST-09 Bill of Material: AE1222-I-WB-Subassembly-B.2
23 3 1 AE1222-I-WB-ST-10 Number Quantity Part Number Type Nomenclature Revision
"This is the Production Plan in Appendix E"

9 1 AE1222-I-WB-Plate-11 24 1 AE1222-I-WB-Plate-19 Part Sheet C Outer Top A4


1 AE1222-I-WB-Plate-12 Plate
1 1 AE1222-I-WB-ST-13 25 1 AE1222-I-WB-ST-20 Part Outer TE Top Corner A4

4
L Stringer 4
2 AE1222-I-WB-ST-14
8 26 1 AE1222-I-WB-ST-21 Part Outer LE Top Corner A4
1 AE1222-I-WB-RIB-15
26 L Stringer
1 AE1222-I-WB-RIB-16
7 27 1 AE1222-I-WB-ST-22 Part L Stringer A4
1 AE1222-I-WB-Plate-17
28 1 AE1222-I-WB-ST-23 Part L Stringer A4
1 AE1222-I-WB-ST-32
29 1 AE1222-I-WB-ST-24 Part L Stringer A4
1 AE1222-I-WB-ST-18
30 1 AE1222-I-WB-ST-25 Part L Stringer A4
1 AE1222-I-WB-Plate-19
32

3
1 AE1222-I-WB-ST-20 3
6 Bill of Material: AE1222-I-WB-Subassembly-C
1 AE1222-I-WB-ST-21
1 AE1222-I-WB-ST-22 Number Quantity Part Number Type Nomenclature Revision
31 11 1 AE1222-I-WB-ST-23 31 2 AE1222-I-WB-Wall-30 Part Sleeve Side Wall A4
12 10
1 AE1222-I-WB-ST-24 32 1 AE1222-I-WB-SPL-31 Part Sleeve Top/Bottom A4
13 Plate
1 AE1222-I-WB-ST-25
33 1 Symmetry of AE1222- Part Sleeve Top/Bottom A4
1 AE1222-I-WB-Plate-26
I-WB-SPL-31 Plate
1 AE1222-I-WB-Plate-27

2
2
1 AE1222-I-WB-RIB-28 This drawing is our property.
It can't be reproduced DASSAULT SYSTEMES
1 AE1222-I-WB-RIB-29 or communicated without
2 AE1222-I-WB-Wall-30 our written agreement.
2 DRAWING TITLE
1 AE1222-I-WB-SPL-31 DRAWN BY DATE
14 16 Exploded view
5 1 Symmetry of AE1222- dnikolov 12-May-21 Wingbox assembly
4 Scale: 1:10 I-WB-SPL-31 CHECKED BY DATE SIZE DRAWING NUMBER REV
XXX XXX A0 AE1222-I-WB-Assembly A5
DESIGNED BY DATE
24

1
SCALE 1:10 WEIGHT(kg) 7.82 SHEET 1/1 1
Group E10E XXX

P O N M L K F E D C B A
P
O
N
M
L
K
J
I
H
G
F
E
D
C
B
A

8
8

4
10

Isometric view[2]
Scale: 1:10

7
7
Isometric view
5 Scale: 1:10

3
13

12

6
6
11

6
Isometric view
Scale: 1:10
16
Bill of Material: AE1222-I-WB-Subassembly-A

5
5
Number Quantity Part Number Type Nomenclature Revision
1 1 AE1222-I-WB- Assembly Inner Top Panel A4
Subassembly-A.1
15 1 AE1222-I-WB- Assembly Inner Bottom Panel A4
Subassembly-A.2
14 1 1 AE1222-I-WB-Plate-11 Part Sheet B Inner TE A4
Side Plate
2 1 AE1222-I-WB-Plate-12 Part Sheet A Inner LE A4
Side Plate
3 1 AE1222-I-WB-ST-13 Part TE Side Panel L A4
Stringer

4
4
4 2 AE1222-I-WB-ST-14 Part LE Side Panel L A4
290
stringer
140 5 1 AE1222-I-WB-RIB-15 Part Inner Rib 1 A4
6 1 AE1222-I-WB-RIB-16 Part Inner Rib 2 A4

