You are on page 1of 86

REGRESSION RATE STUDY OF HTPB / N

2
O HYBRID ROCKETS

by
HEATHER L. CHLUDA
B.S.M.E., Ohio University, 1999


A thesis submitted to the
Faculty of the Graduate School of the
University of Colorado in partial fulfillment
of the requirement for the degree of
Masters of Science
Department of Mechanical Engineering
2006




UMI Number: 1433501
1433501
2007
UMI Microform
Copyright
All rights reserved. This microform edition is protected against
unauthorized copying under Title 17, United States Code.
ProQuest Information and Learning Company
300 North Zeeb Road
P.O. Box 1346
Ann Arbor, MI 48106-1346
by ProQuest Information and Learning Company.
This thesis entitled:
Regression Rate Study of HTPB / N
2
O Hybrid Rockets
Written by Heather Chluda
has been approved for the Department of Mechanical Engineering




___________________________________
David Kassoy






____________________________________
Melvin Branch



Date__________________

The final copy of this thesis has been examined by the signatories, and we find that
both the content and the form meet acceptable presentation standards of scholarly
work in the above mentioned discipline.











ii
Chluda, Heather (M.S. Mechanical Engineering)
Regression Rate Study of HTPB / N
2
O Hybrid Rockets
Thesis directed by Professor David Kassoy

ABSTRACT
Hybrid rockets are once again gaining distinction in the rocket industry and
are being recognized as a simpler and safer rocket for upper stage commercial and
human space flight. The current study explores regression rates for the specific solid
fuel and liquid oxidizer combination, Hydroxyl-terminated Polybutadiene (HTPB)
and Nitrous oxide (N
2
O). Recent studies show this propellant combination has
benign environmental impacts and gives ultimate safety for human space flight.
Different methods are explored to derive empirical constants for regression rate. The
regression rate equation found for the current study is

r =0.36Gox
1.02
, which
correlates to previous tests performed at the University of Colorado, Boulder.
Conclusions are made in comparing regression rates for different oxidizers. The
results of chemical kinetics codes are compared. An approach is presented to
determine the thermochemistry for reactants HTPB, as hydrocarbon molecule
C
4
H
8
O
2
, and N
2
O, used for predicting solid fuel regression rates from ideal rocket
equations.

iii
ACKNOWLEDGEMENTS

I would like to express my great appreciation to my thesis advisor, Dr. David
Kassoy, for his support and patience throughout the learning phases of this thesis
project. Dr. Melvyn Branch and Dr. Sedat Biringen are also to be acknowledged for
their input and insight into the different aspects of this research that spanned the
combustion and aerospace realms. I would also like to thank the members of the
MaCH-SR1 seniors who have been continuously expanding hybrid rocket research.
Finally, I would like express my great appreciation for my parents and siblings.
Their unconditional support and love has been a constant driver throughout this
endeavor.





iv
CONTENTS

ABSTRACT...............................................................................................................iii
ACKNOWLEDGEMENTS....................................................................................... iv
CONTENTS................................................................................................................ v
LIST OF TABLES..................................................................................................... vi
LIST OF FIGURES ..................................................................................................vii
NOMENCLATURE .................................................................................................. ix
1. INTRODUCTION.............................................................................................. 1
1.1 PROPELLANT PROPERTIES AND CHARACTERISTICS ................... 3
1.2 OUTSTANDING PROBLEMS WITH HYBRID MOTORS .................... 7
1.3 FUEL REGRESSION FOR HTPB/ N
2
O COMBINATION...................... 9
2. LITERATURE SUMMARY............................................................................ 11
2.1 DIFFUSION COMBUSTION FLAME THEORY .................................. 12
2.3 HISTORICAL HYBRID ROCKET EXPERIMENTS............................. 19
2.4 DESIGN CALCULATIONS .................................................................... 25
3. CHEMICAL KINETICS MODEL................................................................... 28
3.1 STANJAN MODELING .......................................................................... 29
3.2 CHEMKIN MODELING.......................................................................... 33
3.3 MODEL COMPARISONS AND CONCLUSIONS ................................ 36
4. EXPERIMENT................................................................................................. 39
4.1 EXPERIMENTAL DESIGN AND SETUP ............................................. 41
4.2 EXPERIMENTAL SETUP....................................................................... 46
4.3 EXPERIMENT RESULTS....................................................................... 54
4.4 PREDICTIONS USING IDEAL ROCKET CALCULATIONS.............. 62
4.5 DISCUSSION OF RESULTS................................................................... 66
5. CONCLUSIONS AND FUTURE CONSIDERATIONS................................. 70
6. REFERENCES ................................................................................................. 74



v
LIST OF TABLES


Table 1. Thermo chemical Properties of Representative Hybrid Propellants............. 5
Table 2. Chemical Kinetics of Hybrid Rockets ........................................................ 19
Table 3. Regression Rate Equations ......................................................................... 21
Table 4. Oxidizer/Fuel Ratio vs. Chamber Temperature and Specific Impulse ....... 26
Table 5. HTPB fuel mixture ingredients used in all MaCH-SR1 HRM hot-fires..... 37
Table 6. GDL Propep Input for HTPB and Nitrous oxide with an O/F = 4 ............. 37
Table 7. GDL Propep Results and Oxidizer Comparisons ....................................... 38
Table 8. Instrumentation List for Baseline Tests...................................................... 53
Table 9. Regression Rate Tables for k-Type thermocouples.................................... 56
Table 10. Steady State System Analysis Results Summary for Planned and Actual
Hot Fire Parameters .................................................................................................. 65
Table 11. Regression Rate Equations in Literature .................................................. 67
Table 12. Summary of Thermochemistry Parameters with Calculated Regression
Rate Results .............................................................................................................. 71

vi
LIST OF FIGURES

Figure 1. 250,000 lbf Hybrid Motor Test at Stennis Space Center on August 13,
1999 [Picture courtesy of Rocketdyne]....................................................................... 3
Figure 2. Thesis Objectives....................................................................................... 11
Figure 3. Hybrid Rocket Combustion - Boundary Layer with Embedded Flame .... 13
Figure 4. Typical Hybrid Rocket Layout .................................................................. 16
Figure 5. Pyrolysis of Liquifying Fuels, ie. Paraffin (Karabeyoglu, et al., 2002) .... 17
Figure 6. Ames Research Center Test of Paraffin/Oxygen Hybrid Rocket
(Karabeyoglu, et al. 2002) ........................................................................................ 20
Figure 7. Regression Rate Data for Paraffin and HTPB Fuels (Karabeyoglu, et al.,
2002) ......................................................................................................................... 23
Figure 8. C
4
H
8
O
2
-N
2
O Adiabatic flame temperature ............................................... 31
Figure 9. Major product species for optimized C
4
H
8
O
2
-N2O fuel - oxidizer
combustion................................................................................................................ 31
Figure 10. Formaldehyde (CH
2
O)/ N
2
O Adiabatic flame temperature..................... 32
Figure 11. Major product species for optimized CH
2
O-N
2
O fuel - oxidizer
combustion................................................................................................................ 33
Figure 12. Species mole fractions and temperature as a function of time for an
adiabatic and constant pressure results for Eq. (11) reaction mechanism for HTPB /
N
2
O combustion........................................................................................................ 35
Figure 13. Species mole fractions and temperature as a function of time for an
adiabatic and constant pressure results for Eq. (12) reaction mechanism for HTPB /
N
2
O combustion........................................................................................................ 36
Figure 14. Hybrid Rocket Flight Configuration ....................................................... 40
Figure 15. Full-scale Star Port Fuel Regression Characteristics (MaCH-SR1 2004)
................................................................................................................................... 42
Figure 16. Lab-scale Mutli Port Regression Characteristics (MaCH-SR1 2004)..... 42
Figure 17. Lab-scale 4-Point Star Port Fuel Grain (MaCH-SR1 2005).................... 43
Figure 18. Nanmac Eroding Thermocouple.............................................................. 45
Figure 19. K-type Thermocouples ............................................................................ 46
Figure 20. Baseline Hybrid Rocket Motors with Instrumentation............................ 50
Figure 21. Hybrid Motor Test Setup Sketch............................................................. 51
Figure 22. Baseline Hybrid Rocket Motor and Instrumentation Ready for Testing. 51
Figure 23. Data Acquisition System Setup............................................................... 54
Figure 24. Test Matrix for MaCH-SR1 2005............................................................ 55
Figure 25. Baseline 1 Regression Rate Raw data ..................................................... 57
Figure 26. Baseline 2 Regression Rate Raw Data .................................................... 57
Figure 27. MaCH-SR1 2005 Regression Data and empirical constants found by
Power Curve Method................................................................................................ 58
Figure 28. Log - Log Plot to determine a and n constants........................................ 58
Figure 29. HTPB Regression Rate Results............................................................... 59
Figure 30. Modified Regression Rate Equation for Percentage of Oxygen in N2O
Oxidizer..................................................................................................................... 60
Figure 31. Percentage of Oxidizer in the Trough Area of a 4-Point Star Fuel Port
Geometry................................................................................................................... 61

vii
Figure 32. NANMAC Eroding Thermocouple Attempt at Measuring Combustion
Temperature .............................................................................................................. 62
Figure 33. Regression Rate for Single Round Port HRMs ....................................... 69
Figure 34. MaCH-SR1 April 2004 hot fires Injector End View, both after 18
second burn time....................................................................................................... 70
Figure 35. Nozzle End View, a)after 5 sec burn time, b) after 18 sec burn time ..... 70

viii
NOMENCLATURE
A = Arrhenius constant
port
A = fuel port area (m
2
)
a, n, m = regression rate constants
E = activation energy (kcal/mol)
S
p
= circumference of individual ports (m)
D = fuel port diameter
G
o
= mass flux (kg/m
2
s)
HTPB = hydroxyl-terminated polybutadeine
HRM = hybrid rocket motor
ID = inner diameter
IPDI = isophorone diisocyanate
ISP = specific impulse (s)
K = kelvin
L = port length (m)
LM = Lockheed Martin
lb
f
= pound-force
lbm = pound-mass
m = mass (kg)

ox
m = oxidizer mass flow rate (kg/s)

f
m = fuel mass flow rate (kg/s)
MR = mixture ratio
N = number of fuel ports
N
2
O = nitrous oxide
O
2
= oxygen
O/F = oxidizer to fuel ratio
OD = outer diameter
= equivalence ratio, defined as O/F
theoretical
/ O/F
actual

r = fuel port radial regression rate (mm/s)


= degrees
f
= fuel density (kg/m
3
)





ix
1. INTRODUCTION

The rocket family consists of four main types; Liquid propellant, solid
propellant, hybrid, and monopropellant. These rockets have been the workhorses of
the commercial, aerospace, defense, and government research businesses. Each of
these rockets has their unique advantages and disadvantages. An important
parameter to characterize a rocket is the specific impulse or thrust to weight ratio,
which is used for comparison among these different types of rockets (Larson and
Wertz, 1999).
The liquid rocket engine propulsion system combines two liquid propellants
as the oxidizer and fuel for combustion. The advantages of these engines are the
high specific impulse (ISP) ranging from 300-450 seconds

(Larson and Wertz,
1999), and throttle ability (controllability). The disadvantages are the complexity of
the system which includes thousands of pipes, valves, connections, and high speed
rotating machinery to pump the liquid fuel and oxidizer, and safe storage of the
liquid and often cryogenic propellants. These myriad of parts are often assembled
with redundancies to increase the safety and reliability of the system due to the
catastrophic failure if an integral part should fail.
The solid rocket motor propulsion system combines the fuel and oxidizer
into a single solid grain with varying geometries. The advantages are the simplicity
of the motor design, no mixing of oxidizer and fuel, and reliability. However, solid
rocket motors produce a lower ISP than liquid rockets engines, ranging from 280-
300 seconds and have a higher rocket mass (Larson and Wertz, 1999). Other

1
disadvantages include; combustion instabilities with increasing complexity in grain
geometry, no throttling abilities, and no shutdown capabilities after the solid
propellant has been ignited.
Monopropellants use one liquid propellant for combustion. The advantages
of these rockets are the simplified propellant-feed and control system. However, the
low thrust output limits the use for mainly small attitude control thrusters on
orbiting satellites or spacecraft. The most common monopropellant is hydrazine that
has an ISP of about 200-260 seconds. Hydrazine can be stored for over 15 years and
has good combustion characteristics. The Space Shuttle uses 38 different thrusters
for attitude control, orbit corrections, station keeping, or for reentry (Sutton and
Biblarz, 2001).
Hybrid rocket motors often use a solid fuel grain that lines the combustion
chamber and a liquid or gaseous oxidizer injected into the combustion chamber that
can be pressure fed or turbopump fed. Hybrid rocket motors combine the safety of
separating the fuel and oxidizer, as in liquid rockets, with the simplicity of a solid
rocket motor design. Hybrid rockets have several advantages including; simplicity
of design, controllability of the thrust during flight and re-ignition capabilities
(Larson and Wertz, 1999). The potential for higher ISP values has been established
using liquid oxygen and HTPB oxidizer-fuel combinations for hybrid motor
combustion. A typical HTPB/LOX hybrid rocket has an equivalent ISP to a RP-1 /
LOX engine. One recorded vacuum ISP value for a 250,000 lb
f
thrust, liquid
oxygen HTPB/Escorez hybrid that has been tested and validated at NASAs
Stennis Space Center was 276 seconds (Story, et al., 2003), making hybrid rockets a

2
feasible alternative to existing space launch vehicles. A picture of this hybrid rocket
motor hot-fire is seen in Figure 1.

