Introduction To Airplane Design
Introduction To Airplane Design
FT
RA
D
D
RA
FT
I NTRODUCTION TO A IRPLANE D ESIGN
FT
RA
by
D
Roelof V OS
FT
RA
D
A CKNOWLEDGMENTS
• Dr. ir. Willem Anemaat (DARcorporation) for stimulating conversations about this
textbook and for offering me a desk during my sabbatical in Lawrence, Kansas in
the summer of 2022. This gave me the rest to focus on this new endeavour without
distraction of daily activities.
• Harry Kwik (Airbus, GmbH) for help on the certification requirements of the emer-
gency exits.
• Inès van Teeffelen, Berend Domhof, Albert Chou for proof-reading the text, testing
the assignments, and providing feedback.
• Dr. Maurice Hoogreef for reviewing and proving valuable feedback.
FT
RA
D
v
D
RA
FT
C ONTENTS
Preface ix
1 Introduction 1
1.1 Engineering Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Learning Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 Outline of the Textbook . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2 Design Process 7
2.1 Airplane Design Process . . . . . . . . . . . . . . . . . . . . . . . . . . 7
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
9
2.3 Sizing versus Designing . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3 Requirements and Objectives
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
15
15
17
24
39
RA
4 Configuration Selection 43
4.1 Wing Positioning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
4.2 Propulsion System Integration . . . . . . . . . . . . . . . . . . . . . . . 51
4.3 Landing Gear Configuration . . . . . . . . . . . . . . . . . . . . . . . . 58
4.4 Tail Configuration. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
5 Fuselage Design 67
5.1 Design Considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
5.2 Preliminary Payload Volume Analysis . . . . . . . . . . . . . . . . . . . . 77
D
vii
viii C ONTENTS
FT
10 Aircraft Analysis and Design Iteration I
10.1 Estimation of the Drag Polar . . . . . . . . . .
10.2 Estimation of Propulsion System Characteristics
10.3 Result of Design Iteration on Design Objective .
10.4 Effect of Design Parameters on Design Objective
11 Wing Design Revisited and Design Iteration II
11.1 Design of Roll Control Surfaces . . . . . . . . . . . .
11.2 Design of High-Lift System . . . . . . . . . . . . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
213
213
213
213
213
215
215
215
RA
11.3 Estimation of Maximum Lift Coefficient . . . . . . . . . . . . . . . . . . 215
11.4 Result of Design Iteration on Design Objective . . . . . . . . . . . . . . . 215
11.5 Effect of Wing Design Parameters on Design Objective . . . . . . . . . . . 215
12 Empennage Design Revisited and Design Iteration III 217
12.1 Stability and Control Requirements . . . . . . . . . . . . . . . . . . . . . 217
12.2 Component Weight Estimation . . . . . . . . . . . . . . . . . . . . . . . 217
12.3 Generation of Loading Diagram. . . . . . . . . . . . . . . . . . . . . . . 217
12.4 Sizing for longitudinal stability, equilibrium and control . . . . . . . . . . 217
D
FT
nautical engineering studies. Typically, freshman or sophomore students who have a
good grasp of high-school mathematics and physics, but have limited knowledge of the
fundamental disciplines that underpin an airplane design: aerodynamics, structures,
propulsion, stability & control. In this textbook it is assumed that the reader is familiar
with the basic concept of flight of a powered fixed-wing aircraft: the concepts of lift, drag,
weight, and thrust. Furthermore, it is assumed that the reader has a basic understand-
ing of the flight mechanics of such a vehicle. All other concepts required to successfully
design an airplane are explained within the text and are used directly within the design
RA
process.
This book has been written to allow students in the very beginning of their studies
to enjoy the process of designing an airplane and learn about the integration challenges
that arise when doing so. While the book hardly introduces any novel design methods, it
does familiarize the student with a process that can be built upon in more advanced air-
plane design courses. It allows the reader to understand how the individual aeronautical
disciplines come together when a vehicle is designed.
As designing is a process, this process is best internalized by practicing it. Therefore,
D
this textbook has been written including multiple assignments to allow the student to
design their own airplane in a step-by-step manner. A particular focus is being placed on
the iterative nature of the design process: when or where are assumptions to be replaced
by analysis? What “fidelity” of analysis is required at what stage of the design process?
And, when do you stop the iterative process and consider a design to have converged?
These are some of the questions that are answered in the current text book.
As the amount of different flying vehicles is vast and ever expanding, in this text we
only treat the design of fixed-wing aircraft (i.e. airplanes) to limit the scope of the text-
book. The aim of this book is to give the reader an introduction into the airplane-design
process, rather than covering the design of as many flying objects as possible. On the
other hand, we do distinguish between two different energy sources on board the air-
craft: electric energy stored in batteries or chemical energy stored in liquid fuel. While
electric aircraft are, at the time of writing, still relatively new, it allows for flights without
emissions and could therefore contribute to climate-neutral aviation in the future.
ix
x P REFACE
A final unique aspect of this text book is that we also explicitly show how a change
in design objective affects the outcome of the design process. For example, an airplane
designed for minimal operating cost is different from an airplane designed for minimal
climate impact, even when either airplane has been designed for the same set of require-
ments. While it is acknowledged that the design of a fuel-burning engine has a substan-
tial impact on the emissions of an airplane, this has been left out of the scope of this
book.
With this book we provide anyone who likes to design an airplane a set of basic meth-
ods to do so. Once familiarized with the design process presented herein, they can ex-
plore more intricate analysis methods, include more design variables, or alter the order
of the design steps during a more advanced Airplane Design course. This book therefore
acts as a stepping stone to explore more aircraft design. As the title suggests, this book is
intended to really give the student an Introduction to Airplane Design.
Roelof Vos
Lawrence, KS, August 2022
FT
RA
D
FT
The design of any artifact is typically a creative process. Whether you are designing a
coffee mug or an airplane, designing inherently comes down to making decisions. For
the coffee mug, these could be the decisions about the size of the mug, the material it is
made of, and its color. While the mug needs to be manufactured, very few engineering
calculations are required during the design process. For the design of an airplane, de-
cisions also need to be made: what type of propulsion system is used, how is the cabin
RA
arranged, or how is the wing attached to the fuselage? The difference with designing a
coffee mug is that during the design of an airplane, many engineering calculations need
to be made on which many of the design decisions are based. Airplane design, as used
in the context of this book, is therefore a version of engineering design.
ABET, the Accreditation Board for Engineering and Technology, defines engineering
design as follows:1 Engineering design is a process of devising a system, component, or
process to meet desired needs and specifications within constraints. It is an iterative, cre-
ative, decision-making process in which the basic sciences, mathematics, and engineering
D
sciences are applied to convert resources into solutions. In our case, we consider the de-
sign of a product: an airplane. What is evident from this definition of engineering design
is that it is not a fundamental science but a process in which engineering science is being
applied. Therefore, in this textbook, mathematical formulas are often presented without
derivation. Where possible, we offer a reference to where the derivation of the formula
can be found.
A structured engineering design process is split into several phases in time. Each of
these phases consists of similar steps but with an increasing level of detail in the design
and increased fidelity in the applied analysis methods. Now, let us consider the typical
steps that are to be found in engineering design in each phase of the design process:
1. Define the problem
2. Establish requirements and objective
1 From: www.abet.org, retrieved 18 May, 2022.
1
2 1. I NTRODUCTION
FT
lem statement is as follows: “More and more people will be flying between city pairs
that are relatively close to each other. As each flight contributes to global warming, the
greenhouse-gas emissions of regional air travel need to be reduced.” As you can see, this
problem statement has an apparent problem (“contribution to global warming”) but also
hints towards the solution (i.e., airplanes with reduced greenhouse-gas emissions). This
problem statement would be a good starting point for a design process.
Requirements on the design can come from various sources: customers who will be
using the product, regulatory bodies that dictate safety standards, or subsystem suppli-
RA
ers whose systems need to be installed in the product. Typically, not all requirements
are used in all of the phases of the design process. It is up to you to state all relevant
requirements for that particular design phase. A design objective must also be stated to
enable the comparison of various design options. This usually means one would like to
minimize or maximize a specific performance aspect of the product. The word “perfor-
mance” should be interpreted here in the broadest sense, as it can relate to any design
aspect. A thorough analysis of the requirements and a statement of the design objective
is important before progressing to the next step in the design process. In Chapter 3, we
will tell you more about how to use requirements and objectives in the design process.
D
Setting up the design options is the creative part of the design process. In this step,
you might look at design solutions that have been proposed in the past. Why were some
designs successful and why did other designs fail? Are there some lessons learned that
can be applied in the current design? In this step, it is good to think in terms of functions:
what should your product be able to do? Should you use a variation of an existing design
solution, or are conceiving a completely original design solution? It is good to explore
multiple ideas, but you should also keep in mind that the solutions you propose in this
step must be engineered in the following steps. To describe each design option, we use
the concept of design variables. Design variables are entities that can change the shape
or properties of the design.
Your design options can be analyzed using engineering methods of various fidelity
and accuracy. In this step, it is key to select analysis methods that are suitable to the
phase of the design that you are in. The result of the analysis methods should allow you
to evaluate how the performance of the design compares to the requirements as well as
the design objective. In this textbook, we are at the beginning of the design process of an
airplane. Therefore, we select relatively simple analysis methods: empirical or analyti-
cal methods. However, in later phases of the design process, the methods can be more
complex and might require dedicated software, hardware, and/or experts to perform the
analysis. To select the right fidelity of the analysis method for each design stage, ensure
that the analysis method can distinguish between various design options such that a
comparison between the design options can be performed in the next step.
To make sure the various design options can be compared, it must be verified that
each of the design options meets the requirements. If a design option does not fulfill the
requirements, there is the possibility of making changes to the design variables. These
changes result in a different performance. For example, if your analysis tells you that
the airplane you are designing is unable to take off within the required take-off distance,
you could select an engine with more thrust to comply with this requirement. In many
instances, the verification of requirement compliance is closely tied to the analysis of the
performance aspects of the design. When we use requirements directly to determine the
FT
“right” dimensions of (a component of ) the design, we call this sizing.
To compare the performance of different design options, a list of performance cri-
teria should be made. Note that the performance criteria are complementary to the re-
quirements. If a design solution cannot meet the design requirements, it is disqualified
from the trade-off process that is being performed in this step. Usually, the performance
criteria are closely aligned with the design objective, i.e., the design’s performance as-
pect(s) that you try to minimize or maximize. For example, the design objective for an
RA
airplane could be to minimize the fuel required for a given mission while minimizing
its noise footprint around airports. In that case, the performance criteria could be fuel
burn and landing and take-off (LTO) noise. To enable the evaluation of these criteria, the
analyze step should, therefore, include analysis methods that can quantify these metrics
for each of the design options.
The next step is to make a choice. Given the previous step’s output, this might seem
like a trivial task. However, in practice, this choice can be relatively difficult for several
reasons. First of all, you might realize that the comparison performed in the previous
step is based on analysis methods that are inherently flawed. Analysis methods cap-
D
competing products? These are the types of questions that are answered at this stage.
It is a critical reflection on the design solution. If you see room for improvement, the
design can be further iterated, i.e., the design variables can be changed to increase the
performance of the design in a certain aspect. If you cannot further improve the design
by changing the design variables, you have finished this part of the design process.
A SSIGNMENT 1.1
In this assignment, you are going to define a design problem. As this is a creative
assignment, there is no right or wrong answer. To help you along in setting up the
problem statement, we have prepared the following questions. If you have trou-
ble answering these questions, you could also use an existing airplane instead.
a. What air transportation needs do you foresee in the near future?
b. What transportation need would you like to focus on during the course of
this book?
c. If there are existing design solutions for these transportation needs, how
FT
do you want your design solution to distinguish itself from what is already
existing?
d. Having answered these questions, write down your design problem state-
ment.
Unconventional airplanes, such as the blended-wing-body or the flying wing, are outside
the scope of this book. However, the design process applies to both conventional and
unconventional airplanes.
After having read this textbook and having performed the end-of-chapter assign-
ments, you should be able to:
• Formulate an airplane design problem
• Perform a structured airplane design process
• Perform sizing for components: propulsion system, wing, fuselage, and empen-
nage
• Integrate components and make a three-view drawing
• Iterate a design at various stages of the design process
• Know when and how to replace assumptions with analysis results
• Present your design results in a report
FT
are presented in the upcoming chapters can be used for other aircraft types as well, there
are also parts that only apply to this category of aircraft. It is therefore recommended to
select a design problem that fits within the scope of this book.
RA
D
FT
In this chapter, the airplane design process that we follow in this textbook is presented.
By completing the assignments that are presented in the subsequent chapters, you will
follow this design process and produce your own unique airplane design. Before you get
started, this chapter shows how the airplane design process is organized in the industry
(Section 2.1. Furthermore, we explain the process you are going to follow when designing
your own airplane(Section 2.2). We also introduce the concept of sizing and how it differs
from designing (Section 2.3).
RA
2.1. A IRPLANE D ESIGN P ROCESS
The airplane development process in the industry is organized in different phases. Fig-
ure 2.1 shows schematically which distinct phases can be recognized and which mile-
stones are present. It is important to distinguish between the different design phases of
an airplane as very different activities take place, different tools are used, different num-
bers of people are involved, and very different amounts of costs are associated with each
D
7
8 2. D ESIGN P ROCESS
Detailed Design
Manufacturing
Authorization to
proceed Testing
Major milestone
Support
First delivery
Service
Configuration design
investigation, requirements for future products can be derived. For example, the eco-
Engineering
Figure 2.1: Schematic of design and development process in industry (adapted from [15]). Note that the boxes
RA
nomic growth in parts of the world might lead to more demand for air travel between two
city pairs. A future airplane should be able to fly between those city pairs and, therefore,
be able to cover the range, i.e., the distance between those two cities. The combination
of market requirements and selected technologies are often prerequisites to commence
the design process of a new airplane.
The design process itself is divided into three non-equal phases: Conceptual Design,
Preliminary Design, and Detailed Design. In conceptual design, an airplane is being de-
signed taking into account the so-called top-level aircraft requirements (TLARs) . Using
a combination of sizing methods, creative thinking, and critical evaluation, an airplane
D
is conceived that meets these requirements and optimizes a well-defined design objec-
tive. The output of this process is typically a report that comprises a three-view drawing
of the airplane, including its major components (engines, landing gear, flaps, etc.), a list
of airplane specifications (i.e. components masses, center-of-gravity location, stability
and control properties, etc.), and a set of assumptions that have been made in order to
perform the calculations during the design process. The conceptual design of an air-
plane is typically performed within a matter of days to weeks by a relatively small team
of people. Simple analytical or empirical methods are used in the analysis and sizing
methods, which are often embedded in software tools such as AAA.1
In the preliminary design phase, the conceptual design is further analyzed and de-
tailed by a variety of disciplinary teams within the company. Within each discipline,
parametric studies are performed, and the baseline design concept is refined further.
1 AAA is a comprehensive airplane design program developed over 30+ years by DARcorporation
The preliminary design phase typically takes months to a few years and is performed by
much larger teams compared to the Conceptual Design team. The dimensions, specifi-
cations and assumptions from the Conceptual Design phase become part of the require-
ment set for the preliminary design phase. If you think about the aerodynamic design of
the wing, for example, the planform dimensions, thickness distribution, and structural
weight constrain the preliminary design of the wing. Furthermore, its performance in
terms of lifting capability or the drag it produces also constrains the design of the wing,
i.e., the wing needs to possess these capabilities in order to satisfy the performance spec-
ifications resulting from the conceptual design process. This also explains why there is
an overlap between the conceptual design process and the preliminary design process:
specifications from the conceptual design process might be too optimistic or too con-
servative, leading to an unfeasible or noncompetitive design, respectively.
At the end of the Preliminary Design phase, the design is frozen, meaning that the
top-level design variables, along with the performance specifications, are kept constant
from that moment onward. At this point, it needs to be decided whether the design is
progressing to the next phase of detailed design. This is an important decision as the
FT
amount of engineering that is required to perform the subsequent phases is relatively
high compared to the previous two phases. For many designs, this is the terminal sta-
tion of the design process and the design is shelved, only to be revisited in the future if
circumstances change. However, if the go-ahead is given to proceed with the design, the
outcome of the preliminary design process becomes the specifications for the Detailed
Design phase.
In the Detailed Design phase , the manufacturing drawings are being produced for
every component of the airplane. Here, a strong interaction with the manufacturing
RA
and testing of newly designed components exists because the airplane needs to demon-
strate compliance with the specifications stemming from the previous design phases as
well as additional specifications coming from the airworthiness regulations (see Chap-
ter 3). This phase typically takes multiple years to complete. Each airplane component
is broken down into parts, and an engineering design process is being performed for
each part. In this process, suppliers are often included as the expertise for any of the
subsystems, which does not necessarily exist at the Original Equipment Manufacturer
(OEM). This makes the detailed design process a multi-team, multi-disciplinary process
D
that requires good project management to ensure all the results of all efforts can be prop-
erly integrated into a successful product. Systems Engineering tools and methods have
proved themselves very useful to enhance this process.2 These methods are effective in
organizing and controlling the product development process.
Design Structure Matrix (DSM3 ). On the diagonal of this matrix the activities are shown:
sizing, analyzing, evaluation. The feed-forward connections are shown above and to the
right of the diagonal. Meaning that the output of one activity is the input of a subsequent
activity. Below and to the left of the diagonal are the feedback connections. The feedback
connections show that the process is iterative. If you look carefully, you can distinguish
the 3 through 5 of the engineering design process in the design structure matrix. In other
words, this process needs to be repeated for every design option before the options can
be compared.
The design process starts at the top left of the figure with the definition of the concept
and the selection of the design variables. It is assumed that the TLARs have already been
defined. The first activities that are proposed here are the estimation of the mass of the
airplane as well as the sizing of the wing and propulsion system. These activities are
often referred to as Preliminary Sizing and are the subject of Chapter 7. In the second
block, the fuselage and wing are being designed. Then, the wing is being positioned on
the fuselage based on the predicted location of the component centers of gravity. Also,
the propulsion system is further designed and integrated with the vehicle. Finally, the
can be constructed.
FT
landing gear and empennage are being integrated. Then, a three-view of the full airplane
Once the dimensions of the airplane have been established, a rudimentary aerody-
namic analysis can be performed to estimate the drag polar of the airplane. This can be
done in various configurations, i.e., with or without the gear extended and with or with-
out the flaps extended. Also, a component-based mass estimation can be performed to
estimate the mass of each of the components, i.e., the fuselage, the wing, or the propul-
sion system. Then, the aircraft’s center of gravity excursion during loading and unload-
RA
ing can be computed. The center of gravity is a common term for what is formally known
as the center of mass of an object, i.e., the spatial location where the resultant weight vec-
tor of the system acts. Furthermore, the maximum lift capability of the airplane can be
analyzed for various configurations, as well as its stability and controlability properties.
The output of all these analysis modules can be combined with the output from the
previous two activity modules to give us all the information we need to evaluate whether
all the constraints are satisfied and what the value is of the design objective. However,
we can also use the output of these analysis methods to replace assumptions that have
D
been made earlier in the design process to perform, for example, the mass estimation of
the airplane. If this feedback loop is utilized and the process is repeated, the outcome
from the analysis methods can be different from the first iteration. Usually, a small num-
ber of iterations (2 - 3) are needed to ensure that the output of the analysis methods has
converged. Convergence means that the difference in analysis results from two subse-
quent iterations is below a predefined threshold. For conceptual design, a threshold of
5% is typically sufficient. It is up to you to decide a priori what convergence threshold
you would like to use.
At the end of the design process, the constraints and objective values are evaluated.
When you follow the process all the way to the evaluation step, you have already used
many of the constraints and TLARs to size various aspects of the aircraft (see also Section
3 A design structure matrix is an example of a Systems Engineering tool to unambiguously describe a design
process.
o
o Propulsion system
Propulsion sizing
system sizing Take-off power or thrust
o Wing design
o Propulsion integra!on
Three-views Three-views
o Tail sizing
o Landing gear integra!on
Figure 2.2: Design structure matrix showing the design process proposed in this textbook.
FT
11
12 2. D ESIGN P ROCESS
2.3). Demonstrating constraint compliance for these requirements and constraints is,
therefore, trivial. However, there might be additional design constraints that need to be
evaluated at this stage. Furthermore, the output of the analysis methods should allow
you to compute the value of the design objective. If you are satisfied with the outcome
of these two evaluation processes, you can stop the process and document your design.
However, if either constraints are not satisfied or you think you can further improve the
design objective, there is a second feedback loop that brings you back to the beginning of
the design process. There, you can change some of the top-level design variables without
altering the concept.
In the second rectangular block of Figure 2.2, the words wing design and empennage
sizing are displayed in italic. We do this to emphasize that we use various fidelity meth-
ods to design the wing and the empennage surfaces. For the wing design we propose
here to first design the wing without the explicit definition of high-lift devices and/or
control surfaces. Then, after we have iterated the design once or twice, we refine the
wing design by including the wing design variables. For the empennage, we start out
by using a very simple method to dimension the size of the tail surfaces. After having
FT
done at least one design iteration of the airplane, including its wing movables, we em-
ploy a more sophisticated method to size the empennage. We do this to familiarize you
with the concept of multi-fidelity analysis. There are two reasons not to start with the
higher-fidelity analysis right from the start. First of all, there is not sufficient design in-
formation available at the beginning of the design process. Secondly, the higher-fidelity
analysis that is required to perform the sizing requires more time and effort to compute.
Therefore, you would like to limit the number of design iterations for which you have to
RA
perform this sizing process.
Example 2.1
We are designing a wing for an airplane, which only has one requirement: at sea level
and a speed of V = 100 m/s, the wing should be able to lift a mass of 50 metric tonnes.
The question is: how large should the wing area (S) be in order to fulfill this requirement?
It may be assumed that the lift coefficient in this condition is C L = 2.0.
To solve this problem, we employ two approaches, i.e., the design approach and the siz-
ing approach:
1. In the design approach, we would propose a design, analyze it, and compare it to
the requirement. Let’s propose the wing size is S = 25 m2 . The lift (L) that this wing
1
L = ρV 2C L S = 310 kN ∼ 31 t < 50 t
2
Here we have used a sea-level density of ρ = 1.225 kg/m3 . In other words, the wing
that we have proposed is not large enough to carry the mass of the airplane. We
can repeat this process multiple times until we find a wing size that is large enough.
However, that becomes quite tedious and there is an obvious way in which we can
do this more efficiently.
2. In the sizing approach we use the requirement directly. We know that 50 metric
tonnes equals 490 kN of weight and with lift equaling weight (W ), we rewrite the
lift equation as follows:
2W
S= = 40 m2
ρV 2C L
FT
As you can see, we have rearranged the lift equation and combined it with the lift re-
quirement to find a lower bound for the required wing size. Clearly, the sizing approach
gives us a precise answer to our design question much quicker than the design approach.
Sizing can often be a much faster way to dimension a component than using the design
approach. Therefore, in many steps of the airplane design process, one or more design
requirements are directly used to size a component.
In the previous example, we sized a wing. However, sizing applies not only to com-
ponents with physical dimensions that we measure in meters or feet. It can also apply to
RA
the thrust of an engine or the pressure in a tire. If a requirement has been used to size a
component using a simple algorithm like the re-written lift equation in Example 2.1, the
same equation can be used at the end of the design process (“constraint evaluation” in
Figure 2.2) to demonstrate that the design complies with that requirement.
D
FT
A problem statement is the starting point of any design process. However, if we need
to design an airplane, we have to know what requirements the design solution needs
to fulfill in order to solve the design problem. Therefore, this chapter is about design
requirements and design objectives. The requirements and objectives guide the design
process and, together with your design choices, result in your airplane design. If you
are designing an airplane for a customer, understanding what their requirements are is
an important part of the design project. Apart from these so-called top-level airplane
RA
requirements (TLARs), many requirements come from airworthiness regulations. These
also need to be taken into account during the design process.
that is cheaper or less polluting, respectively, than the competition. Alternatively, your
design could serve a completely original need. For example, the need to transport a
thousand passengers over just 500 km. This would be addressing part of the market for
which no dedicated airplanes exist. If there is a desire to address this market segment,
then the design objective might be less clearly defined. However, for a new airplane
design to be attractive to the market, the cost of the airplane and/or the cost to operate
the airplane should be as low as possible. Therefore, minimizing the life-cycle cost of
the airplane would be a suitable objective. The following example demonstrates how a
design objective can be formulated based on a problem statement.
Example 3.1
P ROBLEM S TATEMENT:
The CO2 produced by airplanes contributes to 3% of the total global CO2 emissions.
15
16 3. R EQUIREMENTS AND O BJECTIVES
Long-haul passenger airplanes are responsible for 45% of the global CO2 emissions. There-
fore, long-haul transport airplanes are an important contributor to global warming.
D ESIGN OBJECTIVE :
To design a long-haul passenger airplane that minimizes CO2 emissions
Apart from solving a design problem, there might be other triggers to start a design
process. For example, the maturation of new technologies that could improve a new air-
plane significantly. A well-known historical example is the introduction of the jet engine,
which increased the speed of passenger airplanes by a factor of two in the early 1950s.
Rather than being triggered by a need, the design process is then triggered by a seed.
In the case of military programs, the design process typically starts with a Request for
Proposal (RFP). On the base of the actual and foreseen political scenarios and the cur-
rent capabilities, the problem statement, top-level requirements and design objectives
for a new airplane type are included in the RFP and sent to various Original Equipment
proposal.
FT
Manufacturers (OEMs). Each of the OEMs is asked to respond to the RFP with a design
In the statement of the design objective, it is important that the objective is specific
and measurable. In Example 3.1, the objective is to minimize CO2 emissions. These
emissions can be quantified in the design process as long as analysis methods are in-
cluded that can compute this. In other words, the objective can be measured at the end
of the design process. However, it is not very specific. The objective does not state what
mission the airplane needs to perform for which the CO2 emissions should be mini-
RA
mized. An improved design-objective statement would therefore be: “To design a long-
haul passenger airplane that minimizes CO2 emissions per seat-kilometer over a 10,000-
km mission.” This design objective allows you to quantify the CO2 emissions in tonnes
by analyzing a mission of the airplane. To enable this, a mission-analysis method should
therefore be included in the design process as well as a method to quantify how much
CO2 is produced when fuel is combusted within the engine.
In general, the following aspects should be carefully considered before starting a new
airplane program:
D
• Introducing an airplane to the market at the right moment (not too early, not too
late), which is “better” than all competitors, if there are any.
• Capability to address the largest market share (customization vs. standardization).
Designing a commercial airplane for the special needs of a single customer might
only give you access to only a small portion of the market.
• Enter the market with the right technologies. The implemented technologies need
to be at the right level of maturity such that they can be manufactured by the OEM,
operated by customers, and repaired by a maintenance company.
• Clear understanding of development risks. Everything can be solved when time
and money are infinite, but in practice both resources are limited.
To investigate each of these aspects, continuous communication with potential cus-
tomers is often key. In the end, they should be the ones purchasing the airplane once
it has gone through its design, manufacturing, and testing cycle. Other stakeholders in
the process include maintenance companies, passengers, regulatory bodies, and suppli-
ers.
Example 3.2
FT
gardless of the type of airplane you are designing, the payload requirement is an impor-
tant first step.
For a commercial transport airplane, the payload requirement typically comprises pas-
sengers and cargo. Based on the interaction with potential customers, there might be a
list of payload requirements. This is an example of such a list:
RA
1. The airplane shall be able to transport 300 passengers and their luggage in a high-
density cabin configuration with a 31” seat pitch and 16.5” seat width.
2. The airplane shall be able to transport 250 passengers and their luggage in a typical
two-class configuration with 15% business class (55” seat pitch, 20” seat width)
and 85% in economy class with a 32” seat pitch and 16.5” seat width.
3. The airplane shall be able to store 20 m3 of cargo (excluding volume for luggage).
4. The airplane shall have a maximum structural payload of 35 tonnes.
D
The previous example shows that there are different payload requirements that do
not necessarily act at the same time. Often a so-called design payload is specified. The
design payload is the payload that you decide to be the most flown. In the example
above, the design payload would be the second item: 250 passengers plus their luggage.
The design payload can be the same as the maximum structural payload, but could also
be lower, as is the case in the previous example. In specifying the range requirements of
the aircraft, the design payload is coupled to the so-called design range, which we discuss
below.
You might have noticed from the previous example that the payload requirement
list is quite specific. The dimensions of the seat width and seat pitch are important to
determine the required volume for the passengers in the fuselage. Also, the required
volume for cargo is specified. Many transport airplanes carry cargo and luggage together
in the same cargo holds. If that is indeed the case, you have to compute how much
volume is required in the cargo hold to store the luggage and the cargo together. Note,
that for large transport airplanes, the carry-on luggage is stored in the main passenger
cabin. In other words, you need to investigate how much volume is required for luggage
in the cabin and how much volume is required for luggage in the cargo hold.
The average mass of a passenger at European airports in 2022 was 76 kg, and their av-
erage carry-on luggage weighed 8 kg. The mass of checked luggage is typically bounded
by a maximum set by the airline (i.e. 23 kg for long-haul flights). However, the average
checked luggage mass in 2022 at European airports was 16 kg. A single average passenger
at a European airport therefore has a payload mass of 100 kg. To enable the estimation of
the required volumes for storing luggage and cargo, you may assume an average density
of luggage of ρ luggage = 170 kg/m3 and an average density of cargo of ρ cargo = 160 kg/m3 .
A SSIGNMENT 3.1
FT
i. How many passengers are required to be transported?
ii. What is the mass of the passengers plus their luggage?
iii. How are the passengers divided among different classes?
iv. What seat width and seat pitch do you require?
v. How many cubic meters of luggage do you need?
vi. How much volume is required in the cabin for carry-on luggage?
vii. How much volume is needed in a cargo hold for luggage?
RA
viii. What is the cargo mass?
ix. How much cargo volume is required?
x. What is the maximum structural payload mass?
b. Answer the following questions relating to the design condition of the air-
plane: a
i. How many passengers are required to be transported?
ii. What is the mass of the passengers plus their luggage?
iii. How are the passengers divided among different classes?
D
To derive part of the flight performance requirements, you should define a mission
profile for the envisioned design. This can graphically show what it is that the airplane
needs to do. The mission profile can be a starting point for deriving the top-level aircraft
requirements. The following examples illustrate how one sketches a mission profile and
Example 3.3
The mission profile for a transport airplane is shown in Figure 3.1. It shows the two-
dimensional profile of the mission with the horizontal axis representing the distance
traveled and the vertical axis representing the altitude of the airplane. As can seen, each
phase of the flight is annotated to stipulate what is happening during that phase.
cruise to destination
taxi take-off
cruise altitude and speed
attempt to land
descent
land + taxi
RA
• Taxi: the airplane needs to be able to taxi on a paved taxiway
• Take-off: the airplane needs to be able to take off from a 2.5-km paved runway at
sea-level conditions
• Climb: the airplane needs to have a climb rate of 1 m/s at top-of-climb at cruise
speed and cruise altitude at 98% of its maximum take-off mass.
• The airplane shall cruise at a Mach number of M = 0.8
• The nominal range of the airplane with all passengers on board in a typical seating
D
The mission profile of Figure 3.1 shows two distinct ‘hops’. The first hop is the nom-
inal mission; i.e. the mission that is flown from origin to destination. The distance be-
tween these two points is referred to as the nominal range. The second hop is the diver-
sion part of the mission, which is only used when the airplane is unable to land at the
end of its nominal mission. It then needs to fly to an alternate airport, hold (i.e. loiter),
and land. To be able to fly to an alternate airport is often required by airworthiness reg-
ulations. However, the top-level aircraft requirements typically only specify the nominal
range. When designing the airplane, sufficient fuel needs to be taken on board to fly the
complete mission profile, including the diversion leg. Specifying the diversion range is
therefore required.
If we examine the requirements that we have derived based on the mission profile,
you can see that they are specific and measurable. Of course, these requirements are just
an example, but you can see how we use the mission profile to start deriving the mis-
sion requirements. It shows that from each phase of the mission, different requirements
might result. You need to gather all these requirements to the best of your ability prior
to starting the design process. Quantification of the requirements is important such that
you can verify compliance with the requirement at the end of the design process. There-
fore, you also need to select analysis methods that can quantify the performance aspects
that are listed in the requirements. That would allow you to determine whether your
design meets all the requirements.
Some of the requirements might seem somewhat ambiguous at first sight and we
need to make them more specific in order to be able to check whether they have been
met. For example, a ‘typical seating configuration’ is not specific. Therefore, we need to
FT
turn to the payload requirement to understand what is implied by this statement. The
other example of a seemingly ambiguous requirement is the statement to be able to land
in ‘all-weather conditions.’ However, this typically implies that you need to consider the
worst-case weather scenario and still make sure you comply with this requirement. It is
evident that on a dry runway, you can break much harder without slipping than on an
icy runway. Therefore, this requirement implies that you should be able to land within
the specified distance on a dry, wet, and icy runway.
RA
Example 3.4
Figure 3.2 shows the mission profile for a high-altitude reconnaissance airplane. Such
an airplane is typically used by the military to gather information. The airplane flies
for a long period of time over the same area and relays photographs through a satellite
connection to the ground station.
FT
The mission profile for the reconnaissance airplane is almost symmetric. This is be-
cause the airplane typically flies to a particular location, stays there, and flies back. For
any given airplane, there is a single lift coefficient at which the airplane cruises most
efficiently, i.e. it consumes the least amount of energy over a certain range. The com-
bination of speed, altitude, and weight determines the instantaneous lift coefficient of
an airplane in steady, symmetric rectilinear flight (see the lift equation in Example 2.1
RA
on page 12). For a given cruise speed and airplane weight, there exists a unique altitude
where the airplane cruises most efficiently. That is the altitude one would select to cruise
to the destination. When arrived, the airplane then ascents to the observation altitude to
stay there for an extended period of time. Then it flies back home using the same route
but then in reverse.
If we look at the requirements that have been deduced in Example 3.4, we recognize
some similarities with the ones that have been derived for the transport airplane. How-
ever, there are also a few distinct differences. First of all, the mission range is replaced
by a mission radius. This is because a military reconnaissance airplane typically does
D
not fly from A to B, but flies from airfield A towards its target and then back to airfield
A. When specifying a mission radius, you basically specify a circular domain around the
airfield that can be reached by the airplane while performing the mission specified in the
mission profile.
Secondly, we see that two climb-rate requirements are set at the end of each climb
phase. The reason why the climb rate is typically specified at the top-of-climb is because
at that altitude, the air has the lowest density. Air-breathing engines have less available
power at high altitudes because of the lower density. Therefore, if you can meet the top-
of-climb climb-rate requirement, you can sustain that climb rate also at lower altitudes.
In the requirement evaluation, it should be verified that the airplane fulfills both climb
rate requirements.
Finally, it might be observed that there is no speed specified to loiter. Therefore, this
speed might be freely chosen. If some of the performance aspects are not completely
specified, you might turn to the design objective for guidance on how to make this de-
cision. If we use the design objective of Example 3.1, we would try to minimize the CO2
emissions during the loiter phase. To minimize this, we should try to find a speed at
which the loitering flight requires minimum power from the engine, while still flying at
the observation altitude.
The payload and performance requirements form the top-level aircraft requirements
for your design problem. In practice, we often see that the requirements of payload and
range are specified in as a pair, i.e. a range specified for a given payload (as shown in
Example 3.3. Sometimes multiple payload-range pairs are specified in the requirements
and you need to demonstrate that your design complies with all of them. In this book,
we limit ourselves to a single payload-range pair. For this, we use the design payload
and the design range. The design range is the nominal range that needs to be flown with
the design payload.
A SSIGNMENT 3.2
airplane.
FT
In this assignment you will derive the mission-related requirements by perform-
tain altitude above sea level. For example, a take-off needs to take place from an airfield
that is elevated 1.5 kilometers above sea level. The second condition is the difference in
local temperature with respect to the International Standard Atmosphere (ISA). For ex-
ample, a requirement might need to be fulfilled on a warm day when the temperature
is 10 degrees Celcius higher than normal. We then write: ISA+10◦ C. For a given pres-
sure, an increase in temperature reduces the air density, which affects the performance
of the airplane. The final condition is the instantaneous mass at which the requirement
needs to be fulfilled. Examples of the mass are the maximum take-off mass, the maxi-
mum landing mass, or a fraction of the maximum take-off mass. More information on
the various aircraft mass definitions is provided in Chapter 6.
Example 3.5
In this example, we derive a fictional landing field length requirement for a jet airplane.
As is shown in Figure 7.5, the landing field length comprises an air distance and a ground
roll. We ask ourselves the following questions:
i. What is the minimum required landing field length?
ii. What runway condition is considered, i.e. dry, wet or icy?
iii. What runway surface is considered, i.e. grass, gravel, tarmacadam, concrete?
iv. What airfield altitude is considered?
v. What temperature w.r.t. ISA is considered?
vi. What airplane mass is considered?
To answer these questions we have to envision what airfields we have to land. As we
are designing a jet airplane, we envision it has to land at airports with a paved runway.
Therefore we select tarmacadam as the runway surface as this is present on many paved
runways around the world. Secondly, we choose to provide a landing field length re-
quirement for a dry runway, while understanding that the actual landing field length in
wet or icy conditions will be longer. We do this because our design methods for wing
sizing are tuned to dry runway requirements (see Chapter 7). We investigate the run-
way lengths at major airports and find many of them have airfields in excess of 3 km
FT
in length. However, we envision operating this airplane also from smaller airfields such
as London City Airport, which has a runway length of 1500 meters and is located at sea
level. We chose London City Airport as our reference airport for this airplane. We want to
be able to operate the airplane also on hot days when the local temperature is 30◦ C. This
is 15 degrees higher than the average temperature at sea level in ISA conditions (288 K or
15◦ C). Finally, we have to think about the mass of the airplane during landing. Here we
envision the scenario that the airplane needs to be able to take off and, in an emergency
situation, be able to turn around and land directly on the same airfield. Therefore, we
RA
require the airplane to land at its maximum take-off mass. The following list summarizes
the landing field length requirement:
i. The minimum landing field length, LLF = 1500 m.
ii. The runway surface is considered to be dry.
iii. The runway surface is considered to be tarmacadam.
iv. The airfield altitude is 0 m above sea level.
v. The temperature is ISA +15◦ .
vi. The landing mass is equal to the maximum take-off mass.
D
The previous example shows how you can derive a particular requirement. You can de-
rive these requirements on your own, using straightforward investigative tools. You can
also derive these requirements by interacting with other people. The questions that are
listed in the previous example can help you to make sure that the requirement formu-
lation becomes specific and measurable. This is important because we need these re-
quirements in the sizing of the wing and powerplant (see Chapter 8).
A SSIGNMENT 3.3
In this assignment, you are going to derive specific flight performance require-
ments based on the phases of the mission profile of Assignment 3.2. As you
are deriving the requirements, there is no single correct answer to each of these
questions. Also, for one or more phases of the flight, you might not have a spe-
cific performance requirement. These phases you can skip.
a. Take-off
i. What is the minimum required take-off distance?
ii. What runway surface is considered, i.e. grass, gravel, tarmac, con-
crete?
iii. What airfield altitude is considered?
iv. What temperature w.r.t. ISA is considered?
v. What airplane mass is considered?
b. Climb
i. What is the minimum required climb rate?
ii. What altitude is considered for this climb rate?
iii. What temperature w.r.t. ISA is considered?
iv. What airplane mass is considered?
c. Cruise
FT
i. What is the minimum required cruise speed (or cruise Mach num-
ber)?
ii. What cruise altitude is considered?
iii. What temperature w.r.t. ISA is considered?
iv. What airplane mass is considered?
d. Landing
i. What is the minimum required landing field length?
RA
ii. What runway condition is considered, i.e. dry, wet, or icy?
iii. What runway surface is considered, i.e. grass, gravel, tarmac, con-
crete?
iv. What airfield altitude is considered?
v. What temperature w.r.t. ISA is considered?
vi. What airplane mass is considered?
D
FT
(i.e. the kind of maneuvers it makes), and the powerplant technology (either turbine of
propeller). In Table 3.1 on page 26 we show the various airworthiness standards. For the
smallest airplanes, the US regulations and EU regulations differ slightly where the Very
Light Aircraft specifications have a maximum take-off mass limitation of 750 kg and the
Light Sport Aircraft specifications have this limit at 600 kg. For either category, a max-
imum stall speed of 45 kts is specified. The stall speed is the lowest speed an airplane
can sustain in level flight. Airplanes that are larger can be certified under CS/FAR-23
when they weigh less than 5670 kg (12,500 lb), have 9 seats or less (excluding the pilots),
RA
and fall in the normal, utility or aerobatic category. Commuter aircraft can still be cer-
tified under CS/FAR-23 when they weigh less than 8618 kg (19,000 lb) and have 19 or
less seats (excluding the pilots). All larger airplanes that do not fall under CS/FAR-23
are certified under CS/FAR-25. Note that CS/FAR 23 has a stall-speed requirement for
all single-engine airplanes as well as twin-engine airplanes of less than 2722 kg (6000 lb)
that are unable to meet a climb-gradient requirement of 1.5% at an altitude of 5000 ft
with one engine inoperative. There is no stall speed requirement for other twin-engine
airplanes, commuter airplanes, and large airplanes.