Figure A.2: Subassembly A (own work)


Bill of Material: AE1222-I-WB-Subassembly-A.1
Number Quantity Part Number Type Nomenclature Revision
7 1 AE1222-I-WB-Plate-01 Part Sheet C Inner Top A4

130.2
1.8
1
Plate
8 1 AE1222-I-WB-ST-02 Part Inner TE Top Corner A4
Left view Front view Right view Rear view Stringer
Scale: 1:10 Scale: 1:10 Scale: 1:10 Scale: 1:10

3
9 1 AE1222-I-WB-ST-03 Part Inner LE Top Corner A4 3
Stringer
10 5 AE1222-I-WB-ST-04 Part L Stringer A4
Recapitulation of: AE1222-I- 11 1 AE1222-I-WB-ST-05 Part L Stringer A4
WB-Subassembly-A 12 1 AE1222-I-WB-ST-06 Part L Stringer A4
Different parts: 16 13 1 AE1222-I-WB-ST-07 Part L Stringer A4
Total parts: 21
"This is the Production Plan in Appendix E"

Quantity Part Number


Bill of Material: AE1222-I-WB-Subassembly-A.2
1 AE1222-I-WB-Plate-01
Number Quantity Part Number Type Nomenclature Revision
1 AE1222-I-WB-ST-02
14 1 AE1222-I-WB-Plate-08 Part Sheet C Inner A4
1 AE1222-I-WB-ST-03 Bottom Plate

2
5 AE1222-I-WB-ST-04 15 1 AE1222-I-WB-ST-09 Part Inner TE Bottom A4 2
1 AE1222-I-WB-ST-05 Corner Stringer
Bottom view 1 AE1222-I-WB-ST-06 16 1 AE1222-I-WB-ST-10 Part Inner LE Bottom A4
Scale: 1:10 Corner Stringer
1 AE1222-I-WB-ST-07
215
1 AE1222-I-WB-Plate-08
1 AE1222-I-WB-ST-09 This drawing is our property.
It can't be reproduced
1 AE1222-I-WB-ST-10 or communicated without
DASSAULT SYSTEMES
1 AE1222-I-WB-Plate-11 our written agreement.
DRAWING TITLE
1 AE1222-I-WB-Plate-12 DRAWN BY DATE
1 AE1222-I-WB-ST-13 dnikolov 12-May-21 Inner wingbox Assembly
2 AE1222-I-WB-ST-14 CHECKED BY DATE SIZE DRAWING NUMBER REV

1
XXX 1
1 AE1222-I-WB-RIB-15 XXX A1 AE1222-I-WB-Subassembly-A A4
DESIGNED BY DATE
1 AE1222-I-WB-RIB-16 SCALE 1:10 WEIGHT(kg) 2.03 SHEET 1/1
Group E10E XXX
25

P O N M L K F E D C B A
P
O
N
M
L
K
J
I
H
G
F
E
D
C
B
A

8
8
4

7
7

6
6
Isometric view
5 Scale: 1:5
Exploded view
Scale: 1:5

6 7

5
5

0.8
56

4
( 67 )
4

379.2

(4x67=) 268
Right view

Figure A.3: Subassembly A.1 (own work)


Front view Scale: 1:5
Scale: 1:5

Bill of Material: AE1222-I-WB-Subassembly-A.1


Number Quantity Part Number Type Nomenclature Revision

3
3
1 1 AE1222-I-WB-Plate-01 Part Sheet C Inner Top A4
Plate
2 1 AE1222-I-WB-ST-02 Part Inner TE Top Corner A4
Stringer
3 1 AE1222-I-WB-ST-03 Part Inner LE Top Corner A4
Stringer
"This is the Production Plan in Appendix E"

4 5 AE1222-I-WB-ST-04 Part L Stringer A4


Bottom view
Scale: 1:5 5 1 AE1222-I-WB-ST-05 Part L Stringer A4
6 1 AE1222-I-WB-ST-06 Part L Stringer A4
7 1 AE1222-I-WB-ST-07 Part L Stringer A4