Figure 1. 250,000 lbf Hybrid Motor Test at Stennis Space Center on August 13,
1999 [Picture courtesy of Rocketdyne]
The disadvantages are the combustion instability effects seen in large scale hybrids,
the scale up prediction difficulties from lab scale experiments (i.e. the 100 - 1000 lb
f

range) to large scale tests (i.e. the 10,000 - 250,000 lb
f
range), and the
inconsistencies in fuel regression rate due to testing variations (Boardman, et al.,
1997 and Chiaverini, et al., 2001). Hybrid rocket motors are among the safest and
simplest rockets and have the potential to contribute widely to the space industry.
1.1 PROPELLANT PROPERTIES AND CHARACTERISTICS
The use of benign oxidizers and fuels makes the hybrid rocket motor a
viable candidate for wide range of uses due to the immense safety, low hazardous
exhaust emissions, lower cost design and higher reliability. The potential increase in
ISP over solid motors is also an important attribute that makes this rocket type a

3
desirable candidate for booster rockets as well as upper stage motors. Currently,
upper stage rockets are dominated by the efficient but expensive liquid bi-propellant
rockets.
Several different fuel and oxidizer combinations for hybrid rocket motors
have been tested starting in the early 1960s to the present. Some common fuels
used are Plexiglas, HTPB, and paraffin wax, while the oxidizers are oxygen (O
2
),
nitrous oxide (N
2
O), hydrogen peroxide (H
2
O
2
), and nitrogen tetroxide (N
2
O
4
).
Plexiglas was first used as a solid fuel for hybrid rockets in the 1960s by Marxman,
et al., (1964) and gives a high regression rate. HTPB has an added safety benefit of
being nearly inert in the absence of an oxidizer so it will not chemically react by
itself. Paraffin wax, the same wax used in household hurricane candles, has been
recently researched at Stanford University and tested for a high regression rate that
is 3 times faster than HTPB (Karabeyoglu et al., 2003).
The oxidizers commonly used have benign impacts on the environment.
Liquid or gaseous oxygen is the most widely used oxidizer due to its high ISP
output with any of the fuels mentioned. Liquid oxygen requires high maintenance to
keep it refrigerated to cryogenic temperatures. Dinitrogen monoxide (N
2
O) or
laughing gas is self pressurizing to 700 psi at the liquid state so it is easy to
handle. Hydrogen peroxide (H
2
O
2
) is difficult to use because it is a hypergolic (like
the monopropellant, hydrazine) but the decomposition product is essentially water
(Sutton and Biblarz 2001). Nitrogen tetroxide (N
2
O
4
) has a high density and can be
stored almost indefinitely. N
2
O
4
is only mildly corrosive when pure. The hybrid
rocket solid fuels are in extreme contrast to the environmentally unfriendly solid

4
rocket fuels which produce acidic hydrogen chloride and other polluting chemicals
(Sutton and Biblarz, 2001).
The theoretical ISP data in Table 1 is taken from Sutton and Biblarz (2001)
in Chapter 15 on hybrid rockets. The table shows the ideal specific impulse at
vacuum for different oxidizers reacting with HTPB. It is shown that ideally
HPTB/O
2
propellant combination has comparable characteristics to the RP-1/O
2

liquid bi-propellant rocket engine, as mentioned before. However, HTPB/N
2
O
propellant combination has less actual ISP capability, due to the low oxidizer
density and high flame temperatures needed to vaporize the fuel. High-energy
density materials can also be added to the solid propellant to increase the ISP or
thrust of a rocket (Frisbee 2003). HTPB with burn enhancing additives, like
aluminum, mixed into the fuel may make the fuel a higher performer and better
candidate for large scale hybrid rocket motors (Chiaverini, et al., 2000) but these
materials will not be further researched in this thesis.
Table 1. Thermo chemical Properties of Representative Hybrid Propellants
Fuel / Oxidizer
Isp
Ideal
(sec)
Isp
Actual
(sec)
Density
Fuel
(kg/m^3)
Density
Oxidizer
(kg/m^3)
HTPB / O2 360 331 900 1227
Paraffin / O2 369 314 930 1227
Plexiglas / O2 270 310 1190 1227
HTPB / N2O 224 200 900 745
HTPB / H2O2 295 *NF 900 1414
HTPB / N2O4 297 *NF 900 1440
HEDM+HTPB / O2 400 *NF 990 1227
RP-1 / O2 325 311 810 1227
*NF: Data Not Found

An HTPB/N
2
O hybrid rocket combines all the positive features of hybrid
motors and is the focus of this thesis research. Liquid nitrous oxide provides the

5
added benefit of being self-pressurized to about 700 psi at room temperature. The
self-pressurizing attribute eliminates the need to design a complicated oxidizer turbo
pump fed system, and instead can rely on a pressure fed technique to draw the
oxidizer into the combustion chamber. A pressure fed oxidizer system reduces the
overall weight of the launch vehicle and simplifies the rocket. Although,
performance potential is diminished due to the low density of N
2
O, an increase in
safety is gained by not dealing with cryogenic needs as in oxygen, to keep it
liquefied. HTPB also has an added safety benefit of being nearly inert in the absence
of an oxidizer. HTPB must be first vaporized in the presence of an atomized
oxidizer at a high temperature (~1000 K) in order for it to burn. Because an oxidizer
needs to be present to make HTPB burn, it is not a hypergolic mixture and therefore
can be transported and handled without special shipping and safety techniques
unlike solid rocket motor propellant (Story, et al., 2003). As a solid fuel, HTPB has
a rubberized consistency so unwanted cracking from impact damage is not an issue.
Most hybrid fuels have a nearly indefinite shelf life. The fuel casting process can
impart voids or bubbles into the fuel which can cause small performance or
instability problems during hot-fire but filling the chamber under vacuum conditions
is one preventative method. Unlike solid rocket motor propellant, cracks, voids, or
other defects in the propellant of a hybrid motor is not a catastrophic failure. All
these attributes give confidence that HTPB and N
2
O is a promising propellant
combination for hybrids.

6
1.2 OUTSTANDING PROBLEMS WITH HYBRID MOTORS
Hybrid rocket technology is fairly underdeveloped because of the lack in
consistent experimental data and understanding of key combustion characteristics.
Several parameters of hybrid motor performance include thermo chemistry
properties of fuel and oxidizer combination, fuel regression rate, ISP, thrust to
weight ratio, design of the fuel grain for complete burning, and combustion stability.
Many of these performance metrics have been observed experimentally but are not
well defined or repeatable, which will be addressed in more detail in Section 2.
Hybrid fuel regression rate, or the rate at which solid fuel burns in the radial
direction, is one major technical problem among several attributed to hybrid rockets.
Fuel regression is expected to be an integral measurement of hybrid rocket motor
combustion. Fuel regression rate contributes to overall motor performance, thrust,
and fuel usage and is defined as the velocity at which solid fuel vaporizes and burns
away. The vaporization of the fuel is directly dependant on the oxidizer flow rate or
the velocity of oxidizer over the solid fuel (Pacchioli, 1997). Fuel regression rate is
highly variable because of the many contributing factors such as, the
thermodynamic properties of the specific fuel and oxidizer combination, the
oxidizer to fuel ratio (O/F), the port geometry (size and number), the oxidizer
impingement angle on the solid fuel, the oxidizer atomization through the injector,
the turbulent flow or mixing within the fuel port, the mass flux or mass flow rate of
the oxidizer, the mass flux of the vaporized fuel, and the length of the motor with
respect to the port diameter (L/D). Knowing the exact regression rate for a given
hybrid rocket will determine the exact amount of fuel needed for a given burn time.

7
Furthermore, knowing the exact amount of fuel needed for a given burn time will
reduce the amount of unused fuel, which reduces the mass of the overall launch
vehicle and therefore reduces the cost of the mission. The University of Colorados
MaCH-SR1 2002-2003 hybrid motor project designed the fuel grain thickness three
times larger than needed for a 15 second burn time due to the unknown regression
rates. The motor only reached a thrust to weight ratio of 1.2, of which the fuel was a
contributing factor of the relatively large mass (MaCH-SR1, 2003). Typical thrust to
weight ratios for the space Shuttle Main Engine liquid oxygen/liquid hydrogen
system is 66 while the Shuttles Solid Rocket Booster is 2.5. SpaceShipOnes
launch vehicle with a single hybrid rocket motor has a thrust to weight ratio of about
2.1 which is capable of launching off the ground but was instead carried to 15
kilometer and launched from a carrier plane called the White Knight to boost the
ballistic trajectory height of the launch vehicle. These examples show the
importance of accurately defining the regression rate of the fuel to eliminate excess
weight that can be costly to the mission, knowing that the regression rate is hybrid
motor specific.
Compared to the well-known combustion analysis of solid and liquid
rockets, only a limited amount of comparative analysis and data has been recorded
for hybrid rocket combustion. In addition, there is an abundance of data for HTPB
and liquid oxygen hybrid rockets but limited information using N
2
O as the oxidizer
due to the proprietary nature of SpaceShipOnes hybrid rocket data. The focus of
most hybrid rocket studies has not involved the specific solid fuel and liquid
oxidizer combination, Hydroxyl-terminated Polybutadiene (HTPB) and Nitrous

8
oxide (N
2
O). Furthermore, the author knows of no published regression data for this
type of HRM. This thesis is designed to increase the understanding of this specific
hybrid rocket motor.
1.3 FUEL REGRESSION FOR HTPB/ N
2
O COMBINATION
Regression rate of a hybrid rocket motor can be presented in terms of either
the convective and radiative heat transfer from a diffusion-limited flame or the
kinetics of the gas and solid phase reactions. This thesis focuses on studying the
fuel regression rate of the lesser-known HTPB/N
2
O fuel and oxidizer hybrid rocket
motors. Most regression research uses a simplified regression rate formula to fit the
experimental data as seen in this equation , where the r is the regression
rate or the velocity at which the fuel burns toward the wall of the combustion
chamber, a is the constant including the blowing coefficient, n is the regression rate
oxidizer exponent, and G
n
o
aG r =

o
is the localized mass flux of the oxidizer in the hybrid
motor or oxidizer mass flow divided by the port area. Solid rocket motor regression
rate behavior is defined by r = aPc
m
which is shown to be dependant on chamber
pressure, Pc. Some hybrid rockets have shown pressure dependence when using
metalized fuels and defines regression rate as Pc
n
o
aG r =

m
(Sutton and Biblarz,
2003).
The solid fuel regression rate empirical constants, a and n, for non-metalized
hybrid rocket motor will be derived from experimental measurements of the
University of Colorado, Boulders MaCH-SR1 hybrid rocket motor tests from 2005.
Experimental fuel regression data from four MaCH-SR1 hybrid rocket tests was

9
measured along with other combustion parameters to determine the rate of fuel
burned for each hybrid motor. The results are compared to existing regression rate
empirical constants from HTPB/Oxygen and Plexiglas/Oxygen hybrid rocket
motors. The layout of the thesis is shown in Figure 2. Furthermore, a simplified look
into the chemical kinetics of HTPB/N
2
O is developed to determine thermodynamic
characteristics, equilibrium flame temperature, the ratio of specific heats, and
combustion products. Simple chemical kinetics codes are compared and an
approach is presented to determine the thermochemistry for the reactants HTPB, as
hydrocarbon molecule C
4
H
8
O
2
, and N
2
O. The fundamental thermochemistry
parameters, flame temperature and ratio of specific heats, of various reactants are
then used in the ideal rocket equations to help determine the performance or ISP of
the hybrid rocket motor. The chemical kinetics analysis is carried out using
available combustion software codes including, STANJAN, CHEMKIN, and GDL
Propep. STANJAN is used to determine the adiabatic flame temperature and the
ratio of specific heats. CHEMKIN is used to analyze equilibrium products by
assuming a few simple reaction mechanisms produced by specified reactants. The
thermodynamic parameters are used in the ideal rocket equations to establish a
correlation between the test data and the model. An understanding of the
fundamental chemical and thermodynamic properties of these reactants is important
to prove that hybrid rockets can produce high combustion performance needed for
space applications and are indeed safe rockets that can even be used for human
space flight.


10
Literature Survey on Hybrid Rockets
Hybrid Rocket Regression Rate Experiment
Topic Chosen
Fuel / Oxidizer Trade Study
Obtain Regression Rate Experimental Measurements
From MaCH-SR1 HRMs (Primary Thesis Objective)
MaCH-SR1 HRM Hot-fires
(2 Tests w/ Sensors)
Literature Summary
On Regression Rate
Chemical Kinetics Model to predict Regression Rate
(Secondary Objective)
Regression Rate Sensor Trade Study
Results Empirical Constants
(a & n) Determined
STANJAN & GDL Propep Oxidizer Comparison
(Chemical Equilibrium Programs)
Adiabatic Flame Temp and Gamma Determined
(Assumed Adiabatic, Constant Pressure System)
Empirical Constants Compared to
Limited Literature Database
Basic Thermochemistry Review
Temperature and Gamma used in Ideal Rocket
Equation to calculate thrust & verify measured r
Literature Survey on Hybrid Rockets
Hybrid Rocket Regression Rate Experiment
Topic Chosen
Fuel / Oxidizer Trade Study
Obtain Regression Rate Experimental Measurements
From MaCH-SR1 HRMs (Primary Thesis Objective)
MaCH-SR1 HRM Hot-fires
(2 Tests w/ Sensors)
Literature Summary
On Regression Rate
Chemical Kinetics Model to predict Regression Rate
(Secondary Objective)
Regression Rate Sensor Trade Study
Results Empirical Constants
(a & n) Determined
STANJAN & GDL Propep Oxidizer Comparison
(Chemical Equilibrium Programs)
Adiabatic Flame Temp and Gamma Determined
(Assumed Adiabatic, Constant Pressure System)
Empirical Constants Compared to
Limited Literature Database
Basic Thermochemistry Review
Temperature and Gamma used in Ideal Rocket
Equation to calculate thrust & verify measured r

Figure 2. Thesis Objectives
2. LITERATURE SUMMARY
A literature review has been researched for hybrid rocket diffusion flame
combustion theories, fuel regression equations with empirical data, historical tests of
hybrid rocket motors, scaling effects from lab-scale to full-scale HRM tests, and
chemical kinetics overview of hybrid propellant reactants. Until recently,
experimental tests for hybrid rockets have been almost exclusively done with
HTPB, PMMA and Plexiglas fuels and oxygen oxidizer propellant systems. The
purpose of this thesis is to extend this research to determine the regression rate
empirical constants of the HTPB and nitrous oxide propellant combination as well
as develop a basic review of the chemical kinetics within a hybrid rocket motor.

