Sie sind auf Seite 1von 59

Flight & orbital mechanics summary: Flight

mechanics
Based on lectures by Dr. Ir. M. Voskuijl

Sam van Elsloo


February - April 2017

Version 1.0
2

©Sam van Elsloo


Preface
Note that I haven’t included any examples this time. The reason for this is that a) I don’t have time to include and
b) you should really just practice with old exam questions (it’s literally the only practice material available, and
there are full solutions available for them already so no real point in me making them). It is for this reason that I
instead made a section at the end of each chapter (except chapter 1) saying which exams you could practice with
for each chapter.
Furthermore, I already stated it in the preface of the propulsion and power summary: I’d like you to keep
something in mind regarding the summaries and solution manuals: when I made a solution manual and excel
sheet for the fourth propulsion bonus assignment, I sent it to a handful of people, but unfortunately it got leaked,
and after that it was forwarded a few times by other people. This was absolutely not my intention at all and you
really shouldn’t do so. Let me make it as clear as I can be: if I don’t publish something on facebook, you are
not allowed to forward it without my explicit permission. This holds for any kind of document (summary,
solution manual or something else) for any course. Bear in mind that if you do so anyway, literally the only
options for me are to do either do nothing at all (which would be rather unfair to me in my opinion), or to not
publish the next summary/solution manual at all.
Finally, I’ve split this summary into two separate documents: one for flight mechanics, and one for orbital
mechanics, because otherwise it’d just become too long for me to recompile every time.
But at least you now have not one, but two beautiful front pages to print.

The summary is not completely finished yet; I need to add the lecture on wind gradients still (idk when I’ll do
that, so don’t bother me asking when I’ll add it). However, this is a rather short chapter and is barely ever asked
(out of 500 points you could achieve in past flight mechanics exams, only 5 or 10 or so points were about wind
gradients).

3
4

©Sam van Elsloo


Contents

1 Refreshment 7
1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.2 Equations of motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.3 Performance diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.3.1 Jet aircraft . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.3.2 Propeller aircraft . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.3.3 Effect of altitude . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2 Lecture 1 15
2.1 Steady climb performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.1.1 Maximum rate of climb . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.2 How do we climb? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.2.1 Rate of climb . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.3 Total time and fuel consumption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.4 Minimum time to climb . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.4.1 Introduction of energy height . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.4.2 Use of energy height . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.4.3 Supersonic climb . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.5 Descent . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.6 Gliding flight . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.7 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.8 Relevant old exam questions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.8.1 Full exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.8.2 Parts of exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.8.3 Exercises combining previously discussed material . . . . . . . . . . . . . . . . . . . 25

3 Turning flight 27
3.1 Equations of motion and load factor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.2 Standard turns . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.2.1 Turn coordinator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.3 Maximum aircraft turning performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.3.1 Steepest turn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.3.2 Minimum turn radius . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.3.3 Minimum time to turn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.4 Altitude effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.5 Relevant old exam questions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.5.1 Full exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.5.2 Parts of exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.5.3 Exercises combining previously discussed material . . . . . . . . . . . . . . . . . . . 33

4 Airfield performance 35
4.1 Balanced field length . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
4.2 Take-off performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
4.2.1 Ground run . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
4.2.2 Airborne distance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.3 Landing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
4.3.1 Airborne phase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
4.3.2 Ground run distance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4.3.3 Brake enhancers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
4.4 Factors neglected so far . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
4.4.1 Wind . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

5
CONTENTS 6

4.4.2 Runway slope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42


4.5 Take-off speeds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
4.6 Relevant old exam questions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.6.1 Full exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.6.2 Parts of exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.6.3 Exercises combining previously discussed material . . . . . . . . . . . . . . . . . . . 43

5 Cruise 45
5.1 Optimum cruise for simplified jet aircraft . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
5.1.1 Optimum flight condition for a single point in time . . . . . . . . . . . . . . . . . . . 45
5.1.2 Cruise strategy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
5.1.3 Range equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
5.1.4 Operational limitations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
5.2 Range for propeller aircraft . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
5.3 Unified range equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
5.3.1 Jet aircraft . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
5.3.2 Propeller aircraft . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
5.4 Economics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
5.4.1 Operating costs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
5.5 Wind . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
5.6 Cruise at transonic flight conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
5.7 Relevant old exam questions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
5.7.1 Full exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
5.7.2 Parts of exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
5.7.3 Exercises combining previously discussed material . . . . . . . . . . . . . . . . . . . 59

©Sam van Elsloo


1 Refreshment
This chapter focusses on stuff of last year.

1.1 Introduction

D EFINITION Two performance parameters:


• Rate of climb: vertical component of velocity: 𝑉 sin (𝛾) = 𝑅𝐶.
• Climb angle: flight path angle.
A few flight angles:
• Angle of attack: angle between nose and speed vector; indicated by 𝛼.
• Flight path angle: angle between speed vector and horizontal; indicated by 𝛾.
Two optimal speeds indicated in the manual of an aircraft.
• 𝑉𝑥 : IAS (indicated airspeed) for maximum climb angle.
• 𝑉𝑦 : IAS for maximum rate of climb.
These are generally not the same.

The first four definitions are also depicted in figure 1.1.

Figure 1.1: Flight angles and rate of climb (𝑅𝐶)

Another set of definitions:

D EFINITION Symmetric flight means that the entire flight trajectory can be described in a 2D plane; i.e. he angle of
slideslip is zero and the plane of symmetry of the aircraft is perpendicular to the Earth (so there is
neither yaw nor roll). This means that the roll angle 𝜇 = 0 and the turn rate 𝑑𝜉∕𝑑𝑡 = 0.

Steady flight means that 𝑑𝑉 ∕𝑑𝑡 = 0.

Straight flight means that 𝑑𝛾∕𝑑𝑡 = 0.

1.2 Equations of motion


Let us derive the equations of motion for symmetric flight (but we do not assume steady flight). This means that
we only have to deal with 2D equation of motions (making our lives easier). Now, the equation of motions start
from F = 𝑚a, so let’s start by identifying a. To do this, we use figure 1.2a: here, the altitude is indicated by the
vertical direction; the horizontal distance is indicated by the horizontal direction. At a time 𝑡1 , the velocity vector
is V1 ; at a time 𝑡2 , it is V2 . The flight angle 𝛾 is simply the angle between V and the horizontal. Additionally, ̂i𝑎
and k̂ 𝑎 are indicated for both points in time: ̂i𝑎 is the unit vector along the velocity vector; k̂ 𝑎 is the unit vector

7
CHAPTER 1. REFRESHMENT 8

perpendicular to V, pointing down to Earth. The reason for the subscript 𝑎 will become clear when we’re gonna
draw the FBD and KD.
Now, we obviously have
𝑑V
a=
𝑑𝑡
Furthermore, what happens when 𝑑𝑡 → 0 is indicated in figure 1.2b: the difference in angle is simple 𝑑𝛾. We
then have, as V = 𝑉 ̂i𝑎 , we have, by virtue of the chain rule:
( )
𝑑 𝑉 ̂i𝑎 𝑑 ̂i
𝑑V 𝑑𝑉 ̂
a= = = i𝑎 + 𝑉 𝑎
𝑑𝑡 𝑑𝑡 𝑑𝑡 𝑑𝑡
So, the acceleration in the flight direction is simply equal to 𝑑𝑉 ∕𝑑𝑡 (as indicated by the first term). The
acceleration normal to the flight direction is given by 𝑉 𝑑 ̂i𝑎 ∕𝑑𝑡, but we’d like get rid of this vector as it’s not
really something obvious. What we actually see from figure 1.2b is that

𝑑 ̂i𝑎 = −𝑑𝛾 k̂ 𝑎1

as 𝑑𝛾 is small, and 𝑑 ̂i𝑎 points in the opposite direction of k̂ 𝑎1 (hence the minus sign). We can thus write

𝑑𝑉 ̂ 𝑑 ̂i 𝑑𝑉 ̂ 𝑑𝛾 ̂
a= i𝑎 + 𝑉 𝑎 = i −𝑉 k
𝑑𝑡 𝑑𝑡 𝑑𝑡 𝑎 𝑑𝑡 𝑎1

(b) Sketch of velocity vectors and unit vectors very shortly


(a) Sketch of velocity vectors and unit vectors.
after each other.

Figure 1.2: Sketches.

So, now that we have clearly have decomposed a into one vector pointing in ̂i𝑎 direction and one in k̂ 𝑎1 direction,
we can focus on drawing the FBD. For this, it is first important to remember that we distinguish 4 reference
frames:
• Earth axis or ground system: axis system that is fixed on the Earth instead of the aircraft; denoted by
subscript 𝑔 and is never used actually (not in this course, at least).
• Moving Earth or local horizon system: axis system that moves along with plane, with its origin located at
the centre of gravity of the plane and for which the 𝑧-axis points downward, perpendicular to the Earth’s
surface; denoted by subscript 𝑒.
• Body axis system: axis system for which the 𝑥-axis points through the nose; denoted by subscript 𝑏.
• Air-path axis system: axis system for which the 𝑥-axis points along the flight direction; denoted by
subscript 𝑎.
The latter three coordinate system are also shown in figure 1.3a, together with the accelerations (note that
𝑑𝛾∕𝑑𝑡 = 𝜔 = 𝑉 ∕𝑅, but this is not an expression we’ll actually be using so just ignore it), to form the kinetic
diagram.
Then, from this FBD and KD, we can easily set up the equations of motion:

©Sam van Elsloo


9 1.3. PERFORMANCE DIAGRAM

(b) Free body diagram.

(a) Kinetic diagram. Note that the angle between 𝑥𝑏 and 𝑥𝑎


equals 𝛼, and the angle between 𝑥𝑎 and 𝑥𝑒 equals 𝛾.

Figure 1.3: Sketches.

FORMULAS: 𝑇 cos 𝛼𝑖 − 𝐷 − 𝑊 sin 𝛾 =


𝑊 𝑑𝑉
(1.1)
GENERAL 𝑔 𝑑𝑡
EQUATIONS OF 𝑊 𝑑𝛾
𝐿 − 𝑊 cos 𝛾 + 𝑇 sin 𝛼𝑖 = 𝑉 (1.2)
MOTION FOR 𝑔 𝑑𝑡
SYMMETRIC
FLIGHT
Now, we’ll make use of the small angle approximation to simplify matters a bit: first of all, we’ll say that
cos ≈ 1. Additionally, as 𝑇 will be comparatively small compared to 𝐿 and 𝑊 , and 𝛼𝑖 is typically small as
well, we’ll say that 𝑇 𝛼𝑖 ≈ 0 as well. Note that we do not use sin 𝛾 ≈ 𝛾! This is an approximation that does not
have any real benefits regarding computations, so we better just don’t make it. This leads to the approximated
equations of motion:

FORMULAS: 𝑇 − 𝐷 − 𝑊 sin 𝛾 =
𝑊 𝑑𝑉
(1.3)
APPROXI- 𝑔 𝑑𝑡
MATED 𝑊 𝑑𝛾
𝐿−𝑊 = 𝑉 (1.4)
EQUATIONS OF 𝑔 𝑑𝑡
MOTION FOR
SYMMETRIC
These are the equations of motion we’ll be using in the rest of this chapter, so if there’s an exam question
FLIGHT
about deriving the equations of motion, you already know how to (do note: I’m pretty sure you don’t need to
know how to derive the acceleration terms, you can just immediately write down what they are in the kinetic
diagram).

1.3 Performance diagram


We have two important performance diagrams that we’ll be seeing often in this course and it is essential that
you understand why these look like they look like. We can distinguish between jet aircraft and propeller
aircraft.

1.3.1 Jet aircraft

We have two performance diagrams of interest: a graph where force (both thrust and drag) are plotted against
airspeed, and a graph where powers (both power available, 𝑃𝑎 = 𝑇 𝑉 , and power required, 𝑃𝑟 = 𝐷𝑉 ) are plotted
against airspeed. The first graph is drawn in figure 1.4a, the second in figure 1.4b.
Now, let’s analyse all lines one by one:
• For a jet engine, thrust is more or less independent of airspeed, hence it’s a straight horizontal line.

samvanelsloo97@icloud.com
CHAPTER 1. REFRESHMENT 10

(a) Force vs. velocity diagram. (b) Power vs. velocity diagram.

Figure 1.4: Performance diagrams of a jet aircraft.

• For the drag, remember that we have

1
𝐷 = 𝐶𝐷 ⋅ 𝜌 ⋅ 𝑉 2 ⋅ 𝑆
2
where

𝐶𝐿2
𝐶𝐷 = 𝐶𝐷,0 + = zero-lift drag coefficient + induced drag coefficient
𝜋𝐴𝑒
so that essentially,

1 𝐶2
𝐷 = 𝐶𝐷,0 ⋅ 𝜌 ⋅ 𝑉 2 ⋅ 𝑆 + 𝐿 ⋅ 𝜌 ⋅ 𝑉 2 ⋅ 𝑆 = zero-lift drag + induced drag
2 𝜋𝐴𝑒
Now, what happens if we increase the velocity? Then quite clearly, the zero-lift drag increases with the
square of the velocity, as 𝐶𝐷,0 , 𝜌 and 𝑆 are all constant. However, the induced drag is slightly more
complex. If we increase the velocity, the lift coefficient must decrease, otherwise we’d create excess lift
and we’d start ascending which we do not want. In fact, we have the relation

1
𝐶𝐿 ⋅ 𝜌 ⋅ 𝑉 2 ⋅ 𝑆 = 𝑊
2
2𝑊
𝐶𝐿 =
𝜌𝑉 2 ⋅ 𝑆

Plugging this into the equation for the drag yields

1 4𝑊 2
𝐷 = 𝐶𝐷,0 ⋅ 𝜌 ⋅ 𝑉 2 ⋅ 𝑆 + = zero-lift drag + induced drag
2 𝜋𝐴𝑒𝜌𝑉 2 𝑆

so that clearly, there’s an hyperbolic relation between induced drag and velocity. This is sketched in figure
1.5. This explains why in figure 1.4a, the drag first decreases and then increases; also note that the drag
does not start at 𝑉 = 0, but rather at 𝑉 = 𝑉stall .
• The power available is simply 𝑃𝑎 = 𝑇 𝑉 , and thus the power available is a linear relation going through
the origin.
• For the power required, we have 𝑃𝑟 = 𝐷𝑉 : basically, this means that the drag-curve is shifted upwards,
but that for higher values of 𝑉 , it is shifted more upwards than for lower values of 𝐷. This also means
that the point at which the power required is minimum lays to the left (thus at a lower airspeed) than the
point at which the drag is minimum.