D
The propulsion system that is used in each category can differ. For small airplanes,
propulsion is provided through propellers. Large airplanes either use propellers or jet
propulsion. Propeller torque can be created by a reciprocating engine, an electric motor,
or a turboprop engine. A turboprop engine is a gas turbine that produces shaft power to
spin a propeller. In the VLA/LSA category, airplanes are required to have (a maximum
of) one engine or motor. CS/FAR-23 airplanes in the normal, utility, and aerobatic cate-
gories are also required to have a single powerplant, which can be either a reciprocating
engine or a turboprop. For the commuter category, one or two engines can be present.
Examples include so-called twin props as well as very-light jets (VLJs). Large airplanes,
certified under CS/FAR-25 are turbine-powered, either employing turboprop engines or
turbofan engines. In a turbofan engine,, the gas turbine is connected to a ducted fan, and
thrust is produced as a reaction to the jets coming from the exhaust of the gas turbine
and the ducted fan. The word ‘turbine’ in Table 3.1 refers to thrust being produced from
a turbofan engine. The following examples show the category under which an airplane
qualifies.
Example 3.6
The Pipistrel Velis Electro (Figure 3.3) is a two-seater airplane. The function of this air-
plane is to train pilots. It has a single electric motor powering a propeller. Its maximum
take-off mass is 600 kg. In terms of maneuvers, it falls in the ‘normal’ category meaning
FT
it can do: (1) Any maneuver incident to normal flying; (2) Stalls (except whip stalls); (3)
Lazy eights, chandelles and steep turns or similar maneuvers, in which the angle of bank
is not more than 60◦ . 4 Finally, it has a stall speed of 45 knots. The Velis Electro can
therefore be certified under CS-VLA as well as in the FAR-LSA.
Example 3.7
RA
The de Havilland Canada DHC-6 Twin Otter is a twin turboprop airplane. It can seat
a maximum of 19 passengers and has a maximum take-off mass of 5670kg. It is of the
‘utility’ category, meaning it can be used for transporting people as well as goods. In
terms of maneuvers, it should be able to handle the same maneuvers as in the ‘normal’
category plus (1) Spins (if approved for the particular type of aeroplane); and (2) Lazy
eights, chandelles, and steep turns, or similar maneuvers in which the angle of bank is
more than 60◦ but not more than 90◦ .
D
Example 3.8
The Cirrus SF50 (Figure 3.5) is a single-engine Very Light Jet. It is controlled by a single
pilot and can seat up to six passengers. It has a single turbofan engine and a maximum
take-off weight of 2,722kg (6000lb). It is certified under CS-23 in the ‘normal’ category.
It features a ballistic parachute system that can rescue the entire airplane in case of an
emergency. Such a parachute system is not required by CS-23, demonstrating that the
regulations merely provide a lower limit of safety. An airplane manufacturer can choose
to improve the safety of the airplane by adding systems like this parachute.
1 n =normal, u = utility, a = aerobatic
2 1.0 [kts] = 0.51 [m/s]
3 See main text for applicable conditions
4 See https://www.easa.europa.eu/sites/default/files/dfu/decision_ED_2003_18_RM.pdf, Ac-
cessed 2 June, 2022.
Figure 3.3: The Pipistrel Velis Electro is certified un- Figure 3.4: The DHC-6 is certified under CS/FAR-23
der CS-VLA. Photo: Airjuice Photography in the ‘utility’ category. Photo: Timo Breidenstein
Example 3.9
FT
The Embraer E190E2 (Figure 3.6) is a transport airplane seating 96 passengers in a typi-
cal two-class configuration. It has two turbofan engines and a maximum take-off mass
of 56,400 kg. This airplane clearly qualifies as a Large Airplane and is certified under
CS/FAR 25.
RA
D
Figure 3.5: The Cirrus SF-50 is certified under CS/FAR- Figure 3.6: The Embraer E190E2 is certified under
23 in the ‘normal’ category. Photo: Markus Eigenheer. CS/FAR-25. Photo: Valentin Hintikka.
When you define the mission of the airplane, it is important to also clearly state for
what airworthiness category the airplane should be designed. CS/FAR 25 contains more
and stricter requirements than CS/FAR 23. The same holds for CS-VLA compared to CS-
23. The latter contains more and stricter requirements than the former. If you design a
short-range two-seater airplane, you have to comply with fewer requirements if you in-
clude the top-level constraints stemming from CS-VLA. In practice, this implies that you
have to do fewer tests to demonstrate that the airplane complies with all certification re-
quirements. This reduces the development cost of the airplane compared to certification
under CS-23.
A SSIGNMENT 3.4
Once you have determined which airworthiness regulations apply to the design of
your airplane, it is time to think about which requirements need to be included at the
earliest stage of the design process. Depending on the regulations that apply, there can
be hundreds to thousands of requirements that the airplane needs to demonstrate com-
pliance with. To demonstrate compliance, each set of airworthiness regulations consists
of two books. Book 1 is the Airworthiness Code and contains all the requirements. It
is divided into several subparts to distinguish between various aspects of the airplane,
including structure, equipment, and power plant. Book 2 holds the Acceptable Means
of Compliance (AMC). It has the same subparts as Book 1 and explains how one can
demonstrate compliance with the requirements specified in Book 1. In the examples be-
FT
low, we present requirements that are important for the design process that is covered in
the present textbook. The curious reader is encouraged to explore the various CS or FAR
documents that can be found online.
The flight performance requirements are specified in subpart B of CS/FAR-23 and
CS/FAR 25. These requirements are going to be important for sizing the wing and power
plant in Chapter 7. We focus here on two sets of requirements: requirements that specify
the minimum speed the airplane should be able to fly and requirements that specify the
RA
minimum climb gradient the airplane should be able to achieve.
First, we look at the stall speed requirements in the airworthiness regulations. For
CS/FAR-23-certified airplanes a minimum stall speed is prescribed as is shown in the
following example:
Example 3.10
The previous example shows that a 61-knot (31 m/s) stall-speed (VS0) requirement ap-
plies to all CS-23 airplanes, but also that there are some exceptions. Reference is made
to other paragraphs in the regulations that detail these exceptions. If you wish to cer-
tify your airplane as a very light aircraft (VLA) or light-sport aircraft (LSA), the minimum
speed cannot exceed 45 knots (23 m/s). CS/FAR 25 does not state a minimum speed re-
quirement. However, as large airplanes need to be able to fly a specific traffic pattern in
order to land safely, they do need to fly slow enough to do so safely. Therefore, typically a
final approach speed should be specified (VAPP ) in the TLARs. According to CS-25 there
needs to be a speed margin of 23% between the minimum final approach speed (VREF)
and the stall speed (VSR0) as specified in CS/FAR 25.125:
Example 3.11
CS 25.125 Landing
(a) The horizontal distance necessary to land and to come to a complete stop from a
point 15 m (50 ft) above the landing surface must be determined [. . . ]
(b) In determining the distance in (a):
(1) The aeroplane must be in the landing configuration.
(2) A stabilised approach, with a calibrated airspeed of not less than VREF, must
be maintained down to the 15 m (50 ft) height.
(i) In non-icing conditions, VREF may not be less than:
(A) 1.23 VSR0;
FT
Here, the ‘landing configuration’ implies that the flaps are fully extended and the landing
gear is deployed. If a final approach speed is specified in the TLARs, the stall speed in the
landing configuration can easily be derived by dividing the final approach speed by 1.23.
The relation between approach speed (Vapp ) and stall speed (VS0 ) is defined in CS-25:
Vapp = 1.23VS0
While the conditions for the stall speed for CS/FAR-23 airplanes are listed in the para-
(3.1)
RA
graph describing the stall, the conditions can be chosen for the CS-25 airplane. For ex-
ample, one could choose to specify a minimum final approach speed at a lower mass
than the maximum take-off mass (i.e. at “landing mass”), at a higher altitude, or at an
increased temperature. The increase in altitude and temperature both result in a low-
ering of the density, which affects the generation of lift. This is why we need to be spe-
cific when listing our requirements. In the following assignment, you will derive the stall
speed requirement or approach speed requirement for your airplane.
A SSIGNMENT 3.5
D
Now, we will look at the climb gradient requirements that are specified in the airwor-
thiness regulations. The climb gradient tells you how steep you can climb (see Figure
3.7). You can imagine that this requirement relates to being able to clear obstacles after
a take-off. Example 3.12 is an example of a climb gradient requirement.
Example 3.12
FT
You can see that this requirement is very specific : four conditions must be met when the
requirement is verified. These conditions are related to the available power, the landing
gear configuration, the wing flap configuration, and the speed relative to the stall speed
in cruise configuration (VS1 ) and, when applicable, the minimum control speed5 . Sec-
ondly, you can also tell that the requirement is measurable, i.e. the climb gradient can be
quantified by dividing the climb rate by the forward speed. Both of these quantities can
RA
be measured in flight.
climb gradient:
c
G= ≈c /V ed: V
Vcos γ airspe climb rate:
c= V sin γ
climb angle: γ
horizon
D
Figure 3.7: Definition of the climb gradient (G), climb speed (c), and climb angle (γ).
The climb gradient requirement of Example 3.12 is just one climb-gradient example
stemming from CS/FAR-23. Several climb gradient requirements might be applicable to
your design and need to be verified. It is up to you to discover which climb-gradient re-
quirements apply to your design. The following paragraphs in CS/FAR-23 should there-
fore be consulted:6
These paragraphs cover the climb gradient requirements that must be met in various
parts of the mission profile. In paragraph 57, the take-off maneuver and requirements
are presented. Then, in paragraph 63 the conditions are specified, which apply in para-
graphs 65, 67, and 77. Then, paragraph 65 details the requirement in case of an all-
engines-operative (AEO) condition (see 3.12). For multi-engine airplanes, CS/FAR 23.67
specifies the requirements that need to be satisfied in case of a one-engine inoperative
(OEI) condition. Finally, paragraph 77 specifies the climb gradient requirements in case
of balked landing, i.e. a go-around near or from the ground. The latter needs to be eval-
uated in the AEO condition.
As CS-23 has four different categories of airplanes, there is a differentiation between
the climb gradient requirements for each of these categories. Furthermore, within the
FT
aerobatic, normal, and utility category, another distinction is made between airplanes of
2722 kg (6,000 lb) or less and airplanes heavier than 2722 kg. If you are at the beginning of
the design process and you do not yet know what the mass of the airplane is going to be,
it is recommended to state the set of climb-gradient requirements for both cases. After
the initial weight estimation (Chapter 7) it can be determined which of the sets should
be used. Similar to CS/FAR 23, CS/FAR-25 also has a set of climb gradient requirements.
Here is an example from CS-25.
RA
Example 3.13
The previous example describes the landing climb, i.e. the climb gradient after the
pilot has decided to abort the landing. In that case, the airplane is in landing configura-
tion, meaning that the flaps are fully extended and the landing gear is deployed. All of
the climb gradient requirements that need to be satisfied can be found in the following
CS/FAR paragraphs:7
In short, paragraph 111 describes the take-off path in detail, showing that it consists of
various segments of climb and acceleration. Paragraph 117 states that each of the re-
quirements in paragraphs 119 and 121 need to be shown at each weight, altitude, and
ambient temperature within the operational limits of the airplane. Paragraph 119 de-
scribes the climb gradient requirement in the landing phase in the AEO condition (see
Example 3.13). For each of the climb segments of the take-off path, paragraph 121 de-
fines the gradient requirement in case of an OEI condition. Note that in paragraph 121(d)
also a climb gradient requirement is provided for the approach segment, where the air-
plane is landing configuration. This ensures that in case of an OEI condition, after a
missed approach the airplane is able to climb. A summary of these requirements is
shown in Table 3.2.
Table 3.2: Summary of climb gradient requirements of CS/FAR 25.119 and CS/FAR 25.121. L = landing, TO =
take-off, CR = cruise, APP = approach.
climb gradient
CS/FAR condition Segment flaps gear mass Ne = 2 Ne = 3 Ne = 4
119
121 (a)
121 (b)
121 (c)
121 (d)
AEO
OEI
OEI
OEI
OEI
L
1st TO
2nd TO
3rd TO
APP
FT L
TO
TO
CR
L
↓
↓
↑
↑
↑
m MTO
m MTO
m MTO
m MTO
mL
3.2%
0.0%
2.4%
1.2%
2.1%
3.2%
0.3%
2.7%
1.5%
2.4%
Note that the requirements are different depending on the number of engines (Ne )
3.2%
0.5%
3.0%
1.7%
2.7%
RA
that you specify for your design. A two-engined airplane needs to comply with less strict
requirements compared to a three-engined or four-engined airplane in case of an OEI
condition. For each requirement, particular conditions apply: the flap configuration of
the airplane may be different, the landing gear may be up or down, and the mass of the
airplane is either the maximum take-off mass (m MTO ) or the maximum landing mass
mL .
A SSIGNMENT 3.6
D
Define the climb gradient requirements stemming from the airworthiness regu-
lations. If the design requirement depends on the number of engines, revisit this
assignment after you have completed Assignment 4.3.
a. Which paragraphs and subparagraphs from the airworthiness regulations
define the climb gradient requirements?
b. For each applicable paragraph and subparagraph, state the climb gradient
requirement along with the following conditions:
i. Weight
ii. Thrust available relative to maximum thrust available
iii. Landing gear configuration (i.e. extended or stowed)
iv. Wing flap configuration (i.e. cruise, take-off, or landing)
The climb-gradient requirements discussed above are part of the flight performance re-
quirements of the airplane, which also include requirements on the take-off distance,
the cruise speed, the climb rate, and the landing field length. However, you will also
need other requirements to design your airplane. To design your fuselage, you need to
know how many doors are required. Requirements for emergency evacuation can be
found in subpart D ‘Design and Construction’ of CS/FAR-23/25. A top-level requirement
for both CS/FAR-23 and CS/FAR-25 airplanes is that passengers should be able to leave
the airplane in case of an emergency situation within 90 seconds (see paragraph 803
CS/FAR-23/25). In the conceptual design stage, you cannot easily test this requirement.
Luckily there are other requirements that you can use to make sure this 90-second re-
quirement can be met. The example below shows a shortened version of a paragraph in
CS/FAR-23 describing the requirements for the placement of emergency exits.
Example 3.14
FT
crowding in any probable crash attitude. The aeroplane must have at least the
(a) For all aeroplanes with a seating capacity of two or more, excluding aero-
planes with canopies, at least one emergency exit on the opposite side of the
cabin from the main door specified in CS 23.783. [. . . ]
(b) Type and operation. Emergency exits must be movable windows, panels, canopies,
or external doors, openable from both inside and outside the aeroplane, that pro-
vide a clear unobstructed opening large enough to admit a 48- by-66 cm (19-by-26
RA
in) ellipse. [. . . ]
(d) Doors and exits. In addition, for commuter category aeroplanes the following re-
quirements apply:
(1) In addition to the passenger-entry door
(i) For an aeroplane with a total passenger seating capacity of 15 or fewer,
an emergency exit, as defined in subparagraph (b) , is required on each
side of the cabin; and
D
(ii) For an aeroplane with a total passenger seating capacity of 16 through 19,
three emergency exits, as defined in subparagraph (b), are required with
one on the same side as the passenger entry door and two on the side
opposite the door. [. . . ]
It may be noted that this requirement specifies the number and required dimensions of
the emergency exits for CS-23 airplanes. Note, that sub d gives special attention to com-
muter category airplanes, which again emphasizes the differences between the various
categories. When adhering to these requirements it may be assumed that the 90-second
rule of paragraph 803 may be complied with. Similarly, CS-25 also has regulations re-
garding the number of doors and their placement. In particular, paragraph 807 covers
this explicitly. There exist several standardized door types, which are defined in para-
graph 807(a) and are further detailed in Chapter 5. The largest standard exit is the Type
A exit, measuring 61 cm in width and 122 cm in height. The smallest is a Type IV exit,
which is 48 cm wide and 66 cm high, the same size as CS-23 emergency exits. In the
following example, the minimum number of emergency exits, along with their type, is
specified.
Example 3.15
(e) Uniformity. Exits must be distributed as uniformly as practical, taking into account
passenger seat distribution.
(g) Type and number required. The maximum number of passenger seats permitted
depends on the type and number of exits installed on each side of the fuselage. Ex-
cept as further restricted in subparagraphs (g)(1) through (g)(9) of this paragraph,
the maximum number of passenger seats permitted for each exit of a specific type
installed on each side of the fuselage is as follows:
Type B
Type C
Type I
FT
Emergency exits
(each side of fuselage)
Type A
Maximum number of
passenger seats allowed
110
75
55
45
RA
Type II 40
Type III 35
Type IV 9
A SSIGNMENT 3.7
The next requirements that stem from the regulations stem from subpart E ‘Powerplant’.
In this subpart, you typically find the requirements that have to do with the installation of
the powerplant. However, it must be noted that there are separate regulations regarding
the powerplant itself (e.g. CS-E or FAR-33 for engines and CS-P or FAR-35 for propellers).
Subpart E also specifies the requirements with respect to cooling, fuel systems, and the
placement of firewalls. Here is an example of a regulation regarding propeller clearance
in CS/FAR-25, which can also be found in CS/FAR 23.
Example 3.16
FT
Unless smaller clearances are substantiated, propeller clearances with the aeroplane at
maximum weight, with the most adverse centre of gravity, and with the propeller in the
most adverse pitch position, may not be less than the following:
(a) Ground clearance. There must be a clearance of at least 18 cm (7 inches) for
each aeroplane with nose wheel landing gear) or 23 cm (9 inches) for each aero-
plane with tail-wheel landing gear, between each propeller and the ground with
RA
the landing gear statically deflected and in the level take-off, or taxiing attitude,
whichever is most critical. In addition, there must be positive clearance between
the propeller and the ground when in the level take-off attitude with the critical
tyre(s) completely deflated and the corresponding landing gear strut bottomed.
[. . . ]
These examples from CS-23 and CS-25 show how detailed the requirements are speci-
fied. To comply with these requirements the AMCs of Book 2 can be consulted. How-
D
ever, in addition to the AMCs, the FAA and EASA have issued so-called advisory circulars
(ACs). These documents provide guidance for compliance with the airworthiness regu-
lations. Some of these ACs consist of merely a few pages, but others consist of tens of
pages and are quite extensive. While the ACs are not regulations and are not mandatory
to comply with, following the guidelines typically leads to compliance with the CS/FAR
regulations. Therefore, they could be perceived as defacto airworthiness regulations. Ad-
visory Circulars can be found online at www.faa.gov. The following example describes
the design considerations that need to be taken into account when positioning turbine
engines and auxiliary power units (APUs). An APU is a gas turbine that provides power
to the airplane in case the main engines are off.
Example 3.17
FT ~
RA
D
(7) Design Considerations. [. . . ] The most effective methods for minimizing the haz-
ards from uncontained rotor fragments include location of critical components
outside the fragment impact areas or separation, isolation, redundancy, and shield-
ing of critical airplane components and/or systems. The following design consid-
erations are recommended:
a. Consider the location of the engine and APU rotors relative to critical com-
FT
ment as defined in Paragraph 9a.
(ii) Intermediate Fragment. There is not more than a 1 in 40 chance of catas-
trophe resulting from the release of a piece of debris as defined in Para-
graph 9. [. . . ]
When installing propellers, you must also take into account that a propeller blade can
break off during rotation. Similar to the turbine discs in the previous example, these
RA
broken propeller blades could cause major damage when they impact with airframe.
Therefore, several design requirements in CS/FAR 23 and CS/FAR 25 pertain to the instal-
lation of the propeller with respect to critical systems. In CS/FAR 25.771, the propeller
location with respect to the pilot compartment is presented. CS/FAR-25.905 specifies
what precautions need to be taken in order to minimize the risk of catastrophic failure.
In addition, AC 25.905-1 presents further design guidelines for the safe installation of
propellers. The following example presents text from these two documents. When we
discuss power plant installation in Chapter 4, we will address what implications these
D
Example 3.18
CS 25.905 Propellers
FT
(d) Design precautions must be taken to minimise the hazards to the aeroplane in the
event a propeller blade fails or is released by a hub failure. The hazards which must
be considered include damage to structure and critical systems due to impact of a
failed or released blade and the unbalance created by such failure or release.
AC 25.905-1 - Minimizing the Hazards from Propeller Blade and Hub Failures
7. Design Practices to Minimize Blade Fragment Hazard
a. General
RA
(1) Techniques defined in AC 20-128A for minimizing the hazards following
an uncontained engine rotor failure (i.e. separation of critical systems,
isolation of functions, redundancy of functional elements, or shielding)
are also applicable when minimizing damage from propeller blade frag-
ments. However, the numerical assessment of fragment size defined in
paragraph 9 of AC 20-128A is not applicable for the propeller.
(2) Applicants should take all practical precautions in the airplane’s design
to minimize, based on good engineering judgment, the risk of catas-
trophic effects due to the release of part of a blade or a complete blade.
D
A final example that we include here is a set of requirements defined by the Interna-
tional Civil Aviation Organisation or ICAO. ICAO fosters the development of interna-
tional air transport to ensure safe and orderly growth.8 Their Annex 6 “Operations of
Aircraft” prescribes how much fuel an airplane needs to take on board to ensure a safe
flight. A distinction is made between commercial aviation and general aviation. Com-
mercial aviation is that part of aviation where operators provide a transportation service
through the purchase of a ticket. General aviation is everything else, except for military
aviation. For general aviation, the distinction is made between flying under Instrument
Flight Rules (IFR) and Visual Flight Rules (VFR). Whether you fly an airplane under IFR
or VFR depends on the weather conditions. Note that the wording in the example below
only implies fuel as an energy source. In this text, we assume that the same requirements
8 See: online description of ICAO.
would apply to battery-powered airplanes and that the word “fuel” would be replaced by
the word “energy.”
Example 3.19
For commercial aviation, ICAO Annex 6, Part I, Section 4.3.6 “Fuel Requirements” re-
quires the fuel quantity ahead of a flight to include:
• Taxi fuel
• Trip fuel (to reach intended destination)
• Contingency fuel (5% of “trip fuel” or 5 minutes of holding flight, whichever one is
more)
• Alternate destination fuel (to fly a missed approach and reach an alternate airport)
• Final reserve fuel (45 minutes of holding flight for reciprocating engines, 30 min-
utes for turbine engines)
• Additional fuel (if needed to guarantee ability to reach an alternate with an engine
failure or at lower altitude due to a pressurization loss)
FT
• Discretionary fuel (if the pilot in command wants it)
For general aviation, ICAO Annex 6 Part II, Section 2.2.3.6 “Fuel and oil supply” requires
the fuel quantity ahead of a flight to include:
• For IFR,9 enough fuel to reach destination, then alternate (if required), plus 45
minutes
• For day VFR,10 enough fuel to reach destination plus 30 minutes
• For night VFR, enough fuel to reach destination plus 45 minutes
RA
Note that the fuel requirements from the example above partially overlap with the ones
we deduced from our mission profile of a transport airplane (Example 3.3). You can
use the requirements from ICAO’s Annex 6 to harmonize the overall requirements you
defined based on your mission profile. Note that ICAO does, in some cases, require the
airplane to reach an alternate airport in case of emergency without specifying how far
this alternate airport is away from the destination airport. Therefore, this range should
be decided and made explicit in the top-level requirements. Finally, in case you design
an electric airplane, it can be assumed that the word fuel should be replaced with energy.
D
In this section, we have provided you with numerous examples stemming from the
airworthiness regulations as presented in the CS/FAR documents. Many of the presented
examples are going to be used in the design process of your airplane. Also during the de-
sign process itself, we might introduce more requirements stemming from the airwor-
thiness regulations to ensure compliance later on. It is by no means necessary to know
all the regulations by heart, but it is important to know where to find the relevant docu-
ments and how to use them in the design process.
typical chicken-and-egg problem where you need an airplane to analyze something and
analysis results to design an airplane. To avoid this problem, you need to make assump-
tions about some aspects of the design before you can perform any sizing or design step.
During the design process, these assumptions might be replaced in subsequent itera-
tions of the design. If an assumption is replaced by an analysis result, it does not need to
be communicated in a design report. However, if an assumption is used throughout the
design process, you should report it explicitly in a design report. This is important be-
cause the assumption might relate to an analysis result of the design and therefore affect
one or more performance metrics of the design.
The assumptions that are needed to start the design process pertain to aspects of the
design that are not known yet. For airplanes, we typically look at previous airplanes to
quantify these aspects. The best airplanes to choose in this respect are airplanes that
have been designed for a similar set of TLARs as the ones that you are using for your own
design. For example, if you need to design a 6-seater passenger airplane, it makes sense
to look for other airplanes in the 4-8 seat range. The airplane that you use in the design
process to derive assumptions from are your reference airplanes. The following example
illustrates this.
Example 3.20
FT
You have been tasked to design an airplane to transport 50 passengers over a distance of
1000 km (560 nmi). Which reference airplanes would you choose to base your assump-
tions on?
A quick search on the internet tells us that we are dealing here with a regional airliner.
RA
Based on the TLARs of range and pax, we can quickly find the following five airplanes:
In Example 3.20, we only selected five reference airplanes. In practice, it is often ben-
eficial to have more than that. Typically, between 5 and 10 airplanes are sufficient for
the design process that we use here. Depending on the design specification, you might
be able to find more or less reference airplanes. Regional airliners of 50 passengers (or
around that number) have been around since the 1950s. Therefore, you could have a
long list of potential reference airplanes. However, older airplanes might not have the
same technology as newer airplanes. Therefore, newer airplanes get priority over older
airplanes when you have to choose. Usually, the reference airplanes can differ some-
what from the specifications you have been given. In the previous example, you might
have noticed that the range of most airplanes is larger than 1000 km. Similarly, the num-
ber of passengers is also somewhat higher than specified. This is almost unavoidable
and should not negatively influence the design process. However, when selecting refer-
ence airplanes, priority should be given to the airplanes that mimic the specifications as
closely as possible. In Example 3.20, the ATR 42-600 comes closest to the specification
(see Fig. 3.9).
FT
Figure 3.9: The ATR 42 is selected as one of the reference airplanes in Example 3.20. Photo: KlausF.
As is evident from the design process of Figure 2.2, multiple steps in the design pro-
RA
cess require input from assumptions. When we discuss the sizing and analysis methods
in the Chapters 7 through 9, this is made more explicit. In a design report, it is advised
to present your reference airplane in more detail, including data that is relevant for the
design process. The following example illustrates this for the ATR-42.
Example 3.21
The ATR 42-600 is a regional airliner. It has turboprop engines, a high-wing configura-
tion and a t-tail. The main landing gear is connected to the fuselage and is retractable.
D
The maximum payload mass is 5.45 tonnes, the empty mass is 11.5 tonnes, and the max-
imum take-off mass is 18.6 tonnes. It typically seats 48 passengers and has a cruise speed
of 300 kts. The wing area is 54.5 m2 and the engines each produce a maximum of 1,300
kW of shaft power at sea-level.11
while for others only a handful were built. This provides some context to the design you
are making and could potentially help you in making design decisions later in the design
process.
A SSIGNMENT 3.8
Based on the TLARs of the airplane you need to design, you need to select and
present a set of five reference airplanes. The following tasks and questions will
help you in the process:
a. List the following TLARs for your airplane design: number of passengers,
payload mass, and mission range.
b. Find five airplanes whose specifications resemble the TLARs of the air-
plane you need to design. For each reference airplane list the following
items in a table:
i. Manufacturer
ii. Airplane name
iii. Number of passengers
iv. Maximum payload mass (kg)
v. Maximum range
vi. Maximum take-off mass (kg)
FT
vii. Empty mass or operating empty mass (kg)a
viii. Wing area (m2 )
ix. Wing span (m)
x. Maximum take-off power or maximum take-off thrustb
RA
c. For each airplane, add a photo that shows what the airplane looks like.
a Whichever one is listed in your source
b Whichever one is applicable to your design choice
D
FT
In the previous chapters, you have estimated the lift, drag, and efficiency characteristics
of your airplane. You have made design choices on the type of energy carrier, the number
of engines, the aspect ratio, etc. You have computed the characteristic masses of the
airplane and you have sized the wing and the propulsion system. The question that you
will answer in this chapter is: what will your airplane look like?
There is a large variation in airplane configurations that have appeared over the last
century. In Figure 4.1 we show three examples. We can see that the wing can be posi-
RA
tioned high or low and that the engines can be mounted on the wing, on the fuselage, or
buried in the nose of the airplane. We also notice some variation in the tail design. The
reason why these airplanes have different configurations has to do with their function-
ality. Each of them needs to comply with a different set of requirements. Some of these
requirements might play an important role in the choice of the most appropriate config-
uration. The second reason to prefer one configuration over another is the optimization
of the design objective. The assignments in this chapter help you to select an airplane
configuration that meets your design requirements while optimizing your design objec-
tive.
D
In Section 1.1, you have seen that the third step of the engineering design process is to
set up design options that can be a solution to your design problem. The choice of your
airplane configuration is an example of such a design option. If you select more than one
Figure 4.1: Examples of various airplane configurations. Photos by Alan Wilson, Pedro Aragao, and Anna
Zvereva.
43
44 4. C ONFIGURATION S ELECTION
Figure 4.2: Examples of different wing positions: high-wing (left, mid-wing (center), low-wing (right). Photos:
Péter Czégény, Alan Wilson, Textron
design option, you need to complete the entire design process for each option before
you can perform a trade-off and choose one of them as your design solution. In this
chapter, we ask you to choose a single design option to get experience with this process.
Contrary to the quantitative work that we performed in Chapters 6 and 7, this chapter is
much more qualitative in nature. We also limit the discussion of airplane configurations
FT
to “conventional” configurations that all have the same ingredients: a fuselage to store
the payload, a wing to provide lift, a propulsion system to provide thrust, a tail to provide
stability and control, and a landing gear to interface with the ground surface.
In this chapter, we have broken down the airplane configuration in four main as-
pects: the wing positioning (Sec. 4.1), the engine integration (Sec. 4.2, the landing gear
configuration (Sec. 4.3) and the tail configuration (Sec. 4.4). While we treat these config-
urational aspects in distinct sections of this chapter, we will also stress important cross-
links between them. By going through the examples and assignments in this chapter,
RA
you can make a well-funded decision about the configuration of the airplane you are
designing.
later in Chapter 9. We will show the advantages and disadvantages of these three differ-
ent wing positions. At the end of the section, you will be asked to make a well-funded
decision about the position of the wing of the airplane you are designing.
The vertical position of the wing with respect to the fuselage influences many aspects
of the airplane such as:
1. Aerodynamic interaction between wing and fuselage
2. Aerodynamic interaction between wing and horizontal tailplane
3. Accessibility to the fuselage
4. Type and positioning of engines
5. Positioning of landing gear
6. Airplane lateral stability
7. Airframe structural design
8. Crashworthiness
9. Pilot visibility
We will present each of the three wing positions and highlight how each of these aspects
is affected.
FT
RA
D
Figure 4.3: The Lockheed C-5 Galaxy is a military transport airplane with a high-wing configuration. Photos:
Phylyp, Roland Balik, US Army Europe, and Eric Purcell.
The high-wing configuration of this airplane also allows for a larger vertical distance
between the engines and the ground. This makes it easy to comply with ground clear-
ance requirements without the need for tall landing gear. For airplanes that need to
take off from unpaved runways, sufficient clearance between the ground and the en-
gines needs to be preserved to prevent the engines from ingesting dirt or debris from the
runway. This can cause so-called foreign object damage (FOD) to the engine. Figure 8.14
illustrates the dust and debris that is kicked up during take-off from a gravel runway. An-
other reason to increase the ground clearance could be to increase the distance from the
deployed flaps to the ground. Seaplanes need to keep the wing and the installed engines
out of the reach of the water spray. They also benefit from a high-wing configuration.
Figure 4.4: High-wing airplanes with wing-mounted engines allow for a larger ground clearance and reduce
the risk of debris ingestion during landing and take-off. Photos: US Airforce.
If a high-wing configuration is selected, the main landing gear can either be con-
nected to the wing or mounted to the fuselage (see Figure 4.5). The latter option has been
chosen for the C-5 because the resulting landing gear height would be very large when
connected to the high wing. When integrating the main landing gear with the fuselage,
a support structure is added to the side of the fuselage, which ensures the landing gear
FT
has a sufficiently wide track. A fairing covering this structure and the retracted landing
gear is called a sponson and causes an increase in (friction) drag. The additional support
structure between the landing gear struts and the frames also increases the structural
weight.
RA
Figure 4.5: For a high-wing configuration, the landing gear can be mounted to the wing (left) or to the fuselage
(right). Photos: Aero Mongolia, Anna Zvereva
D
A high-wing configuration also increases the lateral stability of the airplane, com-
pared to the other two configurations. To understand the concept of lateral stability, we
first examine what happens to an airplane when it experiences a sideslip condition. In
a sideslip, the airplane is moving laterally with respect to the airflow direction. Such an
event could occur when a lateral wind gust occurs or when the airplane is landing in
cross-wind conditions. As the airplane is sideslipping, a rolling moment could develop.
1
This is notionally shown in Figure 4.6. In this figure, three scenarios are shown: one
where a restoring rolling moment due to sideslip occurs (stable), one where no rolling
moment occurs (neutral), and one where a rolling moment occurs that aggravates the
bank angle and increases the sideslip of the airplane (unstable). CS/FAR-23.177 and
CS/FAR-25.177 require all non-aerobatic airplanes to have a stable rolling moment due
to sideslip.
1 A rolling moment is a moment about the longitudinal axis of the airplane that creates a bank angle.
neutral
unstable roll
due to sideslip
Figure 4.6: The rolling moment due to sideslip is dependent on the vertical position of the wing. Image after
Whitford [22].
lift increases. This causes a rolling moment in the counter-clockwise direction. A high-
wing configuration of the wing therefore increases the lateral stability of the airplane,
while a low-wing airplane decreases the lateral stability.
ΔL
ΔL
ΔL
Figure 4.7: Change in lift over the wing due to side-wind for a high-wing configuration (left) and a low-wing
configuration (right)
expense of little wetted area increase. Alternatively, the landing gear can be stored in the
nacelle behind the engine for airplanes with a wing-mounted propeller engine. In other
words, the low-wing configuration allows for a low-drag and low-weight integration on
the landing gear.
FT
RA
Figure 4.8: A low-wing configuration allows for an efficient integration with the landing gear. Photos:
Poudou99, Bob Adams, Peter Haas.
Another advantage of the low-wing configuration is that the wing is easily accessible.
For large airplanes, this allows the airplane to be easily fueled from below. Also, inspec-
tion of the wing can be easily performed on the platform prior to every departure. If
engines are installed on or below the wing, they, too, can be easily checked and serviced.
Large passenger airplanes often have a low-wing configuration. Therefore, the fuse-
D
lage is relatively far off the ground. This means that airports need to be equipped with
all kinds of lifts and conveyor belts to transport luggage and cargo onto the airplane’s
cargo deck (Figure 4.9). Passengers need to board via air stairs or through a dedicated air
bridge. Also, the trucks that service the galleys need a lifting platform that can interface
with the main passenger deck. Without all of this ground equipment, the airplane cannot
be properly serviced. However, there are also examples of airplanes with a low-wing con-
figuration that are self-supporting, indicating that a low-wing transport airplane could
also do without all of this equipment. In that case, creative solutions need to be found to
elevate passengers and goods to the passenger deck and freight deck, respectively.
The gear should be tall enough for a low-wing configuration with a wing-mounted
propulsion system to provide sufficient ground clearance. You can imagine that when
the airplane’s shock absorber and tire are completely compressed, the engine cowling
or propeller blade should still have a proper clearance to the ground (see also Chapter
9). For wing-mounted jet engines with a high bypass ratio, requires the engines to be
Figure 4.9: Low-wing passenger airplanes need elevators and conveyor belts to transport cargo and luggage to
the freight deck. Photos: Downtowngal, Jamesshliu
positioned ahead of the wing and require a long crane beam to connect the engines to
the wing box. To increase the ground clearance between the engine and the ground, a
few degrees of wing dihedral can help.2 Alternatively, one can choose to position the
FT
engines on the fuselage, which might allow a shorter landing gear. However, this has
other implications, as we discuss in Section 4.2. Figure 4.10 shows these two options.
RA
Figure 4.10: For low-wing configurations, jet engines can either be integrated under the wing (left) or on the
fuselage (right). Photos by Anna Zvereva and Dmitry Mottl.
D
The low-wing configuration reduces the lateral stability of the airplane, as we dis-
cussed in Sec. 4.1.1. However, increasing the dihedral angle of the wing can offset this
effect and allow the airplane to comply with the lateral stability requirements.
From a structural perspective, the easiest way to integrate a low wing with the fuse-
lage is to mount the fuselage on top of the wing. This allows for an unobstructed fuselage
structure connected to an unobstructed wing structure. For (near-) circular pressurized
fuselages, a relatively large streamlined fairing between the wing and the fuselage is re-
quired. Such a wing-body fairing is common among business jets, and it increases the
wetted area of the airplane. However, the unobstructed fuselage structure is relatively
light. Another commonly found solution is to let the wing go through the fuselage below
the main deck. This requires a smaller wing-body fairing and also brings the fuselage
closer to the ground (see Figure 4.11). However, it does increase the mass of the fuselage
2 The dihedral angle is the angle between the wing and a horizontal line in the front view.
structure as the structurally efficient circular tube has a large cut-out. A beneficial ef-
fect of the wing-body fairing is the increased available volume to stow the landing gear,
house part of the flap-deployment mechanism, and provide space for the pressurization
and air conditioning kit (PACK).
Figure 4.11: For pressurized fuselages, the wing structure can be positioned below the fuselage (left) or pierce
FT
through the fuselage below the passenger deck (right). Photos: James, Leo067
Figure 4.12: A military combat airplane (left) and general-aviation airplane (right) with a mid-wing configura-
tion. Photos: José Luis Celada Euba and Tim Felce
From a structural point of view, the mid-wing configuration can pose a challenge.
The highest bending moments occur at the root of the wing. Therefore, we prefer an
unobstructed wing box from tip to tip. A passenger airplane with a mid-wing configu-
ration would then have the wing box through the passenger cabin. For military combat
airplanes, this problem is less acute because there are no passengers there. However, a
large intake duct to the engine can be present, which could interfere with the wing struc-
ture. Another possibility is to position the wing behind the passenger cabin as is shown
on the right-hand side of Figure 4.12. However, you then have to be sure that you can
balance the airplane properly. In other words, you must make sure that the center of
gravity is close to the center of lift in terms of their respective longitudinal positions. In
the example of Figure 4.12 this is ensured by adding a foreplane.
A SSIGNMENT 4.1
In this section, you have seen examples of three wing configurations. Based on
your top-level requirements and your design objective, what wing configuration
do you choose for your airplane? Explain why you have chosen that configura-
tion.
FT
All powered airplanes have a propulsion system on board that provides thrust to the air-
plane. To provide thrust, we need an energy source as well as a machine to convert that
energy source into thrust. In this textbook, we consider two types of energy sources:
chemical energy and electrical energy. Chemical energy is stored in fuel (e.g., kerosene,
hydrogen, or avgas), while electrical energy is stored in a battery. Fuel can be converted
to mechanical power (or shaft power) by combusting it in a piston engine or a gas tur-
bine. A gas turbine that drives a propeller is termed a turboprop engine. An example of
RA
a turboprop engine is the Pratt & Whitney PW120, which is depicted in Figure 6.9. Elec-
trical power can be converted to shaft power by an electric motor. With the available
shaft power, a propeller can be driven, which produces thrust. In a turbojet engine, fuel
is combusted to produce thrust by accelerating the combustion products through a noz-
zle. If a fan (i.e., a multi-bladed ducted propeller) is connected to a shaft of the turbojet
engine, we call this a turbofan engine. These four propulsion systems are schematically
depicted in Figure 4.13.
In this section, you will decide what energy carrier your airplane will have, what
propulsion system your airplane will have, and how the propulsion system will be in-
D
tegrated. We will present various engine integration options and discuss the advantages
and disadvantages of each. Then, at the end of the chapter, you are asked to decide how
you wish to integrate your propulsion system with your airframe.
In Chapter 3 we presented some of the requirements that pertain to the integration
of propellers and jet engines. If you are designing a propeller airplane, it is advised to
read through Examples 3.16 and 3.18. They describe the clearance requirements and
how to prevent catastrophic failure in case of a blade failure of the propeller. You can
consult Example 3.17 for turbine engines such as turbofans and turboprops. It tells you
to position the engine in such a way that structural failure of a turbine disk fragment
does not lead to catastrophic failure of the airplane. This should be kept in mind when
positioning a propeller/engine with respect to other propellers/engines on the airplane
or with respect to critical systems such as the fuel system, the flight control system, or
the fuel tanks.
fuel
C T
fuel
C T
battery
T = turbine.
plane of the airplane. Most single-prop airplanes have a tractor configuration, meaning
the propeller is positioned in the nose of the airplane, ahead of the center of gravity.
The propeller airplanes in Figure 4.2 are examples of single-prop airplanes where the
propulsion system is integrated in the nose of the airplane. If the propeller is positioned
behind the center of gravity, we call this a pusher configuration. . An example of either
configuration is provided in Figure 4.14.
