You are on page 1of 82

Abstract

This project aimed to explore the requirements needed to take a payload into a Low Earth
Orbit at an altitude of 100, and the concept of a Hohmann transfer to re-position the
payload with the International Space Station.
After concluding the final velocity of a single stage rocket to be 4.7118!"!!! , it was
discovered that the payload needed a propellant burn out velocity of 7.8490!"!!! , and
therefore the concept of multi-staging was introduced. Exploring 2-stage rockets resulted in a
final speed of 6.1145!"!!! , and 7.6463!"!!! when the mass of the 1st stage was 10.0499
times greater than the 2nd. This eventually indicated the need for an optimized 3-stage rocket.
However when investigating the effects of external forces on a single stage rocket,
0.5227!"!!! of final velocity had been lost to gravity and drag. This prompt the decision to
design a rocket capable of reaching 10!"!!! in hope to provide more realistic results, as
external forces had not been incorporated. Given a mass ratio of 6.1921 and a final mass of
236.4181, a comparison was made with the Saturn V rocket, and from this a more realistic
rocket was then designed that had a mass ratio of 3.0843, and final mass of 28.3407.
Assuming a payload mass of 1000 and that the flight path was known, the distance
traveled by the 3-Stage rocket was calculated to be 229.5587, which required a propellant
flow rate of 207.3962!"!!! , and a burn time of at least 122.9850!"#$%&! to reach the orbits
requirements.

Furthermore,

using

the

individual

stage

masses

19,827.8, 6,428.6 & 2,084.4 , the distance, burn time, altitude and horizontal
distance for each stage was calculated as displayed in the Summary of Findings.
Finally, the velocity required to complete the Hohmann transfer for the 1st and 2nd maneuver
were found to be 0.0798!"!!!

& 0.079!"!!! , with a time delay in-between of

2,671.1019!"#$%&! .
Acknowledgements
I would like to thank Dr Nigel Atkins for his help in constructing this project; he was always
available I when required assistance, he made sure I kept inline and didnt stray off path, and
made sure I thoroughly understood the concepts behind my research. Because of this, my
paper has turned out to be far better than I could have ever imagined, and I sincerely thank
him for his guidance.

Contents
Chapter 1: Introduction

3-4

1.2 - The use of rockets

Chapter 2: Aim and Assumptions

2.1 - Aims and Objectives

2.2 Assumptions

Chapter 3: Single Stage rocket Motion

6-18

3.1 - General solution for velocity at any time t without external forces

6-9

3.2 - General solution for velocity at any time t: Including external forces

9-11

3.3 - Distance Traveled: Single stage displacement

11-15

3.4 - Final Velocity at Propellant burn out

16

3.5 - Newtonian Orbital Velocity for a circular orbit

16-18

Chapter 4: 2-Stage rocket


4.1 - 2-Stage rocket velocity
Chapter 5: Optimized 2-Stage Rocket

19-21
18-20
22-25

5.1 - Optimized 2-stage rocket velocity

22-24

5.2 - Comparison of all the model rocket velocities

25

Chapter 6: 3-Stage Rocket mass

26-32

6.1 - 3 Stage optimized equation

26-28

6.2 - The Method of Lagrange Multipliers

28-32

Chapter 7: 3-Stage Evaluation

33-34

7.1 - Comparing model rocket to Saturn V

33

7.2 - Improved overall Mass

34

Chapter 8: 3-Stage Flight path & Distance

35-42

8.1 Flight path

35-37

8.2 Evaluating the Distance equation

37-42

Chapter 9: Individual Distance, Altitude & Range

43-46

9.1 - Individual Stage Burn duration

43

9.2 - Individual Mass distances

43-44

9.3 - Arc Length formula Altitude & Horizontal range

45-47

9.4 - Launch Log

47

Chapter 10: Orbital Transfer


10.1 - Hohmann Transfer Orbit
Chapter 11: Conclusion

48-52
48-52
53-54

11.1 Summary of Findings

53

11.2 Summary

53-54

11.3 Evaluation

54

11.4 Recommendations

54

Chapter 12: References, Bibliography & Appendices

55-78

12.1 References

55-56

12.2 Bibliography

57

12.3 Appendices

58-78

Table of Figures
Glossary

3-4

1.1 - The History of Rockets

79
80-82

Chapter 1: Introduction
1.1 - The History of Rockets
The basic principles of rockets span back thousands of years ago to 400 BCE, and all begin
with the writings of Aulus Gellius, a Roman who told a story about a Greek man called
Archytas. Archytas designed a wooden flying pigeon that ejected steam and adopted the later
stated action-reaction principle, causing it to moving along the wires it was suspended by.
After 300 years, another Greek man named Hero adopted this steam invention, creating a
steam powered devise which gave thrust to a spherical object causing it to rotate, which he
called, a Aeolipile. Without intentionally designing anything remotely symbolizing a rocket,
these 2 men opened the door to propulsive systems, and ultimately what we base rockets off
today. Soon after, their work had been discovered around Asia and Europe, and over the next
2,300 years the world became fascinated with the idea of static objects reaching impressive
velocities due to propulsion (Taylor, 2009).
In 1903 a Russian high school teacher by the name Tsiolkovsky published a book called The
Exploration of Cosmic Space by Means of Reaction Devices, which later became known as
the first true book on rocket science. The book detailed most of the ideas behind rocket
science today, and through his life he published more papers about rockets and especially the
idea of multi-staging them. Over the next few decades the real first successful rocket flights
occurred, which were fueled heavily by World War 1 & 2, and gave birth to what is famously
know as the Soviet and America Space race (Taylor, 2009).
On October the 4th 1957, Russia successfully launched Sputnik, which was the first manmade
object to reach space and successfully enter an orbit for 3 months before burning up in the
atmosphere. This fete forced the Americans to concentrate their efforts more on rockets and
ultimately gave birth to the most successful space agency to date, the National Aeronautics
and Space Administration or NASA. From here on NASA Successfully sent satellites to the
Moon and back, produced the first rocket to ever reach escape velocity, and successfully
placed a human into a LEO (Low Earth Orbit). However, it was on July 16 1969 that is
arguably the most important date in space exploration, the date of launch for the most famous
3-Stage rocket in history, the Saturn V. 4 days after Saturn Vs launch, Neil Armstrong and
his colleagues had been successfully escorted to the Moon and back on the Apollo 11
Mission, which became to be known as the greatest human fete in history (Taylor, 2009).
Over the next 40 years rocket exploration drastically decelerated due to enormous economic
strains put on NASA and FSA (Russians equivalent), and this provided a strong reason to

begin research into optimizing rockets to improve their efficiency. Which now brings us to
modern day rockets, which are now capable of reaching velocities and altitudes that exceed
all previous expectations, with improved efficiency, and are able further aid mankind in the
exploration of our universe (Taylor, 2009).
1.2 - The use of rockets
The idea behind modern rockets today is to place a chosen payload in outer-space/orbit, but to
do so efficiently, effectively and quickly. However due to Gravity, anything travelling
vertically experiences huge velocity losses, and therefore needs to accelerate as quickly as
possible to reduce this and reached orbital velcoity. Entering an orbit around a gravitational
body cancels out any negative effect on the velocity and provides ample time to calculate any
additional maneuvers. Therefore, rockets essentially create a controlled explosion in the
combustion camber and expel accelerating gas out, pushing itself forward and towards its
destination (Taylor, 2009).

Chapter 2: Aim and Assumptions


2.1 - Aims and Objectives
This project aims to explore the use of rockets in delivering a desired payload to a LEO (Low
Earth Orbit) in a realistic scenario, and the concept of orbital transfers.
The project shall aim to achieve this by calculating the final velocity needed for a successful
orbit at 100!" altitude, and work through the use of a single stage, 2-stage, and optimized 2stage rocket in reaching it. The progression throughout each model will also be shown, and
the estimated amount of velocity lost due to external forces will be investigated. Additionally,
the use of Lagrange Multipliers shall be explored in order to calculate an optimum 3-stage
final rocket mass in terms of the payload for the chosen velocity.
Furthermore, using the 3-stage rockets mass and some assumptions about the flight path, the
minimum requirements needed to understand the rockets journey into the LEO will be
investigated, in hope to build a simplified Launch Log specifying the rockets status at each
stage.
Finally, once the payload has successfully entered a LEO, the necessary requirements needed
for an orbital transfer, to reposition the payload to the same orbit as the International Space
Station, will be proposed.
2.2 - Assumptions
The following assumptions will be made about the paper, and only changed if mentioned:

The rocket is operating in an atmospheric and gravity free environment to begin with,
were external forces will be added if/when necessary.

The rockets thrust will always be at its maximum, and this can be reached
instantaneous. The rocket requires no fire time before lift off.

It will also be assumed that all the propellant flow rates for all the models are kept
constant for their entire burn time, and for all stages regardless of size, even if the
model is multi-staged.

The separating of stages for a multi-stage rocket happens instantaneously after the
previous stage has burnt out, and the next stage instantly produces the same exhaust
velocity as the previous one.

The rocket will have exhausted all its fuel at the point it reaches the LEO altitude.

All answers will be rounded to 4 decimal places

Chapter 3: Single Stage rocket Motion


3.1 - General solution for velocity at any time t without external forces
As stated in the assumptions, the rocket is operating in an atmospheric and gravity free
environment, so the effect of external forces sum to 0. Therefore the only forces that would
affect a rocket would be ones created by it.
If the rockets mass moved with velocity , the rockets momentum would equal,
= = .

1.1

and it would remain this as the momentum is conserved.