1.3.2 Propeller aircraft

For a propeller aircraft, a similar story can be told, looking at figure 1.6a and 1.6b:

©Sam van Elsloo


11 1.3. PERFORMANCE DIAGRAM

Figure 1.5: Variation of zero-lift drag and induced drag with velocity.

(b) Power vs. velocity diagram. The dotted line is the real case,
(a) Force vs. velocity diagram. the full line is the idealized case.

Figure 1.6: Performance diagrams of a propeller aircraft.

samvanelsloo97@icloud.com
CHAPTER 1. REFRESHMENT 12

• The thrust is inversely related to the velocity. In reality, it is as shown by the dotted line; idealized, it is as
the full red line.
• For the drag, the exact same story holds as for the jet aircraft: it does not matter a single fuck whether it’s
a jet or propeller aircraft to draw the 𝐷 − 𝑉 line.
• The power available is more or less constant with airspeed (note the difference with jet aircraft, there the
thrust was constant), except for very low and very high airspeeds, where performance significantly drops.
However, we won’t really deal with those regions most of the time. However, just remember that there’s a
significant performance drop for small speeds.
• The power required is again exactly the same as for jet aircraft.
Please, please remember these four graphs: they’re not that hard at all, but it saves me a lot of time explaining
everytime why the drag curve looks like that, etc. You should also be able to recognize whether it’s a propeller
aircraft or jet aircraft, although we will usually work with propeller aircraft in the lectures themselves (not
necessarily on exam questions, though). Please note that the same methods will still apply for jet aircraft, it’s
only on the "engine part" (i.e. thrust and power available) that things change, but if you are remotely smart you
should be able to deal with this.

1.3.3 Effect of altitude

With increasing altitude, the air becomes thinner, so 𝜌 decreases. This means that the thrust decreases, and thus
the thrust lines in figures 1.4a and 1.6a move downward, and the same happens for the power available lines in
figures 1.4b and 1.6b. What happens to the drag and power required is shown in figures 1.7a and 1.7b. First, the
drag can be explained as follows: if we increase the altitude, 𝜌 decreases, but the lift needs to stay the same. We
can take care of that by either increasing 𝐶𝐿 or by increasing 𝑉 . Assume we do it by letting 𝐶𝐿 stay constant
and increase 𝑉 . Then,

𝐶𝐿2
𝐶𝐷 = 𝐶𝐷,0 +
𝜋𝐴𝑒
will stay constant as well. What happens then with

1
𝐷 = 𝐶𝐷 ⋅ 𝜌 ⋅ 𝑉 2 ⋅ 𝑆
2
you may wonder? To maintain the same amount of lift, we must have

2𝑊
𝑉2 =
𝐶𝐿 𝜌𝑆

and thus
1 2𝑊 𝐶
𝐷 = 𝐶𝐷 ⋅ 𝜌 ⋅ 𝑆 = 𝐷𝑊
2 𝜌𝑆 𝐶𝐿

If 𝐶𝐿 and 𝐶𝐷 are both constant, this means that 𝐷 must be constant as well (as 𝑊 is also constant). Thus, the
drag is constant, but the velocity is increased. This means the drag curve simply shifts to the right.
The power required curve then simply follows from 𝑃𝑟 = 𝐷𝑉 : for larger values of 𝑉 , the drag is increased by a
larger number, meaning the curve moves upwards there more than for lower values of 𝑉 . This explains the top
right movement of figure 1.7b.
This basically concludes our basic knowledge of last year’s stuff. Especially this last section is important as I’ll
often use information in this section to quickly explain something.

©Sam van Elsloo


13 1.3. PERFORMANCE DIAGRAM

(a) Drag upon increasing altitude. (b) Power required upon increasing altitude.

Figure 1.7: Drag and power required upon increasing altitude.

samvanelsloo97@icloud.com
CHAPTER 1. REFRESHMENT 14

©Sam van Elsloo


2 Lecture 1

2.1 Steady climb performance

2.1.1 Maximum rate of climb

If we have steady climb (i.e., the accelerations are zero), our equations of motion simply become

𝑇 − 𝐷 − 𝑊 sin 𝛾 = 0
𝐿−𝑊 = 0

Then, if we multiply the first equation by 𝑉 , we get

𝑇 𝑉 − 𝐷𝑉 − 𝑊 𝑉 sin 𝛾 = 𝑃𝑎 − 𝑃𝑟 − 𝑊 𝑅𝐶 = 0

as 𝑃𝑎 = 𝑇 𝑉 (power available), 𝑃𝑟 = 𝐷𝑉 (power required) and 𝑅𝐶 = 𝑉 sin 𝛾 (rate of climb). This equation is
called the power equation. Now, if we assume that our pilot is capable enough of selecting 𝑇 and 𝑉 himself,
we are able to compute all of this:
• The drag 𝐷 is given by 𝐷 = 𝐶𝐷 ∕2 ⋅ 𝜌𝑉 2 ⋅ 𝑆, where only 𝐶𝐷 is unknown; but this is given by

𝐶𝐿2
𝐶𝐷 = 𝐶𝐷0 +
𝜋𝐴𝑒
where only 𝐶𝐿 is unknown.
• 𝐶𝐿 can be computed by solving the second equation, leading to
2𝑊
𝐶𝐿 =
𝜌𝑉 2 𝑆
• All other values are simply given to you in the question.
Using this, we can thus write that

FORMULA: 𝑃𝑎 − 𝑃𝑟
𝑅𝐶𝑠 = (2.1)
RATE OF CLIMB 𝑊
FOR STEADY where the 𝑠 subscript denotes the rate of climb in steady climb.
CLIMB

FINDING THE 1. Establish a coordinate system.


MAXIMUM 2. Draw the free body and kinetic diagram.
RATE OF CLIMB 3. Set up the (approximated) equations of motion.
FOR STEADY 4. Multiply the first equation by 𝑉 to set up the power equation,
CLIMB
𝑃𝑎 − 𝑃𝑟
𝑃𝑎 − 𝑃𝑟 − 𝑊 𝑅𝐶𝑠 = 0 → 𝑅𝐶𝑠 =
𝑊
5. Solve the second equation (normal to the flight path) to find 𝐶𝐿 .
6. Find 𝐶𝐷 by plugging in the numbers for
𝐶𝐿2
𝐶𝐷 = 𝐶𝐷0 +
𝜋𝐴𝑒
7. Find the maximum rate of climb by plugging in
𝑃𝑎 − 𝑃𝑟
𝑅𝐶𝑠 =
𝑊

15
CHAPTER 2. LECTURE 1 16

This is called the power equation. Now, it is important to understand what the power equation actually tells
us: it tells you that you have a power available delivered by the engine (taking into account all efficiencies
etc.). Part of this is wasted on drag (taken into account in 𝑃𝑟 ). The surplus is then converted into potential
energy (𝐸𝑝 = 𝑊 ⋅ 𝐻), where the rate of climb obviously is directly related to the potential energy. From the
performance diagram, we can easily find the RC, by simply taking the distance between the two curves and
dividing it by the aircraft weight, as shown in figure 2.1a. Doing this for all airspeeds leads to figure 2.1b.

(a) Power surplus. (b) Rate of climb.

Figure 2.1: Power surplus and rate of climb.

Note that the performance diagram only holds for one specific altitude: for higher altitudes, the power available
decreases as the air is thinner. The power required graph shifts to the right and upwards. This means that
the rate of climb decreases with increasing altitude, as shown in figure 2.2; clearly, the maximum distance
decreases.

Figure 2.2: Performance diagram with increasing altitude: white dots represent lower altitude, black dots
represent higher altitude.

This rather logically leads to the diagram shown in figure 2.3: for a fixed altitude, the rate of climb is a function
of airspeed; its maximum is somewhere in the middle, it reaches a minimum at very low and very high airspeeds
(as shown in figure 2.1b. Furthermore, the maximum rate of climb decreases for increasing altitude, and occurs
at a higher true airspeed than at lower altitude (as shown in figure 2.2).

Figure 2.3: Rate of climb as a function of altitude.

Finally, this leads to the flight ceiling is depicted in figure 2.4. It is clear that there is a lower velocity limit
on the left, initially caused by stall limits, but then by a thrust limit (remember that for lower velocities, drag
increased again: at high altitude, this means that thrust actually may become too small to fly at such a small
velocity, hence the thrust limit). Then there’s the theoretical ceiling and then there’s the thrust limit for the

©Sam van Elsloo


17 2.2. HOW DO WE CLIMB?

maximum velocity. At this theoretical ceiling, the power required curve has moved all the way up and the power
available curve all the way down so that there is no rate of climb possible any more. Furthermore, it should be
noted that the theoretical ceiling is never really achieved in life: just before the theoretical ceiling, your rate of
climb is very low, meaning it takes ages to get to there. Furthermore, if a gust appears, you’re fucked and your
altitude quickly lowers. Instead, we usually speak of a service ceiling, slightly below the theoretical ceiling, but
at the service ceiling, we can actually fly in a steady flight.

Figure 2.4: Theoretical ceiling.

2.2 How do we climb?


So far, we assumed steady flight: the true airspeed is constant. However, for a change, let’s keep the indicated
airspeed (IAS) constant, and let’s see whether how well the performance then is (we never actually discussed
whether steady climb was the optimal way to climb). Before we start, let’s recall what indicated vs. true airspeed
meant. The indicated airspeed is also known as the calibrated airspeed or the equivalent airspeed (calibrated for
compressible flow, equivalent for incompressible flow). For equivalent airspeed, we have

𝜌0
𝑉TAS = 𝑉
𝜌 EAS

so the true airspeed will be higher than the equivalent airspeed anywhere above sea-level. Thus, if a pilot flies at
constant indicated airspeed, the true airspeed will actually increase with increasing altitude, meaning that the
flight is unsteady. However, this approach is quite clearly easy for the pilot to execute: he just has to keep the
indicated airspeed constant. If he wanted to keep the true airspeed constant, he would continuously need to
calculate what the true airspeed was, which he can’t obviously (not because he’s stupid but because he’s not a
computer).
This acceleration can be plotted in figure 2.5: we see that for increasing altitude, the 𝑉TAS initially increases.
However, there’s a certain Mach limit: a Mach number that may not be exceeded due to various reasons.
Although this Mach number is constant, as the temperature decreases with increasing altitude (and thus the
speed of sound), the corresponding 𝑉TAS decreases as well. At a certain moment, there’s a cross-over point after
which the Mach number becomes the deciding factor. Note that at the very top, the Mach number becomes a
vertical line; this is because we then enter the stratosphere where the temperature is constant.

Figure 2.5: Crossover speed.

samvanelsloo97@icloud.com
CHAPTER 2. LECTURE 1 18

Furthermore, the climb path is slightly curved, as shown in figure 2.6: the 𝛾 decreases with increasing altitude.
This is because if we remember 𝑅𝐶 = 𝑉 sin 𝛾, we have that 𝑅𝐶 decreases with increasing altitude, but 𝑉
increases, and thus 𝛾 is apparently decreasing. However, this change in 𝛾 over time is really really small, so we
can set it almost equal to zero, i.e. 𝑑𝛾∕𝑑𝑡 ≈ 0.

Figure 2.6: Climb path curvature.

Thus, to conclude, we have for the equations of motion, when we keep 𝑉EAS constant:
• Unsteady climb: 𝑑𝑉 ∕𝑑𝑡 ≠ 0.
• Quasi-rectilinear: 𝑑𝛾∕𝑑𝑡 ≈ 0.

2.2.1 Rate of climb

Let us now consider what the rate of climb will be. We have that the equations of motion equal
𝑊 𝑑𝑉
= 𝑇 − 𝐷 − 𝑊 sin 𝛾
𝑔 𝑑𝑡
𝑊 𝑑𝛾
𝑉 = 𝐿−𝑊
𝑔 𝑑𝑡
but as 𝑑𝛾∕𝑑𝑡 ≈ 0, the second equation reduces to 𝐿 = 𝑊 . Again, multiply the first row by 𝑉 , then divide by
𝑊 to get the power equation:

𝑊 𝑑𝑉
𝑉 = 𝑇 𝑉 − 𝐷𝑉 − 𝑊 𝑉 sin 𝛾 = 𝑃𝑎 − 𝑃𝑟 − 𝑊 𝑉 sin 𝛾
𝑔 𝑑𝑡
𝑃𝑎 − 𝑃𝑟 𝑉 𝑑𝑉
= + 𝑅𝐶
𝑊 𝑔 𝑑𝑡
So, we now see that the specific excess power is both converted into kinetic energy (𝑉 ∕𝑔 𝑑𝑉 ∕𝑑𝑡) and potential
energy (𝑅𝐶). Furthermore, note that we already derived that the specific excess power also equalled the steady
rate of climb, i.e. the rate of climb that would have been achieved, had the flight been steady (in other words, if
you would have used all your excess power merely to have a rate of climb and not accelerate the aircraft). In
fact, we can derive that (remember that 𝑑𝐻∕𝑑𝑡 is simply equal to the rate of climb):
𝑃𝑎 − 𝑃𝑟 𝑉 𝑑𝑉
= 𝑅𝐶𝑠 = + 𝑅𝐶
𝑊 𝑔 𝑑𝑡
𝑉 𝑑𝑉 𝑑𝐻 𝑉 𝑑𝑉
𝑅𝐶𝑠 = + 𝑅𝐶 = 𝑅𝐶 + 𝑅𝐶
𝑔 𝑑𝐻 𝑑𝑡 𝑔 𝑑𝐻
( )
𝑉 𝑑𝑉
𝑅𝐶𝑠 = 𝑅𝐶 +1
𝑔 𝑑𝐻
𝑅𝐶 1
= 𝑉 𝑑𝑉
𝑅𝐶𝑠 1+ 𝑔 𝑑𝐻

Note that this means that always 𝑅𝐶 ≤ 𝑅𝐶𝑠 : i.e., the actual rate of climb in unsteady flight is smaller than
the rate of climb that would have been achieved if the pilot would have kept the true airspeed constant, so that

©Sam van Elsloo


19 2.3. TOTAL TIME AND FUEL CONSUMPTION

𝑑𝑉 ∕𝑑𝐻 would have been constant. Furthermore, we have the relation


( √ ) (√ )
𝜌 𝜌0
𝑑 𝑉EAS 𝜌0 𝑑
𝑑𝑉 𝜌
= = 𝑉EAS
𝑑𝐻 𝑑𝐻 𝑑𝐻
as 𝑉EAS is constant. Furthermore, this derivative can be computed using ISA-relations.

FINDING THE 1. Establish a coordinate system.


