You are on page 1of 14

The UCLA Rocket Project

1
The Development of a Paraffin Based Experimental Hybrid
Sounding Rocket
Kurt U. Zimmerman
1
, Brian C. Kentosh
2
, Peter C. Chang
3
, Mackenzie J. Booth
4
, Richard E. Abrantes
5
, Phuoc Hai N. Tran
6
,
Craig P. McGrath
6
, Brandon A. Dizon
7
, Adel M. Shalabi
7
, William A. Silva
7
, Matthew J. Marshall
7

University of California, Los Angeles, CA, 90095
The UCLA Rocket Project entered the third year of design and testing of its custom hybrid rocket engine
which will be used as the propulsion system for UCLAs entry into the 6
th
Intercollegiate Rocket Engineering
Competition put on by the Experimental Sounding Rocket Association of Green River, Utah. The hybrid
engine combines paraffin wax with 5 micron aluminum powder as its fuel and liquid nitrous oxide as its
oxidizer. It was designed by student members of the UCLA Rocket Project and has a specific impulse of 202
seconds, a thrust of 1004 pound force, and a burn time of 9.9 seconds. In conjunction with the 15 foot long
carbon fiber and fiberglass air frame designed and fabricated by the Rocket Project, the hybrid engine is
capable of propelling a 10 pound payload to 25,000 feet above ground level. An on board avionics package
comprised of off the shelf accelerometers and microcontrollers is capable of accurately predicting the rockets
apogee and initiating an engine shut down in order to reach exactly 25,000 feet. Students in the Rocket
Project also designed, developed and constructed all of the electrical interfaces for the various sensors and
electronic valves used in both testing and launch of the experimental rocket. For recovery and reusability, the
rocket employs a dual deployment drogue and main parachute system, which allows the rocket to land safely
within 6 minutes of reaching apogee.
Nomenclature
a = regression rate coefficient l = port length t = time
a
k
= acceleration L = nose cone length T
0
= stagnation temperature
A
cs
= cross section area m = mass T
n
= temperature bulk air
A
inj
= area injector orifice = mass flow v = velocity
C
D
= coefficient of drag M = mach number x = distance from tip
C
P
= specific heat constant pressure n = regression rate exponent = shock angle
c
*
= characteristic velocity P = pressure = density
F
D

= drag force r = radius
g = gravitational constant = change in radius
k = flow loss coefficient R = base radius
I. Introduction
HE UCLA Rocket Project has continued its three year project to design and fabricate its custom experimental hybrid
rocket engine, known as the Hybrid Propulsion Experiment (HyPE). The goal of this development has been to compete
in the Experimental Sounding Rocket Associations (ESRA) Intercollegiate Rocket Engineering Competition (IREC). The
objective of the advanced competition in the 6
th
IREC includes carrying a 10 pound payload to 25,000 feet above ground
level (AGL) and returning the rocket and payload in reusable condition sans expendables such as propellant. The UCLA
Rocket Project started working on the HyPE during the 2008-2009 year with six active members participating in the project.

1
President, UCLA Rocket Project
2
Propulsion Lead, UCLA Rocket Project
3
Electronics Lead, UCLA Rocket Project
4
Aerodynamics and Structures Lead, UCLA Rocket Project
5
Recovery Manager, UCLA Rocket Project
6
Propulsion Team Member, UCLA Rocket Project
7
Electronics Team Member, UCLA Rocket Project
T

The UCLA Rocket Project


2

Figure 2. Carbon fiber
tube lay-up process


Figure 1. Quarter Cutaway View of
UCLAs Entrant into the 6
th
IREC


Figure 3. Coefficient of Drag vs Mach Number Plot Generated
using RASAero
5

Today there are over 30 active members with diverse backgrounds from aerospace and mechanical engineers, to electrical
engineers, material scientists, and computer scientists.
In addition to completion of the design, fabrication and testing of the HyPE engine, the UCLA Rocket Project also
designed and fabricated its own carbon fiber and fiberglass airframe, designed the dual deployment recovery system, all of
the electronic controls and interfaces, the test equipment and custom avionics package capable of predicting the apogee of the
rocket in flight for accurate and dynamic engine shut down capability. All of these projects were undertaken with the intent of
competing in and winning the advanced category of the 6
th
annual IREC competition.
II. Structural and Aerodynamic Systems
A. Airframe
The primary purpose of the airframe is to provide a rigid structure
through which the thrust can be transferred, to house the various
subsystems while inducing minimal drag, and to allow easy access to
components within the airframe. Accessibility of internal components is
important for integration in the field, so the airframe diameter was
increased to 8 in. The increased diameter is also beneficial in shortening
the overall rocket length to 15ft, giving an aspect ratio of 22.5, which
eliminates the extreme bending moment seen in UCLAs entrant to the 5
th

IREC, the RATTworks rocket.
There were three key requirements that were considered in choosing a
material for the airframe: strength, weight, and cost. In order to achieve
both high strength and low weight, composite materials were deemed the
best option. After conducting extensive
research on available composites, a
unidirectional pre-impregnated (pre-preg) carbon fiber was chosen for the coupling tubes,
and a woven pre-preg was chosen for the body tubes. Not only do these materials have a
high strength to weight ratio, but they are relatively easy to work with. In addition, their
widespread use in the aerospace and marine industries made it possible to eliminate material
costs through in-kind material donations.
It is necessary to cure pre-preg composite parts in either an oven or autoclave. The
UCLA Rocket Projects oven size limited the lengths of tubes to 23 in. This necessitated the
use of coupling tubes, made of 8 layers, to link together outer body tubes, made of 3 layers.
The process of fabricating the carbon fiber tubes for the airframe was relatively simple,
but tedious. Prior to each lay-up, the carbon fiber was cut to size, and an aluminum mandrel
was wrapped in a release film to prevent the cured carbon fiber tube from adhering to the
mandrel. Layers of carbon fiber were then wrapped around the mandrel individually. A
layer of pressure tape was applied after every other layer and then subsequently removed
after ten minutes. The pressure tape provided compression on the fibers and prevented the
formation of creases in the final part by reducing the bulk factor of the material
1
. After the
final layer of carbon fiber was applied, a layer of
porous peel ply was added. This facilitated the
bleeding of excess resin out of the tube, which
decreased weight and increased strength. A final
layer of pressure tape was added, and then the
tube was allowed to cure within the oven for 2.25
hours at 275 degrees F.
B. Drag Analysis
Extensive research was conducted on
various types of drag to optimize the performance
of the rocket. Three primary forms of drag had to
be taken into account while designing the rocket:
skin friction drag, pressure drag and base drag
2,3,4
.
Skin friction drag is affected by the wetted area
and surface smoothness
2
, with the wetted area