After some time the bodys change in momentum can be described by:
= +
or
= !"#$%& + !"!!"#$

1.2

But after , the rockets momentum would have changed as it would of expelled and
gained + .
!"#$%& = . + .
Now to calculate the !"!!"#$ , the difference between the rockets velocity , and the
effective exhaust velocity speed , is required, as figure 1 demonstrates:

Figure 1: Single stage rocket motion

The gas expelling from the rocket would be perceived to have a velocity of to an
observer, therefore this will need to be added onto !"#$%& .
However, due to Conservation of momentum, both momentums !"#$%& and !"!!"#$
should equal one another or subsequently cancel each other out.
!"#$%& = !"!!"#$

1.3

or
!!!"#$ !"!!"#$ = = 0

1.4

(Notice how the !"!!"#$ in ! 1.2 has changed to a sign, but after all this would be
expected as the exhaust gas is travelling in the opposite direction to the rocket.)
At , the exhaust would have expelled , so the momentum of the exhaust !"!!"#$
would actually be:
!"!!"#$ =
Subbing these into our momentum ! 1.4 yields:
0 = . + .
=
Taking the limit as 0, will give the equation the ability to calculate a velocity for any
time .

1.5

As the velocity and mass are with respect to time, the equation can be integrated between the
boundaries 0, , to find the velocity of the rocket at any time:
!
!

! =

!
!

1.6

!
!

= !

!
!

Evaluating between the boundaries yields:


! ! = ! + !

1.7

At = 0 the rockets initial velocity ! = 0, therefore after further simplifying, (1.7) then
becomes:
! =

!
!

1.8

(Khin et al, 2008)


This now is the final ! for describing the velocity at any time for a single stage rocket.
Where ! denotes the complete initial rockets mass including payload ! + , and !
denotes the instantaneous mass of the rocket.

Figure 2: Illustrates the velocity in !! of a Single stage rocket with a ! = 101,000!" ,


effective exhaust velocity = 3!"!!! , propellant flow rate of 1,500!"!!! and burn time
=

!"#
!

(Smith, R.C., Smith, P., 1990). (Please see Appendix 1 for Maple code)

As seen in Figure 2, the velocity of the rocket beings to accelerate more quickly towards the
end of its flight, which is due to the rocket loosing mass from exhausting its fuel, but still
producing the same thrust.
3.2 - General solution for velocity at any time t: Including external forces
Now looking at the case where the rocket is inside the Earths atmosphere and is being
directly affected by the Earth's gravitational pull, to see the effect of external forces on the
single stage model.
Taking the external force ! from Khin et al (2008) and adding on !"# for aerodynamic
Drag, the equation becomes:
Note: Both Gravity and Drag have signs as both apply their forces in the opposite
direction to motion.
(Its assumed that any force created by Lift is 0 for simplicity.)

=
!"#

1.9

Re-writing,
=

!"#

1.10

and integrating 1.10 just like (1.5) was, between the same boundaries 0,
!

=
!

!"#

1.11

will yield.

!
!

= !

!
!

!
!

!"#
!
! !

1.12

As the rate of change ! ! of the rockets mass is just the propellant flow rate , it can be
substituted in to get:
! !!

! !!

!
!

!"#

1.13

Now as the initial velocity of the rocket and gravitational force are both 0, the final ! for
the velocity now including external factors can become:

! =

!
!

!"#
!

1.14

Therefore subbing in !"# = ! ! which is the equation for drag force, and using
!

conditions typical of maximum drag.


Where the air mass density is = 0.25!"!!! ! , = 0.7!"!!! , cross section area =
10! !! , and the drag coefficient is ! = 0.2. The velocity lost if the rocket has to endure
maximum drag for the entire burn time can be found. (Taylor, 2009)
Hence:
!"# = 3.8385
Thus, for a complete final mass including payload ! of 101,000!" , propellant flow rate
of 1,500 !! , gravitational acceleration to Earth = 0.0098!"!!! , and burn time of
=

!"#
!

seconds, the effect of gravity and drag on the rockets final velocity are found to beK

= 0.5227!"!!!
!"#
!

3.4885
21,000

= 0.0040!"!!!
1,500
101,000

Thus thesingle stage rocket will loose 0.5267!"!!! of final velocity due to external forces

10

Figure 3: Demonstrates the velocity in !! of a Single stage rocket with external forces
over time, using the same rocket values as before. (Please see Appendix 2 for Maple code)
Now looking at Figure 3, its clear to see the effect of eternal forces on the rockets final speed
over time, as the velocity accelerates slightly slower than that in Figure 2. However in real
life, the amount of drag also increases with a rockets acceleration, thus a graph showing a real
rockets velocity graph may not accelerate as quickly.
Additionally,
0.0040!"!!!
= 0.0077
0.5227!"!!!
drag is only about 0.77% of the effect of gravity on the final velocity.
However it has been assumed that the velocity for the drag component is a constant to help
simplify the equation, but because of this, the calculated value above is actually smaller than
what should be expected realistically. Nevertheless, this value would still be very small
compared to gravity in the end (Gravity Loss, 2008 & Gatech, 2012).
3.3 - Distance Traveled: Single stage displacement
Integrating the velocity equation earlier derived for time, will lead to an equation for the
distance-traveled for any time.
This will be done for ! 1.14 as it incorporates external forces, which will of course lead
to a more realistic ! . However as previously identified, the effect of drag on the rockets

11

final velocity is fairly small compared to gravity, hence it will be neglected the distance
equation as the solution will be greatly simplified, for the cost of a very small error.
Therefore ! 1.14 without Drag is:
! =

!

!

1.15

Now, with the propellant mass flow rate a constant the instantaneous mass ! at any time
can then be described with:
! = !

1.16
(Smith, R.C., Smith, P., 1990)

Which essentially states, the mass at any time ! is the complete initial rocket mass ! ,
take away the duration the propellant mass is constantly flowing for.
Subbing ! 1.16 for ! in the velocity ! , and applying the integrals leads to:

!

!

!

!

!
!

1.17

1
+ !
!
2

1.18

Please see Appendix 3


Where is our constant from integration.
Now subbing in

1
!
!
Please see Appendix 4

The ! becomes,

12

!
!

!
1
+ !
!
2

1.19

and after integrating the inside of the bracket, the formula can become:

!
!

!
!

1
+ !
2

1.20

Please see Appendix 5


Using the knowledge that at = 0, = 0.

!
!

1.21

Therefore subbing ! 1.21 into 1.20 and simplifying yields:

= +

1
!
!

!
2

1.22

(Smith, R.C., Smith, P., 1990)


Please see Appendix 6
Which leaves an equation for the distance traveled, and if all the variables and a value for
time are known, they can all be subbed into this function to give a displacement value for
the rocket at that time.

13

Figure 4: Demonstrates the distance traveled in kilometers for a Single stage rocket over time
with a ! = 101,000!" , effective exhaust velocity = 3!"!!! propellant flow rate of
1,500!"!!! , gravitational acceleration to Earth = 0.0098!"!!! and burn time =
!"#
!

(Please see Appendix 7 for Maple code)

Similarly to Figures 1 & 2, as the rocket exhausts more fuel over time but produces the same
thrust, the distance covered over time beings to accelerate despite the increasing loss from
gravity. This however portrays the benefit of exhausting fuel quickly, as the increase in
acceleration can greatly outweigh the effects of gravity.
However, it will be more useful to derive an equation that calculates the altitude reached at
propellant burnout.
This then of course means the will become ! = , at propellant burn out.

= ! +

! !

1
! !
! !

!
2

1.23

Now slightly modifying the formula for the burn duration from Smith and Smith (1990),

14

! =

1.24

where ! is the rockets mass, is the fuel ratio, and is the propellant flow rate, it can be
subbed in to calculate a distance at propellant burn out.
Subbing this for ! in 1.23 leads to

=
! + ( 1 ! + ) 1

1+
!

1 !

2

1.25

Please see Appendix 8

and setting =

!
!!

which is also known as the Payload Fraction, ! 1.25 can become:

! + ( 1 ! + ) 1

1+

1 !

2

1.26

This is the final distance equation, and calculates the final propelled distance the rocket will
reach, while under the effects of gravity.
Using: ! = 100,000!" = 1,500!" = 3!"!! = 0.8 = 0.0098!"!!!
= 1000!" =

!
!,!""

!
!""

80,000 + (20,000 + 1,000) 1

!.!
!
!!
!""

!",!!! !

!,!""

1
80,000 32,982.5597 0.0049
1,000

6,400
9

= 94.0349 13.9378 = 80.0971!"


It can be seen that the rocket travels 80.0971!" until all the fuel is exhausted, but is of course
the distance traveled while being propelled, and does not include the coasting stage.

15

3.4 - Final Velocity at Propellant burn out


Subbing the ! for the time of propellant burn out 1.24 into the velocity equation earlier
derived, will generate the rockets velocity at propellant burn out.
However this will be done for the equation neglecting external forces as it will greatly
simplify the solution.

Taking ! 1.8 , ! =

!!

and subbing in ! 1.16 for ! yields:

!!

! =

!
!

1.27

Now replacing ! 1.24 with and simplifying, leads to

!"#$%"& = 1

1.28

1+

!
Please see Appendix 9

and again subbing =

!
!!

yields:

!"#$%"& = 1

1+

1.29

! 1.29 is the equation for the velocity at propellant burnout, depending only on, the
effective exhaust velocity , the fuel ratio and the payload fraction .
Again subbing some values in, the speed at propellant burn out can be generated.
Using: = 3!"!! = 0.8 =

!
!""

!"#$%"& = 4.7118!"!!!
(Atkins, 2013)
Therefore, after all the fuel has been exhausted the rocket will be travelling at 4.7118!"!!!

16

3.5 - Newtonian Orbital Velocity for a circular orbit


Now that the final velocity for the rocket is known, it would be sensible to calculate the
orbital velocity required to orbit the earth at 100!" .
Taken from Newtons principles of physics, the centripetal acceleration of a body about a
center of gravity is:
=

1.30

where is the acceleration of the body towards the center of gravity, is the constant speed
traveled along the orbits circumference , and is the radius from the orbiting body to the
center of gravity.
Taking Newtons 2nd law of motion, where is the bodys acceleration, and is the bodys
mass,
=

1.31

! 1.30 can be substituted in for


!
=

1.32
(Rybczyk, 2009)

to get the Centripetal Force equation.