As can be seen in Figure 4.14, a propeller in the back of the fuselage prevents the in-
tegration of a conventional tail. Therefore, a foreplane is added to balance the airplane
around the center of gravity. Due to the relatively heavy propulsion system, the wing
is positioned aft on the fuselage to provide lift as well as stability. For a tractor config-
uration, the wing is located further forward on the fuselage, and a long arm is formed
between the tail and the wing. One advantage of the tractor configuration is the fact that
the propeller receives clean flow, while the air to a pusher propeller has been influenced
by the presence of the fuselage upstream. This means that the propeller efficiency for a
Figure 4.14: Example of single-engine propeller airplanes with a tractor configuration (left) and a pusher con-
figuration (right). Photos: Pipistrel and Stephen Kearney.
FT
4.2.3. I NTEGRATION OF M ULTI -P ROP P ROPULSION S YSTEM
If you design an airplane with two propellers, there are several options to integrate the
propeller and engine/motor with the airframe. The most common configuration has
the two propeller engines mounted to the wing in a tractor configuration. Alternatively,
one can combine a tractor propeller and a pusher propeller in the in-line configuration
(see Figure 4.15). The in-line configuration has the advantage that a single engine failure
does not result in a yawing moment.3 This could reduce the required size of the vertical
tailplane. As for all pusher propellers, care must be taken that the propeller does not
RA
touch the ground when the airplane rotates during take-off. Therefore, the main land-
ing gear must be sized appropriately. Furthermore, integration with the tail could prove
challenging. Therefore, the example of Figure 4.15 shows a twin vertical tail supported
by two tail booms.
D
Figure 4.15: Integration of two propeller engines can be done in line with the fuselage (left) or attached to the
wing (right). Photos: Wally Cacsabre and Michael Miley.
The configuration with twin propellers mounted below the wing allows the nose to
be tailored for excellent pilot visibility. Furthermore, the mass of the engine causes a
bending moment in the root of the wing that opposes the bending moment introduced
3 A yawing moment is a moment about the airplane’s vertical axis
by the distributed lift force on the wing. The resulting wing-root bending moment is,
therefore, lower compared to the in-line engine configuration. This results in a lower
mass of the wing box. When positioned in a tractor configuration as in Figure 4.15, the
slipstream of the propeller causes the lift over the wing to increase. This has implications
for the stability and control of the airplane that goes beyond the scope of this book but is
explained in Ref. [6]. The airplane of Figure 4.15 also shows a strut supporting the wing.
While the strut increases the wetted area of the airplane and, thereby, the drag, it does
reduce the bending moments in the wing significantly and, thereby, the wing mass.
Figure 4.12 shows that also a wing-mounted pusher configuration is possible for a
twin prop. This has the advantage that the propeller noise is less high in the passen-
ger cabin. However, care must be taken that the wing wake and engine exhaust do not
cause an increase in noise when they pass through the propeller disk. This could create
a higher fly-over noise for people on the ground. Two other examples where the engines
are located on the rear fuselage are shown in Figure 4.16. To support the engine and
propeller, a relatively large pylon is required. The pylon is the structural component that
connects the engine (or motor) to the airframe. The propfan shown in the right-hand
FT
picture of Figure 4.16 is a cross-over between a turboprop engine and a turbofan engine.
RA
Figure 4.16: Example of fuselage-mounted propellers (left) and fuselage-mounted propfans (right). Photos by
Tim Rees and Andrew Thomas, respectively.
If more than two propellers are selected, then they are typically mounted to the wing.
D
Figure 4.17 shows examples of four-engined airplanes with wing-mounted engine instal-
lation. In a four-engine configuration there is statistically a larger chance that one of
the engines fails. Furthermore, more engines also mean that the maintenance costs are
higher. Finally, smaller engines suffer from a reduction in thermal efficiency when they
are smaller. However, there can still be good reasons to choose more than two engines.
One advantage is that in case of an engine failure, the loss in thrust is only 25% for a
four-engined airplane. Secondly, there might not be an engine that is powerful enough
to provide half of the required take-off power. In that case, you have to spread the take-
off power over (at least) four engines. A final reason could be that the combination of
flight speed and engine power would result in a poor match between the propeller and
the engine/motor. In Chapter 8, we will further detail how the propeller can be designed
to absorb the power of the engine/motor.
Although less common, choosing more than four engines can prove a feasible design
solution. Electric motors, for example, do not suffer from a loss in efficiency with size.
Figure 4.17: Four-engined airplanes with a high-wing configuration (left) and a low-wing configuration (right).
Photos by Ronnie Macdonald and RuthAS, respectively.
Furthermore, due to the lack of hinging and sliding components within the motor, they
are also less maintenance prone. To have a good match between the propeller and the
motor, it can be desirable to distribute the power over more than four motors. This con-
FT
cept of distributed propulsion has additional benefits: the aerodynamic interaction with
the wing is spread over the span, which improves the span-wise lift distribution in cruise
flight, the one-engine inoperative condition has little impact on the flight performance,
and the smaller propellers make it easier to integrate a short landing gear in case of a
low-wing configuration.
For wing-mounted engines, the mass of the engines relieves the bending moment
of the wing, which reduces its structural mass. On a low-wing, the engine can be easily
reached for inspection or for maintenance. However, the aerodynamic and structural
integration with the wing are quite challenging. While the engine receives clean airflow
regardless of the angle of attack or sideslip angle of the airplane, the wing experiences a
flow field that is affected by the engine. Care must be taken to prevent so-called interfer-
ence drag in the vicinity of the pylon, wing, and nacelle. Also, aerodynamic modifications
are required to prevent an early onset of stall at the location where the pylon interrupts
the leading edge of the wing. Finally, engines that are close to the ground can easily suck
in debris when they are operated at low (taxi) speeds and relatively high thrust.
The fuselage-mounted engines result in a clean wing, which prevents the compli-
cated aerodynamic and structural interaction of the wing-mounted engine. However,
the structural mass of the wing is higher for this engine configuration. On the other hand,
a one-engine inoperative condition results in a smaller yawing moment because the en-
gines are located closer to the symmetry plane. Also, the distance from the thrust vector
to the center of gravity is smaller, which reduces the change in pitching moment due to
thrust.4 Positioning the engine on the rear fuselage reduces the cabin noise for most
passengers. Also, the wing partially shields the fan noise to ground observers, which re-
duces the fly-over noise. On the other hand, the large mass of the propulsion system
causes the wing to be positioned further rearward for balancing reasons. This leads to
a larger center-of-gravity excursion during the loading and unloading of fuel and pas-
sengers (see Chapters 9, which 12) results in a larger horizontal tailplane. Finally, due
to the exhaust of the engine, the horizontal tailplane needs to be raised, resulting in a
cruciform tail or a T-tail (see Section 4.4).
An alternative integration solution is to use the so-called over-the-wing engines (see
Figure 4.18). In such a configuration, the engines are positioned on pylons on the up-
per surface of the wing. This allows for a short landing gear, the engine mass reduces
the wing bending moment, and the thrust vector is aligned with the center of gravity
of the airplane. There are two main challenges with this configuration: (1) the aerody-
namic interaction between the upper surface of the wing and the engine inlet and (2) the
FT
structural integration of a vibrating engine on top of a flexible pylon. To reduce the aero-
dynamic interaction, the engines are vertically separated from the wing with a relatively
large pylon, which increases the structural mass and the friction drag of the airplane.
RA
D
Figure 4.18: Integration of a jet engine on a large passenger airplane (left) and a business jet (right). Photos by
Alan Wilson and Michael Pereckas, respectively.
Three-engined jets and four-engined jets reduce the impact of an engine failure on
the flight performance of the airplane. On a three-engined airplane, the integration of
the third engine is typically done in the tail of the airplane. A critical aspect of the in-
tegration of this engine is the fact that the engine needs to be easily removable from
the airframe for maintenance. The engine can either be embedded in the aft fuselage
or positioned in the vertical tailplane as shown in Figure 4.19. When integrated into
the fuselage, an s-duct connects the inlet to the engine. Both configurations have their
challenges regarding the structural integration of the aft fuselage, tailplanes, and engine.
Also, aerodynamically, there is a strong interaction between these components, which
requires careful aerodynamic design in the preliminary design phase.
FT
Figure 4.19: The integration of the center-engine can be done in the tail (left) or embedded in the aft fuselage
(right). Photos by Boushh-TFA and New York-air, respectively
Modern four-engined jet airplanes all have the same configuration: four engines
hanging under the wing (see Figure 4.20). The engines are distributed to find a balance
RA
between a low wing mass on the one hand and an acceptable vertical tail size to balance
a one-engine-inoperative yawing moment. For very large airplanes, the selection of four
engines can be the only option to have sufficient thrust with the available engines. While
the under-the-wing engines are the only feasible option for high-bypass-ratio engines,
other integration solutions have been found for low-bypass-ratio engines such as inte-
gration in the wing root (de Havilland Comet), pairwise installation on the rear fuselage
(Ilyushin 62 and Vickers VC-10), or pairwise installation under the wing (Concorde). The
B-2 stealth bomber even has its four engines completely embedded in the fuselage.
D
A SSIGNMENT 4.3
In this section, we have shown the advantages and disadvantages of various op-
tions to integrate the propulsion system with the airframe.
a. What propulsion system do you choose for your airplane? Why?
b. What is the cruise speed or cruise Mach number of your airplane?
c. How many engines/motors have you selected for your design? Please, mo-
tivate this decision.
d. How do you intend to integrate the engines/motors with your airframe?
Why do you choose this integration solution?
e. For propeller airplanes, how do you intend to comply with blade failure
requirements for the chosen configuration? See Example 3.18 for more de-
tails on blade failure.
Figure 4.20: Two examples of four-engined jet airplanes: the Airbus A380-800 (in the air) and the Boeing 747-8
(on the ground). Photo by Kiefer.
g. Has the integration of the propulsion system changed your design decision
regarding the wing configuration? If so, explain.
If you have not stated your climb gradient requirement(s) because these are de-
RA
pendent on the number of engines, you can now revisit Assignment 3.6 and ex-
plicitly state the climb gradient requirement(s) for your airplane.
gear (also known as “tail dragger”). Both configurations are shown in Figure 4.21. At the
end of the section, you will be tasked to make a decision on the configuration of your
landing gear. In Chapter 9 you will subsequently learn how to integrate your chosen
landing gear with the airframe.
Figure 4.21: Example of a conventional landing gear for an aerobatic airplane (left) and a tricycle landing gear
for a very large transport airplane (right). Photos by Velodenz and Vasiliy Koba, respectively.
FT
wheel on the runway until the wing stalls, and then the main gear touches the ground.
The nose-high attitude on the ground with the stalled wing causes a lot of drag, quickly
reducing the airspeed and the landing field length. A three-point landing is similar but
with all three wheels touching the runway at the same time. Whether a three-point land-
ing or a tail wheel landing is performed depends on the stall angle of the wing. Another
advantage of the nose-high attitude is that during taxiing, the propeller has large ground
clearance, preventing gravel or debris to be sucked into the propeller.
RA
On the negative side, the high attitude during taxiing results in poor over-the-nose
visibility for the pilot. Also, the advantage of high drag during landing can be a disad-
vantage during take-off, which reduces the acceleration of the airplane until the speed
is high enough to raise the tail. In the case of a two-wheel landing with a large decent
rate, the inertia of the airplane causes an increase in angle-of-attack, which might result
in the airplane bouncing off the runway again.
While braking is more effective with conventional landing gear, it also comes with
a nose-over and a ground loop risk. Both events are due to the fact that the airplane’s
center of gravity is located behind the main landing gear. The nose-over event can occur
D
when the main wheels suddenly come to a standstill, and the momentum of the airplane
tips the airplane on its nose. This causes the propeller to hit the ground or sometimes
even to flip the airplane on its back. A ground loop can occur when asymmetric braking
is applied during deceleration. The resulting curved track that the airplane follows can
be aggravated by the centrifugal force that acts at the center of gravity of the airplane,
which resides behind the main landing gear. This destabilizing force can lead to large
yawing and rolling motions if not appropriately corrected for by the pilot. In short, land-
ing an airplane with a conventional landing gear is more difficult compared to landing
an airplane with a tricycle landing gear.
ensures that the airplane is in an upright position, which reduces the drag and provides
a horizontal floor for the passengers and freight when on the ground. Furthermore, the
pilot has good over-the-nose visibility during take-off and landing. Finally, in case of a
large decent rate and a two-wheel landing, the inertia of the airplane causes the angle
of attack to decrease, thereby shedding the lift and putting the nose wheel firmly on the
ground.
All of these advantages make the tricycle configuration the most widely adopted
landing gear layout. However, it comes with a price: mass. The nose landing gear (NLG)
is much heavier than the tail wheel because it is larger and needs to sustain higher (dy-
namic) dynamic loads. Furthermore, to carry the point load of the nose wheel with the
surrounding structure, local strengthening of the airframe is required, which increases
the mass even further.
Despite the higher landing gear mass, the tricycle landing gear is used in all airplanes,
ranging from very small two-seater airplanes (see Figure 3.3) to very large passenger air-
planes (Figure 4.21). While some large airplanes feature multiple struts combined with
FT
many wheels on the main landing gear (as in Figure 4.20), we still characterize this as a
tricycle landing gear. If you look at the examples that have been presented in this chap-
ter, you can see many more tricycle landing gear layouts.
the cost of increased complexity, cost, weight, and maintenance. It also requires a (large)
volume inside the airplane to be stowed, which complicates the integration of all the
subsystems on the airframe. Apart from the kinematic mechanism required to safely
deploy the landing gear, this is often combined with doors that open and close, which
further increases the complexity. Finally, it is inherently less reliable than the fixed land-
ing gear.
Whether you choose a fixed or retractable landing gear for your airplane depends on
the type of airplane you are designing. Using the methods Chapters 6 and 7, you can
quantify what the impact is of using a fixed or retractable landing gear. Because these
methods are not sensitive to qualitative aspects such as complexity and maintenance,
you need to make a (qualitative) trade-off. The question is: are the improvements in
the performance metrics (maximum take-off mass, take-off power, wing size) worth the
increase in complexity, part count, and maintenance cost?
Figure 4.22: A retractable landing gear requires volume to be stowed (left) and comprises many hinging parts
(right). Photos by Victor and Julian Herzog, respectively.
A SSIGNMENT 4.4
answer.
c. Has the decision of your landing-gear layout changed your decision on the
vertical wing position or how to integrate the propulsion system?
d. Is your landing retractable or fixed? Explain your decision.
RA
4.4. TAIL C ONFIGURATION
The function of the tail is three-fold: to balance, to stabilize, and to control the airplane.
In this section, we present the design options for the tail. We will present the conven-
tional tail configurations: low tail, cruciform tail, and T-tail, as well as more unconven-
tional tail configurations like the V-tail and the inverted Y-tail. For each of these configu-
rations, we will present the advantages and disadvantages. We will also present possible
synergies between the tail configuration and the wing configuration. Or, oppositely, we
D
will explain which combinations of tail configuration and wing position could lead to
problematic interactions between these two lifting surfaces. Finally, we will prompt you
to make a decision on the tail configuration for your airplane.
tail and the T-tail (see Figure 4.23). The low tail also allows for simple staggering of the
horizontal and vertical tailplane on the fuselage, meaning one is positioned ahead of the
other. This reduces the aerodynamic interference between the vertical and horizontal
tail.
Figure 4.23: Three commonly used tail configurations: low tail (left), cruciform tail (center), and T-tail (right).
Photos by Caribb c b n d, dylan3300 c z, and Nabil Molinari c b n a, respectively.
While the low-tail configuration might be the most common, there can be a good
FT
reason to raise the horizontal tailplane and mount it on the vertical tailplane. For twin-
prop airplanes with wing-mounted tractor propellers, the propeller slipstream can re-
duce the stabilizing effectiveness of the horizontal tailplane. Raising the horizontal tail
improves the stabilizing effectiveness of the tailplane. Jet-powered airplanes with en-
gines on the fuselage need to raise the horizontal tailplane to get it out of the hot exhaust
plume. Also, the tail cone of the fuselage could be occupied by other systems, such as
an air brake system or an auxiliary power unit that complicates the integration of the
horizontal tailplane with the fuselage.
RA
If a low tail is not possible, you can either select a T-tail or a cruciform tail. A cruci-
form tail has the advantage that it is still relatively close to the fuselage. This implies that
only part of the vertical tail structure needs to be reinforced in order to transfer the loads
of the horizontal tailplane to the fuselage. This also allows the vertical tailplane to have
a lower taper ratio5 which reduces its weight.
The T-tail configuration ensures the largest vertical separation between the wing and
the horizontal tailplane. This maximizes the stabilizing effectiveness of the horizontal
tailplane. When positioned on a vertical tailplane with a sweep-back angle, the arm be-
D
tween the horizontal tailplane and the airplane’s center of gravity is maximized. This
implies that the horizontal tailplane can be smaller, which reduces drag. The horizontal
tailplane also acts as an end plate to the vertical tailplane, improving the effectiveness of
the vertical tailplane as a stabilizing surface.
On the other hand, the T-tail requires a strong vertical tail to transfer the loads to the
fuselage. A low aspect ratio helps to reduce the weight of the vertical tailplane but also
reduces its effectiveness as a stabilizing surface. The taper ratio of the vertical tailplane
is close to one in order to provide sufficient volume at the tip of the vertical tailplane to
mount the horizontal tailplane.
T-tails are also notorious because of the so-called deep stall that they can cause. A
deep stall is an unstable stall where the airplane keeps on increasing its angle of attack
without the pilot being able to correct for this by means of the elevator. Such a situation
5 The taper ratio of a lifting surface is the ratio of tip chord to root chord
can occur when the wake of a stalled wing immerses the horizontal tailplane (Figure
4.24). The low dynamic pressure in the wake results in a reduction in the aerodynamic
force that can be generated by the horizontal tailplane and the elevator. To avoid this,
many airplanes use a system that pushes the nose down before the deep-stall angle of
attack is reached. Such a system might be mechanical (i.e., a stick pusher) or embedded
in software as part of the flight envelope protection system.
FT
RA
Figure 4.24: When stalling, the wing wake immerses the horizontal tailplane causing a deep stall. Image by GrahamUK,
Mysid, and Michael32710 c b a.
let us consider the possibility of having multiple vertical tail surfaces. Examples of twin
vertical tails and triple vertical tails are shown in Figure 4.25. Distributing the vertical
tails adds redundancy to the directional stability and control system, which is important
for military vehicles. When positioned at the end of the horizontal tailplane, the effec-
tiveness of the horizontal tail as a stabilizing surface increases due to the end-plating
effect. As the vertical tail forms the highest point on the airplane, distributing the ver-
tical tail area over multiple tails can ensure the airplane can meet a maximum height
requirement. Such a requirement could stem from the height of available hangars for
the airplane. A final motivation to choose multiple vertical tail surfaces is the aerody-
namic interaction with the wing or fuselage at high angles of attack. Depending on the
stall characteristics of the airplane, a twin fin might be more effective than a single fin in
such a situation.
The downside of having multiple tails is that the number of individual parts increases.
A higher part count typically implies higher manufacturing costs. If vertical tails are close
Figure 4.25: Examples of twin vertical tails (left and right) and a triple tail (center). Photos by Robert Sullivan
c p, Jez c b n d, and pqgw c b n a, respectively.
together, they also influence each other aerodynamically. This means that two small tails
are less effective than a single large tail with the combined size of the two small tails. The
further you position your tails apart, the lower this interference effect becomes. Plac-
ing the vertical tails at the tips of the horizontal tail therefore minimizes the interference
between the two vertical tailplanes. Also in supersonic conditions, the interference be-
FT
tween the two tailplanes is reduced due to the formation of oblique shock waves from
In Figure 4.26 you can see a V-tail, an inverted Y-tail, and a twin-boom tail. The V-tail
provides stability and control in pitch and yaw through two tailplane surfaces. This re-
duces the required total tail area and thereby reduces the (friction) drag of the tail. As the
control surfaces on the V-tail are used to control pitch and yaw, they are typically revered
as ruddervators. When deployed to yaw the airplane to the left, the port-side ruddervator
deflects down, while the starboard-side ruddervator deflects up. This creates a yawing
RA
moment to the left but also a rolling moment to the right! This effect is not present on
the inverted Y-tail, where a yawing moment to the left is accompanied by a rolling mo-
ment to the left. The downside of the inverted Y tail is its close proximity to the ground.
This reduces the rotation angle that the airplane can make on the ground. The inverted
Y-tail in Figure 4.26 also protects the pusher propeller from striking the ground during
take-off and landing maneuvers.
D
Figure 4.26: Examples of less common tail configurations: the V-tail (left), the inverted-Y tail (center) and the
twin-boom tail (right). Photos by Huhu Uet c b, NASA/GA-ASI c p, and Wally Cacsabre c b, respectively.
The twin-boom tail configuration shown in Figure 4.26 is actually closely related to
the jet engine location of the airplane. In Figure 4.15, we saw another example of the
twin-boom configuration for a twin-prop airplane with a pusher propeller at the back of
the fuselage. The engine location prevents a single tail boom extending from the cen-
terline of the airplane. Therefore, two tail booms are attached to the wings, each having
a vertical tailplane at the end. The T-tail configuration of the twin-boom configuration
in Figure 4.26 prevents the hot exhaust of the jet engine from immersing the horizontal
tailplane.
FT
ble as winglets. This reduces the lift-induced drag resulting in a higher Oswald factor
(e). Also, per unit of useful volume, each of the tailless airplanes of Figure 4.27 have less
wetted area (S wet ), which reduces their zero-lift drag. Therefore, tailless configurations
like the blended-wing-body and the flying-V airplane are efficient for carrying large vol-
umes of passengers, cargo, and fuel. For hydrogen-fueled airplanes, which require at
least four times the volume for the same amount of energy, these configurations could
be advantageous over the traditional distinct wing-fuselage configuration.
RA
D
Figure 4.27: Examples of tailless airplanes: a flying-wing airplane (left), a blended-wing-body airplane (center)
and a flying-V airplane (right). Photos by Harveyellis c b a, S. Ramadier for Airbus, and J. van Oppen for TU Delft, respectively.
Tailless airplanes have many challenges when it comes to integrating all the func-
tionalities of a wing and a horizontal tailplane into a single wing. The generation of lift,
providing trim, providing pitch control, and providing longitudinal stability needs to be
performed with the same lifting surface. This requires compromises in the design of the
wing. As the stability and control functionalities cannot be directly mapped onto the de-
sign of distinct tail surfaces, the design process is also more complicated. Therefore, this
textbook does not provide methods to design tailless airplanes. The examples of Figure
4.27 show that while the design process is more complicated, tailless airplanes can be
successfully flown in a stable and controlled manner.
A SSIGNMENT 4.5
FT
RA
D
FT
In this chapter, you are going to design the fuselage of your airplane. The primary func-
tion of a fuselage is to house the payload. The volume of the payload is, therefore, a
driving requirement for the size of the fuselage. The type of payload also influences the
design of the fuselage. If the payload consists only of cargo, it can be very densely packed,
resulting in a relatively small fuselage compared to the cargo volume. Passengers, on the
other hand, require more volume than the volume that is occupied by their bodies. They
need sufficient space to feel comfortable. Furthermore, additional space is needed for
RA
lavatories, galleys, and carry-on luggage. Also, aisles are needed to reach the seats and
provide a path for emergency evacuation. To facilitate the loading and unloading of the
passengers and/or cargo, the fuselage also needs to provide access doors.
The fuselage does not only provide a volume to store the payload, it also provides
the payload protection from the environment in terms of wind, extreme temperatures
(both hot and cold), and noise. For airplanes that fly at high altitudes, it also ensures
that passengers are protected from the low ambient pressure. At this low pressure, the
air does not contain enough oxygen, which can lead to hypoxia.
D
The fuselage also provides space for the cockpit and an array of systems. For small
airplanes, this includes the avionics (aviation electronics), the engine or motor, a fire ex-
tinguisher, part of the fuel system, part of the flight control system, fuel or batteries, and
a battery. Larger airplanes can have an environmental control system (ECS), an auxil-
iary power unit (APU),, and landing-gear bays. Also, hydraulic, pneumatic, and electric
lines are running throughout the fuselage to connect all these systems. Furthermore,
a weather radar and various antennas are often installed to facilitate communications.
Finally, the fuselage also provides connection points for the wing, the landing gear, and
the tail.
To design the fuselage of your airplane, we are going to take a step-by-step approach.
In Section 5.1, we first discuss some design considerations regarding payload, emer-
gency evacuation, aerodynamics, ground clearance, structures, ground handling, and
pilot visibility. Prior to designing the fuselage, we first analyze the payload in terms of
its volume (Sec. 5.2). Then, we first show how to design the cross-section of the fuselage
67
68 5. F USELAGE D ESIGN
(Sec. ??). We subsequently design the fuselage in the top view (Sec. 5.4) and in the side
view (Sec. 5.5).
The design of the fuselage is the starting point of a three-view design of the complete
airplane. In Chapters 8 and 9, you will dimension the wing, the propulsion system, the
landing gear, and the tail. The fuselage acts as the connecting element between all of
these components and is, therefore, our starting point for the three-view drawing of the
airplane.
5.1.1. PAYLOAD
FT
which considerations are at play in the mind of the airplane designer when designing
the fuselage.
As stated above, the fuselage size is primarily determined by the payload requirements.
Figure 5.1 shows the unloading of fuselage barrels from the fuselage of a cargo airplane.
The cargo plane has clearly been designed to fit large components inside its fuselage. As
RA
these components do not require oxygen, only the cockpit of the cargo airplane needs to
be pressurized. The size of the fuselage is determined by the size or volume of the cargo
that you want to transport. Most passenger airplanes also have storage for goods. This
can be luggage or cargo. In either case, it is the volume of these goods that determines
how much volume is needed in the fuselage to store them.
When your airplane transports passengers, you need to think about the level of com-
fort you wish to include in your design. More space per passenger means a more com-
fortable flying experience. However, it also means a larger, heavier, draggier fuselage. In
other words, more comfort for the passengers means more energy is required to trans-
D
port them. This goes against the objective of minimum fuel burn, minimum operating
cost, or minimum climate impact. You should therefore make a careful trade-off be-
tween your design objective and the comfort level you want to provide.
FT
Figure 5.1: The payload requirements for a cargo airplane can be completely different from a passenger air-
plane leading to different design solutions. Photo by David c b a.
out.
• If you have a pair of emergency exits, position them across from each other, such
RA
that they can be connected via a cross aisle.
• If you have over-the-wing exits, make sure they are positioned at the location of
the wing. If you do not know the wing location, position them at a longitudinal
location where you expect the wing root to be positioned. Update the longitudinal
location of the exit(s) once the position of the wing root is known.
• One or more emergency exits are likely to be used for boarding of passengers and/or
crew. Other emergency exits might be used for servicing the airplane (see Section
5.1.7). Therefore, you may want to increase the size of the emergency exits to im-
prove these processes.
D
For CS-23 airplanes, the minimum size of the emergency exit should be 48-by-66 cm
(width × height). This corresponds to a so-called Type IV exit according to CS-25 regu-
lations. This minimum size is not very comfortable for boarding an airplane. Therefore,
the boarding door is usually larger than the minimum size. In Figure 5.2, you can see an
example of a CS-23 airplane with its main exit open. According to the regulations, this
airplane should have at least two Type IV exits. However, the designers have chosen to
have one Type IV exit and one Type II exit. The type II exit is larger and is located on the
left-hand side of the fuselage. The Type IV exit is an overwing exit and is located on the
opposite side of the fuselage.
CS-25 specifies various types of exits. In Example 3.15, we showed how many pas-
sengers are allocated per pair of emergency exits of a given type. Most common are exits
which are located at cabin floor level and overwing exits. Their dimensions are shown in
Figure 5.3. For the overwing exits, a maximum step height is defined from the inside to
Figure 5.2: Example of a Type II exit on a business jet. Photo: Anna Zvereva c b a
the door (h 1 ) and from the door onto the wing (h 2 ). Table 5.1 shows the dimensions of
R
FT
the most common door types. An exhaustive list of emergency exit types and associated
dimensions can be found in CS-25.807.
H
R H
RA
B h1 h2
floor level wing level
B
Table 5.1: Dimensions of emergency exit types as defined in CS-25.807. Dimensions are defined in Figure 5.3.
1 inch = 2.54 cm.
5.1.3. A ERODYNAMICS
The fuselage is generally not intended to generate any lift force. However, when sub-
jected to an angle of attack or angle of sideslip, it generates resulting forces. These forces
impact the longitudinal, lateral, and directional stability of the airplane. In addition, the
fuselage has a large contribution to the drag of the airplane, mainly due to its large wet-
ted area. How the fuselage shape relates to the drag is further expanded below.
In Figure 5.4, we show how the total drag of a fuselage-like body scales with the in-
verse of its so-called fineness ratio. The fineness ratio is the ratio between the length of
the body (l ) and the diameter of the body (d ). We note that the friction drag reduces
with the inverse of the fineness ratio and has a minimum when the body is a sphere.
This makes sense because a sphere has the smallest ratio of surface area to internal vol-
ume. However, a sphere is not an aerodynamically shaped body. Therefore, it has a high
pressure drag. If we stretch the sphere into a tube, the pressure drag reduces. Simultane-
ously, the friction drag starts to increase because the wetted area increases for the same
internal volume. A minimum in total drag is found when the fineness ratio is l /d ≈ 3.
Typical subsonic
jet transports
[0.08 - 0.12]
(l/d: 8 - 12)
CD0
d
FT
Typical single
engine
[0.13 - 0.2]
(l/d: 5 - 8)
l
Resultant zero-lift lift
drag coefficient:
CD0
Figure 5.4: The zero-lift drag coefficient of the fuselage is dependent on the fineness ratio, l /d . After Corke [2]
.
D
Superimposed on the curves of Figure 5.4 are colored bands that indicate typical fine-
ness ratios of actual fuselage shapes. These all have a considerably higher fineness ratio
than the one for minimal drag. For all airplanes, this has to do with the fact that a longer
fuselage creates a longer arm from the center of gravity to the tailplanes. A longer arm
means that a smaller horizontal and vertical tail is required to provide adequate stability
and control (see also Chapter 9). For pressurized fuselages, the circumferential (hoop)
and axial loads in the skin that result from pressurization scale linearly with the diame-
ter. Therefore, a higher fineness ratio results in a lower diameter and, therefore, a thinner
and lighter fuselage skin. Supersonic airplanes have a very high fineness ratio because
the pressure drag stemming from conical shock waves correlates quadratically to the
fineness ratio of the nose and tail of the fuselage [14]. A higher fineness ratio therefore
reduces the supersonic drag of these airplanes.
horizontal plane
84 mm horizo
n tal line
eye trajectory
eye location
of sigh
t
RA
downward view angle
If we concentrate on the symmetry plane of the airplane, we show how the require-
D
ments on the downward view angle and upward view angle influence the shape of the
nose and the size of the cockpit window. For simplicity, we assume that the pilot’s eye is
located on the symmetry plane. We also have the cross-section of the nose of the fuse-
lage, along with the location and dimensions of the cockpit windows (see Figure 5.6).
Then, we can draw a straight line from the pilot’s eye through the upper left corner of the
window and measure the upward view angle. Similarly, we can draw a straight line from
the pilot’s eye through the lower left corner of the window and find the downward view
angle. The latter angle is referred to as the over-nose angle, because the nose shape can
influence how large the downward view angle can be. The angle between the window
pane and the downward view angle is called the grazing angle. To prevent distortion,
this angle is recommended to stay below 30 degrees, [9] although values below 20 de-
grees are also present in existing airplanes as can be seen in Figure 4.14.
The requirements for pilot visibility can come from the regulations or from the cus-
tomer. For example, EASA specifies the requirements in Ref. [3]. You can imagine that
cockpit windows
upward view angle
overnose angle
grazing angle
FT
the pilot of a combat airplane requires a much larger field of view than the pilot of a
small general aviation airplane. How the requirements affect the shape of the nose and
the size of the cockpit windows depends on the visibility requirements and the attitude
variation of the airplane during flight. An airplane that lands at a very high pitch atti-
RA
tude might require a high over-nose angle such that the pilot can see part of the runway
during landing in foggy weather. The attitude of the airplane during flight is determined
by the design of its wing, its speed, its flight path angle, its weight, and the deployment
of trailing-edge flaps. In the first iteration of the airplane design, we cannot yet assess
all of these aspects. However, after you have designed the wing and its high-lift devices
(Chapter 11), you can derive what the over-nose angle needs to be.
The design of the fuselage tail cone is influenced by the attitude of the airplane during
landing. To prevent the tail of the fuselage from striking the ground during take-off or
landing, the tail cone often has an upsweep. Figure 5.7 shows how the upsweep angle is
defined for two different tail cone geometries. As you can see, a horizontal line is defined
in the middle of the fuselage. The tail cone upsweep angle is defined with respect to this
horizontal line. The upsweep allows for an increase in clearance between the ground
and the fuselage when the airplane has an increased pitch attitude while the wheel of
the main landing gear (MLG) is on the ground. The required upsweep angle to have suf-
ficient ground clearance is, therefore, dependent on the same design parameters and
operational parameters that determine the over-nose angle. In addition, the upsweep is
dependent on the length of the main landing gear struts, the tire diameter, and the char-
acteristics of the shock absorber. Therefore, in the first iteration of the design, you should
choose a tail cone upsweep angle and re-evaluate this choice after having designed the
landing gear (Chapter 9).
tc tc
MLG wheel MLG wheel
Figure 5.7: Example of a tail-cone upsweep angle on a general aviation airplane (left) and a transport airplane
(right). MLG = main landing gear.
For an airplane with a tail wheel (i.e. conventional landing gear) instead of a nose
wheel, an upsweep of the tail cone is not needed to ensure ground clearance. However,
as the main landing gear is positioned more forward for an airplane with a conventional
landing gear configuration, tail cone upsweep might still be advantageous to allow for an
increase in maximum pitch attitude. An increase in pitch attitude can result in a lower
FT
stall speed in case of a tail-wheel landing (see also Section 4.3.1 on page 58). On the other
hand, many airplanes with a conventional landing gear do not have upsweep in the tail
cone, such as the airplane shown on the left side in Figure 4.21 on page 59. Therefore, you
can decide whether or not to include an upsweep in your tail cone if you are designing
an airplane with a conventional landing gear configuration.
5.1.6. S TRUCTURES
The function of the structure of the fuselage is to protect the payload from wind, low
RA
pressure, and impact during a crash. It also provides attachment points for the wing, tail,
and possibly the landing gear. It must be strong enough to carry all the loads, be durable,
and be lightweight. The weight of the fuselage itself and the payload inside it produces
a distributed load in the structure. The wing introduces a more concentrated load at the
intersection between the wing and the fuselage. Also, the tail surfaces introduce loads.
If the fuselage is pressurized, this causes additional loads. The fuselage structure should
therefore be sized to withstand various combinations of these loads.
A fuselage structure for an airplane can be characterized as a stiffened, thin-walled
D
structure. The skin panels transfer shear stresses and normal stresses. Connected to the
skin panels are stringers that run in the longitudinal direction, also known as longerons.
They carry part of the normal loads and prevent the skin from local buckling under com-
pression load. Frames are added perpendicular to the centerline to prevent global buck-
ling of the fuselage under bending or torsion loads. Frames are also used to introduce
concentrated loads into the thin-walled structure. Some of the frames are strengthened
with plates to carry loads from the wing or landing gear. These frames are called bulk-
heads. A separate floor structure is present on larger airplanes to transfer the payload
loads to the frames. This floor structure typically consists of lateral floor beams and lon-
gitudinal seat tracks. Figure 5.8 shows the main structural elements in a cross-section of
a pressurized fuselage.
For airplanes with pressurized cabins, a circular cross-section is preferred because
it minimizes the structural weight that is required to sustain the pressurization loads.
Any non-circular shell structure that is pressurized needs to be strengthened with ad-
frame
A-A
skin
twall
stringer
r1, inner
A
A
floor beam
seat track
strut
frame
r2, inner
floor
beam
cargo floor beam
Figure 5.8: cross-section of a double-bubble fuselage with annotation of structural elements. After Niu [5].
FT
ditional structural elements to prevent large deformations. In Figure 5.8, a so-called
double-bubble structure is shown where the floor beam connects the upper and lower
fuselage shells that each have a circular shape. When pressurized, the floor beam is
loaded in tension and prevents the shape from deforming. A double-bubble is some-
times preferred over a pure circular cross-section as it can result in a smaller perimeter
of the fuselage cross-section and a smaller radius of the upper and lower arcs. This re-
sults in lower friction drag and lower pressure-induced loads in the skin, respectively.
RA
In the design of the cross-section, you need to account for appropriate distances for
the structural depth of the side wall. We propose the following design rules:
where d f, outer is the outer diameter of the fuselage, and d f, inner is the inner diameter
of the fuselage. For a double-bubble fuselage (referring to Figure 5.8), you can use an
equivalent inner diameter:
d f, inner = r 1, inner + r 2, inner (5.4)
The depth of the passenger floor also depends on various factors, including the fuselage
diameter, the passenger load, and how the floor is supported. We propose to use a struc-
tural depth of the passenger floor between 100 mm and 300 mm, depending on the size
of the airplane.
A pressurized fuselage has a bulkhead ahead of the flight deck as well as at the end of
the cabin. The latter bulkhead is referred to as the aft pressure bulkhead. This bulkhead
is often spherical in shape, although some airplanes have a flat, stiffened aft pressure
bulkhead.
FEET
0 5 10 15 20
0 2 4 6 8
METERS
PBB
FTAC
TOW
G
PU
CAT
LD CL
AS
ULD
RA
FUEL
ULD
LD CL
WV CB
D
CAT
BULK
LV
Figure 5.9: Layout of servicing vehicles at the airport stand for a large passenger airplane [1].
We shortly go over all the ground equipment that is connected to the airplane of Fig-
ure 5.9 starting at the top left quarter and making our way clockwise around the airplane.
First of all, we see an air-conditioning (AC) cart, which provides fresh air to the cabin
when the engines and APU are shut off. Then, we see the passenger boarding bridge
(PBB), which can also be replaced by stairs at the front and/or the aft passenger door.
Subsequently, we have the tow tractor (TOW), which can provide a push-back to the
airplane from the gate. Next to it is the ground-power unit (GPU), which provides elec-
tricity to the airplane when engines and APU are off. Next to the GPU, we have the first
catering truck. This truck has a lifted van body that connects to one of the emergency
exits. From this van, the front galley is supplied. Next to it is the lower-deck cargo lift (LD
CL), which can lift a Unit Load Device (ULD) to the level of the lower deck. A ULD is a
cargo container with standardized dimensions. We also see a dolly train that brings the
ULDs to the airplane. The fuel truck interfaces with the wing, but behind the wing, we
see another cargo lift and a conveyor belt (CB) for bulk cargo, i.e. cargo that is not stored
in a ULD. Next to it, another catering truck is present that services the aft galley. At the
tail, we find the lavatory vehicle (LV) that dispenses the wastewater. Finally, we have the
potable water vehicle (WV) that provides fresh water ahead of every flight. What is not
shown is a cleaning truck, which can be connected at one of the exits to allow a cleaning
crew and their equipment to enter the cabin.
FT
In the conceptual design of the fuselage, we include the location of all emergency
exits. Positioning of these exits can therefore be influenced by the accessibility require-
ments that are imposed by the ground servicing vehicles. Producing a drawing like the
example in Figure 5.9 can help in deciding where the various emergency exits, as well as
cargo doors need to be positioned. In Section 5.5, we show how the emergency exits can
be positioned.
RA
5.2. P RELIMINARY PAYLOAD V OLUME A NALYSIS
The total payload of an airplane consists of goods and/or passengers. The first question
that we ask ourselves is: where is the payload stored in the fuselage? If you have passen-
gers, luggage, and cargo, you have to decide to have that all on the same floor, to store
(part of) the luggage and cargo below the floor in a separate cargo compartment, or to
use separate cargo holds ahead and/or behind the passenger cabin. It is instructive to
look at how cargo and luggage are stored on your reference airplanes. This could help
D
you in deciding what works best for your design. To help you decide, it is instructive to
have an estimation of the total cargo volume in relation to the total passenger volume.
Example 5.1
Consider the airplane of Example 6.7 of the four-person twin-prop airplane. We have a
payload mass of 400 kg for four passengers. We make the following crude assessment of
the required volume for luggage and passengers:
Here we have implicitly assumed that each passenger weighs 80 kg and requires 0.7 m3
of volume. So we conclude that 85% of the payload volume is taken by the passengers,
and we only need 15% of the payload volume for the luggage. If we examine how the
reference airplanes of Example 6.7 provide storage for luggage, we can see that cargo
volumes are made ahead or behind the passenger cabin. For now, we decide to store the
luggage behind the passenger cabin.
In the previous example, it has been decided to position the luggage behind the passen-
gers. Therefore, the dimensions of the luggage do not play a role when defining the cabin
cross-section. However, in the next example, we show that for larger airplanes, it makes
sense to store the luggage and cargo below the passenger floor.