11
2.1 DIFFUSION COMBUSTION FLAME THEORY
Fuel regression analysis is formulated as a coupled fluids and heat transfer
problem. The fundamental theory of the turbulent boundary layer with an embedded
flame was developed by Marxman, et al. (1964) and applied to hybrid rocket motor
design. The theory is based on flat plate approximations and relates fuel regression
rate to heat transfer between the surface of the fuel and the flame inside the
boundary layer (Incopera and DeWitt, 1990). Hybrid combustion theory has been
improved upon in later years but has not strayed far from the analysis work
originally done by Marxman, et al. (1964).
The analytical approaches to hybrid rockets combine turbulent flow of
gaseous oxidizer over a solid fuel with convective, conductive and radiative heat
transfer. As the oxidizer flows over the solid fuel, a reactive boundary layer is
produced with a flame inside and positioned near the fuel. The flame provides heat
to the solid fuel which vaporizes the fuel. The fuel and oxidizer react and mix at the
flame front. The assumptions in hybrid rocket combustion are the Lewis number,
Le=1, the combustion occurs near the wall, and combustion only occupies a small
portion of the boundary layer so that the flame sheet model can be used. Hybrid
rocket motor combustion combines the momentum equation with the energy and
species equations to encompass the complexities of the combustion phenomenon
(Marxman, et al., 1964). A sketch of this diffusion flame combustion with mass
injection or vaporization from the solid fuel grain is shown below in Figure 3.

12

Figure 3. Hybrid Rocket Combustion - Boundary Layer with Embedded Flame
The temperature profile inside the boundary layer increases to its peak at the
flame zone. Below the flame zone, the temperature gradient decreases exponentially
toward the fuel surface and similarly decreases above the flame zone towards the
oxidizer flow and edge of the boundary layer. The velocity and temperature profiles
in this diffusion flame boundary layer above the flame are directionally opposite and
the profiles are similar below the flame (Marxman et al., 1964). The detailed
chemical kinetics of the solid fuel and liquid oxidizer reactions can be employed in
modeling regression rate to increase the accuracy of the prediction and thus increase
the performance of an HRM.
The engineering-related physics of a diffusion boundary layer with an
embedded flame have been studied and applied to hybrid rocket combustion.
Regression rate derivations have usually come from the heat transfer mechanism
from a diffusion-limited flame, suggesting that the aerodynamics of the turbulent
boundary layer at the solid fuel surface out weighs the kinetics mechanism of the

13
gas and solid phase. The following equations show the different regression rate
equations for theoretical and empirical models (Serin 2003 and Marxman, et al,
1964). Hybrid combustion rate is dominated by conductive heat transfer and not
radiation effects.
3 / 2
Pr


=
p
e e
v
t
f
u
h
h
C r

(1)
Equation (1) is the developed regression rate equation from flat plate
approximations where C
f
is the skin friction coefficient with mass injection related
to Reynolds number, and h is the heat of reaction divided by the heat of vaporization
when the fuel surface temperature is known. These two parameters make up the
blowing coefficient, , seen in equation (2). The blowing coefficient, also depicted
as C
f
, is defined as a non-dimensionalized rate of the vaporized fuel expelled from
the fuel surface and (Sutton and Biblarz, 2001). The third parameter in the
regression equation (1) is the ratio of the mass flow of the oxidizer to the density of
the fuel. Finally Prandlt number, Pr, is the turbulence parameter and it is a ratio
between the momentum and thermal diffusivity. In equation (2) turbulent flow is
assumed over the entire axial length of the hybrid rocket motor so Pr =1.
23 . 0 2 . 0
8 . 0
) ( 036 . 0

x
G
r
f
=

(2)
Equation (2) is a modified version of Eq (1) that reflects empirical data for
non-radiative, Plexiglas and oxygen hybrid combustion. Eq (2) shows that
regression rate is highly dependent on the total mass flow of the oxidizer and fuel
flow per unit area, G, at a given distance from the start of the fuel and less

14
dependant on the rate at which the fuel is burning off normal to the fuel grain, or
blowing coefficient.
The above equations describe the thermal and fluid dynamics used to
characterize fuel regression. The equations use the pyrolysis law which considers an
energy balance at the fuel grain surface and regression rate is controlled by heat
transfer from the flame to the surface. Equations (1) and (2) are obtained from an
idealized turbulent boundary layer combustion model for non-radiative systems with
flat plate approximations (Marxman, et al., 1964). The turbulent boundary layer
increases the mixing of the fuel and oxidizer at the flame front, which increases the
regression rate of the solid fuel. Radiation transfer is omitted in the two equations.
Marxman et al. (1964) show that radiation does not significantly contribute to
changing the magnitude of the regression rate for non-metallized grains, like
Plexiglas and HTPB without additives. Equation (2) incorporates turbulent flow
over the entire length of the fuel surface. The Reynolds analogy and Prandlt
number, Pr = 1, is valid outside of the flame. Chiaverini, et al., (2001) uses a similar
analysis approach.
Figure 4 represents the boundary layer and flame zone that develops in a
hybrid rocket with a single round port fuel grain (Karabeyoglu, et al., 2003).

15

Figure 4. Typical Hybrid Rocket Layout
The pre-combustion chamber is used to create circulation of oxidizer before flowing
into the fuel grain geometry. The post-combustion mixing chamber is needed to
have complete combustion of fuel and oxidizer before the gas is expanded in the
nozzle area to create thrust. The figure shows the boundary layer interactions due to
fuel grain geometry. More complicated fuel grain geometry can lead to an increased
complication in combustion effects and modeling.
A detailed characterization of the regression rate will determine the exact
amount of fuel needed for a given burn time. Hybrid motor tests can give detailed
regression rate for the specific hybrid system design. As shown earlier in Eq. (2)
fuel regression or vaporization of the fuel is directly dependant on the oxidizer flow
rate

(Marxman, et al., 1964). The mass flow rate is dependant on the area of the fuel
port geometry. All of the regression results derived from Marxmans analysis are
compared with experimental studies of oxygen flow over a flat plate of Plexiglas
fuel (Marxman, et al., 1964). Accurate and repeatable regression rate predictions
occur when all dependent parameters are unchanged from test to test.

16
Hybrid combustion theory can be applied to solid fuels that form a liquid
layer on top of the burning surface. The layer sketch as shown in Figure 5 includes
the hydrodynamic instability which produces droplets from the liquefying layer to
efficiently mix with the gaseous oxidizer stream and create a higher regression rate.

Figure 5. Pyrolysis of Liquifying Fuels, ie. Paraffin (Karabeyoglu, et al., 2002)
Karabeyoglu (2002) had discovered this during his research at Stanford. With high
enough oxygen mass flow rate, the paraffin wax melts fast enough to form roll
waves and droplets that burn rapidly, about three times faster than HTPB
(Karabeyoglu et al., 2002). Knowing the chemical characteristics of a fuel and
oxidizer will increase the accuracy of the fuel regression rate predictions. Rapid
chemical reactions like in paraffin wax and Plexiglas consume most of the solid fuel
quickly and thus increase the fuel regression rate. However, the structural integrity
of paraffin waxes is still being researched for large scale hybrids (Karabeyoglu, et
al., 2002).
2.2 CHEMICAL KINETICS
Chemical kinetics plays an important role in the modeling of a hybrid rocket.
Basic chemical kinetics can estimate equilibrium combustion products, adiabatic

17
flame temperature, molecular weight of the mixture, and reaction rates for
stoichiometric and non-stoichiometric mixtures. The complicated hybrid
combustion establishes the need to understand the chemical kinetics involved in
such systems to determine detailed regression rate analysis. Solid rocket motor burn
rates are highly dependant on reaction kinetics and are independent of the port
geometry. Hybrid rocket motors combine the heat transfer mechanisms coupled
with the kinetics mechanism in the solid and gas phase of combustion.
Serin and Gogus (2003) analyzed the chemical reactions of HTPB and O
2
.
HTPB is known to gasify into butadiene (C
2
H
4
) fuel and then into methane (CH
4
),
which is released at a surface temperature ranging from 913 K to 942 K (Serin and
Gogus, 2003) . The make up of the combustion products also included C
4
H
6
, CH
4
,
CO, O
2
, H
2
O, CO
2
, OH, O, C
2
H
4
, (Serin and Gogus, 2003). Branch and Cor

(1993)
have analyzed the reaction mechanism for the burning of CH
4
and N
2
O. This
approximation will be applied to nitrous oxide and is discussed in Section 3.
Activation energy, E, plays an important role on regression rate as shown in
the equation . Regression rate is exponentially dependent on
temperature which means that if the temperature is high then the regression rate is
faster. The Arrhenius constant, A, is an experimentally derived value that tells us
the thermochemistry of the reactants and is proportionally related to regression rate.
The data shown in Table 2 gives a brief understanding of how activation energy, E,
is related to regression rate of a specific hybrid rocket.
RT E
Ae r
/

=

18
Table 2. Chemical Kinetics of Hybrid Rockets
Fuel / Oxidizer
Flame
Temp
(K)
E
(kcal/mol)
Regression Rate
(Avg) (mm/sec)
HTPB / O2 3800 48.6 1.65
Paraffin / O2 750 17 3.85
Plexiglas / O2 3500 33 3.05
HTPB / N2O 3500 *NF 1.25
*NF: Data Not Found

The low E of paraffin wax shows that less energy is required to break the bonds and
vaporize the fuel. On the other hand, HTPB has a high E and thus more energy and
higher temperatures are required to vaporize the solid fuel, resulting in a lower
regression rate.
2.3 HISTORICAL HYBRID ROCKET EXPERIMENTS
Experimental techniques for measuring regression rates and determining
chemical kinetics in a hybrid rocket are described in the literature. Early hybrid
rocket motor experiments were performed by Marxman, et al., (1964) using oxygen
flow over a single plate of solid Plexiglas. Several decades went by without much
hybrid rocket interest until Boardman, et al., (1997) performed lab-scale and full-
scale HTPB/oxygen hybrid motor testing with funding from NASA called the
Hybrid Propulsion Development Program. It was found that fuel regression in lab-
scale hybrid rockets is not easily scaled up to large full-scale designs. Therefore,
regression rate values are primarily based on experimental test data from a
controlled laboratory environment (Boardman et al, 1997, Chiaverini, et al., 2001,
and Estey, et al. 1991). Several universities have developed their own hybrid rocket

19
experiments. Paraffin wax has been recently researched at Stanford and tested at
Ames Research Center (Figure 6) and provides a very high regression rate with
oxygen oxidizer. Unity IV was a joint project with Utah universities that built and
tested HTPB/N
2
O rockets (2001). However, no regression rate information was
published. The MaCH-SR1 student built rocket at the University of Colorado,
Boulder has been testing HTPB fuel hybrids since 2002 with liquid oxygen oxidizer
and N
2
O.


Figure 6. Ames Research Center Test of Paraffin/Oxygen Hybrid Rocket
(Karabeyoglu, et al. 2002)
Table 2 shows different regression rate equations from both theoretical and
empirical models. A wide range of exponents can be seen.

20
Table 3. Regression Rate Equations
Reference Description Regression Eqn Fuel / Oxidizer Eqn #
Marxman & Gilbert
Diffusion-limited
heat transfer
theory Plexiglas / Oxygen 1
Marxman & Gilbert
Improved version
of Eqn (1) for
turbulent flow and
non-radiative
systems Plexiglas / Oxygen 2
Serin & Gogus
Navier - Stokes
model using CFD-
ACE and
REGRESS
software HTPB / Oxygen 3
Thiokol Corp.
Small rocket
firings HTPB / Oxygen 4
Karabeyoglu , et. al.
Small rocket
firings Paraffin / Oxygen 5
3 / 2
Pr


=
p
e e
v
t
f
u
h
h
C r

15 . 0
39 . 0
o
G r =

23 . 0 2 . 0
8 . 0
) ( 036 . 0

x
G
r
f
=

62 . 0
488 . 0
o
Paraffin G r =

681 . 0
146 . 0
o
HTPB G r =


Significant differences can be seen in regression rates for different oxidizer and fuel
combinations, making the reactants a key parameter. A complete boundary layer
and combustion analysis of the oxidizer/fuel interactions is required to properly
simulate the time-dependant regression of the fuel grain. One approach requires the
use of numerical codes to calculate the flow field properties at all points. Typical
results are given in a model by Serin & Gogus (2003). Their rocket firing results and
theoretical analysis were compared with the three regression rate equations (1), (2),
and (3) in Table 3. The model results fit Eq. (3) best, while there are large
discrepancies within the Marxman & Gilbert regression equations due to the
different oxidizer and fuel combination and assumptions in the REGRESS software
(Serin & Gogus, 2003).
Averaged regression rates have been more recently calculated in the 1990s
during full scale testing of hybrid rockets, all using HTPB and liquid O
2
(Boardman,
et al., 1997). The vaporization of the fuel is directly dependant on the oxidizer flow

21
rate (Sutton and Biblarz,, 2001). Very few hybrid rocket motor tests measure the
instantaneous regression rate in a hybrid rocket using ultrasonic and x-ray methods.
The instantaneous data measured by Gramer and Taagen (2001) shows that
regression rate decreases as the burn time increases.
A series of HTPB/O
2
small rocket hot fires were conducted at the Thiokol
Corp (Sutton and Biblarz, 2001). The data collected was used to derive the
following equation in the form of . This simplified regression rate equation
is the most fundamental regression equation from pyrolysis (mass conservation at
fuel surface), where a and n are derived from empirical data. This equation has
shown up in literature as a way to model regression rate but is very hybrid motor
system specific where changes in motor size, fuel grain geometry, fuel and oxidizer
combination, oxidizer flow rate, and combustion chamber pressure. Eq. (4) is a
result of this study.
n
o
aG r =

681 . 0
146 . 0
o
HTPB G = r

(4)
where G
o
is the oxidizer mass velocity, with a and n as free constants fit by
experimental data. The discrepancy between this regression rate equation and Eq. 1
relates to the flat plate approximations used by Marxman, et al (1964). A similar
regression equation was developed by Karabeyoglu, et al. (2002) for paraffin wax,
62 . 0
488 . 0
o
Paraffin G r =

(5)
Note that the regression rate for both Eq (4) and Eq (5) is in mm/sec and the
oxidizer mass flux is in g/cm
2
. Figure 7 compares the regression rates of both

22
Paraffin/GOX and HTPB / GOX hybrids. Paraffin wax is shown to have a much
higher regression rate. However, Paraffin has inherent disadvantages which include
slumping of the fuel during storage, and non uniform burning or chugging.