Then taking Newtons law of universal gravitation,

1.33

setting ! 1.32 & 1.33 equal to each other, and making the subject, the equation
becomes:
=

1.34

Where is known as the universal gravitational constant, and represents the mass of the
body gravitational body, which in our case both apply for Earth.
Assuming both & stay constant, can be subtitled in for both of them:

17

1.35
(Rocket & Space Technology, 2013)

now denotes what is know as the Standard Gravitational parameter of Earth, which is
approximately 398,600!"!!! .
Additionally, as the desired radius is known ( !"#$! + desired altitude), it can be
substituted in to work out the required velocity .

398,600!"!!!
= 7.8490!"!!!
6,370!"# + 100!"#

Therefore, it is evident that once the rocket has exhausted all its fuel, it must have a velocity
of 7.8490!"!!! in order to stay in a orbit at 100!"# . If a greater velocity is reached, the
rocket could very easily be sling-shotted out of orbit completely, and if a smaller velocity
were achieved, the rocket would face plummeting back to earth. So ensuring the desired
velocity is actually met is essential.
But it was previously concluded that the single stage model has a final velocity of
4.7118!"!!! , which is no way near the velocity required.
S in this case the rocket could be given a greater mass in hope to produce a greater velocity,
but instead the concept of staged rockets will be now explored.

18

Chapter 4: 2-Stage rocket


4.1 - 2-Stage rocket velocity
To further increase the final velocity, the concept of staging the rocket will be explored,
where the rockets mass is simply split into 2 equal stages.
So,
! = ! + !
where:
! = 1 +
! = 2 +
This will of course lead to 2 separate velocities for each different stage, and the new final
velocity will be the sum of these 2 values, like so:
! = ! + !

2.0

Because theyre staged, both stages will now require a new payload fraction , to take into
account the separate payloads each stage must carry.

!"#$%! =

! +
!

2.1

! 2.1 is the payload fraction that must be subbed into the 1st stage equation, as when
stage 1 is active, ! + is essentially going to be the actual payload burden for it,

!"#$%! =

2.2

and ! 2.2 is the payload fraction for the 2nd stage. Notice how its very similar to the
original for a single stage rocket, but this should be expect this as once the 2nd stage is
active; it is effectively back to a single stage rocket just with respect to ! .
Now taking ! 2.0 and subbing in ! 1.29 along with the appropriate for each
stage, the final velocity of a 2 stage rocket can be worked out.

! = 1

1
1 + !"#$%!
1 + !"#$%!

2.3

19

Subbing in the equation for the different s will lead to:

! = 1

!
!
1
! + ! +
! +

2.4

Now as it was earlier assumed that each stage was equal, ! = ! , can be evenly
distrusted over both stages in terms of the payload . Before =

!
!""

!
!

, but it shall now

become:
=

=
=
50 ! !
! !
=
=
50
50

Subbing everything now known into ! 2.4 and simplifying will yield:

! = 1

2+

1
50

2.5

1+

1
50

Finally, using the same values as the single stage rocket the final velocity of the 2-stage
rocket is found to be:
= 3!"!! = 0.8
! = 1.5127!"!!! + 4.6018!"!!! = 6.1145!"!!!
(Atkins, 2013)
Its clear that a greater final velocity was reached without actually adding any additional fuel
but by just staging the rocket, as the mass containing the exhausted fuel was discarded and
therefore led to a lighter and more efficient rocket overall.

20

Figure 5: Demonstrates the velocity in !! of the 2-stage rocket with a


! = 101,000!" , effective exhaust velocity = 3!"!!! propellant flow rate of
1,500!"!!! and burn time =

!"#
!

. (Please see Appendix 10 Maple code)

Figure 5 shows a substantial increase in the rockets final velocity by just being staged, and is
due to the 2nd stage weighing considerably less than the 1st. This is because the structure from
the 1st stage has been separated, as it no longer holds any fuel. Thus allowing the 2nd stage to
accelerate much more quickly, and reach a greater velocity.
Now the rocket can be further staged, which will lead to an even greater velocity, however
before that is done there is another optimization technique that can be carried out.

21

Chapter 5: Optimized 2-Stage Rocket


5.1 - Optimized 2-stage rocket velocity:
The 2-Stage rocket can be further improved by optimizing the value of one mass with respect
to the other, but this will of course mean a new assumption in term of the 2 masses.
Therefore assuming that ! ! , and instead
! = ! !

2.6

it can be substituted into ! 2.4 and simplify to get:

! = 1

! !
! +

!
! +

2.7

This is the new formula for ! , in terms of just the 1 variable ! , as , , ! & are all
constants. This then means it can differentiate with respect to ! , to produce a maximum
value for it when ! is at its maximum.
Therefore, differentiating ! 2.7 and setting it = 0.


! !!
!!!

!
!!

! !! !!!
!! !!

!
!! !!

!
=0
!

!!

!!!
!! !!

! !! !! !!!!
!! !! !

2.8

Simplifying the above ! leads to:


1 ! ! + 2! ! = 0

2.9
Please see Appendix 11

But realistically 0, and thus the equation can be further simplified to,
! ! + 2! ! = 0

2.10

where the Quadratic formula can then be applied to yield the solution.

! = + ! + !

2.11

22

(The solution has been neglected as it leads to a ! , which of course cannot


realistically happen.)
Now dividing the ! by ! will put it in terms of ,
!
= + ! +
!

!
!

2.12

where ! 2.6 can again be used to get a function for ! :

!
!
=1
!
!

2.13

Subbing ! 2.12 into ! 2.13 and factorizing out 1 +


!
= 1+
!

!
!

1+

!
!

!
!

will lead to,

2.14

and to get the desired ! & ! ratio, ! 2.14 must be divided by 2.12 .
Which then yields
!
=
!

1+

2.15

Please see Appendix 12


Subbing in =

!
!""

, the mass ratio is found to be:


!
= 10.0499
!

2.16

or
! = 10.0499!

2.17

Therefore, the mass of the 1st stage must be 10.0499 times greater than the 2nd stage for it to
give an optimized final velocity.
Now subbing ! 2.17 into 2.4 ,

! = 1

10.0499!
!
1
10.0499! + ! +
! +

2.18

and simplifying
= 1

10.0499

1
11.0499 +
1+

2.19

23

will lead to the final equation for the optimized 2-stage rocket.
Again taking = 3!"!! = 0.8 and =

!
!!

!
!.!"##

! = 3.8232 + 3.8232 = 7.6463!"!!!


(Atkins, 2013)
Thus, after propellant burn out the optimized 2-stage rocket will have a final velocity of
7.6463!"!!! , which is a greatly increased final velocity, and one that nearly meets the
required orbital velocity of 7.8490!"!!! .

Figure 6: Exhibits the velocity in !! of the optimized 2-Stage rocket over time, using the
exact same variables as before. (Please see Appendix 13 for Maple code)
The optimized 2-Stage model in Figure 6 shows a ever greater velocity increase than just the
2-Stage, and is the result of calculating the optimum mass for the 2nd stage, where it acquires
a greater amount of , in a even shorter time by having a optimized mass.

24

5.1 - Comparison of all the model rocket velocities

Black Single stage


Blue - Single stage with
external forces
Red 2-stage rocket
Green Optimized 2stage rocket

Figure 7: Displays a comparison of the velocities for our single stage, single stage with
external forces, 2-stage and optimized 2-stage rocket over time. (Please see Appendix 14 for
Maple code)
Looking at Figure 7 its clear to see how the rockets velocity over time has improvement by
simply staging or optimizing it without the need to change variables. Additionally, the effects
of external forces are clearly evident and it justifies their importance for being included when
calculating a rockets velocity in a real life situation.
However, despite reaching a final velocity just short of the required orbital speed, the
optimized 2-stage rocket would fail stay in an orbit and risk falling back to earth. Therefore to
ensure the payload is guaranteed to successfully enter a LEO, the final optimization method
of Lagrange Multipliers will now be explored

25

Chapter 6: 3-Stage Rocket Mass


6.1 - 3 Stage optimized Rocket equation
To achieve the new final velocity of 10 !! , a 3-stage rocket optimized by the method of
Lagrange Multipliers will be used, as it allows a desired final velocity to be chosen and
optimize for.
Thus a new 3-stage rocket equation is required.
Using a re-written version of ! 1.28 from the Single stage rocket at propellant burnout,

= 1

!
! +

which can again be re-written as:


=

! + !

! +
! +
! + !

3.0

Now like before, must be split into the necessary amount of equations with respect to the
appropriate masses, where ! = ! + ! + ! for the 3 different stage masses.

! =

! + ! + ! +
! + ! + ! + !

! + ! + ! +
(1 )! + ! + ! +

! =

3.1

! + ! +
! + ! + !

! + ! +
(1 )! + ! +

3.2

26

! =

! +
! + !

! +
(1 )! +

3.3

Therefore, the final velocity will be the summation of the 3 velocities


! = ! + ! + !

! =

!! !!! ! !! !!

!!! !! !!! ! !! !!

!! ! !! !!
!!! !! ! !! !!

!! !!
!!! !! !!

3.4

Now setting,
! + ! + ! +
1 ! + ! + ! +

3.5

! + ! +
1 ! + ! +

3.6

! +
1 ! +

3.7

! = ! + ! + !

3.8

! =

! =

! =

! 3.4 can be simplified down to,

where ! 3.8 will later become the Constraint Equation.


However moving on and setting,
=

! +
1 ! +

3.9

where ! 3.9 is now the general in terms of the whole rockets mass, it can be simplified
and re-written down to:
! +

1 1

3.10
Please see Appendix 15

And now expanding the RHS in terms of ! , ! & ! will give:


! +
!
=

1 ! 1

!
1 ! 1

!
1 ! 1

3.11

27

This is the equation that now requires minimizing, but due to its complexity it will be best to
take the natural logarithm of both sides, and minimize that function as it greatly simplifies
the workings.
This can be done because the minimum of the original function has the exact same minimum
of the function with in place, so minimizing both will result in the same soltuion.
Therefore ! 3.11 can become,

! +
!
=

1 ! 1

!
1 ! 1

!
1 ! 1

and using the Laws of Logarithms that = + this can be changed to:

!! !!
!