The UCLA Rocket Project


3

Figure 4. Drag comparison for various nose
cone shapes at a fineness ratio of 3:1
2


being a design parameter that is determined by the nose cone shape and the fineness ratio of the nose cone. Pressure drag is
based upon the cross sectional area of the body and is relatively small at subsonic speeds. One of the specific types of
pressure drag is wave drag, which begins to affect objects moving at velocities greater than Mach 0.8. Base drag is caused by
the low pressure region created at the aft end of the rocket, and can have a substantial contribution to overall drag once the
engine has stopped thrusting
2
. In addition to these main three forms of drag, the rocket also experiences interference drag
around the camera fairing and where the fins are attached.
To determine the overall drag of the rocket, shown in Figure 3, an excel calculator was created, which took in
various geometric inputs and then plotted the different types of drag. This made it possible to quantify which types of drag
had the greatest effect. This calculator was used in conjunction with the free software RASAero
5
.
C. Nose cone
After studying the different nose cone shapes available, an LD Haack
shape was selected, as research showed that it outperformed most other
shapes in the transonic region, shown in Figure 4
2,4,6
. The LD Haack isnt
a simple geometric form; rather it is a mathematically derived shape that
has been optimized to reduce wave drag. The equation for an LD Haack
is:
sin 2
2
R
r
|
|
t
=
(1)
Where
( )
1
2
cos 1
x
L
|

= (2)
Where R is the base radius, L is the nose cone length and x is the distance
from the tip of the nose cone.
The optimum length ultimately came to be a compromise between skin friction drag and wave drag, as increasing the
length increased the skin friction drag, yet reduced the wave drag by reducing the curvature. Data from the both the in house
excel calculator and RASAero led to a fineness ratio of 4.5:1.
Aerodynamic heating was taken into consideration due to the rocket attaining supersonic velocities. The maximum
temperature that a fluid can be heated to near a moving body is the stagnation temperature
2,3,7
, where T
n
is the temperature of
the bulk air, v is the rockets velocity (1340 ft/s) and C
P
is the specific heat at constant pressure:

2
0
2
n
p
v
T T
C
= +
(3)
Equation (3) gave a maximum stagnation temperature of 144 F (335K). To alleviate problems with aerodynamic heating the
nose cone tip was slightly rounded to increase the surface area through which heat is absorbed.
Due to the presence of radio frequency transmitters in the nose cone, it was necessary to choose a material that did not
block radio waves. Furthermore, this material should be capable of withstanding temperatures up to 144 F. Fiberglass with a
room temperature curing resin fulfilled these requirements. Many of these resins have relatively low glass transition
temperatures, T
g
. However, after contacting several companies in the resin industry, a suitable mixture was found in 820
green laminating system resin from Adtech. In order to manufacture the nose cone, a male plug followed by a female
fiberglass mold had to be made before the actual nose cone could be fabricated. The plug, made of wood, polystyrene foam
and Bondo, was used to provide a shape around which the mold could be made. The mold was made into two halves for ease
of lay-up. The majority of the nose cone was fiberglass, however some carbon fiber was added in the tip and around the
shoulder to alleviate aerodynamic heating and provide additional strength.
D. Boattail
Boattails reduce base drag by reducing the cross sectional area at the base of the rocket, and thereby minimizing the low
pressure region. Boattails may also incur extra wave drag if curvature isnt kept to a minimum
2,8
. In theory, the absolute
minimum wave drag for a boattail is provided by a shape that has zero change in cross sectional area at the base; however for
ease of fabrication, a simple LD Haack was used
9
. This shape has better wave drag characteristics than a conical boattail,
however, and enabled it to be fabricated from the same mold that was used to make the nose cone.
E. Fins
Fins are an important element for ensuring that the rocket remains stable during flight by moving the center of pressure
behind the center of gravity. A clipped delta shape was chosen to maximize stability. The clipped delta is a shape that is easy

The UCLA Rocket Project


4

Figure 5. Dimensioned Fin


Figure 7. Stress Simulation Showing a
Factor of Safety Plot

Figure 8. Assembled HyPE 1B Engine.

Figure 6. Typical Supersonic
Airfoil Configurations
2




to manufacture, provides a good amount of rigidity to counter fin flutter, and has relatively low drag compared to several
other designs
2
.
The primary design factor for the fins was to provide a stability margin of 1.5 to 2.0 calipers, or reference diameters, for
both the wet and dry configurations of the rocket. The Barrowman Equations and the open source software OpenRocket
10

were used to determine the fin dimensions and initial estimate of the rockets center of pressure location relative to its center
of gravity. The RASAero software was utilized to determine a more comprehensive center of pressure location as a function
of Mach number
11
. With the aid of the above software, the optimum fin dimensions, shown in Figure 5, were determined to
be 12 inches for the root chord, 4 inches for the tip chord and 8 inches for the height. In order to ensure that the fins stay
within the mach cone, equation (4) was used, where M is the Mach number and is the shock angle
2
:

1
1
90 sin
M
| |
A =
|
\ .
(4)
With a maximum mach number of 1.25, a sweep angle of 36.9 is determined. For the
dry configuration, a stability margin of 2.31 calipers was achieved, while for the wet
configuration, the weight increases to 180.16 lbs with a stability margin of 1.86.
The second primary design criterion was to minimize drag and fin flutter. Fin flutter
is a major concern as it can be severe enough that enough stress can be put on the fins
that the material fails causing the fins to be violently ripped off of the rocket body.
Additionally fin flutter also creates a turbulent flow and increase drag. Several of the
factors that affect fin flutter are the dynamic pressure, Mach number, material stiffness
and fin mass
12,13
. To prevent fin flutter, the fins must be prevented from bending or
twisting. Smooth flow around the fin will act to
dampen out the oscillations
12,13
.
Four aluminum brackets were made for each fin, in order to keep them rigidly
attached. The fins were then bolted to the engine mount tube and bonded to both the
engine mount and the boattail. Fiberglass was added to the bonds to provide strength.
The fins were made of 20 layers of unidirectional pre-preg, alternating the layers
fiber direction by 90 for increased strength. To reduce drag, the fins used a modified
double wedge cross section (Figure 6), which is optimum for supersonic flights
2
. The
tip of the fin was left flat to prevent flow around the tip.
F. Launch Rail and Test Stand
A test stand was built in order to conduct static hot fire tests of the
experimental engine. Based on the projected weight of the rocket, it became
obvious that a reinforced launch rail would be necessary. A modular setup
was designed and built to accommodate both requirements and reduce
material cost.
After several design studies, a triangular cross section was chosen, where
two side by side 6ft long sections would make up the test stand setup and
four 6ft sections stacked end to end would become the launch rail. The
modularity of the system not only decreased material costs, but it also
improved transportability. The test stand version has a factor of safety
(FOS) of 3.88 based on a 1000 lb load (Figure 7) and in the launch rail configuration the system deflects 3.31 in (84mm)
based on a 100lb load from off axis thrust applied perpendicularly to the top of the rail, the worst case scenario. The system
was designed in SolidWorks and has shown an ability to handle the expected loads. The final launch rail/test stand was
constructed from ASTM A36 steel and assembled using a combination of welds and bolts.
III. Propulsion Systems
One of the primary focuses
of this project was the Hybrid
Propulsion Experiment 1B
(HyPE 1B) propulsion system.
The overall layout is shown in
Figure 8.
The HyPE 1B is a student-
designed hybrid rocket engine, using aluminized paraffin as the solid fuel and liquid nitrous oxide as the oxidizer. Regulated