RATE OF CLIMB 2. Draw the free body and kinetic diagram.
FOR AN 3. Set up the (approximated) equations of motion.
UNSTEADY 4. Multiply the first equation by 𝑉 to get
CLIMB
𝑊 𝑑𝑉
𝑃𝑎 − 𝑃𝑟 − 𝑊 𝑉 sin 𝛾 = 𝑉
𝑔 𝑑𝑡
𝑃𝑎 − 𝑃𝑟 𝑉 𝑑𝑉
= 𝑅𝐶 +
𝑊 𝑔 𝑑𝑡
5. Rewrite this as follows:
𝑉 𝑑𝑉 𝑑𝐻 𝑉 𝑑𝑉
𝑅𝐶𝑠 = 𝑅𝐶 + = 𝑅𝐶 + 𝑅𝐶
𝑔 𝑑𝐻 𝑑𝑡 𝑔 𝑑𝐻
( )
𝑉 𝑑𝑉
𝑅𝐶𝑠 = 𝑅𝐶 1 +
𝑔 𝑑𝐻
𝑅𝐶𝑠
𝑅𝐶 = 𝑉 𝑑𝑉
1+ 𝑔 𝑑𝐻

6. Find 𝑅𝐶𝑠 using the previous problem solving guide.


7. Find 𝑑𝑉 ∕𝑑𝐻 by applying ISA-relations to
( √ ) (√ )
𝜌 𝜌0
𝑑 𝑉EAS 𝜌0 𝑑
𝑑𝑉 𝜌
= = 𝑉EAS
𝑑𝐻 𝑑𝐻 𝑑𝐻

2.3 Total time and fuel consumption


Now that we’ve seen those nice things about descend and gliding, let’s go back to our method of climbing where
we keep the indicated airspeed constant. How can we determine the time and fuel it takes to reach a certain
altitude? This is an iterative process. One starts at a certain altitude, and then calculates the associated rate of
climb, using the method described before:
𝑅𝐶𝑠
𝑅𝐶 = 𝑉 𝑑𝑉
1+ 𝑔 𝑑𝐻
where
𝑃𝑎 − 𝑃𝑟
𝑅𝐶𝑠 =
𝑊
The term 𝑑𝑉 ∕𝑑𝐻 can be calculated using ISA-relations, as it could also be written as
( √ ) (√ )
𝜌 𝜌0
𝑑 𝑉IAS 𝜌0 𝑑
𝑑𝑉 𝜌
= = 𝑉IAS
𝑑𝐻 𝑑𝐻 𝑑𝐻
Then, one computes the corresponding fuel flow, based on the thrust and specific thrust (we have 𝑚̇ = 𝑐𝑝 ⋅ 𝑃𝑏𝑟 or
𝑚̇ = 𝑐𝑗 ⋅ 𝑇 ). One then computes the ground speed (true airspeed) based on the indicated airspeed. Based on
the rate of climb, you then compute how long it takes to reach a certain following altitude (e.g. 500 m higher),
compute how much fuel has been burned and compute the new aircraft weight. After that, the rate of climb, fuel
flow and ground speed are computed again, and one computes the time it takes to reach the next altitude (e.g. if
the previous two steps were 0 m and 500 m, then the third step would become 1000 m, etc.).

samvanelsloo97@icloud.com
CHAPTER 2. LECTURE 1 20

2.4 Minimum time to climb

2.4.1 Introduction of energy height

So, now the longest thing we’ll be discussing this chapter, namely what is actually the most optimal way to
climb? How close does the simple way of keeping the indicated airspeed constant come to this?
For this, let us reconsider figure 2.3, as shown in figure 2.7a: we want to go from start to finish. You may be
inclined to think, the fastest way to do this is as shown in figure 2.7b: you accelerate a bit at constant altitude,
then at each altitude, you climb at the maximum rate of climb possible, until you reach the desired altitude,
where you accelerate or decelerate some more to end up at the desired velocity. This is also approximately
what we’re doing when keeping the indicated airspeed constant: after all, as visible from graph 2.7b, we’re
continuously increasing true airspeed during ascent (which also happens when we keep the indicated airspeed
constant).

(a) From departure to destination. (b) Possible solution.

Figure 2.7: Graphs.

However, this is not the best way to climb. To find the best way to climb, we must introduce the concept of
energy height: note that the total energy of an aircraft is given by
1
𝐸𝑡𝑜𝑡 = 𝑚𝑔𝐻 + 𝑚𝑉 2
2
Dividing by 𝑚𝑔 gives
𝑉2
𝐸𝐻 = 𝐻 + = 𝐻𝑒
2𝑔
where 𝐻𝑒 denotes the energy height. Note that the unit of this is simply meters and not joule or something. Why
is this an interesting property? Note that it is very easy to interchange between potential and kinetic energy:
if you want to increase the velocity a lot in a very short amount of time, the best way to do this is by simply
pointing the nose of the aircraft down, descending a lot very quickly and using this loss in potential energy to
accelerate the aircraft. This goes much faster than waiting for your propulsion system to add sufficient kinetic
energy to your aircraft whilst flying at constant altitude. Indeed, if we plot lines of constant energy height, as
done in figure 2.8, we see that all of the process II, III and IV take a shorter amount of time than I; even IV,
which seems to be a lot longer, takes a shorter while (note that for IV, the aircraft has a very high velocity at
zero altitude; the pilot pitches up the nose of the aircraft significantly (almost vertically), and this kinetic energy
is converted into potential energy. In conclusion: speed and altitude can be exchanged rapidly, but increasing
the total energy level is time consuming.
Now, let’s start calculating stuff. First of all, the time to climb from altitude 𝐻1 to 𝐻2 equals
𝐻2
𝑑𝐻
𝑡=
∫ 𝑅𝐶
𝐻1

©Sam van Elsloo


21 2.4. MINIMUM TIME TO CLIMB

Figure 2.8: Lines of constant energy height.

It should be pretty obvious why: the higher the 𝑅𝐶, the lower the time it’ll take to climb (it’s inversely related
to each other). Now, suppose one climbs at the steady rate of climb: then the kinetic energy does not change;
only the potential energy changes. So, an increase in altitude 𝑑𝐻 is equal to the the change in energy height
𝑑𝐻𝑒 : essentially, our integral now becomes

𝐻𝑒
2
𝑑𝐻𝑒
𝑡=
∫ 𝑅𝐶𝑠
𝐻𝑒
1

An alternative way of deriving this is as follows: we start with

𝐻2
𝑑𝐻
𝑡 =
∫ 𝑅𝐶
𝐻1
𝑅𝐶𝑠 𝑅𝐶𝑠
𝑅𝐶 = 𝑉 𝑑𝑉
=
1+ 1 𝑑𝑉 2
𝑔 𝑑𝐻 1+ 2𝑔 𝑑𝐻

Substituting gives:

𝐻2 𝐻2 ( ) 𝐻2 ( )
𝑑𝐻 𝑑𝐻 1 𝑑𝑉 2 1 1 2
𝑡= = ⋅ 1+ = 𝑑𝐻 + 𝑑𝑉
∫ 𝑅𝐶𝑠 ∫ 𝑅𝐶𝑠 2𝑔 𝑑𝐻 ∫ 𝑅𝐶𝑠 2𝑔
𝐻1 1 𝑑𝑉 2 𝐻1 𝐻1
1+ 2𝑔 𝑑𝐻

but we already defined before that

𝑉2
𝐻𝑒 = 𝐻 +
2𝑔

so that

𝑑𝑉 2
𝑑𝐻𝑒 = 𝑑𝐻 +
2𝑔

and thus
𝐻𝑒
𝐻2 ( ) 2
1 1 𝑑𝐻𝑒
𝑡= 𝑑𝐻 + 𝑑𝑉 2 =
∫ 𝑅𝐶𝑠 2𝑔 ∫ 𝑅𝐶𝑠
𝐻1 𝐻𝑒
1

samvanelsloo97@icloud.com
CHAPTER 2. LECTURE 1 22

2.4.2 Use of energy height

Why is this interesting? This means that to minimize the flight time, we have to maximize 𝑅𝐶𝑠 for a constant
𝐻𝑒 : that is, we have to fly through points in figure 2.7a that have maximum 𝑅𝐶𝑠 for a constant energy height,
not for a constant altitude. More graphically, this means that we have to follow the following path in the graph:
we set the 𝑥-axis to 𝑉 2 ∕2∕𝑔0 , and we draw lines of constant energy height (as we now use 𝑉 2 ∕2∕𝑔0 , these
are now straight lines), as shown in figure 2.9a: you then have to travel through the points for which the lines
of constant 𝑅𝐶𝑠 (those parabola thingies) are tangent to lines of constant energy height. This is more clearly
depicted in figure 2.9b. Note that before and after the "real" climb (I mean, the parts where you travel from
your initial velocity to the starting point of the steep climb, and from the finish of the steep climb to the final
destination), you descend along lines of constant energy height, so that you get there really quickly.

(a) Optimal way to climb. (b) Optimal way to climb.

Figure 2.9: Optimal way to climb.

So, what does this tell us?


• The optimum way of climbing is to first descent along a line of constant energy height, then climb through
points where the 𝑅𝐶𝑠 curve is tangent to a line of constant energy height, then descent along a line of
constant energy height again (you may also want to ascent along a line of constant energy height, if the
desired velocity is lower than the velocity at the end of the climb).
• The easy way of having constant indicated airspeed comes pretty close to this optimum way of climbing,
nevertheless it’s still suboptimal.
Now, you may be amazed by now, how in the name of the lord is that possible? How can anything be faster than
continuously climbing at maximum 𝑅𝐶𝑠 for that altitude? How can it be faster to climb at a lower 𝑅𝐶𝑠 ? Well,
what you have to realize: if you’re continuously accelerating (which is happening), then you’re not experiencing
steady climb, so your rate of climb does not equal the steady rate of climb. Rather, this steady rate of climb is
merely an indication of what your real rate of climb will be, as we know that
𝑅𝐶𝑠
𝑅𝐶 = 𝑉 𝑑𝑉
1+ 𝑔 𝑑𝑡

It should be obvious that this is not a simply linear relation, but one that also depends on 𝑉TAS and 𝑑𝑉 ∕𝑑𝑡,
which implies that there is indeed maybe a faster way of climbing.

2.4.3 Supersonic climb

In supersonic climb, we pretty much have the same idea: in the diagram, we want fly through points where
𝑅𝐶𝑠 is tangent to lines of constant energy height. However, for Mach numbers close to 1, the figure is entirely
messed up as we have to deal with the Mach divergence number. The result is shown in figure 2.10.
The ideal path is now as follows: you descent along a line of constant energy height, then climb passing through
points that are tangent to lines of constant energy height. Then, after a certain while, you dive through the sound
barrier, to quickly increase your velocity so that you don’t have to deal with the sound barrier any more. We

©Sam van Elsloo


23 2.5. DESCENT

Figure 2.10: Supersonic climb.

then again climb passing through points tangent to lines of constant energy height, and then finally ascend along
a line of constant energy height. Now, how do we determine when we need to dive through the sound barrier?
We do that by looking a bit where it’s best to dive: in the dive path shown in the figure, we dive from a climb
rate of about 40 to a climb rate of about 45: if we’d continue climbing in the subsonic part of the graph, we’d
climb at a lower rate than if we’d be climbing in the supersonic part, and thus we make the dive. Similarly, if we
dived earlier than this, we would have to pay the punishment by climbing in the supersonic zone at a lower rate
of climb.

2.5 Descent
Descent is done in exactly the same way as ascent, but just in the reverse direction (you use a deficit in power
available vs. power required). First you have constant Mach number, then once you’ve hit the cross over speed
you take constant IAS. The only difference is the approach phase: at a certain point, you reach a glideslope
where your flight angle is 3°; the constant velocity is determined by air traffic control.
So, suppose we have 𝛾 = 3° and 𝑉 , 𝑊 , 𝐻 and 𝜌 are known. Then we can compute the required thrust as follows.
We have (steady flight):
𝑇 − 𝐷 − 𝑊 sin 𝛾 = 0
𝐿 = 𝑊
We can compute 𝐶𝐷 from
𝐶𝐿2
𝐶𝐷 = 𝐶𝐷0 +
𝜋𝐴𝑒
where 𝐶𝐿 is known from
2𝑊
𝐶𝐿 =
𝜌𝑉 2 𝑆
then 𝐷 follows from
1
𝐷 = 𝐶𝐷 𝜌𝑉 2 𝑆
2
and then we can just plug in
𝑇 = 𝐷 + 𝑊 sin 𝛾

2.6 Gliding flight


Gliders always go down relative to the air. However, if the air has a vertical component, this means that gliders
can go up as well. This happens in thermals, as shown in figure 2.11.

samvanelsloo97@icloud.com
CHAPTER 2. LECTURE 1 24

Figure 2.11: Thermals.

As a glider aircraft has no thrust, the equations of motion becomes

𝑊 𝑑𝑉
−𝐷 − 𝑊 sin 𝛾 =
𝑔 𝑑𝑡
𝐿 = 𝑊

but we know already that the minimum glide angle occurs when 𝑑𝑉 ∕𝑑𝑡 = 0; furthermore, we use 𝛾̄ to indicate
the conjugate of 𝛾 (i.e. 𝛾̄ = −𝛾) (we know that 𝛾 will be negative, and we all prefer positive numbers), so that
we get

−𝐷 + 𝐿 sin 𝛾̄ = 0
𝐷 𝐷
𝛾̄ = arcsin = arcsin
𝐿 𝐿
( )
So, to minimize 𝛾̄ , 𝐶𝐿 ∕𝐶𝐷 must be maximized. We can find the associated 𝐶𝐿𝑜𝑝𝑡 as follows:
( )
𝐶𝐿 𝑑𝐶𝐷
𝑑 𝐶𝐷
𝐶𝐷 − 𝐶
𝑑𝐶𝐿 𝐿
= =0
𝐶𝐿 2
𝐶𝐷

with 𝐶𝐷 = 𝐶𝐷0 + 𝐶𝐿2 ∕ (𝜋𝐴𝑒), this means that

𝑑𝐶𝐷 2𝐶𝐿
=
𝑑𝐶𝐿 𝜋𝐴𝑒

and thus
2𝐶𝐿
𝐶𝐷0 + 𝐶𝐿2 ∕ (𝜋𝐴𝑒) − ⋅ 𝐶𝐿 = 0
𝜋𝐴𝑒
𝐶2
𝐶𝐷0 − 𝐿 = 0
𝜋𝐴𝑒

𝐶𝐿opt = 𝐶𝐷0 𝜋𝐴𝑒

Please note: the minimum glide angle has nothing to do with weight of the aircraft (as 𝛾̃ is only a function of
𝐷∕𝐿); however, a heavier aircraft will have a higher flight path velocity.