Figure 7. Regression Rate Data for Paraffin and HTPB Fuels (Karabeyoglu, et
al., 2002)
A N
2
O/HTPB hybrid rocket, which combines all the positive features of
hybrid motors, is a new oxidizer and fuel combination that has been researched
extensively by SpaceDev of Poway, California for the privately funded
SpaceShipOne launch vehicle. The data collected at SpaceDev is proprietary.
Lockheed Martin also developed a N
2
O/HTPB sounding rocket (Arves, et
al., 1997). Several university projects, including the Unity hybrids in collaboration
with the universities in Utah (States, 2001 and Horton, et al., 1996) tested HTPB
and N
2
O but no regression data was reported.

23
Several experimental hybrid rocket tests using the HTPB/N
2
O propellant
were performed as part of the MaCH-SR1 hybrid rocket senior design project. The
tests are part of the MaCH-SR1 (Multi-disciplinary University of Colorado High-
altitude Student Rocket) hybrid rocket project is designed, built and tested by senior
design students in the Aerospace Engineering Department at University of
Colorado, Boulder. The continuous MaCH-SR1 senior design project, which began
in 2002, has successfully tested one 100lb
f
, seven 300lb
f
, one 1000 lb
f
, and one
5000lb
f
thrust hybrid rockets at various establishments in Colorado. The MaCH-
SR1 2004-2005 team hot fired four 300 lb
f
thrust hybrid rockets at the Lockheed
Martin facility in Littleton, Colorado. The MaCH-SR1 student built hybrid rockets
are increasing the amount of experimental data for HTPB/ N
2
O rockets, which is
important for the rocket community (MaCH-SR1, 2005).
Additional experimental data is required to establish trends and fully
characterize the performance for different sized hybrid motors with different fuel
and oxidizer combinations as well as different fuel geometries. An understanding of
the trends is especially important when adapting the engineering practice of using
lab scale tests results to establish full scale predictions, such as in testing a 5000 lb
f

thrust lab scale hybrid rocket motor to predict a 250,000 lb
f
thrust full size hybrid
rocket motor, where fuel regression variations become much more apparent
(Boardman, et al., 1997). Fuel regression rate is shown to be difficult to predict and
is non-repeatable when comparing lab scale tests to full scale tests (Boardman, et al.
1997 and Swami and Gany, 2003). A common way to predict scaling effects is from
a simplified engineering approximation approach by using non-dimensional

24
parameters, which is not a true representation of the complex combustion in a
hybrid motor

(Swami and Gany, 2003). Characterizing fuel regression rates from lab
scales to substantially larger full-scale designs, is addressed but mostly from an
engineering viewpoint rather than analysis and does not produce consistent trends
(Boardman, et al., 1997, Swami and Gany, 2003 and Estey, et al., 1991).
2.4 DESIGN CALCULATIONS
The characteristics of fuel combustion can be described in terms of fuel
grain regression rate and the mass flux of the fuel. First, the fuel grain regression
rate is used to determine that the design will maintain a desirable geometry and
thrust profile, for as much of the burn as possible, and that no burn-through of the
combustion chamber will occur. Second, the mass-flux of the fuel is determined by
the oxidizer flow rate and exposed perimeter of the port. The mass flux determines
how fast the port will regress and the value of the oxidizer to fuel (O/F) ratio. The
O/F ratio directly affects combustion chamber temperature and thus, thrust and
specific impulse. The oxidizer/fuel ratio trade study in Table 4 incrementally
changes input parameters to achieve a wide range of O/F ratio and related ISP. An
O/F of 4 gives a desirable ISP value while keeping the chamber temperature to a
minimum.

25
Table 4. Oxidizer/Fuel Ratio vs. Chamber Temperature and Specific Impulse
O/F Chamber Temperature (K) Isp (s)
2 1673 187.3
2.5 1892 197.1
3 2109 204.3
3.5 2470 216.3
4 2758 222.5
4.5 2980 224.7
5 3141 224.9
5.5 3244 224.2
6 3302 223.2
6.5 3329 222.1
7 3337 221
5.0g Castor Oil
O/F x 100g N2O
All Calculations performed in GDL Propep
83.56g HTPB
10.34g IPDI
.09g Carbon Black

Experimental measurements in lab-scale and full-scale hot fires have been
used to identify regression rate values, fuel and oxidizer combinations, and port
geometry in a hybrid motor. However, the regression rate is primarily controlled by
the designed oxidizer mass flux and also changes with length of the chamber as seen
in Equations (6) and (7). The following simplified empirical equations for
regression rate (Incopera and DeWitt, 1990) are often used

n
o
aG r =

( )
s
m
(6)
x
n
o
aG r =

m
(7)
where a is the an empirical parameter that combines the blowing coefficient, x
dependency, fuel density, and gas viscosity into one constant. G
o
is the oxidizer
mass flux, which is equal throughout the entire test to the oxidizer flow rate divided
by the combustion port area. The value of x is the length of the chamber, and n and

26
m are the empirically derived exponents. The coefficient a, and exponents n and m
vary according to the fuel and oxidizer combination used. In nearly all cases, they
must be determined experimentally, and the scaling from one thrust level to another
is very non-linear, though it is dependent on the port geometry used (Incopera and
DeWitt, 1990). The regression parameter is used to size the diameter of the grain
and for geometry analysis to determine how the grain will perform throughout the
burn.
The main design parameter is the oxidizer mass flux through the chamber at
the beginning of the burn. Mass flux is given by Equation (8) from Sutton and
Biblarz (2001).
port
ox
o
A
m
G

s m
kg
2
(8)
The mass flux parameter is controlled by varying the mass-flow rate of the oxidizer
and the port area. In the case of a multi-port, the port area is simply the area of an
individual port multiplied by the number of ports. The mass flux parameter also
determines the O/F ratio.
Finally, given the port geometry, mass flow rate of fuel, fuel density, and
regression rate, the fuel grain length can be determined with the equation below
(Sutton and Biblarz, 2001),

p f
f
S r
N m
L

( ) m (9)
where N is the number of ports, and S
p
is the circumference of an individual port.

27
The final performance metric is the mass of fuel required, determined by
considering both the initial volume of fuel and the final volume of fuel. The optimal
design will offer the lightest overall combination of chamber and fuel grain. Also, a
very wide combustion chamber is unacceptable for drag reasons. However, a trade
study would be needed to determine the drag on a wide combustion chamber versus
a longer combustion chamber with excess (heavier) fuel.
Fuel regression characteristics can be analyzed in more detail using the
results from experimental testing of lab scale and full scale hot fires. To aid in the
fuel regression analysis for this current study, low-cost measurement techniques are
implemented. These measurements include thrust, pressure, mass flow rate, and
local regression rate with thermocouples. In addition, the fuel grain can be weighed
and measured before and after each hot fire to evaluate the average fuel regression
rate. These regression observations for the lab scale motor tests are noted in the
experiment section.
3. CHEMICAL KINETICS MODEL

Several analytical models for determining the thermochemistry and
equilibrium compositions of reactants are currently being used and most are
supplied as free software that can be found on the internet. Basic chemical kinetics
models can estimate equilibrium combustion products, adiabatic flame temperature,
molecular weight of the mixture and reaction rates for stoichiometric and non-
stoichiometric mixtures. This research uses STANJAN, CHEMKIN, and GDL
Propep software, developed by Stanford University, Kee (1985) and Martin Marietta
(now Lockheed Martin) respectively. STANJAN uses the element potential method

28
to find the chemical equilibrium composition of specified reactants. The key outputs
of this software and the outputs used in this research is the adiabatic flame
temperature of the combustion and molecular mass or weight of the products.
CHEMKIN uses elementary reaction mechanism to generate a table of temperature
and species mole fractions as a function of time. CHEMKIN also shows the
importance of each reaction in the mechanism at each time step relative to each
other.
The current study focuses on a hybrid rocket with a Hydroxyl-terminated
Polybutadiene (HTPB) fuel and a nitrous oxide (N
2
O) oxidizer combination and
establishes fundamental chemical kinetics and thermodynamic quantities such as,
equilibrium equations, reaction rates and rate constants. Simple reaction
mechanisms were inputted to see the effects of HTPB and N
2
O combustion. The
focus of most hybrid rocket studies does not involve this specific solid fuel and
liquid oxidizer combination, consequently the specific thermochemistry details are
not well known.
3.1 STANJAN MODELING

STANJAN uses the element potential method to find the chemical
equilibrium composition of specified reactants. The element potential method used
in STANJAN finds the Gibbs function for the system but only at the minimum state.
Thermochemsitry data must be inputted into the software before inputting the
reactants. The Gibbs function is defined as G = H-Ts. The second law shows that
the Gibbs function is true for decreasing isothermal, isobaric and fixed mass
systems in absence of all work (Cengel and Boles, 1994).

29
STANJAN software is used to calculate the adiabatic flame temperature and
equilibrium composition of the products at different equivalence ratios for two trials
of reactants, a simplified form of HTPB and Formaldehyde. The equivalence ratio,
, is a non-dimensional parameter that defines a mixture as stoichiometric
(balanced equation of fuel and oxidizer reactant mixture assuming complete
combustion), fuel rich, or oxidizer rich. The inputs to STANJAN are the specified
fuel and oxidizer originally at 298 K and 34 atm (500 psi), which is the chamber
pressure assumed in the 300 lb
f
MaCHSR1 hybrid rocket motor. An adiabatic and
constant pressure system was assumed to produce the following results.
Two trials of STANJAN results have been carried out. The first trial inputs
the reactants as a simplified form of HTPB with a chemical formula of C
4
H
8
O
2

which is the chemical composition for Plexiglas (PMMA) and N
2
O. Equation (10)
shows the stoichiometric mixture for HTPB (C
4
H
8
O
2
)

from Wangmin (2002) and
nitrous oxide (N
2
O) that is used in the Stanjan analysis.
C
4
H
8
O
2
+ 10 N
2
O 4 CO
2
+ 4 H
2
O +10 N
2
(10)
A stoichiometric mixture with complete combustion is an approximation
because optimal hybrid combustion reacts in a fuel rich mixture ratio. Dissociation
of the products occur and change the combustion flame temperature. However,
assuming that the equivalence ratio, , is unity is a good start to establish the
combustion characteristics of this specific fuel and oxidizer. Figure 8 shows the
adiabatic flame temperature of the above reactants versus equivalence ratio.

30
C4H8O2-N2O ADIABATIC FLAME TEMPERATURE VS. EQUIVALENCE RATIO
3800
3900
4000
4100
4200
4300
4400
0.6 0.8 1 1.2
EQUIVALENCE RATIO
T
E
M
P
E
R
A
T
U
R
E

(
K
)


Figure 8. C
4
H
8
O
2
-N
2
O Adiabatic flame temperature
Figure 9 shows the mole fraction of the products for the different
equivalence ratios.
C4H8O2 - N2O MOLE FRACTION VS. EQUIVALENCE RATIO (MAJOR SPECIES)
N2/10
CO
H2O
H2O
O2
H2
CO2
2.00E-02
4.00E-02
6.00E-02
8.00E-02
1.00E-01
1.20E-01
0.7 0.8 0.9 1 1.1 1.2 1.3
EQUIVALENCE RATIO
M
O
L
E

F
R
A
C
T
I
O
N
,

X
i

Figure 9. Major product species for optimized C
4
H
8
O
2
-N2O fuel - oxidizer
combustion

31
The second trial inputs the reactants as formaldehyde that has a chemical
formula of CH
2
O. As specified earlier, HTPB has been empirically found to quickly
decompose into butadiene, methane and then formaldehyde. Using formaldehyde
gives a more accurate result when finding the adiabatic flame temperature and
equilibrium compositions. Figure 10 shows the adiabatic flame temperature of the
above formaldehyde and nitrous oxide versus equivalence ratio.
CH2O-N2O ADIABATIC FLAME TEMPERATURE VS. EQUIVALENCE RATIO
3720
3740
3760
3780
3800
3820
3840
0.4 0.6 0.8 1 1.2 1.4 1.6
EQUIVALENCE RATIO
T
E
M
P
E
R
A
T
U
R
E

(
K
)


Figure 10. Formaldehyde (CH
2
O)/ N
2
O Adiabatic flame temperature
Figure 11 shows the mole fraction of the products for different equivalence
ratios.

32
CH2O - N2O MOLE FRACTION VS. EQUIVALENCE RATIO (MAJOR SPECIES)
H2O
O2
H2
N2/10
CO
CO2
O2
0.00E+00
2.00E-02
4.00E-02
6.00E-02
8.00E-02
1.00E-01
1.20E-01
0.5 0.6 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4 1.
EQUIVALENCE RATIO
M
O
L
E

F
R
A
C
T
I
O
N
,

X
i
5

Figure 11. Major product species for optimized CH
2
O-N
2
O fuel - oxidizer
combustion
The adiabatic flame temperature for formaldehyde and nitrous oxide more
closely resembles the flame temperature in the literature (Serin 2003) and also
matches the flame temperature and molecular weight the GDL Propep software
produces. The differences between the two different models are small. More
complex reactions need to be taken into account that is not possible to model in
STANJAN.
3.2 CHEMKIN MODELING
CHEMKIN is a useful computer code that has been widely used in the
combustion industries. CHEMKIN software was used in this research to compare
the species mole fractions by assuming a few simple reaction mechanisms produced
by the reactants HTPB, as hydrocarbon molecule C
4
H
8
O
2
, and N
2
O. Since the
complex details of hybrid combustion of HTPB/N
2
O is beyond the scope of this

33
research, two simplified reactions were chosen and an adiabatic and constant
pressure system was assumed that fundamentally represent HTPB / N
2
O combustion
in a hybrid rocket. The simple reaction mechanisms are described in the equations
below.
C
4
H
8
O
2
+ 10 N
2
O 2 CH
2
O + 2CO + 2 H
2
+ 8 N
2
O + N
2
(11)
CH
4
+ 10 N
2
O 2 CH
4
+ 2CO + 10 N
2
O (12)
The two equations signify the simple but important reaction mechanisms that are
assumed to be significant in the combustion of HTPB / N
2
O.
The following trials input the reactants as a simplified form of HTPB with a
chemical formula of C
4
H
8
O
2
, which is similar to the chemical composition for
Plexiglas (PMMA), and N
2
O (nitrous oxide). The first trial decomposed HTPB into
formaldehyde and carbon monoxide and other products as shows in Eq (11). Figure
12 shows the adiabatic and constant pressure results of this equation.