The UCLA Rocket Project


5

Figure 9. Fuel grain port detail. The
minor rippling is believed to result from
slight misalignment during spin casting
operations.

nitrogen pressurizes the oxidizer tank, flowing liquid nitrous oxide into the combustion chamber where it combusts with the
paraffin. D-class hobby rocket motors serve as the ignition source. A graphite bell nozzle accelerates the hot gases and
produces thrust. Loads are carried from the combustion chamber to the oxidizer tank through a thrust structure and
subsequently transferred to the rocket body. All tanking is performed remotely, and in-flight engine cutoff is made possible
by an on-board pneumatic system. Table 1 provides a summary of the engines predicted performance characteristics.
Table 1. HyPE 1B Performance Summary
Thrust [lbf] Isp [s] C* [ft/s] O/F Ratio Burn Time [s] P
c
[psi] A
e
/A
t

1004 202 4,619 4.30 9.90 400 5.21
A. Previous Research
The previous year, the UCLA Rocket Project developed and test-fired the first HyPE engine, designated the HyPE 1A, to
represent that this was the Block I design of the HyPE engine. The aluminum combustion chamber of the HyPE 1A deformed
during its first and only hot fire test, which occurred after the 5
th
IREC in Green River, Utah. To mitigate this, the combustion
chamber was converted from in thick aluminum to in thick steel. This variant was designated the HyPE 1AT (HyPE 1,
Block I, Test Configuration). The HyPE 1AT allowed for research to continue while the upgraded HyPE 1B (Block II) was
being designed.
B. Propellant
The HyPE 1B fuel grain was based on research performed by Stanford
University, which showed that aluminized paraffin had a high regression
rate, allowing for high thrust via a single port design
14-30
. This, coupled with
moderately high performance, made paraffin an attractive option.
The final composition of the HyPE 1B fuel was 74% paraffin wax, 18%
aluminum powder, 6% Vybar 103, and 2% carbon black. The addition of
aluminum increases the regression rate and Isp of the engine
14
. Vybar and
carbon black were used to reduce sloughing, which is the expulsion of
unburned fuel out the nozzle. Vybar, commonly used as an additive in
candles, increases the melting temperature and helps slow the melting of the
fuel. Carbon black acts as an opacifier and absorbs thermal radiation,
preventing the heat from conducting past the surface layer.
To cast the fuel grain, the components were melted together in an electric
convection oven. The mixture was then poured into an ablative liner and
spun on a lathe until it cooled. This produces a grain with an even port down the central axis. Weights were taken before and
after casting to verify the mass of the fuel grain.
Characteristic velocity, c*, was determined through NASA Chemical Equilibrium with Applications (CEA). CEA did not
include data for paraffin, so this was input from data given by the Journal of Chemistry
31
. From this analysis the optimal O/F
ratio was found to be 4.30, and this value was used to size the fuel grain for optimal combustion.
C. Pressurant System
The HyPE 1B employs a pressure regulated system to keep the oxidizer flow constant for the burn duration. This
performs better than the simpler blow down system where the pressurant gas is unregulated, causing oxidizer pressure to drop
as the gas expands.
The HyPE 1B utilizes a L45M SCBA tank manufactured by Luxfer Gas Cylinders. Specifications of this tank are given in
Table 2
32
.
Table 2: Specifications of Luxfer L45M
32

Service Pressure Volume Diameter Length Weight
4500 psi 285 in
3
5.4 in 18.4 in 6.6 lb
The L45M was sized for the HyPE 1A and as such is slightly undersized for the HyPE 1B, however the performance gain
of upgrading the pressurant was not substantial enough to justify the cost. The pressurant tank is filled with gaseous nitrogen
from a standard K-size gas cylinder. The nominal fill pressure is 2250 psia, which is the standard pressure of the nitrogen gas
cylinder.
The pressurant flow is controlled by a Mighty Mite high flow regulator manufactured and generously sponsored by
Dresser. A more stable oxidizer pressure is achieved due to its high flow rate.
D. Oxidizer Tank

The UCLA Rocket Project


6

Figure 11. Ox Valve Actuation. Valves
are mounted to thrust rods (not
pictured for clarity)




Figure 10. HyPE 1B Oxidizer Tank.
Left hand side connects to pressurant
tank, right hand side connects to
combustion chamber



Figure 13. HyPE 1B Combustion Chamber

Figure 12. Impinging injector
during a water flow test in a
demonstration of atomization.

The oxidizer tank used in the HyPE 1B was designed in-house and
holds 36 lbs of liquid nitrous oxide. The entire tank was constructed from
6061-T6 aluminum alloy for its affordability and availability. The layout of
the oxidizer tank is shown in Figure 10. The outside diameter, excluding
bulkhead flanges, is 7.5 inches, and the overall length including bulkheads
is 26.6 inches.
The nitrous oxide is filled as a soft-cryogen, between -40 and -60 F.
This increases the oxidizer density and lowers the pressure required to
maintain it as a liquid. To help maintain low temperatures, the tank is
insulated with low temperature foam insulation. In the event that the
oxidizer heats and over pressurizes, the system is equipped with a burst disk
which activates at 675 psig to vent pressure. This condition requires a
complete reset of the system, but prevents catastrophic failure.
The nominal oxidizer pressure is 500 psia. This puts the tank under 15 ksi of hoop stress. At -50 F operating temperature,
the wall has a yield stress of 38 ksi
33
, and a FOS of 2.5. At temperatures of 100 F, the FOS drops to 2.3. This large FOS
primarily results from the expense needed to acquire a thinner walled tube of the required dimensions.
Additionally, the tank features a diffuser to even out the flow of gaseous nitrogen into the oxidizer tank and minimize the
introduction of gas into the liquid nitrous oxide. The tank bulkheads contain flanges to center the tank inside the rocket. Each
flange has four channels around the circumference to allow for wire harnesses
to pass alongside the tank and past the flanges.
E. Oxidizer Control Valve
To control the oxidizer flow valve actuation, a pneumatic piston system
was developed which is able to open or close the ball valve. This system uses
a dedicated on board compressed air reservoir and actuates in both directions,
achieving at least two full open close cycles. Since the HyPE 1B was not
designed to restart, the valve is only required to perform one cycle. The piston
must open the valve as part of the ignition sequence, and must close the valve
as part of the main engine cut off.
Due to the cost prohibitive nature of 3-way solenoid valves, this system
uses two simple solenoid valves arranged in two pairs. Each valve pair is
connected to one end of the piston.
F. Combustion Chamber
The HyPE 1B combustion chamber consists of a
chamber wall, ablative liner, injector, igniters, pre-
combustion chamber, fuel grain, post-combustion
chamber, and nozzle. The layout is shown in Figure 13.
All parts are made from 6061-T6 aluminum alloy unless
otherwise specified. The chamber itself is made from an
extruded aluminum tube. The outside diameter of the
chamber is 5 inches and the overall length from thrust
plate to nozzle exit is 32.1 inches.