!"!
!!!! !!!

!"!
!!!! !!!

!"!

3.12

!!!! !!!

This is now the stage to implement the optimization method and construct the Lagrange
function.
6.2 - The Method of Lagrange Multipliers:
Firstly a Lagrange function must be deduced, which holds the form:
, , , = , , + , ,
, , - This is what required maximizing. The Objective function
, , - And this is the constraint, which will be kept a constant.
- Is the actual Lagrange Multiplier.
(Harvard, 2009)
Hence,
, , =

!"!
!!!! !!!

!"!
!!!! !!!

, , = ! = ! + ! + !

!"!
!!!! !!!

3.13

3.14

and so forth the Lagrange equation is:

28

!
1 ! 1

!
1 ! 1

!
1 ! 1

3.15

+ ! + ! + !
This is now the new function that required minimizing, which can be begun by partially
differentiating with respect to ! , ! & ! .

Evaluating

!
!!!

!
=

! !
1 ! 1

!
!

3.16

!
=

! !
1 ! 1

!
!

3.17

!
=

! !
1 ! 1

!
!

3.18

leads to,
! =

1 +
1

3.19
Please see Appendix 16

and applying the same method for ! & ! also leads to the exact same answer, which proves
that ! = ! = ! .
Thus each ! , ! & ! can be replaced with just :

1 +
1

3.20

Furthermore, using this newfound knowledge for , ! 3.14 can re-written as,
! = 3

3.21

and rearranged for to make:


=

!!
!!

3.22

Now using the value = 3!"!!! like before, and setting ! to be the final velocity desired by
propellant burnout of the 3rd stage, a value for can be derived.
However, as external forces have been neglected from the equations, to ensure the rocket
definitely acquires the necessary velocity, the masses required for a final speed of 10 !!

29

shall be calculated. As up to 2 !! can be lost in velocity from gravity and drag when
getting a payload into orbit (Otaski, 2012).
= 3!!!!! , ! = 10
!"

= ! = 3.0377
Furthermore,
Taking ! = =

!! !!
!!! !! !!

and making ! the subject will lead to,


! =

3.23

1
1 1

3.24

where,
=

Please see Appendix 17


and if the same method is applied to ! & ! , it will lead to the following answers.
! = + 1

3.25

! = + 1 !

3.26
Please see Appendix 18

Thus, the optimized mass ratios and masses can now be calculated.
Dividing ! by ! ,
! + 1
=
=+1
!

subbing ! 3.24 in for ,


=

1
+1
1 1

and simplifying,
=

1
1 1
1
+
=
1 1
1 1
1 1

yields the first mass ratio:


!

=
! 1 1
Evaluating

!!
!!

3.27

also results in,


! + 1 !
=
=+1
!
+1

which has already been calculated to be.


!

=
! 1 1

30

Therefore it can be concluded that the general optimum mass ratio for any 2 stages can be
described by the following formula:
!

=
!!! 1 1

3.28

Using the values = 3.0377 & = 0.8, the value for the mass ratios are found to be:

! !
=
= 6.1921
! !
! = 6.1921!
! = 6.1921!
! = 38.3421!

Thus, the mass of stage 1 needs to be 6.1921 times greater than the mass of stage 2, and the
mass of stage 2 also needs to be 6.1921 times greater than stage 3. This then means, stage 1
need to be 38.3421 times greater than stage 3, in order to produce a optimized final velocity
of 10 !!
Furthermore, using the 3 ! s

(3.23, 3.25 & 3.26) for ! , ! & ! , the equation

describing the overall complete rocket mass ! is established to be:


! = ! + ! + !
! = + 1 ! + + 1 +
1 + + 1 + + 1

! = ! + 3 ! + 3

3.29

Additionally, due to the fixed ! and , is then also a constant, allowing to be calculated.
For

= 3.0377, = 0.8 and ! 3.24 , calculates to be:


= 5.1921

Therefore, using ! s (3.23, 3.25 & 3.26), the sizes of the individual masses are:

31

! = 5.1921
! = 32.1500
! = 199.0760
Thus,
! = 236.4181
And the final Mass of the whole rocket just before takeoff is then equal to,
! = ! + 236.4181 + = 237.4181
Hence, the rocket would need an initial mass of 236.4181 to get a payload to its required
final velocity of 7.8490 !! (assuming external forces affected the ! by 2.151 !! ),
and into an orbit at 100!" above the Earths surface.

Figure 8: Illustrates the change in final rockets final mass ! is terms of P, between the old
and new final desired velocity . (Please see Appendix 19 for Maple code)
Looking at figure 8 its clear to see how the rockets final mass ! exponentially increases as
the desired final velocity goes up. (assuming the effective exhaust velocity remains the
same) However this would be expected, as is to the power

!"
!

in ! 3.22 , and this

increasing will cause to exponentially increase, which will in turn greatly increase our
rockets final mass.

32

Chapter 7: 3-Stage Evaluation


7.1 - Comparing model rocket to Saturn V:
Now that an equation for the rockets final mass ! has been derived, comparing the model to
another 3-stage rocket will indicate how realistic it is.
Hence, comparing the model to the Saturn V, which was also a 3-stage rocket, and was
capable of delivering an 118,000!" payload to a LEO orbit while only weighing
2,800,000!" (Nasa, 2014).
Putting the following data in the rocket mass ! s (3.22, 3.24, 3.25 & 3.28) its clear that
from the following values,
! = 5.1921 118,000!" = 612,667.8!"
! = 32.15 118,000!" = 3,793,700!"
! = 199.076 118,000!" = 23,490,968!"
! = 236.4181 118,000!" = 27,897,335.8!"

in order to get an 118,000!" payload to a final velocity of 7.8490 !! , the 3-stage rocket
would require an initial mass ! of 27,897,335.8!!
Furthermore, when comparing the final mass of the models ! to Saturn Vs actual value,
27,897,335.8!"
!
=
= 9.9633
!
2,800,000!"
its apparent that that model requires an initial mass nearly 10 times greater than the mass of
the Saturn V rocket.

33

7.2 - Improved overall Mass:


Seeing as the rocket designed is very inefficient, the case when the model uses a new
improved fuel ratio, and effective exhaust velocity shall be explored to see how much the
rockets mass improves.
Substituting in = 3.5574!"!!! for the effective exhaust velocity, and = 0.9 for the fuel
ratio, a new set of mass values can be calculated (Taylor, 2009 & Gatech, 2012).
Using ! 3.21
= !

!"
!.!!"#

= 2.5523

and ! 3.24 , the new becomes:


= 2.0843
Therefore, again using ! s (3.22, 3.23 & 3.25) the size of new the individual masses are:
! = 2.0843
! = 6.4286
! = 19.8278


! = 28.3407
So for = 118,000 like before, the rocket would now have a final mass of ! =
3,344,202.6!" , which is a lot closer to the mass of the Saturn V.
So it can be concluded that that mass equations are indeed realistic, and the new effective
exhaust velocity and fuel ratio lead to a more efficient and realistic rocket than the previous
one.
Therefore, for the remainder of the paper these new values for the effective exhaust velocity
and fuel ratio shall only be used.

34

Chapter 8: 3-Stage Flight path & Distance


Now that the overall rocket mass equation has been verified, the other variables required to
place the payload into a LEO can be calculated.
8.1 Flight path
Assuming the rocket has a flight path described by,
=

!
,
400

0 200

and reaches an altitude of 100 at a horizontal distance of 200, its path would look like
Figure 9:
(Launch Sequence Details, 2011)

Figure 9: Shows the 3-Stage rockets flight path where = altitude, and = horizontal
distance. (Please see Appendix 20 for Maple code)
Thus, to calculate the distance the rocket will travel once it reaches an altitude of 100, the
Arc Length formula in Cartesian form must be used:

35

!"#$%! =

4.0

1+

(WCAS, 2004)

Substituting in the path equation and its conditions, the above function will lead to:
!""

1+ 1
200

=
!

4.1

This is now where the trigonometric substitution

must be made for 1

!
!""

to yield:
!""

4.2

1 + !

Likewise, the must also be written in terms of and subbed in.

1
=

200

= 200 !
Therefore, ! 4.2 becomes:
!""

1 +

200 !

!
!""

= 200

4.3

Finally, calculating the new boundaries in terms of


=01

!
!""

=1

= 1 =

= 200 1 1 = 0

!
!

= 0 = 0

and subbing them in will return the function that must be integrated in order calculate the
flight paths length.
!

= 200

!
!

4.4

Therefore integrating the above equation leads to,


= 200

1
+ +
2

4.5

!
!

(Paul, 2014)
which is then evaluated between its boundaries and simplified to yield:

36

= 100

2 + 1 + 2

= 229.5587!"
Thus, once the rocket has travelled exactly 229.5587!" , it must be travelling at
7.8490 !! (assuming velocity lost due to external forces), which also means all the fuel
needs to be exhausted at this point as well.
8.2 Evaluating the Distance equation
With this value for the distance travelled, can be set equal to = 229.5587!" , and the
distance equation from chapter 3 can be evaluated.
Recall ! 1.26 :

=
! + ( 1 ! + ) 1

1+

1 !

2

As the value for all the variables in the above equation apart from are known, the function
can be manipulate to calculate a suitable value for , that will cause the rocket to exhaust all
its fuel at = 229.5587!" . Therefore, rearranging ! 1.26 for leads to

! =

! + ( 1 ! + ) 1

1+

and making the substitution =


=

!
!!

!
!

! + ( 1 ! + ) 1

1
!
2
!
!!!

&

will lead to the simplified version:


! + = 0

Using the Quadratic formula:

!! ! ! !!!"
!!

! =

4.6

, both values fore function can be calculated:

! 4.1.
2

+ ! 4
! 4
& ! =
2
2

37

Now to get values for the propellant mass flows, values need to be substituted into each
function.
This is the point where the rest of the calculations are done specifically for a single type of
rocket, and its specifications only.
Thus the propellant flow rates required to place a payload of 1000 in orbit, shall be
calculated.
Using the values,
= 3.5574!"!!! ! = 28.3407 = 28,340.7!" = 0.9

= 229.5587!" =

!
!!