Example 5.2
Consider the large transport jet for which a payload-range diagram has been made in
Example 6.10. The design payload is 19 tonnes, corresponding to 180 passengers in
economy class. We assume that each passenger weighs 80 kg, has 5 kg of carry-on lug-
required:
payload
passengers
FT
gage, and 20 kg of cargo-hold luggage. In addition, the maximal structural payload is 23
tonnes. This implies that an additional 4000 kg of cargo can be added, for which vol-
ume should be available. We make the following crude assessment of the volume that is
m (t)
23
14
ρ kg/m3
-
110
V (m3 )
-
130
RA
carry-on luggage 0.9 150 6.0
cargo-hold luggage 3.6 170 21
cargo 4.0 160 25
This crude assessment indicates that we need approximately 46 m3 for cargo and lug-
gage, against 136 m3 for passengers with their carry-on luggage. If we decide to position
the cargo holds ahead and/or behind the passengers, we would end up with a fuselage
that is roughly 33% longer compared to a fuselage with only passengers on the main
D
deck. Therefore, we decide to add a cargo compartment below the passenger floor. This
is also in line with reference airplanes with a similar mission.
The previous two examples have shown how a crude preliminary estimation of the re-
quired volume can be deduced from the top-level requirements. To that extent, we have
used assumed values for the density of cargo and luggage. These values are only indica-
tive but should be confirmed later on. In practice you might be able to find more explicit
requirements on the amount of volume that is desired. For example, you could have the
requirement that every passenger should be able to take a trolley suitcase into the cabin
and store it in the overhead bins. Based on such a requirement, you can directly compute
how much volume is required in the overhead bins.
Luggage and cargo can be stored in an airplane in two ways: in bulk or in standard-
ized containers or pallets. If you choose to load luggage in bulk, you can effectively
use the available cargo-hold volume. In other words, the available cargo-hold volume
Figure 5.10: Examples of cargo holds below the passenger floor for bulk cargo (left) and unit load devices
(right). Photos by User:Dtom c b a and Asiir c b a, respectively.
is equal to the required cargo-hold volume based on the sum of luggage and cargo vol-
FT
ume. However, loading bulk luggage is a laborious process, which is can be quite strain-
ing for the ground-handling crew due to the constrained space they need to work in.
Therefore, you can also choose to load luggage and cargo in standardized containers or
pallets. These containers are termed unit load devices or ULDs. A luggage container can
be filled with luggage inside the airport terminal and can be loaded onto the airplane
relatively quickly. The downside of using ULDs for luggage is their added weight. This
results in a higher fuel consumption compared to loading luggage and cargo in bulk.
Small airplanes up to single-aisle transport airplanes typically have bulk loading, while
RA
all twin-aisle transports have containers. Some single-aisle airplanes are offered with
either option. Then, it is up to the airline to decide whether they prefer bulk loading
or containers. An overview of the weights and dimensions of ULDs used in civil aviation
can be found on wikipedia.org/ULD. Figure 5.10 shows how bulk cargo and ULDs can be
stored below deck. Note the use of nets for bulk cargo to prevent it from shifting during
flight.
A SSIGNMENT 5.1
D
In this assignment, you will decide on the distribution of your payload within the
fuselage.
a State how your reference airplanes store cargo and luggage.
b State the design payload mass and maximum payload mass (if applicable).
c Assume a mass per passenger and compute the total passenger mass ex-
cluding luggage.
d Assume the average carry-on luggage mass per passenger and compute the
total carry-on luggage mass.
e Assume the average cargo-hold luggage mass per passenger and compute
the total cargo-hold luggage mass.
f Compute the remaining cargo mass for the design payload mass.
g Compute the remaining cargo mass for the maximum structural payload
mass. Assume the same number of passengers and luggage mass as above.
FT
minimizes the friction drag of the fuselage as well as the weight.
To understand what the cross-section of the payload volume looks like, we first need
to establish the dimensions of our payload volume. For a passenger airplane, we need
to decide how many people are sitting abreast in the airplane. For small airplanes, this
choice can be straightforward: 2-abreast is the obvious choice. However, for larger air-
planes, it might be less obvious. How many seats abreast do you choose for a 100-
passenger airplane? Or for a 300-passenger airplane? The following formula can provide
guidance to make that decision:
RA
p
n SA = 0.45 n pax (5.5)
where n SA is the number of seats abreast and n pax is the number of passengers in the
airplane. Note that you need to round the result of (5.5) to an integer! You can choose
to round up, which results in a wider but shorter fuselage. Or you can choose to round
down, which results in a narrower but longer fuselage. For airplanes where passengers
do not have direct access to an emergency exit, an aisle is required. The minimum di-
D
mensions of the aisle are specified in CS/FAR 23/25 and are printed in Table 5.2. You can
see that depending on the number of passengers, the requirements change. An airplane
with more passengers requires a wider aisle. We use h as a measure for the vertical dis-
tance from the floor. The aisle width below the armrest height (i.e. h < 64 cm) may be
narrower than above the armrest (h ≥ 64 cm). As a designer, you have the freedom to
choose a larger aisle width to improve the boarding process as well as the level of com-
fort experienced by the passengers. You need to trade this against the resulting increase
in weight and drag.
If we know how many seats are positioned abreast and the width of the aisle, we can
start to construct a set of rectangular boxes that bound the cross-sectional area of the
passenger volume. This is notionally shown in Figure 5.11, where various dimensions
are defined, such as the width of the seat cushion (w seat ), the shoulder height (h shoulder )
and the armrest width (w armrest ). The value of dimensions are neither requirements nor
assumptions. You have to choose a value for each of them based on available data on
Table 5.2: Required aisle dimensions according to CS/FAR 23.815 and CS/FAR 25.815. Some conditions apply.
seats as well as common sense. Table 5.3 shows an example of seat dimensions and
pitches for three different passenger airplanes.
Table 5.3: Example of seat dimensions and pitches for three different airplanes belonging to the same airline.
Data from seatguru.com, retrieved on December 12, 2022. Using common airline fair basis codes: J - Business
Class, W - Premium Economy, Y - Economy/Coach.
5
W
81 81 76
43 43 43
5
Y
5
FT J W
84 84
43 43
5 5
Y
76
43
5
J
51
8
W
107 89 79
44 44
5
Having chosen a value for each dimension, we define the width of the cabin as fol-
lows:
Y
w aisle . The third box ensures that the feet of the passengers stay within the inner wall. It
has a width of:
w floor = w cabin − 2 (w armrest + w clearance ) (5.7)
The final box ensures that there is enough head space for the passengers. It has its corner
points at a height of h headroom above the floor and a width of:
Now that we have defined the bounding box for the passengers on the floor, it is time
to turn our attention to the volume below the floor. If there is no under-floor volume
required for cargo or luggage, then this step can be skipped. If you choose to have bulk
cargo under the floor, it is advised to select a minimum distance between the passen-
ger floor and the cargo floor such that the ground-handling crew has enough space to
haisle
wheadroom
wclearance hheadroom
waisle
wcabin
hshoulder
harmrest
wseat
FT
wfloor
move and distribute the luggage. If you choose to use a ULD for luggage and cargo, the
floor
RA
dimensions of the ULD should be included in the cross-section.
In the following examples, we will show how to construct a cross-section for various
airplane types. To structure this process, we use the following sequence:
Example 5.3
In this example, we compute the cross-section of the four-seat airplane of Example 5.1.
Prior to performing the design steps of the cross-section, we perform a preliminary ex-
perimental analysis on how we organize the cabin and what dimensions we need to ac-
1 in case of a double-bubble fuselage, choose two circles
commodate passengers comfortably within the cross-section. We decide to let two pas-
sengers sit side-by-side without the presence of an aisle. Instead, we decide to have 5
cm of clearance between the two passengers. We assume that the people are sitting with
their knees slightly bent, to reduce the necessary height of the cabin. We chose a seat
that has a width of 45 cm and an armrest for each person measuring 6 cm. Finally, we
assume a clearance of 2 cm between the armrest and the inner cabin wall. We choose the
armrest height to be 50 cm above the floor, a shoulder height of 100 cm, and a headroom
height of 135 cm.
Step 1 With the dimensions mentioned above and using w aisle = 5 cm, we can use
(5.6) to find that w cabin = 123 cm. The floor width is computed with (5.7) and
is w floor = 107 cm. The headroom width is computed with (5.8) and amounts to
w headroom = 62 cm. The armrest height, shoulder height, and headroom height
FT
have dimensions h armrest = 59 cm, h shoulder = 100 cm, and h headroom = 135 cm,
respectively. The dimensions are drawn in the illustration below.
Step 2 Based on the dimensions of step 1, we draw the two-dimensional space alloca-
tion using rectangular blocks (see below).
Step 3 We draw the floor as a single line and decide to draw the floor structure once the
RA
frame thickness has been determined.
Step 4 We skip this step because there is no allocation of cargo in the cross-section (see
Example 5.1).
Step 5 We identify the corner points of the rectangular blocks with dots (see below).
D
Step 6 We draw the inner perimeter of the fuselage slightly outside of the dots. As this
fuselage is not going to be pressurized, the inner perimeter is not perfectly cir-
cular. However, we do try to minimize the perimeter to minimize the wetted
area of the fuselage as well as the structural mass.
Step 7 Based on (5.1), we have a wall thickness of t wall = 40 mm. We draw the outer
perimeter with an offset of 40mm from the inner perimeter (see below). The
resulting fuselage cross-section has the following dimensions: w fus = 131 cm
and h fus = 169 cm.
Step 8 We skip this step because there is no allocation of cargo in the cross-section (see
Example 5.1).
Step 1 Step 2
headroom
ℎshoulder
cabin
ℎheadroom
ℎarmrest
�loor
FT
In the previous example we inserted for w aisle the desired width between the two pas-
wfus
RA
sengers. Also, note that the dimensions that we set in Step 1 were derived from an exper-
imental analysis where the body dimensions of a random person were used to define the
various heights and widths of the cabin. In practice, you might have to do some research
on the dimensions of the passengers that you wish to transport in your airplane. An an-
thropometric database such as DINED2 could be employed to perform such research.
Alternatively, you can research the dimensions of seats that are used in, for example,
your reference airplanes. If dimensional data of these seats can be found, they can be
used to determine the required dimensions in Step 1. In the following example, we show
D
In this example, we draw the cross-section of a passenger transport airplane. The fol-
lowing requirements have been derived. This airplane needs to seat 130 passengers in a
single class. We have a seat width of 46 cm, an armrest width of 5 cm, a stand-up height
in the aisle of 190 cm, and an aisle width of 45 cm, measured between two seat cushions.
To store luggage and cargo, an LD3-45 container is to be used below the passenger floor.
This container has a height of 114 cm (45 inches), a roof width of 242 cm, and a base
width of 156 cm with 45◦ walls between the base and either side.
Step 1 Using (5.5), we decide that we use a five-abreast cabin, with a single aisle. This
2 For more information on anthropometric data, visit dined.io.tudelft.nl.
implies that the aisle is not going to be in the center of the cross-section. Note
that the aisle width is defined between the two seat cushions, while (5.6) defines
the aisle width between the two armrests. If we subtract the armrest width, we
would have an aisle width between the armrests of 35cm. However, we have to
use w aisle = 38 cm according to the data of Table 5.2 in order to comply with CS-
25. Now, we employ (5.6), (5.7), and (5.8), to find w cabin = 307 cm, w floor = 293
cm, and w headroom = 247 cm. For the height of the armrest, shoulder and head-
room, we have no explicit requirements. Therefore, we decide to add 5cm to the
height values of Example 5.3 such that the seat cushion is raised. The armrest
height, shoulder height, and headroom height have dimensions h armrest = 64
cm, h shoulder = 105 cm, and h headroom = 140 cm, respectively. A representation
of the passengers and a standing person in the aisle is shown in the graphic be-
low.
Step 2 Based on the dimensions of the previous step, we draw four rectangular blocks.
Three blocks are positioned on top of each other. The block representing the
FT
aisle is partially overlapping with the three blocks that cover the seats (see be-
low). A symmetry line is added to show where the symmetry plane of the cross-
section is located.
Step 3 For the passenger floor, we choose a floor depth of 20 cm. This is added below
the passengers.
Step 4 The dimensions of an LD3-45 container are found online3 and its cross-sectional
geometry is added below the passenger floor.
RA
Step 5 The corner points of the rectangular blocks above the floor and the cargo con-
tainer below the floor are marked with dots.
Step 6 It is decided to design a double-bubble cross-section with an upper arc en-
closing the passenger compartment and a lower enclosing the cargo compart-
ment. By means of trial and error4 , the upper and lower arcs are constructed
(see below). We measure the radius of the upper arc and find 158 cm. This re-
sults in d f, inner = 316 cm. Using the same approach for the lower arc, we find
d f, inner = 294 cm.
D
Step 7 Using (5.3) and the largest of the two diameters, we find an outer diameter,
d f, outer = 330 cm. We have a wall thickness of 7 cm. We choose to keep the wall
thickness constant over the entire perimeter of the fuselage, including the cargo
hold. We draw the outer fuselage arcs by increasing the radius of each arc by 7
cm. The resulting fuselage width is w fus = 330 cm, and the height is h fus = 384
cm, which we can measure from the drawing. We can also measure the height
of the lower deck (ld) and main deck (md), measured from the lowest part of the
cross-section: h ld = 45 cm, h md = 1.58 m.
Step 8 Respecting the headroom height, the aisle width, and the aisle height, we draw
a notional perimeter of the overhead bins. Note that the overhead bins are not
the same on either side of the aisle.
3 Information on ULDs can be found at wikipedia.org
4 A process of finding the smallest circle is shown in this video
wfus
FT hfus
As you can see from the examples, the design of the fuselage cross-section requires you
to perform calculations as well as to draw. To draw accurately, we recommend using grid
paper in combination with a ruler and a geometric compass. Finding the origin of the
hld
hmd
RA
inner circle can be done using trial-and-error with your geometric compass. You choose
one of the corner points and select a location for the origin of your circular arc. When
you then draw the circular arc, you can evaluate whether all corner points fall within
the arc and how much clearance is left between the arc and the the other corner points.
Once you have found the smallest circular arc that encloses all the corner points, you
have finished. 5
A SSIGNMENT 5.2
D
tors. First of all, there is the explicit requirement that the fuselage should have sufficient
volume to store the payload. The level of comfort to the passengers can be driving the
length of the cabin through a required seat pitch. For large airplanes, we might also have
two or three different classes, each with their individually defined seat pitch. In addi-
tion, we have the requirement that we need to accommodate the pilot(s) in a (separate)
cockpit. Finally, we have the implicit requirement that a tail should be mounted to the
fuselage and that the fuselage should be connected to the wing.
In addition, we need a design objective statement such that we can justify the choices
we make when we design the top view of the fuselage. We propose to use the minimiza-
tion of drag as a design objective. With a given cross-section, minimization of the (fric-
tion) drag implies we need to minimize the wetted area (S wet ). However, we also realize
that an aerodynamically shaped nose and tail are needed to have low pressure drag. In
addition, the size of the horizontal and vertical tails are also dependent on their respec-
tive distances to the center of gravity of the airplane. In other words, a short fuselage
might result in large tail surfaces, which themselves would have a large wetted area. So,
we refine the design objective as follows: to minimize the drag of the fuselage while hav-
FT
ing an acceptable tail arm to the tail surfaces.
Before we present you with a possible design sequence to design the fuselage in the
top view, let us first present some general design features of the fuselage that we can
recognize in the top view of most tubular fuselage geometries. We have visualized the
top view of a single-prop, small passenger airplane in Figure 5.12. You can see that the
fuselage outer mold line consists of three components that are attached to each other:
the nose cone, the cylindrical part, and the tail cone. The addition of these three compo-
RA
nents forms the complete fuselage length. Even though we use the word ‘cone’ in relation
to the nose and tail geometry, the three-dimensional shapes are not necessarily conical.
If look at the inside of the fuselage, we see that the various internal components are dis-
tributed a bit differently. For example, within the nose cone we have the nose as well as
part of the cockpit. The perimeter of the nose is indicated with a dashed line because
the nose consists of an engine cowling. The design of the engine cowling is not part of
the fuselage design and is discussed in Chapter 8. However, it does still contribute to the
fuselage length. Since this is a small airplane, the cockpit is deemed part of the cabin.
The cabin is located mostly in the cylindrical part of the fuselage but also extends into
D
the nose cone and tail cone. In this example, we have chosen to give the tail cone the
shape of a tadpole tail, where we have used a simplified trapezoid shape as a drawing
guide (also shown in dashed lines in Figure 5.12). You can see that behind the cabin,
there is still ample volume for storing luggage or cargo.
If we look at the definitions for a large passenger airplane (Figure 5.13), we can dis-
tinguish the same three components that make up the outer mold line of the fuselage:
the nose cone, the cylindrical part, and the tail cone. Also, the nose, cabin, and tail are
present. However, the cockpit is no longer part of the cabin but is a separate component
inside the nose and the nose cone. The cabin starts behind the cockpit. For the large
passenger airplane, we have given the tail cone a more conical shape. When drawing
the tail, we have made a smooth transition from the cylindrical part to the tail cone to
avoid local flow separation as well as to provide ample cabin width at the start of the tail
cone. The shape of the nose cone is constructed as an ellipse with a blunt nose, which is
wfus
Figure 5.12: Top view of a fuselage design with distinct sections indicated. The dashed lines are for reference
only.
wfus
nose
FT
fuselage length, lfus
cabin tail
RA
nose cone cylindrical part tail cone
The length of the cabin is dominated by the size of the payload. When only pas-
sengers occupy the cabin, you can use the seat pitch to estimate how much length is
D
needed to accommodate all the passenger seats. Example values for seat pitch can be
found in Table 5.3. For small airplanes, the seat pitch might be enough to accommodate
the length of the cabin. For large passenger airplanes, the addition of cross-aisles (for
evacuation), crew seats, galleys and lavatories also needs to be added. Each of these has
standardized dimensions that can be easily found online. To determine the length of the
cabin, you can position all required cabin items in the top view of the cabin and mea-
sure the resulting length. To perform this exercise, you need to decide on the number of
galley carts you need, the number of lavatories to install, and how many cross-aisles are
required. If you decide to also store cargo on the same deck as the passengers, the length
of the passenger cabin might also increase. In other words, determining the length of the
cabin requires the generation of a floor plan to accommodate the various items within
the cabin. To perform this exercise can be quite labor intensive for large passenger air-
planes.
In lieu of this laborious effort, we propose a statistical method to determine the
length of the cabin for large passenger airplanes without making a floor plan. We use
a statistical factor k cabin , which represents the length in meters per available seat row in
the cabin. The cabin length can then be computed as:
n pax
l cabin = k cabin (5.9)
n SA
with ½
1.08 [m] for short/medium-range airplanes;
k cabin =
1.17 [m] for long-range airplanes.
The value of k cabin is subject to statistical validation.
Now, that we have determined the length of the cabin, we compute the length of
the nose, nose cone, tail cone, and tail, respectively. To support this process, we have
proposed values for various dimensions and fineness ratios, which are grouped in Table
5.4.
The length of the nose is a function of human dimensions, where we assume that
FT
there are two people sitting side-by-side in the cockpit. For propeller airplanes, the
length of the nose should be chosen between 2.0 and 4.0 meters, whereas smaller air-
planes typically have a shorter nose. For jet airplanes, which generally fly faster than
propeller airplanes, the nose is typically a bit longer. However, the maximum length
from the tip of the nose to the back of the cockpit should still be chosen between 3.0
and 4.5 meters. The nose cone can either be longer than the nose (e.g. for small air-
planes) or shorter (e.g. for large airplanes). Typical values for the nose-cone fineness
ratio (l nc /d fus ) are listed in Table 5.4. Here we use the equivalent fuselage diameter, d fus ,
RA
which we define as:
w fus + h fus
d fus = (5.10)
2
You can see that for jet airplanes, the nose cones have a higher suggested fineness-ratio
range than for propeller airplanes. This is because jet airplanes typically have a high
subsonic cruise Mach number, which can cause additional drag if the nose is too blunt.
Table 5.4: Proposed range of values for dimensional parameters of the nose and tail cones
D
The length of the tail cone can be computed by choosing an adequate tail cone fine-
ness ratio (l tc /d fus ), for which suggested ranges are given in Table 5.4. Even though small
propeller airplanes might not require a tail cone with a high finesse ratio from a drag
point of view, a high fineness ratio results in a long tail arm, which reduces the neces-
sary size of the horizontal and vertical tailplanes (see Chapter 9). Once you have chosen
a value of the tail cone fineness ratio, you must decide how much of the cabin is still
located inside the tail cone. This is reflected in the tail-to-tail-cone ratio, l t /l tc . For ex-
ample, when l t /l tc = 0.7, it means that the first 30% of the tail cone is used as part of the
cabin. We suggest to choose a tail-to-tail-cone ratio between 0.5 and 1.0. With a selected
tail fineness ratio and tail-to-tail-cone ratio, you can compute the length of the tail cone
and tail.
In summary, we propose the following design sequence for the fuselage perimeter in
the top view:
In the following two examples, we demonstrate how this process works for the four-seat
airplane and the 130-seat airplane.
Example 5.5
FT
In this example, we design the fuselage of a 4-seat airplane in top view. We follow the
design sequence that is shown above as well as data from Example 5.3.
Step 1 The width of the fuselage measures 131 cm. We draw this dimension together
with the symmetry line in the drawing below.
RA
Step 1 wfus
pitch of 85 cm, which becomes the length of the cabin. Note that the cockpit is
Step 2 located in the nose and that we do not count the two front seats as part of the
0 0.5 1.0 [m]
cabin. We draw the dimensions in the drawing.
lcabin
ln
Step 2
Step 3
n
l
Step 3 We choose a nose
l length of 3.0 meters, based on Table 5.4. Secondly, we choose
a nose finenessncratio of 1.4. This results in a nose length of 1.83 m. We add these
two lengths to the drawing.
Step 3 lt
Step 4 F OR R EVIEW O NLY. D O N OT D ISTRIBUTE
lnc
ltc
lt
wfus
lcabin
ln
Step 2
Step 3
1 wfus
lt
ln
Step 4
Step 5
3
lnc
Step 5 Because this four-seat
FT ltc
Step 5
lfus
In the previous example, we have made a few choices that ask for some elaboration. First
of all, we have allocated part of the cabin in the tail cone. If you do this, you should make
sure that the narrower cross-section is still wide enough to accommodate the passengers
in the cross-section. Secondly, we have not yet drawn the nose of the airplane. In Chap-
ter 8, we will determine the physical dimensions of the electric motor. Then, we will also
add a cowling around the engine and finish the nose of the airplane.
In the following example, we show how the fuselage of a large passenger airplane is
designed in top view.
Example 5.6
In this example, we design the fuselage of Example 5.4 in the top view. We follow the
step-by-step procedure as in Example 5.5, but we make different choices.
Step 1 The width of the cross-section is 3.3 m. We first draw a symmetry line, and then
we draw two parallel lines showing the width of the fuselage.
Step 1
0 2 4 [m]
Step 1 wfus
Step 2 Using (5.9), we compute the length of the cabin to be 28.1 m. We0draw
lcabin 2 the
4 [m]be-
wfus
Step 1 ginning and end of the cabin in our drawing.
Step 2
Step 1
Step 2
Step 2
wfus
wfus
ln
FT lcabin
lcabin
lcabin
0 2
0 2
Step 3 We select a nose length of l n = 4.0 m and a nose fineness ratio of 1.8. The result-
ln
4 [m]
4 [m]
RA
ing nose-cone length is l nc = 1.8 · 3.3 = 5.9 m. We draw the bounds for the nose
Step 3
Step 2 cone and tail cone in the drawing below.
Step 3 ln
lnc
Step 3 lt
ln lnc
Step 4 lt
Step 3 lnc
D
Step 4
Step 4 We select a tail-cone fineness ratio of 3.2 and a tail-to-tail-cone l
ltc t ratio of 0.65.
This gives alnctail-cone length of l tc = 3.2 · 3.3 = 10.6 m and l t = 0.65 · 10.6 = 6.9 m.
Step 4 ltc l
t
Step 5
Step 4 ltc
Step 5
lfus
0l 2 4 [m]
tc
Step 5 lfus28.1 + 4.0 + 6.9 = 39 m. We construct the fuse-
Step 5 The length of the fuselage is l fus = 0 2 4 [m]
lage perimeter in the top view in the drawing below. We make sure we have
Step 5 a blunt nose cone that is slightlylfus elliptical and a tail cone that is tangent to the
0 2 4 [m]
fuselage. Note that the resulting cabin resides partially in the nose and tail cone.
lfus
F OR R EVIEW O NLY . D O N OT D ISTRIBUTE 0 2 4 [m]
lnc
5.5. D ESIGN OF THE F USELAGE IN S IDE V IEW 93
lt
Step 4 Particularly in the tail cone, there is a noticeable decrease in the width of the
cabin within the tail cone. Since the standard five-abreast seats would no longer
fit, this part of the floor space could be effectively used for a galley and/or lava-
tory. ltc
Step 5
lfus
0 2 4 [m]
Examples 5.5 and 5.6 have shown that the step-by-step procedure can be used for dimen-
sioning the fuselage in the top view, regardless of its actual size. Of course, you should
still employ common sense when making choices. If you struggle with deciding what val-
A SSIGNMENT 5.3
FT
ues to choose for the parameters of Table 5.4, you can consult your reference airplanes
(see Section 3.4). By measuring the dimensions of their nose-cone and tail-cone lengths,
you might get a better idea of acceptable values for these parameters.
In this assignment, you will design your fuselage in the top view.
a Choose a value for the following dimensional parameters of the nose and
tail cone of the fuselage. Substantiate your choice in a single sentence.
RA
i nose length: l n
ii nose-cone slenderness ratio: l nc /d fus
iii tail-cone slenderness ratio: l tc /l fus
iv tail-to-tail-cone length ratio: l t /l tc
b Use the sequence of the design steps at the beginning of this section to
design your fuselage in the top view.
c What is the resulting cabin length, l cabin ?
d What is the resulting length of your fuselage, l fus ?
D
with the results from the design of the fuselage cross-section and the design of the fuse-
lage in top view, we can design the fuselage in side view by performing the following
design steps:
FT
Step 9 Copy doors and windshield design to top view.
By following this step-by-step process, you arrive at a conceptual design of the fuse-
lage the in side view. The distribution of the emergency exits could still change at a later
stage due to the integration of the wing and/or the propulsion system. In the following
two examples, the presented procedure is applied to our two demonstration airplanes.
Example 5.7
RA
In this example, we show how the step-by-step procedure can be used to design the
fuselage of the four-seat airplane of Example 5.5 in side view.
Step 1 We take the height of the fuselage (h fus ) from the design of the cross-section and
draw two horizontal lines that mark the upper and lower sides of the fuselage.
Step 2 We copy all the longitudinal dimensions from the top view and add them as
vertical lines to our drawing.
D
ln lcabin lt
Step 1 & 2
hfus
0 0.5 1.0 [m]
lnc ltc
Step 3 The two front seats are for the pilot and possibly a co-pilot. We term this part of
the fuselage the “flight deck.” . Based on the anthropometrics, we estimate the
Step 3length of the flight deck to be 1.6 meters from the beginning of the fuselage. We
draw this dimension into the drawing.
15°
Step 4 & 5 10°
Step 1 & 2
hfus
5.5. D ESIGN OF THE F USELAGE IN S IDE V IEW 0 0.5 1.0 [m] 95
lnc ltc
ln lcabin lt
Step 3
Step 1 & 2
hfus
0 0.5 1.0 [m]
lfd
lnc ltc
Step 4 In this small airplane, there is no real distinction between the main deck and the
15°deck. We drawl a line for the complete
flight lcabin floor, 15 cm above
lt the bottom curve
Step 4 & 5 10° n
of the fuselage, to allow for structure.
Step
Step 5 3Based on anthropometrics, we hpeestimate that the pilot’s eye is located 1.2 m be-
Step 1 & 2 hind the front of the fuselage and 1.1 m above the floor. We assume there is an
hmd=hfd
over-nose requirement of 10 degrees and an upward-viewing requirement of 15
hfus
l
degrees. We draw
fd
the pilot’s eye and field of view. 0 0.5 1.0 [m]
Step 3
lnc
30°
hpe FT 5°
hmd=hfd
ltc
RA
Step 6 We choose an upsweep angle of 5 degrees and construct a straight line from the
lfd tail cone until the end of the fuselage. We place a dot at the
Step 8beginning of the
Step 6 & 7resulting endpoint of the fuselage. We do not place a5°single point for the start of
the15°
30°
fuselage, because we know that an engine still needs to be placed ahead of
Step 4 & 5the10°
fuselage.
0 0.5 1.0 [m]
Step 7 We construct the crown and belly curve of the fuselage by using all the points
hpe
and lines that we have drawn so far as guides for our drawing. Note that con-
struct a distinct cockpit shape, with a windshieldhmd =hhas
that fd a grazing angle that
D
Step 6 & 7 5°
30°
0 0.5 1.0 [m]
Step
Step 8 9Referring to the regulatory requirements regarding emergency evacuation of Ex-
ample 3.14, we need a minimum of two exit doors, each measuring at least 48-
Step 8by-66 cm. We choose to have a larger passenger door on the left-hand side of the
airplane to enable easy boarding of the airplane. In addition, we choose a door
of minimum size on the right-hand side of the airplane. The boarding door is
0 0.5 1.0 [m]
F OR R EVIEW O NLY. D O N OT D ISTRIBUTE
Step 9
hmd=hfd
Step 6&7 5°
96 30° 5. F USELAGE D ESIGN
Step 6 & 7 5°
30°to the pilot’s seat. The emergency exit is positioned in between
positioned close
the pilot seats and rear seats.
Step 8
Step 8
0 0.5 1.0 [m]
Step 9 We copy the location of the two doors in the top view of the airplane to show on
which side each of the emergency exits is located. 0 0.5 1.0 [m]
Step 9
Step 9
FT
Let us briefly reflect on the design we have just constructed. The design fulfills the re-
quirements in terms of pilot visibility and emergency evacuation. Regarding the clear-
ance of the tail, we have already provided some upsweep in the tail cone. However, only
RA
when we integrate the landing gear will we know whether this is an adequate angle. Fur-
thermore, the over-nose angle of 10 degrees shall also impact the shape of the engine
cowling that is yet to be added ahead of the fuselage. The following example shows how
the design sequence is applied to our 130-seater.
Example 5.8
In this example, we design the fuselage of the large passenger airplane of Example 5.6 in
side view.
D
Step 1 From Example 5.4, we know that the height of the fuselage is h fus = 3.84 m. We
draw this height on our canvas.
Step 2 We copy all dimensions of the fuselage from the top view design of Example 5.6
to our drawing in side vies.
ln lcabin lt
Step 3
ln lcabin lt
lfd
Step 1 & 2
Step 4 From Example 5.4,hpethe distance 0from2 the
4[m]
keel of the fuselage to the main deck
is25°
h1.58 m. The height of the lower deck is hmd of the cross-
0.45 m above the keel
fus
Step 4 & 5 section.
20° Wel decide to raise the flight deck by 25 cm with respect hld the cabin
nc
ltc to
deck to enable more volume below the floor for stowage of the nose landing
gear. We draw the main hfddeck, lower deck, and flight deck in our drawing.
3 5 The pilot’slneye is located at 1.0 m behind
StepStep lcabin l
the start of the flight deckt and 1.1 m
4°
above the flight deck. We draw the position of the pilot’s eye along with an over-
Step 61 &
Step & 72 lfd 20 degrees and an upward
0 2 4[m]
nose angle of viewing angle of 25 degrees.
h
25°
fus
hpe
ln
1 &62 We choose an
l
We also draw
hfus
l
lnc
Type A
hfd
FT
lcabin
Type III
hmdltc
hld
lt
0 2 4[m]
0 2 4[m]
Step 9 We copy the doors from the side view of the airplane to the top view.
Step 9
The previous examples have demonstrated that the design sequence presented at the
beginning of this section can be used for both small and large airplanes. In the following
assignment, you will practice with this sequence.
A SSIGNMENT 5.4
In this assignment, you will design your fuselage in the side view.
a What upward view angle and over-nose angle do you choose for the pilot
of your airplane and why?
b What upsweep angle do you choose for the tail cone of your fuselage and
why?
FT
c Follow the steps laid out at the beginning of this section and design your
fuselage in the side view.
d Motivate why your decision on the distribution of the emergency exits.
RA
D
FT
Now that the design process has been defined (Figure 2.1) and the set of top-level re-
quirements has been compiled, it is time to take our first steps on the design ladder.
These first two steps are collectively termed “preliminary sizing,” i.e. using a subset of
our top-level requirements to determine the size of three key aspects of the airplane we
are designing: the airplane mass, the wing area, and the thrust (or power) of the power-
plant. However, to perform the sizing activities, we need to know how the lift and drag of
the airplane are correlated. In other words, we need a drag polar, a mathematical expres-
RA
sion that expresses the drag coefficient of an airplane as a function of its lift coefficient.
Here, we arrive at our first chicken-and-egg problem: how to derive a drag polar from
an airplane for which we do not know anything yet? Fig. 2.1 already gives us a hint: we
need to make assumptions. Assumptions that we replace by analysis results later on in
the design process. Not only do we need assumptions for the drag polar, we also need
assumptions for the characteristics of the powerplant. As the outcome of the preliminary
sizing process depends on the mathematical expression of the drag polar, it is inherently
correlated to the assumptions that we make. Badly chosen assumptions might result in
D
an airplane design with either too favorable or too pessimistic performance. So how do
you make sure your assumptions are good when you embark on the sizing process?
For that, we turn to our reference airplanes that we have defined in Section 3.4. We
will use both the qualitative and quantitative information that we have gathered in As-
signment 3.8. The fact that we use information from the past to start the design process,
does not mean that the resulting airplane design will be simply an average of our refer-
ence airplanes. On the contrary, we only use it to get the process started. Just think of it
as an “egg” that we need to produce our first chicken. Once we have our first chicken, we
can ask it to lay its own eggs, i.e. replace the initial assumptions by analysis results. The
closer the initial assumptions are to the analysis results, the lower the number of design
iterations that are needed for a converged design. This is yet another reason to choose
your reference airplanes carefully.
In this section, we explain how you can estimate the overall mass properties of the air-
plane. First, we show what the various mass definitions mean. We subsequently show
99
100 6. P RELIMINARY M ASS E STIMATION
how, based on the top-level airplane requirements, you can estimate the maximum take-
off mass, the fuel mass, and the operating empty mass. Finally, we demonstrate how to
construct the payload-range diagram.
For the preliminary sizing of the maximum take-off mass of the airplane (m MTO ), we
consider that it comprises three components:
m MTO = m pl + m OE + m ec (6.1)
where m pl is the payload mass, m OE is the (operating) empty mass, and m ec is the mass
of the energy carrier on board. We intentionally have the letter “O” in grey because for
small airplanes, we typically use the empty mass, while for larger airplanes, we use the
operating empty mass in this equation, as will be explained below. Energy can be stored
in a fuel or it can be stored in an energy system, such as a battery. By dividing the left-
hand side and right-hand side by m MTO , we find the so-called unity equation:
m pl m OE m ec
+ + =1 (6.2)
m MTO m MTO m MTO
FT
As m pl is known from the top-level airplane requirements, we need to make an estima-
tion for (m OE /m MTO ) and (m ec /m MTO ) in order to solve (6.2).
where et a em and η eng are the electric motor and engine efficiency, respectively, η p is the
propulsive efficiency, L and D are the lift and drag force, respectively, e f and e bat are the
specific energy of fuel and battery, respectively, and g is the gravitational acceleration.
In both equations, the aerodynamic efficiency (L/D) is linearly related to the range
as well as to the propulsive efficiency (η p ) of the powerplant. Note that within the mass
fraction of (6.3), the maximum take-off mass is in the numerator, while it is in the de-
nominator of the mass fraction of (6.4). For the fuel-burning airplane, the engine (η eng )
in (6.3) is a thermodynamic efficiency for converting fuel into mechanical power. The
electric motor efficiency (η em ) in (6.4) is an electric efficiency of converting electric cur-
rent into mechanical power. The mass-specific energy or simply specific energy of fuel is
denoted with e f in (6.3), which is a property of the type of fuel that you have selected (As-
signment ??). For batteries, the specific energy (e bat ) is a function of the battery chem-
istry that you select as well as the packaging of the battery pack. In the context of this
text, e bat refers to the specific energy of the battery pack, rather than the individual cells
that are in the battery pack. Finally, g is the gravitational acceleration (i.e. 9.81 m/s2 ).
You can use Equation 6.3 when your selected energy carrier is fuel, while (6.4) should
be used when you select batteries as the energy carrier for your airplane.
The Breguet equations relate the cruise range R to the energy consumption. How-
ever, we can imagine that we need even more energy to account for acceleration, climb,
diversion, and contingency. Therefore, we define an equivalent range, R eq that is the
cruise range that could be flown on the total energy that is on board the airplane, i.e.
without acceleration, climb, diversion, and contingency (see also Section 6.1.3). We use
the equivalent range in (6.1) and (6.6) to compute the energy-mass fraction (m ec /m MTO )
in (6.2). We rewrite (6.3) and (6.4) and substitute R = R eq :
mf −R eq
· ¸
=1 − exp (6.5)
m MTO η eng η p (e f /g )(L/D)
m bat R eq
= (6.6)
m MTO η EM η p (e bat /g )(L/D)
FT
Let us examine the ingredients of these equations and how we can compute them. The
energy carrier of your airplane is a design choice, and you can look up the specific energy
of the energy carrier of your choice (e f or e bat ). The lift-to-drag ratio (L/D) is a property
of the airplane. In Section 6.1.1, we show how to estimate the lift-to-drag ratio in cruise
conditions. The engine efficiency (η eng ) or motor efficiency (η EM ) are a function of the
propulsion system. In Section 6.1.2, we show how to estimate these properties along with
the propulsive efficiency (η p ). Finally, in Section 6.1.3, we show how you can estimate an
RA
equivalent range (R eq ).
CL = (6.7)
ρV 2 S w
2D
CD = (6.8)
ρV 2 S w
where ρ is the density of air, V is the velocity and S w is the wing reference area. The
reference area is typically based on the planform area of the wing. Without derivation,
we present that for subsonic, fixed-wing airplanes up to moderate angles of attack, the
drag coefficient of the airplane can be related to the lift coefficient according to:
C L2
C D = C D 0 +C D i = C D 0 + (6.9)
πAw e
where C D 0 is the zero-lift drag coefficient and C D i the induced-drag coefficient. The
latter is a function of the lift coefficient, the aspect ratio (A), and the Oswald factor e.
The aspect ratio is a non-dimensional parameter that describes the wing slenderness. It
is defined as follows:
b2
Aw = w (6.10)
Sw
Equation 6.9 is the drag polar, a two-term, second-order equation in C L . As you can
see, the equation describes a parabola with a minimum at C L = 0. The value of the drag
coefficient at an arbitrary lift coefficient depends on two things: the value of C D 0 and the
value of 1/ (πAw e). Figure 6.1 shows how we define these reference dimensions. Note
that part of the wing reference area is located inside the fuselage.
Sw bw
FT
RA
Figure 6.1: Definition of the reference wing area (S w ) and the reference span (b w ) for a generic airplane
Now that we have established a relationship between the lift and drag coefficient,
we can estimate the maximum lift-to-drag ratio, which is equivalent to the minimum
drag-to-lift ratio. We derive the following:
D
∂ CD
µ ¶ q
=0 → C D = 2C D 0 and C L = πAw eC D 0 (6.11)
∂C L C L
Substituting these values into the expression for lift-to-drag ratio results in a maximum
lift-to-drag ratio of:
s
L 1 πAw e
µ ¶
= (6.12)
D max 2 CD0
This is the lift-to-drag ratio that we substitute in (6.5) and (6.6) when we compute the
fuel mass fraction or battery mass fraction, respectively. There are three ingredients that
we need to estimate in order to quantify the maximum lift-to-drag ratio: Aw , C D 0 , and
e. We will subsequently discuss how to determine each of them.
Example 6.1
Assume that we are designing a jet transport airplane and that we have estimated a
zero-lift drag coefficient and Oswald factor in cruise configuration of C D 0 = 0.0180 and
e = 0.80, respectively. Compute the maximum lift-to-drag ratio.
solution We employ (6.12) to compute the lift-to-drag ratio. We choose an aspect ratio
of Aw = 8. Then we have:
π · 8 · 0.80
r
L
µ ¶
1
= = 16.7
D max 2 0.0180
FT
heavy structure inside to carry the weight of the fuselage and its payload. In general, we
can say that the lower the aspect ratio, the lighter the wing. While we cannot quantify
the effect of your decision at this stage, it is something to keep in the back of your mind
when making a choice for the aspect ratio. Naturally, you can look at your reference
airplanes and chart their wing aspect ratios. This could give you some indication of what
is practical. However, it is up to you to choose the value of your aspect ratio.
A SSIGNMENT 6.1
RA
1. For each of your reference airplanes, compute the reference aspect ratio,
Aw .
2. Decide which aspect ratio you choose for your airplane. Motivate your de-
sign decision by including the following aspects:
• How does the choice in aspect ratio compare to the ones found on
your reference airplanes?
• What is the envisioned effect of your choice on the lift-induced drag?