1. Injector
The injector is the heart of the hybrid rocket engine. This is where the liquid
oxidizer is introduced to the solid fuel grain. The injector determines the flow rate
of oxidizer, which in turn controls many key parameters including: regression rate,
fuel burn rate, O/F ratio, and chamber pressure. Further, the effectiveness of the
injector determines how efficient combustion will be. Ideally the injector will
atomize the liquid into a gas but in practice the goal is to achieve the finest droplet
size possible.
The HyPE 1B uses an impinging showerhead injector as shown in Figure 12. In
this configuration, the oxidizer flow is split into streams which are all directed at a
single point, causing the streams to collide, creating very fine droplets. This
produces a highly atomized flow with enhanced combustion efficiency compared to

The UCLA Rocket Project


7

Figure 14. Ignition
Ring (Pyrotechnic
charge not shown)

a standard straight showerhead
34
.
The injector orifice determines the pressure drop across the injector, or the P. A large ratio of pressure drop to chamber
pressure, P/P, stabilizes combustion and prevents backflow of hot combustion gases into the oxidizer tank. Meanwhile, a
lower P/P saves weight by allowing for a less structural oxidizer tank as well as a smaller pressurant tank. The HyPE 1B
optimizes the P/P at 25%. A check valve between the oxidizer tank and the combustion chamber prevents gases from
moving back up to the tank in the event of a backflow. Based on the oxidizer pressure of 500 psia and losses in the oxidizer
plumbing line, the HyPE 1Bs chamber pressure is designed at 400 psia.
The total injector orifice area, or the combined area of all the injector holes, is related to mass flow, oxidizer flow rate,
and pressure drop
35
:

2
inj ox
ox
k
A m
P
=
A
(5)
K is the flow loss coefficient of 1.7
35
which has been verified in-house by cold flow testing. Design oxidizer flow rate is
3.53 lb/s (1.60 kg/s) and nitrous density is 68 lb/ft
3
(1089 kg/m
3
)
36
. This produces a design injector area of .0834 in
2
. Four
concentric rings were used, with 6, 12, 18, and 24 holes each (60 in total). The impingement point was chosen to be 2.0 in
from the injector to allow space for the igniters.

2. Ignition System
Ignition energy requirements in a hybrid rocket depend on initial oxidizer flow rate and fuel
volatility, which can be met very simply through adequate heating of the fuel grain in the
presence of an oxidizer
37
.
The HyPE utilizes a simple pyrogenic ignition system which provides hot gasses at high
pressure for initiation of the combustion process. An empirical formula for the heat required
for hybrid ignition, similar to the Bryan-Lawrence Equation for estimation of ignition energy
requirements for solid rocket motors
38
, was not found. However, required heat can be estimated
using equilibrium combustion thermodynamics
39
. If heat available from igniter combustion
exceeds the estimated required heat by an appreciable margin then ignition can be ensured
40
.
The ignition system consists of an assembled ignition ring that lies in the pre-combustion
chamber just aft of and flush against the injector. Six half-slices of D-class Aerotech model
rocket motors act as the pyrogenic igniters seen in Figure 14. They are ignited pyrotechnically
by commercial string-light bulbs filled with FFFFG black powder triggered by a 12 volt power
source.
In practice, the ignition ring houses two alternating sets of three igniters wired in parallel, adding a layer of redundancy
to the system. Each igniter has its own pyrotechnic charge. The wires run through the fuel grain port and out of the nozzle
and are cleanly forced out upon engine start.
Design of the ignition system was motivated by simplicity, accuracy, reproducibility, and reliability, which have been
improved over three major development phases since its inception for the HyPE 1A engine. The concept has been shown
effective in its application in solid rocket motors and in the similar cylindrical pyrogenic preheater grain system used in the
RATTworks M900 that the UCLA Rocket Project launched at the 5
th
ESRA IREC. Proof of concept and design optimizations
have been achieved through multiple firings of the Hype 1AT engine.

3. Pre/Post Combustion Chamber
The pre-combustion chamber holds the igniter system and also provides a space between the injector and fuel grain. This
allows for the oxidizer to impinge before reaching the grain, and also allows a recirculation area for enhanced mixing and
heating
34,35
.The pre-combustion chamber length was characterized based on the ignition system.
The post-combustion chamber provides a space between the fuel grain and nozzle, allowing combustion to complete
before the products enter the nozzle. A longer post-combustion chamber leads to better combustion efficiency, but also to
increased weight of the engine. The size used was based on a rule of thumb presented in Humble which suggests the optimal
post-combustion length is twice the fuel port diameter
35
. This gives the post-combustion length as 2.25 in.

4. Ablative Liner
The ablative pre-preg liner protects the combustion chamber wall from the hot combustion gases. By burning sacrificially,
the ablative produces a film of cooler gases which surround the chamber wall, protecting it. This is one of the simplest and
most cost effective methods of cooling an engine. The liner is expendable and must be replaced after each launch.
The best performing and most reasonable alternative was identified to be a silica-polyimide composite
41
. However, for
convenience, the use of carbon fiber was investigated for an ablative. From preliminary in-house torch tests, carbon fiber was
found to work as an adequate ablative.

The UCLA Rocket Project


8

Figure 15. MOC: Bell Nozzle Design

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16
0
0.01
0.02
0.03
0.04
0.05
0.06
length [m]
h
e
i
g
h
t