!
!".!"#$

= 0.0098!"!!!

& calculate to be
= 274.3351 & = 13,887.0017
! = 207.3962!"!!! & ! = 66.9783!"!!!
Now evaluating the distance equation with these calculated ! & ! values should both lead
to burn out distance desired:
!! = 303.6729!" 74.1140!" = 229.5587!"
!! = 940.3136!" 710.6144!" = 229.6992!"
As both propellant flow rates lead to answers to the same distance value earlier calculated
using the Arc Length formula, (despite the small rounding errors) it can be fair to assume they
will realistically allow the rocket to meet its objective without any problems.
Note: The RHS of both equations show the effect of gravity on the rockets distance, and
interestingly how the effect of gravity increases as the burn time goes up due to a decreasing
. This justifies the need to reach orbital velocity as soon as possible when factoring in
gravity, as the faster orbital velocity is reached; the less fuel is then wasted.

38

Using both ! & ! , the rockets burn time for the separate propellant flow rates can be
worked out by using ! 1.24 :
!!!! = 122.9850!"#$%&!

!!!! = 380.8193!"#$%&!

Therefore, after 122.9850 380.8193 seconds, the rocket would of exhausted all of its fuel
at an altitude of 100!" , and will have reached its desired final velocity.

Advantage

Advantage

Leads to the shorter and more ideal burn time

Leads to the longer and more realistic burn

as the fuel is exhausted at a faster rate, which

time, and allows the rocket to be made with

reduces the amount of ! / that is

much lighter and weaker materials, as the

lost to the increasing gravity component.

rocket accelerates slower, putting it under a


lot less mechanical stress.
The longer burn time will give the rocket
more time to complete any maneuvers like a
gravity turn. (Taylor, 2009)

Disadvantage

Disadvantage

Exhausting the fuel too quickly can put the

The longer the time spent propelling, will

rocket under immense mechanical stress,

result in more distance lost due to gravity,

which can break parts, alter the rockets

and even though both lead to a distance

trajectory or put the astronauts under to much

of 229.5587 , ! lost 710.6144!" due to

G-force.

gravity.

(Taylor, 2009)

39

Figure 10: Presents the quadratic function 1.26 for ! & ! . The values for that lead to
the solutions above the vertical axis provide distances less that 229.5587!" , while the values
underneath the vertical axis lead to answers greater than 229.5587!" . (Please see Appendix
21 for Maple code)
Note: Looking at figure 10, its clear to see ! & ! as they are the roots when the vertical
axis equals 0, and thus lead to a propellant burnout at = 229.5587!" . However as the
function clearly has a local minimum, an optimum value for that would then lead to a even
greater distance can be calculated.
Differentiating the quadratic function for , setting it = 0 and then solving for leads to a
of 137.1776!"!!! , a burn time of 185.9388!"#$%&! , and subsequently a of 289.7088!" .
This new value for would now be the greatest distance the rocket could achieve with the
same specifications, but with an optimum . This would yield very useful in a scenario
where the greatest distance was desired without altering anything else apart from , which
in real life is very easy to change. But as the objective/flight path is still the same, the
previous are needed, and proceeding with the of 137.1776!"!!! , will not be
necessary.
Please see Appendix 22 for Maple calculations

40

Figure 11: Illustrates the 3-Stage rockets Figure 12: Illustrates the 3-Stage rockets
distance against time for ! (Please see distance against time for ! . (Please see
Appendix 23 for Maple code)

Appendix 24 for Maple code)

Furthermore as Figure 12 highlights, the acceleration, and subsequently thrust created by !


is so small that the rocket fails to lift-off for around 200 seconds, and must exhaust some fuel
to loose weight before it can. This is of course very unrealistic and inefficient, and as ! s
performance as seen in Figure 11 is a lot more realistic and effective, ! shall be neglected
while ! is proceeded with.
However ! is fairly small compared to real rocket propellant flow rates, but when used on
this model, the rocket is capable of reaching a velocity of 10!"!!! in just over 2 minutes,
when in real life it would around 8 minutes (Nasa, 2007).
The reason why this small ! is able to produce such a high velocity is most likely due to the
fact that the rocket at hand is very efficient, it has a fuel ratio of 0.9, and effective exhaust
velocity of 3.5574 !! . Meaning the rocket is made from a strong, lightweight material
that can hold 9 times its own weight, and the high-energy fuel is capable of producing
extremely fast gas velocities. It was also concluded in chapter 6 that these values are
potentially similar to those used to take the Saturn V to the moon, which will certainly
possess the ability to take this rocket into a Low Earth Orbit (LEO) with a small .
In addition, most rockets do not require orbital velocity by the time theyve reached their
desired orbit, however from the assumptions thats exactly what this rocket is designed to do.

41

This has undoubtedly increased the with respect to the rockets size and has produced a
large , which explains why the burn time is so short.
Finally, as external forces have not incorporated in the equations and instead the decision was
made to aim for a greater velocity, the only thing slowing its acceleration down is its mass,
which is also rapidly decreasing. Hence, its probably fair to assume if the effects of external
forces where incorporated in the velocity equations, it would lead a greater burn time.
Now that the nature of ! is better understood, the altitude, range, horizontal distance and
burn time for the individual stages can now be established.

42

Chapter 9: Individual Distance, Altitude & Range


9.1 - Individual Stage Burn duration
Given a propellant flow rate ! = 207.3962!"!!! , = 0.9, and using a modified version of
! 1.24 from before, but altering it for a function of any individual stage,

!! =

4.7

the individual stage burn durations can be calculated, as displayed in Table 1.


Table 1- Individual stage burn times
!

! = 19,827.8!"

!!

86.0431!"#$%&!

! = 6,428.6!"

!!

27.8970!"#$%&!

! = 2,084.3!"

!!

9.0449!"#$%&!

9.2 - Individual Mass distances


Now that a set of values for how long each stage was active for has been produced, they can
be used in the distance equation again to calculate roughly how much each stage traveled.
But before this is done, the distance equation in terms of ! needs to be formulated.
Substituting in ! + ! for ! in ! 1.15 leads to

! =

! + !

!

where denotes the additional distributed amount of payload the individual stage must now
carry.
Integrating the above equation like how ! 1.15 was for !! , will this time lead to,

! = !! +

! + ! !!

1
! + ! !!
!!

! + !
2

4.7

(Please see Chapter 3.3- Distance Traveled: Single stage displacement for method)

43

and substituting in ! !! for !! in the gravity component, will correct the additional effect
of gravity that would have then been be lost from the above equation.

! = !! +

! + ! !!

1
! + ! !!
! !!

! + !
2

4.8

Please see Appendix 25 for a small explanation


Finally, subbing in ! 1.24 for ! and simplifying will lead to

! =
! + ( 1 ! + ! ) 1

1+ !
!

1

2

4.9

! !

To calculate a suitable value for ! , the rockets optimized masses can be used to distribute P
as a percentage accordingly to them, as done in Table 2.
Table 2 Distribution of !
Stage mass in

Percentage of !

! Value

Stage 1

19,827.8

69.9623%

0.6996

Stage 2

6,428.6

22.6833%

0.2268

Stage 3

2,084.3

7.3544%

0.0735

28,340.7

100%

0.9999

Subbing in all of the rockets values = 3.5574!"!!! = 207.3962!"!!! = 0.9


= 1000!" = 0.0098!"!!! plus the appropriate ! & ! values, the distance values for
the individual rocket stages are calculated to be:
Table 3 Distance traveled for each stage
Stage

Distance Travelled

Stage 1 - !

! = 160.605!"

Stage 2 - !

! = 52.0179!"

Stage 3 - !

! = 16.8836!"

Overall Distance

229.5065!"
Please see Appendix 26 for calculations

44

9.3 - Arc Length formula Altitude & Horizontal range


As the distance each stage travels is known, they can be used in conjunction with the Arc
Length formula to calculate the altitude and horizontal range for each stage, by calculating the
boundaries at each stage for the flight path.

Blue Distance 1
c

Red Distance 2
Green Distance 3

Figure 13: Shows the distance travelled by the individual stages along the flight path, where
the x-coordinate boundaries are designated by a,b and c accordingly. (Please see Appendix 27
for Maple code)
For stage 1 the boundary = 0 is already known, therefore only the other boundary a is
required, as shown on figure 13. Once the flight path equation is integrated between 0 and a,
the answer should then equal the value ! = 160.605!" .
Thus working through the equation like before but with the new boundaries, equating it to ! ,
and then working backwards should provide an estimate for a.
Starting from ! 4.5 , only the new boundaries need to be calculated:
=01
= 1

!
!""
!
!""

=1

= 1 =

= = tan!!

Subbing these into to ! (4.5) yields,

45

1
= 200 + +
2

!"#!! !
!
!

As is equal to ! , it can be subbed in and the equation can then be simplified to get:
1.6061 = + +

!"#!! !
!
!

4.10

Evaluating the above equation between its boundaries and then using the substitutions.
tan!! =

! + 1 & tan (tan!! ) = 1

will lead to:


1.6061 = 2 + 1 + 2

! + 1 . +

! + 1 +

which subsequently leads to the approximated solution:


0.3384
(Please see Appendix 28 for Maple calculations)
Subbing this back in for the boundary condition allows a value to be calculated for :

1
= 0.3384
200
= 132.32!"
This is the horizontal distance (x value) for the 1st stage, and subbing it back into the original
flight path ! will produce the altitude at that stage:
132.32!" !
= 132.32!"
= 88.5485!"
400
Therefore, at 1st stage propellant burnout the rocket will have travelled 132.32!"
horizontally, and 88.5485!" vertically.
Furthermore, as the starting boundary for stage 2 is now known, the same method be can
applied to calculate the horizontal distance and altitude for the 2nd stage, and subsequently the
3rd stage.
Calculating these values will lead to the answers presented in Table 4

46

Table 4 Individual stage horizontal/distance values


1st Stage

132.32!"

88.5485!"

2nd Stage

183.1!"

