thesis

214
AUTOIGNITION OF KEROSENE BY DECOMPOSED HYDROGEN PEROXIDE IN A DUMP COMBUSTOR CONFIGURATION A Thesis Submitted to the Faculty of Purdue University by James C. Sisco In Partial Fulfillment of the Requirements for the Degree of Master of Science in Aeronautics and Astronautics August 2003

Upload: will-black

Post on 23-Nov-2014

282 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Thesis

AUTOIGNITION OF KEROSENE BY DECOMPOSED HYDROGEN PEROXIDE

IN A DUMP COMBUSTOR CONFIGURATION

A Thesis

Submitted to the Faculty

of

Purdue University

by

James C. Sisco

In Partial Fulfillment of the

Requirements for the Degree

of

Master of Science in Aeronautics and Astronautics

August 2003

Page 2: Thesis

ii

Page 3: Thesis

iii

ACKNOWLEDGEMENTS

Page 4: Thesis

iv

TABLE OF CONTENTS

Page

LIST OF TABLES ............................................................................................................ vii

LIST OF FIGURES ............................................................................................................ix

ABSTRACT ...............................................................................................................xiv

CHAPTER 1: INTRODUCTION .................................................................................. 1

1.1 Hydrogen Peroxide ..............................................................................................3

1.1.1 Benefits of Hydrogen Peroxide as a Propellant ...........................................4

1.1.2 Catalyst Bed Design ....................................................................................6

1.2 Background on Hydrogen Peroxide Bipropellant Engines..................................7

1.2.1 Hypergolic Bipropellant Engines ................................................................7

1.2.2 Staged-Bipropellant Engines .......................................................................8

1.3 Design of Staged-Bipropellant Engines.............................................................11

1.3.1 Advantages of Staged-Engines ..................................................................12

1.3.2 Challenges in Staged Engine Design .........................................................17

CHAPTER 2: INJECTOR DESIGN ............................................................................ 20

2.1 Engine Design....................................................................................................21

2.2 Transverse Injector Design Considerations .......................................................22

2.2.1 Orifice Sizing .............................................................................................23

2.2.2 Decomposed Gas Flow Calculations .........................................................27

2.2.3 Rearward-Facing Step Sizing ....................................................................29

2.2.4 Trajectory Model .......................................................................................32

2.2.5 Baseline Injector Design............................................................................35

CHAPTER 3: AUTOIGNITION.................................................................................. 40

3.1 Basic Chemistry of Autoignition.......................................................................40

Page 5: Thesis

v

Page

3.1.1 Composition of Kerosene Fuel..................................................................41

3.1.2 Combustion of Hydrocarbons ....................................................................43

3.1.3 Kinetics of Autoignition............................................................................45

3.2 Autoignition Studies of Kerosene Fuel in Air ...................................................47

3.3 Dump Combustor Autoignition.........................................................................51

3.4 Autoignition Studies in Staged Combustors ......................................................52

3.5 Goals of this Autoignition Study.......................................................................56

CHAPTER 4: EXPERIMENTAL SETUP................................................................... 58

4.1 Test Facility Overview ......................................................................................58

4.2 Cavitating Venturi Flow Control.......................................................................60

4.3 Data Acquisition and Control............................................................................63

4.4 Instrumentation..................................................................................................65

4.5 Test Article and Setup .......................................................................................65

4.6 Hydrogen Peroxide Dilution..............................................................................71

4.7 Pressure Budget .................................................................................................72

4.8 Test Procedure ...................................................................................................74

4.9 Firing Sequence .................................................................................................77

CHAPTER 5: EXPERIMENTAL RESULTS.............................................................. 81

5.1 Test Plan Overview............................................................................................81

5.2 Data Reduction ..................................................................................................84

5.2.1 Pressure Transducer Data ..........................................................................84

5.2.2 FFT and Filtering .......................................................................................86

5.2.3 Calculations using Measured Data ............................................................88

5.3 Uncertainty Ana lysis .........................................................................................93

5.4 Data Summary ...................................................................................................96

5.4.1 Strong Autoignition...................................................................................96

5.4.2 Weak Autoignition...................................................................................101

5.4.3 No Autoignition.......................................................................................104

5.4.4 Trends in Autoignition Data ....................................................................107

Page 6: Thesis

vi

Page

5.4.5 Temperature Data ....................................................................................109

5.5 Catalyst Bed Performance ...............................................................................111

5.6 DMAZ Fuel .....................................................................................................116

5.7 Data Analysis ...................................................................................................118

5.7.1 Pressure Effects .......................................................................................118

5.7.2 Jet Trajectory ...........................................................................................120

5.7.3 Shear Layer Residence Time ...................................................................123

5.7.4 Autoignition Correlation..........................................................................123

CHAPTER 6: SUMMARY AND CONCLUSIONS ................................................. 125

LIST OF REFERENCES ................................................................................................ 131

APPENDICES

Appendix A: Part Drawings ............................................................................... 137

Appendix B: DMAZ Material Safety Data Sheet .............................................. 159

Appendix C: Test Cell P&ID............................................................................. 165

Appendix D: Data Reduction............................................................................. 167

Appendix E: Uncertainty Analysis .................................................................... 183

Appendix F: Test Data ....................................................................................... 189

Page 7: Thesis

vii

LIST OF TABLES

Table Page

Table ?1.1: Performance comparison of 90% H2O2/JP-8 system to common rocket propellant combinations. Specific impulse calculated assuming a chamber pressure of 1000 psia and equilibrium expansion to sea level pressure of 14.7 psia. ............... 2

Table ?1.2: Properties of liquid and decomposed hydrogen peroxide with concentration.3-5

..................................................................................................................................... 4

Table ?2.1: RBCC engine operating conditions. 26 ............................................................. 22

Table ?2.2: Fuel orifice and manifold geometry of transverse injector design including flow parameters at baseline operating conditions. .................................................... 36

Table ?2.3: Oxidizer port geometry and flow parameters in baseline monopropellant and bipropellant operational modes. ................................................................................ 36

Table ?3.1: Basic hydrocarbon families as found in kerosene-based fuels.51 ................... 42

Table ?3.2: Physical properties of JP-8 fuel. All properties stated at ambient temperature and pressure unless indicated otherwise.47,50 ............................................................ 42

Table ?3.3: Comparison between some physical properties of DMAZ and JP-8............... 57

Table ?4.1: List of cavitating venturis available at APCL. ................................................ 63

Table ?4.2: Comparison between catalyst bed designs used during testing. ..................... 66

Table ?4.3: Typical instrumentation list for autoignition testing. ...................................... 70

Table ?4.4: Pressure budget for a test at baseline design conditions using the GK catalyst bed, see Chapter 2.................................................................................................... 74

Page 8: Thesis

viii

Table Page

Table ?4.5: Typical valve firing sequence for each catalyst bed design. ........................... 80

Table ?5.1: Variation in chamber gas properties based on chamber contraction ratio. ..... 91

Table ?5.2: Decomposition properties of hydrogen peroxide as a function of concentration.................................................................................................................................... 92

Table ?5.3: Estimated random error for all measured variables. ........................................ 95

Table ?F.1: Measured and calculated test data during bipropellant operation. ................ 190

Table ?F.2: Measured and calculated bipropellant test data (cont.) ................................. 191

Table ?F.3: Measured and calculated monopropellant test data. ...................................... 192

Table ?F.4: Calculated decomposed gas and fuel flow conditions................................... 193

Table ?F.5: FFT results for bipropellant portion of each test including maximum range of pressure oscillations. ............................................................................................... 194

Table ?F.6: FFT result for monopropellant section of each test including maximum range of pressure oscillations............................................................................................ 195

Table ?F.7: Calculated uncertainty in densities and mass flow rates. .............................. 196

Table ?F.8: Calculated uncertainty in equivalence ratio and bipropellant performance parameters. .............................................................................................................. 197

Table ?F.9: Calculated uncertainty in monopropellant performance parameters. ............ 198

Table ?F.10: Calculated uncertainties in fuel and oxidizer flow parameter as well as momentum ratio and residence time. ...................................................................... 199

Page 9: Thesis

ix

LIST OF FIGURES

Figure Page

Figure ?1.1: Typical staged-bipropellant engine using H2O2/kerosene. This is commonly called a ‘dump’ combustor configuration. ................................................................ 11

Figure ?1.2: Comparison of C* vs. mixture ratio curves of 90% H2O2/JP-8 to common rocket propellant combinations.4 C* calculated assuming a chamber pressure of 1000 psia and equilibrium expansion. ...................................................................... 13

Figure ?1.3: Schematic of a Gamma class research engine indicating gas ports and fuel injection points.15 ...................................................................................................... 14

Figure ?1.4: Picture of a Gamma class gas port injector.15 ................................................ 15

Figure ?1.5: Early injector design used in thermal ignition research.11 This injector uses a combination of basic injection concepts such as swirl and gas port injectors. ......... 15

Figure ?1.6: Comparison of chamber temperature vs. mixture ratio of 90% H2O2/JP-8 against other common propellant combinations.4 Combustion temperature calculated assuming a chamber pressure of 1000 psia and equilibrium expansion. . 16

Figure ?2.1: Schematic of a transverse injector indicating important design parameters. . 23

Figure ?2.2: Dependence of discharge coefficient on orifice length over diameter ratio for a full flowing square-edged inlet.44........................................................................... 26

Figure ?2.3: Reacting flow behind a rearward-facing step showing turbulent eddies.45 .... 29

Figure ?2.4: Schematic of injector assembly indicating important geometry and dimensions. All dimensions are in inches. ............................................................... 37

Page 10: Thesis

x

Figure Page

Figure ?2.5: Comparison between JP-8 jet trajectories produced by transverse injector design during monoprop and biprop operation. Fuel orifices are located at x = 0, y =0 and x = 0, y = 1.707 inches centerline is at y = 0.854 inches............................... 38

Figure ?4.1: Schematic of a cavitating venturi indicating important parameters. ............. 62

Figure ?4.2: Simplified schematic of test stand with instrumentation locations. .............. 68

Figure ?4.3: Photos of installed test article for: a) GK catalyst bed and b) PCI catalyst bed assemblies. ................................................................................................................ 68

Figure ?4.4: Comparison of engine assemblies for each catalyst bed design. Axial locations of instrumentation are indicated in schematic. .......................................... 69

Figure ?5.1: Plot of pc2 from an autoignition test using PCI catalyst bed. Test conditions: 98% H2O2, φ = 1.84, CR = 3.0. ................................................................................. 85

Figure ?5.2: Plots of pc2 in a) unfiltered and b) filtered form. Test conditions: 87.5% H2O2, φ = 1.40, and CR = 3.0. ............................................................................................. 87

Figure ?5.3: a) Full and b) partial power spectrum of monoprop chamber pressure data shown boxed in Figure ?5.2a)..................................................................................... 88

Figure ?5.4: Variation in C* with φ for H2O2 combusting with JP-8.................................. 92

Figure ?5.5: Plots of pfu_inj and pc2 at the point of injection for a strong autoignition test using: a) the GK catalyst bed using 90% H2O2 at φ = 1.59 and CR = 3.0 and b) the PCI catalyst bed using 90% H2O2 at φ = 1.58 and CR = 3.0..................................... 98

Figure ?5.6: Frame-by-frame view of a strong autoignition test using GK catalyst bed. Test conditions: 90% H2O2, φ = 1.59, CR = 3.0. ...................................................... 99

Figure ?5.7: Frame-by-frame view of a strong autoignition test using PCI catalyst bed. Test conditions: 90% H2O2, φ = 1.58, CR = 3.0. .................................................... 100

Figure ?5.8: a) Unfiltered and b) filtered plots of pfu_inj and pc2 at the point of injection for a weak autoignition test using the GK catalyst bed. Test conditions: 87.5% H2O2, φ = 1.62, and CR = 3.0. ........................................................................................... 101

Page 11: Thesis

xi

Figure Page

Figure ?5.9: Frame-by-frame view of a weak autoignition test using GK catalyst bed. Test conditions: 87.5% H2O2, φ = 1.62, CR = 3.0. ......................................................... 102

Figure ?5.10: Plot of pfu_inj and pc2 at the point of injection for a test using PCI catalyst bed. Test conditions: 94% H2O2, φ = 2.15, and CR = 3.0. ............................................. 105

Figure ?5.11: Frame-by-frame view of a test with no autoignition using PCI catalyst bed. Test conditions: 94% H2O2, φ = 2.15, CR = 3.0. .................................................... 106

Figure ?5.12: Autoignition of JP-8 as a function of φ and H2O2 concentration at CR = 3.0.................................................................................................................................. 108

Figure ?5.13: Autoignition of JP-8 as a function of equivalence ratio and contraction ratio at a constant H2O2 concentration of 85%................................................................ 109

Figure ?5.14: Measured Td over the course of tests run at a) 90% H2O2 at φ = 1.59 and CR = 3.0 with GK bed and b) 98% H2O2 at φ = 1.84 and CR = 3.0 with PCI bed. 110

Figure ?5.15: Comparison in decomposition efficiency produced by GK and PCI catalyst beds using 90% H2O2 at approximately equivalent monoprop operating conditions.................................................................................................................................. 112

Figure ?5.16: Comparison in pressure drop across GK and PCI catalyst beds using 90% H2O2 at approximately equivalent monoprop operating conditions. ...................... 113

Figure ?5.17: Plot of the chamber pressure noise parameter against the dominant frequency during the monoprop mode each autoignition test................................. 114

Figure ?5.18: Plot of the chamber pressure noise parameter against the dominant frequency during the biprop mode each autoignition test....................................... 114

Figure ?5.19: Comparison of DMAZ autoignition points to those of JP-8 using the GK catalyst bed at a CR = 3.0. ...................................................................................... 117

Figure ?5.20: Autoignition of JP-8 as a function of equivalence ratio and decomposition temperature at a contraction ratio of 3.0. ................................................................ 119

Figure ?5.21: Autoignition of JP-8 as a function of pc2_tot_mono and decomposition temperature at a contraction ratio of 3.0. ................................................................ 120

Page 12: Thesis

xii

Figure Page

Figure ?5.22: Trajectory variations in tests run at a contraction ratio of 3.0.................... 122

Figure ?5.23: Trajectory variations in tests run at a contraction ratio of 5.0.................... 122

Figure ?A.1: Engine assembly using GK catalyst bed, page one. .................................... 138

Figure ?A.2: Engine assembly using GK catalyst bed, page two. .................................... 139

Figure ?A.3: Extension piece for GK catalyst bed, allows temperature and pressure measurement upstream of the fuel injector. ............................................................ 140

Figure ?A.4: Mounting plate for engine assembly using GK catalyst bed. ...................... 141

Figure ?A.5: Page one of transverse fuel injector drawing, indicates manifold dimensions.................................................................................................................................. 142

Figure ?A.6: Page two of transverse fuel injector drawing, indicates orifice dimensions.................................................................................................................................. 143

Figure ?A.7: Transverse fuel injector seat, the fuel feed line is attached to this piece..... 144

Figure ?A.8: Drawing of fuel film cooling injector seat, fuel feed line was capped for autoignition testing. ................................................................................................. 145

Figure ?A.9: Drawing of fuel film cooling injector, fuel was not flowed through this piece during autoignition testing. ..................................................................................... 146

Figure ?A.10: Page one of combustion chamber part drawing. ....................................... 147

Figure ?A.11: Page two of combustion chamber part drawing. ....................................... 148

Figure ?A.12: Page three of combustion chamber part drawing. ..................................... 149

Figure ?A.13: Page four of combustion chamber part drawing. ....................................... 150

Figure ?A.14: Drawing of nozzle piece with a contraction ratio of 3.0. .......................... 151

Figure ?A.15: Drawing of nozzle piece with contraction ratio of 5.0. ............................. 152

Page 13: Thesis

xiii

Figure Page

Figure ?A.16: Drawing of nozzle piece with contraction ratio of 6.5. ............................. 153

Figure ?A.17: Assembly drawing of water cooled deflection plate, water cooling apparatus not shown. ............................................................................................................... 154

Figure ?A.18: Engine assembly using PCI catalyst bed. .................................................. 155

Figure ?A.19: Mounting plate used for engine assembly using PCI catalyst bed, this piece was attached to the top of the catalyst bed.............................................................. 156

Figure ?A.20: Page one of drawing of transition piece between PCI catalyst bed and transverse fuel injector. This piece also allowed the measurement of pressure and temperature at the exit of the catalyst bed............................................................... 157

Figure ?A.21: Page two of drawing of PCI transition piece, indicates dimension of V-shaped groove for metal o-ring. .......................................................................... 158

Figure ?C.1: Plumbing & Instrumentation Diagram of Test Cell A at APCL. ................ 166

Page 14: Thesis

xiv

ABSTRACT

Sisco, James C. M. S., Purdue University, August, 2003. Autoignition of Kerosene by Decomposed Hydrogen Peroxide in a Dump Combustor Configuration. Major Professor: William E. Anderson.

In staged-bipropellant rocket combustors that use decomposed hydrogen peroxide as

the oxidizer, a liquid fuel is injected into the hot decomposition products comprising

oxygen and water vapor. The oxidizer is at a sufficiently high temperature to vaporize

and to autoignite the liquid fuel. Although the need for a separate ignition system is

eliminated with this configuration, two other issues arise: it is difficult to efficiently mix

a relatively small amount of liquid fuel into a large volumetric flow of oxidizer at the

performance-optimized mixture ratios of about eight; and the combustor design must

provide residence times sufficient for autoignition. The latter issue typically results in the

use of high combustion chamber contraction ratios with their attendant higher weight and

surface area cooling requirements. In this study a transverse injector was used in a dump

combustor configuration, which incorporates a rearward-facing step, to investigate the

autoignition characteristics of JP-8 in decomposed hydrogen peroxide. The goals of the

investigation were to develop a greater understanding of the autoignition process and, if

possible, develop autoignition model for a staged combustor. The chamber contraction

ratio was varied between three and five to evaluate the effects of chamber gas Mach

number, and the hydrogen peroxide concentration was varied from 85 to 98% to evaluate

the effects of oxidizer temperature. Results showed that as hydrogen peroxide

concentration and/or contraction ratio was increased the fuel-rich equivalence ratio which

defined the autoignition boundary increased as well. At a contraction ratio of 3.0, no

autoignition was achieved down to an equivalence ratio of 1.37 using 85% hydrogen

Page 15: Thesis

xv

peroxide, but at 98% hydrogen peroxide autoignition occurred up to an equivalence ratio

of 2.06. When the contraction ratio was increased to 5.0 autoignition was achieved at an

equivalence ratio of 1.38 using 85% hydrogen peroxide. More data is needed rega rding

the effects of pressure and decomposed gas Mach number to develop an accurate

autoignition model. However, the use of flame stabilization provided by the

rearward-facing step improved the autoignition range of the combustor according to past

engine data.