2.7 Conclusion
What are things you need to be able to do or explain from this chapter?
• You need to be able to explain the operational practice of climb and descent (so how do they it, with
constant indicated airspeed and all).

©Sam van Elsloo


25 2.8. RELEVANT OLD EXAM QUESTIONS

• You need to be able to calculate the aircraft performance: i.e., you need to be able to calculate what the
rate of climb of an aircraft climbing at constant IAS is.
• You need to be able to determine (graphically) the optimum flight profiles.
• You need to be able to explain the concept of energy (so make sure you understand the concept).
• You need to be able to evaluate glider descent performance, i.e. minimum glide angle, glide distance,
time it takes to reach the ground, etc.

2.8 Relevant old exam questions

2.8.1 Full exercises

• Exam January 2013, question 5.


• Exam February 2012, question 4.
• Exam October 2010, question 2.

2.8.2 Parts of exercises

• Exam April 2015, question 4b-4d.


• Exam January 2011, question 1a.

2.8.3 Exercises combining previously discussed material

None.

samvanelsloo97@icloud.com
CHAPTER 2. LECTURE 1 26

©Sam van Elsloo


3 Turning flight

Of course, we want to be able to turn our aircraft when it’s in the air as well. That’s exactly what we’re gonna
discuss here, in case you couldn’t deduce that from how this chapter is called.

3.1 Equations of motion and load factor

First, let’s take a closer look at the equations of motion in turning flight. The free body diagrams are shown in
figure 3.1. Steady means that 𝑑𝑉 ∕𝑑𝑡 = 0, coordinated means that there is no sideslip, horizontal means that
you maintain constant altitude.

Figure 3.1: Free body diagrams for steady, coordinated, horizontal turn. On the left, we take the aircraft from
behind, on the right, we see it from above. Note that 𝐹𝑐 is a fictious force.

The equations of motion then become:

𝑊 𝑑𝑉
= 𝑇 −𝐷
𝑔 𝑑𝑡
𝑊 𝑉2
= 𝐿 sin Φ
𝑔 𝑅
𝑊 𝑑𝛾
𝑉 = 𝐿 cos Φ − 𝑊
𝑔 𝑑𝑡

Applying the flight conditions, 𝑑𝑉 ∕𝑑𝑡 = 0, 𝑑𝛾∕𝑑𝑡 = 0, we end up at

E QUATIONS OF 0 = 𝑇 −𝐷 (3.1)
MOTION FOR
𝑊 𝑉2
STEADY, = 𝐿 sin Φ (3.2)
𝑔 𝑅
COORDINATED ,
0 = 𝐿 cos Φ − 𝑊 (3.3)
HORIZONTAL
TURN
Furthermore, we have the definition

L OAD FACTOR The load factor equals


𝐿 𝐿 1
𝑛≡ = = (3.4)
𝑊 𝐿 cos Φ cos Φ

27
CHAPTER 3. TURNING FLIGHT 28

3.2 Standard turns


Air traffic control wants aircraft to perform turns with a fixed turn rate and a fixed flight velocity.

STANDARD A rate 1 turn corresponds to a turn that takes 2 minutes to do a full revolution; i.e. a turn rate of 3 °∕s.
TURNS A rate 2 turn is then a turn that occurs at twice the speed, i.e. turn rate of 6 °∕s. A rate 4 turn is a turn
that occurs at four times the speed, i.e. turn rate of 24 °∕s. A rate 1/2 turn is a turn that occurs at half
the speed, i.e. turn rate of 1.5 °∕s. It should be pretty clear by now.

So, suppose we have a rate one turn to perform at an airspeed of 50 m∕s, what will be the turn radius, what
should be our bank angle and what will be the load factor? For the radius, we have

𝑉 = Ω𝑅

where Ω is the turn rate, i.e. 3 °∕s, leading to


𝑉 50
𝑅= = 𝜋 = 955 m
Ω 3 ⋅ 180

For the bank angle, we simply use the latter two of the equations of motion:

𝑊 𝑉2
= 𝐿 sin Φ
𝑔 𝑅
0 = 𝐿 cos Φ − 𝑊

From the second one, we have 𝑊 = 𝐿 cos Φ and thus

𝐿 cos Φ 𝑉 2
= 𝐿 sin Φ
𝑔 𝑅
𝑉2
= tan Φ
𝑔𝑅
𝑉2 502
Φ = arctan = arctan = 15°
𝑔𝑅 9.81 ⋅ 955
The load factor is then simply
1 1
𝑛= = = 1.035
cos Φ cos 15°

3.2.1 Turn coordinator

Airplanes are equipped with turn coordinators, as shown in figure 3.2. This indicates two things: first of all,
the bank angle is shown by the white aircraft: if it’s wing tip touches one of the lower bars, the bank angle is
sufficient for a rate 1 turn. Additionally, it shows whether the turn is coordinated or not.

Figure 3.2: A turn coordinator.

If a turn is not coordinated, it also slips or skids: what this means is shown in figures 3.3a and 3.3b. The effects
occur due to wrong use of the rudder: remember that if we bank, we only cause a centripetal force, meaning our
trajectory becomes circular. However, we also like our aircraft to rotate at the appropriate rate (so that if we’ve

©Sam van Elsloo


29 3.3. MAXIMUM AIRCRAFT TURNING PERFORMANCE

turned 90°, our aircraft has also turned 90°. You need to use your rudder for this. If you apply too much rudder,
you generate a too large moment around the vertical axis of the aircraft, meaning that the nose turns too much in
(a bit like oversteer on a regular car). If you apply too little rudder, not enough of a moment around the vertical
axis of the aircraft is created, meaning that the aircraft tries to fly on straight (even though your trajectory is
circular) (similar to understeer on a regular car).

(a) Skidding turn. (b) Slipping turn.

Figure 3.3: Skidding and slipping.

Now, when you’re skidding, which is basically oversteer, you’re basically turning into much, meaning you
experience a larger centrifugal force, causing the lower ball in the turn coordinator to move to the outside, as
shown in figure 3.3a as well. Note that this means you are experiencing outwards lateral accelerations.
Similarly, for slipping (understeer), you’re not turning in enough, meaning you’re experiencing only a very small
centrifugal force, causing the lower ball to move all the way to the inside, as shown in figure 3.3b. Note that this
means that you are experiencing inwards lateral accelerations (honestly, if you just compare it with how you
experience understeer and oversteer in a car, it’s so much clearer).
All of this is drawn one final time in figure 3.4.

Figure 3.4: Slip indicator.

3.3 Maximum aircraft turning performance

Now, an aircraft is probably able to be do more than standard turns. In fact, we can distinguish three performance
characteristics with regards to turning performance:
• Steepest turn, i.e. what is the maximum bank angle the aircraft can perform a turn at and what is the
associated load factor.
• Minimum turn radius.
• Minimum turn time.
We’ll discuss them one-by-one in this section.

samvanelsloo97@icloud.com
CHAPTER 3. TURNING FLIGHT 30

3.3.1 Steepest turn

What is the maximum load factor we can achieve (and thus maximum bank angle)? First, consider the regular
performance diagram of a jet aircraft as shown in figure 3.5. What happens to it if we increase the load
factor?

Figure 3.5: Performance diagram.

Remember that 𝐿 = 𝑛𝑊 , and thus

𝐷 𝐶𝐷 21 𝜌𝑉 2 𝑆 𝐶
𝐷= 𝐿= 𝑛𝑊 = 𝑛 𝐷 𝑊
𝐿 1 𝐶𝐿
𝐶𝐿 2 𝜌𝑉 2 𝑆
So, if we double the load factor, the drag will become twice as high. Furthermore, the velocity will also change:
we have
1
𝐿 = 𝑛𝑊 = 𝐶𝐿 𝜌𝑉 2 𝑆
2

𝑊 2 𝑛
𝑉 =
𝑆 𝑆𝑆

So, if we double the load factor, the required airspeed becomes 2 as high. This causes the shift of the graph as
shown in figure 3.6.

Figure 3.6: Performance diagram for higher load factor.

You can do this for multiple load factors, leading to the graph shown in figure 3.7a. We can then deduce what
the maximum load factor is for each specific flight speed: initially, this is mostly limited by the stall speed; when
you’re very close to stall speed, you can’t have much roll as this wastes part of your lift, meaning you have to
increase the lift coefficient, but you can’t do this when you’re already close to stall speed of course. After that,
the thrust is the limiting factor, and increasing the roll angle beyond what’s allowed results in too much drag to
survive. Taking the intersections for each flight speed results in the graph shown in figure 3.7b.
From figure 3.7b, one can then straightforwardly deduce 𝑛max . Note that due to the dependence on the thrust, it
is nearly impossible to come up with a mathematical solution for this problem, so most of the time, you’ll need
to do it graphically as shown above. However, if you assume constant thrust, it is pretty easy to compute the
maximum load factor: you calculate the 𝐶𝐷 ∕𝐶𝐿 that minimizes drag (which occurs at minimum 𝐶𝐷 ∕𝐶𝐿 ), and
then solve
𝐶𝐷
𝑇 =𝐷=𝑛 𝑊
𝐶𝐿

©Sam van Elsloo


31 3.3. MAXIMUM AIRCRAFT TURNING PERFORMANCE

(a) Load factor lines in performance diagram. (b) Maximum load factor at different air speeds.

Figure 3.7: Maximum load factor at different air speeds.

for 𝑛, as 𝑇 , 𝐶𝐷 ∕𝐶𝐿 and 𝑊 are known.

3.3.2 Minimum turn radius

Minimum turn radius is pretty easy. We have

𝑊 𝑉2
= 𝐿 sin Φ
𝑔 𝑅
With 𝑊 = 𝐿 cos Φ, this can be written as

𝐿 cos Φ 𝑉 2
= 𝐿 sin Φ
𝑔 𝑅
𝑉2
= tan Φ
𝑔𝑅
𝑉2
𝑅 =
𝑔 tan Φ

Look again at figure 3.1; we have that 𝐿 = 𝑛𝑊 and 𝐿 cos Φ = 𝑊 , so that 𝐿 sin 𝜙 equals (Pythagoras)
√ √
𝐿 sin 𝜙 = (𝑛𝑊 )2 − 𝑊 2 = 𝑊 𝑛2 − 1

so that

𝐿 sin 𝜙 𝑊 𝑛2 − 1 √ 2
tan 𝜙 = = = 𝑛 −1
𝐿 cos 𝜙 𝑊
This means that the radius of a turn at airspeed 𝑉 and load factor 𝑛 equals

T URN RADIUS 𝑉2
𝑅= √ (3.5)
FOR GIVEN
𝑔 𝑛2 − 1
AIRSPEED AND
LOAD FACTOR
We can use the same 𝑛max − 𝑉 relation we found in the previous subsection; for each point on this graph, we
take 𝑛max and 𝑉 from figure 3.8a and calculate the associated 𝑅, leading to the graph shown in figure 3.8b.
Again, the minimum radius must be determined graphically instead of analytically.

3.3.3 Minimum time to turn

The time to turn 360° logically equals

2𝜋𝑅 2𝜋 𝑉2
𝑇 = = √
𝑉 𝑉 𝑔 𝑛2 − 1

samvanelsloo97@icloud.com
CHAPTER 3. TURNING FLIGHT 32

(a) Maximum load factor at different air speeds. (b) Turn radius at different air speeds.

Figure 3.8: Turn radius at different air speeds.

or

FULL TURN 𝑇2𝜋 =


2𝜋𝑅
= √
2𝜋𝑉
(3.6)
TIME 𝑉 𝑔 𝑛2 − 1

Again, one must use the graphical approach. One can either construct the 𝑛max − 𝑉 diagram, then the 𝑅 − 𝑉
diagram and then the 𝑇 − 𝑉 diagram using the first part of the equation above, or one can directly transform the
𝑛max − 𝑉 diagram into a 𝑇 − 𝑉 diagram. The results are shown in figures 3.9a and 3.9b.

(a) Turn radius at different air speeds. (b) Turn time at different air speeds.

Figure 3.9: Turn time at different air speeds.

3.4 Altitude effect


Of course, the altitude has some effect. The performance diagram will change as shown in figure 3.10a: the thrust
will shift downwards with increasing altitude, and the drag curve will move to the right. The reasons for this
are exactly explained in chapter 1 of this summary, so read that again if you don’t remember exactly why. This
results in a new 𝑛max − 𝑉 diagram, shifted down and to the right, as shown in figure 3.10b. This affects the 𝑅 − 𝑉
diagram, as shown in figure 3.11a: the graph shifts to the upperright. This increases the minimum turn radius,
and increases the velocity at which it occurs. Finally, the variation of the three performance characteristics are
shown in figure 3.11b and should speak for themselves. Note that the smallest turn radius occurs at a lower flight
velocity than the smallest turn time, which again occurs at a lower flight velocity than the steepest turn.

3.5 Relevant old exam questions

3.5.1 Full exercises

• Exam April 2016, question 1. Question 1d requires some knowledge of climb performance, but not much.
• Exam April 2014, question 1.
• Exam January 2013, question 3.
• Exam November 2011, question 2.

©Sam van Elsloo


33 3.5. RELEVANT OLD EXAM QUESTIONS

(a) Altitude effect on performance diagram. (b) Altitude effect on 𝑛max − 𝑉 diagram.

Figure 3.10: Altitude effect on performance diagram and 𝑛max − 𝑉 diagram.

(a) Altitude effects on 𝑅 − 𝑉 diagram. (b) Altitude effects on performance characteristics.

Figure 3.11: Altitude effects on 𝑅 − 𝑉 diagram and performance characteristics.

• Exam November 2011, question 3.


• Exam January 2011, question 1. Question 1a requires some knowledge of climb performance, but not
much.
• Exam October 2010, question 1.

3.5.2 Parts of exercises

None.

3.5.3 Exercises combining previously discussed material

• April 2015, question 4e-4g.

samvanelsloo97@icloud.com
CHAPTER 3. TURNING FLIGHT 34

©Sam van Elsloo


4 Airfield performance

4.1 Balanced field length

Suppose you’re the pilot of an aircraft, and you’ve started applying full throttle, and you’re moving over the
runway at a decent speed. Suddenly, one of your engine dies. What do you do? Obviously, that depends on how
fast you’re going and how close you are to the end of the runway. If you’re only going 1 km∕h and you still have
2 km of runway left, obviously it’s safe to brake and just stop and return to the gate. However, if you’re already
going 50 m∕s and you’re very close to the end of the runway, you’ll probably be unable to brake in time so you
should better just continue and hope you’re able to fly.