99.2860!"

3rd Stage

199.96!" 200!"

100!"

Stage

Please see Appendix 29 for full calculations


9.4 - Launch Log
With all the information now calculated, a basic launch log detailing the information of the
time, horizontal distance, altitude, flight path distance and speed at the end of each stage can
be constructed.
()

()

()

()

86.0431

132.32

88.5485

160.605

3.3333

113.940

183.1

99.2860

212.6229

6.6666

122.985

199.96 200

100

229.5065

9.9999

122.985

200

100

229.5587

10

+
*Neglecting rockets initial height.

Its clear that due to the huge mass, stage 1 requires the most time/distance in order to reach its
required velocity, and that the later stages require less as they produce the same thrust but
have a reduced mass. This is however expected, as in chapter 6 is was found that is the
same for all the stages, meaning despite being heavier or not each stage must produce the
same velocity before separation.
Additionally, the 3rd stage travels mostly horizontally and only slightly vertically, which is
due to the chosen flight path slowly converging towards 100km.

47

Chapter 10: Orbital Transfer


10.1 - Hohmann Transfer Orbit
Assuming the payload is orbiting the Earth at 100!" , and the next stage of the mission is to
deliver the payload to the International Space Station, which orbits at an altitude of
approximately 370!" . (Hutchinson, 2013). The payload must make an orbital maneuver to
re-position itself, and the most basic and effective for simple maneuvers, is called a Hohmann
Transfer Orbit.
A Hohmann transfer orbit involves a body applying a certain amount of to change its
circular orbit, into an elliptical orbit where the apogee (the furthest point in the orbit to the
focus point) matches the altitude desired for the final orbit. Once the body orbits half way
around the new elliptical orbit, and has now reached the desired altitude, more will then
be required to circularize the orbit, increasing the perigee (the closest point in the orbit to the
focus point) so its equal the apogee, and ultimately completing the orbital maneuver (Orbital
& Celestial Mechanics, 2014).
As Figure 14 shows, assuming the body starts
in perfect circular orbit of radius (),
the Hohmann orbit procedure can be broken up
into 3 stages:
1. When the orbiting body is at the
perigee, apply to increase the
apogee, so it matches ! , creating the
elliptical orbit
2. The body will then travel along one
half of the new elliptical orbit, where it
will reach the orbits apogee
Figure 14: Demonstrates the 3 stages

3. Once the body has reached the apogee,

within a Hohmann transfer

will then need to be applied to

(Hohmann Transfer Orbit, 2013)

circularize the orbit, and finally


complete the Hohmann orbital transfer.

48

Hohmann Orbit Equations:


Assuming the payload is capable of producing a thrust, then the necessary amount of the
payload must exert to be repositioned must be calculated. Therefore, the velocities for the
following 4 orbits need to be found:

Velocity of current orbit

Velocity of desired orbit

Velocity of first transfer orbit (first half of elliptical orbit)

Velocity of second transfer orbit (second half of elliptical orbit)

Furthermore, Figure 14 also highlights how the payload must gain both required tangent
to the orbit, in the direction of motion. So its assumed that the payload will automatically
angle itself.
To calculate the 4 ! the initial radius ! , the final radius ! , and the semi-major axis are
required,
! = !"#$! + 100!" = 6,370!" + 100!" = 6,470!"
! = !"#$! + 370!" = 6,370!" + 370!" = 6,740!"

= =

! + ! 6,470!" + 6,740!"
=
= 6,605!"
2
2

along with the velocity for the current circular orbit,

5.0

!! =

5.1

!! =

and the velocity for the final orbit:

= 398,600!"!!!

49

The Transfer Velocities (elliptical orbits) are:


Perigee Velocity
! =

2
2

=
! ! + !

! =

5.2

2 1

Apogee Velocity
2
2

=
! ! + !

5.3

2 1

(Aerospace Engineering and Engineering Mechanics, N/A)


Now using the values for ! , ! & , the velocities for

5.0, 5.1, 5.2 & 5.3 are calculated to

be:
!! =

398,600!"!!!
= 7.8490!"!!!
6,470!"

!! =

398,600!"!!!
= 7.6902!"!!!
6,740!"

! =

398,600!"!!!

2
1

= 7.9288!"!!!
6,470!" 6,605!"

! =

398,600!"!!!

2
1

= 7.6112!"!!!
6,740!" 6,605!"

Following on, the payload needs to be put into an elliptic orbit with an altitude of ! , but the
additional velocity needed to move the orbits apogee to the altitude of ! when the payload is
at the orbits perigee, is the difference between them both. This however can be described
mathematically by:
! = ! !!

5.4

Furthermore, the velocity needed to circularize the elliptical orbit requires increasing the
orbits perigee equal to the radius of the apogee, thus circularizing the orbit. This once again is
the difference between both, and can be described by:
! = !! !

5.5
(Widnall & Peraire, 2008)

50

Therefore, using the calculated velocity values


! = 7.9288 7.8490 = 0.0798!"!!!
! = 7.6902 7.6112 = 0.079!"!!!
and the total the payload must exert is the summation of these 2, which is
! = ! + ! = 0.1588!"!!!

Figure 15: Illustrates the amount of required to completed a Hohmann transfer orbit,
starting from radius of 100km and a desired radius between 200km - 36,000km (geostationary
orbit). (Please see Appendix 30 for Maple code)
Interestingly, as the desired radius increases the acceleration of the required ! begins to
decrease; however when looking at equations 5.0, 5.1, 5.2 & 5.3 the reason for this becomes
clearer. As ! , and consequently increase, equations 5.0 & 5.2 either slightly increase or
stay the same, while both 5.1 & 5.3 greatly decrease. Therefore, ! & ! will continue to
decrease at a growing rate, and will in turn result in a decelerating ! .
Finally, to ensure the 2 are fired at the correct times, it must be assumed that when the
payload has completed at least 1 full rotation and is back to the start of its orbit, it will

51

exhaust enough fuel to gain ! . Furthermore to ensure ! is gained correctly, the time it
takes for the payload to orbit around to the apogee must also be considered.

Hence, using the following equation that calculates the time it takes to travel around an
elliptical orbit once,
!"#$% =

4 !

= 2

(Widnall & Peraire, 2008)

the semi-major axis can be substituted in to find the orbit duration:


2

6,605!
398,600

= 5,342.2038!"#$%&! 44.5184!"#$%&'
However, the payload only needs to complete half an orbit before it must fire up again to gain
! .
Therefore, after !"#$% = 2,671.1019 !"#$%&! = 22.2592!"#$%&' the payload must fire
again, and do so towards the direction of motion/tangent to the orbit.
Note: if the payload desired to return to its original orbit, the same and !"#$% would be
needed, but the payload must fire them in the opposite direction this time (Widnall & Peraire,
2008).

52

Chapter 11: Conclusion


11.1 Summary of Findings

7.8490!"!!!

4.7118!"!!!

4.7678!"!!!

6.1145!"!!!

7.6463!"!!!

10!"!!!

19,827.8!"

6,428.6!"

2,084.3!"

28,340.7!"


()

86.0431

27.8970

9.0449

207.3962!"!!!
122.985

160.605!"

52.0179!"

16.8836!"

229.5587!"

132.32!"

50.78!"

16.9!"

200!"

88.5485!"

10.7375!"

0.714!"

100!"

3.3333!"!!!

3.3333!"!!!

3.3333!"!!!

10!"!!!


()

()

0.0798!"!!!

0.079!"!!!

0.1588!"!!!

2,671.1019

11.2 - Summary
Its evident that the single stage rocket lacked the capability to reach the desired orbital speed;
and that the method of staging rockets is required to reach such velocities. Therefore, by
splitting the rocket in half the final results were successfully increased, but the produced
results where still smaller than the required velocity. Therefore rather than split the rocket
again hoping to further increase on this, the decision was made to optimize the rocket stages.
Nevertheless the new results were still short of the desired velocity, and called upon the use
of a 3-stage rocket and a completely new approach, the method Lagrange Multipliers. It was
earlier established that gravity and drag impacted heavily on a rockets final speed, and
consequently led to a decision to aim for a greater final velocity in hope to produce more
realistic results. With a final rockets mass, a comparison was made between the Lagrange
rocket model and the Saturn V, where some parameters were found to be ineffective/outdated
and led to re-calculating the rockets mass but with new parameters. This verified that the
model and new parameters where indeed realistic, along with provide a new improved rocket

53

mass. This improved value, along with the flight path meant the overall propelled distance
could be calculated using the Arc length formula, and provided two propellant flow rates ,
where only one was found to be reliable. Proceeding with the more realistic the overall
burn time was established, and combining the overall distance and burn time, an equation
describing the individual stage distances was formulated, along with their individual burn
times. Going back to the Arc length formula with this information allowed the horizontal
distance and altitude to be estimated, and with all this information a basic launch log detailing
the rockets status at each stages separation was constructed.
With the rocket now in a stable orbit, the necessary additional velocitys required to deliver
the payload to the International Space Station was established, along with the required time
delay between maneuvers.
11.3 - Evaluation
I believe the models derived give a fairly realistic representation of how the velocities for real
rockets are calculated, and the step-by-step approach undertaken helps to portray how rockets
have evolved with the use of mathematics. Furthermore, having put my payload in a scenario
that is very typical of actual space missions that take place today, and pushing for additional
realism wherever possible, my answers are more realistic than I had hoped to have achieved
when initially undertaking a project of this difficulty. Finally, the launch log in chapter nine
represent a simple version of the actual launch logs used by space agencies before a rocket
takeoff, and demonstrates how only a few models can be used to produce a sufficient amount
of data.
11.4 - Recommendations
If I had to do this project again Id do the following things differently:
1. Consider incorporating gravity and drag in my multi-stage rocket models, in order to
produce more realistic answers
2. Look at the case were the rocket is required to deliver a satellite (payload) to
geostationary orbit.
3. Construct a new equation that better describes the flight path of a rocket more
realistically
4.

Consider orbital plane changes, and interplanetary trajectories.