Page 16: Thesis

1

CHAPTER 1: INTRODUCTION Hydrogen peroxide and kerosene rocket engines have a long history of use in

propulsion systems dating back prior to World War II.1-3,11 Although the performance of

this propellant combination is not as high as liquid oxygen/liquid hydrogen, LOx/LH2, or

nitrogen tetroxide/mono-methyl hydrazine, NTO/MMH, systems it is still a very

appealing option for a number of reasons. Hydrogen peroxide, H2O2, is a very versatile,

highly reactive, high density, storable, and non-toxic oxidizer. The versatility of

hydrogen peroxide is, in a way, a result of its reactivity. It can be decomposed and used

as a monopropellant for reaction control, to drive a turbine, or as a pressurant. Kerosene

based fuels such as Jet-A, JP-8, and RP-1 are very commonly used in the aviation and

rocket industries. These fuels are also storable and non-toxic. An important feature of

this propellant combination is its high density specific impulse, which is defined as the

total impulse delivered per unit volume of propellant. The density specific impulse of

H2O2/kerosene when compared to typical rocket propellant combinations is exceeded

only by the NTO/MMH system. Table 1.1 outlines performance parameters of common

rocket systems operating at similar conditions.4,5

Spacecraft reaction control, RCS, and orbital maneuvering systems, OMS, have

typically used hydrazine and NTO/MMH rocket systems since the 1960’s.6 This was due

to the storability of the propellants, hydrazine’s high performance as a monopropellant,

and the fact that nitrogen tetroxide and mono-methyl hydrazine are hypergolic or ignite

on contact. These factors made hydrazine and NTO/MMH systems very simple and

reliable.6 However, all three propellants are toxic and corrosive while hydrazine is a

carcinogen. This creates significant safety hazards when trying to handle the propellants.

As a result, there is significant interest in developing rocket systems using non-toxic

propellants to replace hydrazine and NTO/MMH systems.6 There is also increased

Page 17: Thesis

2

interest to develop low cost, reusable satellite launch vehicles to replace the expendable

vehicles currently used in industry.19 Many of these expendable vehicles use toxic or

cryogenic propellants, such as liquid oxygen and hydrogen. Storable, non-toxic

propellants are also preferred for these launch vehicle applications for ease of handling

on the ground.

Table 1.1: Performance comparison of 90% H2O2/JP-8 system to common rocket propellant combinations. Specific impulse calculated assuming a chamber pressure of

1000 psia and equilibrium expansion to sea level pressure of 14.7 psia.

Oxidizer Fuel

90% H2O2 JP-8

NTO MMH

LOx LH2

LOx RP-1

Optimum O/F Ratio 7.8 2.2 3.5 2.6

Characteristic Velocity, C* (ft/s) 5300 5710 7940 5890

Chamber Temperature, Tc (°F) 4600 5650 4450 6160

Specific Impulse, Isp (sec)

267 288 386 300

Density Specific Impulse, Density Isp (sec) 344 346 101 308

Hydrogen peroxide and kerosene rocket systems have the potential to replace

their toxic and cryogenic predecessors. Table 1.1 shows that both NTO/MMH and

H2O2/kerosene systems have comparable density specific impulse, density Isp, and are

both higher than cryogenic systems. This means that per unit volume of propellant a

H2O2/kerosene systems offer similar if not superior performance compared to

conventional propellant combinations. On a per unit mass basis hydrogen peroxide

systems are not quite as good performers. As a monopropellant hydrogen peroxide has a

lower specific impulse, Isp, than hydrazine and a bipropellant H2O2/kerosene system also

has lower Isp than NTO/MMH, LOx/RP-1, and LOx/LH2. However, analyses have

shown that hydrogen peroxide/kerosene systems may be the most cost effective for future

launch vehicles regardless of mass-based performance.19

Page 18: Thesis

3

There are some technical issues associated with these bipropellant systems that

must be resolved to make it a viable replacement for NTO/MMH and current launch

vehicle propellants. Since NTO and MMH are hypergolic it makes the system very

simple in design, it is desired that a H2O2/kerosene system have similar simplicity as

well.6 Hydrogen peroxide and kerosene are not hypergolic by themselves, and there is

research being done make these propellants ignite on contact.6,7,12,13 Alternatively, an

H2O2/kerosene engine can operate in a staged configuration. In this configuration the

hydrogen peroxide is decomposed in a catalyst bed and the kerosene fuel is injected into

the hot decomposed gases. If conditions are correct the oxidizer/fuel mixture can

autoignite eliminating the need for a complex ignition system. However, autoignition is

dependent on a number of different factors such as fuel injector design, hydrogen

peroxide concentration, decomposed gas velocity, chamber pressure, and mixture

ratio.8-11 A better understanding of the autoignition process in these staged

H2O2/kerosene rocket engines is required. The goals of the research described in this

thesis include; outlining a design method for a staged engine injector, generating

experimental data on autoignition under varying engine operating conditions, and

creating a model to aid in the prediction of autoignition. Results of this research may

make the staged-bipropellant H2O2/kerosene rocket a lighter, more reliable, and higher

performing engine in the future.

1.1 Hydrogen Peroxide

Hydrogen peroxide is an inherently unstable chemical compound that

exothermically reacts, or decomposes, into hot oxygen gas and water vapor. Hydrogen

peroxide is miscible in water and is commercially manufactured as an aqueous solution in

a variety of concentrations. Concentrations are usually designated as percent H2O2 by

weight of solution. Propellant-grade H2O2, or HTP, is greater than 70% in concentration

and most modern engines tend to use 85, 90, or 98%. The decomposition rate of

propellant-grade H2O2 is less than 0.1% per year over normal atmospheric temperature

and pressure ranges.3 Decomposition is significantly accelerated as the temperature of the

Page 19: Thesis

4

H2O2 and/or its environment is increased and/or when the liquid is in contact with certain

materials or contaminants. These factors can potentially cause a chain reaction of

decomposition since the heat released during a reaction can provide the energy necessary

to decompose the surrounding H2O2 and so on. This is a very dangerous situation in most

cases, however, when controlled it can be advantageous quality. Table 1.2 outlines the

variation in physical and decomposition properties of hydrogen peroxide with

concentration.3,4

1.1.1 Benefits of Hydrogen Peroxide as a Propellant

There are many properties of hydrogen peroxide that make it an appealing choice

as an oxidizer in a rocket propulsion system.1 The density of hydrogen peroxide at a

concentration of 90%, for example, is equivalent to or greater than most common rocket

oxidizers. It has a specific gravity of approximately 1.4, see Table 1.2, whereas liquid

oxygen and nitrogen tetroxide have specific gravities of 1.14 and 1.44 respectively.5

High density propellants are extremely beneficial in rocket systems because more

propellant mass can be stored in a smaller volume. This generally reduces tank volume

and also system mass.

Table 1.2: Properties of liquid and decomposed hydrogen peroxide with concentration.3-5

Concentration 70 % H2O2 80 % H2O2 90 % H2O2 98 % H2O2 Liquid Properties (@ STP)

Molecular Weight 26.86 28.89 31.29 33.42 Specific Gravity 1.283 1.333 1.387 1.432 Boiling Point (F) 257 -- 287 299

Vapor Pressure (psia) 0.137 -- 0.065 0.045 Heat Capacity (Btu/lbm-R) 0.738 -- 0.663 0.633

Decomposed Gas Properties Temperature (F) 504 952 1393 1746

Molecular Weight 21.04 21.56 22.11 22.56 Specific Heat Ratio 1.315 1.287 1.265 1.251 Mass Fraction O2 0.341 0.376 0.423 0.461

Mass Fraction H2O 0.659 0.624 0.577 0.539

Page 20: Thesis

5

Hydrogen peroxide has a low vapor pressure, as Table 1.2 shows, on the order of

one-tenth of a psi. This is significantly lower than the vapor pressure of other common

oxidizers such as liquid oxygen, 735 psia at -193 °F, and nitrogen tetroxide, 110 psia at

160 °F.5 It is advantageous to use a propellant with a low vapor pressure in rocket

system for several reasons.1,14 In turbo-pump systems the propellant can be fed to the

pumps at a low pressure without risking cavitation. In addition, only a low absolute

pressure is required in the propellant tank to prevent the liquid from vaporizing. As a

result, use of a propellant with a low vapor pressure leads to low tank and system

pressures which reduce tank and system mass. Another attractive feature of hydrogen

peroxide is its high heat capacity, 0.66 Btu/lbm-R for 90% H2O2 as shown in Table 1.2.

This is comparable to the heat capacity of water, 1.0 Btu/lbm-R, which is considered a

very good coolant and is used for a number of applications. The high heat capacity of

hydrogen peroxide suggests that it would be an excellent coolant for a rocket system.

Hydrogen peroxide also possesses the properties of a storable propellant. It is a

stable liquid over a reasonable range of temperature and pressure, and it is sufficiently

non-reactive with tank material, when properly passivated, for significant lengths of time,

although the concentration will gradually decrease.1-3 It is considered to be a non-toxic

propellant as well. Toxic propellants are poisonous to humans through inhalation or

contact with the body tissue. However, hydrogen peroxide solutions and vapors are

irritating to body tissue. Solutions can cause skin burns and vapors can inflame the

respiratory tract, however, exposure is only lethal in extremely high doses especially

through ingestion.3

The most important feature of hydrogen peroxide as propellant is its reactivity.

The hot gases produced when H2O2 is decomposed contain a significant amount of

energy, see Table 1.2. This makes hydrogen peroxide an excellent monopropellant.

Monopropellant thrusters are typically used for low thrust applications such as reaction

control systems (RCS). Hydrogen peroxide of 85 or 90% concentration has been used in

RCS systems in the past, such as the Mercury space capsule, and new systems using

H2O2 are currently in development.1,27 These gases can also be expanded through a row

of turbine blades imparting its energy to generate turbine rotation. This is important

Page 21: Thesis

6

since many rocket systems use turbo-pumps driven by turbines to feed propellants to the

combustion chamber. Decomposed hydrogen peroxide could potentially be used as a

tank pressurant as well. The reactivity of H2O2 also makes it a versatile propellant that

can be used for a number of different propulsion systems. Using hydrogen peroxide as an

oxidizer in a bipropellant main engine as well as a monopropellant for reaction control

and turbine power eliminates the need for separate systems for each of these applications.

This greatly simplifies the overall propulsion system design.

1.1.2 Catalyst Bed Design

Hydrogen peroxide is typically decomposed using a catalyst bed. Most modern

catalyst beds consist of many layers of wire mesh screens, typically stainless steel or

nickel, coated with a catalyst material, typically silver.1,2 Initially, before ‘dry’ catalyst

beds were used, liquid catalysts such as potassium permanganate were injected with

liquid hydrogen peroxide to accelerate decomposition.2,6,11 These liquid catalysts worked

quite well however the need for separate systems for supplying the liquid catalyst made it

operationally cumbersome. Later, ‘dry’ catalyst beds used beads or pellets impregnated

with a catalyst material to react with the hydrogen peroxide.1,11 These catalyst beds also

worked well, but had a very limited operational life.11 A significant drawback to silver

plated catalyst beds is that they are limited to concentrations of approximately 92% or

less. At higher concentrations the decomposition temperature of the H2O2 exceeds the

melting temperature of silver, which is approximately 1760°F, and leads to screen

degradation. Catalyst beds have been tested with catalytic materials with higher melting

temperatures, such as platinum and silver-palladium, for use with up to 98% H2O2.28

In many cases only the first third to half of the screens are coated with silver.2

This is due to the unstable nature of the H2O2, as the peroxide begins to decompose on

the first few screens it releases hot gases. These hot gases heat the surrounding liquid

hydrogen peroxide and this H2O2 begins to decompose and so on. Thus as the hydrogen

peroxide makes its way through the catalyst bed screens the decomposition process relies

more and more heavily on thermal decomposition rather than catalyst material. This is a

Page 22: Thesis

7

very complex design dependent on a number of parameters such as hydrogen peroxide

concentration, bed loading (mass flow rate/cross-sectional area), screen mesh or open

area, screen wire diameter, catalyst coating process, and back pressure among others.28

Ideally a good catalyst bed design should produce highly decomposed peroxide at as

small of a pressure drop as possible such that it is still isolated from oscillations in back

pressure.

1.2 Background on Hydrogen Peroxide Bipropellant Engines

1.2.1 Hypergolic Bipropellant Engines

There are two main configurations that have been used in the past and that are

currently being considered for bipropellant hydrogen peroxide engines. The most

simplistic type is a hypergolic configuration whereby a fuel is used that ignites with

liquid hydrogen peroxide on contact in the combustion chamber. In fact, this

configuration was used on the first bipropellant rocket developed using hydrogen

peroxide in Germany in 1935.1,2,6,11 The purpose of this engine, developed by Hellmuth

Walter, was to provide additional thrust at takeoff for an early jet-powered aircraft, the

Messerschmidt 163B.2 At that time only 80% H2O2 was available and its decomposition

temperature was too low to ignite kerosene-based fuels. Therefore Walter developed a

fuel, called C-Stoff, consisting of 30% hydrazine hydrate and 70% methanol, which was

hypergolic with H2O2.2,6,11 Since that time hydrogen peroxide has become available in

higher concentrations. As a result, most engine designers chose to use decomposed

hydrogen peroxide as an ignition source rather than hypergolic fuels.

Recently there has been increased interest in developing rocket systems that use

fuels hypergolic with hydrogen peroxide to replace NTO/MMH systems. These fuels

have been mainly kerosene or methanol based containing additives that make them

hypergolic with H2O2.6,7,12,13 Typically, complex injector designs are required in these

hypergolic engines in order to adequately atomize the propellants. In addition, the

Page 23: Thesis

8

injector must deliver the propellants in such a way that they can mix readily. Recent

studies have used pintle, splash plate, and coaxial swirl designs to achieve the necessary

mixing and atomization.12,13 To this date there has been no hypergolic hydrogen peroxide

engine used in a flight-rated system.

1.2.2 Staged-Bipropellant Engines

The second major bipropellant configuration, and also the most often used to date,

is the staged design. This configuration uses a catalyst bed to decompose the hydrogen

peroxide creating hot oxygen gas and water vapor. Further downstream kerosene fuel is

injected into the high velocity gas where it is atomized, vaporized, and, under the correct

conditions, autoignited in the combustion chamber. This engine configuration grew in

popularity following the development of the silver screen catalyst bed.

1.2.2.1 Staged Engine Development in the United Kingdom

Work on these types of engines began in the United Kingdom, U.K., in the late

1940’s following World War II and was derived from the pre-war work done in Germany.

Early research centered on 80% H2O2 decomposed using catalyst stones and aviation

kerosene fuel.10,11 Much of this early development work was done to better understand

how these rocket systems operated and more importantly what conditions were necessary

for autoignition. By the end of 1950 silver screen catalyst beds had been developed and

were being used to decompose 80% hydrogen peroxide.10 These rocket systems were

designed to provide thrust augmentation for jet-powered manned interceptor aircraft and

to power ground-to-air missiles.2,11 Since the interceptor aircraft bearing these systems

were fueled by kerosene it was only logical to use kerosene as a rocket fuel as well.

Two engine classes came out of these early efforts in the United Kingdom were

designated as the KP and Gamma series. The KP-3 engine, developed for the Red Shoes

ground-to-air missile, used kerosene fuel and a silver screen catalyst bed to decompose

Page 24: Thesis

9

83% H2O2 and produced a maximum thrust of approximately 3300 lbf.11 The Gamma

class of engines was developed as a backup to the de Havilland Spectre engine, which

used H2O2 and kerosene, designed for the Saunders-Roe SR 53 interceptor.1,2 The early

Gamma engines used 85% H2O2 and silver screen catalyst beds.2,11 The Gamma class

was very successful and numerous variants of the engine were developed most notably

the Gamma 2, Gamma 201, Gamma 301, Gamma 304, Gamma 8, BS 606, and BS

625.1,14,15 Each of these engines was developed for use on the U.K.’s Black Knight re-

entry vehicle and/or the Black Arrow satellite launch vehicle. The engines ran on 85 or

87% H2O2 decomposed in a silver screen catalyst bed. These engines are considered the

highest performing hydrogen peroxide engines ever put into production.1 The Black

Arrow was the only launch vehicle, until the late 1990’s, ever to be developed using

hydrogen peroxide as its oxidizer. The U.K. drew on its wealth of experience, gained

from the late 1940’s through the early 1970’s, to produce these high performance

H2O2/kerosene engines. Gradually these engines were replaced by NTO/MMH,

LOx/RP-1, and LOX/LH2 systems. Currently, there is new work being done in the U.K.

to develop small thrusters for satellite propulsion using a staged-bipropellant design and

kerosene fuel.16,17

1.2.2.2 Staged Engine Development in the United States

In the United States, U.S., as in the U.K. work on hydrogen peroxide engines

began in the 1950’s to develop rocket-assist systems for jet-powered aircraft.1

Rocketdyne, through the U.S. Air Force, developed the AR series of engines that used

90% H2O2 and kerosene. The final result of this effort was the AR2-3 engine which

developed approximately 6000 lbf thrust and had impressive records of reliability and

reusability.1 Independently, at the same time Reaction Motors Incorporated, through the

U.S. Navy, developed the LR-40 engine for use on the F8U.1,18 This engine also used

90% H2O2 and kerosene fuel and generated somewhere between 3500 and 10000 lbf

thrust.18 As aircraft jet engines became more powerful these engines were gradually

phased out of use.

Page 25: Thesis

10

Bipropellant hydrogen peroxide engine research and development remained

dormant in the U.S. until the mid 1990’s. At that time low cost launch vehicles were in

demand to replace existing launch systems. Cost analyses, such as that done by Frazier

and Moser, showed that hydrogen peroxide/kerosene systems could be a less expensive

alternative.19 In 1998, Beal Aerospace was formed with a goal to develop both an

expendable and reusable satellite launch vehicle, named the BA-1 and BA-2R

respectively.20 Both launchers were of a three-stage configuration and used a

staged-bipropellant rocket design, using 90% hydrogen peroxide and kerosene, for each

stage.20 Testing began in 1998 with the first firing of the BA-44 third stage engine,

44,000 lbf vacuum thrust.21 Then, in early 2000, Beal ground tested its 810,000 lbf

vacuum thrust second stage engine, the BA-810.21 This engine was the second largest

liquid rocket engine fired in the United States since the Apollo program.21 The first stage

engine was designed to produce 4.1 million- lbf vacuum thrust and would have been the

largest rocket engine ever developed.21 However, in late 2000 the company was

dissolved due to financial reasons and Beal was never able to fire neither a launch vehicle

nor a first-stage engine.

At the roughly the same time the National Aeronautics and Space Administration,

NASA, commissioned the Orbital Sciences Corporation to begin development of a

10,000 lbf vacuum thrust hydrogen peroxide/kerosene stage-bipropellant engine for the

Upper Stage Flight Experiment, USFE.22,23 The purpose of the USFE was to focus on

key technical issues and to demonstrate the operation of a hydrogen peroxide/kerosene

system for a low cost liquid upper stage rocket system.23 Some of these key issues

included catalyst bed design for use with 90% H2O2, kerosene fuel injector design, and

ablative chamber design.22,23 Tests proved that the engine was able to meet project

requirements.23 Currently, the U.S. is still in the midst of staged-bipropellant engine

development and a complete system has yet to be flown.

Page 26: Thesis

11

1.3 Design of Staged-Bipropellant Engines

A staged-bipropellant rocket using hydrogen peroxide and kerosene has a number

of design related advantages. As previously described, in these staged engines the H2O2

is decomposed in a catalyst bed. Downstream of the catalyst bed liquid kerosene is

injected into the high velocity decomposed gas flow. The fuel is atomized and vaporized

creating a mixture of fuel vapor, oxygen gas, and water vapor. Depending on the

operating conditions the fuel/oxidizer mixture can autoignite in the combustion chamber.

A typical staged combustor design is shown in Figure 1.1. In this design all of

the decomposed hydrogen peroxide flows through a central gas or steam port. Liquid

kerosene is injected perpendicularly, or transversely, to the gas flow in the port through a

number of circumferentially arranged orifices. The momentum possessed by the gas

causes the liquid fuel jets to deflect and enhances atomization. Downstream of the fuel

injection point the gas port is suddenly expanded to match the combustion chamber

diameter. The expansion is commonly called a rearward-facing step. This step acts as a

flame holder and creates a recirculation zone that continually feeds reactants to the flame.