D
τ
cf = 1
(6.13)
2 ρV
2
The local skin friction coefficient varies over the airplane as does the local pressure drag
at zero lift. Therefore, we introduce the concept of an equivalent skin-friction coefficient,
C̄ f : the average friction coefficient over all of the wetted surface area of the airplane. The
wetted area (S wet ) is defined as all of the airplane’s surface area that is touched by the
airflow during flight. The zero-lift drag coefficient and the equivalent friction coefficient
are related as follows:
C̄ f S wet = C D 0 S w (6.14)
The product of C D 0 and S w is referred to as the parasite drag area. We can rearrange
this equation to have an estimate for the zero-lift drag coefficient as a function of the
equivalent friction coefficient and the ratio between the wetted area and reference area:
S wet
C D 0 = C̄ f (6.15)
Sw
C̄ f and (S wet /S w ).
FT
In order to estimate the zero-lift drag coefficient, we “only” need to find an estimate for
The ratio between the wetted area and reference area is a function of the type of
airplane that you are designing. You can imagine that if you design a pure flying-wing
airplane, this ratio is a bit larger than 2. However, if you design a classical airplane with
a fuselage and a tail, the ratio increases quite rapidly. Figure 6.2 shows examples of air-
planes and where they fall on the (S wet /S w )-chart. You can see that commercial airplanes
have a ratio of around 6, a twin-prop commuter is around 5, and a small single-engine
RA
airplane sits around 4. Based on this very limited information, you can make an assump-
tion of (S wet /S w ) for your airplane. Keep in mind that this assumption will be replaced by
analysis once we have finished the first iteration of the airplane geometry (see Chapter
10).
The equivalent friction coefficient is dependent on the size of the airplane, the rough-
ness of the skin, and the shape of the components. In addition, manufacturing imper-
fections, rivet heads, gaps, air inlets and exhausts, steps in the exterior geometry, and ex-
ternal protrusions (like antennas) also contribute to the equivalent friction coefficient.
D
None of that information is known at this stage, but we can still make an estimate of
the equivalent friction coefficient based on the estimated wetted area of the airplane.
In assignment 6.2, you can see which steps to take in order to do this. One key aspect
is the relation between size and equivalent friction coefficient. For a given production
technology, it can be generalized that the larger the airplane, the smaller its equivalent
friction coefficient.
Figure 6.3 shows the friction-drag coefficient as a function of wetted area for a large
number of subsonic business jets and jet transports. The trend line that is given shows
that there is a nonlinear relation between the equivalent friction coefficient and the wet-
ted area. It shows that for very small airplanes, the equivalent friction drag coefficient
can be as high as 0.0045, while many single-aisle commercial transports are around
0.0030. The asymptotic behavior that is suggested by the trend line shows that larger
airplanes, have a lower average friction coefficient. However, you should keep in mind
that this trend only holds as long as the surface finish is the same [6].
FT
Figure 6.2: The ratio between wetted area and reference area (S wet /S w ) depends on the type of airplane you
design. Adapted from [9].
RA
D
Figure 6.3: Equivalent friction coefficient versus wetted area for turbine-powered airplanes. Data from [6].
A SSIGNMENT 6.2
In this assignment, you are going to estimate the zero-lift drag coefficient of your
airplane. To complete this assignment, you need data from Assignment 3.8.
a. Make an estimate of the ratio of wetted-area to reference area, i.e.
(S wet /S w ) based on Fig. 6.2, your reference airplanes, and your own notion
of what the airplane will look like.
b. Using the value of (S wet /S w ) from a. and the wing areas of your reference
airplanes, compute for each reference airplane the approximate wetted
area.
c. Using Figure 6.3 and the estimated wetted areas of your reference air-
planes, estimate the equivalent friction coefficient for your airplane.
d. Using (6.15), estimate the zero-lift drag coefficient of your airplane.
FT
Using the steps of Assignment 6.2, you can now make an estimation of the first unknown
parameter of the drag polar. The second parameter is known as the Oswald efficiency
factor, e, and is a correction factor that represents the change in lift-induced drag of
an airplane compared to an ideal wing of minimum lift-induced drag having the same
aspect ratio as the airplane’s wing. The Oswald efficiency factor pertains to a complete
airplane and is therefore different from the span efficiency factor,, which pertains only
to a wing. While the Oswald factor is related to the span efficiency factor, they are not
the same. Apart from the spanwise lift distribution, the Oswald factor depends on the
RA
orientation of the fuselage, the flap setting, as well as the presence of engine nacelles.
For an airplane without flap deflection, the lift-induced drag can therefore be estimated
as follows: µ ¶
1 1
CDi = ψ + C2 = C2 (6.16)
πAw ϕ L πAw e L
where ψ is the parasite drag dependent on the lift coefficient and ϕ is the span efficiency
factor of the wing. Figure 6.4 shows how πA1 e changes as a function of aspect ratio
w
for large transport airplanes. The trend line that is shown assumes a span efficiency of
D
1
e= (6.17)
πAw ψ + ϕ1
A SSIGNMENT 6.3
In this assignment, you are going to estimate the Oswald factor of your airplane.
a. What value for the parasite drag parameter, ψ, do you assume?
b. What value of the span efficiency factor, ϕ, do you assume?
c. Based on these assumptions, what is the Oswald efficiency factor, e of your
airplane?
FT
Figure 6.4: The Oswald factor, e, as a function of aspect ratio, Aw , for a variety of transport airplanes. Data
from [6].
RA
6.1.2. E STIMATING THE P ROPULSION S YSTEM E FFICIENCIES
In this section, we show how you can estimate the propulsion system efficiencies, η EM ,
η f , and η p . We distinguish between electric motors and fuel burning engines but also
between propeller propulsion and jet propulsion.
For electric motors, η EM can often be found in online databases for electric motors.
Typically, electric motors have a very high efficiency (i.e. 95%). However, extracting elec-
tricity from a battery and transporting it through cables also leads to losses. If cables
D
and batteries are properly designed, this discharge efficiency can be close to 100%. In
this textbook we neglect the transmission efficiencies and only consider the electric mo-
tor efficiency.
For fuel-burning engines, often the thrust-specific fuel consumption (TSFC) or power-
specific fuel consumption (PSFC) can be found in the open literature. For piston engines
and turboprop engines, the PSFC is related to the (thermal) efficiency of the engine, η eng .
For a given engine, the following relations can be used in (6.3):
1
e f η eng = (6.18)
PSFC
where e f is the specific energy (i.e. the energy per unit mass) of the fuel that is burned
in the engine. The power-specific fuel consumption for propeller engines is often ex-
pressed in lb/hp/hr. To use these published values in the context of the methods pre-
sented above, these units need to be converted to SI units. In the following example, we
show how the efficiency of a reciprocating engine can be computed.
Example 6.2
The Lycoming O-320 (Fig. 6.6) is a piston engine with a maximum power of 160hp (120
kW) that burns avgas (aviation gasoline). The Lycoming O-320 power-specif fuel con-
sumption is approximately 0.48 lb/hp/hr. Converting that to SI units, this results in
0.45 [kg/lb]
PSFC = 0.48 [lb/hp/hr] · 745 [W/hp]·3600 [s/hr] = 8.05 · 10−8 kg/W/s. We can now directly
compute (e f η eng ) using (6.18):
1
e f η eng = = 1.24 · 107 J/kg
8.05 · 10−8
Using a specific energy of 44.7 MJ/kg for avgas, we can also compute the engine effi-
ciency:
1.24 · 107
η eng =
FT
4.47 · 107
= 0.28 (6.19)
To compute the engine efficiency for a turboprop engine, the same procedure can be
used for a piston engine.
Propulsion systems that drive a propeller are typically rated by their mechanical power
(P ). The thrust (T ) is the resultant force that is generated by a propeller. To maintain a
certain level of thrust at a given speed (V ), the propeller is spun by an engine or motor.
Therefore part of the shaft power is required to overcome the aerodynamic torque that
RA
is produced by the propeller. We therefore use the propulsive efficiency of the propeller
(η p ) to correlate the shaft power to the thrust and speed of the airplane as follows:
ηp P = T V (6.20)
Here, η p is the propulsive efficiency of the propeller. It should be noted that this relation
does not work when V → 0 but works well for all flying conditions.
The propulsive efficiency, η p , is a measure for how much of the shaft power is con-
D
Table 6.1: Tabulated values for η p for various airplane types. Data from Ref. [10].
airplane type ηp
homebuilts 0.70
single engine/motor props 0.80
twin engine props 0.82
regional turboprops 0.85
value, i.e. η j = η eng η p . The published value of TSFC is typically found in the design cruise
condition. We can relate η j to the TSFC as follows:
VCR
e f η eng η p = e f η j = (6.21)
TSFC
where VCR is the cruise speed. Often imperial units are used to express the TSFC, which
results in lb/lb/hr. The following example shows how to compute the engine efficiency
of a turbofan engine.
Example 6.3
FT
The CFM-56 (Fig. 6.5) is a turbofan engine that burns kerosene and has a thrust of 30,000
lb (132 kN) at sea level. A turbofan engine has a gas turbine at its core that drives a ducted
fan. The thrust-specific fuel consumption of a CFM-56 is 0.55 lb/lb/hr.1 If we convert
0.45 [kg/lb]
that to SI units, we have TSFC = 0.55 [lb/lb/hr] · 4.4 [N/lb]·3600 [s/hr] = 1.56 · 10−5 [kg/N/s] =
15.6 [g/kN/s]. We assume that this engine is used on a transport airplane that has a
RA
cruise speed of Vcr = 230 m/s. We can now compute the value of (e f η eng η p ) according to
(6.21):
230
e f η eng η p = = 14.7 MJ/kg
15.6 · 10−6
The resulting engine efficiency is then:
14.7
η j = η eng η p = = 0.34
43
D
The TSFC of turbofan engines is dependent on the altitude and Mach number at which
the engine is operated. In the context of this book, we only consider the cruise con-
ditions when we consider the TSFC. Naturally, the TSFC of an engine depends on the
characteristics of the engine components. In particular, the overall pressure ratio, the
total temperature at the entry of the turbine, and the bypass ratio (B ) have an impact
on the fuel consumption of the engine. The bypass ratio is the ratio between the mass
flow going through the gasturbine core (h, for hot) and the mass flow that bypasses the
gasturbine (c for cold) and only flows through the fan:
ṁ c
B= (6.22)
ṁ h
1 Source: wikipedia.org
Figure 6.5: The CFM56 is a turbofan engine belong- Figure 6.6: The Lycoming O-320 is a piston engine
ing to Example 6.3. Photo: David Monniaux. belonging to Example 6.2. Photo: A. Hunt.
The bypass ratio has a strong impact on the efficiency of the engine as well as the size
FT
of the engine. For the same thrust, increasing the bypass ratio improves the engine effi-
ciency, but also increases the engine diameter.
Despite the complex relation between engine design parameters and engine effi-
ciency, the bypass ratio is the only turbofan design variable that we consider in this book.
Based on engine data from the open literature, we have deduced the following relation
between TSFC in cruise conditions and the bypass ratio, B :
A SSIGNMENT 6.4
D
In this assignment, you will determine the efficiency of your propulsion system.
Depending on your choice of engine (from Assignment 7.4), answer the following
questions:
a. For electric motors, what electric-motor efficiency, η em , do you assume?
b. For piston engines or turboprop engines, What (thermal) efficiency, η eng ,
do you assume?
c. When using a propeller, what propulsive efficiency, η p , do you assume?
d. For turbofan engines, what bypass ratio do you choose? What jet efficiency,
η j , do you compute?
24
22
TSFC=22B-0.19
20
TSFC (g/kN/s)
18
16
14
12
0 5 10 15
Bypass Ratio, B (-)
wikipedia.org.
FT
Figure 6.7: Published values cruise TSFC and bypass ratio for existing turbofan engines.
Let us return to the mission profile that we introduced in Section 3.2. Examples 3.3
and 3.4 show that apart from the cruise phase, other phases make up the complete mis-
sion of the airplane. You can imagine that during take-off and climb, we throttle up the
engine or motor to accelerate and climb to the cruise altitude. However, during descent
Data source:
RA
and deceleration, we throttle down. From experience, we know that over the complete
mission, we use more energy than the energy that is needed to fly between two points in
cruise conditions. In other words, part of the energy is “lost,” in the sense that the energy
is not converted to mission range. This lost range can be computed as a function of the
cruise speed and cruise altitude as follows [16]:
µ ¶ Ã 2
!
1 L VCR
R lost ≈ h CR + (6.24)
0.7 D CR 2g
D
where 0.7 accounts for the fact that we do not fly as efficiently during acceleration and
climb as we do during cruise.
If we know approximately how much this lost range (R lost ) is, we can apply the Breguet
range equations and include the complete mission profile in a single equation. We there-
fore define the equivalent range as follows:
¡ ¢
R eq = (R nom + R lost ) 1 + f con + 1.2R div + t E VCR (6.25)
In this equation, the nominal range (R nom ) and the lost range in the first bracket deter-
mine the trip fuel. This is multiplied by (1 + f con ), where f con is the fraction of trip fuel
that is used for contingency (e.g. headwind or rerouting). Example 3.19 showed that
f con = 0 for general-aviation airplanes and f con = 5% for commercial-aviation airplanes.
The value of R div determines the fuel required for a missed approach and the diversion
to an alternate airport. The factor of 1.2 is added because the diversion is typically flown
at a lower aerodynamic efficiency than during the cruise condition [16]. Finally, the en-
durance time (t E ) to hold (loiter) is multiplied by the cruise speed to arrive at the equiv-
alent range that is covered during the holding time. The following example shows how
to compute the equivalent range for a turbofan-powered airplane.
Example 6.4
FT
cruise, hold, descent, land, and taxi. We choose a cruise altitude of h CR = 11 km. The
cruise speed can be computed to be VCR = 236 m/s when employing (7.24) in combi-
nation with (7.19). We compute the lost range using (6.24) while using the maximum
aerodynamic efficiency of Example 6.1: (L/D)CR = 16.7:
R lost = 330 km
As this is a commercial jet we have f con = 0.05. So, with R nom = 3000 km, t E = 1800 s, and
R div = 400 km, we compute the equivalent range using (6.25):
RA
R eq = 4400 [km]
Note that the equivalent range is almost fifty percent more than the nominal range. Also,
note that we round the value of the range from R = 4401.8 km to R = 4400 km. We do
this because the numbers that we use in the computation also do not have more than
two significant figures. The number of significant figures represents the accuracy of a
calculation result. In our case, much of the input of the calculation does not have more
D
than two significant figures. We therefore also report the result of the calculation in two
significant figures. Not only does this do justice to the calculation accuracy, but it also
allows for a much quicker interpretation of the results.
Example 6.5
In this example, we compute the fuel mass fraction for a turbofan-powered airplane.
Let us use the equivalent range of Example 6.4: R eq = 4400 km. We assume the same ef-
ficiency as the CFM56 of Example 6.3: η j = 0.33 and a maximum aerodynamic efficiency
of Example 6.1: (L/D)CR = 16.7. Now, the fuel mass ratio can be computed using (6.5)
knowing that η j = η eng η p and using R = R eq :
mf
= 0.14
m MTO
The previous example shows that the fuel-mass fraction (m f /m MTO ) is 14 percent. By
varying the input values to this calculation, you can determine how sensitive the fuel-
mass fraction is to the assumed values of L/D and η j . In the next example, we consider a
general-aviation airplane that is battery-powered.
Example 6.6
FT
rules (VFR). Therefore, an energy reserve of 30 minutes is required according to ICAO
Annex 6 (see Example 3.19). It is to have a nominal range of 450 km. Furthermore, it has
a cruise lift-to-drag ratio of L/D = 13.8, a motor efficiency of η EM = 0.94, and a battery
pack with a specific energy of e bat = 350 Wh/kg. To compute (m bat /m MTO ), we first need
to estimate the equivalent range with (6.25). The lost range is computed with (6.24) for
h CR = 1800 m and VCR = 70 m/s:
R lost = 40 km
With f con = 0 and R div = 0, and t E = 1800 seconds we have:
RA
R eq = 620 km
We then convert e sp = 350 [Wh/kg] = 1.26 [MJ/kg]. We subsequently employ (6.6) to find:
m bat
= 0.46
m MTO
D
You may observe that the battery-mass fraction of the previous example is much higher
than the fuel-mass fraction of Example 6.5, even though the range is much smaller.
This is due to the lower specific energy of batteries compared to fuel. However, the
battery-mass fraction is much more sensitive to changes in L/D than the fuel-powered
airplane. This means that the battery-mass fraction would decrease quite rapidly with an
increased lift-to-drag ratio. Furthermore, you might have seen that the assumed η EM is
quite high compared to the ones we computed for the CFM56 (Example 6.3) and the Ly-
coming O-320 (Example 6.2). The Pipistrel E-811 electric motor (Fig. 6.8) is an example
of an electric motor with a high transfer efficiency.
The previous examples featured a turbofan engine and an electric motor. The proce-
dure for estimating the fuel mass fraction for an airplane with a reciprocating engine or
a turboprop engine is the same as for the turbofan engine (see Example 6.5).
We have shown how you can compute the fuel-mass fraction and battery-mass frac-
tion of the airplane you are designing. The examples above have demonstrated that
Figure 6.8: Pipistrel E-811 60kW electric motor. Figure 6.9: Pratt and Whitney PW120 1.3 MW turbo-
Photo: Pipistrel. prop engine. Photo: Imnop88a
these fractions are sensitive to assumed input parameters such as (L/D), η p and η EM ,
but also to the design choice on the aspect ratio (A) and properties of the energy carrier
FT
(e fuel or e bat ) in the next assignment you are going to practice with this.
A SSIGNMENT 6.5
known as the (operating) empty-mass fraction. However, before we compute this, let
us first look at the difference between empty mass and operating empty mass. The (ba-
sic) empty mass of an airplane comprises the mass of the structure (i.e. the airframe),
the propulsion group, and the fixed and removable equipment. The operating empty
mass also includes the mass of all the items and people that are needed to operate the
airplane: cabin crew, cockpit crew, drinks, meals, potable water, non-potable water, op-
erating manuals, etc. Figure 6.10 shows how the various mass definitions relate to one
another. Now, for small airplanes, the crew and the payload typically overlap. For exam-
ple, a four-seater airplane would have a payload mass corresponding to four passengers.
Little additional mass is added for operational items. Therefore, for small airplanes, we
typically find that their empty mass is used in (6.2). On the other hand, larger airplanes
typically do operate with a crew that is separate from the payload. In that case, we count
the crew plus all additional operational items to the operating empty mass. Whether you
should use m E or m OE can be decided by looking at your reference airplanes. Can you
FT
Figure 6.10: Graphical interpretation of the various mass definitions for an airplane. Note that typically the
word “weight” is used in the open literature. After: Torenbeek [15]
only find empty-mass quotes? Then use empty mass in (6.2). Can you only find operat-
ing empty mass quotes? Then you had better use operating empty mass in (6.2).
To compute m OE /M MTO , we use the data from our reference airplanes. While we
design a new airplane, we need to have a good starting point for the estimation of the
(operating) empty mass fraction. The best we can do at this stage is to use data from air-
RA
planes that have come before, knowing that in Chapter 12, we will analyze our airplane
and come up with an improved estimation of the (operating) empty mass. The following
example demonstrates how to estimate this.
Example 6.7
In this example, we estimate the (operating) empty mass fraction for the design of a twin
piston-prop airplane. For this airplane, we have the following top-level requirements:
• Number of passengers: 4
D
Note that the open literature quotes the empty mass (m E ) of the reference airplanes.
Therefore, we also use the empty mass in our computations. To determine m E /m MTO ,
we can simply take the average of the values listed for the reference airplanes. However,
we can also take two other factors into account: the range and the year the airplane was
first flown. The range of the airplane impacts the fuel-mass fraction of an airplane and,
therefore, also the empty-mass fraction. The year an airplane was first flown says some-
thing about the technology that was used in constructing the airplane. Newer airplanes
are likely to be more representative than older airplanes. In this instance, the two newest
planes (DA-42 and PT2006) have an empty-mass ratio of 0.63 and 0.67, respectively. The
nominal range of our airplane is in between the ranges of those airplanes. The Let-200D
and PA-44 are closest in terms of range and have an empty-mass ratio of 0.68 and 0.63,
respectively. But also the Duchess and Cessna 310 are not far off in terms of range, having
an empty-mass ratio of 0.63 and 0.62, respectively. Based on this qualitative assessment,
we choose to assume a value of m E /m MTO = 0.64, which is close to the arithmetic mean
of all the airplanes.
FT
This example has demonstrated how one can make a well-funded assumption for the
empty-mass fraction of the airplane you are designing. Not only is it important to select
the right reference airplanes, but it is equally important to assess how close the range of
the airplane is to the range that is specified in the top-level requirements. If the range of
a reference airplane is much less or much more than the required range of your airplane,
you might want to discard that airplane when computing the (operating) empty-mass
fraction. Secondly, it is important to take the year of first flight into account. Newer
RA
airplanes typically take advantage of improved technology levels, which typically reduce
the weight of the airframe, the powerplant, and the systems. In the following assignment,
you estimate the (operating) empty-mass fraction for your airplane.
A SSIGNMENT 6.6
In this assignment, you will estimate the (operating) empty mass of your air-
plane. You need to consult the reference airplanes that you have selected in As-
signment 3.8.
D
a. For each reference airplane, find the maximum take-off mass value in kg.
b. For each reference airplane, find the empty mass or operating empty mass
in kg.a
c. Compute the (operating) empty-mass fraction for each reference airplane.
d. Make an estimation of the (operating) empty-mass fraction of your air-
plane.
a whichever one is listed in the open literature
upper end, we can imagine an airplane with zero payload and zero battery mass (essen-
tially a glider), which has a theoretical m OE /m MTO = 1. On the lower end, we can think
of airplanes that have been operated over an extremely long range (i.e. high m f /m MTO )
and therefore had to have a low m OE /m MTO . The Rutan Voyager (Figure 6.11) and Rutan
Global Flyer (Figure 6.12), were both designed to fly around the world. Their respective
empty-mass ratios were 0.23 and 0.17.3 Naturally, these airplanes were experimental,
did not need compliance with the airworthiness regulations, and had little comfort for
the crew. Therefore, they were very lightweight. Nonetheless, the empty-mass fractions
of these airplanes could serve as an indicative lower bound for the feasible (operating)
empty-mass fraction range. Secondly, you might want to consult studies into battery-
electric airplanes. You can use data from design reports or scientific publications to es-
timate your (operating) empty-mass fraction.
FT
RA
Figure 6.11: The Rutan Voyager has an empty-mass Figure 6.12: The Rutan Global Flyer has an empty-
fraction of 0.23. Photo: NASA/Thomas Harrop mass fraction of 0.17. Photo: Alan Radecki
m pl
m MTO = ³ ´ ³ ´ (6.26)
m m
1 − m OE − m f
MTO MTO
The right-hand side of this equation is the payload mass, for which the design payload
mass should be used. The other two mass fractions have been explained in the previous
text. The next example demonstrates this.
Example 6.8
In this example, we consider the twin prop of Example 6.7. We have m pl = 400 kg,
3 Source: wikipedia.org
(m E /m MTO = 0.64) and (m f /m MTO ) = 0.15. Substituting this in (6.26) results in:
m MTO = 1900 kg
When comparing this value to the reference airplanes, we see that it falls within the range
of values that are listed under m MTO . With the payload mass fractions, we can now di-
rectly compute the fuel mass and empty mass for this airplane:
m OE = 1200 kg m f = 290 kg
Again, we round the numbers in these calculations to two significant figures to represent
the accuracy of our calculations. However, we store the unrounded numbers to employ
them in the calculations that are yet to come. In the next assignment, you compute the
characteristic masses for your airplane. Use the correct number of significant figures
when reporting the masses.
A SSIGNMENT 6.7
FT
In this assignment you are going to compute the characteristic masses of your
airplane. Perform the following tasks:
a. Compute the maximum-take-off mass of your airplane
b. Compute the (operating) empty mass of your airplane
Now that you have computed the take-off mass of your airplane look back at the
design process shown in Figure 2.1. You can see that the maximum-take-off mass is
RA
one of the important metrics that is needed for many of the next steps to come. For
CS/FAR 23 airplanes, you now also know whether your airplane is going to be above or
below 6,000 lb (2722 kg), telling you which set of climb-gradient requirements to select
(see Example 3.12). Also, think about the decisions you have already made. You have
selected the type of energy carrier. You have selected an aspect ratio. You have selected
a propulsion system. You may have even selected whether you wish to fly this airplane
under IFR or only under VFR. These decisions are all design choices and are an important
step in the design of your airplane.
D
payload and the maximum structural payload. The maximum structural payload is the
maximum payload that the structure of the airplane can carry. The latter is almost twice
as much as the former.
MTOW − 316 000 kg (696 661 lb); MFC 158 790 l (41 949 US gal)
RANGE (km)
2 000 4 000 6 000 8 000 10 000 12 000 14 000 16 000 18 000
70
120
50
PAYLOAD (x 1 000 kg)
100
30
20
10
366 PASSENGERS
60
40
20
RA
95 kg PER PASSENGER INCLUDING BAGGAGE
BASIC CONFIGURATION WITH CREW REST COMPARTMENTS
AND OTHER OPTIONAL FEATURES
0 0
500 1 500 2 500 3 500 4 500 5 500 6 500 7 500 8 500 9 500 10 500
RANGE (nm)
Figure 6.13: Payload-range diagram of the A350-1000 as published by Airbus in the document “airplane Char-
acteristics Airport and Maintenance Planning.”
D
auxiliary range, i.e. the difference between the equivalent range and the nominal range:
The auxiliary range is the range of the airplane that would be covered if the energy that is
required for climbing, accelerating, holding and diverting would be used to fly at cruise
conditions. The following example shows how this works for a battery-powered airplane:
Example 6.9
Let us take the airplane of Example 6.6 and produce the payload-range diagram.
1. The design range of this airplane is 450 km with a payload of 200 kg. Since the
battery mass does not change if we fly a shorter mission, the design payload mass
equals the maximum structural payload mass. This gives us two points in the pay-
load range diagram, which are tabulated below.
2. Before we compute the range at zero payload, we first compute the auxiliary range,
i.e. the equivalent range that is needed for climbing, accelerating, and holding
(6.27):
R = η em η p
D
FT
R aux = 620 − 450 = 170 [km]
3. We employ (6.4) to find the range that can be flown at m pl = 0:
L e bat
µ ¶µ
g
¶µ
m bat
m OE + m bat
¶
− R aux = 530 [km]
We add this value to the table below as the theoretical ferry range of this airplane.
RA
4. We realize that this is a two-person airplane and that at least one person is needed
to fly the airplane. In other words, the payload cannot go to zero when we have
a pilot that is part of the payload mass. When we assume that a single pilot (plus
luggage) has a mass of 100 kg, we find a practical ferry range of 490 km.
5. We use the data points to construct the payload range diagram of Figure 6.14.
The previous example showed that the maximum structural payload of a battery-powered
airplane equals the design payload. Fuel-powered airplanes, however, have the ability to
trade fuel mass for payload mass while keeping the maximum take-off mass constant.
If there exists a requirement to fly a higher payload mass than the design payload mass,
that payload mass becomes the maximum structural payload mass m pl max . The fuel
mass available to fly with maximum structural payload mass can be computed as fol-
lows:
m f = m MTO − m OE − m pl max (6.28)
To compute the associated range, we can employ (6.3). However, we need to be careful
as the resulting range is the equivalent range. In order to compute the nominal range,
we still need to subtract the auxiliary range R aux .
Example 6.10
100
design range
practical minimum payload: single pilot
50
0
0 100 200 300 400 500 600
Range, R (km)
63
FT
We construct the payload-range diagram for a jet airplane with a design range of 3000
km and a design payload of 19 tonnes. Furthermore, this airplane has a maximum struc-
tural payload requirement of 23 tonnes. Using the methods introduced in the previous
section and the values of Example 6.5, we have the following mass characteristics for the
design condition (all in tonnes):
m MTO m OE
35
mf
9.1
m pl
19
RA
We can now start constructing the payload range diagram.
1. The first point in the payload range diagram that can be easily constructed is zero
range combined with our maximum structural payload (first column in table be-
low).
2. The second point is the combination of design range and design payload (third
column in the table below).
3. Before we compute any of the other non-zero range points, we first compute the
auxiliary range with (6.27):
D
4. We compute the available fuel mass at the maximum structural payload using
(6.28):
m f = 5.1 [t]
5. Now, we compute the range at maximum structural payload with (6.3) for m pl = 23
tonnes:
L ef m OE + m pl + m f
µ ¶µ ¶ µ ¶
R = η eng η p ln − R aux = 990 [km]
D g m OE + m pl
We can add this combination of payload mass and range to the table below (sec-
ond column)
6. We compute the ferry range by employing (6.3) with m pl = 0 and m f = 9.1 tonnes:
L ef m OE + m pl + m f
µ ¶µ ¶ µ ¶
R = η eng η p ln − R aux = 5100 [km]
D g m OE + m pl
We add this data point to our table below in the last column.
7. Based on the payload-range points in the table below, we can now construct the
payload-range diagram in Figure 6.15.
m pl [t] 23 23 19 0
R [km] 0 990 3000 5100
20
FT
Payload mass, mpl (t)
15
harmonic range
design range
ferry range
10
5
mTO<mMTO mTO=mMTO mf=Vf, maxρf
0
RA
0 1000 2000 3000 4000 5000 6000
Range, R (km)
The previous example produced a payload-range diagram that looks similar to the one
we saw for the A350-1000 in Figure 6.13. We can observe that an increase in payload
mass rapidly reduces the available range of the airplane. On the other hand, if we fly
this airplane with a payload mass of only 10 tonnes, we can achieve a range of 4000 km.
D
You can see that below the design payload, the available range is limited by the tank
volume, Vf . We have implicitly assumed that the tanks of our airplane are sized to the
design-range requirement. However, we might find out in Chapter 10, that there is more
volume available in the airplane to store fuel. If this is the case, we can make the choice
to trade payload for fuel beyond the design range. We can do this until the tank volume
is completely filled with fuel. The second kink, as well as the ferry-range point, will then
shift to the right in our payload-range diagram.
A SSIGNMENT 6.8
In this assignment, you will construct the payload-range diagram of your air-
plane:
a. State the combination of design range and design payload.
FT
RA
D
FT
Many of the requirements that we have derived from the mission profile relate to the
flight performance of your airplane (see Example 3.3). Also, many of the airworthiness
regulations dictate flight performance minima (see Table 3.2). This includes require-
ments on the climb rate, the climb gradient, cruise speed, landing distance, take-off dis-
RA
tance, approach speed, and sometimes stall speed. In this chapter, we use these require-
ments to size the wing and the propulsion system. Section 2.3 has explained how the siz-
ing process works in general. In this chapter, we are going to make this more concrete.
We employ flight mechanics relations to estimate the wing loading and power loading
(or thrust-to-weight ratio) of your airplane. These, in turn, are required to compute the
wing area and the power or thrust of the propulsion system at take-off.
Let us first define “wing loading,” “power loading,” and “thrust-to-weight” ratio. The
wing loading is defined as the amount of weight (W ) per unit of wing area (S w ). The
D
weight is the mass of the airplane times the gravitational acceleration (g ). During flight,
fuel-burning airplanes become lighter. Therefore, their wing loading changes. However,
in the context of the design of a fixed-wing airplane, when we talk about “wing loading,”
we imply the maximum take-off weight per unit area, i.e.:
WTO m MTO g
= (7.1)
Sw Sw
The power loading is a term that we associate with propeller-powered airplanes. The
power (P ) produced by electric motors is virtually independent of the altitude, ambient
temperature, or flight speed. While the power produced by air-breathing engines is in-
dependent of speed, they produce less power with increasing altitude and temperature
(see Section 7.4.3). When we mention “power loading” in the context of this textbook,
we mean the maximum take-off weight per unit of available take-off power at mean sea-
125
126 7. W ING AND P ROPULSION S YSTEM S IZING
jective. In the subsequent paragraphs, we show how you can construct this diagram and
select the appropriate design point.
As the flight performance of airplanes is related to their aerodynamic performance,
the drag polar (Section 6.1.1) is required. Furthermore, the maximum lift coefficients also
needs to be assumed along with the assumed propeller efficiency (if necessary).
The following example introduces the two airplanes and associated aerodynamic
characteristics that are used throughout the remainder of this section.
Example 7.1
In this section, we consider the design of a propeller-powered airplane and a jet air-
plane. Based on the TLARs, we know that the propeller airplane is going to be certified
1 The international standard atmosphere defines the pressure, density, and temperature as a function of alti-
tude above mean sea level (SL). See wikipedia.org.
LFL
LFL
W/P T/W
feasible TOFL
Vcr Mcr
G
G
feasible TOFL
W/S W/S
Figure 7.1: Schematic examples of the constraint diagrams for propeller-powered airplanes (left) and jet-
powered airplanes (right). Each line represents a flight performance requirement and imposes a constraint
on the feasible design space.
according to CS/FAR-23 and the jet airplane according to CS/FAR 25. We have made the
following design decisions for each of them:
energy carrier
Number of engines/motors (Ne )
aspect ratio (A)
landing gear
Bypass ratio (B )
Cruise altitude (h CR ) [m]
FT propeller
battery
1
9.0
fixed
-
1800
jet
kerosene
2
8.0
retractable
10
11,000
RA
For the propeller airplane, the performance requirements are listed below. Note that all
but one requirement is to be evaluated at sea-level (SL) altitude assuming ISA conditions
and at maximum take-off mass. Requirements include engine operative (EO) and engine
inoperative (EI) conditions as well ass three configurations regarding the flap settings:
cruise (CR), take-off (TO), and landing (L).
requirement alt. (m) temp (K) eng. flaps gear m/m MTO value
VS0 (m/s) 0 TISA EI L ↓ 1.0 31
LLF (m) 0 TISA EI L ↓ 1.0 750
D
The requirements for the jet airplane are a bit more elaborate and are shown in the table
below. Here, we also consider “hot” conditions for two requirements, where the temper-
ature is 15◦ higher than standard ISA conditions. Also, we have two one-engine inopera-
tive (OEI) requirements: one for the climb gradient and one for the take-off field length.
Furthermore, we also consider “hot-and-high” conditions for the take-off field length.
Furthermore, the landing gear (gear) can either be up or down, which is indicated by an
arrow in the table below. Finally, the ratio m/m MTO is lower than one for some of the
requirements as we assume that part of the fuel has been burned when this requirement
needs to be satisfied.
requirement alt. (m) temp. (K) eng. flaps gear m/m MTO value
VAPP (m/s) 0 TISA AEI L ↓ 0.85 68
LLF (m) 1600 TISA AEI L ↓ 0.85 1800
M CR (-) 11 · 103 TISA AEO CR ↑ 0.95 0.81
c (m/s) 11.5 · 103 TISA + 15◦ AEO CR ↑ 0.95 0.50
G (%) 0 TISA OEI TO ↑ 1.0 2.4
LTO (m) 1600 TISA + 15◦ OEI TO ↓ 1.0 2500
The requirements presented in Example 7.1 are made specific and measurable according
to the guidelines provided in Section 3.2. The cruise altitude, h CR , needs to be assumed
at this stage. In Chapter 10, we will show how the optimal altitude can be computed. If
the optimal altitude is known, this should be used instead of the assumed altitude we
use here. The elevated temperature and altitude of the jet airplane lead to a reduction in
air density. The lift and drag directly depend on the density. The available shaft power
of piston engines and turboprops also scales with density according to (7.23). Depend-
ing on the bypass ratio, the thrust of a turbofan engine scales with ambient pressure,
FT
temperature, and Mach number according to Eqs. (7.29)-(7.32).
To evaluate the requirements, we introduce the mass ratio β as follows:
β=
m
m MTO
where m is the mass at which the requirement needs to be evaluated. For example, a
landing field length requirement might specify a landing mass below the maximum take-
off mass. At this stage of the design, we therefore use the mass ratio. However, we need
(7.4)
RA
to be careful that 1 − (m f /m MTO ) < β ≤ 1 as the reduction in mass is due to the fuel that
is burned off. Many transport airplanes are designed with a maximum landing mass
(MLM) that is significantly lower than the maximum take-off mass. When such an air-
plane needs to land right after take-off in an emergency situation, it has to quickly dump
fuel to arrive at its MLM. On the other hand, for full-electric airplanes that use batteries
as their energy carrier, β ≡ 1 for all requirements.
region, the lift is still increasing with the angle of attack. At the stall angle-of-attack, the
maximum lift coefficient is reached, beyond which the lift coefficient decreases with the
angle of attack.
CL max
stall onset
lift-curve slope
Sw
< V CL
FT
If we examine the lift equation (6.7), substitute V = VS0 and C L = C L max and know that
L = βWTO , we can interpret this requirement as a constraint on the wing loading:
WTO 1 ρ 2
β 2 S0 max
(7.5)
RA
For the jet airplane, we have an approach speed requirement, which, according to CS/-
FAR 25.125, should be higher or equal to 1.23 times the reference stall speed. In other
words, for the jet airplane, the wing loading should not be higher than:
WTO 1 ρ Vapp 2
µ ¶
< C L max (7.6)
Sw β 2 1.23
As you can see, for either requirement, we need a value for the maximum lift coefficient
in order to evaluate the constraint. Many airplanes feature so-called high-lift devices to
D
increase the maximum lift coefficient of the wing. As you can see, from the equations
above, increasing the maximum lift coefficient leads to a higher wing loading for the
same speed requirement. In the subsequent section, we explain how you can estimate
the maximum lift coefficient.
selecting the value for C L max in cruise, take-off, and landing configuration for a variety
of airplane types. Note, that minimum and maximum values are given that are typically
found on the types of airplanes that are listed. In Chapter 10, you will design the high-lift
system for your airplane, and the assumption that you make here will be updated with a
value based on the actual flap design of your airplane.
Table 7.1: Tabulated minimum and maximum values for C L max for various airplane types. Data from [10]
FT
The maximum lift coefficient in landing configuration is usually the highest. The
maximum lift coefficient in cruise configuration is the lowest. The take-off flap setting is
usually somewhere in between the cruise configuration and the landing configuration.
This is done to strike a balance between the increase in lift and the decrease in lift-to-
drag ratio due to the deployment of high-lift devices. In the subsequent assignment, you
will make assumptions on the maximum lift coefficient of your airplane. The questions
will guide you in making the assumptions.
RA
A SSIGNMENT 7.1
In this assignment, you make explicit assumptions about the maximum lift coef-
ficients of your airplane.
a. Considering Table 7.1, what airplane type does your design fall under?
b. In the cruise configuration, i.e., with all high-lift devices stowed, what value
of C L max, CR do you assume?
c. Do you envision your airplane to have high-lift devices on the trailing edge?
D
Example 7.2
We use the requirements, specified in Example 7.1. Furthermore, we use the following
assumptions for the maximum lift coefficients:
P ROPELLER A IRPLANE For the propeller airplane, we utilize the stall speed require-
ment in landing configuration (VS0 ), given at sea-level conditions assuming standard ISA
temperature. The wing loading required to fulfill this requirement is readily obtained by
substituting all relevant data from Example 7.1 into (7.5). We then find:
WTO
< 1200 N/m2
Sw
have:
WTO /S w
FT
As you can see, the wing loading is independent of the power loading of the airplane, as
this requirement is evaluated assuming the engine is inoperative (EI). We now are going
to set up the matching diagram. We set the bounds for this diagram at WTO /S w = 1900
[N/m2 ] and WTO /P TO = 0.5 [N/W]. For minimum and maximum power loading, we then
Obviously, this becomes a straight vertical line in the matching diagram. This is shown
in Figure 7.3, where the hash markings indicate the side of the boundary for which the
stall-speed constraint is not satisfied.
J ET A IRPLANE For the jet airplane, we have an approach speed of 68 m/s specified at
sea level conditions and a landing mass which is 85% of the take-off mass (i.e., β = 0.85).
We compute the required wing loading by evaluating (7.6):
D
WTO
< 5500 N/m2
Sw
To construct the diagram, we set the bounds of this diagram to WTO /S w = 7000 N/m2
and TTO /WTO = 0.5. For the approach speed constraint, we have:
In the matching diagram, this constraint becomes a straight line again. This is shown in
Figure 7.4, where the hash marking indicates the infeasible side of the approach-speed
boundary.
0.5 0.5
0.4 0.4
stall speed
0.3 0.3
approach speed
0.2 0.2
0.1 0.1
0 0
0 500 1000 1500 0 2000 4000 6000
Wing loading, W TO/S w (N/m 2) Wing loading, WTO/S w (N/m 2)
FT
Figure 7.5: Definition of the landing field length, LLF , according to CS/FAR 23.
The total landing field length depends on the final approach speed VAPP that the air-
plane has when flying over the obstacle height as well as on the deceleration in the air
and on the ground. While there is not much deceleration in the air, the air distance is
dominated by the glide slope during the landing maneuver. Most of the deceleration
takes place on the ground when the brakes are applied. The total landing distance is,
RA
therefore, largely dependent on the number of wheels and tire pressure, the runway type
(grass, gravel, tarmac), the runway condition (dry, wet, icy), the available brake power,
the application of lift dumpers (e.g. spoilers), the application of aerodynamic brakes (e.g.
spoilers), pilot technique, and weather conditions. All of these aspects make it unavoid-
able to use statistics to estimate the landing field length.