54

Chapter 12: References, Bibliography & Appendices


12.1 References
Aerospace Engineering and Engineering Mechanics (N/A) Hohmann Transfers [Online]. Available at
<http://design.ae.utexas.edu/mission_planning/mission_resources/orbital_mechanics/Back_of_the_Env
elope_Bklt.pdf> [Accessed 21 March 2014]
Akins, N. (2013) Concepts of Mathematics [Online]. Available at <www.studyspace.kingston.ac.uk>
[Accessed 10 February 2013]
Gatech (2012) Introduction to Rocketry [Online]. Available at
<http://soliton.ae.gatech.edu/people/ejohnson/ae1350-Spring2012/14.rockets.pdf> [Accessed 21 March
2014]
Gravity Loss (2008) Drag: Lost in Ascent, Gained in Decent, and What It Means for Scalability
[Online]. Available at <http://gravityloss.wordpress.com/2008/01/10/drag-loss-in-ascent-gain-indescent-and-what-it-means-for-scalability> [Accessed 21 March 2014]
Harvard (2009) Lagrange Multipliers [Online]. Available at
<http://www.math.harvard.edu/archive/21a_spring_09/PDF/11-08-Lagrange-Multipliers.pdf>
[Accessed 21 March 2013]
Hohmann Transfer Orbit, (2013). [Online Image], Available at
<http://culv3r.files.wordpress.com/2013/05/hohmann_transfer_orbit.png> [Accessed 12 March 2014]
Hutchinson, L. (2013) How NASA steers the International Space Station around junk. Science &
Exploration, [Online]. Available at <http://arstechnica.com/science/2013/07/how-nasa-steers-theinternational-space-station-around-space-junk/> [Accessed 21 March 2014]
Rybczyk, J. (2009) Relativistic Orbit Velocity [Online]. Available at
<http://www.mrelativity.net/relorbitalvelocity/relativistic%20orbital%20velocity.htm> [Accessed 21
March 2014]
Khin, Z. Myint, O, Aye, K. (2008) Mathematical Techniques in Rocket Motion. World Academy of
Science, Engineering and Technology [Online]. Available at
<http://waset.org/publications/8913/mathematical-techniques-in-rocket-motionnal> [Accessed 21
March 2014]

55

Launch Sequence Details (2011). [Online Image].


<http://mars.jpl.nasa.gov/msl/mission/timeline/launch/launchsequencedetails/> [Accessed 21 March
2014]
Mathcentre (2004) The Quotient Rule [Online]. Available at
<http://www.mash.dept.shef.ac.uk/Resources/web-thequotientrule.pdf> [Accessed 21 March 2014]
Nasa (2014) What Was the Saturn V? [Online]. Available from
<http://www.nasa.gov/audience/foreducators/rocketry/home/what-was-the-saturn-v58.html#.Uyx9Y3lq4dt> [Accessed 21 March 2014]
Nasa (2007) Nasa Direct: Ask the Mission Team [Online]. Available at
<http://www.nasa.gov/mission_pages/shuttle/shuttlemissions/sts121/launch/qa-leinbach.html>
[Accessed 21 March 2014]
Orbital and Celestial Mechanics (2014) The Hohmann Orbit Transfer [Online]. Available at
<http://www.cdeagle.com/omnum/pdf/hohmann.pdf> [Accessed 21 March 2014]
Otaski (2012) Gravity drag [Online]. Available at
<http://lcas.otaski.org/index.php/Rocket:First_approximations#Gravity_drag> [Accessed 21 March
2014]
Pauls Online Math Notes (2014) Arc Length [Online]. Available at
<http://tutorial.math.lamar.edu/Classes/CalcII/ArcLength.aspx> [Accessed 21 March 2014]
Rocket and Space Technology (2013) Orbital Mechanics [Online]. Available at
<http://www.braeunig.us/space/orbmech.htm#types> [Accessed 21 March 2014]
Smith, R.C. Smith, P. (1990) Mechanics. 2nd ed. West Sussex, John Wiley & Sons.
Taylor, T. (2009) Introduction to Rocket Science and Engineering. Boca Raton, Taylor & Francis
Group
WCAS (2004) Arc Length, Parametric Curves [Online]. Available at
<http://www.math.northwestern.edu/~mlerma/courses/math214-2-04f/notes/c2-arclength.pdf>
[Accessed 21 March 2014]
Widnall, S. Peraire, J. (2008) Orbit Transfers and Interplanetary Trajectories [Online]. Available at
<http://ocw.mit.edu/courses/aeronautics-and-astronautics/16-07-dynamics-fall-2009/lecturenotes/MIT16_07F09_Lec17.pdf> [Accessed 21 March 2014]

56

12.2 Bibliography
Mathematics in Motion
http://www.macs.hw.ac.uk/~bernd/F12MR2/momnotes08.pdf
Simple Rocket Science
http://psas.pdx.edu/simplerockets_1d/#index2h4
Rocket and Spacecraft Propulsion: Principles, Practice and New Developments
http://goo.gl/8P2F6z
The Atlas Family
http://www.braeunig.us/space/specs/atlas.htm
Lagrange Multipliers
http://www.cs.iastate.edu/~cs577/handouts/lagrange-multiplier.pdf
Space Launch Report: Kosmos M3 data sheet
http://www.spacelaunchreport.com/kosmos.html
Rocket Flight Simulation
http://www.thrustcurve.org/simulation.shtml
Arc Length of a Curve
http://www.mathwords.com/a/arc_length_of_a_curve.htm
Centripetal Force
http://www.lhup.edu/~dsimanek/scenario/labman1/centrip.htm
Burghes, D. Graham, A. (1980) Control and Optimal Control Theories with Applications. Chicester,
Horwood Publishing.

57

12.3 - Appendices
1:
Single Stage Velocity for t

>

2:
Velocity with external forces

>

3:
Staring with,
=

!

!

it can be solved using Integration by parts:

Setting = =

d =

!
!"

! ! !!"
!!

! ! !!"
!!

! !

58

! !
! ! !!"

plus the integration for the gravity component, the equation becomes :

!
!

1
+ !
!
2

4:
Where,
= ! !
both sides can be divided by ! :
!

=1
!
!
and therefore can be replaced by each other.
5:
Staring with
=

!
!

and utilizing the Laws of Logarithms that

!
!

! ! (!)
! (!)

!
1
+ !
!
2

= the inside integral will result in:

! !

1
+ !
2

Now to check, differentiate back:

! !

=1

!
!

59

Due to the extra , the RHS to the answer above must be divided by . This means the
solution to the integral is:
=

!
!

!
!

1
+ !
2

!
!

!
!

1
+ !
2

6:
Given

where is
=

!
!

it can be subbed into the earlier equation to become:

!
!

!
!

!
1
! !

Now simplifying the equation will lead to the required formula.

!
!
!
1
+ +
!
! !
!

!
!
!
1
+ +

!
!

!
2

= +

!
!
1

!

!
2

7:
Distance over time

60

8:
Subbing
! =

into
= ! +

! !

1
! !
! !

!
2

will produce,
!
!
!

!
1 !

=
+
!

!
2

this can however be simplified down to


!

!
1 !
+
! ! 1

!
2

!
! + ( ! ! ) 1

1 !

2

Recall: ! = ! +
=

!
! + (! + ! ) 1

1 !

2

!
! + ( 1 ! + ) 1

! +

1 !

2

! + ( 1 ! + ) 1

1+
!

1 !

2

9:
After subbing in for the equation becomes,

61

!"#$%"& =

!
!

which can be simplified down to


!"#$%"& =

!"#$%"& =

!"#$%"& =

!
! !

! +
! + !
! + !
! +

!"#$%"& = 1

!"#$%"& = 1

!
! +

1+

10:
2-Stage Rocket velocity over time:

>

62

11:
Starting with,
! = 1

! !
! +

and again using a Law of Logarithms that

!
=
!

1
! !
1
! +

!
!"

! ! (!)
! (!)

!
! +

, the above equation becomes:

1
! + !

!
! +
! + !
1
! +

The brackets can then be expanded to give,


!

=
+
!
! + ! !

! + !
! ! +
! + !
! +

and after following through with simplifications


=

+
! + ! !
! + ! + !

=
! + ! + !
! + ! !

= ! + ! !

= ! + ! + !

= ! + ! !

= ! + ! + !

= ! + ! ! + ! ! ! 2! + ! ! + ! !
= ! ! ! ! + 2! + ! ! 2!
= ! ! 1 + 2! 1 ! 1
the answer is found to be:
1 ! ! + 2! ! = 0

63

12:
Given both
!
!
=1
!
!
and
!
= + ! +
!

!
!

! can be divided by ! to yield:


!
=
!

1+

!
!

1+

!
!

+ ! +

!
!
!

This can be further simplified down to

1+
=

!
!

!
!

1+

!
!

! . ! + + 1 ! . !

1+

!
!

!
!

1+
!

! + + 1

1+

!
!
!

!
!

!
!
!

!
=
!

=1+

1+

13:
2-Stage optimised:

64

>

14:
Plotting single stage velocity without external forces against single with external forces, 2stage rocket, and the optimised 2-stage rocket:

>

15:
Starting from
=

! +
1 ! +

can be simplified to:


1 ! + = ! +
= ! + 1 !

!
1 !
+ 1

If 1 is taken away from both sides, and

!!
!

+ 1 is made the subject, it will eventually

become:

1 =

1 1

!
1 !
+ 1
1

!
+ 1 1

!
+1

65

!
1 1
+ 1 =

1 1
Re-writing and simplifying the above ! will yield the solution:
! +

1 1
16:
To differentiate

!
=

! !
1 ! 1

one of the Law of Logarithms that

!
!"

This highlights that the differential of

!"!
!!!! !!!

! ! (!)
! (!)

!
!

must be used.

is required, and to calculate this the Quotient

rule will be used.


Utilizing the Quotient rule formula

!
(Mathcentre, 2004)
the

!"!

!!! !!!! !!!

of the equation becomes,

! 1 !! 2! !! 1 2 1 !
!

! !! 1

2!! 1 2!! !! + 2 1 !!
!