34
0.00E+00
1.00E-01
2.00E-01
3.00E-01
4.00E-01
5.00E-01
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
0
500
1000
1500
2000
2500
3000
3500
4000
H2
O2
H
O
OH
HO2
H2O2
H2O
CO
CO2
CH4
CH3
CH2O
HCO
CH3O
N2O
NO
N2
N
T[K]

S
p
e
c
i
e
s

m
o
l
e

f
r
a
c
t
i
o
n
s

Time (ms)
Figure 12. Species mole fractions and temperature as a function of time for an
adiabatic and constant pressure results for Eq. (11) reaction mechanism for
HTPB / N
2
O combustion
The adiabatic flame temperature or the temperature at which the reaction
mechanism reaches equilibrium is at 3520 K. Equation (11) reaction mechanism
gives a higher flame temperature than what has been determined by other chemical
kinetics methods.
The second trial decomposes HTPB into methane and carbon monoxide and
nitrous oxide as shown in Eq (12). Figure 13 shows the adiabatic and constant
pressure results for this equation.

35
0.00E+00
1.00E-01
2.00E-01
3.00E-01
4.00E-01
5.00E-01
6.00E-01
7.00E-01
15 17 19 21 23 25 27 29 31 33 35
0
500
1000
1500
2000
2500
3000
3500
4000
H2
O2
H
O
OH
HO2
H2O2
H2O
CO
CO2
CH4
CH3
CH2O
HCO
CH3O
N2O
NO
N2
T[K]

S
p
e
c
i
e
s

m
o
l
e

f
r
a
c
t
i
o
n
s

Time (ms)
Figure 13. Species mole fractions and temperature as a function of time for an
adiabatic and constant pressure results for Eq. (12) reaction mechanism for
HTPB / N
2
O combustion
The combustion case for Eq (12) shows the consumption of CH
4
and N
2
O
and the production of N
2
, H
2
O, CO and CO
2
among others. The temperature reaches
equilibrium at 3350K which correlates to the GDL Propep combustion temperature
for the specific HTPB fuel mixture and N
2
O oxidizer known include a more
accurate account of the HTPB fuel mixture chemical formula.
3.3 MODEL COMPARISONS AND CONCLUSIONS

The combustion thermochemistry of HTPB and several oxidizers is analyzed
and compared in this section. GDL Propep is a rocket combustion software program
developed at The Martin Marietta Company, now Lockheed Martin, and is available
as freeware (Lanier 1999). GDL Propep software was used to compare the
thermochemistry calculations for a hybrid rocket and validate the simplified

36
thermochemistry inputs given to STANJAN and CHEMKIN codes. GDL Propep
has an extensive fuel and oxidizer thermodynamic database that includes many of
the detailed ingredients used in hybrid rocket motors. This software program was
used to compare the results from STANJAN and CHEMKIN codes and also
compare HTPB with different oxidizers at an O/F ratio of 4. Table 5 shows the
HTPB fuel ingredients used by the MaCH-SR1 team. These same ingredients along
with different oxidizers were inputted into GDL Propep to find the combustion
temperature (Tc), Molecular Weight (M), and Specific heat ratio (Cp/Cv).
Table 5. HTPB fuel mixture ingredients used in all MaCH-SR1 HRM hot-fires
HTPB 84.57%
IPDI 10.34%
Castor Oil 0.09%
Carbon black 5.00%
TOTAL 100.00%
HTPB Fuel Ingredients (% by Mass)

Table 6 shows the inputs for the GDL Propep software.
Table 6. GDL Propep Input for HTPB and Nitrous oxide with an O/F = 4
Parameter Val ue
HTPB - R45 84.57 g
IPDI 10.34 g
Castor Oil 0.09 g
Carbon black 5.00 g
Nitrous Oxide 363.58 g
Chamber Pressure 500 psia
Ambient Pressure 12.2 psia
Ambient Temperature 298 K
Combustion Ingredients

The design chamber pressure for the MaCH-SR1 (2005) hybrid rockets is 500 psia,
the ambient pressure represents the pressure at Denver, Colorados 5250 elevation
and ambient temperature is assumed to be 298 K.

37
Table 7 shows the results from GDL Propep with HTPB R45M fuel and
three different oxidizers, liquid nitrous oxide, gaseous oxygen, and liquid oxygen.
Table 7. GDL Propep Results and Oxidizer Comparisons
Parameter Value
Combustion Temperature (Tc) 2718 K
Cp / Cv () 1.282
Molecular Weight (M) 22.46 gm/mol
c* 4912.3 ft/sec
Isp 229.3
Combustion Temperature (Tc) 3627 K
Cp / Cv () 1.214
Molecular Weight (M) 26.73 gm/mol
c* 5324.4 ft/sec
Isp 252.1
Combustion Temperature (Tc) 3667 K
Cp / Cv () 1.217
Molecular Weight (M) 26.41 gm/mol
c* 5380.3 ft/sec
Isp 254.6
HTPB and Gaseous Oxygen
HTPB and Nitrous Oxide
HTPB and Liqui d Oxygen

The above table shows that the combustion temperature for nitrous oxide is about
1000 degrees lower than that of oxygen. As a result to the lower combustion
temperature, the ISP and c* values are also lower. The c* parameter is described as
the combustion efficiency and is derived from the ideal rocket equations. This
information is good comparison with the STANJAN and CHEMKIN temperature
results. It also coincides with the STANJAN Cp/Cv ratio for formaldehyde and
nitrous oxide. GDL Propep also outputs several combustion performance
calculations for rockets. These are also supplied in the table for later comparison
with the actual measured data from the MaCH-SR1 hot fire. It can be noted that the
results from GDL Propep in the current study compare to the results from MaCH-
SR1 (2005).

38
The simplified thermochemistry inputs into STANJAN and CHEMKIN were
validated by the GDL Propep software. STANJAN produced a higher adiabatic
flame temperature than that of CHEMKIN by 500 degrees and GDL Propep by
1000 degrees in temperature. However, STANJAN and CHEMKIN codes are still
valid tools for calculating combustion thermochemistry because they can
incorporate detailed reaction rates that are inherently prominent in the combustion
of specific oxidizers and fuels. Additional research into the details of gaseous
produces can be performed to understand the reactions between HTPB and N
2
O. A
small plug of fuel can be burned in a simple combustion chamber with a uniformly
mixed bath of nitrous oxide. The detailed measurements of the combustion products
will help anchor the chemical kinetics models. GDL Propep is a canned code that
includes not only HTPB but the IPDI and Carbon black ingredients that may
contribute to this difference in temperature. Carbon black is added to the HTPB fuel
to reduce radiative effects from the flame to the fuel. IPDI added as a bonding agent
for the HTPB polymer (MaCH-SR1 2004) also contributes to the complexity of the
HTPB chemical formula and thus adds additional complexity to the chemical
kinetics and combustion products which could not be modeled in STANJAN or
CHEMKIN.
4. EXPERIMENT
The main objective of the hybrid rocket motor hot fires for this current stuffy
is to obtain experimental measurements of the solid fuel to identify a regression rate
equation for a HTPB/N
2
O HRM. Embedded thermocouples in the solid HTPB fuel
that lines the combustion chamber will help determine a more accurate regression

39
rate measurement. An average regression rate for the entire burn will also be
measured by weighing the combustion chamber before and after the hot fire. These
measurements will be compared to average regression rate equations for different
fuel and oxidizer combinations and different fuel grain geometry found in the
literature.
The flight configuration of a typical pressure fed hybrid rocket motor is
shown in Figure 14. Using a self-pressurizing liquid oxidizer simplifies the oxidizer
system to an oxidizer tank and a pressurant tank. The pressurant tank, often filled
with Helium, pressurizes the oxidizer tank throughout the hot fire to keep the
oxidizer tank inlet pressure the same and thus keeping the same mass flow of
oxidizer into the combustion chamber throughout the hot fire. The combustion
chamber includes the HTPB fuel port with a specific geometry needed to produce
the desired thrust profile and regression rate for the hot fire duration.
Chamber
(with HTPB Fuel)
Oxidizer
(N
2
O) Tank
Pressurant
(He) Tank
Nozzle

Figure 14. Hybrid Rocket Flight Configuration

40
4.1 EXPERIMENTAL DESIGN AND SETUP
Regression rate determination is the important parameter being measured in
the hybrid rocket motor experiments performed by the MaCH-SR1 senior design
project team. A predicted regression rate steady state analysis is performed with test
data from the previous year with a single, round port fuel geometry to give a
conservative prediction. A trade study is performed on different ways to measure
regression that will obtain accurate localized regression rates at different axial
points along the combustion chamber. For the 4-point star fuel port configuration,
both star peak and trough regression rates will be analyzed and differences are
expected.
The fuel geometry can change the performance characteristics of a hybrid
rocket motor. A trade study on different fuel geometries is performed to determine
which geometry gives the most efficient use of the fuel and still provides high
performance and low fuel mass. A visualization method of the regression analysis
was performed by the MaCH-SR1 team (2004) using Solidworks, by regressing the
fuel grain contours at the predicted regression rate over two-second intervals, as
shown in Figure 15, Figure 16, and Figure 17. The fuel regression visualization
assumes uniform regression even for irregular geometry. Hybrid rocket motor
regression results show that uniform regression is not the case for star shape
geometry. A star shape fuel grain has two different regression rates, one for the star
points and one for the troughs. Differences in star and trough regression rates are
confirmed with hot fire data and the results are presented in the Experiment Results
section.

41

Star Trough
Star Peak
Figure 15. Full-scale Star Port Fuel Regression Characteristics (MaCH-SR1
2004)


Figure 16. Lab-scale Mutli Port Regression Characteristics (MaCH-SR1 2004)


42


Figure 17. Lab-scale 4-Point Star Port Fuel Grain (MaCH-SR1 2005)

The previous figures give an approximate regression characteristic of the star and
multi port geometries. Notice that the multi port grain has large unusable fuel once
the port diameters burn to their maximum. Excessive unburned fuel is undesirable
because it increases the weight of the rocket. The approximate regression rate
visualization allows the mass flux, internal surface area, and required port length to
be simulated. However in an actual hot fire differences in localized regression rates
have been observed. For example, the forward end of the chamber will often regress
faster than aft end of the chamber as the O/F ratio changes with axial position. The
4-point star fuel port configuration was chosen and tested for this current study due
to the simplicity of the port design over the multi-port and efficient use of the fuel
for the entire burn.
The MaCH-SR1 team has not previously experimentally tested the 4-star
fuel port geometry; therefore an approximate regression rate of 0.056 in/sec from
previous MaCH-SR1 experimental results with a round fuel port is used to design
and size the star fuel grain with conservatism. The star port is expected to regress

43
faster than a round port design. The O/F ratio of the combustion is designed for 4,
yielding an optimal ISP without a large increase in combustion temperature, as
shown previously in Section 3. Using the desired O/F in the ideal rocket equations,
the designed fuel flow rate is calculated to be 0.26 lbm/sec and oxidizer flow rate is
1.00 lbm/sec. The fuel port length is 17.5" long as stated in the MaCH-SR1 teams
final report (2005). The 4-point star is chosen for the efficient use of the fuel. The
star shape eventually burns the fuel into a cylindrical shape towards the end of the
hot fire and little unused fuel is left over, as seen in Figure 17. The 4-point star also
allows for easy placement of the different regression rate sensors. The sensor
arrangement for the baseline motors is discussed further in the Experimental Setup
section.
A regression rate sensor trade study is developed to determine the sensor to
produce most accurate measurement, easiest to assemble into the combustion
chamber and cheapest to use in the hybrid rocket motors. The literature is scanned
to determine methods previously used. The X-ray method used by Evans, et al.
(2003) at the Pennsylvania State University to measure real-time x-ray radiography
measurements using a novel X-Ray Transparent Center-perforated hybrid rocket
motor system. Both ultrasound pulse-echo system and real-time X-ray radiography
system was used by Chiaverini, et al. (2000) to obtain instantaneous measurements.
Both of these measurement techniques need extensive development of special
hybrid motor combustion chambers and the equipment set-up is highly expensive
making it not a feasible option for this thesis project. Next, MIRRAS regression rate
sensors developed by Gramer and Taagen (2001) use a thin sensor embedded into

44
the fuel to obtain regression measurements at one axial location per sensor. Several
sensors are setup in parallel at equal spacing and held together by a plastic coating.
The sensor erodes away as the fuel regresses and a step function is recorded where
each sensor breaks and the next starts recording. This gives real-time measurements
of regression rate. Six sensors can be recorded using one Orbitec data acquisition
system. The MIRRAS sensors have been used in small hybrid rockets. The sensors
were not chosen due to their high cost, separate DAQ system needed, and
unavailability for our scheduled hot fires at Lockheed Martin that could not be
delayed. NANMAC eroding thermocouples were pursued because they acted like
thermocouples in parallel and were reasonably prices and available for out testing
timeframe. The sensors measure real time regression and temperature effects.
Unfortunately, the sensors are not proven technology so some risk was inherent in
using this measurement technique.

Figure 18. Nanmac Eroding Thermocouple
The k-type thermocouples were also used in conjunction with the eroding
thermocouples. The k-type thermocouple method has been used often in combustion
of solid fuel and was meant as back-up. The thermocouple wires are thin and fragile
and will often break during placement in the fuel or while the motor is being
transported to the test stand which makes them hard to work with. However, they
are simple and effective measurement devices.