[
m
]
MOC Min Length Bell Nozzle
X: 0.1546
Y: 0.05123

Figure 16. Pneumatic Umbilical
Pusher System

G. Nozzle
In the HyPE 1A and HyPE 1AT, a divergent conical nozzle was
used. The flow at the exit of the conical nozzle is not all parallel to the
path of the rocket and so there are losses in usable thrust
39
. A bell
nozzle has been selected to replace the conical nozzle as it channels all
the thrust parallel to the flight path of the rocket, increasing the
efficiency of the rocket nozzle.
The Method of Characteristics (MOC) was used for designing the
minimum length bell nozzle. The MOC works on the principal that fluid
properties are constant along characteristic lines. As these characteristic
lines intersect, a system of equations can be iteratively solved to update
the fluid properties at the new location, creating the boundaries of the
nozzle wall as shown in Figure 15. In order to minimize the length of
the divergent section of the nozzle, the angle after the throat is based on
the Prantl-Meyer expansion fan
42
.
As a result of the new nozzle design, the same expansion ratio and
thrust has been maintained; however, efficiency has been increased by 2.2% (increasing usable thrust to 1004 lbs) and
decreasing the mass of the nozzle by 15%. The final bell nozzle design has been validated using simulations in SolidWorks
Flow Simulation and an in-house program using a finite difference MacCormack scheme with artificial viscosity
42
.
H. Umbilical Fill System
The HyPE 1B utilizes umbilical fill lines which can be remotely actuated.
To achieve this end, the fill lines for both oxidizer (nitrous oxide), and
pressurant (nitrogen) connect to the rocket through quick disconnect couplings.
These couplings are connected to a pneumatic piston-pusher system, shown in
Figure 16, which mechanically disconnects the fill lines, pushing them away to
avoid damaging the rocket during liftoff.
I. Analysis
For propulsion analysis, development continued on the in-house New Optimization Program (NOP). The program accepts
a variety of engine performance inputs including oxidizer flow rate, fuel grain dimensions, and chamber pressure, and
calculates the static thrust. The program then accepts simulation inputs such as initial rocket mass and initial altitude. The
NOP simulates the rockets trajectory by locking the engines design parameters and allowing flow rates and pressures to
change as the burn progresses.
The NOP uses the 1976 US Standard Atmosphere
43
for atmospheric conditions. It also accepts a Drag vs. Mach table as
an input, discussed in section II.B. Other minor considerations include the buoyant and the gravity forces which decrease
with altitude.
The program operates through a time step method. For each instant in time, the program calculates the current thrust,
atmospheric conditions, drag, weight, acceleration, velocity, position and propagates this information ahead in time, iterating
until apogee is reached. The program then plots the results of the simulation, including plots for altitude, velocity,
acceleration, thrust.
A rough overview of the calculations that are involved in the NOP follows. The increase in port radius for any t:

t t t t
r r r t
+A
= + A (6)
Where r is bore radius. This is used to calculate the burn characteristic of the fuel grain assuming constant oxidizer flow rate.
However, the NOP takes this a step further and allows oxidizer flow rate to change as a function of oxidizer pressure. Toward
the end of the burn, the pressurant tank drops below the regulator pressure and the oxidizer tank pressure begins to drop. This
converts the system from a pressure regulated system into a blow down system.

Rewriting the injector equation (5) gives:

2
2
ox
ox c c
inj ox
m k
P P P P
A
| |
= + A = + |
|
\ .
(7)
The equation for chamber pressure is
34
:

c
t
c
P m
A
-
=
(8)

The UCLA Rocket Project


9

Figure 17. MECO System Flight Test
Platform

Where c* is characteristic velocity, and A
t
is throat area. Substituting fuel and oxidizer mass flows for the total flow rate in
equation (8) and then substituting into equation (7) gives:

*
2 1 1 2
2
2
2
n n n
ox ox ox f p ox
ox inj t t
k c c
P m m ar l m
A A A
t

-

( ( (
= + +
( ( (
(
(9)
This equation is insightful in that everything inside the brackets is known at any time step. P
ox
is also known, as the NOP
keeps track of oxidizer pressure during the transition to a blow-down. This fully defines the engines performance for any
oxidizer pressure and port radius. Every time step, the NOP solves equation (9) numerically for oxidizer flow rate. This
allows the NOP to continue modeling the engine as it runs off-nominal These changes in oxidizer pressure cause changes in
both oxidizer and fuel flow rates, as well as changes in chamber pressure and Isp. This yields an improved performance
analysis.
To size the engine, oxidizer flow rate was varied as the main independent variable. As this changes, the fuel grain scales to
maintain the same O/F ratio. The required volume of nitrous was then found, and the CAD model and drag analysis were
changed based on the new geometries and parameters. The NOP was then run to determine the altitude given by this
configuration. The procedure was iterated until the projected altitude reached 28,000 ft to allow for margin in weight and
drag.
IV. Electronic Systems
A. Avionics
A main engine cutoff (MECO) system was designed in order to ensure that the rocket will not exceed the desired altitude
of 25,000 ft AGL. The system has two primary tasks:
1) Calculate the rockets current altitude, velocity and acceleration while in flight
2) Shut off the engine at the appropriate time to reach the desired altitude of 25,000 feet above ground level given its
current altitude, velocity and acceleration
The system will calculate the rockets current altitude using data from an Analog Devices ADXL345 three-axis
accelerometer. Acceleration readings are then integrated through the use of a trapezoidal Riemanns sum to obtain velocity,
and then again to obtain altitude
44
. The microcontroller also calculates the rockets remaining mass as a function of time,
using mass flow estimates from engine testing, while the rockets current mass is greater than its defined inert mass.
The MECO system will then use the calculated acceleration, velocity, position, and mass values and predict what altitude
the rocket will achieve if its engine were to instantaneously shut-off. Acceleration due to drag and gravity is modeled by
45
:

2
, 1,
1
, where
2
D
k sim D air D CS k sim
F
a g F C A v
m


= =
(10)
The simulation is iteratively solved until velocity reaches zero (simulated
apogee). If the resulting altitude prediction is greater than or equal to the
desired altitude of 25,000 feet AGL, the engine is shut-off by issuing a
signal to the oxidizer control valve. The density of air (

) is calculated
as a function of altitude using a polynomial equation curve-fitted to U.S.
Standard Atmosphere
43
values while the drag coefficient (

) table is
obtained from RASAero as discussed in section II.B.
The ADXL345 can measure changes in acceleration as small as
0.125 ft/s
2
and the onboard microcontroller performs these calculations
every 75 milliseconds. A small scale test of the system was carried out
on a model rocket using an A8-3 Estes motor, shown in Figure 17.
Figure 18 shows the result the tests. The accelerometer accurately calculated the instantaneous altitude to within 7% of the
actual value, which was measured using an angle finder. The error may be explained by the use of the angle finder method,
which carries large uncertainty. The apogee prediction code averaged around 6% error and generally underestimated the
apogee by a few feet.
There are a few non-idealities within the current MECO system that were not addressed. The system does not take into
consideration the attitude of the rocket. Attitude was determined using the remaining two axes of the accelerometer; however
after a preliminary engine noise test on the MECO system using vibrations during a static hot fire, it was found that
transverse vibrational noise on the accelerometer would make attitude correction nearly impossible. The current MECO
algorithm will only be a first-order approximation of the rockets true altitude. A time delay based on engine performance
modeling will be used to prevent premature MECO.

The UCLA Rocket Project


10

Figure 18. Integrated Position (Calculated Apogee = 131 ft.
Actual Apogee = 126 ft.)