Based on References [10] and [13], we relate the stall speed of the airplane in landing
configuration to its landing distance on a dry runway as follows:
(7.9)
where VS0 is the stall speed in landing configuration and C LFL is the landing-field-length
coefficient. We assume C LFL ≈ 0.60 [s2 /m] for CS/FAR-23 airplanes and C LFL ≈ 0.45
[s2 /m] for CS/FAR-25 airplanes. These numbers are subject to statistical validation and
should be used with caution. The lower value of C LFL for CS/FAR-25 airplanes implies
that while their LFL is 67% longer than the sum of the air distance and ground roll dis-
tance, the combination of automated lift dumpers, strong wheel brakes and (suppos-
edly) superior pilot technique results in a shorter field length per unit of stall speed.
Airplanes with variable-pitch propellers or turbofan engines often apply thrust re-
versing to reduce the landing field length and/or to reduce the wear on the wheel brakes.
According to the regulations, these are not to be included in the computation of the mini-
mum landing distance on a dry runway. This is because an engine-inoperative condition
should not impact the minimum landing distance.
We now employ the lift equation to relate the maximum wing loading to the mini-
mum landing distance:
WTO 1 LFL ρC L max
< (7.10)
Sw β C LFL 2
where β = m ML /m MTO and C L max is the maximum lift coefficient in landing configura-
tion. The following example shows how this can be included in the matching diagram.
Example 7.3
We use the requirements and design decisions as specified in Ex. 7.1. We use the same
values for the maximum lift coefficients as in the Example 7.2:
FT
P ROPELLER A IRPLANE For the propeller airplane, we have a landing field length of
LLF = 750 m at sea level conditions. Assuming C LFL = 0.6 and knowing that β = 1, we
find with (7.10):
WTO
Sw
< 1600N/m2
The corresponding constraint line is added to the matching diagram in Figure 7.6.
J ET A IRPLANE For the jet airplane, we have a landing field length of LLF = 1600 m,
RA
β = 0.85, and C LFL = 0.45. Employing (7.10), we find:
WTO
< 5840N/m2
Sw
The corresponding constraint line is added to the matching diagram in Figure 7.7.
In the previous example, you can see that neither for the propeller airplane nor for the
jet airplane does the landing distance become a sizing constraint. The stall speed and
D
A SSIGNMENT 7.3
0.5 0.5
0.4 0.4
stall speed
0.1 0.1
0 0
0 500 1000 1500 0 2000 4000 6000
Wing loading, W TO/S w (N/m 2) Wing loading, WTO/S w (N/m 2)
1
T = D = C D ρV 2 S w (7.11)
D
2
We divide both sides by W and substitute the drag polar (6.9) for C D :
T C D 12 ρV 2 C D 0 12 ρV 2 W /S w
= = + (7.12)
W W /S w W /S w πAe 12 ρV 2
For the propeller-powered airplane, we employ (6.20) to relate the power loading to the
wing loading:
η p P C D 0 21 ρV 3 W /S w
= + (7.13)
W W /S w πAe 12 ρV
Equations (7.12) and (7.13) relate to a balance in thrust and drag at the cruise altitude.
The thrust or power that is produced by air-breathing engines depends on the density of
the air that they are taking in. In addition, the forward speed of the airplane also reduces
the thrust that is produced by the propeller or fan. We refer to these reductions as a
thrust lapse or power lapse. This lapse in thrust or power means that our engines need
more thrust or power at sea level to make sure that there is enough left to meet the cruise
speed requirement. The power lapse is therefore defined as:
P
αP = (7.14)
P SL
where P SL is the power available at sea level. For a jet engine, we define the engine thrust
lapse, αT , as the ratio between the thrust at altitude and Mach number and the thrust at
sea level in static conditions:
T
αT = (7.15)
TSLS
where TO is the abbreviation for “take-off” and SLS for “sea-level static,” implying the
engine is at zero speed. Turbofans are rated by their value of TTO, SLS , although the per-
formance constraints need to be evaluated at nonzero speeds.
WTO
P TO
< ηp
FT
To find the relation between the wing loading and power loading at take-off condi-
tions and maximum take-off mass, we have to include the lapse in power with altitude,
αP , according to (7.23) as well as the mass ratio, β. We take the reciprocal of (7.13) and
1 3
αP C D 0 2 ρVCR
Ã
+
βWTO /S w
β βWTO /S w πAe 12 ρVCR
!−1
(7.16)
RA
For the jet-powered airplane, we take (7.12), adjust the thrust-to-weight ratio for the
thrust lapse, αT , and multiply the weight by the mass ratio, β:
1 2
β C D 0 2 ρV
à !
TTO βWTO /S w
> + (7.17)
WTO αT βWTO /S w πAe 12 ρV 2
CR
In order to use these questions to draw the constraint curves, we first need to compute
the power lapse and thrust lapse of the engine. However, we first need to consider how
the atmospheric properties change with altitude and weather conditions.
D
The static temperature in Kelvin at altitude (h) can be computed using a simplified
version of the international standard atmosphere:
where TSL ISA = 288 K is the sea-level temperature, TTP ISA = 217 K is the temperature at
the tropopause, λ = −0.0065 K/m is the temperature lapse rate with altitude, and ∆T
represents a change in temperature due to weather conditions. The latter allows us to
evaluate the flight performance of airplanes on a hot day, i.e., when ∆T ≥ 15 K. We make
the simplified assumption that the temperature lapse rate of -6.5◦ /km is independent of
the temperature conditions.
The atmospheric pressure (p) in the troposphere and in the stratosphere can be re-
lated to the altitude (h) using the following formulas, respectively:
¶ g
λh − λR
µ
p = p SL ISA 1 + for 0 ≤ h < 11 km (7.20)
T
µ SL ISA
−g (h − h TP ISA )
¶
p = p TP ISA exp for 11 ≤ h < 20 km (7.21)
RTTP ISA
where p SL ISA = 101 kPa is the sea-level pressure under ISA conditions, p TP ISA = 22.6
kPa is the ambient pressure at the tropopause, h TP ISA = 11 km is the altitude of the ISA
ρ=
FT
tropopause, and R = 287 J/kg/K is the gas constant for air. To compute the density, we
p
RT
These equations allow us to compute the temperature, pressure, and density as a func-
tion of altitude (h) and temperature deviation (∆T ).
(7.22)
RA
The pressure and temperature variation with altitude is depicted in Figure 7.8. In
addition, the icons depict the typical altitude where various types of airplanes cruise
efficiently. Between an altitude of 5 km and 6 km, the low pressure can cause hypoxia, a
state where there is insufficient oxygen in the blood leading to a loss of consciousness.
Therefore, a pressurized cabin is required when flying above these altitudes.
sion system. We call this power lapse or thrust lapse, respectively. The power lapse of an
air-breathing engine can be computed quite easily because it is only dependent on the
altitude. The thrust lapse of a jet engine is more elaborate to compute because it also
depends on the flight Mach number. In this section, we present the equations that you
can use to estimate the power lapse and thrust lapse.
P OWER L APSE For turboprop engines and piston engines, Ruijgrok defines the power
lapse with altitude (αP ) as a function of the local density [12]:
¶3
ρ
µ
4
αP = (7.23)
ρ SL ISA
where ρ is the ambient density at altitude and ρ SL ISA = 1.225 kg/m3 is the ambient den-
sity at sea level ISA. Hence, by computing the air density at the cruise altitude, the power
lapse can be computed straight away.
Temperature, T (K)
16
stratosphere
altitude, h (km) 14 business jets
12
tropopause transport jets
10
tem
pr
pe
es
rat
su
8 ur regional turboprops
re
e
6
hypoxia onset
twin engine props
4
troposphere
2
FT
single engine props
0
0 20 40 60 80 100 120
Pressure, p (kPa)
Figure 7.8: Relation between altitude and atmospheric pressure in International Standard Atmosphere. The
airplane icons indicate a typical cruise altitude for various types of airplanes.
RA
T HRUST L APSE The thrust lapse with speed is influenced by two primary factors. First
of all, the increase in flight speed reduces the difference between the exhaust velocity
of the air that is accelerated by the engine and flight speed. This reduces the thrust.
Secondly, an increase in flight speed causes an increase in pressure in the inlet of the
engine, which increases the overall pressure ratio (OPR) of the engine and thereby causes
the exhaust velocity to increase. This increases the thrust. These two effects oppose each
other, and whether the thrust at a particular Mach number is higher or lower than static
thrust depends, among others, on the bypass ratio B of the engine. Therefore, we need a
model for αT that is sensitive to B , M , and h.
D
Before we can present the model of αT , we first need to introduce the total tempera-
ture and total pressure. Let us first define the Mach number:
V
M=p (7.24)
γRT
As you can see, the Mach number is solely determined by the ambient temperature and
is therefore a function of h and ∆T through Equations (7.18) and (7.19). The total tem-
perature is the sum of the static temperature and the dynamic temperature. If the flow
is brought to a standstill under isentropic conditions2 the kinetic energy is converted to
heat, raising the temperature. The total temperature can be shown to be related to the
2 An isentropic process is both reversible and adiabatic, i.e., without the addition or removal of heat.
γ−1 2
µ ¶
Tt = T 1 + M (7.25)
2
where γ is the ratio of specific heat coefficients for air, and T is the ambient temperature
at the cruise altitude. With γ = 1.4, this simplifies to: Tt = T (1 + 0.2M 2 ). Similarly, we can
express the total pressure, p t , as a function of Mach number (M ) as follows:
¶ γ
γ − 1 2 γ−1
µ
pt = p 1 + M (7.26)
2
δt =
θt =
FT pt
p SL ISA
Tt
TSL ISA
With δt and θt properly defined, we can now dive deeper into the workings of a turbo-
fan engine in order to properly model it. A turbofan comprises a fan, a compressor, a
(7.27)
(7.28)
RA
combustion chamber, and a turbine. Up to a threshold value of the total temperature,
the maximum thrust of the turbofan is limited by the maximum pressure ratio that the
compressor can provide. Beyond the maximum T t threshold, the materials that are used
to construct the turbine blades limit the performance of the gasturbine. The engine con-
trols are therefore controlled in such a way that beyond a particular value of θt , the throt-
tle is reduced to protect the turbine blades. This value is termed θt break and is a design
choice for engine designers. According to Ref. [4], modern engines have values of θt break
ranging between 1.06 and 1.08.
Mattingly provides the following empirically derived equations for the thrust lapse of
D
You can see that for low-bypass-ratio turbofans, the thrust lapse is actually positive with
Mach number as long as θt ≤ θt break and h is small, which means the word “lapse” is per-
haps not the most appropriate way to address the change in thrust due to Mach number.
You can also observe that for higher bypass ratios, the thrust lapse with Mach number
gets stronger, particularly when θt ≥ θt break . Finally, you should recognize that the thrust
lapse gets stronger with increasing bypass ratio B .
A SSIGNMENT 7.4
In this assignment you will calculate the thrust or power lapse at cruise altitude
for the engine of your airplane.
a. What is the cruise altitude of your airplane?a
b. Compute the power lapse (αP ) or thrust lapse (αT ) of your engine in cruise
conditions using the cruise altitude and the bypass ratio of your choice.
a If this has not been specified in the requirements, you have to choose it
ing diagrams.
Example 7.4 FT
Equation (7.17) relates the wing loading to the thrust-to-weight ratio. These equations
are employed in the subsequent example to draw the constraint curves in the two match-
P ROPELLER A IRPLANE For the propeller-powered airplane we have cruise speed re-
quirement of VCR = 70 m/s at an altitude of 1800 m. As this airplane has an electric
D
motor, there is no power lapse with altitude, i.e. αP = 1. Also, the mass of the airplane
is constant: β = 1. At an altitude of 1800 m, we employ (7.20) and (7.18) to find that
p = 81.5 kPa and T = 276 K, respectively. With the equation of state (7.22), we then find
that ρ = 1.03 kg/m3 . With C D 0 = 0.0260, e = 0.71, A = 9, and η p = 0.80, we can compute
the power loading as a function of wing loading using (7.16):
WTO WTO /S w −1
· ¸
4577
< 0.8 +
P TO WTO /S w 721
With this equation, we can substitute a range of values for WTO /S w and find the corre-
sponding constraint value for WTO /P TO :
0.5 0.5
cru
Thrust-to-weight ratio, T TO/W TO (-)
Power loading, W TO/P TO (N/W)
i se
sp
0.4 0.4
ee
stall speed
d
0.3 0.3
d
0.1 ee 0.1
sp
se
crui
0 0
0 500 1000 1500 0 2000 4000 6000
Wing loading, W TO/S w (N/m 2) Wing loading, WTO/S w (N/m 2)
We can plot these points in the matching diagram to form a curve that bounds the max-
imum power loading as is done in Figure 7.9.
J ET A IRPLANE For the jet airplane we have M CR = 0.8 at an altitude of h CR = 11 km.
The mass ratio for which this requirement is investigated is β = 0.95. We compute the
FTFigure 7.10: Matching diagram with cruise speed re-
quirement for jet airplane
RA
pressure and temperature with (7.20) and (7.19) with ∆T = 0, respectively, to find: p = 23
kPa and T = 217 K. With (7.22) we find an ambient density at the cruise altitude of ρ =
0.36 kg/m3 . Using the value of M CR and T , we can compute the total temperature and
total pressure in cruise conditions using (7.25) and (7.26), respectively, to find p t = 41
kPa and Tt = 256 K. With (7.28) and (7.27), we can compute the total pressure ratio and
total temperature ratio, respectively: δt = 0.40 and θt = 0.89. Realizing that θt < θt break ,
we can now employ (7.31) for B = 10 to find the thrust lapse in cruise conditions is αT =
0.20. Finally, we use (7.24) to compute the cruise speed at M CR = 0.80 and h CR = 11 km:
D
VCR = 236 m/s. With C D 0 = 0.0180, e = 0.80, and A = 8, we can compute the thrust-to-
weight ratio as a function of wing loading using (7.12):
TTO WTO /S w
· ¸
192
> 4.82 +
WTO WTO /S w 214 · 103
With this equation, we can substitute a range of values for WTO /S w and find the corre-
sponding constraint value for WTO /P TO :
We can subsequently plot the corresponding constraint curve in our matching diagram
(Fig. 7.10). To satisfy the cruise-speed constraint, the thrust-to-weight ratio for any wing
loading should be above this curve.
The example above shows that the cruise-speed requirement translates to a power or
thrust constraint for the airplane. We can see that for lower wing loading, more thrust or
power is required to sustain the cruise speed at the cruise altitude. Through the altitude,
also the atmospheric properties play an important role in the estimation of the thrust
(or power) lapse of the engine. In the next assignment, you will add the cruise-speed
constraint to your own matching diagram.
A SSIGNMENT 7.5
In this assignment, you will add the constraint curve for the cruise speed to your
matching diagram.
a. For the cruise altitude and cruise speed of Assignment 7.4, list the αP or αT
of your airplane.
b. If a cruise Mach number is specified, compute the cruise speed VCR at your
cruise altitude using (7.24).
FT
c. Tabulate the power loading or thrust-to-weight ratioa as a function of wing
loading, similar to Example 7.4.
d. Plot the constraint curve of the cruise speed in your matching diagram.
a whichever one is applicable
c = V sin γ (7.33)
We acknowledge that during a steady climb, there is no acceleration, and the forces on
the airplane are balanced. This is notionally shown in the free-body diagram of the air-
plane in Figure 7.11. By equating the forces in the direction of flight, we find:
T − D − W sin γ = 0 (7.34)
Combining (7.33) and (7.34) gives us the following expression for the climb rate:
T − D ηp P D
c =V = −V (7.35)
W W W
where we employ (6.20) to relate the thrust to the power.
c= V·sin γ
T
Climb angle γ
Horizon
D Wcos γ
γ
W
flight direction:
L = W cos γ ≈ W for γ ≪ 1
FT
Figure 7.11: Free-body diagram of forces acting on an airplane in climbing flight.
Based on (7.35), we first investigate for which lift coefficient the climb rate of a pro-
peller airplane is maximized. We consider the equilibrium of forces perpendicular to the
where γ is measured in radians. Since we have the lift equation, we can employ (6.7)
(7.36)
RA
such that V comes to the left-hand side of the equation:
s
W 2 1
V= (7.37)
Sw ρ CL
We can substitute this expression for V in (7.35) together with D/W ≈ D/L = C D /C L to
find: s
ηp P CD W 2
D
c= − 3/2 (7.38)
W CL S w ρ
By inspecting this equation, we can deduce that the climb rate is maximized when the
fraction C D /C L3/2 is minimized. We can analytically show that this fraction is minimized
when: q
C L = 3C D 0 πAe and C D = 4C D 0
Substituting this in (7.38) yields the equation for the maximum rate of climb:
1/4
s
ηp P 4C D
0 W 2
c= − (7.39)
W (3πAe)3/4 Sw ρ
To arrive at the function that relates the power loading to the wing loading in take-off
conditions, we have to let P = αP P TO and W = βWTO . We can rearrange (7.39) to express
When we employ (7.40) we substitute for c max the climb rate requirement. In order to
evaluate this equation, we also need a value of the density, ρ. In other words, we need to
specify an altitude along with a value for the climb rate.
Now, let us make a similar derivation but now for the jet airplane. Starting from
(7.35), substituting (7.37), and approximating D/W ≈ D/L = C D /C L we have:
¶s
T CD W 2 1
µ
c= − (7.41)
W CL Sw ρ CL
Contrary to the propeller airplane, we cannot analytically derive a lift coefficient that is
independent of the thrust-to-weight ratio for which the climb rate is maximized. Ruij-
CL =
q
FT
grok shows that for a pure jet (i.e. B = 0), the lift coefficient for which the rate of climb
is maximized is less than the lift coefficient for which the lift-to-drag ratio is maximized
[12]. Since we consider a turbofan-powered jet airplane, we propose to use:
C D 0 πAe and C D = 2C D 0
With T = αT TTO and W = βWTO , we can rearrange (7.42) and formulate the constraint
curve for the rate-of-climb requirement:
s
TTO β c2 ρq 4C D 0
> C D 0 πAe + (7.43)
WTO αT βWTO /S w 2 πAe
D
In the next example, we will show how the rate-of-climb constraint for the propeller air-
plane and the jet airplane appear in the matching diagram.
Example 7.5
We use the requirements and design decisions as specified in Ex. 7.1. For the aerody-
namic characteristics in the clean configuration, we assume the following values:
P ROPELLER A IRPLANE For the propeller airplane we have c = 2.0 m/s at sea level. We
perform the following steps:
1. Since we are at sea level, we obtain the density by evaluating (7.18), (7.20), and
(7.22), knowing that ∆T = 0. This results in ρ = 1.23 kg/m3 .
2. We know that β = 1 and because we have electric motor αP = 1. So there is no need
to compute the power lapse with altitude.
3. We can insert these values in (7.40) along with all the numerical values for the pro-
peller airplane of Example 7.1. This results in the following constraint equation:
WTO 0.80
> p
P TO 2.0 + 0.095 WTO /S w
4. We can now tabulate the values of power loading for a range of wing loading values:
5. The resulting constraint curve is added to the matching plot in Figure 7.12.
You can see that the climb-rate curve is coincidentally crossing the stall-speed constraint
at exactly the same point as the cruise-speed constraint. It is, therefore, not actively
bounding the feasible design space in this example.
0.5 FT 0.5
cru
Thrust-to-weight ratio, T TO/W TO (-)
Power loading, W TO/P TO (N/W)
ise
sp
ra te
d
0.3 0.3
d
0.1 ee 0.1
e sp
cr uis
D
0 0
0 500 1000 1500 0 2000 4000 6000
Wing loading, W TO/S w (N/m 2) Wing loading, WTO/S w (N/m 2)
Figure 7.12: Matching diagram with added rate-of- Figure 7.13: Matching diagram with added rate-of-
climb requirement for propeller airplane climb requirement for jet airplane
J ET A IRPLANE For the jet airplane, we have a climb rate requirement of c = 0.5 m/s at
h = 11, 500 m altitude. We perform the following steps:
1. We compute the atmospheric properties at h = 11, 500 m by employing (7.19),
(7.20), and (7.22), knowing that ∆T = 0. We have: T = 217 K, p = 23 kPa and
ρ = 0.36 kg/m3 .
2. We
p compute the p lift coefficient for which we have the best rate of climb: C L =
C D 0 πAe = 0.0180 · π · 8 · 0.80 = 0.60.
3. We compute the speed at which we have the best rate-of-climb using (7.37) as a
function of wing loading, substituting W = 0.95WTO in accordance with the speci-
fication. Since the best speed varies with wing loading, we choose a range of wing
loading values and tabulate the associated best climb speed for each wing loading
(see table below).
4. Using (7.24), we compute the Mach number as a function of the wing loading and
tabulate the best Mach number for a range of wing loading values (see table be-
low).
5. Using (7.28) and (7.27), we compute the value of the total temperature ratio and
total pressure ratio, respectively. We tabulate these values below.
6. Noting that θt < 1.08 for all wing loading values, we employ (7.31) to compute the
trust lapse, αT , corresponding to each wing loading value below.
7. We can now employ (7.43) to compute take-off thrust-to-weight ratio correspond-
ing to each wing loading value. This is tabulated in the bottom row of the table
below.
WTO /S w
V
M
θt
δt
αT
TTO /WTO
(m/s)
(-)
(-)
(-)
(-)
(-)
FT
(N/m2 ) Eq.
(7.37)
(7.24)
(7.28)
(7.27)
(7.31)
(7.43)
1000
97
0.33
0.78
0.23
0.16
0.40
3000
168
0.57
0.82
0.28
0.16
0.37
5000
217
0.73
0.87
0.34
0.17
0.34
7000
256
0.87
0.91
0.41
0.19
0.31
RA
8. Using the tabulated values of wing loading and power loading, we add the con-
straint curve for the climb rate to the matching diagram of the jet airplane in Figure
7.13.
You can see that the climb-rate constraint cuts through the cruise-speed constraint in
this example and is therefore actively constraining the minimum thrust-to-weight-ratio
values for WTO /S w > 3400 N/m2 .
The previous example shows that there is a few extra steps required to compute the con-
straint curve for the jet airplane compared to the propeller airplane. This is due to the
D
more elaborate thrust-lapse relations that we employ for turbofan engines. In the fol-
lowing assignment, you will add the constraint curve to your matching diagram.
A SSIGNMENT 7.6
In this assignment, you will construct the constraint curve for the rate-of-climb
requirement. For propeller airplanes, skip steps c-g as they only pertain to jet
airplanes.
a. State the climb rate requirement for your airplane and specify c, h, β, and
∆T .
b. Compute the atmospheric properties T , p, and ρ for the given h and ∆T .
c. Compute the lift coefficient for which the highest rate of climb occurs.
d. For a chosen range of wing loading values, compute the speed for which
FT
In the foregoing examples, we have considered the climb rate requirement assuming
all engines are operative. A well-known climb rate requirement is the definition of a
minimum flight altitude in case of a one-engine inoperative (OEI) situation. To assess
this requirement, the power loading or thrust-to-weight ratio needs to be corrected for
the inoperative engine. For the propeller airplane with OEI, (7.40) would be modified as
follows:
1/4 !−1
RA
à s
WTO Ne − 1 αP 4C D βWTO 2
< ηp c+ 0
(7.44)
P TO Ne β (3πAe)3/4 Sw ρ
where Ne is the number of engines (or motors) that you have chosen. For the jet airplane
with OEI, (7.43) is modified according to:
s
TTO Ne β c2 ρq 4C D 0
> C D 0 πAe + (7.45)
WTO Ne − 1 αT βWTO /S w 2 πAe
D
c T − D η p P − DV
= = (7.46)
V W VW
Note that the fraction c/V is our climb gradient requirement. In other words, we leave
this fraction intact in the following derivation and substitute a numerical value for it
when we evaluate our constraint at the end of the process.
For the propeller airplane, we know that D/W ≈ D/L = C D /C L . Therefore, (7.46) can
now be written as:
c ηp P 1 C D
= − (7.47)
V W V CL
We can insert (7.37) to substitute the speed to find:
s
c ηp P ρ C L CD
= − (7.48)
V W 2 W /S w C L
CL =
C L max
1.1
FT
The climb gradient is maximized when C L is as high as possible. Practically, we are
bound by the value of the maximum lift coefficient that we assume for the configura-
tion that is chosen to evaluate the climb-gradient requirement. Furthermore, we apply a
safety factor of 1.1 to make sure we evaluate this constraint at a small margin to the stall:
and C D = C D 0 +
C L2max
1.12 πAe
(7.49)
RA
Using the thrust lapse (αP ) and the mass ratio (β), we can rearrange (7.48) to find the
constraint curve for the climb gradient requirement:
¶s
WTO β ρ CL
µ
1
< ηp (7.50)
P TO αP c/V +C D /C L 2 βWTO /S w
where C D and C L are given by (7.49). In case we wish to evaluate the climb gradient
requirement of a multi-engine airplane in an OEI condition, (7.50) can be amended as
D
follows:
¶s
WTO Ne − 1 β ρ CL
µ
1
< ηp (7.51)
P TO Ne αP c/V +C D /C L 2 βWTO /S w
where Ne is the number of engines (or motors).
For the jet airplane, we also start from (7.46). We apply D/W ≈ D/L = C D /C L and
rewrite the climb-gradient equation as follows:
c T CD
= − (7.52)
V D CL
We can substitute the expressions of the lift and drag coefficients in (7.52) to find:
s
c T CD0
= −2 (7.53)
V W πAe
By letting W = βWTO and T = αT TTO , we can now express the take-off thrust-to-weight
ratio as a function of the take-off wing loading:
às !
TTO β c CD0
= +2 (7.54)
WTO αT V πAe
We see that for the climb-gradient requirement, the take-off thrust-to-weight ratio seems
to be independent of the wing loading. However, the thrust lapse, αT , is dependent on
the Mach number. Therefore, there is still a relationship with the wing loading through
the equation relating velocity to lift (7.37). If we consider an OEI condition, we need to
increase the required thrust-weight ratio according to:
TTO
=
WTO Ne − 1 αT V
are specified in the certification documents. For many of them, the airplane is specified
to be in take-off or landing configuration with either the landing gear up or down. For
RA
each configuration, the airplane has a different drag polar. You can imagine that when
the flaps and landing gear are deployed, the airplane produces more drag. So, before
we can compute the constraint curve in the matching diagram, we first need to estimate
how the zero-lift drag coefficient and the Oswald factor change as a function of the flap
deflection and landing gear deployment.
7.6.2. E STIMATING THE D RAG P OLAR FOR VARIOUS F LAP AND L ANDING
G EAR C ONFIGURATIONS
D
So far, we have considered the Oswald factor for the cruise configuration of the airplane.
When an airplane deploys its flaps (see Figure 7.14) to increase the maximum lift co-
efficient and thereby reduce its stall speed, the zero-lift drag coefficient of the airplane
increases. In addition, the lift-dependent parasite drag parameter (ψ) increases while
the span efficiency (ϕ) reduces. The net effect on the Oswald factor is that it typically
increases with flap deflection for low-speed flight. Obert proposes the following change
in Oswald factor as a function of flap deflection, δf [6]:
where ∆f e is the change in Oswald factor due to flap deflection (δ f ), with δf being mea-
sured in degrees. We can use this equation to estimate the Oswald factor for any flap
deflection angle. To achieve the highest maximum lift coefficient, the maximum flap
deflection is obtained in the landing configuration. So, in designing the airplane you
should decide what flap deflection this will be. Secondly, you can choose intermediate
flap deflection angles for the take-off of the airplane that range anywhere from zero de-
flection to maximum deflection.
FT
RA
Figure 7.14: Boeing 737 with flaps partially extended during approach. Photo: Manic Nirvana.
The zero-lift drag coefficient also depends on the flap deflection. The change in drag
depends on the size of the flaps, the type of the flaps, and the amount of flap deflection.
Since we do not have that information at the present time, we use a very rudimentary re-
lation that assumes for every degree of flap deflection, we incur a 13-count3 drag penalty
D
Naturally, the equations above are fairly crude, and in subsequent iterations of the de-
sign, many of these values are replaced by analysis results (see Chapter 10). However,
it allows you to construct a series of drag polars for the airplane, which you can use in
subsequent steps of the design process.
3 A drag count is 1/10,000th of a drag coefficient
Example 7.6
In this example, we produce the drag polars for one airplane in various configurations.
This airplane has jet engines that are mounted on the fuselage (7.56). We have selected
an aspect ratio of A = 8, a take-off flap deflection of δf, take-off = 15◦ , and a landing flap
deflection of δf, landing = 35◦ . Furthermore, we have a zero-lift drag coefficient and Os-
wald factor in cruise configuration of C D 0 = 0.0180 and e = 0.80. We consider four con-
figurations: cruise, take-off (1), take-off (2), and landing. For take-off (1), the landing
gear is up, while for take-off (2) and landing configuration, the landing gear is down. We
perform the following tasks: For each configuration, we compute the zero-lift drag coef-
ficient and the Oswald factor. We employ (7.56),(7.57), and (7.58). The resulting values
for the aerodynamic coefficients are tabulated below:
A SSIGNMENT 7.7
TO
L
L
↓
↑
↓
FT
15
35
35
0.0575
0.0635
0.0835
In this assignment, you will construct the drag polars for your airplane.
0.87
0.96
0.96
RA
a. Using Eq. 6.17, compute the Oswald factor of your airplane in cruise con-
figuration.
b. What is the maximum amount of flap deflection that you choose for your
airplane in the landing condition?
c. What intermediate flap deflection do you choose for the take-off condi-
tion?
d. Using Eq. 7.57 and 7.58 compute the zero-lift drag coefficient in take-off
configuration and in landing configuration.
e. Using Eq. 7.56, compute the Oswald factor in the take-off configuration
D
Example 7.7
We use the requirements as specified in Ex. 7.1. For the climb gradient requirement,
the flaps of the airplane are in the take-off configuration. For the propeller airplane, the
landing gear is fixed, so it is down. The jet airplane is able to retract the landing gear,
and therefore, its climb gradient is evaluated in the gear-up configuration. The following
coefficients apply for either airplane:
FT
diagram, we perform the following steps:
↑ 0.038 0.87 2.1
P ROPELLER A IRPLANE For the propeller, we have c/V = 0.083 at sea level ISA with
the flaps in take-off configuration. To compute the constraint curve in the matching
6. We can now add the climb-gradient constraint curve to our matching diagram.
This has been done in Figure (7.15).
You can see that the resulting climb-gradient constraint curve lies below the cruise-
speed constraint curve for wing loading values above 1000 N/m2 . This implies that this
constraint is actively bounding the feasible design space for this propeller airplane.
J ET A IRPLANE For the jet airplane, we consider a more elaborate climb-gradient re-
quirement: c/V = 0.024 at sea level but with a temperature increase of ∆T = 15◦ above
ISA conditions. Furthermore, we consider one engine to be inoperative, the mass equal
to the maximum take-off mass, and the airplane being in the take-off configuration.
With the following steps, we construct the constraint curve:
0.5 0.5
cru
Thrust-to-weight ratio, T TO/W TO (-)
Power loading, W TO/P TO (N/W)
i se
sp
0.4 0.4 climb
ee
stall speed
ra te
d
b
grad
d
0.1 ee 0.1
sp
se
crui
0 0
0 500 1000 1500 0 2000 4000 6000
Wing loading, W TO/S w (N/m 2) Wing loading, WTO/S w (N/m 2)
1. First, we compute the density at sea level. We employ (7.18) and (7.20) to find
that T = 303 K and p = 101 kPa. Then, with (7.22), we compute the density to be
FT
Figure 7.16: Matching diagram with added climb-
gradient requirement for jet airplane
RA
ρ = 1.16 kg/m3 .
2. We compute the value of the lift and drag coefficient for maximum climb gradient
for the take-off configuration (i.e., C D 0 = 0.038 and e = 0.87):
3. We compute the speed for which we have the best climb gradient by employing
(7.37), for a range of wing-loading values and tabulate that in the table below.
4. Using (7.24), we compute the Mach number associated with each velocity value
and ambient temperature and tabulate that below.
5. Using (7.28) and (7.26), we compute the total temperature ratio, θt , and total pres-
sure ratio, δt associated with the Mach numbers computed in the previous step.
We tabulate this below.
6. Acknowledging that θt < 1.08 for all but the highest wing-loading value, we employ
(7.31) and (7.32) to compute the thrust lapse for each Mach number. This is also
tabulated below.
7. We can now employ (7.54) to compute the thrust-to-weight ratio as a function of
wing loading:
8. We plot the tabulated values above as a constraint curve in the matching diagram
of Figure 7.16.
You can see that the curve that we drew for the climb-gradient constraint lies below the
climb-rate constraint for all wing-loading values left to the approach-speed constraint.
It is, therefore, not an active constraint in this example.
A SSIGNMENT 7.8
FT
In this assignment, you will construct the constraint curve for the climb-gradient
requirement. For propeller airplanes, skip steps f-i as they only pertain to jet
airplanes and serve to compute the thrust lapse.
a. State the climb gradient requirement of your airplane: c/V , h, β and ∆T .
Also, state whether the requirement is to be fulfilled in AEO condition or in
OEI condition and the configuration of the airplane.
b. Based on the configuration, state the value of the zero-lift drag coefficient
RA
C D 0 and Oswald factor e that are applicable.
c. Compute the ambient temperature and density that pertain to the climb
gradient requirement.
d. Compute the lift coefficient for which the climb gradient is maximized. Use
an appropriate margin to the stall speed if this is applicable.
e. Compute the drag coefficient for which the climb gradient is maximized.
f. Using a chosen range of wing-loading values, compute the speed for which
the climb gradient is maximized corresponding to each wing-loading value
D
FT
the air distance is dependent on how steep the airplane is able to climb. In Example
3.12, we have seen that the climb gradient is dependent on the lift-to-drag ratio, and in
Example 7.6, we have seen that the highest lift-to-drag ratio is achieved in clean con-
figuration (i.e. with the flaps stowed). To minimize the combined ground-roll distance
and air distance, the airplane therefore uses an intermediate flap setting that increases
the maximum lift coefficient significantly but with an acceptable penalty in drag. This is
what we refer to as the take-off configuration.
RA
35 or 50 ft
Figure 7.17: Definition of the take-off field length, LTO , according to CS/FAR 23/25.
The take-off maneuver with associated distances, speeds, and climb angles is more
elaborately explained in the airworthiness regulations (CS/FAR-23: paragraphs 51-61;
CS/FAR-25 paragraphs 105-115). You can imagine that for multi-engined airplanes, the
take-off distance in the OEI condition is different from the take-off distance in the AEO
condition. For every multi-engined airplane, a decision speed (V1 ) needs to be estab-
lished. If an engine failure occurs below V1 , the pilot needs to break and come to a
standstill. If an engine failure occurs beyond V1 , the pilot needs to continue the take-
off in OEI condition. So in the case of a continued take-off (CTO), the available thrust
changes during the take-off maneuver. In our analysis of the take-off maneuver, we first
consider the case for the AEO condition, which we subsequently augment to consider a
CTO. The take-off safety speed at the obstacle height is termed V2 . We will use V2 and the
associated C L 2 and TV2 as the reference speed, lift coefficient and thrust, respectively.
Contrary to all of the previous flight performance requirements, the take-off maneu-
ver is not steady, meaning the velocity changes with time. We follow the treatment by
Torenbeek [13], who considers the take-off field length (LTO ) to be the sum of the ground
run(LRUN ) and the air distance(LAIR ) for jet airplanes:
The ground run is terminated at the lift-off speed, VLOF . We assume that the difference
between VLOF and V2 is small and can be neglected. If we assume a constant acceleration
during the ground run due to an average acceleration force F̄ acc (i.e., the resultant of
thrust, ground friction, and drag), we can deduce that:
WTO F̄ acc
LRUN = V2 with k T = (7.60)
2g k T TV2 2 TV 2
The parameter k T relates the thrust at V2 to the average acceleration force. Its value is
assumed to be 0.85 for jet airplanes. Applying the lift equation (7.37), we can substitute
V2 in (7.60) and find:
FT
LRUN =
WTO2
ρg S wC L 2 k T TV2
where ρ is the density at the take-off altitude.
(7.61)
The air distance is measured from the point of lift-off to the obstacle height, h 2 . We
assume that the airplane follows a circular take-off path during this maneuver ending
with climb angle γ2 at the obstacle height. The air distance that is covered during this
RA
maneuver is equal to the distance covered during a steady climb over twice the obstacle
height, as is schematically shown in Figure 7.18. Knowing that the climb angle γ2 ≪ 1
when measured in radians, we can deduce the relation between the climb angle, the
thrust-to-weigh ratio, and the lift-to-drag ratio, similar to (7.46):
¶ ¸−1
TV 2 CD
· µ
2h 2
LAIR = = 2h 2 − (7.62)
γ2 WTO C L V2
The drag coefficient at V2 can be found by substituting C L 2 in the drag polar of the air-
D
b path γ2
dy clim
stea
γ2 2h2
h
ff pat
u lar take-o h2
circ
Figure 7.18: Approximation of air distance Lair by means of steady climb angle γ2 ,
There exists one value of C L 2 for which this take-off field length is minimized. However,
Torenbeek analytically derives that for high-aspect-ratio wings, the minimum take-off
field length is insensitive to the actual value of C L 2 due to the trade-off between ground
run and air distance [13]. The thrust-to-weight ratio at V2 for a minimum take-off field
length of LTO is then approximated as follows:
take-off configuration.
TV 2
WTO
= 1.15
s
FTWTO /S w
LTO kT ρg πAe
+
4h 2
LTO
It should be noted that in this expression, the Oswald factor e should be the one for the
For jet airplanes, we now consider the thrust lapse due to speed, altitude and tem-
(7.66)
RA
perature to arrive at the take-off thrust-to-weight ratio constraint:
s
TTO WTO /S w 4h 2
> 1.15αT + (7.67)
WTO LTO kT ρg πAe LTO
For multi-engined jet airplanes, the take-off field length in OEI condition can be approx-
imated as follows:
s
D
TTO Ne WTO /S w Ne 4h 2
= 1.15αT + (7.68)
WTO Ne − 1 LTO k T ρg πAe Ne − 1 LTO
You might have observed that with this derivation, we have reduced a complicated un-
steady maneuver into a single equation relating the thrust-to-weight ratio to the wing
loading. We therefore emphasize that this is an approximation and that (7.64) should be
used to check whether the take-off constraint is actually met. Secondly, we still need to
choose a value of C L 2 in order to calculate V2 as a function of wing loading and subse-
quently determine the value of αT as a function of wing loading.
For propeller airplanes, we need to know V2 in order to evaluate the required power at
the obstacle height. Since the climb gradient of a propeller airplane is maximized when
the lift coefficient is as high as possible, we choose C L 2 to be as high as possible as well.
The airworthiness regulations give a lower bound to V2 with respect to the stall speed in
the take-off configuration. CS/FAR-23.51 specifies that V2 ≥ 1.2Vs1 and CS/FAR-25.107
specifies that V2 ≥ 1.13Vs1 , where Vs1 is the stall speed in a particular configuration (other
than the landing configuration). In this case, VS1 is the stall speed in the take-off config-
uration. Using the ratios between the take-off safety speed and the stall speed, we can
compute C L 2 as follows:
Vs1 2
µ ¶
C L2 = C L max (7.69)
V2
where we use the minimum value for the speed ratio specified by the regulations, i.e.,
1.2 and 1.13 for CS-23 and CS-25 airplanes, respectively. Using the lift equation (7.37),
we can relate V2 to the wing loading and C L 2 . And, with the relation between power and
thrust (6.20) and (7.66) we can establish the power loading constraint as follows:
" s #−1 s
WTO WTO /S w 4h 2 C L2 ρ
< αP 1.15 + (7.70)
P TO LTO kT ρg πAe LTO WTO /S w 2
where αP accounts for the power lapse of air-breathing engines with altitude. Note that
we are implicitly assuming that β = 1 as we are considering the take-off constraint at
WTO
P TO
"
< αP 1.15
s
Ne
FT
maximum take-off weight. Similar to the multi-engined jet airplane, we can modify
(7.70) to account for an OEI condition as follows:
WTO /S w
+
Ne 4h 2
Ne − 1 LTO k T ρg πAe Ne − 1 LTO
#−1 s
C L2 ρ
WTO /S w 2
(7.71)
where Ne is the number of engines or motors that has been chosen. While the value of
1.15 in (7.70) and (7.71) stems from the analysis of jet airplanes in Ref. [13], we propose
RA
to also use it for propeller airplanes, for the lack of a better source. However, this number
is subject to statistical validation. The following example demonstrates how the take-off
constraint is added to the matching diagram.
Example 7.8
We use the requirements, design decisions, and assumptions as specified in Ex. 7.1. In
addition, we estimate the following aerodynamic characteristics:
D
4. We employ (7.70) to compute the power loading values as a function of wing load-
ing knowing that A = 9 and assuming that k T = 0.85:
WTO ³ p ´−1 0.85
< 0.0029 WTO /S w + 0.081 p
P TO WTO /S w
5. We substitute a range of values for WTO /S w in the equation above and tabulate the
resulting values for WTO /P TO in the table below:
0.5
FT 0.5
cru
Thrust-to-weight ratio, T TO/W TO (-)
Power loading, W TO/P TO (N/W)
ise
sp
0.4 0.4
tak
climb
ee
stall speed
ra
e-o
te
d
ff f
b gr
i eld
l en
cli gth
en
nt
gth
m dl
approach speed
RA
ra
0.2 te 0.2 ff f
ke-o
ta
d
0.1 ee 0.1
sp
se
c rui
0 0
0 500 1000 1500 0 2000 4000 6000
Wing loading, W TO/S w (N/m 2) Wing loading, WTO/S w (N/m 2)
D
Figure 7.19: Matching diagram with added take-off Figure 7.20: Matching diagram with added take-off
field requirement for propeller airplane field requirement for jet airplane
J ET A IRPLANE Our jet airplane has a take-off field length requirement of 2500 m at
an altitude of 1600 m and a temperature increase of 15◦ with respect to ISA conditions.