45

Figure 19. K-type Thermocouples
Finally, the average regression measurement technique from weighing the
combustion chambers before and after hot fire was employed to get an overall
picture of fuel regression for each test. This is the most popular method of recording
regression rate and is often presented this way in the literature so it is a comparable
data point.
4.2 EXPERIMENTAL SETUP
The key objectives for the hybrid rocket experiment are to obtain
measurements to identify regression rate characteristics for star shape fuel geometry
and HTPB/N
2
O fuel and oxidizer combination. The hybrid rocket motor tests are
instrumented to measure injector inlet and combustion chamber temperature and
pressures, nozzle casing temperature, solid fuel temperatures, oxidizer mass flow
rate, and hybrid rocket thrust. The thermocouples embedded in the solid fuel that
lines the combustion chamber will help determine a more accurate regression rate
measurement at determined axial locations down the length of the combustion
chamber. Average regression rate measurements will be assessed measuring the

46
change in mass of the fuel in the combustion chambers before and after hot fire. The
average regression rate, often used in the literature, gives a general view of the
performance of a specific hybrid rocket motor. The measurements recorded will be
compared to regression rates in the literature and observations will be made for fuel
and oxidizer combination, fuel geometry, and size of rocket.
The special regression rate instrumentation for the MaCH-SR1 baseline
engines are as follows:
(2 3) NANMAC eroding thermocouple sensors per hybrid
rocket to measure fuel and flame temperature
(6) K - type thermocouples per hybrid rocket to measure
regression rate at specified axial locations
The baseline engine with the two eroding thermocouples is labeled CG engine,
while the engine with three eroding thermocouples is labeled CGC engine. The
locations of the regression sensors are shown in the Figure 20. Three axial locations
are chosen to determine differences of regression at those points. The numbers
shown in the axial images of the fuel grain give the depth, in inches, from the
injector end of the fuel. The CGC engine has thermocouples axially positioned at 5,
7 and 14 inches, while the CG engine thermocouples are positioned at 5, 12 and 14
inches.
Two k-type thermocouples are embedded at each axial location along with
an eroding thermocouple expect at one location as indicated in the images. The k-
type thermocouples can measure moderate temperatures up to 1000 F. The
thermocouples were located at either star points or troughs to observe differences in
regression rate and fuel temperature at these different geometric features. Radial

47
effects were also considered. The standard thermocouples were embedded in the
fuel as pairs. Each pair was radially positioned with one thermocouple near the fuel
surface and the other 0.2 inches further embedded into the fuel grain. The fixed
radial spacing of 0.2 inches was kept for all pairs at each axial location for
regression rate measurements. When the fuel burns away and reaches the
thermocouple the temperature rapidly rises until the thermocouple fails. The time
between failures of each pair of thermocouple with fixed spacing gives the local
regression rate. The different axial locations will help determine where the fuel
regresses the fastest; near the injector, near the nozzle or the middle section.
The NANMAC eroding thermocouples were used to continuously measure
high temperatures inside the solid fuel and inside the fuel port to measure fuel,
combustion gas, and flame temperatures. The eroding thermocouples use a stainless
steal cover to shield the two thermocouple wires. The cover is designed such that
the thermocouples leads erode at a specific temperature, and as it erodes the
thermocouple bead at the tip of the sensor is continuously renewed. The NANMAC
sensors use a NPT threaded fitting that is secured and sealed through the
combustion chamber walls. The sensor length is also adjustable which
accommodates for complications in different fuel thickness, like the star shape fuel
port geometry. For the baseline engines, two types of eroding thermocouples are
used. The G-type thermocouple is made of tungsten and tungsten-26% Rhenium
material and can measure temperatures up to 6000 F. The G-type thermocouples
are designed to measure combustion gas temperatures for the HTPB and nitrous
oxide rocket. The C-type thermocouple is made of Tungsten-5% Rhenium and

48
Tungsten-26% Rhenium and is more accurate at temperatures below 3000 F. The
C-type thermocouple is designed to burn with the HTPB fuel providing
instantaneous measurements of the burning surface. Five eroding thermocouples
were used and distributed between the two baseline engines, due to cost restrictions.
Two of the eroding thermocouples used are G-type and three are C-type.
The CGC engine uses two C-type eroding thermocouples and one G-type.
The G-type eroding thermocouple is positioned just past the inner dimension of the
solid fuel grain at 14 inches from the injector end. The particular position of this G-
type thermocouple allows for direct measurement of the combustion gases in the
center of the combustion zone. The two C-type thermocouples are positioned
adjacent to each other in a trough or peak of the fuel grain. This positioning of the
thermocouples allows for identifying differences in how the peak and trough areas
of the fuel grain erode. A similar configuration is used in the CG engine except the
G-type thermocouple is placed just below the surface of the solid fuel to help
identify the boundary layer temperature changes. The locations of the regression
sensors are shown below in Figure 20. In these figures, both axial and radial
positioning for each sensor is labeled. The placement of the sensors gives the most
efficient use of the limited sensors due to limited funding. In addition, these sensors
could only be used in the two baseline hybrid motors due to safety issues on the low
weight, thinner walled, composite and aluminum combustion chamber motors.
Despite the limited sensors, good observations were made and the groundwork has
been set for future regression rate measurements.

49



Figure 20. Baseline Hybrid Rocket Motors with Instrumentation



50
Figure 21 shows a sketch of the test stand base and Figure 22 shows the
location of the sensors on the actual baseline hybrid rocket motor. The test stand,
designed and manufactured by LM engineers, places the motor on top of linear
bearings to effectively measure the load applied to the load cell which gives the
actual thrust of the hybrid rocket motor during each hot fire.

Figure 21. Hybrid Motor Test Setup Sketch

Figure 22. Baseline Hybrid Rocket Motor and Instrumentation Ready for
Testing

51
The baseline engines are highly instrumented to achieve several objectives
not only for this study but also for the MaCH-SR1 senior design project. Table 8
shows the instrumentation list for both baseline tests. The baseline motor system
uses six regression rate k-type thermocouples, 2 or 3 eroding thermocouples, three
piping thermocouples (provided by Lockheed Martin to measure gas temperatures),
six pressure transducers, one flow meter and a load cell, along with other facility
measurements. All data is collected and sorted by LabVIEW software provided by
Lockheed Martin. The regression k-type thermocouples and eroding thermocouples
are shown to have a resolution of 0.01526 F and 0.09155 F respectively. The
eroding thermocouples can measure temperatures seen in the combustion chamber
up to 6000 F depending on the type of thermocouple. The high temperature range
for the eroding thermocouples is needed to establish the flame temperature for this
specific fuel and oxidizer combination since the chemical kinetics models show
flame temperatures in the range of 2500 to 3500 K. The other thermocouples (TC
10-12) are line monitoring temperature measurements added by Lockheed Martin
test engineers. The solenoid valves are used at the beginning and end of the test to
start fuel port burning with pure gaseous oxygen and stop the burning using
nitrogen.

52
Table 8. Instrumentation List for Baseline Tests
Type Identifier Channel Placement
P.Ducer P1 B5 Pre-Inlet N2O 2000 psi 0.0305 psi
P.Ducer P2 B1 Pre-Injector 2000 psi 0.0305 psi
P.Ducer P3 B3 Chamber 1000 psi 0.0153 psi
P.Ducer P4 B4 Regulated O2 1000 psi 0.0153 psi
P.Ducer P5 B2 Regulated N2 1000 psi 0.0153 psi
P.Ducer P6 B6 Regulated N2O 2000 psi 0.0305 psi
T.Couple TC1 TC1 AL - Regression 1000 F 0.01526 F
T.Couple TC2 TC2 AS - Regression 1000 F 0.01526 F
T.Couple TC3 TC3 BL - Regression 1000 F 0.01526 F
T.Couple TC4 TC4 BS - Regression 1000 F 0.01526 F
T.Couple TC5 TC5 CL - Regression 1000 F 0.01526 F
T.Couple TC6 TC6 CS - Regression 1000 F 0.01526 F
T.Couple TC7 TC7 A - C Erode 6000 F 0.09155 F
T.Couple TC8 TC8 B - C Erode 6000 F 0.09155 F
T.Couple TC9 TC9 C - G Erode 6000 F 0.09155 F
T.Couple TC10 TC10 Regulated N2O 1000 F 0.01526 F
T.Couple TC11 TC11 Pre-Inlet N2O 1000 F 0.01526 F
T.Couple TC12 TC12 Pre-Inject Gas 1000 F 0.01526 F
Load Cell LC B8 Test Stand 500 lbf 0.0076 lbf
Flow Meter FM FM-1 N2O Line 0.75 kg/s 0.00001 kg/s
Control V. FCV FCV-1 N2O Line 450 psi 0.0069 psi
Solenoid V. SOV-1 GOX O2 Line
Solenoid V. SOV-2 GN2 N2 Line O/F
Sensors
Range Resolution
O/F

A test readiness meeting was held at Lockheed Martin prior to the testing of
the MaCH-SR1 hybrid rocket motors. The purpose of the test readiness meeting was
to get all personnel focused on the task at hand and perform final checkout of the
testing facility. These standard practices test readiness reviews are vital to the
success of the tests. Figure 23 shows the general flow chart of all measurements in
the test facility and the data acquisition system recording setup for the baseline
engines. The data acquisition system was loaned to the MaCH-SR1 hot fires. LM
engineers helped with routing instrumentation and performing the final checkouts
prior to test. The thermocouples identified as TC 1 6 are the k-type thermocouples
to measure regression rate. The thermocouples identified as TC 7 9 are the
NANMAC eroding thermocouples.

53

Figure 23. Data Acquisition System Setup
4.3 EXPERIMENT RESULTS
The fuel regression rate for each rocket is measured. The experimental data
from the two baseline tests help anchor the chemical kinetics model and regression
rate model. The regression data and model of the HTPB/N
2
O hybrid rocket are
compared to HTPB/O
2
hybrid data and existing regression models in the literature.
MaCH-SR1 2005 hot-fired five 300 lb
f
-thrust HTPB/N2O hybrid rocket
motors with a 4-point star fuel grain design. All tests were conducted at the
Lockheed Martin propulsion testing facility during the month of April 2005. Below
is the test matrix completed and test durations for each hybrid motor configuration.
Five series of tests were performed to characterize hybrid rocket motor

54
performance. Two of the tests were highly instrumented to characterize the
regression rate of the fuel, fuel temperatures, and combustion temperatures at
specified axial locations of the combustion chamber. A desired oxidizer flow rate of
1.00 lbm/sec was needed to obtain the 300 lbf thrust which resulted in a nitrous
supply pressure of 800psi initially. The nitrous supply pressure was determined
from water flow tests of the injector to determine the coefficient of discharge and
capabilities of the injector design. The desired thrust not achieved during the first
test so the decision was made to increase the pressure for the remaining tests. The
actual performance during hot fire flushed out the coefficient of discharge errors in
the water flow component testing of the injector, performed in the lab. Two-phase
flow out of the injector was attributed to these inaccuracies and results in oxidizer
mass flow rate measurement error. However, all tests had the same injector and can
be directly compared to each other. The test matrix for the 2005 MaCH-SR1 test
series is shown in Figure 24. Baseline 2 test was a short test due to a shortage of
nitrous oxide in the supply tanks. However, regression rate data was still measured
during the test.

Figure 24. Test Matrix for MaCH-SR1 2005

55
The regression rate test results for the Baseline engines are shown in Table
9. The regression rate, delta time between thermocouple failure, fuel port geometric
feature, average thrust, average chamber pressure and average flow rate are shown
in the table for each pair of thermocouples.
Table 9. Regression Rate Tables for k-Type thermocouples
Time at
Max
Port
Geometry
Axial
Distance Delta time Reg. Rate Thrust
Chamber
Pressure Fl ow Rate
sec in sec in/sec Lbf psia kg/sec
AL -0.38
AS 1.85
BL 0
BS 0.47
CL 0.67
CS 2.77
Time at
Max
Port
Geometry
Axial
Distance Delta time Reg. Rate Thrust
Chamber
Pressure Fl ow Rate
sec in sec in/sec Lbf psia kg/sec
AL 1.88
AS 9.09
BL 2.69
BS 8.12
CL 1.63
CS 2.65
187.00 336.0
326.6 0.342
0.347
1.02 0.196 249.08 440.9 0.526
5.43 0.037
0.0001
2.10 0.095 210.46 356.2 0.401
0.47 0.426 23.95 42.3
TC
Trough
Test Averages
Baseline Engine # 1
2.23 0.090 123.05 209.2 0.193 Peak
Peak
Peak
Trough
Peak
Baseline Engine # 2
Test Averages
TC
7.21 0.028 180.85 5
7
14
5
12
14

The detailed regression rate data measured from the Baseline motors gives insight
into the different rates achieved with star shaped fuel ports. A pair of thermocouples
is embedded into the fuel at a star peak or trough geometric feature, as indicated in
Table 9. The peak data showed faster regression rates for a given mass flow rate.
The one usable trough data point showed slower regression rate even at a similar
oxidizer mass flow rate and same axial distance from the injector as the
corresponding peak data. The trough data point in Baseline Engine 1 is not usable
due to the small mass flow rate measured. This trough point was collected during

56
the ignition phase and the thermocouple was possibly dislodged prior to test. Figure
25and Figure 26 show the raw regression rate data of the two baseline tests.