Figure 1.1: Typical staged-bipropellant engine using H2O2/kerosene. This is commonly called a ‘dump’ combustor configuration.

Page 27: Thesis

12

1.3.1 Advantages of Staged-Engines

1.3.1.1 Combustion Performance

An important aspect of the hydrogen peroxide/kerosene propellant combination is

its relatively flat characteristic velocity, C*, versus mixture ratio curve. Mixture ratio is

defined as the ratio of the oxidizer to fuel flow rate. Characteristic velocity is a measure

of propellant, or combustion, performance in the chamber independent of the rest of the

engine. The shape of the C* curve infers that a H2O2/kerosene system can provide

similar combustion performance over a wide range of mixture ratios. Figure 1.2 shows a

plot of C* versus mixture ratio for 90% H2O2/JP-8, NTO/MMH, LOx/RP-1, and

LOx/LH2.4 The figure shows that the 90% H2O2/JP-8 curve has a very flat, gradual peak

while the remaining propellants combinations have a sharp peak in their C* curves. The

decay of the curve for 90% H2O2/JP-8 at high mixture ratios is also less than the other

propellant combinations. This offers significant flexibility when designing a wall cooling

scheme for these engines. In a fuel or oxidizer film cooling scheme some of the fuel or

oxidizer is used as a buffer flow to maintain the chamber wall at a low temperature. A

majority of this propellant may not mix and combust with the core flow affecting the

overall reacting mixture ratio in the chamber. These small deviations in actual

combusting mixture ratio will not have a large affect on engine performance in a

hydrogen peroxide/kerosene system.

Page 28: Thesis

13

2000

3000

4000

5000

6000

7000

8000

0 2 4 6 8 10 12 14 16 18 20

Mixture Ratio, O/F (--)

Ch

arac

teri

stic

Vel

oci

ty, C

* (f

t/s)

90% H2O2/JP-8NTO/MMHLOx/RP-1LOx/LH2

Figure 1.2: Comparison of C* vs. mixture ratio curves of 90% H2O2/JP-8 to common rocket propellant combinations.4 C* calculated assuming a chamber pressure of 1000

psia and equilibrium expansion.

1.3.1.2 Injection Techniques

In many liquid- liquid systems impinging injector elements are required to atomize

and mix the propellants. These injectors perform very well, but require precise

machining and complicated design methods. In a staged engine design the high velocity

gas flow enhances the atomization of the liquid fuel. As a result a more simplistic

injector design, which in some cases may not provide as good atomization as an

impinging element, can be used to obtain similar engine performance. Good atomization

is important in a rocket engine because smaller fuel drops require less time to vaporize.

As a result, the fuel vapor and gaseous oxidizer have more time to mix and burn in the

chamber producing higher combustion performance. In the past designers have used

swirl, 8-10,24,25 gas port,15,29 or ring injector22,23 designs or combinations thereof. Figure

1.3 shows an engine schematic of a Gamma class engine.15 This engine uses a gas port

design with kerosene fuel injected at an angle to the gas flow from the downstream side

Page 29: Thesis

14

of the injector face. A picture of the injector is shown in Figure 1.4.15 There are 49 gas

ports in the injector and the fuel is well dispersed throughout the chamber by numerous

injection orifices located on the injector face. An injector design that was used in early

thermal ignition research done by Walder and Purchase is shown in Figure 1.5.11 This

injector uses a combination of injection concepts. A fuel swirl element is located in a

central gas port with a small mixing chamber downstream. A number of

circumferentially arranged orifices are used to inject a portion of the decomposed gases

perpendicularly to this central flow. The remaining gas flows through an array of large

gas ports located on the periphery of the injector. A thin annular gap located behind

these ports deflects the gas flow toward the center of the chamber.

Figure 1.3: Schematic of a Gamma class research engine indicating gas ports and fuel injection points.15

Page 30: Thesis

15

Figure 1.4: Picture of a Gamma class gas port injector.15

Figure 1.5: Early injector design used in thermal ignition research.11 This injector uses a

combination of basic injection concepts such as swirl and gas port injectors.

Page 31: Thesis

16

1.3.1.3 Thermal Benefits

Another advantage of this system that may often be overlooked is the relatively

low combustion temperature of the propellants even at the stoichiometric mixture ratio,

see Figure 1.6.4 Although a cooling scheme is still required in the chamber and throat to

prevent damage to wall materials, the overall heat load compared to LOx/RP-1, LOx/LH2,

and NTO/MMH systems may be significantly lower. This could make the lifetime of a

robustly designed H2O2/kerosene engine significantly longer than an engine using

industry standard propellants.

1000

2000

3000

4000

5000

6000

7000

0 2 4 6 8 10 12 14 16 18 20

Mixture Ratio, O/F (--)

Co

mb

ust

ion

Tem

per

atu

re, T

c (F

)

90% H2O2/JP-8NTO/MMHLOx/RP-1LOx/LH2

Figure 1.6: Comparison of chamber temperature vs. mixture ratio of 90% H2O2/JP-8 against other common propellant combinations.4 Combustion temperature calculated

assuming a chamber pressure of 1000 psia and equilibrium expansion.

Page 32: Thesis

17

1.3.1.4 Autoignition

An important benefit of a staged combustor is that the decomposed hydrogen

peroxide eliminates the need for a complex ignition system, such as a torch igniter.

Difficulties in obtaining ignition in the vacuum of space are reduced as well because the

decomposed H2O2 provides a pre-pressurized/pre-heated chamber environment prior to

fuel injection. However, a challenge exists in determining what engine operating

conditions and geometry are sufficient to cause autoignition. In some cases, designers

have found that the propellants will not ignite under certain temperature, pressure, and

chamber geometry conditions.8-10,17,22

1.3.2 Challenges in Staged Engine Design

1.3.2.1 Fuel Distribution and Thermal Management

There are also other challenges that arise when designing a staged-bipropellant

rocket using H2O2 and kerosene. Depending on the concentration of H2O2, the optimum

mixture ratio for this propellant combination is on the order of eight-to-one. This means

that 88% of the total mass in the combustion chamber is decomposed hydrogen peroxide.

A problem can arise, as a result, in distributing the relatively small amount of fuel

uniformly in the chamber. This is especially difficult when attempting to fuel film cool

the combustion chamber wall. Chamber cooling, or thermal management, is also a

problem associated with these engines. Due to the lack of large amounts of fuel

conventional cooling techniques are difficult.

An example of the thermal management problems associated with these engines

was encountered during the development of Orbital Sciences Corporation’s Upper Stage

Flight Experiment, USFE, hydrogen peroxide/kerosene engine.23 The engine design

utilized a silica phenolic chamber liner and as such a cool wall was required to prevent

Page 33: Thesis

18

severe ablation of the liner. A ring injector was developed to provide fuel film cooling

along the chamber wall.23 A picture of this injector is shown in Figure 1.7.23 The

injector design is quite complex and is an example of the types of injection techniques

required to keep a cool chamber wall in these engines. A variation of this injector was

used in a prior study using a ‘dump’ combustor configuration.22 In this study additional

kerosene fuel was injected axially along the chamber wall by orifices on the face of the

rearward step.22 A similar ‘dump’ combustor was developed for a rocket-based

combined cycle application and used an axial injector to provide fuel film cooling.26

The combustion chambers of the U.K.’s Gamma class H2O2/kerosene engines

were regeneratively cooled with hydrogen peroxide.14,15,29 This method of cooling is

potentially dangerous. The walls of the coolant channels can become too hot causing

local H2O2 decomposition in the passages. This phenomenon has been observed in the

U.K. in the past and had not caused any significant failures, but it has caused minor

structural and thermal damage to the coolant channels.2 Interestingly enough, the

decomposition temperature 85, 90, and 98% H2O2 are less than most common chamber

wall materials, such as copper and stainless steel. Alternate designs have be proposed

and tested where decomposed gases are used to provide wall film cooling.17,24

1.3.2.2 Performance and Weight

In the end, the low mass-based performance of these propellants may its biggest

disadvantage. The low specific impulse of this propellant combination translates to a

large mass of propellant. This increases structural weight and makes the overall system

very heavy. Catalysts beds, specifically screen beds, are typically heavy as well and their

weight scales with mass flow and chamber pressure. In the past these engines have had

uncommonly large combustion chambers as well. Nearly all past staged engines have

used long chamber lengths and high contraction ratios, which is the ratio of the chamber

to throat cross-sectional area. A paper by Wu et al reports the contraction ratios used by

several past engines at 4.0 or greater.22 A parameter that is commonly used to describe

the chamber geometry is characteristic length, L*. Characteristic length is defined as the

Page 34: Thesis

19

ratio of the chamber volume to the throat area. The characteristic lengths of past engines

have ranged from 30 to 75 inches.17,22,25 The propellant mixture requires significant time

in the chamber to vaporize and react necessitating a long chamber and low gas velocity.

This not only a problem from the stand point of chamber weight but it also creates a

larger chamber area to cool.

Figure 1.7: Picture of USFE ring injector.23 The tubes and crescent-shaped manifolds

are used to feed fuel along the chamber wall.

Page 35: Thesis

20

CHAPTER 2: INJECTOR DESIGN

In a staged H2O2/kerosene engine simple injection concepts can be designed to

use the decomposed gas flow to enhance fuel atomization. In the past swirl injectors

have seen extensive use in staged engine systems.8-10,24,25 One reason for this is that the

swirl injector produces a thin hollow cone-shaped spray of liquid fuel that is finely

atomized by the decomposed gas flow. In addition, these injectors also have a well

established design methodology.31-33 The geometry of the swirler directly influences the

velocity, thickness, and cone angle of the fuel spray. In a typical design the swirl

geometry is contained in the center of the combustion chamber acting as a bluff body.

Decomposed gas flows co-centrically around the swirler section (sometimes referred to

as a swirl-coaxial design). This bluff body creates a recirculation zone and acts as a

flame stabilizer. An interesting benefit to this injector is it ability to keep the wall of a

combustor cool without additional injector hardware.24 Since the decomposed peroxide

flows along the outside of the injector, or the inside wall of the chamber, it acts as a film

coolant protecting the wall from the hot combustion gases.

Another simple injector design that was used in Gamma class engines is the gas

port injector.15,29 In this design the decomposed gas flows through multiple ports having

large cross-sectional areas. Liquid fuel is injected into the gas flow from orifices located

on the walls of the gas ports or on the injector face. These injectors rely heavily on the

decomposed gas for atomization. A variation of this design is the transverse or crossflow

injector, which has also been used in the past.22 In a typical transverse injector design

multiple jets of liquid fuel are injected normal to the flow of decomposed gas. Drag

forces are exerted on the liquid jet by the gas as it tries to flow around it. These

aerodynamic forces cause the liquid jet to be deflected changing the bulk liquid trajectory.

Numerous empirical studies have been performed to attempt to predict the trajectory of

Page 36: Thesis

21

the liquid stream in non-reacting flow situations.34-36 Past experience has also shown that

jet trajectory can have an effect on engine performance.30 Drag forces create

aerodynamic instabilities along the surface of the liquid jet eventually causing jet breakup

and the formation of ligaments and droplets. Jet breakup can be classified into several

regimes,35,37,38 and empirical relations for average drop size have been developed.39,40 A

rearward-facing step is typically used in tandem with this type of injector configuration.

The step acts as a flame-holder creating a recirculation zone that promotes mixing and

combustion at relatively low velocity. This type of combustor configuration is usually

referred to as a dump combustor. Experimental and theoretical studies have identified an

optimal step height at which one obtains highest combustion efficiency.41-43

The transverse injector was chosen for this autoignition study because of its

heritage, simple design, and well characterized performance. This chapter will explain

the models and analyses used to design a transverse injector for a staged-bipropellant

rocket engine. The theory and methodology used to size the rearward-facing step is

discussed as well.

2.1 Engine Design

The staged combustor that was used for this autoignition study was initially

developed for a Rocket-Based Combined Cycle, RBCC, application.26 In this design a

silver screen catalyst bed was used to decompose 90% H2O2 and JP-8 was used as the

fuel. Design requirements called for a ten-second test firing of the engine. To achieve

this without damaging the combustion chamber and nozzle the engine was designed to

incorporate fuel film cooling, FFC. Two injectors were used; a transverse injector to

provide the core fuel flow and an axial injector to provide the film cooling flow. The

axial injector provided the geometry to create a rearward-facing step for flame

stabilization. The engine design is shown in Figure 1.1. The operating and geometrical

parameters of the RBCC engine are shown in Table 2.1. Mass flow and chamber

pressure conditions for the engine were used as a baseline for the injector design used in

this autoignition study.

Page 37: Thesis

22

Table 2.1: RBCC engine operating conditions. 26

Chamber Pressure (psia) 525 Core Mixture Ra tio (--) 5.0

Overall Mixture Ratio (--) 2.0 90% H2O2 Flow Rate (lbm/s) 2.23 Core JP-8 Flow Rate (lbm/s) 0.45 FFC JP-8 Flow Rate (lbm/s) 0.65

Chamber Diameter (in) 2.56 Chamber Length (in) 7.70 Contraction Ratio (--) 6.5 Expansion Ratio (--) 5.0

Thrust (lbf) 600

2.2 Transverse Injector Design Considerations

A schematic illustrating the important design parameters of the transverse injector

is shown in Figure 2.1. These parameters include the diameter, do, number, No, and the

length, lo, of the fuel orifices as well as the cross-sectional area of the fuel manifold, Am,

the fuel velocity exiting the orifice, Vf, the velocity of the decomposed gases, Vox, the

diameter of the gas port, Aox, the rearward step thickness, h, and the distance between the

injection point and the edge of the step, ls. A number of design considerations contribute

to determining a final value for each of these variables. One of the primary goals of the

design is to ensure that the injector delivers a reasonable jet trajectory and good

atomization. In addition, the injector must provide sufficient pressure drop to avoid

instabilities resulting from pressure fluctuations in the chamber. Also, the

rearward-facing step must be large enough to provide adequate flame holding but not so

large as to produce a large pressure loss. The transverse injector design methodology

described here centers around all three of these requirements.

Page 38: Thesis

23

Figure 2.1: Schematic of a transverse injector indicating important design parameters.

2.2.1 Orifice Sizing

Sizing of the fuel orifices begins by considering the pressure drop across the

orifice. As previously discussed the pressure drop across the injector is an important

parameter in a transverse injector design, or any other injector design for that matter.

Typically, the pressure drop is set to be 20% of the expected engine chamber pressure to

prevent combustion instabilities from affecting injector performance.44 The design

chamber pressure of the staged engine used in this study is about 525 psia, see Table 2.1,

which puts the design injector pressure drop at approximately 100 psid. The equation for

the pressure drop across an injector orifice is derived from Bernoulli’s equation.

Assuming no losses and neglecting gravity Bernoulli’s equation gives the following:

022

22

=−−+ i

f

ie

f

e VpVpρρ

, (2.1)

The properties at the inlet of the orifice are denoted with the subscript ‘i’ and

those at the exit with the subscript ‘e’. The cross-sectional area of the injector manifold

is typically designed to be much larger than that of the injector orifice to produce a very

low fuel velocity in the manifold. Therefore, for this analysis the inlet velocity, Vi, is

Page 39: Thesis

24

assumed to be zero. Equation 2.1 can then be solved for the pressure drop, ?pf, across

the injector orifice:

2

21

efeif Vppp ρ=−=∆ , (2.2)

Here ?f is the fuel density which is approximately 50.6 lbm/ft3 at room

temperature.47 Note that the right-hand side of equation 2.2 represents the dynamic

pressure of the fuel exiting the orifice. Since the design pressure drop has been set and

the fuel density is also known equation 2.2 can be solved for the fuel velocity. The

effective orifice area, ∗oA , is related to the fuel velocity through the mass flow rate

equation: ∗= offf AVm ρ& , (2.3)

The fuel velocity at the exit of the orifice is shown as Vf in equation 2.3 and will

be used in the remainder of the analysis to replace Ve. Fuel mass flow rate is denoted

here as fm& and is set for this injector design, see Table 2.1. Therefore, equation 2.3 can

be used to solve for the required effective orifice area. The reason why it is called the

effective orifice area is due to the fact that the velocity profile of the liquid exiting the

orifice is non-uniform. Typically the physical orifice area is multiplied by a constant

called the discharge coefficient, CD, to account for this effect. As a result of this

phenomenon the effective orifice area is related to the physical orifice area as follows:

2

4 ooDoDo dNCACAπ

==∗ , (2.4)

There are an infinite number of combinations of both orifice diameter and number

of orifices that could be used to satisfy a particular total effective area requirement. One

method of proceeding is to specify the number of orifices and then calculate the diameter.

In many cases the diameter that results may need to be modified slightly to match a

common drill bit diameter. If this is required it is advisable to go backwards through the

calculations to determine the pressure drop using the modified diameter. To be

conservative it would be better to adjust to a smaller diameter to make the actual pressure

drop slightly larger to avoid instability.

Page 40: Thesis

25

2.2.1.1 Discharge Coefficient

When a liquid enters an orifice from a large manifold it can separate from the

walls of the orifice creating a recirculation zone. In some cases the liquid does not

reattach before it is expelled from the orifice or if it does reattach a sizable boundary

layer exists. This effect is largely dependent on the orifice inlet geometry. The boundary

layer creates a non-uniform velocity profile in the orifice which remains when the liquid

is expelled from the orifice. The physical orifice area is multiplied by a constant term

called the discharge coefficient to account for non-uniform exit velocity.

The discharge coefficient for a full- flowing circular orifice is dependent on its

inlet type and length. Well-rounded inlets with a radius of one-half of the orifice

diameter or greater produce discharge coefficients approaching unity.44 For square-edged

inlets the discharge coefficient is primarily a function of the orifice length. Figure 2.2

shows the variation in discharge coefficient for a full- flowing orifice with a square-edged

inlet as a function of orifice length over diameter, lo/do.44

As this factor approaches a

value of 3.0 the discharge coefficient steadies out to a value of approximately 0.8. This

value is used a first approximation in the present analysis. Typically, an injector is

water-flowed to determine the actual discharge coefficient of its orifices.

Note that for a square-edged orifice the discharge coefficient does not reach unity.

Although the flow is attached to the wall for these high lo/do orifices a sizable boundary

layer still exists causing the velocity profile at the orifice exit to be largely non-uniform.

The flow through a well-rounded inlet follows the wall very smoothly and a very minimal

boundary layer is created resulting in a nearly uniform velocity profile at the exit.

Chamfered or gradually contracting inlets such as that shown in Figure 2.1 are also used

to reduce inlet losses.

Page 41: Thesis

26

Figure 2.2: Dependence of discharge coefficient on orifice length over diameter ratio for a full flowing square-edged inlet.44

2.2.1.2 Manifolding

In a multi-orifice transverse injector such as that shown in Figure 2.1 manifolding

the liquid properly is an important issue. It is crucial that the pressure losses in the

manifold be kept to a minimum to keep the injector feed pressure low. High feed

pressures require high tank pressures and generally make system operation more difficult.

Typically a ring manifold fed by a single orifice (or downcomer) is used. A low

downcomer exit velocity is required to minimize pressure loss as the fuel enters the

manifold. In addition, the downcomer should be aligned between two injector orifices to

supply uniform flow to the remaining orifices, as shown in Figure 2.1. The velocity in

the manifold should also be low to minimize pressure loss and to provide uniform flow.

A common rule of thumb is to make the cross-sectional area of the manifold ten times

that of the total orifice area. This will guarantee the manifold velocity to be sufficiently

low. Poor distribution of liquid will lead to unstable engine performance during startup

and shutdown and hot streaks along the combustor wall.