2
2
Recapitulation of: Multi
selection
Different parts: 7
Total parts: 22
Quantity Part Number
This drawing is our property.
2 AE1222-I-WB-Plate-01 It can't be reproduced
or communicated without
DASSAULT SYSTEMES
2 AE1222-I-WB-ST-02
our written agreement.
2 AE1222-I-WB-ST-03 DRAWING TITLE
DRAWN BY DATE
10 AE1222-I-WB-ST-04
dnikolov 12-May-21 Inner Top Panel Assembly
2 AE1222-I-WB-ST-05 CHECKED BY DATE SIZE DRAWING NUMBER REV

1
2 AE1222-I-WB-ST-06 XXX 1
XXX A1 AE1222-I-WB-Subassembly-A.1 A4
2 AE1222-I-WB-ST-07 DESIGNED BY DATE
Group E10E SCALE 1:5 WEIGHT(kg) 0.96 SHEET 1/1
XXX
26

P O N M L K F E D C B A
P
O
N
M
L
K
J
I
H
G
F
E
D
C
B
A

8
8

7
7

6
6
3

Isometric view
Exploded view
Scale: 1:5
Scale: 1:5

5
5

4
4

0.8

Figure A.4: Subassembly A.2 (own work)


379.2

3
3

Front view Right view


Scale: 1:5 Scale: 1:5
"This is the Production Plan in Appendix E"

Bill of Material: AE1222-I-WB-Subassembly-A.2


Number Quantity Part Number Type Nomenclature Revision
1 1 AE1222-I-WB-Plate-08 Part Sheet C Inner A7
Bottom Plate

2
2 1 AE1222-I-WB-ST-09 Part Inner TE Bottom A5 2
Corner Stringer
3 1 AE1222-I-WB-ST-10 Part Inner LE Bottom A4
Corner Stringer

Recapitulation of: AE1222-I- This drawing is our property.


Bottom view WB-Subassembly-A.2 It can't be reproduced DASSAULT SYSTEMES
Scale: 1:5 or communicated without
Different parts: 3 our written agreement.
Total parts: 3 DRAWING TITLE
DRAWN BY DATE
Quantity Part Number dnikolov Inner Bottom Panel Assembly
12-May-21
1 AE1222-I-WB-Plate-08 CHECKED BY DATE SIZE DRAWING NUMBER REV

1
1 AE1222-I-WB-ST-09 XXX 1
XXX A1 AE1222-I-WB-Subassembly-A.2 A4
1 AE1222-I-WB-ST-10 DESIGNED BY DATE
Group E10E SCALE 1:5 WEIGHT(kg) 0.61 SHEET 1/1
XXX
27

P O N M L K F E D C B A
P
O
N
M
L
K
J
I
H
G
F
E
D
C
B
A

8
8

9
8

10

7
7
2
1

14

6
3 6

11

12 Isometric view
Scale: 1:10
13 Exploded view
Scale: 1:10
6

5 7

5
5

Bill of Material: AE1222-I-WB-Subassembly-B


Number Quantity Part Number Type Nomenclature Revision
1 AE1222-I-WB- Assembly Outer Bottom Panel A4
Subassembly-B.1
1 AE1222-I-WB- Assembly Outer Top Panel A4
Subassembly-B.2
1 1 AE1222-I-WB-Plate-26 Part Sheet A Outer TE A4
Side Plate
2 1 AE1222-I-WB-Plate-27 Part Sheet B Outer LE A4

4
4
Side Plate
Top view 3 1 AE1222-I-WB-RIB-28 Part Outer Rib 1 A4
Scale: 1:10 4 1 AE1222-I-WB-RIB-29 Part Outer Rib 2 A4

Figure A.5: Subassembly B (own work)


Bill of Material: AE1222-I-WB-Subassembly-B.1
Number Quantity Part Number Type Nomenclature Revision
5 1 AE1222-I-WB-Plate-17 Part Sheet C Outer A4
Bottom Plate
6 1 AE1222-I-WB-ST-32 Part Outer TE Bottom A4
Corner L Stringer