0
20
40
60
80
100
120
140
0 1 2 3 4 5 6
P
o
s
i
t
i
o
n

(
f
t
)
Time (s)
Instantanious Altitude Predicted Apogee Actual Apogee

Figure 19. Schematic of the Control and DAQ Systems

Figure 20. Final PCB Assembly


The MECO system consists of redundant
3.3V Arduino Pro Minis based off an 8 MHz
Atmega 168 microcontroller, ADXL345 three-
axis accelerometers, an SD card data logger, and
a series of transistor operated relays connected to
a 14.8V Li-Po battery pack, which will be used to
actuate the oxidizer ball valve and vent valve, as
well as a means of switching between ground and
vehicle based control. The MECO command will
simultaneously close the oxidizer control valve
and re-open the oxidizer vent to assure no
oxidizer remains when the rocket is recovered,
safing the rocket. The redundant MECO system
will shut-off only when both microcontrollers
indicate an engine cutoff condition. Acceleration,
velocity, position data, as well as engine status
messages will be logged to the on-board memory.
B. Electronic Hardware Interface
The rocket test/launch control and data
acquisition infrastructure consists of two sets of
electronics bridged by long lengths of CAT5
Ethernet cables, which allows for safe rocket
operation as shown in Figure 19. The launch box is
the user interface for control and data acquisition.
The Switch Board, contains an array of
transistor-relay switches that toggle according to the
signals sent out at the Launch Box, which switch
high voltage control signals to actuate the control
hardware. The op-amp ICs, located on the Switch
Board, amplify the thermocouple and load cell
signals. This circuit uses a bipolar power supply,
which allows transmission of differential signals,
which are less susceptible to ground plane noise
46
.
There have been numerous improvements to the
test/launch infrastructure since 2010
47
. The Switch
Board has rapidly evolved from circuit prototypes to a full-fledged custom PCB, shown in Figure 20, in an effort to improve
circuit organization and reliability. Microcontrollers are still used to manage time sequenced events such as the ignition
sequence, but due to the general unreliability of the microcontroller experienced
during testing, manual overrides have been added to the ignition sequence.
C. Software
The GUI for monitoring live launch and test data was developed in National
Instruments LabVIEW programming language. The programs front panel tabs,
shown in Figure 21, displays information such as volumetric flow rate, pressure
transducer readings, force transducer readings, and thermocouple readings. Raw
sensor data is linearized and then simultaneously recorded to file and displayed on
a time domain graph. The software also has integrated abort case detection code
that monitors pressures and alerts the user if a pressure reading is out of spec,
allowing the operator to take steps to safe the rocket. Future work will be done to
automate the control of the rocket in these abort cases. The GUI also displays which switches and controls have been
activated.
The graphical programming language is organized into modular sections comprised of various calculation threads for each
sensor. The calibrations constants used to linearize the raw data were obtained by manually calibrating sensors in the lab, and
then manually inputting them into the LabVIEW interface on the front panel.

The UCLA Rocket Project


11

Figure 21. Labview Front Panel Showing All
Indicators for a Hot Fire Test

An automated launch/abort system was investigated in order
to simplify the launch control process. A simple Arduino
microcontroller was tested for controlling the launch task
signals between the launch box and the pad. After an initial
field test with minimal automation functions, it was determined
that the microcontroller was too unreliable for the factors of
safety and reliability desired in the mission. The National
Instruments software and hardware are potential candidates for
attempts at automating future launch sequences.
D. Scientific Payload
The 10 lb payload to be carried to 25,000 ft AGL is located
in two separate locations. The nosecone contains scientific
instruments to record the acceleration, rotation, ambient
pressure, and nosecone temperature. The engine bay contains an HD camera compartment for recording the flight. The
nosecone payload houses a small project box that holds a 3.3V Arduino Pro Mini microcontroller based off an 8MHz Atmega
168, which interfaces with a BMP065 barometric pressure sensor and an IMU3000 inertial measurement unit (IMU) via an
I
2
C digital communication protocol as well as a DS18B20 ambient temperature sensor device and k-type thermocouple. An
instrumentation amplifier is required to amplify the signal from the thermocouples. Data from the sensors is stored to an SD
card reader for post flight analysis. The barometric pressure sensor is used to monitor the change in pressure as the rocket
travels through the transonic region and the pressure at the rockets apogee. This data can then be used to design a more
accurate altimeter system which can be used to initiate MECO. The IMU unit is used to characterize the trajectory of the
rocket as well as record the transverse and vibrational forces acting on the rocket during flight. The data will then be used to
help improve future designs of the rockets nosecone, body, and fins. The thermocouples are used to record the temperatures
experienced by the nosecone as a result of air friction, which give insight into the flight environment. The Aiptek PenCam
HD Trio is located in the engine bay and will film the rocket launch, flight, and descent. The camera has a battery life of over
two hours and a recording time of approximately 1 hour 40 minutes, and records to on board memory for post-flight analysis.
The HD camera records video viewed down the length of the rocket by using an aerodynamic faring with a 3/4 square
mirror angled at a 45 degree angle within the faring. Modifications were made to the internal circuitry of the camera to
provide external wire leads to operate the power and record functions using the screw terminals
48
.
V. Recovery Systems
For the reusability criterion to be met, the HyPE utilizes a dual deployment recovery system that includes a drogue and
a main parachute. The purpose of the dual deployment scheme is to prevent the accumulation of large drift distances during a
majority of its descent.
A. Parachute Descent Rate
A parachutes physical parameters are used to determine a suitable descent rate that limits the stress placed upon the
rocket when it impacts the ground. A custom descent rate calculator was created for the drogue and the main parachute
systems. The descent rate output function helps determine the viability of parachutes utilized for recovery. This calculator is
based upon equation (11)
46
. The functions inputs involve the weight of the rocket, the air density at apogee, which is held
constant to simulate a worst-case scenario, the parachutes canopy area, and the coefficient of drag.

2
T
t
ref D
W
v
A C
=
(11)
The desired main parachute descent rate is between 15 and 25 ft/s led to the choice of the Sky Angle XXL parachute. The
drogue parachutes ability to minimize drift while reducing stress during the main parachutes ejection led to the selection of
the TAC-1 parachute
49
.
Table 3. Descent Rate Inputs and Outputs
Main Parachute Drogue Parachute
Cd 2.92
50
1.5
51