Figure 25. Baseline 1 Regression Rate Raw data

Figure 26. Baseline 2 Regression Rate Raw Data
The regression rate empirical constants (a and n) are calculated in two
different ways using the simple regression rate equation. The average mass flow
rates measured during the Baseline tests were used to determine the data. No data
manipulation was used in these cases. In Figure 27 and Figure 28, both plots show
that a = 5.36 and n = 4.36. Only slight variations of a and n values between the two
data reduction approaches. The peak points produced different empirical constants,
1.47 and 2.74. The higher regression rates for the peak points coincide with the

57
predictions that the peak points regress at higher rates than the trough or average
regression rate values.
Power Curve Plot - Data Manipulation to Calculate Emperical Constants
Rdot = 5.3563G
4.3582
R
2
= 0.9343, All Points
a = 5.356, n = 4.358
Rdot = 1.4706G
2.7436
R
2
= 1, Peak Points
0
0.05
0.1
0.15
0.2
0.25
0.0000 0.1000 0.2000 0.3000 0.4000 0.5000 0.6000
G of N20 (l bm/sec-i n^2)
R
e
g
r
e
s
s
i
o
n

R
a
t
e

(
i
n
/
s
e
c
)

Figure 27. MaCH-SR1 2005 Regression Data and empirical constants found by
Power Curve Method
Log Plot - Data Manipulation to Calculate Emperical Constants
y = 4.358239x + 0.728869
2
4326 R = 0.93
n= 4.3582, a = 5.356 (log a = .72887)
-1.800
-1.600
-1.400
-1.200
-1.000
-0.800
-0.600
-0.400
-0.200
0.000
-0.6 -0.5 -0.4 -0.3 -0.2 -0.1 0
log G(N2O)
l
o
g

R
d
o
t
y = 4.3582x + 0.72887
R
2
= 0.9343
n = 4.3582, a = 5.356 (log a
= 0.72887)

Figure 28. Log - Log Plot to determine a and n constants


58
The large empirical constants, in the above figures, do not correlate with
results from the literature so another approach was developed to accommodate for
the more complicated star shape fuel geometry. The regression rate data was
normalized for cylindrical shape fuel geometry so empirical data from the literature
could be better compared to this test data, as shown in Figure 29. The simplified
view breaks up the 4-point star shape fuel geometry into two different cylindrical
shapes, one for the star peaks and one for the star troughs. The radii for the peak and
trough points are 0.5 in and 0.75 in, respectively. This new regression rate equation
shows better correlation to the data in the literature and previous MaCH-SR1 tests
of HTPB/N
2
O HRMs.
y = 0.357x
1.023
0.001
0.01
0.1
1
0.01 0.1 1
Gox (lbm/sec-in^2)
R
e
g
r
e
s
s
i
o
n

R
a
t
e

(
i
n
/
s
e
c
)
r = 0.357Gox^1.023 r = 0.174Gox^1.032 r = 0.39Gox^0.15 r = 0.131G^0.674
Chiaverini, et al. (2001)
Shanks & Hudson (2000)
Serin & Gogus (2003)

Figure 29. HTPB Regression Rate Results

One method to determine regression rate for a nitrogen based oxidizer is to
use the fundamental regression equation for pure oxygen and modify the formula for

59
nitrous oxide. Adding nitrogen to the equation acts as an inhibitor to fuel regression
and the regression rate can be analyzed in this way. A dilution factor for the addition
of nitrogen in the oxidizer is present in the Marxman equation (1964) and his
experiments showed a lower regression rate. Since nitrous oxide has 2 atoms of
nitrogen and 1 atom of oxygen, only 1/3 of the oxidizer mass flux was used to
determine regression rate. The regression rate equation for 1/3 the oxidizer mass
flux is r = 1.098G
1.023
. Figure 30 shows the regression rate equations for 100%
oxidizer and diluted oxidizer.
y = 1.4732x
1.0226
33% Oxidizer
y = 1.0977x
1.0226
y = 0.7251x
1.0226
y = 0.5403x
1.0226
100% Oxidizer
y = 0.3569x
1.0226
0
0.02
0.04
0.06
0.08
0.1
0.12
0.14
0.16
0.18
0.2
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
Dilution of Oxidizer
R
e
g
r
e
s
s
i
o
n

R
a
t
e

(
i
n
/
s
e
c
)

Figure 30. Modified Regression Rate Equation for Percentage of Oxygen in
N2O Oxidizer
Figure 31 shows the different regression rate equations derived for the
trough regions of the star geometry during the beginning of a hot fire. At about 50%
of oxidizer, the regression rate for the trough region is r = 0.1065 G
0.9975
. It is likely
that not all oxidizer reaches the trough regions in the beginning of the burn due to

60
the geometry. However, as peak points regress faster, the troughs are exposed to a
higher percentage of oxidizer and the regression rate will change accordingly.
y = 0.1072x
1.0035
y = 0.1065x
0.9975
y = 0.1067x
1.4094
y = 0.0872x
0.7095
0
0.02
0.04
0.06
0.08
0.1
0.12
0 0.25 0.5 0.75 1
% of Oxidizer in Trough Regions
G
o
x

(
l
b
m
/
s
e
c
-
i
n
^
2
)

Figure 31. Percentage of Oxidizer in the Trough Area of a 4-Point Star Fuel
Port Geometry
The measured combustion temperature using the NANMAC eroding sensors
did not produce useful results. Figure 32 shows the measured temperature data from
the NANMAC eroding thermocouples. The highest temperature was measured
during the Baseline Engine 2 test with the higher chamber pressure. The
temperature reading is much lower than expected most likely due to soot produced
from the fuel inhibiting the reading of the sensor during hot fire. The G-type sensor,
measuring flame temperature by protruded the sensor into the combustion chamber
just past the fuel, failed to work. No usable conclusions could be made from the
eroding thermocouple data.


61
0
500
1000
1500
2000
2500
-1 -0.5 0 0.5 1 1.5 2 2.5 3 3.5 4
Burn time (sec)
C
o
m
b
u
s
tio
n
T
e
m
p
e
r
a
tu
r
e
(F
)
A (C-TYPE) TC7() B (C-TYPE) TC8() C (G-TYPE) TC9()
0
500
1000
1500
2000
2500
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Time (sec)
T
e
m
p
e
r
a
t
u
r
e
(d
e
g
F
)
C-Type (1) C-Type (2) G-Type
T (max) = 2095 F T(max) = 2095 F
T(max) = 2409 F

Figure 32. NANMAC Eroding Thermocouple Attempt at Measuring
Combustion Temperature
Figure 32. NANMAC Eroding Thermocouple Attempt at Measuring
Combustion Temperature
The NANMAC sensors have not been used for the internal hybrid combustion
application and thus needs additional development if used in the future. Future
considerations for using NANMAC eroding thermocouple sensors should include
mounting the sensors outside of the combustion chamber near the nozzle exhaust to
capture the temperature of the hot gases. This might eliminate the excessive build up
of soot and the sensor can be reused for additional motor testing.
The NANMAC sensors have not been used for the internal hybrid combustion
application and thus needs additional development if used in the future. Future
considerations for using NANMAC eroding thermocouple sensors should include
mounting the sensors outside of the combustion chamber near the nozzle exhaust to
capture the temperature of the hot gases. This might eliminate the excessive build up
of soot and the sensor can be reused for additional motor testing.
4.4 PREDICTIONS USING IDEAL ROCKET CALCULATIONS 4.4 PREDICTIONS USING IDEAL ROCKET CALCULATIONS
The ideal rocket equations and ideal thermodynamics are used to verify the
regression rate measurements from actual design and tested parameters for the
MaCH-SR1 2004-2005 300lb
f
thrust rockets. Using GDL propep, the known
oxidizer and fuel mixture ratio and specific heat ratio are found and used in the
nozzle characteristic equations.
The ideal rocket equations and ideal thermodynamics are used to verify the
regression rate measurements from actual design and tested parameters for the
MaCH-SR1 2004-2005 300lb
f
thrust rockets. Using GDL propep, the known
oxidizer and fuel mixture ratio and specific heat ratio are found and used in the
nozzle characteristic equations.
A steady state analysis of the system is determined. Given the combustion
chamber pressure, Pc, the specific heat ratio, , and the exit pressure the Mach
number, expansion ratio, , and nozzle throat area can be found. The exit Mach
number is determined using the aerodynamic compressible flow equations. The
A steady state analysis of the system is determined. Given the combustion
chamber pressure, Pc, the specific heat ratio, , and the exit pressure the Mach
number, expansion ratio, , and nozzle throat area can be found. The exit Mach
number is determined using the aerodynamic compressible flow equations. The

62
nozzle exit pressure is set equal to the ambient pressure to maximize the rocket
thrust. Since the hybrid rocket motor is tested in Denver, CO, at an altitude of 5,250
ft. The ambient pressure at this altitude is found to be 12.2 psi, using the Standard
Atmosphere model. The nozzle characteristics are determined using Equations (13)
(16).
1
1
2
1

e
c
e
P
P
M
(13)
( ) 1 2
1
2
2
1
1
1
2 1

+


+
+
= =

e
e t
e
M
M A
A
(14)
(15)
c f
t
P C
F
A =
(16)
Since the assumption is made that the exit pressure is equal to the atmospheric
pressure the coefficient of thrust, C
f
, is only a function of the specific heat ratio, exit
pressure, and chamber pressure all of which are known. Therefore, using the
equation given above, one can solve for the required throat diameter. The nozzle
throat area, A
t
, is determined to be 0.40 in
2
for this hybrid rocket motor design.
The propellant mass flow rate is determined by using the calculated nozzle
throat area in Equations (17) (19). The specific impulse is a function of the
products of combustion (, M, T
c
) along with the pressure difference between the
nozzle exit pressure, Pe and the chamber pressure, Pc. The characteristic velocity,

63
c*, represents the thermochemical makeup for a particular propellant and may be
indicative of the combustion efficiency. The expression for ideal c-star is given in
Equation 18, and is seen to be a ratio of the ISP and coefficient of thrust.
( )

/ 1
1
1
2
c
e c
o sp
P
P RT
g I (17)
f
o sp
C
g I
c =
*
(18)
*
c
A p
m
t c
= &
(19)
Next, the oxidizer mass flow rate and fuel mass flow rate is calculated using
Equations (20) (21). The mixture ratio, MR, is the desired oxidizer to fuel ratio
and in the case of the MaCH-SR1 hot fires the MR is 4. The thrust of the rocket can
also be predicted using Equation (22).
1 +
=

MR
MR
m mox (20)

+
=

1
1
MR
MR
m m
f
(21)
c f t
P C A F = (22)
Knowing the length of the combustion chamber, the mass flow rate the fuel,
and the initial geometry of the fuel grain port, the regression rate parameter is
calculated using Equation (23).

64

p f
f
LS
N m
r

=
(23)
The regression rate value, r, for the actual hot fire of the first Baseline engine test is
found to be 0.051 in/sec, using Sp of 7.16 inches for the 4-point star geometry, the
specific heat ratio of 1.282 from GDL Propep, the average combustion chamber of
400psi, and the burn time of 18 seconds. The theoretical average regression rate
does correlate with average regression rate of 0.056 in/sec measured by former
MaCH-SR1 teams (2004) for a single, round fuel port geometry. The previous
years 5000lb
f
, full scale hybrid tests with an 8-point star shape fuel port yielded a
measured average regression rate of 0.039 in/sec. Table 10 shows the results of the
steady state analysis calculations for the planned test parameters with a chamber
pressure of 500 psi and a burn time of 15 seconds compared to the actual hot fire
results.
Table 10. Steady State System Analysis Results Summary for Planned and
Actual Hot Fire Parameters
Parameter
Calculated from GDL Propep
Pc=500psi
time=15sec
Pc=400psi
time=18sec
Chamber Temperature (K) 2718 2718
Cp/Cv 1.282 1.282
Molecular Mass (gm/mol) 22.46 22.46
Exit Mach Number 3.88 3.78
Nozzle Expansion Ratio 4.28 4.68
Area of Nozzle Throat, At (in2) 0.4 0.4
Nozzle Mass Flow (lbm) 1.26 1.058
Characteristic Exit Velocity, c* (ft/s) 5088 4860
Oxidizer Mass Flow Rate (lbm/sec) 1.000 0.846
Fuel Mass Flow Rate (lbm/sec) 0.26 0.21
Propellant Mass (lbm) 18.84 19.04
Oxidizer Mass (lbm) 18.1 15.22
Fuel Mass (lbm) 3.75 3.81
Nozzle Calculations
Propell ant
Value


65
The predicted regression rate is comparable to the measured average
regression rate. The obvious differences are due to the assumptions used in the ideal
rocket equations which are as follows:
1. Chemical reaction products are homogeneous
2. All species of the working fluid are gaseous
3. The perfect gas law is obeyed
4. No heat transfer across the rocket walls, so the flow is adiabatic
5. Isentropic expansion flow through nozzle
6. Propellant flow is steady and constant
7. Gas velocity, pressure, temperature and density are all uniform
across any section normal to the nozzle axis
8. Chemical equilibrium is established within the rocket chamber and
frozen flow occurs in the nozzle.
The expected actual performance for a chemical rocket is usually 1-6% below the
ideal calculated value (Sutton and Biblarz, 2001).
4.5 DISCUSSION OF RESULTS
The experimental data from the MaCH-SR1 lab scale hybrid rocket motor
tests is used for comparison with the chemical kinetics and regression models. The
regression data is compared to HTPB/O
2
hybrid data and existing regression models
in the literature. The chamber pressure was designed for 500 psia, however, the
pressure drop was greater than expected due to injector design, pipe losses and
losses through flow valves, and relief valves. The burn time in test #1 was 18.2
seconds.
Table 11 compares the different MaCH-SR1 2005 regression rate equations
(from the 5 usable data points) to regression rate equations in the literature for
similar hybrid rocket motor tests. Similarities in equations can be seen with HTPB

66
and Oxygen systems and with HTPB and nitrous oxide systems. Different
calculations in regression rate equations for star peak and trough geometric features
are attempted.
Table 11. Regression Rate Equations in Literature
Reference Description Regression Eqn Fuel / Oxidizer
Marxman & Gilbert (1964)
Turbulent Diffusion
limited-heat transfer
(non-radiative)
system Plexiglas / Oxygen
Serin & Gogus (2003)
Navier - Stokes
model using CFD-
ACE and REGRESS
software HTPB / Oxygen
Thiokol Corp. Small rocket firings HTPB / Oxygen
Karabeyoglu , et. al. (2002) Small rocket firings Paraffin / Oxygen
Greiner & Frederick (1992) tiny lab firings r = 0.044 G
0.6
Plexiglas / GOX
Chiaverini, et al. (2001)
Flat plate Tube
Burner r = 0.049 G
0.61
(R
2
= .989) HTPB / Oxygen
Shanks & Hudson (2000) Small rocket firings r = 0.131 G
0.674
HTPB / Oxygen
MaCH-SR1 02-03 (2004)
300 lbf - thrust HRM,
Average r = .1741G
1.0315
HTPB / N
2
O
MaCH-SR1 02-03 (2004)
1000 lbf - thrust
HRM, Average r = 0.9146 G
2.3698
HTPB / N
2
O
Chluda (2005)
300 lbf

thrust HRM,
Modified Gox r = 0.357 G
1.023
HTPB / N
2
O
Chluda (2005)
300 lbf-thrust HRM,
Average r = 5.3563 G
4.3582
(R
2
= .934) HTPB / N
2
O
Chluda (2005)
300 lbf-thrust HRM,
Star peak r = 1.4706 G
2.7436
HTPB / N
2
O
Chluda (2005)
300 lbf-thrust HRM,
Star trough, 50%G r = 0.107 G
0.998
HTPB / N
2
O
15 . 0
39 . 0
o
G r =

23 . 0 2 . 0
8 . 0
) ( 036 . 0

x
G
r
f
=

62 . 0
488 . 0
o
Paraffin G r =

681 . 0
146 . 0
o
HTPB G r =



The calculated regression rate value, r, for the MaCH-SR1 baseline HRM is
found to be 0.051 in/sec, using a circumferential port perimeter of 7.16 inches for
the 4-point star geometry, the specific heat ratio of 1.236 from GDL Propep, the
average combustion chamber of 400psi, and the burn time of 18 seconds. The
theoretical average regression rate does correlate with average regression rate of
0.056 in/sec measured by former MaCH-SR1 teams (2004) for a single, round fuel

67
port geometry. The measured average regression rate for the baseline engine is
0.089 in/sec.
Large regression rate differences have been observed between trough and
star points in a star shape fuel in previous MaCH-SR1 tests (2004). The star points
regress at a faster rate, and the outcome is a cylindrical shape fuel towards the end
of the burn which has proven to be fuel efficient and have more combustion stability
than multi port fuel geometry. The k-type thermocouple data from the 300 lb
f
-thurst
lab scale Baseline engine shows a regression rate of 0.098 in/sec for the star peak
feature and 0.037 in/sec for the trough. Measurements of the diameter of the fuel
before and after were made to determine regression rate of the peak and trough for
the duration of the tests. The Baseline HRM test resulted in peak and trough
regression rates of 0.0708 in/sec and 0.0482 in/sec and an average oxidizer mass
flow rate of 1.21 lbm/sec. The aluminum HRM test resulted in peak and trough
regression rates of 0.0642 in/sec and 0.0500 in/sec and an average oxidizer mass
flow rate of 1.444 lbm/sec. The composite HRM test resulted in peak and trough
regression rates of 0.1155 in/sec and 0.0905 in/sec and an average oxidizer mass
flow rate of 1.567 lbm/sec. The previous years 5000lb
f
-thrust, full scale hybrid
tests with an 8-point star shape geometry yielded a measured regression rate of
0.045 in/sec for the peak point and 0.028 in/sec for the trough.
Figure 33 shows a non-linear progression in regression rate for different
sized hybrid rocket motors. The figure uses data from MaCH-SR1 tests of HTPB
and N
2
O hybrids with a single, round fuel port. The MaCH-SR1 data from 2005
could not be plotted here due to the different star fuel shape design.