Page 42: Thesis

27

2.2.2 Decomposed Gas Flow Calculations

The decomposed gas flow properties must also be determined to calculate the

trajectory of the fuel jet. Hydrogen peroxide decomposition products consist of high

temperature oxygen gas and water vapor. In this analysis the decomposition products are

treated as an ideal gas mixture. Since the Mach number of the gas may be in the high

subsonic regime it is treated as a compressible fluid and compressibility effects are

included in the design equations. The most important part of this analysis is to determine

the velocity and Mach number of the decomposed gas. For now, it is assumed that the

flow area of the gas, Aox, is known. In fact, the flow area is set through flame

stabilization requirements. The oxidizer flow rate, oxm& , can be described using the mass

flow rate equation:

oxoxoxox AVm ρ=& , (2.5)

Unlike the liquid fuel the gas density is not a constant but is a function of the

static pressure and temperature of the decomposed gases. Using the ideal gas law the

following equation is used to determine the gas density, ?ox, for subsonic flow:

oxu

oxoxox TR

MWp=ρ , (2.6)

Here pox is the static pressure, Tox is the static temperature, and MWox is the

molecular weight of the decomposed gas. The universal gas constant is represented by Ru

and is equal to 1545 ft- lbf/lbm-mol-°R. An alternate form of equation 2.6, which

includes compressibility effects, can be found using the isentropic relations for stagnation,

or total, properties of ideal gases. Using the isentropic relation for total density, ?tox,

equation 2.6 becomes:

11

21

1

2

21

12

11

−−−−

+=

+=γγ γγ

ρρ oxtoxu

oxtoxoxtoxox M

TRMWp

M , (2.7)

Total properties are represented by a subscript ‘t’ while Mox is the gas Mach

number and ? is the specific heat ratio of the decomposed gases. Going back to equation

Page 43: Thesis

28

2.5 the velocity of the decomposed gas can be alternately expressed using the definition

of the Mach number and the speed of sound:

ox

oxuoxoxoxox MW

TRMaMV

γ== , (2.8)

The speed of sound in the decomposed gas is represented by aox. Substituting the

isentropic relation for total temperature the gas velocity is as follows:

21

2

21

1−

+== oxox

toxuoxoxoxox M

MWTR

MaMVγγ

, (2.9)

Substituting equations 2.7 and 2.9 into equation 2.5 and rearranging gives the

following result:

( )( )12

1

2

21

1−

+−

+==γ

γγ

γ oxoxox

toxu

oxtox

oxox MM

MWTR

Apm

M&

, (2.10)

Each of the parameters on the left-hand side of equation 2.10 is a known quantity.

Decomposed gas mass flow rate and flow area are set. The terms under the square-root

are properties of the decomposed gas and can be determined using a thermo-chemistry

code, these properties are dependent on H2O2 concentration.4 The total pressure of the

decomposed gas can be set equal to the chamber pressure of the engine. This assumes

that the gases do not suffer a total pressure loss as they expand around the

rearward-facing step. The chamber pressure can be calculated using the following

equation:

ct

totthCctox gA

mCpp

&∗

== *η, (2.11)

Here pc is the chamber pressure, ∗thC is the theoretical characteristic velocity

determined using a thermochemistry code, ?C* is the assumed combustion efficiency,

totm& is the total mass flow rate, and gc is the gravitational constant. An iterative method,

such as Newton’s method, can be used to solve equation 2.10 for the gas Mach number

since all the other parameters are known. Once the Mach number has been calculated it

can be used in equations 2.7 and 2.9 to solve for the gas density and velocity

respectively.

Page 44: Thesis

29

It is important to note that the total pressure of the oxidizer will change when fuel

is injected into the chamber. In other words, during bipropellant operation the total mass

flow rate will include the fuel flow rate and the characteristic velocity will change as well.

Using the baseline conditions and assuming a combustion efficiency of 98% the chamber

pressure with decomposed gas flow only is about 265 psia, this is also called

monopropellant operation. During bipropellant operation using baseline conditions and

an efficiency of 98% the chamber pressure rises to 533 psia. This demonstrates the

significant change in gas pressure depending on the operating mode of the engine.

2.2.3 Rearward-Facing Step Sizing

The rearward-facing step serves as a flame stabilization mechanism in the

combustion chamber. It creates a recirculation zone that feeds the incoming mixture of

fuel vapor and decomposed gases with hot products to sustain the combustion reaction.

In addition, a turbulent shear layer exists between the free-stream and the recirculating

flow creating large-scale turbulent eddies. A photo of this interaction is shown in Figure

2.3.45 It has been suggested that these hot eddies must ignite before they are cooled by

the free-stream flow or the flame will blow-off.42 Therefore, the turbulent mixing time

must be maximized to create a stable flame.

Figure 2.3: Reacting flow behind a rearward-facing step showing turbulent eddies.45

Page 45: Thesis

30

The turbulent mixing time is usually taken to be equal to the characteristic

breakdown time of the eddies.42 Knowledge of the characteristic length scale of the

turbulent eddies is required to determine the breakdown time. Experimental results have

shown that the length scale is be equal to the width of the flameholder.42 Using the

notation of Figure 2.1 the flameholder width is equal to twice the height of the

rearward-facing step, 2h. The shear layer characteristic time, t sl, can then be written as:42

gV

hsl

2=τ , (2.12)

Vg is the gas velocity at the edge of the step. The gas velocity is not equivalent to

the velocity of the decomposed gases, Vox, since a mixture of fuel vapor and decomposed

gas exists at the edge of the step. To determine an optimum value for the height of the

step some manipulations must be made to equation 2.12. Using the mass flow equation

the gas velocity can be written as:

oxg

totg A

mV

ρ&

= , (2.13)

Compressibility effects are eliminated by assuming subsonic flow in this analysis.

The oxidizer flow area can be rewritten in terms of the chamber diameter, Dc, and the

step height:

( )224

hDA cox −=π

, (2.14)

Substituting equations 2.13 and 2.14 into equation 2.12 results in the following:

( )222

hDhm c

tot

gsl −=

&

ρπτ , (2.15)

Assuming that the gas density and the total mass flow rate are not functions of the

step height the derivative of equation 2.15 with respect to the step height is:

( ) ( )[ ]hDhhDmdt

dcc

tot

gsl 2422

2 −−−=&

ρπτ, (2.16)

Setting the derivative equal to zero and solving for step height gives the following

relation for the optimum step height, hopt:

Page 46: Thesis

31

6c

opt

Dh = , (2.17)

The second derivative of equation 2.15 is taken to confirm that this step height

does indeed maximize the shear layer characteristic time:

( )[ ]hDhDmdh

dcc

tot

gsl 1642422

2

+−−−=&

ρπτ, (2.18)

Substituting equation 2.17 into equation 2.18 for all h:

tot

gc

sl

mD

dhd

&

ρπ

τ2

2

2

−= , (2.19)

The second derivative at hopt is negative meaning that the shear layer

characteristic time curve is concave downward or a maximum at that point. This

confirms that equation 2.17 does represent the value of the optimum height for a

rearward-facing step. This result has been derived mathematically by others and

confirmed through experimental results.41-43 The relation for optimum step height was

used to size the rearward step in this autoignition study.

A dimensionless geometric parameter, called blockage ratio, BG, is often used to

describe the size of the obstruction created by the rearward step in the flow path of the

gas:

c

oxc

AAA

BG−

= , (2.20)

Combustion chamber cross-sectional area is denoted as Ac. The optimum

geometric blockage ratio assuming the step height described in equation 2.17 is 0.56 for

this injector configuration. This means that the rearward-step takes up over 50% of the

total chamber flow area.

Another parameter of interest concerning the rearward-facing step is the length of

the recirculation zone. This has been studied both numerically and experimentally in

both reacting and non-reacting flows.45,46 Typically the flow reattachment length, la, is

specified in terms of number of step heights downstream of the face of the step. For

non-reacting flows the reattachment length is 6 to 8 step heights downstream of the step

face.45,46 In a reacting flow situation the reattachment length drops to 4 to 6 step heights

Page 47: Thesis

32

downstream.45,46 This averages at an attachment length of approximately one chamber

diameter downstream assuming an optimum step height.

2.2.4 Trajectory Model

There has been a good deal of past research aimed to predict the trajectory of a

liquid jet injected transversely into a crossflow of subsonic gas in non-reacting systems.

Lin et al34 summarized and compared many of the empirical correlations as well as the

measurement techniques used to obtain them. One must exercise caution when using

these empirical correlations in reacting flow situations. This is especially true following

the breakup of the liquid column into ligaments and droplets where vaporization begins

to play important role, i.e., high x/do.

Unfortunately, many experimental studies undertaken to develop jet trajectory

correlations are done far from the injection point, high x/do (> 50).34 A study done by Wu

et al35 evaluated jet trajectory at x/do < 10 using a range of liquids. The testing was

conducted over a wide range of liquid and gas flow conditions as well. Wu et al’s

correlations are most applicable to reacting flow situations encountered in rocket

combustors for these reasons.

The trajectory for a liquid jet in crossflow, from Wu et al, is stated as follows:

oDjo dx

QCd

y π= , (2.21)

Here x and y are the axial and transverse directions respectively, CDj is the drag

coefficient of the liquid jet, and Q is the liquid to gas momentum ratio. Wu et al also

developed a correlation for determining the drag coefficient for any liquid based on its

viscosity:35 364.0

984.0

=

w

f

Dw

Dj

C

C

µ

µ, (2.22)

CDw is drag coefficient for water, µw is the viscosity of water, and µf is the

viscosity of the liquid of interest. Wu et al determined the drag coefficient of water to be

Page 48: Thesis

33

1.51.35 At room temperature the viscosity of water is 1.87*10-5 lbf-s/ft2 while the

viscosity of JP-8 fuel is 2.56*10-5 lbf-s/ft2.47 Using these properties the drag coefficient

for a JP-8 fuel jet is 1.67. Substituting this result into equation 2.21 the trajectory of a

JP-8 jet is described by the following equation:

oo dx

Qdy

37.1= , (2.23)

It is important to note that the jet trajectory is primarily dependent on the liquid to

gas momentum ratio. The higher the momentum ratio the greater the jet will penetrate

into the gas stream at a given axial position. The momentum ratio for this analysis is

defined as:

2

2

oxox

ff

V

VQ

ρ

ρ= , (2.24)

Each of the parameters needed to determine momentum ratio can be calculated

using equations 2.2, 2.7, and 2.9.

2.2.4.1 Jet Breakup

A fuel jet exiting from the orifice of a transverse injector relies heavily on the gas

flow for atomization. Considerable work has been done in analyzing the primary

atomization or breakup processes of a nonturbulent liquid jet.35,37,38 In general there are

two parameters which are most important in dictating the jet breakup regime; the

momentum ratio and the Weber number. Weber number, Wegd, is a non-dimensional

variable relating aerodynamic forces of the free-stream gas to the surface tension forces

of the liquid and is defined here as:

f

ooxoxgd

dVWe

σρ 2

= , (2.25)

The surface tension, s f, of JP-8 fuel is approximately 1.57*10-3 lbf/ft. There are

four generally accepted primary breakup regimes for nonturbulent jets separated by

Weber number.35,37,38 According to Mazallon et al37 at Wegd < 5 aerodynamic forces are

Page 49: Thesis

34

small compared to surface tension forces in the liquid column. Instabilities within the

liquid column itself will dominate in this regime, which is referred to as the column

breakup regime. As Weber number increases the aerodynamic forces play a greater role.

In the bag breakup regime, 5 < Wegd < 60, the liquid column deforms into large baglike

structures with interconnecting liquid columns. Multimode breakup occurs at 60 < Wegd <

110 where the column is broken into baglike droplets and liquid ligaments are formed

due to shearing forces acting on the perimeter of the column. At Wegd > 110, the shear

breakup regime, aerodynamic forces are dominant and liquid ligaments are stripped from

the liquid column. Wu et al35 also investigated the different primary jet breakup regimes

noting that at large momentum ratios surface breakup occurs before breakup of the liquid

column. Wu et al also found that the breakup point in the trajectory of a nonturbulent jet

always occurs at an x/do = 8.0.35 This was also confirmed by Salaam et al.38 This result

can be used as a guideline during transverse injector design. It is preferred that the fuel

jet undergo breakup and begin to vaporize prior to passing the rearward-facing step. To

guarantee this the edge of the rearward step should be at least eight step heights

downstream of the fuel injection point.

2.2.4.2 Drop Size

Much work has also been done in determining empirical correlations for average

drop size in a crossflow injection scheme. Ingebo39 developed two mass median drop

diameter (MMD) correlations based on the column wave regime, i.e. capillary or

acceleration waves. Hautman and Rosfjord40 developed an empirical MMD correlation

using Malvern analysis techniques that was dependent on gas density, gas velocity, and

liquid surface tension. Hautman and Rosfjord compared the developed correlation to past

MMD empirical correlations including that of Ingebo. This comparison served mainly to

investigate the various exponents used on the density, velocity, and surface tension terms.

Kihm et al48 developed a correlation for Sauter mean drop diameter (SMD) based on flow

properties as well as axial and transverse position downstream of the injection point.

Page 50: Thesis

35

Due to the multiplicity of jet breakup regimes and limits in measurement

techniques there are no completely comprehensive drop size correlations. However,

according to Hautman and Rosfjord many of the past drop size correlations for liquid jets

in a crossflow use a similar dependency:

pox

nox

mf

VCSMD

ρ

σ∝ , (2.26)

Here C, m, n, and p are all constants. This equation shows that drop size is

directly proportional to surface tension and inversely proportional to gas density and

velocity. The denominator term is similar in form to gas momentum or dynamic pressure,

which is normally defined as ρoxVox2. According to Hautman and Rosfjord, the constant

exponent p is larger than n making the gas velocity the more dominant factor in

determining drop size. This result can be used to qualitatively compare the atomization

of a transverse jet operating under varying flow conditions.

2.2.5 Baseline Injector Design

The important fuel flow parameters and fuel geometry for the transverse injector

design used in this study are shown in Table 2.2. Twelve orifices were chosen for this

design to provide significant fuel distribution. Assuming an orifice discharge coefficient

of 0.8 and designing for a pressure drop of 100 psid an orifice diameter of 0.0355 inches

was required to provide the necessary effective area. A final orifice diameter of 0.036

inches was chosen to match a common drill bit diameter, which slightly decreased the

design pressure drop to 94 psid. Due to design limitations the cross-sectional area of the

injector manifold came out to be approximately seven times the total orifice area

resulting in a manifold velocity of about 15 ft/s. The manifold velocity is significantly

lower than the velocity of the fuel exiting the orifice, 130 ft/s, and should not produce any

significant pressure loss. In addition a 45° contraction was used to smoothly transition

the fuel from the manifold to the orifices. The orifice length came out to be slightly

greater than four times its diameter putting the discharge coefficient close to 0.8

according to Figure 2.2. Dimensioned injector drawings are contained in Appendix A.

Page 51: Thesis

36

Table 2.2: Fuel orifice and manifold geometry of transverse injector design including flow parameters at baseline operating conditions.

Fuel Mass Flow Rate, fm& (lbm/s) 0.45

Fuel Density (JP-8), ?f (lbm/ft3) 50.6

Number of Orifices, No 12

Orifice Diameter, do (in) 0.036

Assumed Discharge Coefficient, CD 0.80

Orifice Exit Velocity, Vf (ft/s) 131

Orifice Pressure Drop, ?pf (psid) 93.9

Orifice Length/Diameter Ratio, lo/do ~ 4.0

Manifold Area/Orifice Area Ratio, Am/Ao ~ 7.0

Table 2.3: Oxidizer port geometry and flow parameters in baseline monopropellant and bipropellant operational modes.

Chamber Diameter, Dc (in) 2.56

Oxidizer Port Diameter, Dox (in) 1.707

Rearward Step Height, h (in) 0.427 Orifice-to-Step Distance, ls (in) 1.75

Monopropellant Bipropellant Oxidizer Mass Flow Rate, oxm& (lbm/s) 2.23 Gas Mach Number, Mox 0.21 0.10

Gas Velocity, Vox (ft/s) 473 235

Gas Density, ?ox (lbm/ft3) 0.290 0.594

The oxidizer port geometry and decomposed gas flow parameters for

monopropellant, monoprop, and bipropellant, biprop, operation of this injector are shown

in Table 2.3. Using equation 2.17 and a chamber diameter of 2.56 inches, from Table

2.1, the optimum step thickness was determined to be 0.427 inches. This set the oxidizer

port diameter, Dox, at 1.707 inches. Breakup of the fuel jet should occur approximately

0.30 inches downstream of the injection point from the finding of Wu et al and Salaam et

al.35,38 The edge of the rearward step was located 1.75 inches downstream of the

injection point due to the presence of the axial injector. This is approximately 50 orifice

Page 52: Thesis

37

diameters downstream which should give the jet significant time to atomize and vaporize.

A schematic of the injector assembly including some important dimensions in shown in

Figure 2.4, dimensioned hardware drawings are contained in Appendix A. The solution

for the gas flow properties depends on the mode of engine operation and the properties of

the decomposed gas, which are shown in Table 1.2. As Table 2.3 shows the gas velocity

decreases by half, from 470 to 235 ft/s, and the density doubles upon switching from

monoprop to biprop mode.

Figure 2.4: Schematic of injector assembly indicating important geometry and dimensions. All dimensions are in inches.

Figure 2.5 compares the fuel jet trajectory produced by this transverse injector in

monoprop and biprop mode. The plots show the variation in trajectory with changes in

gas velocity. In monoprop mode the fuel- to-oxidizer momentum ratio is 13.4 while in

biprop mode it increases to 26.5. This suggests, as Figure 2.5 shows, that the fuel

penetrates deeper into the gas flow in bipropellant mode. The vertical line in Figure 2.5

indicates the axial position of jet breakup. The trajectories produced in both operational

modes intersect after the breakup point. The Weber numbers in monoprop and biprop

modes are approximately 3850 and 1950 respectively. According to the Mazallon et al37

Page 53: Thesis

38

this puts the jet breakup well into the shear regime. Since the gas velocity is higher in

monoprop mode it suggests that the average fuel drop size is smaller than during biprop

operation.

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Axial Direction, x (in.)

Tra

nsv

erse

Dir

ecti

on

, y (

in.)

Biprop TrajectoryMonoprop TrajectoryAxial Fracture PointChamber WallsChamber Centerline

Figure 2.5: Comparison between JP-8 jet trajectories produced by transverse injector design during monoprop and biprop operation. Fuel orifices are located at x = 0, y =0 and

x = 0, y = 1.707 inches centerline is at y = 0.854 inches.

Typically staged-bipropellant engines operate in monoprop mode prior to fuel

injection. From the point at which the fuel is injected until the combustion reaction

occurs in the chamber to increase the chamber pressure the fuel will follow the monoprop

trajectory curve. (There will be a slight chamber pressure rise due to fuel vaporization

but it does not alter the trajectory significantly). As a result the monoprop trajectory is

crucial in dictating the initial atomization characteristics of the flow and could potentially

affect autoignition as well. It is important that the trajectory be monitored very carefully

such that fuel penetration is not too small or too large, especially at off-design conditions.

If the momentum ratio is too small the liquid will not be dispersed well within the

chamber and will tend to hug tightly to the chamber wall causing a fuel-rich periphery

Page 54: Thesis

39

and an oxidizer-rich core. If the momentum ratio is too large the streams will collide and

coalesce at the centerline of the chamber, resulting in a fuel-rich core and possibly poor

vaporization. This can be seen in Figure 2.5 downstream of the breakup point. Both

cases result in poor dispersion of the liquid within the chamber and a significant decrease

in performance. Data obtained by Helms et al30 using two different transverse injector

designs flowing H2O2 showed performance losses when trajectory analysis indicated

liquid stream collision. Analysis of the predicted trajectories shows that opposing fuel

jets do not appear to collide prior to breakup while fuel penetration seems to be sufficient

enough to produce reasonable fuel dispersion.