It’s clear that there is some cut-off velocity, which is denoted by 𝑉1 . Before you reach this velocity, you should
abort the take-off; after this velocity, you should continue. We can compute this velocity using the following
method: for various values of the velocity at which one engine breaks down, we compute:

• The accelerate-climb distance: the distance it takes to accelerate to the velocity when all engines are
working, and then the distance it takes to take-off when one engine breaks down once this velocity is
reached.
• The accelerate-stop distance: the distance it takes to accelerate to the velocity when all engines are
working, and then the distance it takes to stop the aircraft again once the velocity is reached.

The result of this is shown in figure 4.1. We can see the logical of the shapes of the graphs (suppose we have a
2-engine aircraft): if one engine breaks down at a very low velocity, say 𝑉 = 1 m∕s, then the accelerate-climb
distance will be very long, as you’re only able to use 2 engines to accelerate to 1 m∕s; after that, you can only
use 1 engine, which means it takes longer to accelerate sufficiently to 𝑉LOF (lift-off speed). On the other hand,
the accelerate-stop distance will be very short as it does not take long to reach 1 m∕s and it does not take long to
stop from that velocity. At higher engine failure velocities, the accelerate-climb distance will decrease, as you’re
able to use maximum thrust for a larger part of the acceleration. The accelerate-stop distance will decrease, as it
takes longer to reach the 𝑉𝐸𝐹 (engine failure) velocity and longer to decelerate again.

Figure 4.1: Balanced field length.

At one point, the two lines intersect; the associated velocity is 𝑉1 and the associated field length is the balanced
field length: if the length of the runway is equal to the balanced field length and an engine failure occurs when
the ground speed is equal to 𝑉1 , then this is the cross-over point where you have to start continuing your take-off
instead of aborting (so if your ground speed is larger than 𝑉1 , you have to continue on, if it’s smaller than 𝑉1 ,
you have to abort).

Note that if the actual field length is larger than the balanced field length, this simply allows for some margin on
your decision; for example, if one engine then fails shortly after 𝑉1 , it may still be possible to stop in time before
the runway ends.

35
CHAPTER 4. AIRFIELD PERFORMANCE 36

4.2 Take-off performance


Take-off consists of two main phases, as shown in figure 4.2 (𝑉2 is the minimum speed to ensure an adequate
and safe climbout with the critical engine inoperative and the remaining engines at full take-off thrust): we
distinguish a ground run and an airborne phase (which starts as soon as none of the wheels are touching the
ground anymore).

Figure 4.2: Take-off phases.

4.2.1 Ground run

Remember that we have


𝑑𝑠
𝑉 =
𝑑𝑡
𝑑𝑉
𝑎 =
𝑑𝑡
From the first one, we then have
𝑑𝑠
𝑑𝑡 =
𝑉
so that we can plug that into the second on so that we get
𝑑𝑉
𝑎 = 𝑑𝑠
𝑉
𝑎𝑑𝑠 = 𝑉 𝑑𝑉

which is exactly what we remember from dynamics. Integration gives


𝑉LOF 𝑠

𝑉 𝑑𝑉 = 𝑎𝑑𝑠
∫ ∫
0 0

The integral on the left is easy to calculate, but the integral on the right is much harder if we don;t know the
function for 𝑎; instead, we often assume a constant average acceleration, 𝑎. ̄ In that case, we simply get
2
𝑉LOF
= 𝑎𝑠
̄
2
so that

G ROUND RUN 2
𝑉LOF
DISTANCE FOR 𝑠= (4.1)
2𝑎̄
TAKE -OFF

where the average acceleration is simply (should be fairly obvious, just draw a FBD, then set 𝐹 = 𝑚𝑎 = 𝑊 ∕𝑔 ⋅ 𝑎
and solve for 𝑎)
̄

AVERAGE AC- 𝑔 (̄ )
𝑎̄ = 𝑇 − 𝐷̄ − 𝐷̄ 𝑔 (4.2)
CELERATION 𝑊
DURING
GROUND RUN
©Sam van Elsloo
37 4.2. TAKE-OFF PERFORMANCE

Now, you may wonder, how do we determine 𝑎?


̄ Well, rather simply, we just calculate 𝑇̄ , 𝐷̄ and 𝐷̄ 𝑔 at the
average velocity, which is taken to equal
𝑉LOF
𝑉 = √
2
where
𝑉LOF = 1.05𝑉min
where 𝑉min is the aircraft velocity at maximum lift coefficient. Should be pretty straightforward.

4.2.2 Airborne distance

The airborne phase consists of two phases, as shown in figure 4.3a. First we have the transition, where we follow
a circular path (we just assume it’s circular) to get to the correct flight angle; this is followed by a pure climb at
constant climb angle, until the required 50 f t screen height ℎ𝑠 from the ground has been reached. The free body
diagram during the transition phase is shown in figure 4.3b.

(b) Free body diagram.

(a) Transition and screen height.

Figure 4.3: Transition and climb until screen height.

Now, let’s first analyse the transition phase. Remember that the equation of motion equals

𝑊 𝑑𝛾 𝑊 𝑉2
𝑉 = = 𝐿 − 𝑊 cos 𝛾
𝑔 𝑑𝑡 𝑔 𝑅

Again, approximate cos 𝛾 ≈ 1, and divide both sides of the equation by 𝑊 to get

𝑉2 𝐿
= −1=𝑛−1
𝑅𝑔 𝑊

so that we get

𝑉2
𝑅=
𝑔 (𝑛 − 1)

Now, from empirical data, we usually know that pilots go for a load factor 𝑛 = 1.15 during the transition phase,
leading to
𝑉2
𝑅=
0.15𝑔
Then, per simple geometry from figure 4.3a, we have

samvanelsloo97@icloud.com
CHAPTER 4. AIRFIELD PERFORMANCE 38

T RANSITION The horizontal distance covered during the transition phase equals
HORIZONTAL
2
𝑉LOF
DISTANCE
𝑥trans = 𝑅 sin 𝛾climb = sin 𝛾climb (4.3)
0.15𝑔

Then, onto the horizontal distance covered during the climbout. Again, we use simple geometry from figure
4.3a to get
ℎ − ℎ𝑡
tan 𝛾climb = 𝑠
𝑥climb
where ℎ𝑠 = ℎscr , the screen height, and ℎ𝑡 the height achieved during transition, equal to (again, simple geometry
from figure 4.3a):
( ) 𝑉2 ( )
ℎtrans = 𝑅 ⋅ 1 − cos 𝛾climb = LOF 1 − cos 𝛾climb
0.15𝑔
and thus we can write

C LIMB The horizontal distance covered during the climb phase equals
HORIZONTAL
𝑉2
DISTANCE LOF
ℎscr − (1 − cos 𝛾) 0.15𝑔
𝑥climb = (4.4)
tan 𝛾

The total airborne distance is then logically

TOTAL 𝑥total-airborne = 𝑥trans − 𝑥climb (4.5)


AIRBORNE
DISTANCE
The total distance for take-off, to get to the screen height, simply equals

TOTAL 𝑥total = 𝑥ground + 𝑥airborne (4.6)


TAKE -OFF
DISTANCE
Note: it may be the case that the screen height is already reached during the transition phase; in this case, it is
pretty logical what happens to the airborne distance. First, we have
2
𝑉LOF
ℎ𝑡 = 𝑅 (1 − cos 𝛾) = ⋅ (1 − cos 𝛾) = ℎ𝑠 = 50 f t
0.15𝑔
ℎ𝑠 , 𝑉LOF , 𝑔 are known, and thus 𝛾 is straightforward to calculate. The airborne distance is then simply
2
𝑉LOF
𝑥airborne = 𝑥transition = ⋅ sin 𝛾
0.15𝑔

4.3 Landing
For landing, everything is really similar, it’s just in reverse order.

4.3.1 Airborne phase

Looking at figure 4.4, we see great similarly to what was the case for take-off. However, now 𝛾𝑎𝑝 (ap from
approach) is fixed, namely 3°, and 𝑉ap is fixed to be 1.3𝑉min , where 𝑉min equals

𝑊 2 1
𝑉min =
𝑆 𝜌 𝐶𝐿max

©Sam van Elsloo


39 4.3. LANDING

so that √
𝑊 2 1
𝑉ap = 1.3
𝑆 𝜌 𝐶𝐿max

Figure 4.4: Landing airborne phase.

We once again get (just compare with the formulas for take-off; climb becomes descent and transition becomes
flare)
( )
TOTAL ℎscr − 1 − cos 𝛾𝑎𝑝 𝑅
LANDING 𝑥total-airborne = 𝑥flare + 𝑥descent = 𝑅 sin 𝛾𝑎𝑝 + (4.7)
tan 𝛾𝑎𝑝
DISTANCE
where
𝑊 2 1
𝑆 𝜌 𝐶𝐿max
𝑅 = 1.3 2
(4.8)
Δ𝑛 ⋅ 𝑔

From empirical data, Δ𝑛 = 0.10𝑔 for subsonic jets, and 0.15𝑔 for propeller aircraft.

4.3.2 Ground run distance

For the ground run, it is again very similar to as it was for the take-off. However, once we touchdown, there’s a
short period of time we’re only trying to getting the nose of the airplane down, without actually braking, the
so called "rotating" phase (as the plane is literally rotating). What is the distance covered during this? Well
actually, it’s pretty simple to compute: we assume it takes two seconds to transition, and the approach speed is
𝑉𝑎𝑝 = 1.3𝑉𝑚𝑖𝑛 , thus

H ORIZONTAL 𝑥𝑡𝑟 = 𝑡𝑡𝑟 ⋅ 𝑉𝑎𝑝 = 2 ⋅ 1.3𝑉𝑚𝑖𝑛 = 2.6𝑉𝑚𝑖𝑛 (4.9)


DISTANCE
COVERED
DURING For the braking phase, we can apply the exact same derivation as before: we have
TRANSITION 𝑑𝑠
𝑉 =
𝑑𝑡
𝑑𝑉
𝑎 =
𝑑𝑡
From the first one, we then have
𝑑𝑠
𝑑𝑡 =
𝑉
so that we can plug that into the second on so that we get
𝑑𝑉
𝑎 = 𝑑𝑠
𝑉
𝑎𝑑𝑠 = 𝑉 𝑑𝑉

samvanelsloo97@icloud.com
CHAPTER 4. AIRFIELD PERFORMANCE 40

which is exactly what we remember from dynamics. Integration gives


0 𝑠

𝑉 𝑑𝑉 = 𝑎𝑑𝑠
∫ ∫
𝑉𝑎𝑝 0

Again, taking a constant average acceleration, this becomes


2
𝑉𝑎𝑝

= 𝑎𝑠
̄
2
Now, you can again easily draw the free body diagram and find that
𝑊 (̄ ( ))
𝑎̄ = − 𝑇rev + 𝐷̄ + 𝜇 𝑊 − 𝐿̄
𝑔
where 𝑇̄rev is the reverse thrust (if we have a reverse thruster, otherwise this is just 0 as we switch off the engines
during landing). Furthermore, we already determined that

𝑊 2 1
𝑉𝑎𝑝 = 1.3𝑉𝑚𝑖𝑛 = 1.3
𝑆 𝜌 𝐶𝐿𝑚𝑎𝑥

so that

G ROUND 𝑊 2 2 1.32 1
𝑥brake = ( ) (4.10)
BRAKE 2𝑔𝑆 𝜌 𝐶𝐿max 𝑇̄rev + 𝐷̄ + 𝜇 𝑊 − 𝐿̄
DISTANCE

4.3.3 Brake enhancers

In addition to regular braking, we can use special elements to brake. Most notably, we can use thrust reversers,
as shown in figure 4.5a, or spoilers/lift dumpers as shown in figure 4.5b. The functioning of thrust reversers is
pretty logical; spoilers have two functions: increase aerodynamic drag directly, and decrease lift generated by
the wing, increasing the normal force and thus the ground friction.

(a) Thrust reverser.


(b) Spoiler alert.

Figure 4.5: Thrust reversers and spoilers.

4.4 Factors neglected so far


So far, we have made three important assumptions:
• There is no ground effect; in reality, there is, increasing the lift produced and decreasing the drag of the
aircraft, having a positive influence on take-off performance, but negative on landing performance (you
don’t want any lift produced during landing, as that will decrease the normal force and thus the friction
braking causes; furthermore, you want drag to be as high as possible).

©Sam van Elsloo


41 4.4. FACTORS NEGLECTED SO FAR

• There is no wind; in reality, there may be a head or tail wind (aircraft typically, if able to do so), take off
in the direction such that they have headwind.
• The runway does not have a slope.
The precise effect of ground effect is beyond the scope of this course, but the other two can be asked on the
exam.

4.4.1 Wind

Suppose we have a very strong wind of 80 m∕s (so basically a hurricane). It then may be the case that our
aircraft is already taking off, even though it is standing still on the ground, meaning that the take-off distance is
0 (perfect, isn’t it). Indeed, for an aircraft, it is the speed relative to the air molecules that is important, and not
the ground speed. So, if our aircraft moves at 80 m∕s relative to the ground, and there is a headwind of 20 m∕s,
then the airspeed will be 100 m∕s. Looking at figure 4.6, we realize the relation

A IRSPEED VS. 𝑉air = 𝑉ground + 𝑉wind (4.11)


GROUND AND
WIND SPEED

Figure 4.6: Ground, wind and airspeed.