We have a twin-engined airplane, and we consider the take-off field length to be equal
to the continued take-off distance with one engine inoperative as given by (7.65). To
compute the thrust-to-weight ratio constraint for the take-off field length, we perform
the following steps:
1. Using (7.18), (7.20), and (7.22) we compute the ambient temperature and density
at the take-off condition: T = 293 K, ρ = 0.98 kg/m3 .
2. The airplane is in take-off configuration with the landing gear down, so the zero-
lift drag coefficient, Oswald factor, and maximum lift coefficient are C D 0 = 0.058,
e = 0.87, and C L max = 2.1 respectively.
WTO /S w
V2
M
θt
δt
αT
TTO /WTO
(N/m2 )
(m/s)
(-)
(-)
(-)
(-)
(-)
FT Eq.
(7.37)
(7.24)
(7.28)
(7.27)
(7.31)
(7.68)
1000
35
0.10
1.02
0.83
0.68
0.16
3000
61
0.18
1.03
0.85
0.65
0.26
5000
78
0.23
1.03
0.87
0.63
0.32
7000
93
0.27
1.04
0.89
0.62
0.38
RA
9. We construct the constraint curve for the take-off field length in the matching dia-
gram of Figure 7.20.
The resulting constraint curve exceeds the constraint curve of the climb-rate require-
ment when WTO /S w > 4700 N/m2 . So the take-off constraint actively bounds the feasi-
ble design space for the jet airplane. You might have also noted that the thrust lapse at
V2 can be rather large for this jet airplane. The combination of a 15-degree temperature
rise, an altitude of 1600 m, and a speed of 79 m/s results in a thrust reduction of 37%
(αT = 0.63) for a wing loading of 5000 N/m2 .
D
We have demonstrated how the take-off field length constraint can be added to the match-
ing diagram. The mathematical formulation of the constraint curves is based on the
optimization approach. To check if the constraint is actually satisfied, you can employ
(7.64) or (7.65) once you have selected the take-off power loading or take-off thrust-to-
weight ratio along with the wing loading. In the following assignment, you are going to
construct the take-off constraint curve for your matching diagram.
A SSIGNMENT 7.9
In this assignment, you will construct the constraint curve for the take-off field-
length requirement. For propeller airplanes, skip steps e-h as they only pertain
to jet airplanes.
a. State the take-off field length requirement and specify LTO , h, ∆T , and
whether you consider the take-off distance in AEO condition or a contin-
ued take-off after an OEI condition has occurred.
b. Based on the take-off configuration, state the values of the zero-lift drag
coefficient C D 0 and Oswald factor e that are applicable.
c. Compute the atmospheric properties for the take-off condition: T , p, and
ρ.
d. Using the minimal allowed margin to the stall speed, compute the value of
C L2 .
e. For a chosen range of wing loading values, compute the associated take-off
safety speed V2 .
f. For the computed values of V2 , calculate the corresponding Mach num-
bers.
g. Calculate the total temperature ratios θt and total pressure ratios δt asso-
ciated with the Mach numbers computed in the previous step.
FT
h. For the chosen range of wing loading values, compute the thrust lapse αT
that corresponds to the values of θt and δt of the previous step.
i. Compute the take-off power loading or take-off thrust-to-weight ratioa for
the take-off field length requirement for the chosen range of wing loading
values.
j. Plot the take-off field length constraint curve in the matching diagram.
a whichever one is applicable
RA
7.8. C HOOSING A D ESIGN P OINT IN THE M ATCHING D IAGRAM
When all flight performance constraints have been transformed into constraint curves in
the matching diagram, we need to select a single point in this diagram which determines
the size of the wing as well as the power plant. We call this point the design point. The
design point represents a combination of wing loading and power loading for a propeller
airplane or wing loading and thrust-to-weight ratio for a jet airplane. In this section, we
D
are going to demonstrate how you can select this design point.
With all the constraints plotted in the matching diagram, we need to make sure that
the selected design point lies in the feasible design space as we have schematically shown
in Figure 7.1. Secondly, we would like to choose a design point that optimizes our de-
sign objective. This is a bit problematic at this stage: we have not clearly defined how
each possible design objective relates to the wing loading and power loading (or thrust-
to-weight ratio). Generally speaking, we would like to select a design point that results
in the smallest wing and smallest powerplant while still satisfying all constraints. This
would imply the lowest mass and (friction) drag for the wing and propulsion system. For
a propeller airplane, this implies the highest wing loading and highest power loading
permitted by the constraints. For the jet airplane, this implies the highest wing loading
combined with the lowest thrust-to-weight ratio permitted by the constraints. This is vi-
sually shown in Figure 7.21. In the next example, we demonstrate how to choose a design
point in the matching diagram.
W/P T/W
W/S W/S
Figure 7.21: Notional representation of the design objective in the matching diagram of the propeller airplane
(left) and jet airplane (right).
Example 7.9
We use the matching diagrams that have been constructed in the foregoing examples.
The hash marks that we have added to each of our constraint curves now become im-
portant. They signify that any point on the hashed side of a particular constraint curve
FT
does not satisfy that constraint. Any point on the other side of the constraint does satisfy
P ROPELLER A IRPLANE For our propeller airplane, we first determine the feasible de-
sign space, which needs to be below all of the WTO /P TO curves and to the left of the
constraining WTO /S w lines. This is marked by the green area in Figure 7.22. Based on
this, we can see that the feasible design space is bounded by three constraints:
1. Stall speed
RA
2. Cruise speed
3. Climb gradient
The other constraints are not actively bounding the design space for the chosen design
parameters.
In order to minimize the design objective, we need to choose a point in the top-right
corner of the feasible design space. Since there is no clearly defined top-right corner, we
can select any point on the climb-gradient curve. We choose to go for a design point that
results in the smallest wing at the junction of the stall-speed line and the climb-gradient
line. This implies that the wing of the airplane is sized by the stall-speed constraint, and
D
the propulsion system is sized by the climb-gradient constraint. These are the two active
constraints for the chosen design point. All other constraints are inactive.
J ET A IRPLANE For our jet airplane, the feasible design space resides above all TTO /WTO
curves and to the left of the WTO /S w lines. This is shown in green in Figure 7.23. The fol-
lowing constraints bound the design space:
1. Approach speed
2. Cruise speed
3. Climb rate
4. Take-off field length
The other two constraints are not actively bounding the design space.
To minimize the design objective, we need to choose the design point in the bottom-
right corner of the design space. Since the bottom right corner is not clearly defined, we
can choose any design point on the take-off curve between the junction with the climb-
0.5
0.5
cru
Thrust-to-weight ratio, T TO/W TO (-)
Power loading, W TO/P TO (N/W)
i se
0.4 tak feasible
sp
0.4
stall speed
ee
ra te
d
ff f
b gr
i eld
0.3
adie
cli
gth
nt
gth
m
b l en
approach speed
te ie
0.2 ff f
e-o design point
tak
d
0.1 ee
sp 0.1
se design point
c rui feasible
0
0 500 1000 1500 0
0 2000 4000 6000
2
Wing loading, W TO/S w (N/m )
Wing loading, WTO/S w (N/m 2)
rate curve and the junction with the approach-speed curve. We choose the design point
at the junction between the take-off field length and the approach speed. For this choice,
these two constraints are actively sizing the powerplant and wing of our jet airplane. All
FT
Figure 7.23: Matching diagram with feasible design
space and chosen design point for the jet airplane
RA
other constraints are inactive.
Now that we have chosen a feasible design point in our matching diagram, we can com-
pute the power (or thrust) and wing area of the airplane, if we know the maximum take-
off mass m MTO . The wing area can be computed as follows:
m MTO g
Sw = (7.72)
WTO /S w
D
For jet airplanes, we can compute the sea-level static take-off thrust as follows:
Example 7.10
Consider the design points chosen for the propeller airplane and jet airplane in Example
7.9. We are going to calculate the required wing area for either airplane, along with the
take-off power and take-off thrust, respectively.
P ROPELLER A IRPLANE Assume that the battery-powered electric airplane has a max-
imum take-off mass of m MTO = 1, 830 kg. We can observe from the matching plot in
Figure 7.22 that the design point corresponds to the following properties:
By employing (7.72) and (7.73) we compute wing area, and take-off power:
S w = 14.6 m2 P TO = 152 kW
Since our jet airplane has a single electric motor, this motor should produce at least 152
kW of power.
J ET A IRPLANE Assume that the jet airplane has a take-off mass of m MTO = 63.0 t. We
can observe from the matching plot in Figure 7.22 that the design point corresponds to
the following properties:
FT
By employing (7.72), (8.4), and (7.74) we compute the wing area and the take-off thrust
at sea-level static conditions:
Since the jet airplane has Ne = 2, each engine should have a take-off thrust of at least
TTO /Ne = 105 kN.
RA
You can see that we have used six flight performance requirements to determine the
required size of the wing and powerplant. This process is a sizing process as defined
in Section 2.3. In this process we have been prompted to make assumptions as well
as design choices. This implies that the resulting size of the wing and powerplant is
dependent on these assumptions and choices. By making different choices, the result
would be different. This is inherent to the design process. Some of the assumptions we
have had to make are going to change once we analyze our airplanes. However, the result
of the preliminary sizing process gives us an excellent starting point for the dimensioning
D
of two of the most critical subsystems of an airplane: the wing and the powerplant. In
Chapter 8 we will use the results from this preliminary sizing step as a starting point
for determining the geometrical dimensions of the wing and powerplant. In the next
assignment, you will determine the size of the wing and powerplant of your airplane.
A SSIGNMENT 7.10
In this assignment, you are going to determine the size of your wing as well as the
power or thrust of your powerplant.
a. In your matching diagram, indicate the feasible design space.
b. List the constraints that are actively bounding the feasible design space.
c. Based on your design objective, select a design point in your feasible
design space. Report the value of WTO /S w and WTO /P TO or TTO /WTO ,
FT
RA
D
FT
In this chapter, you are going to determine the design parameters of the wing and the
propulsion system. At this point, we assume that you have determined the wing area
(S w ) in a previous design step and that the aspect ratio of the wing has been chosen (see
Chapter 7). In Chapter 4, you have chosen the type of propulsion system and the num-
ber of engines or motors for your design. Subsequently, in Chapter 2 you have estimated
RA
the required power (P TO ) or thrust (TTO ) that is required to meet the flight and field per-
formance requirements.
In this chapter, you will learn how to determine the design parameters of your wing
and propulsion system. For the wing, this includes determining the sweep angle, taper
ratio, dihedral angle, and average thickness-to-chord ratio. All of these parameters will
be properly introduced in the subsequent sections. We will also show how to position
the main structural elements of the wing and how to compute the so-called mean aero-
dynamic chord (MAC). For the propulsion system, you will learn how to determine the
D
dimensions of an engine or motor. If you have chosen propeller propulsion, you will
learn how to determine the diameter of your propeller. Finally, we will show you how to
integrate your propulsion system with the wing or fuselage.
This chapter is divided into three sections. In Section 8.1.3, we show you how to
determine the geometrical design parameters that describe the wing shape. In Section
8.2, we show you how to size and integrate a propeller propulsion system. Here, we
distinguish between electric motors, reciprocating engines, and turboshaft engines. In
Section 8.3, we demonstrate how to determine the size of a turbofan engine and how to
integrate it with the wing or the fuselage.
167
168 8. D ESIGN OF W ING AND P ROPULSION S YSTEM
to estimate other geometric parameters. Together with the aspect ratio, the sweep angle
and the taper ratio determine the planform shape of the wing, i.e the projection of the
wing onto a horizontal surface. Based on the planform shape, we compute the location
and dimension of the mean aerodynamic chord. The mean aerodynamic chord is the
one-dimensional representation of the wing in the symmetry plane of the airplane. With
the dihedral angle and the thickness-to-chord ratio, one can also draw the wing in side
view and top view. How to determine each of these parameters is explained below.
FT
mation of shock waves over the wing’s upper surface that appear at high subsonic Mach
numbers. These shock waves can cause flow separation, resulting in a large drag force.
Without going into too much detail, the shock waves can form when the flow over the
wing is sped up to supersonic conditions. Imagine that an airplane increases its Mach
number during accelerated flight. For a given speed, the flow over the wing’s upper sur-
face has an even higher peak Mach number due to the wing’s thickness and the lift it
generates. Therefore, there exists a flight Mach number where the local Mach number
reaches a value of 1: sonic flow somewhere on the wing. The flight Mach number where
RA
this occurs is called the critical Mach number. Beyond the critical Mach number, part of
the wing experiences supersonic flow, and shock waves may occur. Introducing sweep
increases the critical Mach number and, therefore, postpones shock wave formation.
[19]
To understand how the critical Mach number is increased by wing sweep, have a look
at Figure 8.1. Here, we have decomposed the Mach vector into a component parallel to
the quarter-chord sweep line of the wing and a component perpendicular to it. Now, the
effect that the parallel Mach vector has on the velocity distribution over the wing is neg-
D
ligible because in this direction, the wing is flat. On the contrary, the flow perpendicular
to the sweep line experiences the airfoil curvature. Therefore, we can conclude that the
perpendicular Mach vector is responsible for the flow acceleration over the wing and the
maximum local Mach number reached. However, the magnitude of the perpendicular
Mach vector is reduced by a factor cos Λc/4 , where Λc/4 is the quarter-chord sweep an-
gle. In other words, the larger the sweep angle, the higher can be the critical flight Mach
number.
This theory of how wing sweep can increase the critical flight Mach number is called
simple sweep theory and is a simplification of the complicated interaction between ge-
ometrical parameters, such as wing sweep and wing thickness, and the mixed subsonic
and supersonic flow that exists around the wing at high subsonic Mach numbers. How-
ever, simple sweep theory captures the most important effect that we are after: increas-
ing the wing sweep can allow airplanes to fly at high subsonic Mach numbers with a
relatively low drag penalty. We need a method that can relate the cruise speed to the
Λc/4
Λc/4
Airfoil
Figure 8.1: Definition of the quarter-chord sweep angle and decomposition of the Mach vector.
sweep angle.
In Ref. [7], the following method is presented to determine the quarter-chord sweep
angle of the wing:
Λc/4 =
(
arccos
³0
1.16
FT
M CR +0.5
´
where M CR is the cruise Mach number that is related to the cruise speed VCR through
the local speed of sound according to (7.24). You can see that for airplanes that have
if M CR < 0.66
if M CR ≥ 0.66
(8.1)
RA
a cruise Mach number below 0.66, we propose to have no sweep angle at all. It is im-
plicitly assumed that with modern airfoils and the present sweep-Mach-number corre-
lation, it is possible to obtain a wing geometry with adequate thickness to accommo-
date a lightweight structure and sufficient volume to store fuel and systems. For cruise
Mach numbers higher than 0.66, the sweep angle increases gradually. For example, at
M CR = 0.78) the proposed quarter-chord sweep angle is Λc/4 = 25◦ .
40
Quarter-Chord Sweep Angle,
D
MCR
30
Λc/4 (deg)
20 Λc/4
10
0
0.2 0.4 0.6 0.8 1
Cruise Mach Number, MCR (~)
Figure 8.2: Proposed relationship between cruise Mach number and quarter-chord sweep angle.
A SSIGNMENT 8.1
In this assignment, you choose the quarter-chord sweep angle of your wing.
a. What is the cruise Mach number of your airplane?
b. What quarter-chord sweep angle do you choose for the wing of your air-
plane?
c. Why do you choose this sweep angle for your wing?
FT
RA
Figure 8.3: The taper ratio is the ratio of the tip chord over the root chord.
A wing with a taper ratio of 1 has the advantage of being easy to produce. The inter-
nal structure of a wing is often reinforced by positioning ribs in the chordwise direction
at regular spanwise intervals. If the wing has no taper, each of the ribs has the same
geometry. Also, all the connections between the wing skin and the ribs are the same at
D
every rib station. This makes the manufacturing of ribs cheaper because you have fewer
unique components. A tapered wing is, therefore, more expensive to produce but has
advantages over an untapered wing.
Tapering the wing towards the tip affects the spanwise distribution of lift. You can
imagine that a smaller chord length in the outboard wing also reduces the local lift. Vice
versa, a larger wing chord at the root increases the lift. In other words, the value of the
taper ratio influences how the lift is distributed. Figure 8.4 shows how the taper ratio
changes the lift distribution for a wing with zero quarter-chord sweep angle. For a wing
without tapering (i.e. λ = 1), there is relatively more lift over the outboard wing, while
a taper ratio of λ = 0 causes more lift inboard. In Figure 8.5, we show that the sweep
angle also affects the spanwise distribution of lift. Sweeping the wing rearwards results
in more lift over the outboard wing and less lift inboard.
The redistribution of spanwise lift as a result of the taper ratio and the sweep angle
has the following effects:
Λ = -45°
λ=0.5
λ=1 Λ = 0°
Λ = +45°
1
clc/CLc (~)
Figure 8.4: Effect of taper on spanwise lift distribu- Figure 8.5: Effect of wing sweep on lift distribution
tion. Reproduced from Ref. [9]. on a wing without taper. Reproduced from Ref. [19].
FT
1. L IFT INDUCED DRAG. The span efficiency factor (ϕ in Equation (6.17) of Section
6.1.1) is changed. The closer the lift distribution is to an elliptical lift distribution,
the closer φ is to 1. An elliptical lift distribution minimizes the lift-induced drag
of a wing. For a wing with zero quarter-chord sweep angle, a taper ratio of λ = 0.4
results in a lift distribution that is almost elliptical.
2. W ING MASS . If more lift is generated outboard of the wing, this results in a higher
bending moment in the wing structure. So tapering the wing can reduce the bend-
RA
ing moments in the wing box. Secondly, a tapered wing results in a thicker airfoil
at the wing root, which has a larger second moment of area about the chord line
of the airfoil. This means that for a given bending moment and given allowable
stress level in the material, a thinner skin or smaller spar cap is required to carry
this moment. The lower bending moment and the larger second moment of area
at the root of a tapered wing result in a lower mass of the wing structure compared
to an untapered wing.
3. T IP STALL The taper ratio results in a lower chord length near the tip. The two-
dimensional lift coefficient of an airfoil section has a dependency on the chord
D
length. In general, airfoils with a smaller chord length have a lower maximum lift
coefficient under the same aerodynamic conditions than the same airfoils with
a larger chord length. This means stall on a tapered wing will likely start more
outboard than on an untapered wing. A stall over the outboard wing might result
in a reduction in roll control authority if an aileron is located in the stalled region
of the wing. This effect is even stronger on tapered and swept wings.
You can choose a taper ratio anywhere between 0 and 1. Based on your design ob-
jective, you should qualitatively weigh the arguments above and come to a rational de-
cision. To guide you in decision-making, we have reproduced a figure from Torenbeek
[15] that plots the value of the quarter-chord sweep angle and taper ratio for various air-
planes. You can see that for propeller airplanes, the taper ratio is scattered between 0.35
and 0.8. Also, you can see that when the quarter-chord sweep angle is increased, the ta-
per ratio seems to decrease. Although there is quite some spread in the data, we propose
a simplistic trend line that relates the taper ratio to the sweep angle:
³ π ´
λ = 0.2 2 − Λc/4 (8.3)
180
where Λc/4 is measured in degrees. You can see that this formula proposes a taper ratio
of 0.4 when the quarter-chord sweep angle is zero. Naturally, you can deviate from this
choice, as many airplane designers have done in the past.
1.0
transport jets and business jets
-20 -10
0.4
0.2
0
FT10
Quarter-chord sweep angle, Λc/4 (deg)
20 30 40 50
RA
Figure 8.6: Sweep angle and taper ratio of various airplanes with notional trend line. Data from Ref. [15].
A SSIGNMENT 8.2
Secondly, we compute the root chord knowing that we have a trapezoidal wing planform:
2S w
cr = (8.5)
(1 + λ)b
c t = λc t (8.6)
With all these geometric characteristics of the wing computed, it is now time to make
a top-view drawing of our wing planform. In summary, we have the following design
process for the wing planform design:
Step 1 Determine the quarter-chord sweep angle based on the cruise Mach number.
FT
In the following example, we demonstrate how to do that.
Example 8.1
RA
In this example, we design the wing planform for a turbofan airplane. We use the top-
level airplane requirements from Example 7.1, which specify a cruise Mach number of
M CR = 0.80.
Step 1 We first determine the sweep angle. Since M CR = 0.80 > 0.66, we decide to apply
a sweep angle to our wing. We employ (8.1) to compute the quarter-chord sweep
angle and report the value in the table below.
D
Step 2 Based on the sweep angle, we decide to use the linear relationship of (8.2) to
compute the taper ratio and report the value in the table below.
Step 3 We subsequently employ Equations (8.4), (8.5), and (8.6) to compute the span,
the root chord, and the tip chord. To do that, we use the wing area that we have
determined in Example 7.10. We report the resulting values in the table below.
Step 4 We first draw the root chord and the span of the wing. Then, we draw the
quarter-chord sweep line and the tip chord. Finally, we connect the leading and
trailing edges of both wing sections, and we have the wing planform.
cr
Λc/4
bw
0 2 4 [m] ct 0 2 4 [m]
A SSIGNMENT 8.3
FT
Draw the planform shape of your wing in the top view. You may also draw only
one wing half, similar to the example above. Use the same scale as you used to
draw your fuselage top view (Assignment 5.3) and side view (Assignment 5.4).
Now that we have defined the wing planform shape, we can graphically construct the
mean aerodynamic chord (MAC). In Figure 8.7, the MAC is drawn alongside the location
of the leading edge of the MAC, i.e. the LEMAC. The longitudinal location of the LEMAC
RA
and the length of the MAC are important properties of the wing. airplane properties such
as the center-of-gravity or the neutral point are often expressed as a fraction of the MAC.
To find the MAC and the LEMAC, we take the following steps:
Step 1 Draw the 50%-chord line
Step 2 From the trailing edge of the root, extend the root chord by a tip-chord length.
Extend the tip chord by a root-chord length from the leading edge of the tip.
Diagonally connect the points of the extended root and tip chords.
Step 3 At the intersection between the diagonal and the 50%-chord line, draw the mean
D
aerodynamic chord line between the leading edge and trailing edge.
Step 4 Draw the MAC of the wing on the symmetry plane.
This graphical method works for all straight-tapered wings. However, technically, we
have only constructed the MAC for only one half of the wing. The MAC of the other wing
half has the same longitudinal position but is mirrored in the symmetry plane. There-
fore, the MAC of the full wing resides on the symmetry plane of the wing. You can per-
ceive it as a one-dimensional representation of the wing. The following example shows
how to construct the MAC using the steps above.
Example 8.2
In this example, we graphically construct the MAC for the wing of Example 8.1. We use
the following steps:
LEMAC
0.5
c MAC
cr
ct
Figure 8.7: Geometric construction of the mean aerodynamic chord (MAC) on a straight-tapered half-wing
planform.
FT
Step 2 Secondly, we extend the root and tip chords by the tip and root chords, respec-
tively, and connect the tip points with a diagonal line.
RA
ct
0.5
D c
cr
0 2 4 [m] 0 2 4 [m]
Step 3 At the intersection of the two dashed lines, we draw the dimension of the MAC
between the local leading and trailing edge.
Step 4 Finally, we draw the MAC on the symmetry line of the wing.
MAC
0 2 4 [m] 0 2 4 [m]
A SSIGNMENT 8.4
FT
Draw the mean aerodynamic chord on the symmetry line of your wing from As-
signment 8.3.
cation point(s) of the load(s). Therefore, movable parts of the wing, such as ailerons and
flaps, should be positioned behind the aft spar, while a movable leading edge should be
positioned ahead of the front spar. The volume between the spars can be utilized for fuel
storage. Therefore, the available fuel volume is also impacted by the location of the front
and rear spar.
A wing can have a single spar, two, three, or even more spars. In the preliminary
design phase (see Chapter 2), a more refined structural layout of the wing is made. In
Figure 8.8, we show a wing profile with a front, mid, and rear spar. Here, we consider our
wings to have a box structure with two spars: a front spar and a rear spar. In between
the spars, we have volume for a fuel tank. We have a fixed leading edge or a leading-edge
high-lift device such as a slat ahead of the front spar. Behind the rear spar, we can have
a fixed trailing edge or volume for a flap or an aileron. In Chapter 11, we present how
to size these wing movables. To allow for sufficient space for these high-lift devices, we
x
c
Figure 8.8: The chordwise location of the front and rear spar in a typical airfoil section.
propose the following positions of the front and rear spar, respectively:
³x ´
= 0.20 (8.7)
c front spar
³x ´
= 0.70 (8.8)
c rear spar
FT
While it is likely that the exact positions of the spars change in the design process,
they are important elements to position at this stage of the design process. The rear spar
is important regarding the positioning of the main landing gear (Chapter 9), while the
front spar is important for positioning a wing-mounted propulsion system such as a jet
engine or an electric motor.
A SSIGNMENT 8.5
RA
In this assignment, you position your wing’s front and rear spar. You may assume
that the spars are straight and are located at a constant chordwise location along
the span of the wing.
a. What is the chordwise location of your front and rear spar, respectively?
b. Draw the spars in your wing planform from Assignment 8.3.
D
First of all, the wing’s drag is a function of the thickness-to-chord ratio. Generally
speaking, the thicker the wing, the higher the so-called form drag. Form drag is drag
associated with the shape of a body. Since the thickness-to-chord ratio of a wing affects
the shape of the wing, it also impacts the form drag. The zero-lift drag of an airfoil can
be correlated to the thickness-to-chord ratio using the following relationship:
t t
µ ¶
c d0 ≈ 0.0035 + 0.018 for 0.06 ≤ ≤ 0.25 (8.9)
c c
The form drag of the airfoil contributes to the zero-lift drag coefficient (C D 0 ) that we
have estimated in Chapter 6 based on the wetted-area-to-reference-area ratio (S wet /S w ).
The wetted area of the wing is approximately twice its wing area minus the part that is
covered by the fuselage, i.e. c r × w fus . We propose to select a thickness-to-chord ratio for
which the two-dimensional zero-lift drag coefficient is:
c r w fus
µ ¶
c d0 < C̄ f 2 − (8.10)
Sw
Chapter 6.
FT
where C̄ f is the average friction drag coefficient of the airplane, which we introduced in
If a wing is too thick, the flow might not be able to ‘stick’ to it and follow its curved
surface all the way to the trailing edge. So, if airfoils are too thick, the flow might easily
separate and cause the wing to stall. This can result in a low value of the maximum lift
coefficient. On the other hand, if a wing is too thin, a sharp leading edge results, which
can also cause premature separation and a relatively low maximum lift coefficient. This
relationship is shown by means of bands for various two-dimensional airfoil families in
RA
Figure 8.9. While the maximum lift coefficient of an airfoil (c l max ) is not identical to that
of a three-dimensional wing, it is correlated to it. The legend refers to various airfoil
families that are further elaborated in Chapter 11. The band of each airfoil family is due
to the fact that the maximum lift coefficient depends on the airfoil’s size and the speed
at which the airplane flies.
enve
pe
lope
D
lo
1.6
ve
en
cl max (-)
1.6
1.2 MS(1)
NACA 4 & 5-Digit NACA 4 & 5-Digit
1.2
NACA 6-Digit NACA 6-Digit
0.8
6 10 14 18 22 6 10 14 18 22
Thickness to Chord Ratio, t/c (%) Thickness to Chord Ratio, t/c (%)
Figure 8.9: Relationship between thickness-to-chord ratio and maximum lift coefficient for various two-
dimensional airfoil families. Modified from Ref. [11]
So at this stage of the wing design, you have to select a thickness-to-chord ratio that
can support a wing shape with said C L max . For swept wings with a moderate to high
aspect ratio, it is known that the maximum lift coefficient deteriorates with increasing
sweep angle. Therefore, we propose the following relation between the two-dimensional
maximum lift coefficient of the airfoil (c l max ) and the quarter-chord sweep angle:
C L max, CR
c l max > 1.1 p (8.11)
cos Λc/4
where C L max, CR is the maximum lift coefficient in the cruise configuration. The factor 1.1
accounts for three-dimensional effects as well as stall characteristics of the wing. We
advise you to select a thickness-to-chord ratio for which the two-dimensional maximum
lift coefficient is below the dashed envelope in the graphs of Figure 8.9.
If a wing is designed for Mach numbers higher than 0.65, it is likely to have wing
sweep. Wing sweep allows airplanes to fly at high subsonic Mach numbers without a
large drag penalty. However, that statement is only valid if the thickness-to-chord ratio
FT
is low enough to prevent the formation of strong shock waves at the combination of the
cruise Mach number (M CR ) and the cruise lift coefficient (C L,CR ). Therefore, for swept
1.5
t cos Λc/2 [0.935 − (M CR + 0.03) cos Λc/2 ] − 0.115C L CR
c
≤
cos2 Λc/2
(8.12)
where Λc/2 is the half-chord sweep angle and C L CR is the cruise lift coefficient. This lift
coefficient is computed from the lift equation (6.7), which can also be written in terms
RA
of M CR and cruise pressure p CR when we use the equation of state (7.22) and the Mach
number definition (7.24):
2 WTO 2 WTO
C L CR = = (8.13)
ρ CR VCR
2 Sw γp CR M CR S w
2
where WTO /S w is the wing loading, which we have estimated in Chapter 7 and γ is the
ratio of specific heats, i.e. γ = 1.4 for air. Note that p CR depends on the cruise altitude
and can be computed using (7.20).
D
Figure 8.10 shows the relationship between the maximum thickness-to-chord ratio
and the half-chord sweep angle for a range of cruise Mach numbers and two cruise lift
coefficients. For a given lift coefficient and Mach number, a higher half-chord sweep
angle allows for a higher thickness-to-chord ratio. Furthermore, you can observe that
the allowable thickness-to-chord ratio is lower for C L CR = 0.7 compared to C L CR = 0.3.
In summary, to minimize the wing mass and maximize the internal volume, it is ad-
vised to choose a thickness-to-chord ratio that is as high as possible. However, aero-
dynamic constraints related to the profile drag, the maximum lift coefficient, and wave
drag limit the thickness-to-chord ratio. Each of these can be perceived as a maximum
lift budget that needs to be respected to arrive at a feasible design. This process consists
of the following steps:
Step 1 Compute the maximum thickness-to-chord ratio to stay within the zero-lift-
drag budget of the wing.
0.75 0.70
0.15
0.80 0.75
0.1 0.85 0.80
0.90 0.85
0.05
0.90
0
0 10 20 30 0 10 20 30
half-chord sweep angle, (~) half-chord sweep angle, (~)
c/2 c/2
Figure 8.10: Relation between maximum thickness-to-chord ratio and half-chord sweep angle for various
cruise Mach numbers and C L CR = 0.3 (left) and C L CR = 0.7 (right).
Step 2 Compute the minimum and maximum thickness-to-chord ratio to stay within
the maximum lift budget
Step 3 For cruise Mach numbers over 0.65, compute the maximum thickness-to-chord
ratio to stay within the wave drag budget.
FT
Step 4 select the highest thickness to chord ratio within the limitations set by the pre-
vious steps.
RA
The next examples show how you can come to a robust choice for the thickness-to-chord
ratio of your wing.
Example 8.3
In this example, we determine the thickness-to-chord ratio of the wing of Example 8.1.
Step 1 Using (8.10) with w fus = 3.30 m and C̄ f = 0.0030, we compute that the two-
D
The previous example has shown that the difference between the minimum and maxi-
mum thickness-to-chord ratio can be quite small. You can imagine that in some cases,
Option 1 You choose a thickness-to-chord ratio between the minimum and maximum
values you have computed. This is advised if these two values are quite close
to each other.
Option 2 You go back to an earlier step in the design process and lower the value for
C L max, CR (see Assignment ??) or increase the value for the equivalent friction
coefficient, C̄ f (see Assignment 6.2). This option is advised if the minimum
value is much higher than the maximum value.
If you choose option 2, you are essentially iterating on your design. You are updating a
value that you estimated in an earlier design step. However, all the calculations that are
performed downstream of the selection of these parameters need to be redone.
A SSIGNMENT 8.6
chord ratio? FT
In this assignment, you will determine the thickness-to-chord ratio of your wing.
a. What is the maximum zero-lift drag coefficient of the two-dimensional air-
foil of your wing, and what is the corresponding maximum thickness-to-
8.1.6. D IHEDRAL
A wing’s dihedral angle, Γ, is the angle the wing forms with the horizon when seen from
the front and is shown in Figure 8.11. The dihedral angle impacts the lateral stability
of the airplane, i.e. its rolling moment due to sideslip. . In Section 4.1.1, we explained
what lateral stability means and why it is important for an airplane. We showed that
a high-wing configuration increases the lateral stability while a low-wing configuration
reduces the lateral stability. Lateral stability is also referred to as the dihedral effect. This
is because an increase in dihedral angle increases the rolling moment due to sideslip.
The dihedral effect is explained in Figure 8.12. Here, it is assumed that the airplane
experiences a small sideslip angle, β, which causes a side wind with respect to the air-
plane. The magnitude of the side wind is V sin β. If we assume that the sideslip angle
is small, we can approximate the side wind component with βV , where β is expressed
in radians. If we decompose the side wind velocity component into a component per-
pendicular to the wing and a component parallel to the wing, we observe that the port
wing experiences a small increase in angle-of-attack, while the starboard wing experi-
ences a decrease in angle-of-attack. Therefore, the port wing experiences an increase in
lift, while the starboard wing experiences a decrease in lift. As a result, the airplane expe-
riences a rolling moment in the counter-clockwise direction when seen from the front.
This is a stabilizing rolling moment.
ΔL
FT
starboard side port side ΔL
RA
βVcosГ βVcosГ
βV βV
Figure 8.12: The dihedral angle Γ causes a change in lift on either wing half when subjected to a sideslip angle
β.
Apart from the dihedral angle, also the sweep angle impacts the rolling moment due
to sideslip. This is illustrated through vector decomposition in Figure 8.13. When we
decompose the velocity vector V into a component perpendicular to the leading edge
D
and a component parallel to the leading edge, we observe that the perpendicular com-
ponent is smaller on the starboard side compared to the port side. If we assume that the
perpendicular velocity component is responsible for the change in lift, we can deduce
that the port-side wing produces more lift than the starboard-side wing. This causes a
rolling moment in the counter-clockwise direction when seen from the front. In other
words, an aft-swept wing produces a stabilizing rolling moment due to sideslip.
In summary, we have seen that there are three aspects that determine the lateral sta-
bility of the airplane:
1. Vertical wing position. A high-wing configuration increases stability.
2. Dihedral angle. The dihedral angle increases stability.
3. Wing sweep. Wing sweep increases stability.
When an airplane has a high-wing configuration, an aft-swept wing, and a positive di-
hedral angle, it could have too much lateral stability. That implies that the airplane
produces a relatively large rolling moment when subjected to a side wind. This is, for
ΛLE
V ΛLE
V
β
β
FT
Figure 8.13: A sideslip angle causes a difference in the magnitude of the velocity vector perpendicular to the
leading edge of the starboard wing and port wing, respectively.
RA
example, problematic when performing a landing in cross-wind conditions. Therefore,
airplanes with a high-wing configuration and wing sweep often apply anhedral, i.e. neg-
ative dihedral.. Applying anhedral to the wing reduces the lateral stability.
To determine the ’right’ dihedral angle, we propose the following guidelines:
• Default: Γ = 3◦ for unswept wings at mid-wing location
• For every 10 degrees of quarter-chord sweep angle, reduce the dihedral angle by 1
degree.
• For high-wing or low-wing airplanes: subtract or add 2 degrees, respectively.
The following example shows how these guidelines can be applied.
D
Example 8.4
In this example, we determine the dihedral angle for the wing of Example 8.1, and we
assume that this wing is made for an airplane with a low-wing configuration.
This wing has a quarter-chord sweep angle of 26◦ . If we follow our guidelines, we
compute the following:
Γ = 3.0◦ − 0.1 · 27◦ + 2.0◦ = 2.3◦
The dihedral angle also has an effect on the clearance between the ground and the wing
when the airplane is close to the ground and has a bank angle. Also, when an engine or
a propeller is located on the wing, the dihedral of the wing affects its ground clearance.
Figure 8.14 shows that the dihedral angle causes a vertical displacement of the propeller
and turbofan propulsion systems when mounted on the wing. It also raises the wing
tip and provides adequate clearance when landing on one wheel with a bank angle of φ.
Typically, a minimum clearance angle of eight degrees is required, i.e. φ > 8◦ , but more is
desired to allow for operational maneuvering in proximity to the ground. Increasing the
dihedral angle of a low-wing airplane can, therefore, result in lower landing-gear height
and associated landing-gear mass. Therefore, choosing a dihedral angle that is a few
degrees higher than suggested by the guidelines above can be justified.
Ф Ф
A SSIGNMENT 8.7 FT
Figure 8.14: The wing’s dihedral can increase the ground clearance of the wing and/or propulsion system when
landing with a bank angle φ.
In this assignment, you are going to determine the dihedral angle for your air-
plane.
RA
a. Using the guidelines presented above, compute the dihedral angle, Γ.
b. Based on ground-clearance considerations, would you increase the dihe-
dral angle? If so, why? What dihedral angle would you then choose?
In the following example, we show how this leads to a front view and side view of your
wing.
Example 8.5
In this example, we draw the wing of Example 8.1 in the front and top views.
Step 2 With a dashed line, we draw the dihedral angle in the front view (below right).
Rather than using 2.3◦ , like in Example, 8.4, we decide to use Γ = 5◦ . This is due
FT
to the fact that turbofan engines are to be mounted to the wing. A somewhat
higher dihedral angle could result in a somewhat shorter landing gear.
side side
view view
D
0 2 4 [m] 0 2 4 [m]
Step 3 We compute a vertical displacement of the tip of ∆z t = 1.3 m, and we mark the
location of the tip in the side view. With t /c = 0.105, we compute a tip thickness
of 0.19 m. Subsequently, we draw the tip profile in the side view and a small line
in the front view (below, left).
Step 4 We finish the front view by connecting lines between the top and bottom of the
airfoil profile, respectively. We finish the side view by connecting the leading
and trailing edges of the root and tip profiles. This is shown below, right.
A SSIGNMENT 8.8
FT
Using the steps presented above, add a front view and side view to the wing draw-
ing you made in Assignment 8.3. Use the American convention to position your
front and side views with respect to your top views, similar to Example 8.5. Use
the same scale as in 8.3.
• blade geometry
• tractor or pusher configuration
The angular speed and the cruise speed determine the helical tip Mach number of the
blade. As the tip of the blade rotates, it also moves forward with respect to the flow.
Therefore, the resultant velocity can be found by velocity vector triangulation. If this
helical tip velocity is close to the speed of sound (i.e. the helical tip Mach number is close
to one), the noise of the propeller can increase dramatically. As the tip speed is governed
by the angular speed and the diameter, the implicit requirement to limit propeller noise
affects the diameter of the propeller.
Statistically, a good first estimate for the diameter of a tractor propeller is given by:
s
P TO
D p = 0.55 4 (8.14)
1000Ne
Where D p is measured in meters, Ne is the number of engines (or motors), and P TO is the
take-off power that we determined in Chapter 7. Equation 8.14 is a modification of the
FT
method presented in Ref. [9]. A short survey among civil and military propeller airplanes
showed that most propellers fall within a 5% accuracy of this estimation, while military
propellers are underestimated in size up to 15%.
Example 8.6
For the battery-powered, four-seat airplane of Example 7.10, we determine the propeller
diameter. We have P TO = 152 kW and Ne = 1. Employing (8.14), we get: D p = 1.9 m.
RA
A SSIGNMENT 8.9
If your airplane has a propeller, use the methods presented above to determine
the diameter of your propeller.
As shown in Figure 4.13, there are three different systems that can provide shaft power
to the propeller: a reciprocating engine, a turboshaft engine, or an electric motor. Also,
hybrid-electric power trains can be conceived, but they are beyond the scope of this book
due to the increased complexity [21]. With a known take-off power, you can choose an
engine or motor from an online engine database. The characteristic dimensions, the en-
gine mass, and the power-specific fuel consumption can then be used in the subsequent
integration and analysis steps. However, in this section, we provide simple relations that
define the dimensions of engines based on their take-off power, along with a volumetric
engine envelope to account for adjacent systems.