Page 55: Thesis

40

CHAPTER 3: AUTOIGNITION

One of the primary advantages of a H2O2/kerosene staged-bipropellant rocket

engine is that the hot decomposed hydrogen peroxide gases can be used as an ignition

source. This eliminates the need for a separate ignition system, such as a torch igniter,

and greatly reduces the complexity of the engine. One drawback to this ignition

approach is that the conditions required for autoignition are not well understood. Past

research has shown that autoignition of kerosene in decomposed hydrogen peroxide is

dependent on temperature, pressure, gas velocity, and injector design.8-11,17,22 This

chapter will attempt to outline the basics of autoignition in hopes of creating a better

understanding of the process occurring in a staged engine. In addition past work on

autoignition of kerosene in air will be discussed as well as autoignition studies of

kerosene in decomposed hydrogen peroxide. Also included is a model for determining

flame stability and autoignition in dump combustor with a rearward-facing step. The

goals of this autoignition study will be outlined as well.

3.1 Basic Chemistry of Autoignition

Autoignition, also known as spontaneous or thermal ignition, is defined as the

initiation of a self-sustained exothermic reaction without the aid of an external energy

source. In other words, the reactants themselves have enough energy to start and

continue a reaction. Ignition is a term commonly used in reference to combustion or

oxidization of a mixture of gases, which in this study is kerosene fuel vapor and

decomposed hydrogen peroxide. In many cases an external energy source is required for

ignition if the energy of the reactants is insufficient to begin the reaction. Typically an

Page 56: Thesis

41

electric spark, laser, or torch is used to provide additional energy for ignition.

Autoignition, by definition, is an event not a process. However, chemical kinetics, in

general, dominates the process which occurs prior to autoignition. This process is

complex and not well understood, but a basic understanding of it may help in determining

which factors are most important to achieving autoignition.

3.1.1 Composition of Kerosene Fuel

Before going into the kinetics of the processes leading to autoignition it helps to

understand the properties and composition of kerosene fuel. Kerosene is a distillate

fraction of petroleum that boils between 300 and 570°F.47,49,50 It contains thousands of

different hydrocarbon molecules and can have many different formulations as a result.

Common kerosene fuel formulations used in the aviation and rocket industries are Jet-A,

JP-4, JP-5, JP-8, and RP-1. Many of these kerosene derivatives were developed for

specific applications that may have required lower freezing point, higher flash point, or

lower concentrations of specific hydrocarbons than available kerosene-based fuels at the

time. These properties are achieved by varying the composition or introducing additives

to the fuel. JP-8 is currently the most widely used fuel for aircraft in the U.S. Army and

Air Force.49,50 It is basically Jet-A, which is the most common commercial aircraft fuel,

with three military specified additives that inhibit corrosion and improve lubricity,

dissipate static, and inhibit fuel system icing.49,50 Since kerosene-based JP-8 is so widely

used in military aviation it was chosen for this autoignition study.

JP-8 is composed of a number of paraffins, naphthenes, aromatics, and olefins of 50,

30, 18, and 2 percent by volume respectively on average.47,49,50 These different families

of hydrocarbons are classified according to their molecular formula, molecular structure,

and type of carbon-carbon bonding. A summary of the properties of each family of

hydrocarbons is shown in Table 3.1.51 The molecular formula of JP-8 is C11H21 which is

an average based on the composition of hydrocarbons found in the fuel. Some important

physical properties of JP-8 are shown in

Table 3.2.47,49,50

Page 57: Thesis

42

Table 3.1: Basic hydrocarbon families as found in kerosene-based fuels.51

Family Name Other Names Molecular

Formula Carbon-Carbon

Bonding

Primary Molecular Structure

Alkanes Paraffins CnH2n+2 Single bonds only Straight or

branched open chains

Alkenes Olefins CnH2n One double bond, all others single

Straight or branched open

chains

Alkynes Acetylenes CnH2n-2 One triple bond, all others single

Straight or branched open

chains

Cyclanes Cycloalkanes, Cycloparaffins,

Naphthenes

C2H2n or (CH2)n

Single bonds only Closed rings

Aromatics Benzene Family

CnH2n-6 Resonance hybrid

bonds Closed rings

Table 3.2: Physical properties of JP-8 fuel. All properties stated at ambient temperature and pressure unless indicated otherwise.47,50

Average Molecular Formula C11H21 Average Molecular Weight 153.3

Boiling Range (°F) 330-510 Freezing Point (°F) -60

Critical Temperature (°F) 770 Critical Pressure (psia) 340

Density (lbm/ft3) 50.6 Kinematic Viscosity (lbf-s/ft2) 2.56e-5

Surface Tension (lbf/ft) 1.57e-3 Vapor Pressure (psia @ 120°F) 0.190

Page 58: Thesis

43

3.1.2 Combustion of Hydrocarbons

Kerosene-based fuels such as JP-8 contain a number of different aromatics,

naphthenes, paraffins, and olefins. The molecular formula of JP-8 suggests that it is

made up of hydrocarbon molecules containing eleven carbons on average. These

molecules are very large with numerous atomic bonds of different types and strengths.

Breaking down such large molecules cannot be done in a single reaction but must occur

in a series of smaller reactions until only simple molecules remain. For example, Turns51

describes the combustion of paraffins, with greater than two carbon atoms, as a sequence

of three processes.

In the first step of the combustion process the paraffin fuel molecule is

bombarded by hydrogen, H, and oxygen, O, atoms and splits into mainly olefins and

diatomic hydrogen, H2. The diatomic hydrogen is oxidized to form water assuming there

is oxygen available. The second step of the reaction involves the oxidation of the olefins

to carbon monoxide, CO, and diatomic hydrogen. Olefins are unsaturated molecules

meaning hydrogen atoms do not occupy all of the carbon’s available bonds. Also, in this

step, all the diatomic hydrogen is converted to water. Finally, in the third step the carbon

monoxide is consumed to form carbon dioxide, CO2. The consumption of carbon

monoxide accounts for most of the heat release produced by the overall combustion

process. The high heat release is usually accompanied by a luminous flame. In a

reacting flow ignition is usually defined to occur at the point in which a flame is visible,

hot- flame ignition. Since this is the point of highest heat release the energy generated

sustains the reaction.

Combustion or oxidation of a JP-8 is a complex multi-step process as described

above, however, to simplify things somewhat the overall reaction is typically written

instead. The overall combustion reaction of JP-8 with 1 gram of hydrogen peroxide at

concentration 100*X % assuming complete combustion can be expressed as follows:

( ) ( ) ( ) ( ) ( )gglOH

lOH

l OcHbCOOHMW

XOH

MWX

HaC 2222

2222

21111

+→−

++ , (3.1)

Page 59: Thesis

44

Here MWH2O2 is the molecular weight of hydrogen peroxide, 34.016 g/mol, and

MWH2O is the molecular weight of water, 18.016 g/mol. Note there is no oxygen, O2(g), or

water vapor, H2O(g), term on the left side of equation 3.1 to represent decomposed

hydrogen peroxide gas. The chemical reaction describing the decomposition of 100*X%

H2O2 has been combined with the complete combustion reaction of JP-8 with the

decomposed gas to form an equivalent overall reaction. The moles of JP-8, carbon

dioxide, CO2(g), and water vapor are represented by a, b, and c respectively in equation

3.1 and are determined from the following equations:

22652

OHMWX

a = , (3.2)

226522

OHMWX

b = , (3.3)

OHOH MWX

MWX

c222

16586 −

+= , (3.4)

The stoichiometric mixture ratio, O/Fs, for JP-8 combusting with hydrogen

peroxide as a function of concentration can be found from:

8

22

265

=JP

OHs MW

MWX

FO , (3.5)

MWJP-8 is the molecular weight of JP-8 fuel, 153.3 g/mol as shown in

Table 3.2. For 90% H2O2 the stoichiometric mixture ratio is approximately equal

to 8.01 while for 98% H2O2 it is 7.36. The relative amounts of fuel and oxidizer present

in a reaction is most often expressed in terms of equivalence ratio, φ, and is defined as

follows:

FOFO s=φ , (3.6)

The actual oxidizer to fuel mixture ratio is represented by O/F. When operated at

stoichiometric mixture ratio the equivalence ratio is equal to one. At fuel rich conditions,

or O/F < O/Fs, the equivalence ratio is greater than one and at oxidizer rich conditions it

is less than one. Representation of mixture ratio in terms of equivalence ratio allows a

Page 60: Thesis

45

comparison between staged engine autoignition test data and much of the past kerosene

autoignition work done in air.

3.1.3 Kinetics of Autoignition

The previous section outlined the basic chemical processes which must occur

prior to ignition of a hydrocarbon fuel such as kerosene. In order for these reactions to

occur the energy of the gaseous fuel and oxidizer mixture must increase by a specific

amount, called the activation energy, before the reactants can be converted to products.

The collision theory of chemical reactions is used to model a reacting mixture by

supposing the system is a collection of molecules of finite size and velocity moving

randomly in space.51-53 As a molecule in the mixture moves in space it will eventually

collide with another molecule. A chemical reaction may occur at the point of collision

depending on the energy and orientation of each molecule. If the energy of the collision

is too low no reaction will occur at all, it will not activate a reaction. If the temperature

and pressure of the mixture are increased the energy of the molecules will increase as

well. Now, at the point of collision the molecules have more energy allowing stronger

molecular bonds to be broken and potentially releasing more energy to the system.

Eventually the energy of the moving molecules and their molecular composition may

become sufficient to activate the final step in the combustion reaction and autoignite the

mixture.

Collision theory can be also used to explain how the rate of a combustion reaction

is affected by the conditions of a system. In the collision theory model molecules gain

more energy with increasing temperature and pressure. As molecular energy increases so

does the velocity of the molecule and the number of collisions occurring in the mixture.

If collisions occur more frequently and at higher energy then reactions will also occur at a

faster rate. The Arrhenius relation states that the rate, kR, at which a chemical reaction

occurs is dependent on the activation energy of the reaction, Ea, and the temperature, T,

of the system:51-53

Page 61: Thesis

46

=

TRE

Zku

aR exp , (3.7)

Here Ru is the universal gas constant, and Z is the pre-exponential factor which is

dependent on the number of collisions between molecules calculated from collision

theory. Equation 3.7 agrees with collision theory in that the higher the temperature of

the mixture and the lower the activation energy of the reaction the faster the reaction rate.

Due to the temperature dependence of the reaction rate it may change significantly over

the course of a reaction as heat is released to the system.

As collision theory suggests and the previous section described a number of

elementary and chain reactions can occur in a system. Hydrocarbons, such as kerosene,

typically contain a significant number of elementary reactions in addition to the three

major steps described previously. The sum of these reactions describes the overall

reaction or reaction mechanism of the system. For such complex cases a global

activation energy is determined experimentally to approximate the kinetics of the system.

It is important to note that the energy of the gaseous mixture determines what reactions

will occur, the rate of these reactions, and the path of subsequent reactions. Thus, the

final result of the reaction depends on the initial conditions of the reactants and may not

necessarily lead to complete combustion or autoignition. This is demonstrated by the fact

that hydrocarbons reacting at low temperatures, < 800°F at ambient pressure, emit a

bluish flame.54,55 Commonly called cool- flame ignition this reaction is characterized by a

relatively low heat release, as compared to hot- flame ignition, and as such the reaction is

not self-sustainable.

The period of time from which a gaseous fuel and oxidizer mixture is formed to the

point of autoignition is termed the ignition delay. Ignition delay is governed by the

chemical kinetics of the reactions leading to carbon monoxide consumption assuming a

well mixed combination of fuel and oxidizer. From collision theory and the Arrhenius

relation one would expect the ignition delay to decrease with increasing temperature and

pressure as the reaction rate increases. One would also expect the concentration of the

reactants to affect the delay as well. Understanding ignition delay in practical

combustion systems is extremely important. In the case of the staged combustor used in

Page 62: Thesis

47

this study the mixture of decomposed peroxide and kerosene must be supplied at the

appropriate conditions to make the ignition delay as short a possible. Since the gases are

flowing at high velocity in the engine if the delay is too long the reactants will not

autoignite in the chamber. There is an added effect in ignition delay for a staged

combustor in that the fuel does not atomize, vaporize, and mix instantaneously. This

process takes some time and this time varies based on operating conditions. Chemical

reactions begin as soon as fuel vapor is present in the mixture and as such the rate of

vaporization may play an important role in the overall ignition delay. The rearward step

in a dump combustor is used to increase the time that the mixture spends in the chamber.

3.2 Autoignition Studies of Kerosene Fuel in Air

Since the 1970’s there has been significant research conducted on the autoignition

behavior of kerosene fuel in air. Many of these studies were performed initially to

simulate conditions in lean, premixed gas turbine engine combustors.55-58 In these

combustors the fuel and air are mixed prior to entering the main combustion chamber and

therefore autoignition must be avoided to prevent engine damage. More recently

research has been focused on autoignition in supersonic flows as encountered in scramjet

combustors.59 Ignition delay must be as short as possible in such engines to minimize

length and weight. Experimental research was aimed, for both engine cases, at

developing correlations and models for predicting ignition delay based on operating

conditions.

There are a number of methods used in determining the ignition delay of a

gaseous fuel and oxidizer combination. These include the constant volume bomb, rapid

compression, reciprocating engine, shock-tube, and continuous flow methods. In gas

turbine combustor research the continuous flow method was most often used because it

most accurately represented the actual flow field in a combustor. There are many

variations to this technique, but the basic concept is the same for all. The setup usually

consists of a tube or duct through which heated air flows at high subsonic velocity. One

of the benefits of such an arrangement is that the temperature, pressure, and/or velocity of

Page 63: Thesis

48

the air can be controlled independently. The fuel is injected into the air stream where it

mixes and reacts with the air. The method of injection is extremely important, especially

for liquid fuels, in minimizing effects of vaporization on ignition delay and for obtaining

a uniform mixture of fuel and air. Ignition is verified through the direct visual

observation of a flame inside the duct, for windowed setups, or exiting the chamber.

Alternate methods include the use of temperature or pressure measurements or

photodectors to indirectly verify the ignition event. Ignition delay is calculated using the

air velocity and the distance from the injection point to the position of the visible flame.

The correlations used to calculate ignition delay vary as much as the test

apparatus used to obtain them, but they too have a common form. Ignition delay

correlations are derived from chemical kinetics and make use of the Arrhenius relation.

The reaction is treated as a whole using the global activation energy determined from

experimental data. Correlations for ignition delay, t i, of hydrocarbons usually take the

following form:57

[ ] [ ] nmj

u

ai pOFuel

TRE

A 2exp

=τ , (3.8)

Here [Fuel] and [O2] are the concentrations of gaseous fuel and oxygen

respectively. The static pressure and temperature of the incoming air is denoted as p and

T. The parameter A is a constant and Ea is the global activation energy as in equation 3.7.

Exponents j, m, and n are constants determined through experimental data. Values of the

constants in equation 3.8 vary from experiment to experiment for a particular fuel, but

the trends are generally similar.57

There have been several autoignition studies performed to date with

kerosene-based fuel and air in a continuous flow apparatus. One such study was

performed by Spadaccini and TeVelde55 with Jet-A fuel for equivalence ratios ranging

from 0.3 to 0.7, temperatures from 800 to 1350°F, pressures from 150 to 450 psia, and air

velocities from 65 to 330 ft/s. In this experiment the pressure and velocity of the air were

set as well as the length of the duct. The temperature of the air was gradually increased

until a visible flame was seen at the exit of the duct. This was considered the point of

autoignition. Ignition delay was assumed equal to the residence time of the gases in the

Page 64: Thesis

49

duct, which was calculated by dividing the velocity of the air by the duct length. The

study resulted in the following general correlation for ignition delay:

=

TRE

pA

u

ani expτ , (3.9)

For Jet-A fuel the coefficient A is equal to 1.68e-8, n is equal to 2.0, Ea is equal to

37.78 kcal/mol, and p is in atmospheres. Equation 3.9 indicates that equivalence ratio

does not affect ignition delay, at least for φ < 1.0. In addition, ignition delay is shown to

be inversely proportional to the square of the air static pressure. Spadaccini and TeVelde

point out that most ignition delay correlations are fitted with a pressure exponent of unity,

n = 1.0, but their data suggests that an exponent of 2.0 is a better fit especially at high

temperatures. The effect of fuel concentration distribution was also investigated in this

study and was found to influence the ignition delay for short residence times. However,

as the residence time was increased the effect became negligible. Spadaccini and

TeVelde also investigated the effects of the temperature, up to 260°F, of the Jet-A fuel

prior to injection and found it had no influence on ignition delay.

A second study was performed by Freeman and Lefebvre56,57 with Jet-A at

atmospheric pressure. Tests were conducted at equivalence ratios from 0.2 to 0.8,

velocities from 30 to 130 ft/s, and temperatures up to 1450°F. Test were conducted

similarly to Spadaccini and TeVelde in that the flow conditions of pressure, velocity, and

equivalence ratio were prescribed and the air temperature was gradually increased until a

stable flame was observed exiting the duct. Freeman and Lefebvre observed that as the

air temperature increased autoignition initially arose as flashes of flame at a small

distance downstream of the duct exit. Gradually as the temperature increased the flame

front moved upstream to the duct exit at which point the air temperature was recorded

and the test was terminated.

The correlation developed by Freeman and Lefebvre was modeled after Equation

3.8. Since the pressure was kept constant during the investigation the value of n was set

equal to zero. A value of 40.9 kcal/mol was determined for Ea which compares very well

with the value of 37.78 kcal/mol determined by Spadaccini and TeVelde. Freeman and

Lefebvre also found that the equivalence ratio has a negligible effect on ignition delay for

Page 65: Thesis

50

φ < 1.0 also in agreement with Spadaccini and TeVelde. Increasing the oxygen

concentration of the incoming air was found to decrease the ignition delay and Freeman

and Lefebvre set the exponent m equal to -0.65 based on experimental data. The fuel

temperature was also found to have negligible effect on ignition delay consistent with

Spadaccini and TeVelde.

Both Spadaccini and TeVelde and Freeman and Lefebvre compared the

autoignition characteristics of kerosene to other fuels, including other hydrocarbons. In

general, kerosene was found to exhibit a shorter ignition delay in air than other typical

gas turbine fuels55 as well as gaseous fuels such as propane.56,57 In addition, the

experimentally determined activation energy of kerosene fuel has also been found to be

lower than other fuels.55 These results are attributed to the large concentration of

paraffins in the fuel. Paraffins, which are straight-chain hydrocarbons (see Table 3.1),

tend to autoignite more quickly than other types of hydrocarbons.55

The last study on the autoignition of kerosene fuel in air that will be discussed

was done by Mestre and Ducourneau.58 The experimental setup for this study differed

slightly from the previous two in that the had a cocentric design whereby a secondary

flow of air heated the main test section. Tests were conducted over a wide range of

equivalence ratios from approximately 0.8 to 8.0, pressures from 80 to 160 psia, an

average air ve locity of 215 ft/s, and temperatures up to 1500°F. As in the previous

studies the air temperature was gradually raised until autoignition was verified by a visual

detection of a flame exiting the duct. A unique correlation was developed to determine

the autoignition temperature based on pressure and equivalence ratio and is stated as

follows:

( )( )φ

φφφ

4.022.09.845000

16.0432.03

2

++−

++=pp

T , (3.10)

Here T is in degrees Celcius and p is in bars. Experimental results of Mestre and

Ducourneau showed that the autoignition temperature plotted as a function of

equivalence ratio for constant pressure had a minimum that varied with pressure. This

minimum temperature moves from an equivalence ratio of almost 3.0 at 5.4 bars to

approximately 1.0 at 11 bars. Data also showed that the required autoignition

Page 66: Thesis

51

temperature decreased with increasing pressure but the decrease became smaller as

pressure neared approximately 10 bars. The study also found that the required

autoignition temperature decreased with increasing residence time. Use of the secondary

air flow to heat the primary tube did not have a significant effect on the autoignition

temperature.