Aref (ft
2
) 129
50
28.3
52

vt (ft/s) 18.8 82.9
Approximated Total Descent Time (s) 339

The UCLA Rocket Project


12

Figure 22. Concept
Behind the Drift
Application
56



Figure 23. Graphical Results of the Drift Application Program

B. Parachute Deployment
At 25,000 feet above ground level, the air density becomes significantly reduced, which
limits the ability of black powder to combust rapidly and to create enough of a pressure
gradient to deploy the drogue parachute
53
. The Rouse-Tech CD3s 25 gram carbon dioxide
deployment system was adopted to cope with these high altitude conditions and ensure
deployment of the drogue parachute at apogee. The CD3 charge consists of a sealed black
powder controlled plunger which punctures a carbon dioxide canister, releasing pressurized
gas
53
. Since the main parachute will be deployed at an altitude of 1,300 feet above ground
level, the air density is high enough for black powder to be effective. As a result of multiple
tests it was shown that 2.5 grams of black powder was sufficient to jettison the main
parachute from its compartment
54
.
A single level of redundancy is employed in the recovery system. The system has two
separate circuits for each set of ejection stages: the drogue deployment stage and main
deployment stage. Two PerfectFlite HiAlt45K altimeters, which are capable of measuring
altitudes of up to 45,000 feet, are used for apogee determination and triggering of ejection
charges. Each altimeter is connected to its own battery, switch, and deployment charges.
C. Rocket Drift during Descent
In an attempt to decrease the recovery time of the rocket, an iteration-based program has
been created that determines a projected landing site and bearing from the projected apogee
towards the ground. The program requires wind velocity, wind bearing, and several physical
rocket parameter inputs. Equating the drag force equation
51
to Newtons second law yields equation (12)
55


2
( , )
2
D ref rel
x y
C A v
a
m

=
(12)
By iterating this equation for a layer of air every 100 feet from apogee to the ground, we obtain outputs which can be used to
determine rockets theoretical drift distance and bearing for the 100 feet layer of air. In order to account for the possibility of
error and fluctuations in the launch days wind patterns, a factor of a sixth has been incorporated into the program to set
upper and lower bounds of error. This is indicated by the red, transparent circle on the graph titled Drift Projection onto the
Cardinal Directional Plane
56
. The results were simulated based upon wind profile data on May 19
th
of 2011 at the Universal
Coordinated Time of 08Z, or 3 a.m. in the state of Utah
57,58
.
VI. Conclusion
The UCLA Rocket Project has spent the past three years researching and developing hybrid propulsion technologies
for a student designed and fabricated rocket to be flown at the 6
th
IREC Competition in Green River, Utah. The completed
engine design burns aluminized paraffin with nitrous oxide to produce 1004 lbs of thrust with an Isp of 202 s and burn time
9.9 s. With this engine, the 15 ft long 8 in diameter rocket will carry a 10 lb payload to 25,000 ft above ground level. An
avionics package calculates the rockets trajectory, analyzing the optimal time to cut thrust to achieve a precise altitude.
Ground control hardware is used to both control the rocket and collect data during tests. The payload takes scientific
measurements and video to aid in the development and construction of future UCLA rockets. Numerous precautions have
been taken to assure the rocket can be operated safely.
Each individual subsystem design within the rocket has been scrutinized and evaluated for weight, reliability,
simplicity, and safety. The electronics, structures, propulsion, and recovery systems have all been optimized both in theory
and design to culminate in UCLA AIAAs most ambitious project to date, the HyPE rocket.


The UCLA Rocket Project


13
Acknowledgments
The UCLA Rocket Project would like to thank its numerous sponsors, without whom this project could never have left
the drawing board. We would like to especially thank C-P Manufacturing Corporation for their generous donation of time
and skill in machining our precision engine components. Thanks to Northrop Grumman, Lockheed Martin, AFRL, Aerospace
Corporation, Engineering Alumni Association, and the UCLA Engineering School for their monetary donations. Thanks to
Dresser for the donation of several great regulators. Thanks to Airtech, SFU, 3M, and SpaceX for the donation of composite
layup related supplies. To ROC Carbon for the donation of the graphite and machining for our nozzles. To Mouser
Electronics for the donation of electronic components and parts. To Quality Precision Clearing Inc. for the oxygen cleaning
our oxidizer system, and to all our other sponsors. Last but not least, a special thanks to Dr. Richard Wirz for his cont inued
guidance and support of the Rocket Project.
References
1.
Jackson, J. (2011, January 6). Working with Composite Materials.
2.
Chin, S. S. (1961). Missile Configuration Design. McGraw-Hill Book Company, Inc.
3.
Krasnov, N. F. (1970). Aerodynamics of Bodies of Revolution. New York: American Elsevier Publishing Company, Inc.
4.
Stoney, J. W. (1958). Collection of Zero-Lift Drag Data on Bodies of Revolution from Free-Flight Investigations.Washington: NACA.
5.
Rogers, C. E., & Cooper, D. (2011). Rogers Aeroscience RASAero Aerodynamic Analysis and Flight Simulation Program. Lancaster:
Rogers Aeroscience.
6.
Stoney, J. W. (1954). Transonic Drag Measurements of Eight Body-Nose Shapes. Washington DC: NACA.
7.
The Great Soviet Encyclopedia, 3rd Edition. (1970-1979). The Gale Group, Inc.
8.
Jack, J. R. Theoretical Pressure Distribution and Wave Drags for Conical Boattails.
9.
Harder, K. C., & Rennemann Jr., C. (1955). On Boattail Bodies of Revolution Having Minumum Wave Drag - NACA Technical Note
3478. Washington DC: NACA.
10.
Niskanen, S. (2009). Development of Open Source model rocket simulation software. Helsinki: Helsinki University of Technology.
11.
Barrowman, J. S., & Barrowman, J. A. (1966). The Theoretical Prediction of the Center of Pressure.
12.
(1964). Flutter, Buzz and Divergence. In NASA Space Vehicle Design Criteria SP - 8003. Washington DC: NASA.
13.
Martin, D. J. (1958). Summary of Flutter Experiences as a Guide to the Preliminary Design of Lifting Surfaces on Missiles.
Washington: NACA.
14.
McCormick, A., Hultgren, E., Lichtman, M., Smith, J., Sneed, R., & Azimi, S. (2005). Design, Optimization, and Launch of a 3"
Diameter N2O/Aluminized Paraffin Rocket. Tucson, Arizona: AIAA.
15.
Van Pelt, D., Hopkins, J., Skinner, M., Buchanan, A., Gulman, R., Chan, H., et al. (2004). Overview of a 4-in OD Paraffin-Based
Hybrid Sounding Rocket Program. Fort Lauderdale, Florida: AIAA.
16.
Dyer, J., Zilliac, G., Doran, E., Marzoa, M. T., Lohner, K., Karlik, E., et al. (2008). Status Update Report for the Peregrine 100km
Sounding Rocket Project. Hartford, CT: AIAA.
17.
Grosse, M., & Schlatzke, G. (2008). Development of a Hybrid Rocket Motor Using a Diaphragm for a Small Test Rocket. Hartford,
CT: AIAA.
18.
Davidy, A. (2008). Development of an Analytical Solution to the Heat Transfer Equation in the Burning Paraffin Fuel. Hartford, CT:
AIAA.
19.
Evans, B., Favorito, N. A., & Kuo, K. K. (2005). Study of Solid Fuel Burning-Rate Enhancement Behavior in an X-ray Translucent
Hybrid Rocket Motor. Tucson, Arizona: AIAA.
20.
Carmicino, C., & Sorge, A. R. (2006). The Effects of Oxidizer Injector Design on Hybrid Rockets Combustion Stability. Sacramento,
CA: AIAA.
21.
Lohner, K., Dyer, J., Doran, E., & Dunn, Z. (2006). Fuel Regression Rate Characterization Using a Laboratory Scale Nitrous Oxide
Hybrid Propulsion System. Sacramento, CA: AIAA.
22.
Dunn, Z., Dyer, J., Lohner, K., Doran, E., Bayart, C., Sadhwani, A., et al. (2007). Test Facility Development for the 15,000 lb Thrust
Peregrine Hybrid Sounding Rocket. Cincinnati, OH: AIAA.
23.
Karabeyoglu, A., Cantwell, B., & Zilliac, G. (2005). Development of Scalable Space-Time Averaged Regression Rate Expressions
for Hybrid Rockets. Tuscon, Arizona: AIAA.
24.
Karabeyoglu, A., Zilliac, G., Cantwell, B., De Zilwa, S., & Castelluci, P. (2003). Scale-Up Tests of High Regression Rate Liquefying
Hybrid Rocket Fuels. AIAA.
25.
Karabeyoglu, A., Zilliac, G., Castellucci, P., Urbanczyk, P., Stevens, J., Inalhan, G., et al. (2003). Development of High-Burning-
Rate Hybrid-Rocket-Fuel Flight Demonstrators. Huntsville, Alabama: AIAA.
26.
Greatrix, D. R. (2007). Model for Predicting Fuel Regression Rate in Hybrid Rocket Engines. Cincinnati, OH: AIAA.
27.
Doran, E., Dyer, J., Lohner, K., Dunn, Z., Cantwell, B., & Zilliac, G. (2007). Nitrous Oxide Hybrid Rocket Motor Fuel Regression
Rate Characterization. Cincinnati, OH: AIAA.
28.
Zilliac, G., & Karabeyoglu, A. (2006). Hybrid Rocket Fuel Regression Rate Data and Modeling. Sacramento, CA: AIAA.
29.
Zilliac, G., & Karabeyoglu, A. (2005). Modeling of Prepellant Tank Pressurization. Tuscon, Arizona: AIAA.
30.
Kilic, S., Karabeyoglu, A., Stevens, J., & Cantwell, B. (2003). Modeling the Slump Characteristics of the Hydrocarbon-Based
Hybrid Rocket Fuels. Huntsville, Alabama: AIAA.