3
7 1 AE1222-I-WB-ST-18 Part Outer LE Bottom A4 3
Left view Front view Right view Rear view Corner L Stringer
Scale: 1:10 Scale: 1:10 Scale: 1:10 Scale: 1:10
Bill of Material: AE1222-I-WB-Subassembly-B.2
Number Quantity Part Number Type Nomenclature Revision
Recapitulation of: AE1222-I-
WB-Subassembly-B 8 1 AE1222-I-WB-Plate-19 Part Sheet C Outer Top A4
Plate
"This is the Production Plan in Appendix E"

Different parts: 14
9 1 AE1222-I-WB-ST-20 Part Outer TE Top Corner A4
Total parts: 14
L Stringer
Quantity Part Number
10 1 AE1222-I-WB-ST-21 Part Outer LE Top Corner A4
1 AE1222-I-WB-Plate-17 L Stringer
1 AE1222-I-WB-ST-32 11 1 AE1222-I-WB-ST-22 Part L Stringer A4

2
2
1 AE1222-I-WB-ST-18 12 1 AE1222-I-WB-ST-23 Part L Stringer A4
1 AE1222-I-WB-Plate-19 13 1 AE1222-I-WB-ST-24 Part L Stringer A4
1 AE1222-I-WB-ST-20 14 1 AE1222-I-WB-ST-25 Part L Stringer A4
1 AE1222-I-WB-ST-21
1 AE1222-I-WB-ST-22
Bottom view 1 AE1222-I-WB-ST-23 This drawing is our property.
Scale: 1:10 It can't be reproduced DASSAULT SYSTEMES
1 AE1222-I-WB-ST-24 or communicated without
our written agreement.
1 AE1222-I-WB-ST-25 DRAWING TITLE
1 AE1222-I-WB-Plate-26 DRAWN BY DATE

1 AE1222-I-WB-Plate-27 dnikolov 12-May-21 Outer Wingbox Assembly


CHECKED BY DATE SIZE DRAWING NUMBER REV

1
1 AE1222-I-WB-RIB-28 XXX 1
XXX A1 AE1222-I-WB-Subassembly-B A4
1 AE1222-I-WB-RIB-29 DESIGNED BY DATE
Group E10E SCALE 1:10 WEIGHT(kg) 1.73 SHEET 1/1
XXX
28

P O N M L K F E D C B A
P
O
N
M
L
K
J
I
H
G
F
E
D
C
B
A

8
8

7
7

2
1

6
6

Isometric view
Scale: 1:5

5
5

Exploded view
Scale: 1:5

4
4

Figure A.6: Subassembly B.1 (own work)


379.2

3
3

0.8
Front view Right view
"This is the Production Plan in Appendix E"

Scale: 1:5 Scale: 1:5


Bill of Material: AE1222-I-WB-Subassembly-B.1
Number Quantity Part Number Type Nomenclature Revision
1 1 AE1222-I-WB-Plate-17 Part Sheet C Outer A9
Bottom Plate

2
2 1 AE1222-I-WB-ST-32 Part Outer TE Bottom A4 2
Corner L Stringer
3 1 AE1222-I-WB-ST-18 Part Outer LE Bottom A4
Corner L Stringer
Bottom view
Scale: 1:5 Recapitulation of: AE1222-I- This drawing is our property.
WB-Subassembly-B.1 It can't be reproduced DASSAULT SYSTEMES
or communicated without
Different parts: 3 our written agreement.
Total parts: 3 DRAWING TITLE
DRAWN BY DATE
Quantity Part Number dnikolov Outer Bottom Panel Assembly
12-May-21
1 AE1222-I-WB-Plate-17 CHECKED BY DATE SIZE DRAWING NUMBER REV

1
1 AE1222-I-WB-ST-32 XXX 1
XXX A1 AE1222-I-WB-Subassembly-B.1 A4
1 AE1222-I-WB-ST-18 DESIGNED BY DATE
Group E10E SCALE 1:5 WEIGHT(kg) 0.60 SHEET 1/1
XXX
29