Now, how does this alter anything? Well, actually, our integral was slightly wrong. You’ll agree with me that it
makes more sense to write
𝑑𝑉𝑔 𝑑𝑉𝑔
𝑎𝑔 = = 𝑉𝑔
𝑑𝑡 𝑑𝑠𝑔
(𝑉𝑔 )LOF
( ) 𝑉𝑔 𝑑𝑉𝑔
𝑠𝑔 𝑤 =
∫ 𝑎𝑔
0
( )
where 𝑠𝑔 𝑤 denotes the ground distance when wind is taking into account (later we’ll see the need for this
extra subscript 𝑤). Now, suppose we have constant wind speed, so that 𝑑𝑉𝑔 = 𝑑𝑉 and 𝑑𝑎𝑔 = 𝑑𝑎. Furthermore,
we have 𝑉𝑔 = 𝑉 − 𝑉𝑤 , where 𝑉 is the actual airspeed. We can then write the integral as

𝑉LOF ( )
( ) 𝑉 − 𝑉𝑤 𝑑𝑉
𝑠𝑔 𝑤 =
∫ 𝑎
𝑉𝑤

Why? We are now integrating over the air speed. Because is we have a headwind velocity 𝑉𝑔 , then the airspeed
will already be equal to 𝑉𝑔 once we start applying thrust. (𝑉𝑔 )LOF , i.e. the ground speed at which lift off occurs,
now simply becomes 𝑉LOF , i.e. the airspeed at which lift-off occurs. With 𝑉𝑤 constant, we can rewrite this as
follows
𝑉LOF ( ) 𝑉LOF 𝑉LOF 𝑉LOF 𝑉𝑤 𝑉LOF
( ) 𝑉 − 𝑉𝑤 𝑑𝑉 𝑉 𝑑𝑉 𝑉𝑤 𝑑𝑉 𝑉 𝑑𝑉 𝑉 𝑑𝑉 𝑑𝑉
𝑠𝑔 𝑤 = = − = − − 𝑉𝑤
∫ 𝑎 ∫ 𝑎 ∫ 𝑎 ∫ 𝑎 ∫ 𝑎 ∫ 𝑎
𝑉𝑤 𝑉𝑤 𝑉𝑤 0 0 𝑉𝑤

Now, carefully look at what each integral represents:

samvanelsloo97@icloud.com
CHAPTER 4. AIRFIELD PERFORMANCE 42

• The first integral represents the ground distance had no wind been present; this can easily be computed
using knowledge you already gathered by reading this chapter.
• The second integral represents the ground distance to reach the wind velocity; you have to subtract this
distance from the previous ground distance: you basically already start at a certain airspeed (equal to the
wind speed), so you don’t have to accelerate from 0 to 𝑉𝑤 , as your initial velocity is already 𝑉𝑤 .
• The third integral represents the distance an air molecule covers from standstill to take-off: 𝑉𝑤 is the
wind velocity, and the integral is simply
𝑉LOF
𝑑𝑉
= 𝑡𝑔
∫ 𝑎
𝑉𝑤

i.e. the time it takes to accelerate from 𝑉𝑤 to 𝑉LOF ; 𝑉𝑤 ⋅ 𝑡𝑔 is then the distance a wind particle travels in
the time it takes to accelerate. It’s rather logical that we have to subtract this.

4.4.2 Runway slope

The presence of a slope in the runway greatly affects performance; a positive slope (so your altitude increases
during the runway) adversely affects performance quite badly (a 4° slope can already double your take-off
distance). The reason for this becomes quite clear if you draw a FBD: the acceleration will now equal
𝑔 (̄ ( ) )
𝑎̄ = 𝑇 − 𝐷̄ − 𝜇 𝑊 − 𝐿̄ − 𝑊 sin 𝜉
𝑊
where 𝜉 is the slope of the runway. Essentially, part of the weight will now act as "gravity drag" (don’t really
know how to describe it), making the acceleration much lower.

4.5 Take-off speeds

Figure 4.7: Important 𝑉 -speeds.

To conclude this chapter, we take a look at figure 4.7, where multiple important speeds that are reached during
take-off are indicated. One-by-one, these are:
• 𝑉𝑚𝑠 : the minimum stalling speed. Speaks for itself.
• 𝑉𝑚𝑐 : the minimum control speed. This speed is of importance when one engine fails: suppose we have a
two-engine aircraft, with engines mounted to the wing. If one engine then fails, the remaining engine will
clearly generate a moment; this moment is counteracted by the rudder. If the speed is too low, the rudder
simply does not produce enough aerodynamic force to regain control over the aircraft. Only once the
speed reaches 𝑉𝑚𝑐 , the rudder is able to generate enough force to counteract the moment generated by the
remaining engine(s). Note that if your engine fails before this speed, you have to shut off all engines or
you die. Note that this also means that 𝑉1 must happen after 𝑉𝑚𝑐 : if we would have had, say 𝑉1 = 50 m∕s
and 𝑉𝑚𝑐 = 60 m∕s, and one engine would fail at 55 m∕s, then according to the definition of 𝑉1 , we’d
need to keep accelerating, but according to 𝑉𝑚𝑐 , we would need to shut off all engines because we’d keep
rotating otherwise, meaning that we die either way.1
1 In case you’re now kicking yourself, but what if our airplane is simply so good that 𝑉 < 𝑉 ? Then you set 𝑉 equal to 𝑉
1 𝑚𝑐 1 𝑚𝑐 and build
your runway based on this new 𝑉1 .

©Sam van Elsloo


43 4.6. RELEVANT OLD EXAM QUESTIONS

• 𝑉𝑚𝑐𝑔 : the minimum control speed on the ground. On the ground, the minimum control speed is slightly
higher than in the air.
• 𝑉𝑒𝑓 : the engine failure speed: the maximum velocity at which an engine may fail; slightly lower than 𝑉1
due to the reaction time of the pilots.
• 𝑉1 : decision speed.
• 𝑉𝑟 : the speed at which you start rotating the aircraft.
• 𝑉𝑚𝑢 : the minimum unstick speed: the minimum speed at which the airplane could take off by rotating as
much as possible, almost getting a tailstrike (that the tail hits the ground). Passengers don’t really like
knowing they’re taking off with the tail almost hitting the ground, so in reality, we accelerate a bit more,
so that we don’t have to rotate as much to generate sufficient lift.
• 𝑉LOF : lift-off speed: slightly higher than 𝑉𝑚𝑢 for above mentioned reason.
• 𝑉2 : free air safety speed.
Note that these speeds for a given aircraft depend on the aircraft weight (so how much payload and fuel is on
board), aircraft configuration and environmental conditions (air density for example).

4.6 Relevant old exam questions

4.6.1 Full exercises

• Exam June 2014, question 1.


• Exam January 2013, question 1.
• Exam January 2013, question 2.
• Exam November 2012, question 2.
• Exam February 2012, question 1.
• Exam January 2011, question 2. Note: the energy method is not discussed in the slides nor in the book,
so I suppose it’s not part of our study material any more.

4.6.2 Parts of exercises

None.

4.6.3 Exercises combining previously discussed material

None.

samvanelsloo97@icloud.com
CHAPTER 4. AIRFIELD PERFORMANCE 44

©Sam van Elsloo


5 Cruise
For cruise, there are three parameters of interest:
• Largest range
• Longest endurance
• Highest speed
We will limit ourselves to the discussion of largest range. For the longest endurance, the derivations are very
similar, and if you just take a tiny bit of effort to understand how we compute the range, you’ll also understand
how to compute the endurance. Furthermore, it is advisable to make the effort anyway to understand why we do
certain things in computing the range, as they can also ask for suboptimal cases (so flight strategies that do not
lead to maximum range). Derivations are then pretty similar in concept, but have small variations depending on
whether something is constant or not, for example.

5.1 Optimum cruise for simplified jet aircraft

5.1.1 Optimum flight condition for a single point in time

We will first focus on what flight conditions we have to select for a simplified jet aircraft at a certain point in
time to achieve the maximum range (thus, for a given altitude (𝜌) and 𝑊 ). For this, it is important to realize
that the specific range is given by 𝑉 ∕𝐹 , where 𝐹 is the fuel flow. You can derive it as follows: the fuel flow is
related to the weight of the aircraft by 𝐹 = −𝑑𝑊 ∕𝑑𝑡 (if the fuel flow is positive, the weight of the aircraft will
go down). The range obviously equals
𝑡2

𝑅= 𝑉 𝑑𝑡

𝑡1

As we have 𝑑𝑡 = −𝑑𝑊 ∕𝐹 , this becomes


𝑡2 𝑊2 𝑊1
𝑉 𝑉
𝑅= 𝑉 𝑑𝑡 = − 𝑑𝑊 = 𝑑𝑊
∫ ∫ 𝐹 ∫ 𝐹
𝑡1 𝑊1 𝑊2

so 𝑉 ∕𝐹 is the specific range. Clearly, we want to maximize this, but how do we do this? Well, first, look at the
performance diagram of figure 5.1. Note that we now have plotted 𝐹 on the vertical axis; however, remember
that
𝐹 ≡ 𝑐𝑇 𝑇
where 𝑐𝑇 is pretty much constant, so it still looks exactly the same as before. Furthermore, remember that the
maximum thrust is constant for a jet aircraft, independent of airspeed.
Now, we clearly want to maximize 𝑉 ∕𝐹 , or minimize 𝐹 ∕𝑉 . We can do this by drawing a line going through
the origin, tangent to the drag-curve: the angle between the 𝑥-axis and this tangent line is
𝐹
tan 𝛽 =
𝑉
Thus, when this angle is minimized, 𝐹 ∕𝑉 is minimized as well. You can draw several lines, but you’ll quickly
see that this angle is minimized when you draw a line tangent to the curve.
Now, how does this look like from a mathematical perspective? We must maximize

𝑉 1 𝑊 2 1
=
𝐹 𝑐𝑇 𝑇 𝑆 𝜌 𝐶𝐿

45
CHAPTER 5. CRUISE 46

Figure 5.1: Performance diagram - simplified jet.

but we have
𝐶𝐷
𝑇 =𝐷= 𝑊
𝐶𝐿
and thus √
( √ ) √ ( )
( ) 𝐶𝐿 1√ 𝐶
= √
𝑉 𝑊 2 1 1 2 𝐿
=
𝐹 max 𝑐𝑡 𝐶𝐿 𝑊 𝑆 𝜌 𝐶𝐿 𝑐𝑡 𝑊 𝑆 𝜌 𝐶 2
max 𝐷 max

as 𝑐𝑡 , 𝑊 , 𝑆 and 𝜌 are all constant for a given point in time. So, we want to maximize 𝐶𝐿 ∕𝐶𝐷
2 . How do we do

that? Well, perhaps you still remember from last year that this can be computed as follows:
( )
𝑑 𝐶𝐿
= 0
𝑑𝐶𝐿 𝐶 2
𝐷
( ) 2 − 𝐶 ⋅ 2𝐶 𝑑𝐶𝐷
𝐶𝐿 1 ⋅ 𝐶𝐷 𝐿 𝐷 𝑑𝐶
𝑑 𝐿
= =0
𝑑𝐶𝐿 𝐶 2 𝐶 4
𝐷 𝐷
𝑑𝐶
𝐶𝐷2
− 2𝐶𝐿 𝐶𝐷 𝐷 = 0
𝑑𝐶𝐿
𝑑𝐶
𝐶𝐷 − 2𝐶𝐿 𝐷 = 0
𝑑𝐶𝐿
Now, we have
𝐶𝐿2
𝐶𝐷 = 𝐶𝐷0 +
𝜋𝐴𝑒
𝑑𝐶𝐷 2𝐶𝐿
=
𝑑𝐶𝐿 𝜋𝐴𝑒
so that

𝑑𝐶𝐷 𝐶2 2𝐶
𝐶𝐷 − 2𝐶𝐿 = 𝐶𝐷0 + 𝐿 − 2𝐶𝐿 𝐿 = 0
𝑑𝐶𝐿 𝜋𝐴𝑒 𝜋𝐴𝑒
3𝐶𝐿2
𝐶𝐷0 − = 0
𝜋𝐴𝑒

1
𝐶𝐿𝑜𝑝𝑡 = 𝐶 𝜋𝐴𝑒
3 𝐷0
√ 2
1
3
𝐶 𝐷 0
𝜋𝐴𝑒 4
𝐶𝐷𝑜𝑝𝑡 = 𝐶𝐷0 + = 𝐶𝐷0
𝜋𝐴𝑒 3
This is a very important result: for a given aircraft, which will have fixed 𝐶𝐷0 , 𝑒 and 𝐴, the optimum angle
of attack (for maximum range), which determines 𝐶𝐿 and 𝐶𝐷 is fixed; it is independent of aircraft weight,

©Sam van Elsloo


47 5.1. OPTIMUM CRUISE FOR SIMPLIFIED JET AIRCRAFT

flight velocity, altitude and everything else. In turn, the flight velocity is directly dependent on the required lift
coefficient for maximum range:

𝑊 2 1
𝑉 =
𝑆 𝜌 𝐶𝐿opt

So, we can easily compute what flight conditions (what angle of attack and what velocity) we have to fly at for
a given altitude and aircraft weight to maximize the specific range: 𝐶𝐿 follows from above derivation, which
determines both 𝛼 and 𝑉 . Additionally, we can easily compute 𝑉 ∕𝐹 now: 𝑉 is known, and 𝐹 follows from
𝐹 = 𝑐𝑇 𝑇 : the pilot simply selects the thrust level such that the thrust equals the drag, which can be readily
computed based on the drag coefficient.

5.1.2 Cruise strategy

So, now we know how to compute everything once we know 𝑊 and the altitude 𝐻. We can’t really do too much
about weight right now, but what is the optimal altitude to fly at when you’re cruising? For that, we remember
what happens with the performance diagram if we increase the altitude. This is shown in figure 5.2 (the red lines
are the maximum thrust levels that can be applied during cruise; this is a bit smaller than the take-off thrust
as applying maximum thrust for the entire duration of the flight is very bad for the engines). Remember from
chapter 1 that the drag curves shift to the right: drag simply equals 𝐷 = 𝐶𝐷 ∕𝐶𝐿 ⋅ 𝑊 , so a decrease in 𝜌 has no
effect on the drag. However,

𝑊 2 1
𝑉 =
𝑆 𝜌 𝐶𝐿

so 𝑉 ∝ 1∕𝜌; thus, the velocity at which one must fly increases with altitude, thus each point is moved
horizontally to the right. Furthermore, the available thrust decreases according to

𝑇 𝜌
=
𝑇0 𝜌0
𝑇0
𝑇 = 𝜌=𝜅⋅𝜌
𝜌0

so 𝑇 ∝ 𝜌, and thus the thrust lines decrease with increasing altitude (this makes sense, as the density of the
mass flow coming through the inlet is smaller).

Figure 5.2: Performance diagram at several altitudes.

Now, what is the point I’m trying to make? Remember that the angle between the tangent and the 𝑥-axis was
an indication for 𝐹 ∕𝑉 ; the smaller the better. From figure 5.2, we see that it is thus best to just fly as high as
possible, until we reach the cruise thrust limit: indeed, it is optimum to fly as high as possible, as high as the
maximum cruise thrust allows you to.

What does this mean in the long term? Remember what happens to the performance diagram if the weight of

samvanelsloo97@icloud.com
CHAPTER 5. CRUISE 48

the aircraft decreases (which happens during cruise flight as fuel is burned). We have
𝐶𝐷
𝐷 = 𝑊 ∝𝑊
𝐶𝐿

𝑊 2 1 √
𝑉 = ∝ 𝑊
𝑆 𝜌 𝐶𝐿

This means that if the weight of the aircraft is halved, the drag is decreased by a factor 0.5, whereas the velocity
is decreased by a factor 0.707. This means that the performance diagram of figure 5.3 shifts to the lowerleft,
shifting more downwards than it shifts to the left.