The UCLA Rocket Project


14
31.
Abu-Awaad, F. M. (April 2004). The Gas-Phase Heats of Formation of n-Alkanes as a Function of the Electrostatic Potential
Extrema on their Molecular Surfaces. Journal of Chemistry , 81-86.
32.
Luxfer Cylinders. Product: LCX composite cylinder specifications.
33.
MIL-HDBK-5J: Metallic Materials and Elements for Aerospace Vehicle Structures. (2003). Department of Defense Handbook.
34.
Sutton, G. P., & Biblarz, O. (2001). Rocket Propulsion Elements. John Wiley & Sons.
35.
Humble, R. W. (1995). Space Propulsion Analysis and Design, 1st Ed. Learning Solutions.
36.
Braker, W., & Mossman, A. L. (1980). Matheson Gas Data Book, 6th Ed. Lyndhurst, NJ: Matheson.
37.
Schuler, A. L., & Wiley, D. R. (1989). Hybrid propulsion technology program: Phase 1. Volume 3: Thiokol Corporation Space
Operations. NASA CASI.
38.
National Aeronautics and Space Administration. (1971). Solid rocket motor igniters. NASA space vehicle design criteria, chemical
propulsion . Cleveland: NASA CASI.
39.
Hill, P. G., & Peterson, C. R. (1992). Mechanics and Thermodynamics of Propulsion, 2nd. Addison-Wesley Publishing Company,
Inc.
40.
Story, G., Zoladz, T., Arves, J., Kearney, D., Abel, T., & Park, O. (2003). Hybrid Propulsion Demonstration Program 250K Hybrid
Motor . Huntsville: NASA CASI.
41.
Richter, G. P., & Smith, T. D. (1995). Ablative Material Testing for Low-Pressure, Low-Cost Rocket Engines. Cleveland, Ohio:
Lewis Research Center.
42.
Anderson, J. (2003). Modern Compressible Flow 3rd Ed. New York, NY: McGraw-Hill Companies, Inc.
43.
National Aeronautics and Space Administration. (1976). US Standard Atmosphere 1976.
44.
Young, H. D., & Freedman, R. A. (2010). University Physics (13 ed.). San Fransisco, CA: Pearson/Addison Wesley.
45.
Anderson, J. D. (2011). Fundamentals of Aerodynamics (5 ed.). Boston: McGraw-Hill.
46.
Lattice Semiconductor Corporation. (2001, May). Differential Signaling. Application Note. Hillsboro, OR.
47.
Prendergast, L. M., Zimmerman, K. U., Silva, W. A., Chang, P. C., Sivaram, S., Kentosh, B. C., et al. (2010). Paraffin Wax Based
Experimental Hybrid Motor.
48.
Chang, P. C. (2011, May 5). 5/5/11 Electronics Team Update. Retrieved May 23, 2011, from HyPE: Hybrid Propulsion Experiment:
http://aiaa.seas.ucla.edu/RP2011
49.
Knacke, T. (1992). Parachute Recovery Systems: Design Manual. Santa Barbara, CA: Para Publishing.
50.
The b2 Rocketry Company. (n.d.). Retrieved February 2011, from b2 Rocketry Web Site: http://www.b2rocketry.com
51.
Munson, B. R. (2009). Fundamentals of Fluid Mechanics (Vol. 6). Wiley.
52.
Giant Leap Rocketry. (n.d.). Retrieved February 2011, from Giant Leap Rocketry Web Site:
http://www.giantleaprocketry.com/products_recovery.asp
53.
Rouse-Tech, Inc. (2003). Rouse-Tech CD3 Instruction Manual. Rouse-Tech, Inc.
54.
Rouse Tech, T. (2011, April). Determination of Ejection Charges for Parachute Deployment. (R. Abrantes, Interviewer)
55.
Kavehpour, H. P. (2011, February). Application of Drag Forces. (R. Abrantes, Interviewer)
56.
Charlesworth, P. (2004). The Impact of Wind Profile on High Altitude Flights by Fin Stabilised Rockets. United Kingdom Rocket
Assocation.
57.
Correlations of Air Pressure to Wind Velocity above 20,000 Feet. (n.d.). IPS Meteostar.
58.
Wind Profile from Ground Level to 20,000 Feet. (n.d.). Salt Lake City: National Oceanic and Atmospheric Administration: National
Weather Service.

You might also like