R ECIPROCATING E NGINE Many reciprocating engines are have an opposed piston con-
figuration. That implies that two rows of pistons are opposite each other in the same
plane. Such a reciprocating engine has the shape of a rectangular box with a height (h e ),
a width (w e ), and a length (l e ). An example of a piston engine with the box dimensions
is shown in Figure 8.15. To estimate these dimensions we slightly modify the method of
Ref. [9] and find:
¶0.10 ¶0.30 ¶0.55
P TO P TO P TO
µ µ µ
h e = 0.30 w e = 0.17 l e = 0.06 (8.15)
1000Ne 1000Ne 1000Ne
where all the lengths are measured in meters. A small survey among opposed piston
engines demonstrates that the predictions of Equation 8.15 to be within 10% of the mea-
sured values.
we
he
e
FT
Figure 8.15: Example of a piston engine with rectangular box dimensions. Photo: Rameshng c b a
RA
A reciprocating engine requires clearance to the adjacent airplane structure to add
cooling systems, including an air intake, to fit auxiliary systems such as an alternator,
and to fit the engine support structure. Therefore, we add 10% of engine width on all
sides of the engine, except on the front side of the engine. We propose the following
relations to define the dimensions of a rectangular box around the engine. We use the
subscript ‘ee’ to refer to the engine envelope:
D
Example 8.7
Assume that the airplane of Example 7.10 would have a reciprocating engine with the
same take-off power. We determine the dimensions of the engine envelope for such an
opposed-piston engine.
We have P TO = 152 kW and Ne = 1. We first employ Equation 8.15 to find:
We then employ (8.16) and find the dimensions of the engine envelope of this opposed-
piston engine:
h ee = 0.65 w ee = 0.92 l ee = 1.03
e
De
FT
Figure 8.16: Example of turboshaft engine with circular tube dimensions. Photo: Matti Blume c b a
An engine envelope is required to allocate volume for the support structure between
the engine and the airframe. Moreover, volume is required for air ducting to and from
RA
the engine. Various shapes can be selected for engine envelopes. One of these shapes is a
rectangular box that encompasses the engine and has an inlet located below the engine.
The dimensions of the engine envelope for a turboshaft engine can be approximated
with:
h ee = 1.5D e w ee = 1.1D e l ee = l e (8.18)
Example 8.8
passengers. In previous steps, we decided that this airplane has Ne = 4 turboshaft en-
gines.
We perform the following steps:
Step 1 Using Equation 8.14, we compute that the propeller diameter for each propeller
is D p = 2.9 m.
Step 2 Using Equation 8.17, we compute that each turboshaft engine has the following
dimensions: D e = 0.67 m and l e = 1.46 m.
Step 3 Using Equation 8.18, we compute the following dimensions for the engine en-
velope: h ee = 1.00 m, w ee = 0.73 m, and l ee = 1.46 m.
E LECTRIC M OTOR In Figure 8.17, we show an example of an electric motor that has been
developed to drive a propeller. The motor itself can be approximated by a cylinder of
low fineness ratio, i.e. the ratio of the length over the diameter.. In addition, you can see
two box-shaped items that are connected to the motor. These are so-called inverters. An
inverter converts the direct current (DC) from the battery into alternating current (AC)
for the motor. The inverter is connected to the cockpit motor controls and regulates the
motor’s rotational speed.
Dm
FT
m
Figure 8.17: Example of an electric motor with circular tube dimensions and attached power electronics.
Photo: Magni-X
To determine the dimensions of the electric motor, we first establish the power den-
RA
sity of the motor, i.e. the amount of power per unit volume. We call this parameter ρ p ,
which has the unit W/m3 . The subscript ‘p’ refers to power. The fineness ratio of the
motor is indicated with f m , and is defined as follows:
lm
fm = (8.19)
Dm
where D m is the motor diameter and l m is the motor length as shown in Figure 8.17.
Using the geometrical relations of a cylinder, we determine the diameter of the engine
D
as follows: s
4P TO
Dm = 3 (8.20)
π fmρp
For a chosen motor fineness ratio, f m , we can then use Equation 8.19 to compute the
length of the motor, l m .
With the motor dimensions established, we now determine the required engine en-
velope.1 As you can see, the inverters in Figure 8.17 occupy quite some volume. This can
also be seen in Figure 6.8, where an electric motor is integrated into the nose of a small
airplane. We consider the volume that is occupied by the inverter and the cables be-
tween the motor and the inverter to be part of the engine envelope. On the other hand,
the liquid cooling system that may be required for electric motors and inverters is not
1 We keep the terminology for the engine envelope the same as for the internal combustion engines, even
though we have an electric motor.
considered to be part of the engine envelope. For the engine envelope, we propose a
rectangular box around the engine with the following dimensions:
s
3 πl e D m
2
h ee = 1.1D e w ee = 1.1D e l ee = l m + (1 + k cable ) (8.21)
4
where k cable is the ratio between inverter volume and the volume occupied by the cables
between the motor and the inverter. If the inverter is attached to the motor, k cable = 0.
Note that we compute the volume required by the inverter by assuming its volume equals
the volume of the electric motor and that it is shaped like a cube. In practice, various
other motor-inverter configurations have been conceived, which are considered outside
the scope of this textbook.
Example 8.9
In this example, we determine the dimensions of the electric motor for the propeller
FT
airplane of Example 7.10. From Example 7.10, we have a take-off power of P TO = 152 kW.
To determine the dimensions of the electric motor, we perform the following steps:
Step 1 We choose a motor fineness ratio of f m = 0.5 and assume a power density of
ρ p = 7.0 MW/m3 .
Step 2 We employ (8.20) and (8.19) to compute D m = 0.38 m and l m = 0.19 m, respec-
tively.
RA
Step 3 We choose k cable = 0.8 and with (8.21) we compute: h ee = 0.42 m, w ee = 0.42 m,
and l ee = 0.69 m.
A SSIGNMENT 8.10
ensures that we have the smallest cowling wetted area and also makes it easier for
the pilot to see over the nose (i.e., the over-nose angle).
3 V ERTICAL POSITIONING
We would like to position the engine as close as possible to the center of the fuse-
lage to minimize the thrust-induced pitching moment w.r.t. the center of gravity.
However, the required over-nose angle limits the vertical position for which the
engine can be installed.
The following example shows how a piston engine can be installed on the nose of a small
airplane.
Example 8.10
The airplane of Example 5.7 is a four-seat airplane with a tractor propeller mounted on
the nose. The fuselage is truncated at the front to leave room for an engine or motor. In
FT
this example, we show how the engine of Example 8.7 is integrated with the fuselage of
Example 5.7. The figure below shows how the engine and the engine envelope (in red)
are integrated with the fuselage. Here are the steps we take to position the engine with
respect to the fuselage:
Step 1 In the side view, we position the engine envelope against the front line of the
RA
fuselage. This line represents the firewall, a protective surface that separates the
passenger cabin from the engine compartment. We try to position the engine as
high as possible to minimize the pitching moment around the airplane’s center
of gravity. However, we are limited by the required over-nose angle.
Step 2 We add the propeller and the spinner, i.e. the aerodynamic cover around the
propeller hub.
D
Step 3 We draw a notional engine cowling that gives the airplane its characteristic nose
shape.
Step 4 We draw the engine, the engine envelope, and the propeller in the top view of
the airplane.
Step 5 We add an engine cowling between the engine and the fuselage. In this example,
the original fuselage design had some tapering towards the nose (still shown in
gray). However, since the engine envelope is wider than the front of the original
fuselage, we decide to modify the fuselage shape to better match the dimensions
of the engine envelope.
Example 8.11
FT
In this example, we integrate the electric motor of Example 8.9 with the fuselage of Ex-
ample 5.7.
To integrate the motor of Example 8.9 we follow the same steps as laid out in Example
8.10 for integrating a piston engine in the nose of an airplane. Compared to the engine
envelope of the piston engine, the volume of the engine envelope of the electric motor
is considerably smaller. This allows for a smaller volume in the nose to house the motor
RA
and the inverters, as can be seen below.
D
A SSIGNMENT 8.11
If your airplane has propeller propulsion and you have chosen for integration in
the nose of the airplane, perform the following tasks:
a. Integrate the engine and propeller with the nose of your fuselage. Do this
in the top view, side view, and front view.
b. Add an appropriate engine cowling to ensure a smooth integration with
the fuselage.
FT
tip and the fuselage to prevent acoustic fatigue on the fuselage structure and
means that the arm between the thrust vector and the airplane’s center of
gravity should be minimized.
ii. Position the engine in such a way as to minimize the additional wetted area
due to the engine. Positioning the engine axis in the middle of the wing re-
sults in the lowest wetted area of the engine cowling and, therefore, the min-
imum friction drag.
iii. Position the engine in such a way as to minimize the landing gear length. In
the case of a low-wing airplane, the propeller clearance to the ground should
be respected also in the case of deflated tires and compressed landing gear
struts (see Example 3.16). Therefore, you may decide to raise the engine in
favor of a shorter landing gear.
You might have noticed that information from other airplane components, such as the
fuselage and landing gear, is required for some of these guidelines. Some of that infor-
mation is not yet present in the first design iteration. Also, some of the guidelines might
be conflicting with each other. This means that you have to find a compromise when
selecting the actual position of your engine with respect to your wing.
When you have positioned the engine and propeller on the wing, it results in a non-
smooth geometry that is likely to create quite some pressure drag. Therefore, you need
to add a fairing between the engine and the wing, just like a cowling was added to the
engine in Example 8.10. In adding such a fairing, the following guidelines should be
adhered to:
1. The fairing should go around the engine envelope
2. The fairing should be aerodynamically shaped (no sharp corners except for the
trailing point)
3. The fairing should be long enough to position a structure within, which connects
to the front spar and/or the rear spar.
Example 8.12
FT
In this example, we integrate a reciprocating engine to a wing of a twin-prop airplane.
The wing, propeller, and engine have been sized ahead of this example. Here we focus
on how to integrate the engine. It is important to know that the wing has been sized for
a low-wing configuration and has a dihedral angle of Γ = 6◦ . We perform the following
steps:
Step 1 In the top view and side view of the wing, we draw the fuselage side-of-body, i.e.
RA
the junction between the wing and the fuselage
Step 2 The longitudinal and lateral positions are determined simultaneously in the top
view. We let the rear end of the engine envelope intersect with the front spar
while the distance between the propeller disk and the side-of-body is 0.15D p .
Step 3 We then draw the engine in the front view. We decide to let the bottom side of
the engine be parallel to the lower surface of the wing. We also rotate the engine
by 6 degrees to align with the wing’s dihedral.
D
Step 4 We make a section of the wing through the shaft of the propeller. Here, we add
the engine and propeller in the side view, putting the rear end of the engine
envelope against the front spar.
Step 5 In the side view, we make a notional fairing between the engine envelope and a
point behind the trailing edge.
Step 6 In the top view, we make a notional fairing between the engine envelope and
the trailing-edge point from the side view.
The steps result in an integrated engine with the wing, as shown below. The engine en-
velope is shown in red.
side-of-body
CL
0.15Dp 0.5wf
A A’
fairing
rear spar
front spar
FT
The previous example has shown that the creation of an additional fairing around the
wing-mounted engines can result in quite a large volume behind the engine. It is up
to the discretion of the designer to allocate functionality to this volume. This volume
could, for example, be used to store luggage or fuel. Note that we have chosen to give
the fairing a horizontal trailing edge. This makes sense because the engine envelope has
a larger width than height. However, it is also an example of some artistic freedom in
RA
shaping this fairing.
Example 8.11 has demonstrated that the integration of the electric motor with the
fuselage follows the same process as for the piston engine. The same applies to the in-
tegration of electric motors on the wing. All the guidelines provided on page 194 with
respect to the integration of propellers with a wing should be adhered to. Due to the
smaller engine envelope for electric motors, you must take care that the distance be-
tween the leading edge and the propeller disk is respected. Regarding the integration of
turboshaft engines, we can also follow the same guidelines as for piston engines. The
following example shows how the turboshaft engine can be integrated into a large pas-
D
Example 8.13
Step 1 We position the inboard engine and engine envelope (in red) in the top view. We
let the end of the engine envelope touch the front spar while we keep 0.15D p of
clearance between the fuselage side-of-body and the tip of the propeller.
Step 2 We position the second engine and its envelope in the top view by placing the
engine envelope against the front spar and letting a very small spacing between
the two propeller disks.
Step 3 We draw the engines, their envelopes, and propellers in the front view. We de-
cide to let the top of the engine envelopes be aligned with the middle of the wing
section such that most of the engine is below the wing. Because this is a high-
wing airplane, this vertical position brings the thrust line of the propeller closer
to the expected center of gravity of the airplane. Furthermore, this position also
brings the engine closer to the ground, making inspection, servicing, and main-
tenance easier. Finally, because this is a high-wing configuration, the ground
clearance of the propeller is expected to be easily maintained with a short land-
ing gear.
Step 4 In the top view, we add a fairing behind the engine envelope. Because the entire
fairing exists below the wing, we draw it as a dashed line in the top view. We add
a fairing behind the engine envelope with the same shape as the first engine.
FT
Due to the smaller chord of the wing at the spanwise location of the outboard
engine, part of the fairing sticks out behind the wing.
Step 5 Based on the front and top views, we construct the side view of the wing with the
two propellers, engines, engine envelopes, and fairings. We decide to construct
a vertical trailing edge of the engine fairing.
0.15Dp 0.5wf
CL
RA
fairing
rear spar
D
front spar
fairing
0 1 2 [m]
The previous example has demonstrated how multiple propellers can be fitted to the
wing. You can see that there is some design freedom in deciding on the exact spacing
between the propellers. In this case, we decided to keep the propellers relatively close to
each other such that the yawing moment in case of a one-engine-inoperative condition
is relatively small. Also, in shaping the fairings, we took some freedom in creating a
shape that ‘looks good.’ While this is not a quantifiable design objective, it does help
in defining shapes that bring realism to the design. If you want to get inspiration for
how to shape such components, it might be instructive to review the pictures of your
reference airplanes (see Assignment 3.8).
A SSIGNMENT 8.12
If your airplane has propeller propulsion and you have chosen wing-mounted
propellers, perform the following tasks:
a. Integrate the engine(s) and propeller(s) with your wing. Do this in the top
view, side view, and front view.
b. Add appropriate fairings to ensure a smooth integration with the wing.
FT
If you have chosen a jet engine to propel your airplane, this section explains how to di-
mension your engine and nacelle and how to integrate them with the airframe. Here,
we only consider jet engines of the turbofan configuration, i.e. a turbojet engine with
an additional fan. A turbojet engine consists of three basic components: the compres-
sor, the combustion chamber, and the turbine. The thrust of the engine is limited by the
amount of compression that the compressor can achieve and the temperature that can
be achieved in the combustion chamber. We call the ratio between the inlet (total) pres-
sure and the total pressure at the exit of the compressor the overall pressure ratio (OPR).
RA
The maximum temperature that can be achieved in the combustion chamber is limited
by the materials and cooling systems used in the turbine. Therefore, the turbine inlet
temperature is an important design parameter of a turbojet or turbofan engine.
In a turbofan engine, part of the ingested air passes only through the fan before ex-
iting the engine through a nozzle. We call this the bypass flow. The other part passes
through the fan and through the core of the engine, i.e. the part with compressor, com-
bustion chamber, turbine, and nozzle. We call this the core flow. The ratio of the bypass
flow over the core flow is called the bypass ratio. A turbofan engine is typically installed
D
in a nacelle, which is a streamlined casing around the engine. The dimensions of the
nacelle are required for the integration of the engine with the airframe.
We first explain how a turbofan engine and its nacelle can be sized based on the
thrust requirement and a handful of design choices. Then, we show how turbofan en-
gines can be integrated with a wing. Finally, we show how to integrate a turbofan engine
with the fuselage.
turbine inlet temperature (TIT) is, therefore, another important design parameter of the
engine. To determine the required size of the turbofan, we follow the method of Toren-
beek and Berenschot [17], who size the turbofan based on these parameters along with
the required take-off thrust, TTO .
At take-off conditions, the thrust of a jet engine is proportional to the exhaust mass
flow (ṁ) and the average jet velocity. Because the mass flow is an important parameter
for the size of the engine, we compute it first:
TTO 1+B
ṁ = (8.22)
Ne a 0 5η G ¡1 + η B ¢
q
noz tf
where Ne is the number of engines, a 0 is the speed of sound at sea level, η noz is the nozzle
efficiency, G is the gas generator function, η tf is the combined efficiency of the turbine
and fan, and B is the bypass ratio.
Let us go through the parameters in Equation 8.22. The number of engines Ne is a
design parameter that you already selected in Assignment 4.3. The speed of sound at sea
FT
level (ISA) is 340 m/s. The bypass ratio is the ratio between the mass flow through the by-
pass duct and the mass flow going through the core of the engine. The nozzle efficiency
is the average value of the nozzle efficiency of the bypass nozzle and the core nozzle. This
efficiency typically ranges between 97% and 99%, depending on the aerodynamic design
of the nozzles. The turbine and fan efficiency, η tf , is the product of the turbine efficiency
η t and the fan efficiency, η f . A typical value of η tf = 0.75 can be assumed. Finally, G is
the gas generator performance expressed in non-dimensional form. The gas generator
RA
function G is statistically related to the turbine inlet temperature according to [17]:
Tt4
G≈ − 1.25 (8.23)
600
where Tt4 is the TIT measured in Kelvin. The higher Tt4 , the lower the mass flow per unit
of thrust and the smaller the engine. It is therefore desired to let the TIT be as high as
possible if you wish to minimize the mass, size, and drag of the engine. To increase the
TIT, turbine blades are therefore made out of heat-resistant metals and are often actively
D
cooled from the inside. Modern jet engines (in 2023) have a TIT somewhere between
1830 K and 1970 K.
The combustion temperature also influences the emissions of the engine. Combust-
ing fuel produces carbon dioxide (CO2 ), which is a greenhouse gas. The amount of CO2
is linearly correlated to the fuel consumption and independent of the TIT. The formation
of nitrogen oxides, on the other hand, is dependent on the TIT. Nitrogen oxides (NOx ) are
formed when nitrogen and oxygen molecules, which naturally occur in air, are exposed
to the high temperature and pressure of the combustion chamber. When emitted at high
altitudes, NOx has an indirect global warming effect, similar to CO2 . The higher the TIT,
the more NOx is produced. Therefore, modern engines feature combustion chambers
that allow for a lean mixture of air and fuel to be combusted. This reduces the peak
temperatures within the combustion chamber and reduces the formation of NOx with
respect to rich combustion. More information on the effect of the various combustion
species on global warming can be found in Ref. [8].
A SSIGNMENT 8.13
In this assignment, you will size the mass flow of your jet engine.
a. What bypass ratio, B , did choose for your engine in Assignment 6.4?
b. What turbine inlet temperature, Tt4 do you choose for your gas generator?
c. What values do you assume for the combined turbine and fan efficiency,
η tf , the nozzle efficiency, η noz , and the gas generator property, G?
d. Compute the mass, ṁ flow through your jet engine.
Now that we have shown how to determine the mass flow of your jet engine, let us start
sizing the nacelle. The nacelle of the jet engine is similar to the engine envelope that we
defined for the turboprop, electric motor, and piston engine. However, here we distin-
guish two types of nacelle geometries:
• Type B: the nacelle completely covers the bypass flow and gas turbine core. The
bypass flow and core flow are mixed ahead of the exhaust. This forced mixing re-
sults in lower thrust-specific fuel consumption and reduced noise. However, the
In Figure 8.18, nacelle Type B is shown while Figure 8.19 shows Type C. Note that the
inlet is slightly drooped downward for both nacelle designs. This is to emphasize that a
practical nacelle design is not axisymmetric. However, in the context of this textbook, we
will approximate the nacelle by an axisymmetric body. Type A nacelles are reserved for
RA
turbojet engines, which are outside the scope of this textbook.
D
Figure 8.18: Turbofan installation Type B, after [17] Figure 8.19: Turbofan installation Type C, after [17]
A SSIGNMENT 8.14
Which type of nacelle do you choose for your turbofan engine? Explain your de-
cision.
In Figure 8.20, you can see a schematic nacelle geometry of Type C with all relevant di-
mensions labeled. Ahead of the turbofan engine is the inlet. The function of the inlet
is to reduce the speed of the incoming flow such that the fan can compress it with high
efficiency. The inlet’s highlight is the leading edge perimeter, and the highlight plane is
the imaginary plane through the highlight perimeter. Like a propeller, the engine fan
has a spinner that guides the flow towards the fan blades. The bypass flow (or fan flow)
is enclosed by the fan cowling. For a Type B nacelle, the fan cowling would extend to the
end of the nacelle, and the engine core would be inside it. However, in this example, part
of the engine core is visible and has its own cowling. Finally, a cone (sometimes referred
to as plug) is located behind the core nozzle.
Deg
Def
Dg
Dc
Dn
Dh
Ds
Di
spinner
βf
n FT g
Figure 8.20: Definitions of the dimensions of a jet engine nacelle. After Ref. [17].
c
cone
RA
Based on the computed mass flow (Equation 8.22), we are now going to compute the
value of each of the nacelle dimensions of Figure 8.20 following Ref. [17]. The spinner-
to-inlet ratio can be computed as follows:
Ds ρ 0 a0
µ ¶
3B
≈ 0.05 1 + 0.1 + (8.24)
Di ṁ 1+B
u ρ a + 0.0050
0 0
D i = 1.65u
t ³ ´2 (8.25)
1− DD
s
i
We assume that the highlight diameter and the inlet diameter are the same, i.e.:
Dh = Di (8.26)
The nacelle length (excluding the cone), can be computed according to:
Ãs !
ṁ 1 + 0.2B
ln = Cl + ∆l (8.27)
ρ 0 a0 1 + B
Type Cl ∆l β
B 9.8 0.05 0.35
C 7.8 0.1 0.21 + p0.12
ϕ−0.3
values for either nacelle type. Based on the provided values, you can see that nacelle
Type B is 20%-25% longer than nacelle Type C. This allows for the integration of a mixer
behind the turbine of the engine.
The nacelle length equals the fan cowl length for Type B nacelles. For Type C nacelles,
the fan cowl length is shorter. You have to choose how much shorter the fan cowl is by
specifying a value for ϕ, the ratio between fan cowl length and nacelle length. A long cowl
length allows for the integration of an acoustic liner on the inside of the nacelle to reduce
the jet noise. However, it also increases the contact area between the flow through the
bypass duct as well as the external flow around the cowling. The latter contributes to the
FT
friction drag of the airplane. Typically, a value of 0.5 < ϕ < 0.8 is selected. The fan cowl
l f = ϕl n
Independent of the nacelle type, the maximum nacelle diameter is computed as follows:
D n = D i + 0.06ϕl n + 0.03
The location of the maximum nacelle thickness is a fraction, β, of the nacelle length.
(8.28)
(8.29)
RA
The value of β is dependent on the cowl type and typical values are suggested in Table
8.1. The length between the highlight and the position of maximum thickness is βl f as
shown in Figure 8.20.
The exit diameter of the fan cowl is:
µ ¶
1
D ef = D n 1 − ϕ2 (8.30)
3
For Type B nacelles (i.e. ϕ = 1), the diameter of the fan exit is the same as the exit diam-
D
eter of the nacelle. For Type C nacelles, we can continue to compute the dimensions of
the exposed cowling of the gas turbine:
l g = (1 − ϕ)l n (8.31)
The cowling diameter of the exposed gas-turbine core at the fan cowling exit is:
!2
0.089 ρ ṁ B + 4.5
Ã
0 a0
D g = D ef (8.32)
0.067 ρ ṁ
0 a0
B + 5.8
D eg ≈ 0.55D g (8.33)
Finally, the cone diameter can be related to the gas-turbine exit diameter:
D c ≈ 0.55D eg (8.34)
l c = 1.5D c (8.35)
With all the nacelle dimensions computed, we can now proceed to draw the nacelle. In
the following example, you see the step-by-step procedure that is followed to draw the
nacelle.
Example 8.14
In this example, we draw the nacelle for the jet airplane of Example 7.10. We choose the
nacelle to be of Type C and to have a fan cowl fraction of ϕ = 0.75. Selecting a TIT of
FT
Tt4 = 1680, we compute ṁ = 380 kg/s per engine when it produces maximum thrust at
sea level. Based on this mass flow, we size the nacelle according to the following steps.
Step 1 Compute inlet diameter from the mass flow. Using (8.24) and (8.25), we com-
pute D s /D i = 0.19 and D i = 1.6 m.
Step 2 Draw highlight diameter (neglect droop). Assuming the highlight diameter and
inlet diameter are the same, we have D h = 1.6 m.
RA
Step 3 Compute nacelle length. Using (8.27) and using the values from Table 8.1 under
Type C, we compute l n = 4.8 m.
Step 5 Compute position of maximum diameter. We compute β = 0.39 using the for-
mula in Table 8.1. The maximum diameter of the nacelle is therefore found at
βl f = 1.4 meters from the highlight plane.
D
Step 6 Compute maximum diameter. The maximum diameter is computed using (8.29)
and is D n = 1.9 m.
Step 7 Compute fan cowling diameter at the exit. We employ (8.30) to find D ef = 1.5 m.
Step 8 Compute gas turbine diameter at fan cowling exit.2 By employing (8.32) we find
D g = 1.0 m.
Step 9 Compute gas turbine diameter at exit. Using (8.33), we have D eg = 0.56 m.
Step 10 Compute cone dimensions. Using (8.34) and (8.35), we have D c = 0.36 and
l c = 0.54 m.
Step 11 Using all the computed dimensions, draw an outline of the nacelle.
Di = Dh
Deg
Def
Dg
Dc
Dn
βf c
f g
n 0 0.5 1
FT
RA
0 0.5 1
In the previous example, you saw how a nacelle can be drawn. Note that we use the
symmetry line (dash-dotted) as a starting point for the drawing because we assume the
nacelle is axisymmetric. In the following assignment, you can practice with the design
steps of Example 8.14
D
A SSIGNMENT 8.15
In this assignment, you will calculate the dimensions of your turbofan nacelle
and make a drawing of it.
a. Choose a cowl fraction, ϕ for your nacelle.a
b. Using the steps from Example 8.14, compute all the relevant
dimensions of your nacelle and fill out the following table:
Dh ln lf βl f D n D ef D g D eg D c lc
... ... ... ... ... ... ... ... ... ...
c. Produce a drawing of your nacelle. Use the same scale as you used for your
fuselage (Assignments 5.3 and 5.4) and wing (Assignments 8.3 and 8.8).
a If you have chosen nacelle Type B in Assignment 8.14, ϕ ≡ 1
2y e /b w = 0.7.
2. L ONGITUDINAL POSITIONING FT
i. With one engine per wing half, a typical lateral position is 35% of the semi-
ii. With two engines per wing half, a typical lateral position is 2y e /b w = 0.4 and
i. The engine should be positioned, such that rotor fragments have a less than
5% chance of causing a catastrophic failure in case of turbine disk failure (see
Example 3.17). Since a fuel tank is often located behind the front spar, it is ad-
RA
vised to keep the rotating parts of the engine ahead of the front spar. In the
case of a Type B nacelle, it can be assumed that the aft 15% of the nacelle
comprises the nozzle and the mixer. For Type C nacelles, it can be assumed
that the last 10% of the nacelle length comprises the (core) nozzle. Alterna-
tively, dry bays can be introduced in the wing to mitigate the risk of catas-
trophic failure in case of a turbine disk failure.
ii. To reduce the aerodynamic interference between the nacelle, pylon, and wing,
the engine should be positioned as far forward as possible.
D
iii. To minimize the mass of the pylon, the engine should be positioned as close
as possible to the wing box.
3. V ERTICAL POSITIONING
i. The closer the engine is positioned to the wing, the stronger is the aerody-
namic interference between wing, pylon and nacelle.
ii. A lower engine position reduces the ground clearance. Similar to Example
3.16, a minimal distance between the nacelle and the ground should be re-
spected. A lower vertical position could therefore result in a taller landing
gear to respect this ground clearance.
iii. A higher engine position reduces the distance between the thrust vector and
the center of gravity of the airplane, thereby reducing the thrust-induced
pitching moment.
In the following example, we show how the nacelle of Example 8.14 can be integrated
Example 8.15
In this Example, we integrate the nacelle of Example Example 8.14 with the wing of Ex-
ample 8.5, We take the following steps:
Step 1 We choose a location at 20% and 70% of the local chord for the front spar (fs)
and rear spar (rs), respectively. We draw the spar locations in the top view of the
wing by means of dashed lines (see below).
FT
Step 2 We choose the lateral location of the engine to be 35% of the semi-span of the
wing, and we mark this location in the top view (see below).
Step 3 We assume the last 10% of the nacelle length contains no rotating components,
and we choose the position of the engine inlet to be 90% of the nacelle length
ahead of the front spar (see below). This would allow a fuel tank between the
RA
front and rear spar to be sufficiently protected from turbine disk failure.
Step 5 We draw the nacelle in the front view of the wing at the same lateral position as
D
in the top view. We produce a circular nacelle shape and we choose to have the
upper surface of the nacelle slightly below the lower surface of the wing in front
view.
Step 6 In the front view of our drawing, we measure the vertical offset of the engine
with respect to the leading edge of the center airfoil: ∆z e = 0.85 m.
Step 7 We copy the longitudinal position of the engine from the top view to the side
view, and we take the vertical position with respect to the leading edge of the
center airfoil from the front view. We draw the engine in the side view of the
wing.
Δze
rs
fs
top view
front
view 0.35bw
0.90ln
0 1 2 [m] Δze
FT side view
RA
The previous example has shown how you can integrate a nacelle with a wing. At mul-
tiple steps, you have to make a choice. For every choice that you make, you should be
able to provide a justification. The content from the first paragraph of this section can
be used to provide justification. However, you might have other reasons that guide your
choice. In any case, each and every decision that you make throughout the design pro-
cess should have some justification such that you can explain why you have made a cer-
tain choice.
D
A SSIGNMENT 8.16
If your airplane has jet propulsion and you have chosen your engines to be under
the wing, perform the following tasks:
a. Integrate the engine(s) with your wing. Do this in the top view, side view,
and front view.
b. Add lines to indicate the location of the pylon in the front view.
the rear of the fuselage, as shown on the right-hand side of Figure 8.21. In this section,
we focus on the latter configuration and present how you can integrate twin jet engines
with the rear of the fuselage. This configuration is common for transport aircraft as well
as for business jets.
Figure 8.21: Installation of jet engines at the front of the fuselage (left) and at the rear of the fuselage. Photos
by: JuergenKlueser c b a and Lars Söderström c b a, respectively.
FT
When we determine the position of jet engines on the rear fuselage, we take the fol-
i. Rotor disk failure in the jet engine can cause a hazard to the fuselage. Rotor
fragments can cause holes in the fuselage, which leads to depressurization.
In addition, hazards to passengers might arise if they are seated between the
engines. Therefore, it is desirable to position the engines aft of the cabin.
RA
ii. Cabin noise can be quite high when the engines are attached to the fuselage.
The short structural path between the engine and the fuselage frames can
cause the noise to travel through the structure into the cabin. The more aft
the engines are positioned, the lower the cabin noise.
iii. Emergency exits can be located at the rear of the fuselage. The engine should
not block the emergency exit. The location of the emergency exit has been
chosen based on the requirement to have the distribution as “uniformly as
practical” (Example 3.15). If necessary, it is, therefore, possible to change the
D
location of the aft-most pair of emergency exits to allow the integration of the
jet engine.
iv. The aerodynamic interference between the wing and the inlet should be con-
sidered. As the velocity over the wing increases, the inlet can be exposed to
a (very) high Mach number when it is positioned over the wing. Therefore, it
is desirable to have the inlet positioned behind the wing. We will see that for
business jets, it can be inevitable to have the inlet slightly over the rear part
of the wing.
v. Jet engines are relatively heavy systems. If they are positioned on the rear of
the fuselage, they cause a large center-of-gravity excursion during the load-
ing of passengers and fuel (see Chapter 9). The center-of-gravity excursion
affects the required size of the tail as well as the landing gear integration. It
is therefore desired to have heavy systems, such as the engines, as close as
possible to the center of gravity of the airplane. For rear-fuselage-mounted
engines, this would translate into the desire to position them as far forward
as possible.
In the reference frame of the fuselage, we use capital X to denote the longitudinal
position along the fuselage. X h is the longitudinal position of the highlight plane
of the engine inlet (see Section 8.3.1). Given these considerations, we propose to
abide by the following guidelines regarding the longitudinal position of the engine.
We want X h to be as small as possible, subject to the following constraints:
X h > X apb − 0.5l n and X h > X exit + D h for transport jets (8.36)
X h > X apb + 0.5D h for business jets (8.37)
where X abp is the longitudinal location of the aft pressure bulkhead and X exit is the
location of the aft-most pair of emergency exits.
2. V ERTICAL POSITIONING
i. Position the engine, such that the thrust vector with respect to the center
of gravity is as small as possible. This minimizes the (nose-down) pitching
FT
moment that is introduced by the thrust of the engine. In practice, this means
positioning the engine close to the middle of the fuselage.
ii. Make sure the engine inlet is (barely) above the wing. This ensures that the
wing wake is not ingested by the engine.
iii. Make sure that the exhaust does protrude below the belly curve of the fuse-
lage. This prevents the engine from scraping the runway when the airplane
rotates during the take-off roll.
RA
3. L ATERAL POSITIONING
i. Since the engine is mounted to the fuselage, the one-engine-inoperative yaw-
ing moment is relatively small. The closer the engine is to the fuselage, the
smaller the yawing moment.
ii. The closer the engine is located to the fuselage, the smaller the pylon and the
lower the structural mass of the pylon can be.
iii. The inlet should be offset from the fuselage to prevent the boundary layer of
the fuselage from entering the engine. This would cause a reduction in inlet
D
In this example, we show how to integrate the engine of Example 8.14 with the fuselage
of Example 5.8. We take the following steps:
Step 1 Since our airplane is a transport jet, we use Equation 8.36 to determine the lon-
gitudinal location of the highlight. There is no emergency exit at the rear of the
airplane, so the placement of the engine does not interfere with the exits.
Step 2 To minimize the thrust-induced pitching moment, we propose to the vertical
position to be in the middle of the fuselage. So the engine, is positioned 0.5h −
fus above the bottom of the fuselage.
Step 3 With the longitudinal and vertical position known, we draw the engine into the
side view of the fuselage (see below). We verify that the nacelle does not pro-
trude beyond the tail cone perimeter. This is not the case.
Step 4 We determine the clearance between the engine and the fuselage to measure
25% of the nacelle diameter. We measure the resulting lateral displacement of
the engine, which is ∆y = 3.1 m.
Step 5 We draw the engine in the top view at the appropriate location and add a pylon
geometry (see below).
Step 6 We draw the engine in the front view and also add lines representing the pylon
(see below).
front view
ΔYe
top view
FT 0 2 4[m]
ΔYe
0.25Dn
RA
0.5hfus side view 0.5ln
0.5hfus
Note that in the previous example, we only draw one engine in the top view and side
D
view. Naturally, we there should also be an engine on the opposite side, as the airplane
is symmetric. The motivation to draw only half of the airplane is to reduce the size of the
drawing, such that we can use the available space on our paper most efficiently.
A SSIGNMENT 8.17
If your airplane has jet propulsion and you have chosen your engines to be at-
tached to the fuselage, perform the following tasks:
a. Integrate the engine(s) with your fuselage. Do this in the top view, side
view, and front view.
b. Add lines to indicate the location of the pylon in the top view and the front
view.
FT
9.1. G ENERATION OF S IMPLE L OADING D IAGRAM
9.2. L ONGITUDINAL W ING P OSITIONING
9.3. U NDERCARRIAGE D ISPOSITION
RA
9.4. E MPENNAGE S IZING
D
211
D
RA
FT
10
A IRCRAFT A NALYSIS AND D ESIGN
I TERATION I
213
D
RA
FT
11
W ING D ESIGN R EVISITED AND
D ESIGN I TERATION II
FT
11.1. D ESIGN OF R OLL C ONTROL S URFACES
11.2. D ESIGN OF H IGH -L IFT S YSTEM
11.3. E STIMATION OF M AXIMUM L IFT C OEFFICIENT
RA
11.4. R ESULT OF D ESIGN I TERATION ON D ESIGN O BJECTIVE
11.5. E FFECT OF W ING D ESIGN PARAMETERS ON D ESIGN O B -
JECTIVE
D
215
D
RA
FT
12
E MPENNAGE D ESIGN R EVISITED
AND D ESIGN I TERATION III
FT
12.1. S TABILITY AND C ONTROL R EQUIREMENTS
12.2. C OMPONENT W EIGHT E STIMATION
12.3. G ENERATION OF L OADING D IAGRAM
RA
12.4. S IZING FOR LONGITUDINAL STABILITY, EQUILIBRIUM AND
CONTROL
12.5. L ONGITUDINAL W ING P OSITIONING R EVISITED
12.6. S IZING FOR DIRECTIONAL STABILITY, EQUILIBRIUM , AND
CONTROL
D
217
D
RA
FT
13
AUTOMATION OF THE A IRPLANE
D ESIGN P ROCESS
P ROBLEM
FT
13.1. F ORMALIZING THE C ONSTRAINED D ESIGN O PTIMIZTION
219
D
RA
FT
B IBLIOGRAPHY
[1] Airbus. A320 Aircraft Characteristics Airport and Maintenance Planning. Tech. rep.
Blagnac, France, Dec. 2020.
[2] T. C. Corke. Design of Aircraft. Prentice Hall, 2003. ISBN: 9780130892348.
[3] EASA. Acceptable Means of Compliance (AMC) and Guidance Material (GM) to An-
nex VIII Specialised Operations [PART-SPO]. Tech. rep. 31 October 2016.
[4] J. D. Mattingly et al. Aircraft Engine Design. Reston, VA: American Institute of Aero-
nautics and Astronautics, 2018.
[5] M. C. Y. Niu. Airframe Structural Design: Practical Design Information and Data on
[6]
[7]
[8]
FT
Aircraft Structures. 2nd ed. Adaso Adastra Engineering Center, 1999.
E. Obert. Aerodynamic Design of Transport Aircraft. Delft: IOS Press, 2009.
Pieter Jan Proesmans. “Aircraft Design for Minimal Climate Impact (working ti-
tle)”. PhD thesis. Delft University of Technology, 2024.
Pieter-Jan Proesmans and Roelof Vos. “Airplane Design Optimization for Minimal
Global Warming Impact”. In: Journal of Aircraft 59.5 (2022), pp. 1363–1381. DOI:
10.2514/1.C036529.
RA
[9] D. P. Raymer. Aircraft Design: A Conceptual Approach. Washington, DC: AIAA, 1989.
[10] J. Roskam. Airplane Design Part I: Preliminary Sizing of Airplanes. Lawrence, KS,
USA: DARcorporation, 1989.
[11] J. Roskam. Airplane Design, Part VI: Preliminary Calculation of Aerodynamic, Thrust
and Power Characteristics. Lawrence, KS: DARcorp, 1990.
[12] G. J. J. Ruijgrok. Elements of Airplane Performance. Delft, The Netherlands: Delft
University Press, 1989.
D
[13] E. Torenbeek. “Elements of Aerodynamic Wing Design”. In: Advanced Aircraft De-
sign. Chichester, West Sussex, UK: John Wiley and Sons, Ltd., 2013. Chap. 10.
[14] E. Torenbeek. Essentials of Supersonic Commercial Aircraft Conceptual Design. Chich-
ester, Uk: John Wiley & Sons Ltd., 2020.
[15] E. Torenbeek. Synthesis of subsonic airplane design. Delft, The Netherlands: Delft
University Press, 1976.
[16] E. Torenbeek. The initial calculation of range and mission fuel during conceptual
design. en. Tech. rep. LR-525. Delft University of Technology, Faculty of Aerospace
Engineering, 1987.
[17] E. Torenbeek and G. H. Berenschot. De berekening van het omspoeld gondelop-
pervlak van enkel- en dubbelstroom straalmotoren voor civiele vliegtuigen. Tech.
rep. Memorandum M-445. Technische Hogeschool Delft, afdeling Luchtvaart- en
Ruimtevaarttechniek, 1983.
221
222 B IBLIOGRAPHY
[18] E. Van Hinte and M. Van Tooren. First Read This. nai010 publishers, 2013. ISBN:
9064506434.
[19] Roelof Vos and Saeed Farokhi. Introduction to Transonic Aerodynamics. Vol. 110.
Fluid Mechanics and Its Applications. Dordrecht: Springer Netherlands, 2015. ISBN:
978-94-017-9746-7. DOI: 10.1007/978- 94- 017- 9747- 4. URL: http://link.
springer.com/10.1007/978-94-017-9747-4.
[20] R. de Vries, M. Hoogreef, and R. Vos. “Range Equation for Hybrid-Electric Aircraft
with Constant Power Split”. In: Journal of Aircraft 57 (3 2020).
[21] Reynard de Vries, Malcom Brown, and Roelof Vos. “Preliminary Sizing Method
for Hybrid-Electric Distributed-Propulsion Aircraft”. In: Journal of Aircraft 56.6
(2019), pp. 2172–2188. DOI: 10 . 2514 / 1 . C035388. URL: https : / / doi . org /
10.2514/1.C035388.
[22] R. Whitford. “Chapter 3: Fuselage Design”. In: Design for Air Combat. London, UK:
Jane’s, 1987, pp. 148–160.
FT
RA
D
223
224 I NDEX
need, 16
seed, 16
sideslip, 46
FT tractor configuration, 52
TSFC, 107
turbine inlet temperature, 199
turbofan, 25, 51, 109
turboprop, 25, 51
ULD, 79
unit load device, 77
RA
sizing, 3, 12, 164
preliminary, 99 unity equation, 100
span efficiency factor, 106 upsweep, 73
specific energy, 100
speed vertical tailplane, 61
cruise, 135 VFR, 38, 39
minimum, 128 view angle
stall, 25 downward, 72
sponson, 46 upward, 72
D