3.3 Dump Combustor Autoignition

As discussed in Chapter 2 in order for a stable flame to exist in a dump

combustor the rearward-facing step must provide sufficient shear layer residence time or

the flame will blow off. A study by Plee and Mellor42 outlines the development of a

model to predict blow off for various flame holder geometries under oxidizer rich

conditions. The model basically states that if the shear layer residence time is greater

than the ignition delay then a stable flame will exist in the shear layer. If the residence

time is to small then the shear layer flame will be unstable and blow off will occur. For

an oxidizer rich, premixed flow the relation can be stated as follows:

bm isl += ττ , (3.11)

The slope and intercept of the equation for the blow off limit are denoted by m

and b respectively. These values are determined through experimental data. Shear layer

residence time is calculated using equation 2.12 for a rearward-facing step configuration.

Plee and Mellor suggest the use of the following relation for ignition delay for oxidizer

rich equivalence ratios:

φτ

1exp

=

TRE

TT

Ku

a

ini , (3.12)

Here T is the adiabatic flame temperature at equivalence ratio φ, Tin is the inlet

temperature of the premixed gases, and K is a constant. The T/Tin term is used to

decouple fluid mechanics from thermochemistry. Data from Mestre and Ducourneau

showed that autoignition temperature increases with increasing equivalence ratio for fuel

rich mixture ratios. This suggests that the ignition delay increases with equivalence ratio

Page 67: Thesis

52

for fuel rich mixtures. Equation 3.12 would show the opposite trend if used for fuel rich

conditions and for that reason can only be used for oxidizer rich equivalence ratios. Note

that this relation for ignition delay does not include a pressure term.

In situations where vaporization may still play an important role in the

recirculation region behind a bluff body a modification to equation 3.11 is required:

bam visl ++= τττ , (3.13)

A constant, a, is multiplied by the vaporization characteristic time, t v, to give a

best fit to the experimental data. The vaporization characteristic time or fuel droplet

lifetime is determined from the D2 law for droplet vaporization:

λτ

2o

v

d= , (3.14)

The vaporization coefficient, ?, is determined through heat and mass transfer

considerations, see Turns,51 and the droplet diameter, do, is determined through

experimental data or correlations. A blow off model based on equation 3.13 has been

developed by Plee and Mellor based on experimental data for Jet-A reacting in air for

several different types of flameholder configurations. However, a model of flame

stability for Jet-A reacting with air behind a rearward-facing step was not developed in

the study by Plee and Mellor. This model is fairly simplistic, especially when

vaporization effects are not important, in that one can compare the calculated values of

shear layer time and ignition delay to determine whether a stable flame or autoignition

will occur.

3.4 Autoignition Studies in Staged Combustors

Research into the autoignition characteristics of kerosene fuel in decomposed

hydrogen peroxide has not been as substantial as it has been in air. Much of the data

concerning autoignition in staged engines are the result of problems encountered during

engine design. Typically the baseline engine design is tested and does not achieve

autoignition.17,22 Design changes are made, usually in residence time, and then the

Page 68: Thesis

53

engine is retested to verify autoignition. The studies do not go into much more depth

than stating the operating conditions and geometry of each design. There were, however,

a series of papers written in the early 1950’s on autoignition of kerosene fuel in

decomposed hydrogen peroxide by Walder,8,9 Walder and Purchase,10 and Walder and

Broughton.25 A summary of this research is found in a paper by Harlow.11

The most extensive of these autoignition studies by Walder is found in his earliest

paper.8 In this study 80% H2O2 was decomposed in a catalyst stone bed giving a

decomposition temperature of approximately 950°F, see Table 1.2. Tests to determine

autoignition temperature were conducted similarly to those done with kerosene in air.

Initially water is mixed with the liquid hydrogen peroxide to drop the decomposition

temperature below 750°F. The water flow was gradually reduced to increase the

decomposed gas temperature until autoignition was verified by visual detection of a

flame exiting the chamber throat. Aviation kerosene fuel was fed to the chamber through

twelve swirl injectors and the residence time of the gas in the chamber was modified by

varying the chamber length and throat diameter. The contraction ratio was varied from

approximately 6.5 to 20.0. All tests were conducted at a mixture ratio of 9.1, which is

approximately stoichiometric for 80% H2O2 according to equation 3.5.

Walder developed a correlation for the autoignition temperature of kerosene in

decomposed hydrogen peroxide similar in form to equation 3.9. The main difference

was that Walder supposed that the ignition delay was proportional to the characteristic

length, L*, of the engine. The basic form of the correlation is as follows:

=∗

TC

pC

Ln

12 exp , (3.15)

Here p is the chamber pressure, T is the decomposition temperature, and n, C1,

and C2 are constants. The constant C1 is basically equivalent to the constant Ea/Ru term

found in equation 3.9. Walder manipulates equation 3.11 by taking the logarithm of

both sides of equation 3.11 and arrives at the following:

( ) DTC

pL n +=∗log , (3.16)

Page 69: Thesis

54

The slope of a line describing autoignition pressure versus characteristic length at

constant temperature gives the value of the exponent n. Walder found n to be equal to

1.15 based on test data. This value is close to unity, which Spadaccini and TeVelde

suggested worked well for autoignition of kerosene in air at low temperatures. The

constants C and D are the slope and intercept of a line plotting autoignition temperature

as a function of L* at constant pressure. Walder found these constants to be equal to 3720

and -2.62 respectively based on test data. From this relation one can see that as L* and/or

chamber pressure are increased the temperature required for autoignition decreases. This

agrees with the results from Mestre and Ducourneau. Walder also investigated the effect

of equivalence ratio on the autoignition temperature at constant pressure of 100 psia and

L* of 50.4 inches. It was found that the required autoignition temperature rose to a value

of about 830°F at an equivalence ratio of 1.25 from 805°F at the stoichiometric value. At

an equivalence ratio of approximately 0.8 the required autoignition temperature fell to

770°F. The trend in these results also agrees with that of Mestre and Ducourneau for fuel

rich equivalence ratios.

A subsequent study by Walder further investigated the effect of L* on

autoignition.9 In this study the characteristic length was artificially increased by placing

a restrictor plate on the throat of the combustion chamber using the same setup as in the

previous study.8 The pressure would gradually increase as the fuel and decomposed gas

entered the chamber until autoignition occurred and blew out the restrictor plate. Another

study, by Walder and Purchase,10 investigated the affect of injector design on autoignition

temperature. A number of complex injector designs were tested which made use of swirl

jet, straight orifice, and combined swirl and straight orifice injection techniques. All tests

were conducted at a stoichiometric mixture ratio of 9.1 for 80% H2O2 and L* ranging

from 68 to 100 inches using the same test setup and procedure as in the previous

investigations.8 Average autoignition temperature was found to vary with injector design,

the injector requiring the lowest temperature is shown in Figure 1.5. The results suggest

that autoignition is dependent on the atomization and vaporization processes which occur

following fuel injection. A fourth study, performed by Walder and Broughton,25

investigated autoignition in larger, hydrogen peroxide cooled chambers. A silver screen

Page 70: Thesis

55

catalyst bed was used during this test series to decompose 80% H2O2. The engine ran at

approximately stoichiometric mixture ratio and a biprop chamber pressure of 400 psia. A

characteristic length of 30 inches was used, corresponding to a contraction ratio of

approximately 6.5, to provide a thrust of approximately 92% of the theoretical value.

A study by Wu et al22 on the development of staged-bipropellant engines using

hydrogen peroxide and JP-8 noted some problems obtaining autoignition. The test

engine was very similar in design to that presented in Chapter 2 of this thesis. A silver

screen catalyst bed was used to decompose 85% H2O2 at 95% efficiency and the engine

used a transverse type injector design with a rearward step. Autoignition was not

achieved at a monoprop chamber pressure of 340 psia and a contraction ratio of 5.4. The

mixture ratio used during this test is unclear but other data from the study suggests that it

was somewhere between 6.0 and 11.0, corresponding to equivalence ratios between 0.8

and 1.4. The contraction ratio was increased to 7.11 to establish reliable autoignition.

In a study by Coxhill et al17 autoignition problems were also experienced during

staged engine development. This engine design used kerosene fuel jet injectors angled at

55° to the decomposed gas flow and did not incorporate flame holding geometry. A

silver screen catalyst bed was used to decompose 90% H2O2 and the engine operated at

mixture ratios from about 6.0 to 10.0, equivalence ratios from 0.8 to 1.3. Autoignition

did not occur at a contraction ratio of approximately 13.0 at chamber pressures less than

160 psia even for fully decomposed hydrogen peroxide. It is unclear as to what the

mixture ratio was during these autoignition tests. At a contraction ratio of about 20.0

autoignition was achieved for all tested conditions and was subsequently used in the final

engine design.

It is interesting to note the high contraction ratios used during autoignition testing

of kerosene in hydrogen peroxide. From the data presented above the smallest

contraction ratio used was at a value of 6.5. Flight -rated engines such as the Gamma Mk

201 and Mk 301 had contraction ratios of approximately 7.529 and the Orbital Sciences

USFE engine had a contraction ratio of about 7.1.23 It is generally desired that the

contraction ratio be as small as possible to minimize the size of the combustion chamber.

Page 71: Thesis

56

This reduces weight as well as cooling requirements. Typical values of contraction ratio

for many other engines and propellant combinations are on the order of 2.0 or 3.0.

3.5 Goals of this Autoignition Study

The primary goal of this autoignition study is to develop a model for autoignition

to aid in the design of staged H2O2/JP-8 engines. Past research on autoignition of

kerosene fuel in air and decomposed hydrogen peroxide has shown good agreement on

the effect of temperature on ignition. The exponential effect of temperature on ignition

delay as derived from chemical kinetics and the Arrhenius relation, equation 3.7, is

agreed upon by most researchers. Most correlations suggest that ignition delay is

dependent on the inverse of pressure while some suggest that this dependence changes at

higher temperatures. It is also generally agreed that mixture ratio does not play a

significant role in autoignition at oxidizer rich equivalence ratios, φ < 1.0. However,

there is little data on the effect of fuel rich equivalence ratios on autoignition. It is also

unclear whether equivalence ratio effects are the result of changes in the chemical

kinetics of the reaction or of changes in the atomization properties of the injector, which

affects vaporization time. In addition, many studies mention that autoignition is

dependent on the residence time of the gaseous mixture, but there are few correlations

that actually take this effect into account. This effect is especially important when bluff

body flame stabilization is used in a combustor. In relation to rocket combustors

residence time, especially contraction ratio, seems to be extremely important to

autoignition. As a result, a design model for autoignition in a H2O2/JP-8 engine should

include the effects of fuel rich equivalence ratios and contraction ratio.

It is proposed that the blow off model derived by Plee and Mellor42 be used as a

model for autoignition for the engine used in this study. It will be assumed that

vaporization is complete prior to entering the recirculation zone behind the

rearward-facing step to eliminate the need for the droplet lifetime term. This may not be

a good assumption but it greatly simplifies the analysis. To use this model the ignition

Page 72: Thesis

57

delay equation, equation 3.12, must be modified to deal with fuel rich equivalence ratios.

The proposed modification is as follows:

j

u

a

ini TR

ETT

K φτ

= exp , (3.17)

The exponent, j, on the equivalence ratio term will need to be determined through

experimental data. The activation energy can be taken as an average of the results of

Spadaccini and TeVelde55 and Freeman and Lefebvre.56,57 Temperature effects on

autoignition can be determined by varying the hydrogen peroxide concentration and thus

the decomposition temperature. Shear layer residence time will be modified by changing

the chamber contraction ratio and thus the chamber pressure and gas velocity. The

rearward-facing step height will remain constant through testing.

A secondary goal of this study is to compare the autoignition characteristics of

JP-8 in decomposed hydrogen peroxide with that of an experimental fuel called DMAZ.

DMAZ is an acronym for Dimethylamino-2-ethyl azide or Dimethyl-2-Azidoethylamine.

Due to its experimental status there is a limited amount of data regarding the physical

properties of DMAZ. Some of the physical properties of the fuel can be found in its

Material Safety Data Sheet, MSDS, which is contained in Appendix B. The known

properties of DMAZ are compared to JP-8 in Table 3.3. DMAZ will be run under

similar test conditions as JP-8 in terms of equivalence ratio, hydrogen peroxide

concentration, and contraction ratio to make the comparison.

Table 3.3: Comparison between some physical properties of DMAZ and JP-8.

Fuel DMAZ JP-8 Molecular Formula (CH3)2NCH2CH2N3 C11H21 Molecular Weight 114.2 153.3

Boiling Point/Range (°F) 275 330-510 Freezing Point (°F) -92 -60 Density (lbm/ft3) 58.2 50.6

Kinematic Viscosity (lbf-s/ft2) < 2.09e-4 2.56e-5 Vapor Pressure (psia) 0.21-0.97 (@ 68°F) 0.190 (@ 120°F)

Flash Point (°F) 86 127 O/Fs with 90% H2O2 4.32 8.02

Page 73: Thesis

58

CHAPTER 4: EXPERIMENTAL SETUP

Autoignition testing was conducted in Test Cell A of Purdue University’s

Advanced Propellants and Combustion Lab, APCL, located at Maurice J. Zucrow

Laboratories. Over the past two years this facility has seen extensive use in testing of

hydrogen peroxide in monopropellant and bipropellant systems.12,13,24,30 An excellent

description and history of the test stand located in Test Cell A is contained in a thesis by

B.J. Austin.60

4.1 Test Facility Overview

Test Cell A at APCL houses a test stand rated to a maximum operating thrust of

1000 lbf. The existing bipropellant system contains two stainless steel tanks, a four-

gallon capacity fuel tank and four-gallon capacity oxidizer tank. The oxidizer tank is

devoted exclusively for use with hydrogen peroxide. Nitrogen is used to pressure-feed

the propellants to the test article and each of the propellant tanks are rated to a maximum

pressure of 2000 psia. Two 6000 psia storage cylinders supply nitrogen to the entire

system and are located outside of the test cell. This nitrogen supply is run through the

test cell to a regulator panel in the control room. Dome-loaded regulators, which are

tapped into the 6000 psi nitrogen supply line, are used to regulate the ullage pressure in

both the fuel and oxidizer tanks. These provide rapid response and maintain the tanks at

a very steady ullage pressure during a test run. Hand- loaded regulators located on the

panel in the control room are used to set the operating pressures of the dome-loaded

regulators as well as the purge pressures for the fuel and oxidizer lines. Typical tank

ullage pressures range from 800 to 1800 psia.

Page 74: Thesis

59

Propellant feed lines as well as nitrogen supply lines consist of a combination of

½” and ¼” diameter stainless steel tubing. Stainless steel, grades 304 or 316, is used

throughout the system due its compatibility with hydrogen peroxide.3 Flow control is

accomplished using a wide range of cavitating venturis for use in both ½” and ¼” feed

line. Propellants are vacuum loaded through fill lines connected to the bottoms of the

tanks. Suction pressure is provided by a 1/3 hp vacuum pump, which can move 1.23

ft3/min of air, and the vacuum system is controlled by manual ball valves. All of the

controllable valves in the system, excluding purge valves, are pneumatically actuated ball

valves that use helium, supplied at 100 psia, for fast response. Solenoid valves are used

to control nitrogen flow through the purge lines.

Each propellant system, fuel and oxidizer, contains six remotely operated valves.

A pressurization valve is used to control the flow of nitrogen pressurant to the tank. The

tank vent valve releases the pressurant to the environment following a test or in the case

of emergency. This is the only normally open valve in each system. A dump valve

provides a means of draining propellant from the tank. This is used if propellant remains

in the tank following a test or to safely drain the tank in emergencies. Drained hydrogen

peroxide is collected in a plastic bucket that is half full of water. This dilutes the H2O2 to

a lower concentration making it safer to handle. Drained fuel is collected in an empty

aluminum bucket. An isolation valve, located downstream of the dump valve, is used to

isolate the tank from the test article. This is only used during a test in the case of a

catastrophic failure. Downstream of the isolation valve is the main valve which controls

the flow of the propellant to the test article. This is the only timed valve in each system.

A complete plumbing and instrumentation diagram, P&ID, of the test facility is shown in

the Appendix C.

The test stand is outfitted with two parallel lengths of Unistrut metal framing

downstream of the oxidizer and fuel main valves. A mounting plate is attached to this

framing and the test article is secured to the plate. Typically the test article is mounted

vertically, but the valve box can be easily reoriented to a horizontal position to

accommodated longer test articles. The metal framing is attached to a pair of flexures

which transfers the thrust produced by the engine to a load cell.

Page 75: Thesis

60

4.2 Cavitating Venturi Flow Control

As previously mentioned flow control at this test facility is accomplished using

cavitating venturis. The internal contour of a cavitating venturi does not vary

significantly from a typical venturi flow meter, but it functions in a distinctly different

manner. The job of a cavitating venturi is to provide a large enough pressure drop to

push the local pressure of a fluid below its vapor pressure. Usually this occurs at or just

downstream of the venturi throat. When the fluid cavitates in the venturi tiny pockets of

vapor are formed that effectively create a choke point in the flow. The choke point

prevents variations in downstream pressure from propagating upstream to affect the mass

flow rate through the venturi. However, if the downstream pressure becomes too large it

will push the venturi out of cavitation and create an unstable flow rate. Generally, if the

back pressure rises to above 75% of the venturi inlet pressure cavitation will not occur.

In a venturi, as in an orifice, the flow does not exactly follow the wall contour and

often separates from the wall downstream of the throat. The separating flow creates a

vena contracta, which acts as a flow restriction and creates a fluid dynamic throat of

smaller diameter. To account for this phenomenon a constant discharge coefficient term

is used in cavitating venturi mass flow rate calculations. A well designed cavitating

venturi contour will have a discharge coefficient equal to unity. Water flows are

performed to determine the discharge coefficient for a particular venturi. The general

water flow procedure consists of flowing water through a venturi of a specific throat

diameter at a constant inlet pressure for some length of time. This information is used to

calculate what the total mass flowed would be assuming a discharge coefficient equal to

one. It is very difficult to maintain a completely constant pressure at the inlet to the

venturi so the mass flow rate is calculated based on inlet pressure at each instant in time

for the duration of the test. The mass flow rate history is integrated in time over the

entire test duration to determine the total theoretical mass. The water is collected in a

bucket and the total mass is recorded. Dividing the mass of the collected water by the

theoretical mass of water gives the discharge coefficient. Typically three water flows are

done to obtain a good average.

Page 76: Thesis

61

Mass flow through a cavitating venturi can be calculated from Bernoulli’s

equation for incompressible flow. Writing Bernoulli’s equation from the venturi inlet to

the throat of the vena contracta gives the following:

022

22

=−−+ cv

ii

Vp

Vp

ρρ, (4.1)

Here pi is the venturi inlet or supply pressure, Vi is the inlet velocity, ? is the

liquid density, pv is the vapor pressure of the liquid, and Vc is the velocity at the throat of

the vena contracta. The crude schematic of a cavitating venturi is shown in Figure 4.1.