68
Fuel Regression for single round ports as a function of thrust
0.048
0.05
0.052
0.054
0.056
0.058
0.06
0.062
0 200 400 600 800 1000 1200
Thrust (l bf)
R
e
g
r
e
s
s
i
o
n

(
i
n
/
s
)
Fuel Regression

Figure 33. Regression Rate for Single Round Port HRMs
Figure 34 and Figure 35 show the remaining unburned fuel after an 18
second and 5 second hot fire for a 5000lb
f
thrust HRM. The fuel observations made
visually and by measurements show that the fuel closer to the injector regresses
faster than the nozzle end. Also it was noted that the 5000lb
f
thrust rocket was built
too long and the flame became so fuel rich toward the end that it stopped burning at
the nozzle end. No difference in regression could be noted after a 5 second or 15
second burn. Typical hybrid rocket motor solid fuel burns more efficiently in an
oxidizer rich environment. A large amount of unused fuel at the end of a hot fire is
not desirable. More research is needed to determine regression rate differences in
axial positions along large (< 1000 lb
f
-thrust) HRM combustion chambers. The
large axial difference in regression rate is not observed in the 300 lb
f
thrust lab
scale hot fires.

69

Figure 34. MaCH-SR1 April 2004 hot fires Injector End View, both after 18
second burn time


Figure 35. Nozzle End View, a)after 5 sec burn time, b) after 18 sec burn time
5. CONCLUSIONS AND FUTURE CONSIDERATIONS
An accurate characterization of the fuel regression rate can reduce the
overall mass of a launch vehicle by reducing the mass of the fuel which directly
contributes to increasing the thrust to weight ratio. Using the ideal rocket equations
presented in Section 4 can help predict regression rate with given fuel grain
geometry and thermochemistry values for a specific fuel and oxidizer combination.
Combining that with detailed thermochemistry and incorporating the details into the
Marxman (1964) regression rate equations will make it possible to solve the

70
complexities of regression rate for different size hybrid motors with different
propellant combinations.
The measured average regression rate and calculated regression rate are
comparable. Knowing that the differences are inherent in the accuracy of the
measurement sensors, and the assumptions set by the ideal rocket equations. The
rocket performance calculations were also verified by the GDL Propep software and
were comparable. This shows that the software used in the chemical kinetics is
comparable to the STANJAN and CHEMKIN models. The table below presents the
fundamental parameters, calculated or measured, that are compared in this thesis.
Table 12. Summary of Thermochemistry Parameters with Calculated
Regression Rate Results
Rocket Parameters CHEMKIN STANJAN GDL Propep Experiment
Combustion Temp, Tc (K) 3535 4303 2718 1590
Molecular Weight, M 21.97 21.97 22.46 N/A
Cp / Cv , 1.33 1.33 1.282 N/A
Avg Regression rate (in/sec) * 0.038 0.035 0.051 0.059
O/F Ratio = 4 for all parameters
* All average regression rate is calculated expect for "Experiment" column


The adiabatic flame temperature, which is the combustion temperature (Tc), and
Cp/Cv are determined from STANJAN, the same Cp/Cv is used in CHEMKIN but a
different combustion temperature is found. The combustion temperatures and Cp/Cv
for all cases is inputted into the ideal rocket equations to calculate mass flow rate
using an O/F = 4. The average regression rate for each model is then calculated and
the results are in Table 12. The measured average regression rate is found by the
change in mass of the baseline combustion chamber measured before and after the
test and the result is recorded in the experiment column.

71
The measured combustion temperature was achieved using the NANMAC
eroding sensors. However, errors can be observed in the measure temperature data.
The temperature was much lower than expected and must have measured a lower
reading due to soot produced on the sensor during hot fire. These sensors have not
been used for this application and thus need additional development if used in the
future. Future considerations for using NANAMAC eroding thermocouple sensors
should mount the sensors outside of the combustion chamber near the nozzle
exhaust to capture the temperature of the hot gases. This might eliminate the
excessive build up of soot and the sensor can be reused for additional motor testing.
A predicted regression rate can be estimated using the ideal rocket equations
to complete the comparison of a highly simplified model to hybrid rocket regression
rate. Future research in regression rate should look at the detailed chemical kinetics
to reach a more accurate regression rate that spans the market of different fuel and
oxidizer combinations. Often, thermochemistry parameters are lumped into one
empirical constant and the variation for different hybrid rocket motors is lost.
Compared to the well-known combustion analysis of solid and liquid rockets, only a
limited amount of analysis and data has been recorded for hybrid rocket
combustion. A detailed understanding of the chemical and thermodynamic
properties of specific reactants is important to prove that hybrid rockets can produce
high combustion performance and are indeed safe rockets that can be used for
human space flight. Chemical kinetics may also be able to solve the complexities of
scaling hybrid rocket motors.

72
After a hiatus in hybrid rocket testing in the late 1990s, hybrid rockets are
once again being acknowledged in the rocket industry and are also being recognized
as a simpler and safer rocket for human space flight. Large scale hybrid rocket
motor testing had reached its peak in the 1990s with NASA funding, however, the
test program was severed prematurely due to combustion instabilities that could not
be solved at that time. Compared to the well-known combustion analysis of solid
and liquid rockets, only a limited amount of analysis and data has been recorded for
hybrid rocket combustion. If a hybrid rocket was to take the place of the Space
Shuttles solid rocket boosters (SRB), additional large scale testing is needed. The
length of an SRB sized hybrid rocket motor would need to be about 20-40% longer,
depending on the oxidizer used, but the diameter of the rocket could remain the
same. Another design attribute of hybrids is a fly-back booster that would safely fly
back to the launch site for refueling, resulting in cost savings and less detrimental
impact on the environment. More testing and analysis is needed before hybrids take
over the solid rocket placeholder in the launch industry. Hybrids are a feasible
alternative to existing launch vehicles due to their increased safety and should be
considered for the next generation Crew Launch Vehicle. Hybrids combine the
safety of separating the fuel and oxidizer, as in liquid rockets, with the simplicity of
a solid rocket motor design. The environmentally friendly propellants also give
hybrids an added advantage for the next generation launch vehicle. Thus additional
research, development, and testing for hybrid rocket motors are warranted for future
space flight endeavors.

73
6. REFERENCES
Boardman TA, Abel TM, Claflin SE, and Shaeffer CW. Design and test planning
for a 250-klbf-thrust hybrid rocket motor under the Hybrid Propulsion
Demonstration Program, AIAA 97-2804, 33rd AIAA/ASME/SAE/ASEE Joint
Propulsion Conference and Exhibit, July 1997.

Boardman TA, Hybrid Propellant Rockets, in: Sutton, G.P., and Biblarz, O.
Rocket Propulsion Elements. 7th ed, John Wiley & Son, New York, 2001, Chap 15
& 13.

Boardman TA, Carpenter RL, and Claflin SE., A Comparative Study of the Effects
of Liquid- Versus Gaseous-Oxygen Injection on Combustion Stability in 11-inch-
Diameter Hybrid Motors, AIAA 97-2936, 33rd AIAA/ASME/SAE/ASEE Joint
Propulsion Conference and Exhibit, July 1997.

Branch, Melvyn C., and Cor, Joseph J., Structure and chemical kinetics of flames
supported by nitrogen oxides, Pure and Applied Chemistry, Vol. 6, No. 2, pp. 277-
283, 1993.

Cengel, YA, and Boles, MA. Thermodynamics An Engineering Approach, 2
nd

edition, McGraw Hill, Inc., New York, 1994.

Chiaverini MJ, Kuo KK, Peretz A, and Harting, GC. Regression-Rate and Heat-
Transfer Correlations for Hybrid Rocket Combustion, Journal of Propulsion and
Power, Vol. 17, No. 1 2001, pp. 99-110.

Chiaverini MJ, Serin N, Johnson DK, Lu YC, Kuo KK, and Risha, GA. Regression
Rate Behavior of Hybrid Rocket Solid Fuels, Journal of Propulsion and Power,
Vol. 16, No. 1, 2000, pp. 125-132.

Chiaverini MJ, Kuo KK, Peretz A, and Harting, GC. Regression-Rate and Heat-
Transfer Correlations for HTPB/GOX Combustion in a Hybrid Rocket Motor,
AIAA 1998-3185.

Chluda, H.L. and McWilliam, K.I., "MaCH-SR1 HTPB/N20 Hybrid Rocket Motor
Fuel Design and Regression Analysis", AIAA Region V Student Paper Conference,
Minneapolis, April 2004

Currie, I.G., Fundamental Mechanics of Fluids, 3rd edition, Marcel Dekker, Inc.,
New York, 2003.

Estey PN, Altman D, and McFarlane JS. An Evaluation of Scaling Effects for
Hybrid Rocket Motors, AIAA 91-2517, 27th AIAA/SAE/ASME Joint Propulsion
Conference, June 1991.


74
Evans B, Risha GA, Favorito N, Boyer E, Wehrman RB, Libis N, and Kuo, KK.
Instantaneous Regression Rate Determination of a Cylindrical X-Ray Transparent
Hybrid Rocket Motor, AIAA 2003-4592, 39
th
AIAA/ASME/SAE/ASEE Joint
Propulsion Conference, July 2003.

Gramer, DJ and Taagen, TJ. Low Cost Surface Regression Sensor for Hybrid
Fuels, Solid Propellants and Ablatives, American Institute of Aeronautics &
Astronautics, AIAA 2001-3529.

Incopera, Frank P., and DeWitt, David P., Fundamentals of Heat and Mass Transfer,
3rd ed., Wiley, New York, 1990, pp. 827-828.

Karabeyoglu MA and Altman D. Dynamic Modeling of Hybrid Rocket
Combustion, Journal of Propulsion and Power, Vol. 15, No. 4, 1999, pp. 562-571.

Karabeyoglu MA, Altman D., and Cantwell, BJ. Combustion of Liquefying Hybrid
Propellants: Part 1, General Theory, Journal of Propulsion and Power, Vol. 18, No.
3, 2002, pp. 610-620.

Lanier, J.E., GDL Propep Front Panel, Version 1.2 Beta, Gas Dynamics Lab, 1999

Larson, W.J., and Wertz, J.R., Space Mission Analysis and Design. 3rd ed.
Microcosm Press, California, 1999, Chap 17.

Marxman GA, Wooldridge CE, and Muzzy RJ, "Fundamentals of Hybrid Boundary
Layer Combustion," Heterogeneous Combustion, edited by H.G. Wolfhard, I.
Glassman, and L. Green, Jr., Vol. 15, AIAA Progress in Astronautics and
Aeronautics, Academic Press, Inc., New York, 1964, p. 485-521.

MaCH-SR1 2003-2004., Final Report, 2004.

MaCH-SR1 2004-2005., Final Report, 2005.

Norr, J., Project Prometheus,
http://stimpy.cen.uiuc.edu/soc/isds/hybrid/combust.htm, October 1996.

Pacchioli, D., Hybrid Rockets. Research/Penn State, Vol. 18, no. 1 January, 1997.

Peters, R.L. Design of Liquid, Solid, and Hybrid Rockets, Hayden Book
Company, Inc., New York, 1965.

RCS Rocket Motor Components, Inc., Air Force Chemical Equilibrium Specific
Impulse code, http://www.rocketmotorparts.com/resources.html, April 2004.


75
Serin N and Gogus, YA. Navier-Stokes Investigation on Reacting Flow Field of
HTPB/O2 Hybrid Motor and Regression Rate Evaluation, AIAA 2003-4462, 39th
AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, July 2003.

Shank, R. and Hudson, MK. A labscale Hybrid Rocket Motor for Instrumentation
Studies, Journal of Pyrotechnics, Issue 11, Summer 2003.

States, D.N., Unity IV Hybrid Rocket Motor, Nozzle, Fuel and Ignition Systems,
AIAA 2001-3533, 2001.

Story, G., Zoladz, T., Arves, J., Kearney, D., Abel, T., and Park, O. Hybrid
Propulsion Demonstration Program 250K Hybrid Motor, AIAA 2003-5190, 39th
AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, July 2003.

Swami RD and Gany A. Analysis and Testing of Similarity and Scale Effects in
Hybrid Rocket Motors, Acta Astronautica, Vol. 52, 2003, pp. 619-628.

Turns, S.R., Introduction to Combustion, 2nd ed., McGraw Hill, New York, 2000.

Wangmin, HTPB, http://htpb.netfirms.com/htpben.htm, January 2002.
Wooldridge CE, and Muzzy RJ, "Internal Ballistic Considerations in Hybrid Rocket
Design," Vol. 4, No. 2, J. Spacecraft, Academic Press, Inc., New York, 1967, p.
255-262.


76

You might also like