! !! 1
!!
!! 1 1 !

1 1 !

and subbing this answer in while also differentiating the RHS will lead to:

66

1 1 !
=
!
!
1 ! 1

=
!
1 1 !

1 ! 1

+
!
!

1
=
+ 0
! 1 1 !
1 = 1 1 !
Which can be re-arranged into the required value for !

! =

1 +
1

17:
Starting with
=

! +
1 ! +

and making ! the subject will eventually lead to


1 ! + = ! +
! ! + = ! +
! ! ! =
! 1 =

! =

were setting =

!!!
!!! !!!

1
1 1

results in the simplified final mass value for !

67

! =
18:
Applying the same method in Appendix 17 to ! & !
Calculating ! :
! = =

!! ! !! !!
!!! !! ! !! !!

1 ! + ! + = ! + ! +
! ! + ! + = ! + ! +
! ! ! = ! ! +
! 1 = ! 1 + 1

! =

! 1 + 1
! 1 + 1

1
1 1

! =

1 ! +
! +
1 1

As ! = it can be subbed into the above solution to yield:


! = +
! = + 1
Calculating ! :
! = =

!! !!! ! !! !!
!!! !! !!! ! !! !!

1 ! + ! + ! + = ! + ! + ! +
! ! + ! + ! + = ! + ! + ! +
! ! ! + ! + ! + = ! ! + ! ! +

68

! 1 = ! 1 + ! 1 + 1

! =

!! !!! !!! !!! !! !!!


!!!"!!

! =

As

! =

!! !!! !!! !!! !! !!!


!!! !!!

1 ! + ! +
! + ! +
1 1

& ! = + 1 , they can both be subbed into the above equation to

give:
! = + 1 + +
! = ! + 2 + 1 = + 1 + 1
! = + 1 !
19:
Rocket Mass for different V in terms of P:

>

>

>

69

20:
Flight Path

>
21:
Propellant flow ratio:

>

22:
Optimized-Max :
Given the equation for
>

70

it can differentiate and set = 0, and solved to find the optimum .

With the new , the earlier burn time equation can be used to get a new burn duration.

Subbing the new and burn time into the distance equation, provides the maximum distance
the rocket can go:

Which is 289.7088!" .
23:
Distance travelled - for m1

>

71

24:
Distance travelled - for m2
>

>
25:
For the original distance equation, the gravity component was
1
!
2

This however can be re-writing is as


1
! !! + !! + !!
2
Which is the total burn time multiplied by the sum of the individual burn times, and after
expanding the brackets the equation becomes:
1
! !! + ! !! + ! !!
2
Therefore its clear to see the individual gravity component equations should be:
Stage 1 Gravity component

1
! !!
2

Stage 2 Gravity component


1
! !!
2
Stage 3 Gravity component

1
! !!
2

72

Putting in the appropriate values and totaling them up, they should add up the amount of
distance lost due to gravity, for the whole rockets distance, which was 74.1140!"
0.0098
122.9850 86.0431 + 27.8970 + 9.0449
2

= 74.1165

Which means
! !!
needs to be subbed in for !! to get the correct values, and if it isnt the solutions will become
incorrect.
26:
Distance traveled for 1st stage:

! =

3.5574
0.9 19,827.8 +
207.3962

1

2

0.1 19,827.8 + 699.6 1

0.9 28,340.7
207.3962

0.9
699.6
1+
19,827.8

0.9 19,827.8
207.3962

3.5574
1
17,845.02 5,458.793832 0.0098
207.3962
2

122.985 86.0431

= 212.4569 51.8519 = 160.605


Distance traveled for the 2nd stage:

! =

3.5574
0.9 6,428.6 +
207.3962

1

2

0.1 6,428.6 + 226.8 1

0.9 28,340.7
207.3962

0.9
226.8
1+
6,428.6

0.9 6,428.6
207.3962

73

3.5574
1
5,785.74 1,769.829 0.0098
207.3962
2

122.985 27.897

= 68.8836 16.8657 = 52.0179


Distance traveled for the 3rd stage:

! =

3.5574
0.9 2,084.3 +
207.3962

1

2

0.1 2,084.3 + 73.5 1

0.9 28,340.7
207.3962

0.9
73.5
1+
2,084.3

0.9 2,084.3
207.3962

3.5574
1
1,875.87 573.7799 0.0098
207.3962
2

122.985 9.0449

= 22.3343 5.4507 = 16.8836


27:
>

>

>

>
>

74

28:
Distance 1 Equation value:
Putting the formula into Maple

>

and solving it numerically yields

29:
Applying the same method used to calculate the end boundary for the 1st stage to then 2nd and
3rd stage, will also provide the approximated values for the horizontal distance and altitude.
Hence using the boundaries = 132.32 and ! = 52.0179!" , the value for for the 2nd
stage can be identified:

= 132.32 1
= 1

!
!""

!"#.!"
!""

= 0.3384

= 0.3384 = 0.3263

= = tan!!

And the ! will simplify down to

0.5202 = + +

!"#!! !
!.!"#!

>

Which leads to the approximated solution

>

75

The negative and imaginary values should all be ignored as they lead to incorrect/unrealistic
solutions, which leaves the 2nd value of 0.08454937 as the answer.
0.0845
Therefore
1

= 0.0845
200

= 183.1!"
Which will lead to a value of
= 99.2860!"
Applying the method again for the 3rd stage

= 181.3 1
= 1

!
!""

!"!.!
!""

= 0.0845

= 0.0845 = 0.0843

= = tan!!

And the ! will simplify down to

0.1688 = + +

!"#!! !
!.!"#$

>

Which leads to the solutions


>

Once again the first negative and the imaginary answers are ignored, taking just the real and
positive value as the solution
0.0002

76

So
1

= 0.0092
200

= 199.96!"
Here its clear to see the error from the approximated solution for as the furthest boundary is
known to be 200, nevertheless it still leads to the final value of
= 100!"
30:
Orbital manoeuvre:

77

>

78

Table of Figures
Figure 1 - Single stage rocket motion
Figure 2 - Single stage rocket velocity over time
Figure 3 - Single stage rocket velocity over time including external forces
Figure 4 - Single stage propelled distance traveled
Figure 5 2-Stage rocket velocity over time
Figure 6 Optimized 2-Stage rocket velocity over time
Figure 7 Velocity for single stage, single stage with external forces, 2-stage and optimized
2-stage rocket over time
Figure 8 Change in final rocket mass against old-new velocity
Figure 9 3-Stage rocket flight path
Figure 10 Quadratic function for the propellant flow rate
Figure 11 3-Stage rocket distance for !
Figure 12 3-Stage rocket distance for !
Figure 13 Individual distances for the 3-Stage rocket
Figure 14 Hohmann Transfer Orbit
Figure 15 Required velocity for Hohmann transfer.

79

Glossary
If a variable/term is used in multiple Chapters, it will only be mentioned in the first section it
appears in.
Chapter 3
LEO Low Earth Orbit (100 1,000)
Momentum
Mass
Velocity
Change in the bodys Momentum
!"#$%& Change in Rockets Momentum
!"!!"#$ Change in exhaust Momentum
Change in Mass
Change in Velocity
Effective Exhaust Velocity
! Mass at time
! Velocity at time
! Initial Rocket takeoff Mass, including the payload
! Initial Rocket Velocity at time 0
Time
Payload
Propellant /fuel flow rate
Gravitational acceleration to Earth at sea level
!"# Aerodynamic drag
Cross-sectional area
! Drag coefficient
Air mass density
Distance traveled
! Initial Rocket Mass (excluding payload )
! Burn duration for propellant burn out
Amount of fuel in the rocket (Fuel ratio)
Payload fraction
!"#$%"& Velocity at propellant burn out
Centripetal acceleration of a body towards the center of gravity
Radius from body
Force

80

Universal Gravitational constant


Mass of gravitational body
Standard Gravitational parameter of Earth 398,600!"!!!
!"#$! Radius of Earth
Chapter 4
! Mass of the 1st stage (fuel + structure)
! Mas of the 2nd stage (fuel + structure)
! Change in Velocity for the 2-stage rocket
! Change in Velocity for the 1st stage of a staged rocket
! Change in Velocity for the 2nd stage of a staged rocket
!"#$%! Payload fraction for the 1st stage of a staged rocket
!"#$%! Payload fraction for the 2nd stage of a staged rocket
Chapter 6
! Mass of the 3rd stage (fuel + structure)
! Change in Velocity for the 3rd stage of a stages rocket
! Final Velocity for the multi-stage rocket
! Initial and burn out mass fraction for the 1st stage of a staged rocket
! Initial and burn out mass fraction for the 2nd stage of a staged rocket
! Initial and burn out mass fraction for the 3rd stage of a staged rocket
Initial and burn out mass fraction for the complete rocket
, , , = , , + , , Lagrange equation format
, , Objective function
, , Constraint function
Lagrange multiplier
Rocket mass to structure ratio
! Mass of stage
!!! Mass of stage + 1
Chapter 8
The arcs length (function length)
Rocket component over distance
Gravity component
!! Distance using propellant flow rate !
!! Distance using propellant flow rate !

81

!!!! Burn duration for propellant burn out using !


!!!! Burn duration for propellant burn out using !
Chapter 9
!! Burn duration for propellant burn out for stage
!! Burn duration for propellant burn out for stage 1
!! Burn duration for propellant burn out for stage 2
!! Burn duration for propellant burn out for stage 3
! Percentage of payload distributed to stage
! 1st stage distance traveled
! 2nd stage distance traveled
! 3rd stage distance traveled
Chapter 10
Apogee Furthest point in the orbit to the focus point
Perigee Closest point in the orbit to the focus point
! Initial/current orbit radius
! Final/desired orbit radius
Semi-major axis
!! Velocity of current orbit
!! Velocity of desired orbit
! Perigee transfer velocity
! Apogee transfer velocity
! Change in Perigee transfer velocity
! Change in Apogee transfer velocity
! Total required velocity
!"#$% Time duration for an elliptical orbit

82

You might also like