The inlet and exit diameter of a cavitating venturi are equal to the inner diameter of the

stainless steel tubing, which is 0.18” for ¼” tubing and 0.37” for ½” tubing. Substituting

the equations for mass flow rate at the inlet and vena contracta throat into equation 4.1

results in the following after some manipulation:

( )21

2

12−

−−=

i

cvic A

AppAm ρ& , (4.2)

The cross-sectional area of the venturi inlet is denoted by Ai, Ac represents the

cross-sectional area of the vena contracta throat, and m& is the mass flow rate of liquid.

Typically the square of the ratio of the area of the contracta to the inlet is a very small

number and as such is usually neglected in mass flow calculations. Also, when compared

to the required supply pressure the vapor pressure of many liquids is extremely small and

is usually neglected as well. The vapor pressure of JP-8 is 0.190 psia at 120°F47 and the

vapor pressure of 90% H2O2 is 0.065 psia at 77°F.3 Since the cross-sectional area of the

vena contracta is not known it is usually represented as in terms of a constant discharge

coefficient, CDcv, and the physical throat cross-sectional area. Substituting this into

equation 4.2 results in the following equation:

( )21

2

12−

−−=

i

tDcvvitDcv A

ACppACm ρ& , (4.3)

Neglecting the area ratio and vapor pressure terms gives an alternate form:

itDcv pACm ρ2=& , (4.4)

Page 77: Thesis

62

This form of the cavitating venturi mass flow rate equation was used during

autoignition testing for both JP-8 and hydrogen peroxide. Typical oxidizer and fuel flow

rates can range from 0.1 to 2.5 lbm/s and 0.1 to 0.5 lbm/s respectively depending on the

venturi and supply pressure. The effective diameter, deff, of a cavitating venturi can be

written as follows:

Dcvteff Cdd = , (4.5)

The effective diameter is an alternate name for the vena contracta throat diameter

of the venturi. A list of venturis available for use at APCL including physical and

effective throat diameters as well as flow rate of water for a 1000 psia supply pressure is

shown in Table 4.1. During testing the 0.111 and 0.094 inch throat diameter venturis

were used to control the flow of H2O2 while the 0.058 inch venturi was used to control

JP-8 flow rate.

Figure 4.1: Schematic of a cavitating venturi indicating important parameters.

Page 78: Thesis

63

Table 4.1: List of cavitating venturis available at APCL.

Throat Diameter,

dt (in.)

Discharge Coefficient,

CDcv (--)

Effective Diameter, deff (in.)

Tubing Diameter

(in.)

Water Mass Flow, m& (lbm/s)

0.018 1.851 0.024 ¼ 0.076 0.019 -- 0.019 ¼ 0.047 0.022 0.998 0.022 ¼ 0.064 0.028 0.934 0.027 ¼ 0.096 0.030 0.814 0.027 ¼ 0.096 0.032 -- 0.032 ¼ 0.134 0.034 0.833 0.031 ¼ 0.126 0.036 -- 0.036 ½ 0.170 0.038 0.705 0.032 ¼ 0.134 0.040 1.018 0.040 ¼ 0.210 0.054 0.861 0.050 ¼ 0.328 0.054 0.807 0.049 ¼ 0.315 0.058 0.844 0.053 ¼ 0.369 0.062 -- 0.062 ½ 0.504 0.068 0.935 0.066 ¼ 0.572 0.078 0.839 0.071 ¼ 0.661 0.094 0.994 0.094 ½ 1.160 0.111 0.989 0.110 ½ 1.588

4.3 Data Acquisition and Control

A LabVIEW, version 6.1, virtual instrument (VI) is used for valve control and

data acquisition at the facility. The program is run on a PC operating Windows 2000

with a Pentium IV 2.0 GHz processor. During a test a VI is used to open and close

valves, to set and carry out valve firing sequences, as well as to monitor and acquire data.

The program creates a file which contains a number of rows and columns of data. Each

column corresponds to a particular channel of temperature or pressure data and each row

represents the data collected from each channel at a particular instant in time.

To actuate a valve in the system the VI is used to command a five volt

(TTL/CMOS) digital ON/OFF signal be generated by the 32-bit high-speed digital I/O

card, National Instruments (NI) model number PCI-6533 (PCI-DIO-32HS). This digital

signal is sent to a 32-channel relay board which transmits a 24 volt signal to the

Page 79: Thesis

64

appropriate valve. In a pure solenoid valve, as used in the purge lines, the 24 volt signal

actuates the solenoid which moves the valve stem and opens the valve. The valve will

remain open as long as the 24 volt signal is in place. In a pneumatic valve the signal

actuates a solenoid which allows helium at 100 psia to enter a chamber forcing the valve

stem to rotate to its open position. When the 24 volt signal is turned off the solenoid

closes and the helium is vented from the chamber.

A 16-bit resolution data acquisition (DAQ) card, NI model number PCI-6052E, is

used for a total collection capability of 333,000 samples per second of 8 differential

inputs. The analog input must be within ±0.05 and ±10.0 volts. A NI SCXI-1000 chassis

is connected to the DAQ card to expand the total channel capacity. The chassis consists

of four modules of up to 32 channels a piece for a maximum of 128 analog inputs. This

sets the maximum scan rate for sampling 128 channels at approximately 2600 samples

per second. During testing all data channels are sampled at 1000 samples per second.

Currently one module in the chassis is set up to collect thermocouple data, NI model

number SCXI-1102. The thermocouples are plugged into a terminal block, NI model

number TC-2095, which is connected to the module. The terminal block can

accommodate up to 32 channels. The voltage signal generated by the thermocouples is

automatically converted to temperature using LabVIEW and the terminal block which

contains cold junction compensation circuitry. Two of the remaining three modules, NI

model number SCXI-1102C, are set aside for pressure transducer data as well as any

other input signals of less than 10 volts, such as load cell, valve position, or heat flux

transducer data. The transducer wires are connected to a terminal block, NI model

number SCXI-1303, one of which is connected to each of the two modules in the chassis.

Each terminal block can accept up to 32 channels of data. Typically pressure

measurements are recorded as voltage in a data file and converted to units of pressure by

a data reduction code.

Page 80: Thesis

65

4.4 Instrumentation

Pressure measurements are made using 3000 psi absolute pressure transducers

manufactured by Druck, model number PMP 1260 or 1265. These transducers follow a

linear calibration curve and output one volt at a pressure of 0 psia and five volts at 3000

psia. Using these two points the slope of the calibration curve is 750 psia per volt. The

transducers are accurate to within ±0.25% of the full scale pressure, which corresponds to

an accuracy of ±7.5 psia. At APCL the transducers are powered by a 24 volt DC supply,

the excitation voltage must be between 8 and 30 volts DC for normal operation. These

transducers can withstand an overpressure of up to 6000 psia and operate nominally

within -40 and 185°F.

Temperature measurements are made using type K thermocouples made by

Omega Engineering, Inc. The temperature sensing junction of these thermocouples is

formed by Chromega and Alomega wire, which are nickel/chromium and

nickel/aluminum alloys respectively. These wires are contained in a stainless steel, SS

304, sheath and are attached to a two-prong connector. Insulation is used to isolate the

wires from each other and the sheath material. This combination of junction and sheath

material gives thermocouple an operating temperature range of up to 1700°F. The probe

diameter is 1/16 inch and is 12 inches in length. Exposed and grounded junctions were

used for this testing. In a grounded thermocouple the wires are actually attached to the

sheath material inside the tip of the probe. The sensing junction protrudes from the tip of

an exposed thermocouple probe leaving it exposed to the environment.

4.5 Test Article and Setup

A simplified schematic of the test stand including the location of pressure

transducers and thermocouples is shown in Figure 4.2. Pressure transducers are located

at the top of both propellant tanks to measure and monitor the ullage pressure.

Temperature of the liquid hydrogen peroxide in the oxidizer tank is measured using a

Page 81: Thesis

66

grounded thermocouple. This temperature is routinely monitored when the tank contains

H2O2 for safety purposes in the event of fluid decomposition. Transducers are also

located just upstream of the fuel and oxidizer cavitating venturis to measure the inlet

pressure. This pressure is used in for mass flow rate calculations, equation 4.4. A

pressure transducer is placed just upstream of the inlet to the fuel injector manifold. This

measurement is used along with chamber pressure to calculate the pressure drop across

the injector. Another transducer is positioned just upstream of the inlet to the catalyst bed.

Pressure drop across the catalyst bed is calculated using this measurement and the

pressure measured downstream of the catalyst bed.

Two different catalyst beds were used during the course of autoignition testing.

The baseline catalyst bed is a silver screen design and was made by General Kinetics,

LLC. Due to its silver screen design hydrogen peroxide concentration is limited to 92%

or less. This bed was designed to operate at a flow rate of 2.5 lbm/s of 90% H2O2 or a

bed loading of approximately 0.38 lbm/in2-s. Bed loading, G, is defined as the mass flow

rate divided by the cross-sectional area that the flow sees in the catalyst bed. The

diameter of the flow path through this bed is about 2.9 inches. A second bed, made by

Precision Combustion, Inc., has a different internal design which allows it to operate

using hydrogen peroxide concentrations up to 98%. This bed was also designed to

operate at a flow rate of approximately 2.5 lbm/s of 90% H2O2, but the loading is double

that of the baseline bed at about 0.80 lbm/in2-s. The diameter of the flow path through

the bed is about 2.0 inches. Table 4.2 shows a comparison between the two catalyst bed

designs.

Table 4.2: Comparison between catalyst bed designs used during testing.

Catalyst Bed General Kinetics, LLC. (GK)

Precision Combustion, Inc. (PCI)

General Design Silver Screen Proprietary H2O2 Concentration Up to 92% Up to 98%

Design Flow Rate (lbm/s) 2.50 2.50 Design Loading (lbm/in2-s) 0.38 0.80

Internal Diameter (in) 2.90 2.00 Exit Diameter (in) 2.90 1.75

Page 82: Thesis

67

Since the catalyst beds have different exit geometry two different engine

assemblies were used during the course of autoignition testing. The difference between

the two is in the method and hardware used to mate each catalyst bed to the transverse

injector. All hardware downstream of and including the transverse injector is identical in

both cases. A schematic of the two assemblies is shown in Figure 4.4 and a photo of

each is shown in Figure 4.3. All hardware excluding the chamber and nozzles are

fabricated from stainless steel, SS 304. Copper is used for the combustion chamber and

nozzles because of its high thermal diffusivity, which means that heat travels through the

material very quickly. As a result, the temperature at the inner walls of the chamber and

nozzles rises slowly extending the time which they can be exposed to high temperatures.

However, even with copper the biprop test duration is limited to five seconds or less, at a

mixture ratio of five with 90% H2O2 and JP-8, to avoid thermal damage to the throat.

The chamber inner diameter is 2.56 inches and three nozzles were fabricated at

contraction ratios of 3.0, 5.0, and 6.5. This corresponds to throat diameters of 1.478,

1.145, and 1.000 inches respectively. These contractions ratios set the characteristic

length of the chamber, measured from the face of the rearward step to the throat, at 24.1,

40.4, and 52.7 inches respectively. The nozzles with contraction ratios of 3.0 and 6.5 are

shown in Figure 4.4. Viton o-rings are used to seal most of the interfaces between

engine hardware. This material is compatible with hydrogen peroxide and has a

maximum operating temperature of 600°F. Typically, for tests of between three and five

seconds, these o-rings will begin to deteriorate from exposure to high temperatures.

Pressure-filled stainless steel o-rings are used for sealing between the catalyst beds and

their mating pieces. These o-rings were specified by the catalyst bed manufacturers. A

water-cooled carbon steel deflection plate was used to direct the exhaust gases from the

engine outside of the test cell. The deflection plate can be seen at the bottom of the photo

in Figure 4.3a. Part drawings for the deflection plate are contained in Appendix A.

Page 83: Thesis

68

Figure 4.2: Simplified schematic of test stand with instrumentation locations.

a) b)

Figure 4.3: Photos of installed test article for: a) GK catalyst bed and b) PCI catalyst bed assemblies.

Page 84: Thesis

69

Figure 4.4: Comparison of engine assemblies for each catalyst bed design. Axial

locations of instrumentation are indicated in schematic.

The locations of pressure transducers and thermocouples on the test article are

shown in Figure 4.4. This can also be seen in the photos in Figure 4.3. General Kinetics,

LLC specifies that the catalyst be heated to approximately 350°F prior to firing. This

was accomplished using a 150 Watt, cloth- insulated electric wrap heater purchased from

Cole Palmer Instrument Company. The wrap heater was two inches wide and two feet

long and can be seen wrapped around the catalyst bed at the top of Figure 4.3a. A

thermocouple was surface welded to the outside wall of the of the GK catalyst bed to

monitor its temperature, Tskin. Immediately downstream of the catalyst bed in both

assemblies an exposed thermocouple is used to measure the decomposition temperature,

Td, of the hydrogen peroxide. This thermocouple is positioned such that the probe is at

the centerline of the duct. The pressure at the exit of the catalyst bed, Pcb_out, is also

Page 85: Thesis

70

measured by a transducer at the same position. Four grounded thermocouples were used

in the combustion chamber to measure the temperature profile along the chamber length,

Tc1-Tc4. These probes are positioned flush with the inner chamber wall and spaced

equally along the length of the chamber. Two pressure transducers are used in the

combustion chamber. The transducer at the first axial position in the chamber measures

the static pressure in the recirculation zone behind the rearward-facing step, Pc1. The

second measures the static pressure just upstream of the chamber contraction, Pc2. This

measurement is used to calculate the characteristic velocity. A typical list of

instrumentation used during autoignition testing is shown in Table 4.3.

Table 4.3: Typical instrumentation list for autoignition testing.

Description Instrument Designation Oxidizer Tank Ullage Pressure Druck 0-3000 psia Pox_ull

Oxidizer Venturi Inlet Pressure Druck 0-3000 psia Pox_cv Catalyst Bed Inlet Pressure Druck 0-3000 psia Pcb_in Fuel Tank Ullage Pressure Druck 0-3000 psia Pfu_ull Fuel Venturi Inlet Pressure Druck 0-3000 psia Pfu_cv Fuel Injector Inlet Pressure Druck 0-3000 psia Pfu_inj

Catalyst Bed Outlet Pressure Druck 0-3000 psia Pcb_out Chamber Pressure 1 Druck 0-3000 psia Pc_1 Chamber Pressure 2 Druck 0-3000 psia Pc_2

Oxidizer Tank Temperature Omega Type K Grounded Tox

Decomposition Temperature Omega Type K Exposed Td Chamber Temperature 1 Omega Type K Grounded Tc1

Chamber Temperature 2 Omega Type K Grounded Tc2

Chamber Temperature 3 Omega Type K Grounded Tc3

Chamber Temperature 4 Omega Type K Grounded Tc4

Catalyst Bed Skin Temperature Omega Type K Surface Welded Tskin

Page 86: Thesis

71

4.6 Hydrogen Peroxide Dilution

Propellant grade hydrogen peroxide from FMC Corporation was used during

autoignition testing. Hydrogen peroxide from FMC comes standard in 90 and 98%

concentrations. These standard concentrations were diluted using deionized water prior

to testing in order to create non-standard concentrations such as 85, 87.5, and 94%. The

amount of H2O2 needed at these lower concentrations is usually known prior to testing.

Therefore, one must calculate the amount of dilution water and hydrogen peroxide at

standard concentration that must be mixed to meet the new mass and concentration

requirements. The required water mass, mH2Oadd, is determined from the following

equation:

2@221

22 1 WOHOaddH m

WW

m

−= , (4.6)

Here W1 is the current or standard concentration (in weight percent), W2 is the

desired concentration, and mH2O2@W2 is the desired mass of hydrogen peroxide as the new

concentration. Note that the mass to be added becomes negative if the desired

concentration is greater than the current concentration, W2 > W1. The mass required at

the current concentration, mH2O2@W1 , is found from:

2@221

21@22 WOHWOH m

WW

m

= , (4.7)

Once these masses were determined the required amounts of water and hydrogen

peroxide were measured, poured into a plastic jug, sealed, and shaken by turning end

over end until the liquids were well mixed. The concentration was then measured to

verify that it matched the desired value. The hydrogen peroxide was then stored in the

jug at room temperature until it was loaded into the run tank during the test. Both the

hydrogen peroxide and JP-8 fuel were stored at room temperature, around 70°F, for these

autoignition tests.

The concentration of hydrogen peroxide is determined by measuring the refractive

index of the liquid. The refractive index can be used to determine the hydrogen peroxide

Page 87: Thesis

72

concentration since the density of H2O2 varies with concentration and the refractive index

of a fluid is dependent on its density. The density of hydrogen peroxide, ?H2O2, varies

with temperature and concentration according to the following curve-fit equation:3

?H2O2 = 66.166 + 1.577*10-1W + 1.112*10-3W 2 – 2.310*10-2T –

4.700*10-6T 2 - 1.380*10-4WT, (4.8)

Here density is in units of lbm/ft3, W is the weight percent of hydrogen peroxide,

and T is the H2O2 temperature in °F. At APCL a refractometer is used to measure the

refractive index of hydrogen peroxide. To determine concentration the measured

refractive index of the hydrogen peroxide is compared against previously recorded values

of refractive index for a range of H2O2 concentrations.3 This data was recorded at a fixed

temperature of 77°F therefore a heater/refrigerator system is used to maintain the

refractometer at this temperature as well.

4.7 Pressure Budget

Prior to beginning test procedures it is customary to calculate the expected

readings from each piece of instrumentation. Estimates of the pressure at each

measurement location in the system are especially important when using cavitating

venturis since the downstream pressure has a significant influence on cavitation. Since

only a certain range of back pressures are allowed based on venturi supply pressure and

flow rate the summary of system pressures is often termed a pressure budget. Calculation

of the pressure budget begins by calculating the chamber pressure, Pc2, during the

monoprop and biprop modes of operation. Equation 2.1 is used to calculate chamber

pressure based on known test parameters of throat diameter, mass flow rate, and C* and

an assumed C* efficiency. An efficiency of 98% is a good upper bound estimate for both

monoprop and biprop cases. The characteristic velocity is dependent on H2O2

concentration for the monoprop case and mixture ratio for the biprop case. Both of these

parameters are set for each test.

The pressure drop across the fuel injector can be calculated using equations 2.2

through 2.4 and added to the chamber pressure to determine the injector inlet pressure,

Page 88: Thesis

73

Pfu_inj. The cavitating venturi supply pressure, Pfu_cv, is calculated with equation 4.4

using the desired mass flow rate and selected cavitating venturi geometry, Table 4.1. To

verify cavitation the injector pressure is divided by the fuel venturi inlet pressure. If this

ratio is greater than 0.75 cavitation will not occur and a new venturi must be selected

which requires a higher supply pressure to meet the necessary flow rate. In some cases

none of the venturis available at APCL meet the mass flow and supply pressure

requirements. A new cavitating venturi may need to be ordered at that point

Calculations for the oxidizer line pressures work similarly to the fuel, but the

catalyst bed pressure drop is more difficult to calculate. Typically, the pressure drop is

estimated based on previous test data. A conservative assumption for pressure drop

across the GK bed is 400 psid and for the PCI bed is 500 psid for both operational modes.