P O N M L K F E D C B A
P
O
N
M
L
K
J
I
H
G
F
E
D
C
B
A

8
8

7
7

6
6

7
Exploded view
Scale: 1:5 Isometric view
Scale: 1:5
6

5
5
4

4
4

Figure A.7: Subassembly B.2 (own work)


Bill of Material: AE1222-I-WB-Subassembly-B.2
Number Quantity Part Number Type Nomenclature Revision

3
3
1 1 AE1222-I-WB-Plate-19 Part Sheet C Outer Top A4
Plate

379.2
2 1 AE1222-I-WB-ST-20 Part Outer TE Top Corner A4

257
L Stringer

190
3 1 AE1222-I-WB-ST-21 Part Outer LE Top Corner A4

123
L Stringer

0.8
"This is the Production Plan in Appendix E"

4 1 AE1222-I-WB-ST-22 Part L Stringer A4


5 1 AE1222-I-WB-ST-23 Part L Stringer A4
Front view Right view
Scale: 1:5 Scale: 1:5 6 1 AE1222-I-WB-ST-24 Part L Stringer A4
7 1 AE1222-I-WB-ST-25 Part L Stringer A4

2
2
Recapitulation of: AE1222-I-
WB-Subassembly-B.2
Different parts: 7
Total parts: 7
Quantity Part Number
This drawing is our property.
1 AE1222-I-WB-Plate-19 It can't be reproduced
Bottom view or communicated without
DASSAULT SYSTEMES
1 AE1222-I-WB-ST-20
Scale: 1:5 our written agreement.
1 AE1222-I-WB-ST-21 DRAWING TITLE
DRAWN BY DATE
1 AE1222-I-WB-ST-22
dnikolov 12-May-21 Outer Top Panel Assembly
1 AE1222-I-WB-ST-23 CHECKED BY DATE SIZE DRAWING NUMBER REV

1
1 AE1222-I-WB-ST-24 XXX 1
XXX A1 AE1222-I-WB-Subassembly-B.2 A4
1 AE1222-I-WB-ST-25 DESIGNED BY DATE
Group E10E SCALE 1:5 WEIGHT(kg) 0.68 SHEET 1/1
XXX
30

P O N M L K F E D C B A
H
G
F
E
D
C
B
A
2

8
8

7
7

Isometric view
3 Scale: 1:5

Exploded View

6
Scale: 1:5 6

5
5

= 170 =
= 150 =
Right view

4
Front view 4
= 400 = Scale: 1:5 Scale: 1:5
125

= 500 = 155

Figure A.8: Subassembly C (own work)


3
3
Bill of Material: AE1222-I-WB-Subassembly-C
Number Quantity Part Number Type Nomenclature Revision
1 2 AE1222-I-WB-Wall-30 Part Sleeve Side Wall A4
2 1 AE1222-I-WB-SPL-31 Part Sleeve Top/Bottom A4
"This is the Production Plan in Appendix E"

Plate
3 1 Symmetry of AE1222- Part Sleeve Top/Bottom A4
I-WB-SPL-31 Plate

2
Recapitulation of: AE1222-I- 2
WB-Subassembly-C
Different parts: 3 This drawing is our property.
It can't be reproduced DASSAULT SYSTEMES
Total parts: 4 or communicated without
our written agreement.
Quantity Part Number DRAWING TITLE
2 AE1222-I-WB-Wall-30 DRAWN BY DATE
Bottom view dnikolov 12-May-21 Sleeve Assembly
1 AE1222-I-WB-SPL-31
Scale: 1:5 CHECKED BY DATE SIZE DRAWING NUMBER REV
1 Symmetry of AE1222-

1
XXX XXX
1
I-WB-SPL-31 A2 AE1222-I-WB-Subassembly-C A4
DESIGNED BY DATE
31

Group E10E SCALE 1:5 WEIGHT(kg) 4.06 SHEET 1/1


XXX

H G B A

You might also like