Figure 5.3: Effect of weight on performance diagram.

Where do we go from here? What do we do? Do we just fly at the new black dot, so at a lower speed, but
constant altitude1 ? Or do we do something else? We indeed do something else: remember that the only reason
why we wouldn’t fly higher is that we did not have enough thrust to fly higher. However, if our weight decreases
by 20%, then our drag would also decrease by 20%. This means we have excess thrust, and thus we can climb a
bit higher as well. In fact, remember that our thrust was related as follows
𝑇2 𝜌2
=
𝑇1 𝜌1
𝜌2
𝑇2 = 𝑇1
𝜌1
So, if our deliverable thrust 𝑇2 may decrease by 20% compared to 𝑇1 , then we can fly at an altitude where the
density 𝜌2 is 20% lower than it was at 𝜌1 too. So, by increasing the altitude from 𝐻1 to 𝐻2 , where the density is
equal to 𝜌2 , we manage to once again fly at our thrust limit, which we discussed before was the bay way to cruise.
Now, what does this do to our velocity? We have 𝑊2 = 0.8𝑊1 (in this example), and 𝜌2 = 0.8𝜌1 . Thus,
√ √ √
𝑊2 2 1 0.8𝑊1 2 1 𝑊1 2 1
𝑉2 = = = = 𝑉1
𝑆 𝜌2 𝐶𝐿 𝑆 0.8𝜌1 𝐶𝐿 𝑆 𝜌1 𝐶𝐿

and thus 𝑉2 = 𝑉1 : we fly at constant airspeed! So, to summarize:


• We fly at constant angle of attack, as we keep 𝐶𝐿 ∕𝐶𝐷
2 maximum.

• We fly at constant velocity.


• We gradually climb throughout our cruise: we continuously increase our altitude such that we can fly at
our maximum deliverable cruise thrust.
Note: you may think it’s strange that we’re able to fly at constant velocity, as this would mean that lift is constant,
but as the weight increases, this causes our airplane to accelerate upwards, which would surely be wrong? Well,
fortunately, this does not happen: the decrease in density caused by climbing precisely offsets the decrease in
weight.
1 Note that the angle of attack is also constant; in both cases, the black dot corresponds to maximum 𝐶 ∕𝐶 2 , which corresponds to 1
𝐿 𝐷
angle of attack, which only depends on the geometry of the aircraft and not on the flight conditions.

©Sam van Elsloo


49 5.1. OPTIMUM CRUISE FOR SIMPLIFIED JET AIRCRAFT

5.1.3 Range equation

So, now, what is actually our maximum range? Remember the formula we saw at the beginning,

𝑡2 𝑊2 𝑊1
𝑉 𝑉
𝑅= 𝑉 𝑑𝑡 = − 𝑑𝑊 = 𝑑𝑊
∫ ∫ 𝐹 ∫ 𝐹
𝑡1 𝑊1 𝑊2

We have 𝐹 = 𝑐𝑇 𝑇 = 𝑐𝑇 𝐷 = 𝑐𝑇 𝐶𝐷 ∕𝐶𝐿 ⋅ 𝑊 :

𝑊1
𝑉 𝐶𝐿 1
𝑅= 𝑑𝑊
∫ 𝑐 𝑇 𝐶𝐷 𝑊
𝑊2

Now, 𝑐𝑇 is pretty much always constant, and as we fly at constant angle of attack and constant velocity, 𝑉 , 𝐶𝐿
and 𝐶𝐷 are all constant as well. Thus, integration simply leads to

( )
O PTIMUM 𝑉 𝐶𝐿 𝑊1
𝑅= ln (5.1)
RANGE FOR 𝑐𝑇 𝐶𝐷 𝑊2
SIMPLIFIED JET
AIRCRAFT
Note that this only holds when we fly at constant angle of attack and constant velocity! If your exam question
says you don’t fly at constant velocity, you need to take appropriate measures, depending on how exactly the
aircraft flies. For example, if we instead fly at constant altitude and angle of attack (instead of constant velocity),
the derivation becomes

𝑊1 𝑊 2 1 √ 𝑊1
𝑆 𝜌 𝐶𝐿 𝐶𝐿 1 1 12 1
𝑅 = 𝑑𝑊 = ⋅ 𝐶𝐿 √ 𝑑𝑊
∫ 𝑐𝑇 𝐶𝐷 𝑊 𝑐𝑇 ⋅ 𝐶𝐷 𝑆𝜌 ∫ 𝑊
𝑊2 2𝑊
√ (√ )
2 12 √
= ⋅ 𝐶𝐿 𝑊1 − 𝑊2
𝑐 𝑇 ⋅ 𝐶𝐷 𝑆𝜌

5.1.4 Operational limitations

There are two important practical problems to remark about what we’ve done so far:

• First of all, it may be that you run into an operational speed velocity, as shown in figure 5.4a. We can’t
increase our speed beyond this limit; otherwise, a wind gust would increase the load factor too much and
we’d all crash and die, which is not really how you reach maximum range (also it sucks if you die, but
that aside). So what do we do then? Well, you may be inclined to think that we should our altitude such
that when we fly at 𝑉lim , we fly at the point of maximum 𝐶𝐿 ∕𝐶𝐷 2 (this point corresponds to the ’middle’

blue dot, and the middle drag curve). However, we can lower our 𝐷∕𝑉 (which directly indicates how
low 𝐹 ∕𝑉 is, which we want to keep as low as possible), by flying at a slightly higher altitude, such that
when we fly at 𝑉lim , we fly at minimum drag. In other words: we set the velocity equal to 𝑉lim , and then
compute at what altitude we must fly so that this corresponds to the point of minimum drag (it’s basically
drawing the associated drag curves for several altitudes, and then checking for which altitude the point
of minimum drag lies exactly above 𝑉lim . Now, what is the drag associated with minimum drag again?
Remember that we have
𝐶𝐷
𝐷= 𝑊
𝐶𝐿

samvanelsloo97@icloud.com
CHAPTER 5. CRUISE 50

thus we must minimize 𝐶𝐷 ∕𝐶𝐿 : we have

( ) 𝑑𝐶𝐷
𝐶 − 𝐶𝐷
𝑑 𝐶𝐷 𝑑𝐶𝐿 𝐿
= = 0
𝑑𝐶𝐿 𝐶𝐿 𝐶𝐿2
𝑑𝐶𝐷
𝐶 − 𝐶𝐷 = 0
𝑑𝐶𝐿 𝐿
𝐶𝐿2
𝐶𝐷 = 𝐶𝐷0 +
𝜋𝐴𝑒
𝑑𝐶𝐷 2𝐶𝐿
=
𝑑𝐶𝐿 𝜋𝐴𝑒
𝑑𝐶𝐷 2𝐶𝐿2 𝐶2
𝐶𝐿 − 𝐶𝐷 = − 𝐶𝐷0 − 𝐿 = 0
𝑑𝐶𝐿 𝜋𝐴𝑒 𝜋𝐴𝑒

𝐶𝐿 = 𝐶𝐷0 𝜋𝐴𝑒
√ 2
𝐶𝐷0 𝜋𝐴𝑒
𝐶𝐷 = 𝐶𝐷0 + = 2𝐶𝐷0
𝜋𝐴𝑒

and you now have pretty much everything you need to know to calculate stuff.
• Secondly, gradually climbing throughout your entire flight is generally not appreciated by air traffic
control, as they like everyone to just stay at constant altitude so not everyone flies into each other, as this
would again destroy your range. Therefore, we usually follow the strategy depicted in figure 5.4b: we
climb in steps; note that to fly at constant altitude, we now do have to continuously decrease our velocity,
as otherwise we’d start creating excess lift.

(b) Operational cruise implementation.


(a) Operational speed limit.

Figure 5.4: Operational speed limit and cruise implementation.

5.2 Range for propeller aircraft

For propeller aircraft, we can do a very similar derivation: we again must maximize 𝑉 ∕𝐹 , but this time, we
have
𝑃 𝐶𝑝
𝐹 = 𝐶𝑝 𝑃𝑏𝑟 = 𝐶𝑝 𝑎 = ⋅ 𝐷𝑉
𝜂 𝜂

and thus we must maximize


𝑉 𝑉 𝜂 1
= 𝐶𝑝
=
𝐹 ⋅ 𝐷𝑉 𝐶𝑝 𝐷
𝜂

©Sam van Elsloo


51 5.3. UNIFIED RANGE EQUATION

and thus we must minimize 𝐷. We already proved that this occurs for

𝐶𝐿 = 𝐶𝐷0 𝜋𝐴𝑒
√ 2
𝐶𝐷0 𝜋𝐴𝑒
𝐶𝐷 = 𝐶𝐷0 + = 2𝐶𝐷0
𝜋𝐴𝑒
Now, what is the optimum cruise strategy? Must we again fly as high as possible? The answer is yes, but not for
the same reason as before. Look at figure 5.5, where the effect of increasing altitude is shown: the drag curve
shifts to the right. However, 𝐷min remains constant, so 𝑉 ∕𝐹 is not affected by altitude, so it does not matter at
what altitude we fly, the minimum fuel flow remains constant. However, it is nonetheless beneficial to fly at
higher altitude: we then simply fly faster, so we’ll be there faster. Anyway, we now get for the range

𝑊2 𝑊2 𝑊2
𝑉 𝜂 1 𝜂 𝐶𝐿 1
𝑅 = 𝑑𝑊 = 𝑑𝑊 = 𝑑𝑊
∫ 𝐹 ∫ 𝐶𝑝 𝐷 ∫ 𝐶𝑝 𝐶𝐷 𝑊
𝑊1 𝑊1 𝑊1

O PTIMUM 𝜂 𝐶𝐿 𝑊 2
ln (5.2)
RANGE FOR A 𝐶𝑃 𝐶𝐷 𝑊 1
TURBOPROP

Figure 5.5: Effect of increasing altitude on propeller performance diagram.

5.3 Unified range equation


Now that we’ve seen both, you must surely (probably not) be wondering whether there’s a way to combine these
equations into a unified range equation. Yes, we can, actually.
First of all, the total efficiency of a propulsion system is simply

𝑃𝑎
𝜂total =
𝑊
where 𝑄 is the thermal power, equal to
𝐻
𝑄=𝐹
𝑔
where 𝐹 is the fuel flow, and 𝐻 the joule energy per kilogram of the fuel (energy density). Note that the 𝜂 you
saw for propeller aircraft is not 𝜂total , as 𝜂 = 𝑃𝑎 ∕𝑃𝑏𝑟 . Let’s now try to rewrite stuff for the range equation of both
the jet aircraft and the propeller aircraft.

samvanelsloo97@icloud.com
CHAPTER 5. CRUISE 52

5.3.1 Jet aircraft

For jet aircraft, we have


𝑃𝑎 𝑇𝑉
𝜂𝑡𝑜𝑡𝑎𝑙 = =
𝑄 𝐹 ⋅ 𝐻𝑔

so that
𝑉 𝐻 1
= 𝜂𝑡𝑜𝑡𝑎𝑙
𝐹 𝑔 𝐷
Furthermore, 𝐹 = 𝑐𝑇 𝑇 and 𝑇 = 𝐷, and thus

𝑉 𝐻 1
= 𝜂𝑡𝑜𝑡𝑎𝑙
𝑐𝑇 𝑇 𝑔 𝑇
𝑉 𝐻
= 𝜂𝑡𝑜𝑡𝑎𝑙
𝑐𝑇 𝑔

So, if we have ( ) ( )
𝑉 𝐶𝐿 𝑊0 𝐻 𝐶𝐿 𝑊0
𝑅𝑗𝑒𝑡 = ln = 𝜂𝑡𝑜𝑡𝑎𝑙 ⋅ ln
𝑐𝑇 𝐶𝐷 𝑊0 𝑔 𝐶𝐷 𝑊1

5.3.2 Propeller aircraft

For a propeller aircraft, we once again have


𝑃𝑎 𝑇𝑉
𝜂𝑡𝑜𝑡𝑎𝑙 = =
𝑄 𝐹 ⋅ 𝐻𝑔

so that
𝑉 𝐻 1
= 𝜂𝑡𝑜𝑡𝑎𝑙
𝐹 𝑔 𝐷
We can also write
𝑐𝑃 𝑐
𝐹 = 𝑃𝑎 = 𝑃 𝑇 𝑉
𝜂 𝜂
𝑉 𝑉 𝜂 𝑉 𝜂 1 𝐻 1
= 𝑐𝑃 = = = 𝜂𝑡𝑜𝑡𝑎𝑙
𝐹 𝑇 𝑉 𝑐𝑃 𝑇 𝑉 𝑐𝑃 𝑇 𝑔 𝐷
𝜂

where the final part came from what we just derived above. Again, we have have 𝑇 = 𝐷, and thus
𝜂 1 𝐻 1 𝐻 1
= 𝜂𝑡𝑜𝑡𝑎𝑙 = 𝜂𝑡𝑜𝑡𝑎𝑙
𝑐𝑃 𝑇 𝑔 𝐷 𝑔 𝑇
𝜂 𝐻
= 𝜂𝑡𝑜𝑡𝑎𝑙
𝑐𝑃 𝑔

so that we can write ( ) ( )


𝜂 𝐶𝐿 𝑊1 𝐻 𝐶𝐿 𝑊1
𝑅𝑝𝑟𝑜𝑝 = ln = 𝜂𝑡𝑜𝑡𝑎𝑙 ⋅ ln
𝑐 𝑝 𝐶𝐷 𝑊2 𝑔 𝐶𝐷 𝑊2

5.4 Economics
You probably still remember the beautiful diagram of figure 5.6. In there, we can distinguish a few important
things. The operative empty weight is obviously constant for all possible ranges. For 0 range, we can completely
fill up the payload, leading to the maximum zero fuel weight. On top of that, we are required to have some
reserve fuel; the required amount is constant for all desired ranges. To increase the range, we can simply take
some fuel on board; this leads to the line AB, which denotes the take-off weight. After B, the sum of OEW,
payload and fuel weight reaches the maximum take-off weight; if you want to increase the range, you’ll need to

©Sam van Elsloo


53 5.4. ECONOMICS

replace payload with fuel. Point B is therefore called the design range. The amount of payload that can be taken
on board then steadily reduces as we increase the range. In theory, the line would continue on until point D,
where there is no payload on board and the reserve fuel is still unused; however, in practice, the fuel tanks have
a finite size and at point E, we reach this fuel tank capacity limit. The only way to increase the range then is
to get rid of some payload, but one must get rid of a lot of payload to increase the range a bit, hence the steep
decrease in slope. Point F is the actual maximum range of the aircraft, when the fuel tanks are fully filled and
there is no payload on board.
The progress of the payload is once again visualized in the right part of figure 5.6: first, it is only limited by
volumetric or seating capacity; after that, the MTOW becomes the limiting factor, and finally it becomes limited
by fuel tank capacity.

Figure 5.6: Payload-weight diagram.

Now, onto the real business aspect of this section (at least, as businessy as it’s gonna get, unfortunately). In
commercial air transport, several parameters are of interest, which will be explained in the order listed:
• 𝐸𝐵 : block time
• 𝑉𝐵 : block speed
• 𝑃𝑟 : transport product
• 𝑃ℎ : transport productivity
• 𝑃𝑦 : revenue-earning capacity
The block time is the total time elapsing from starting engines at the departure airport to engines off at the
destination place, so including taxiing, takeoff, climb, etc. The block speed is the block distance divided by the
block time, i.e.
𝑅
𝑉𝐵 =
𝐸𝐵
which will clearly be lower than the cruise speed. In fact, if we write the block time as
𝑅
𝐸𝐵 = + Δ𝑡
𝑉𝑐𝑟
where Δ𝑡 is the length of time that accounts for the field operations and the lower airspeeds in flight phases
other than the cruise and 𝑉𝑐𝑟 the cruise speed, then we can write
𝑅
𝑉𝐵 = 𝑅
𝑉𝑐𝑟
+ Δ𝑡

If we fix 𝑉𝑐𝑟 and fix Δ𝑡 (e.g. to 50 minutes), the graph of figure 5.7c results.
Obviously, the revenue of a flight primarily depend on the payload weight and the range. Thus, it makes sense
to define the transport product as
𝑃𝑟 = 𝑊𝑃 𝑅

samvanelsloo97@icloud.com
CHAPTER 5. CRUISE 54

Figure 5.7: Various economic performance parameters.

Using figure 5.7a as reference, one can calculate this for each range, resulting in figure 5.7b.

The transport productivity is simply the transport product divided by the block time:

𝑊𝑃 𝑅
𝑃ℎ = = 𝑊𝑃 𝑉 𝐵
𝐸𝐵

From the combination of figure 5.7b and 5.7b, we get the graph of figure 5.7d. We clearly see that maximum 𝑃ℎ
occurs at the design range.

Finally, the revenue-earning capacity is the transport product per year; if 𝑈 is the annual flight utilization of the
airplane in hours, we have
𝑃𝑦 = 𝑃ℎ 𝑈 = 𝑊𝑃 𝑉 𝐵 𝑈

5.4.1 Operating costs

We distinguish two types of operating costs2 :

• Direct operational costs: these are directly associated with the operation of a single airplane. For example,
fuel costs is a direct operational cost: had you not operated a certain airplane, you also wouldn’t have had
to pay the fuel costs for that specific flight. Same can be said for the salary of the crew, landing fees etc.
• Indirect operational costs: these are costs for a company that can’t be directly blamed on one specific
flight/airplane; for example, the management of a company would still have gotten salary, even if that
airplane did not fly. Other such costs are costs related to advertising, housing, administration, and depre-
ciation (afschrijven) of ground properties and equipment. Remarkable, the slides only give depreciation
of the aircraft as example of indirect operational costs, however, I’m pretty sure those are not part of
indirect operational costs, but are direct operational costs (and although I may not have a degree in flight
mechanics yet, I do have one in accounting): depreciation of your airplane is directly associated with
using the airplane so it should be a direct operational cost.

5.5 Wind

Again, just like for take-off, wind has an effect on cruise flight. In fact, rather than optimising 𝑉 ∕𝐹 , we ought to
maximize 𝑉𝑔 ∕𝐹 , where
𝑉𝑔 = 𝑉 − 𝑉𝑊

2 Apparently, the airliner industry is quite rebellious in how it names these costs. In pretty much every other industry, we simply use the

terms ‘direct costs’ and ‘indirect costs’ instead of putting operating costs in there.

©Sam van Elsloo


55 5.6. CRUISE AT TRANSONIC FLIGHT CONDITIONS

using the same logic as we did previously, using figure 5.8 (headwind is positive, tailwind is negative). We then
get for propeller and jet aircraft, respectively:
[ ]
𝑉𝑔 𝑉𝑔 𝜂𝑗 𝑉 − 𝑉 𝑊
= =
𝐹 𝑐𝑃 𝑃𝑏𝑟 𝑐𝑃 𝑃𝑟
[ ]
𝑉𝑔 𝑉𝑔 1 𝑉 − 𝑉𝑊
= =
𝐹 𝑐𝑇 𝑇 𝑐𝑇 𝐷

Figure 5.8: Ground, wind and airspeed.

Thus, we most minimize 𝑃𝑟 ∕(𝑉 − 𝑉𝑊 ) and 𝐷∕(𝑉 − 𝑉𝑊 ), respectively. What difference does this make? Well,
look at figure ??. The left figure corresponds to propeller-driven airplanes, and the right figure to jet airplanes.
Remember that we must now minimize the ratio 𝑃𝑟 ∕(𝑉 − 𝑉𝑊 ) (the very same can be said about the right figure,
but for brevity’s sake I’ll limit myself to the propeller one). If we have no wind present, nothing changes,
obviously. If we have headwind present, then 𝑉𝑊 is positive; thus, if you’d stay on the same point in the 𝑃𝑟 -curve,
the ground speed would be reduced by the magnitude of the head wind, thus the vertical axis shifts to the right.
We then draw a line through the new origin, tangent to the curve; the point where they are tangent is the new
cruise condition: the new air speed is then the distance to the old vertical axis (whereas the distance to the new
vertical axis is the ground speed); this is the point where 𝑃𝑟 ∕𝑉𝑔 is minimized.
Similarly, if we have a tailwind, then 𝑉𝑤 will be negative and for a single point on the 𝑃𝑟 curve, the ground
speed would be the magnitude of 𝑉𝑤 larger. Thus, the vertical axis shifts to the left (by a distance |𝑉𝑊 |); the
distance with respect to this new axis is the ground speed in case of tailwind, the distance to the old axis is
still the air speed. However, it is 𝑃𝑏𝑟 ∕𝑉𝑔 that we need to minimize, and thus we draw the line through the new
origin.
What does it change to our actual range? Well, we can readily compute that quite easily:
𝑊1 𝑊1 ( ) 𝑊1 𝑊1
𝑉𝑔 𝑉 − 𝑉𝑊 𝑉 𝑑𝑊
𝑅= 𝑑𝑊 = 𝑑𝑊 = 𝑑𝑊 − 𝑉𝑊
∫ 𝐹 ∫ 𝐹 ∫ 𝐹 ∫ 𝐹
𝑊2 𝑊2 𝑊2 𝑊2

Now, the integral on the left is simply equal to the range had no wind been present, 𝑅(𝑉𝑊 =0) ; the integral on the
right is simply the time the flight takes (note that this is not equal to the maximum endurance, but simply the
endurance associated with maximum range). We can thus write
𝑅 = 𝑅(𝑉𝑊 =0) − 𝑉𝑊 𝐸

5.6 Cruise at transonic flight conditions


Classical cruise performance assumes
• Single, constant lift drag polar
• Maximum power / thrust independent of airspeed
• Power /thrust specific fuel consumption constant
First, let’s look at the specific range. For this, the total efficiency equals
𝑃𝑎 𝑇𝑉
𝜂total = =
𝑄 𝐹 ⋅ 𝐻𝑔

samvanelsloo97@icloud.com
CHAPTER 5. CRUISE 56

With 𝑇 = 𝐷 = 𝐶𝐷 ∕𝐶𝐿 𝑊 , this becomes

𝑉 𝐶 𝐻 1
= 𝜂total 𝐿
𝐹 𝐶𝐷 𝑔 𝑊

Now, the total efficiency depends on the Mach number and 𝐻, and 𝐶𝐿 ∕𝐶𝐷 depends on the Mach number. As
things get rather ugly, we’ll just look for a graphical solution in the end.
Let’s first discuss only the effect of the dependence of 𝐶𝐿 ∕𝐶𝐷 on the Mach (but we assume a simplified jet).
We then have:
𝑉 𝑉 𝑉 𝐶𝐿 1 𝑎 𝐶 1
= = = 𝑀 𝐿
𝐹 𝑐𝑇 𝑇 𝑐𝑇 𝐶𝐷 𝑊 𝑐𝑇 𝐶𝐷 𝑊
( )
as 𝑉 = 𝑀𝑎. So, we need to maximize 𝑀 𝐶𝐿 ∕𝐶𝐷 . For a single Mach number, the drag polar looks like as
shown in figure 5.9. The optimum then follows from as drawn.

Figure 5.9: Drag polar for single Mach number.

Now, note that



𝑊 2 1
𝑉 = =𝑀 ⋅𝑎
𝑆 𝜌 𝐶𝐿

so that
1 𝑊 2 1
𝑀2 =
𝑎2 𝑆 𝜌 𝐶𝐿
Note that for a given altitude, 𝑎, 𝑊 , 𝑆 and 𝜌 are constant: thus, to have horizontal flight (no climb or descent,
𝐿 = 𝑊 ), you can draw a graph as shown in figure 5.10, which makes clear that for each value of the lift
coefficient, there is only one unique 𝑀 at which may be flown.

Figure 5.10: Required lift coefficient.

So, for a given 𝐶𝐿,𝑜𝑝𝑡 , we can determine the optimum Mach number by

1 𝑊 2 1
𝑀=
𝑎 𝑆 𝜌 𝐶𝐿,𝑜𝑝𝑡

©Sam van Elsloo


57 5.6. CRUISE AT TRANSONIC FLIGHT CONDITIONS

Now, suppose we complicate stuff and realize that we have different drag polars for different Mach num-
bers:
𝐶𝐷 = 𝐶𝐷0 (𝑀) + 𝑘 (𝑀) 𝐶𝐿2
where the (𝑀) indicate that it is a function of 𝑀 (not that you’re multiplying). This can be plotted as shown in
figures 5.11a and 5.11b. It is clear the the maximum 𝐶𝐿 ∕𝐶𝐷 decreases as 𝑀 increases. So, for any point in
figure 5.10, we can look up the 𝐶𝐷 associated with it from figure 5.11a. So, you can do this for all combinations
of 𝐶𝐿 and 𝑀 (even the ones that are not on the line (yes in real life you have to fly at a combination on that line,
but we ignore that for one second)), and then compute 𝑀 ⋅ (𝐶𝐿 ∕𝐶𝐷 ) for each combination, and plot it as shown
in figure 5.12a: we can then simply look the optimum 𝐶𝐿 ∕𝐶𝐷 by drawing again the 𝐶𝐿 − 𝑀 2 line (the red
dotted line) (which is the line which contains the combinations that we are allowed to fly one: anything else will
cause us to descend or climb, which will cause death and that would be a relatively unfortunate event). Note
that the blue lines are independent of altitude, as they only depend on the aircraft geometry and aerodynamic
parameters (it is the multiplication of 𝑀 with 𝐶𝐿 ∕𝐶𝐷 , neither of them depend on atmospheric conditions (𝑀 is
something you choose yourself here, it’s not determined by the speed of sound)). However, the red dotted line is
dependent on altitude, as it depends on 𝜌, thus we can plot several of these lines for different altitudes, as shown
in figure 5.12b: for each altitude, we should fly at the point which yields us the highest value for 𝑀𝐶𝐿 ∕𝐶𝐷 , i.e.
the denoted green dots.

(a) High speed drag polars. (b) High speed drag polars.

Figure 5.11: High speed drag polars.

(a) 𝑀 ⋅ 𝐶𝐿 ∕𝐶𝐷 for combinations of 𝐶𝐿 and 𝑀, including(b) 𝑀 ⋅ 𝐶𝐿 ∕𝐶𝐷 for combinations of 𝐶𝐿 and 𝑀, including
𝐶𝐿 𝑀 2 line for specific altitude. multiple𝐶𝐿 𝑀 2 lines for multiple altitudes.

Figure 5.12: 𝑀 ⋅ 𝐶𝐿 ∕𝐶𝐷 plots.

Now, the thrust specific fuel consumption also varies with altitude (and it increases with Mach number), as
shown in figure 5.13. We use a numerical analysis to find the optimum condition to fly at to have the best range.
We do this by looking at figure 5.14a: we can fly at any of the points enclosed by the blue line; each point relates
to a specific altitude and specific speed. We can thus know 𝐶𝐿 , and then calculate 𝐶𝐷 , so that we know the drag,
which is equal to the thrust, so that we can compute the specific fuel consumption, by looking up the applicable
value of 𝐶𝑇 . You can do this for all the points, and you get a figure as shown in figure 5.14b.
The blue point is then the point where 𝑉 ∕𝐹 is maximized; iso-lines of lower 𝑉 ∕𝐹 -values are also drawn.
So, in conclusion: the optimum turbofan cruise conditions
• are unrealistic if compressibility effects are ignored

samvanelsloo97@icloud.com
CHAPTER 5. CRUISE 58

Figure 5.13: Variation of 𝐶𝑇 with Mach number.

(b) 𝑉 ∕𝐹 for all combinations of 𝑀 and 𝐻.


(a) 𝐻 − 𝑉 diagram.

Figure 5.14: 𝐻 − 𝑉 and 𝑉 ∕𝐹 plots.

• are at Mach number well within drag rise


• can be computed using aforementioned scheme:
– One takes a point in the 𝐻 − 𝑉 diagram of figure 5.14a.
– One calculates the required lift coefficient.
– One calculates the associated drag coefficient, taking into account the compressibility effects.
– One calculates the drag and thus the required thrust.
– One looks up the correct value of 𝑐𝑇 and computes the fuel flow 𝐹 = 𝑐𝑇 𝑇 .
– One computes the associated 𝑉 ∕𝐹 , and repeats this procedure for every point enclosed in figure
5.14a. The maximum value of 𝑉 ∕𝐹 is then the optimum flight condition.

5.7 Relevant old exam questions

5.7.1 Full exercises

• Exam June 2014, question 2.


• Exam April 2014, question 2.
• Exam January 2013, question 4.
• Exam November 2012, question 1.
• Exam November 2011, question 4.
• Exam October 2010, question 3.

5.7.2 Parts of exercises

None.

©Sam van Elsloo


59 5.7. RELEVANT OLD EXAM QUESTIONS

5.7.3 Exercises combining previously discussed material

None.

samvanelsloo97@icloud.com

Das könnte Ihnen auch gefallen