The pressure drop will be lower during biprop operation than monoprop due to the

increased back pressure on the bed. This decreases the flow rate through the bed and as a

result the pressure loss. The pressure drop is added to the catalyst bed outlet pressure,

which can be assumed to be roughly equivalent to the chamber pressure, to determine the

catalyst bed inlet pressure, Pcb_in . The venturi calculations for the oxidizer line proceed in

the same manner as for the fuel line. Ullage pressures for both the oxidizer and fuel

system, Pox_ull and Pfu_ull, are typically anywhere from 0-150 psi greater than the venturi

inlet pressures. This is due to pressure losses associated with the valves and bends in the

lines between the tank and the venturi. These losses increase as the flow rate, or velocity

in the feed line, increases. A conservative estimate would be anywhere between 50 and

150 psid, but it is better to guess high than low to ensure cavitation. Deviations of actual

pressure drop from these estimates will affect the mass flow rate through the system.

The upstream chamber pressure, Pc1, is more difficult to estimate due to its

location in the recirculation zone behind the rearward-facing step. It is usually estimated

to be equal to the downstream chamber pressure. The decomposition temperature can be

estimated using a thermochemistry code4 and is dependent on concentration. The

chamber skin temperature should be roughly equal to 350°F and oxidizer temperature

should be around room temperature, 77°F. The chamber temperatures, Tc1-Tc4, are very

difficult to estimate and typically fail due to high temperatures in the chamber. They

Page 89: Thesis

74

roughly measured the wall temperature and a conservative estimate would be around

1500°F for biprop operation and around 800°F for monoprop. A table of estimated

conditions for the baseline design conditions, see Chapter 2, is shown in Table 4.4. The

baseline engine used the 0.111 inch throat diameter cavitating venturi for the oxidizer line

and the 0.058 inch venturi for the fuel line, see Table 4.1.

Table 4.4: Pressure budget for a test at baseline design conditions using the GK catalyst bed, see Chapter 2.

Designation Monoprop Biprop Pox_ull (psia) 1550 1550 Pox_cv (psia) 1400 1400 Pcb_in (psia) 667 920

Pcb_in/Pox_cv (--) 0.43 0.66 Pfu_ull (psia) -- 1750 Pfu_cv (psia) -- 1700 Pfu_inj (psia) -- 608

Pfu_inj/Pfu_cv (--) -- 0.36 Pcb_out, Pc_1, Pc_2 (psia) 267 520

Tox (°F) 77 77 Td (°F) 1390 1390

Tc1-Tc4 (°F) 800 1500 Tskin (°F) 350 350

4.8 Test Procedure

Four people are required to conduct a rocket test at APCL, and each of these

people has a specific set of responsibilities during the test. The test conductor is

responsible for all test operations, reads the test procedures, and maintains the list of test

conditions. The test operator loads propellants, operates manual valves and regulators,

and performs other functions related to propellant or pressurant as dictated by the test

conductor. The data system operator runs the LabVIEW program, monitors and records

test data, and maintains operability of all instrumentation and controls for each test. The

site safety director maintains functionality of safety equipment, keeps site clear of

unauthorized personnel, and ensures that test personnel follow safety procedures at all

times.

Page 90: Thesis

75

A typical test procedure can be split into four parts: test preparation, propellant

loading, test firing, and shutdown. Test preparation begins by measuring the

concentration of the hydrogen peroxide and, if required, diluting to a lower concentration.

Next, the catalyst bed heater is turned on when using the GK catalyst bed. Typically

one-half hour is required for the bed to heat up sufficiently, which is roughly the amount

of time it takes to complete the test procedures. Following this step the computer,

instrumentation, valves, relay board, and chassis are powered up. Next the LabVIEW

program file is opened and the file name, scan rate, and channel list are set. Once this

information has been verified by the test conductor the data system operator can begin

running the program. Next the communication between the computer and the

instrumentation is checked as well as the readings from each instrument to verify they are

functioning properly. Then the outside doors to the test cell are opened, and the cameras

are set up and checked for functionality. Then the VHS tapes used to record each camera

view are queued to the correct position and the test number is recorded on each tape.

Following this all of the manual regulators on the panel in the control room are fully

unloaded, the nitrogen and helium supplies are checked for adequacy, and if so the

helium regulator is loaded to 100 psia. For most tests a nitrogen supply of at least 2200

psia is required to maintain ullage pressure. Next, the remotely controlled valves are

cycled to make sure they are functioning properly. During the final step of test

preparation each of the manual and remote valves are verified to be in their power off

position.

The fuel is loaded first during the test procedure because it can safely sit in its

tank for a longer period of time than hydrogen peroxide. The loading procedure is

identical for both propellants, but extra care is required for handling hydrogen peroxide

and is described here. Following test preparation the oxidizer tank vent valve is closed

and the isolation valve is opened, this allows the propellant to fill the system all the way

down to the main valve. The vacuum isolation valve is opened, the vacuum line is

connected to the pump, and the pump is turned on. At this point the warning light is

turned on to alert anyone outside the building that a test is about to take place. The

pressure in the tank is monitored and when it reaches approximately 4 psia the vacuum

Page 91: Thesis

76

isolation valve is closed, the pump is turned off, and the vacuum line is disconnected

from the pump. Next, the test operator puts on the appropriate safety gear for handling

hydrogen peroxide and brings the propellant into the test cell. The jug of propellant is

placed on a scale and the initial mass is recorded, the fill valve is opened, and the

propellant is sucked into the tank. The data system operator must monitor the

temperature and pressure in the tank at all times while hydrogen peroxide is being loaded.

In the event that rapid decomposition begins in the tank these are the first signs of

problems. When the required amount of mass has been loaded into the tank the fill valve

is closed and the final mass is recorded. Then the vent valve is opened and the tank is

allowed to return to ambient pressure. Before returning the propellant jug to storage and

taking off safety gear the test operator must make a thorough review of the propellant

tank and feed lines to make sure there are no obvious propellant leaks. Since fuel loads

are usually small a graduated cylinder is used to load the propellant into the tank. This

requires volume-based loading rather than mass-based, which is the case for H2O2.

While all the pressure transducers are measuring ambient pressure the data system

operator acquires two seconds of data. This data is used to zero the pressure transducer

measurements in the data reduction code. This is explained in the next section. Next, the

door between the test cell and control room is bolted shut. Following this the oxidizer

and fuel purge regulators are loaded to the desired pressure. Typically, the oxidizer purge

is loaded to 100 psia while the fuel purge is load to 1.5 to 2.0 times the monoprop

chamber pressure. This is to ensure that hot decomposed gases do not enter the fuel

injector during the monoprop stage of the test. Then the tank vent valves are closed, the

pressurization valves are opened, and the ullage pressure regulators for both tanks are

loaded to the desired test pressures. As previously described, these regulators actually set

the pressure in the dome of the dome-loaded regulators located in the test cell. Once the

tanks have reached the desired ullage pressure the valve timing sequence is entered into

the LabVIEW code. The timing sequence is unique to each catalyst bed and will be

explained in more detail later. Next, the VCR’s are set to begin recording, data

acquisition is started, and then the fire button in LabVIEW is pressed to start the fire

sequence. Once the sequence is complete the VCR’s are stopped.

Page 92: Thesis

77

After the completion of the test the ullage pressure regulators are unloaded, the

tank vent valves are opened, and the tank pressure is allowed to return to ambient. While

the tank is venting the purge valves are opened to purge any remaining propellant from

the engine. The oxidizer purge valve is always opened before the fuel purge to prevent

fuel from being blown into the catalyst bed. This could potentially cause significant

internal damage to the catalyst bed. While the purge valves are still open the regulators

are unloaded, fuel first, and then the valves are closed. This guarantees that all the

pressure has been vented from the purge lines. At this point the tanks should be fully

vented and the pressurization and isolation valves are closed. In most cases, the

propellants are run to completion during a test. This means that there is no propellant left

in the tanks or lines following the test. If propellant is suspected to remain in the lines a

dump procedure, not described here, must be followed to completely drain the

appropriate tank or tanks. At this point the test is completed, the LabVIEW program can

be stopped, and all the power can be turned off. The nitrogen and helium bottles must be

closed and each line must be vented to ambient pressure.

4.9 Firing Sequence

Each catalyst bed required some form of warm-up prior to running an autoignition

test. This was done to make sure that the bed was fully decomposing the hydrogen

peroxide before any fuel was supplied to the engine. The decomposition efficiency of the

catalyst bed can be determined in a number of ways including monitoring the chamber

pressure, decomposition temperature, or by visual observation of the exhaust gases. A

clear exhaust plume indicates full decomposition while a cloudy plume indicates poor

decomposition. The GK catalyst bed, as previously mentioned, is pre-heated for

approximately one-half hour prior the actual test firing. In addition to this the catalyst

bed is fed a number of short pulses of hydrogen peroxide to warm it further. A typical

pulse sequence consists of two half second pulses followed by two one second pulses

with at least ten seconds in between. The ten second waiting period allows the heat

generated during the pulse to soak into the bed. A minimum of ten seconds wait period is

Page 93: Thesis

78

required before initiating the autoignition test as well. This pulse sequence is usually

sufficient to produce good decomposition. The PCI catalyst bed, on the other hand, is not

built to handle short duration, cyclic pulsing. For this reason the bed is warmed by

running it for five seconds and followed by at a least ten second wait before commencing

the autoignition test. By the end of the ten second warm-up period the catalyst bed

should be adequately decomposing the hydrogen peroxide. For both catalyst beds, the

mass flow rate of hydrogen peroxide used during the warm-up period should be the same

as required for the autoignition test. A benefit of the catalyst bed warm-up period is that

is offers the opportunity to evaluate the catalyst bed performance prior to the autoignition

test. If the catalyst beds are not functioning properly at the end of the warm-up period the

test can be aborted at that point. Whereas, if there were no pulse sequence there would

not be enough time recognize poor decomposition and then abort before the fuel flow was

turned on.

Typically before any pulsing has been performed the purge lines are opened in the

system. As previously mentioned, the oxidizer purge valve should always be opened

before the fuel purge to prevent fuel vapor from entering the catalyst bed. Fuel droplets

usually remain in the injector manifold from previous tests and can be blown into to the

engine when the nitrogen purge is turned on. Therefore, the oxidizer purge valve is

opened first followed by the fuel purge valve. Once it is visually verified that there is no

fuel vapor exiting the engine the oxidizer purge valve is turned off. This is to prevent the

nitrogen from cooling the catalyst bed and from contaminating the catalyst material. The

fuel purge usually remains on during the pulse sequence as well as the autoignition test.

There are check valves in both the oxidizer and fuel purge lines to prevent liquid

propellant or other gases from penetrating into the lines during the test. If at any point

during test operations the fuel purge valve is closed the oxidizer purge must be opened

before the fuel purge can be opened again. As soon as any fuel vapor has dissipated the

oxidizer purge is closed. Also, before the autoignition test the water flow to the

deflection plate is turned on. It takes approximately ten seconds for the water to fill the

manifold behind the plate at which point it begins to flow full. After the test the water

flow is turned off.

Page 94: Thesis

79

Following the pulse sequence the autoignition test is begun by starting the

hydrogen peroxide flow to the catalyst bed. After one-half second the fuel flow is turned

on and remains flowing for another second. Following fuel shutdown the hydrogen

peroxide flows for two more seconds and then is turned off. By lagging the fuel by

one-half of a second any transients involved with the start-up of the catalyst beds are

bypassed allowing the beds to reach steady-state decomposition. This lag is also short

enough that it does not allow the decomposed gases to heat the engine significantly,

which may affect autoignition behavior. If the conditions in the engine are sufficient

autoignition should occur almost instantaneously upon fuel injection. For this reason the

biprop test duration is limited to one second in length. Usually, after the fuel flow is

turned off, there is residual fuel remaining in the feed lines that trickles into the engine.

To burn off any residual fuel in the engine, injector, or feed lines the hydrogen peroxide

flow remains on for two additional seconds following fuel shut down. Typically the

oxidizer and fuel main valves remain open for one second following propellant depletion

to allow nitrogen in the tank to exhaust through the engine. This thoroughly purges the

entire feed system of any remaining drops of propellant. The typical valve sequencing

for each catalyst bed engine assembly is shown in Table 4.5.

Page 95: Thesis

80

Table 4.5: Typical valve firing sequence for each catalyst bed design.

Valve (ON/OFF)

GK (seconds)

PCI (seconds)

Oxidizer Purge ON t – 58.0 t – 30.0 Fuel Purge ON t – 56.0 t – 28.0 Oxidizer Purge OFF t – 51.0 t – 23.0 Oxidizer Main ON t – 48.0 -- Oxidizer Main OFF t – 47.5 -- Oxidizer Main ON t – 37.5 -- Oxidizer Main OFF t – 37.0 -- Oxidizer Main ON t – 27.0 -- Oxidizer Main OFF t – 26.0 -- Oxidizer Main ON t – 16.0 t – 20.0 Oxidizer Main OFF t – 15.0 t – 15.0 Deflection Plate Water ON t – 10.0 t – 10.0 Oxidizer Main ON t – 0.5 t – 0.5 Fuel Main ON t – 0.0 t – 0.0 Fuel Main OFF t + 1.0 (+1.0) t + 1.0 (+1.0) Oxidizer Main OFF t + 2.0 (+1.0) t + 2.0 (+1.0) Deflection Plate Water OFF t + 5.0 t + 5.0 Fuel Purge OFF t + 10.0 t + 10.0

Page 96: Thesis

81

CHAPTER 5: EXPERIMENTAL RESULTS

As discussed in Chapter 3 the primary objective of this study was to develop an

autoignition model that could be used to aid in the design of a H2O2/kerosene staged

rocket engine. Based on past data the decomposed gas temperature, equivalence ratio,

and gas velocity were reasoned to be the most influential to autoignition in these engines.

Testing was structured to investigate each of these factors through the hydrogen peroxide

concentration, mixture ratio of H2O2 to JP-8, and chamber contraction ratio respectively.

A secondary objective of this study was to develop a better understanding of the design

and operation of these staged-bipropellant engines. The data and experience accumulated

through each autoignition test builds toward this goal.

5.1 Test Plan Overview

In most kerosene autoignition studies the gas temperature is gradually increased

over the course of a test to the point of autoignition. Operating conditions such as

pressure, gas flow velocity, and equivalence ratio remain relatively constant during this

process. When air is used as the working gas it can be gradually heated to increase its

temperature, but modifying the temperature of decomposed hydrogen peroxide requires a

different approach. Since the decomposition temperature of hydrogen peroxide is

dependent on its concentration in liquid form the concentration must be varied to alter the

decomposition temperature. Walder8 mixed water with hydrogen peroxide prior to

entering a catalyst pack to continuously control its concentration. Unfortunately, the

setup of Test Cell A at APCL at the time of testing was not conducive to this type of

operation. In addition, if modifications were made to the stand there still would have

Page 97: Thesis

82

been a number of issues to consider regarding the mixing of water and hydrogen peroxide.

Some of these include: determining the time and method required to completely mix the

liquids, damaging the catalyst beds through water soaking, altering the control program

for active adjustment of water flow rate, and sizing new propellant tanks to meet total

expected run time during a test. To avoid these issues hydrogen peroxide of 90 and 98%

concentrations were diluted to lower concentrations, as described in Chapter 4, prior to

testing. The advantage of this approach is that the concentration can be measured very

precisely prior to loading. However, a disadvantage exists in that the decomposition

temperature is set for each test.

The test plan for this autoignition study was not predetermined, but instead

evolved based on the outcome of each test. Testing began by running the engine at

baseline conditions, Chapter 2, but using a nozzle with a contraction ratio of 3.0. If

autoignition were achieved at these conditions the equivalence ratio would be increased

for the following test. If autoignition occurred at these new conditions then the

equivalence ratio would be increased again for the next test and again until a value was

found at which no autoignition occurred. In event that autoignition was not achieved

during the initial test the equivalence ratio would be decreased for the next test.

Subsequent tests would follow a similar pattern as the first case but now the equivalence

ratio would continually be decreased until autoignition occurred. This procedure

determines the autoignition boundary in terms of equivalence ratio at a specific

decomposition temperature and contraction ratio. After defining the autoignition

boundary for 90% H2O2 at a contraction ratio of 3.0 this process could be extended to

other concentrations. In addition, the contraction ratio could be changed and the process

repeated at a constant concentration.

To modify the equivalence ratio the hydrogen peroxide mass flow rate was varied

while the fuel mass flow rate was kept at the baseline value. The reasoning behind this

was that if the hydrogen peroxide flow were kept constant the fuel flow would need to be

increased to increase equivalence ratio. Since pressure drop varies with the square of the

fuel flow rate the injector pressure drop would increase substantially with flow rate along

with the venturi back pressure as a result. If the fuel flow rate were increased past a

Page 98: Thesis

83

certain point cavitation would be impossible with any venturi because the back pressure

on the venturi would be too large. On the other hand, the fuel flow rate would need to be

decreased to decrease equivalence ratio. Thus, pressure drop would decrease as well,

which would ensure cavitation. However, the injection velocity would also decrease

lowering the fuel momentum and potentially causing poor fuel distribution. In addition,

if the injector pressure drop fell too low in relation to chamber pressure it could affect the

stability of the engine. New injectors could be fabricated with new geometry to correct

these problems, but the engine would need to be disassembled to insert the new hardware.

As a result, it was decided to vary the hydrogen peroxide flow rate. One benefit of

varying H2O2 flow rate over fuel flow rate is that the pressure drop across a catalyst bed

is typically much less sensitive to changes in mass flow rate than a fuel injector. In

addition, catalyst beds are usually more resistant to instability over a fairly large range of

flow rates.

Three methods were used to determine whether or not autoignition occurred

during each test. The first method of evaluation was through real-time visual observation

of the exhaust plume. This was accomplished by viewing two television monitors,

located in the control room, during the test. These views were provided by two cameras

located in the test cell, one on the north side and one on the south side of the cell. In

most cases it is difficult to observe the presence of a delay between the instant of fuel

injection and the appearance of a flame in real-time. For this reason, a second evaluation

of autoignition was achieved through frame-by-frame playback of the test. The cameras

used during testing recorded at approximately 30 frames per second or about one frame

every 30 milliseconds. Significant delays in autoignition or unstable combustion can be

identified at these slow playback speeds. Finally, autoignition could also be scrutinized

by viewing plots of chamber pressure data versus time. The rise in magnitude of

chamber pressure following fuel injection can be evaluated using this data as well as the

lag between the chamber pressure and fuel injector pressure rise.

Page 99: Thesis

84

5.2 Data Reduction

As shown in Table 4.3, there are sixteen different measurements that are made

during a typical autoignition test. The output from each of these pressure transducers and

thermocouples are connected to a specific terminal or channel in the SCXI chassis.

During a test LabVIEW is used to sample the reading from each channel and record it to

a data file 1000 times each second. Each successive sample of data is placed in a new

row in the data file and each channel of data is placed in its own column in the file. As a

result, sampling all sixteen channels for one second would create a data file with 1000

rows and 16 columns. Typically each column of data is referred to as an array.

5.2.1 Pressure Transducer Data

Since the thermocouple readings are converted to units of degrees Fahrenheit by

LabVIEW it is recorded in temperature units in the data file. Pressure data, on the other

hand, is recorded in units of volts and must be converted to pressure units, psia, using a

MATLAB data reduction file, shown in Appendix D. As described in Chapter 4, the

Druck pressure transducers have a linear calibration curve of pressure versus voltage.

For a transducer with a range of 0-3000 psia the slope of the calibration curve is 750

psi/volt. The first step in conversion to pressure units is to multiply the voltage data by

the slope of the calibration curve, mpt:

[ ] [ ]ipti Vmp = , (5.1)

Here [V]i is an array of voltage data and [p]i is the calculated array of pressure

data. Brackets, [], are used to denote that voltage and pressure are one-dimensional

arrays. The channel of pressure transducer data used in the calculation is denoted by the

subscript ‘i.’ As it currently stands this equation does not give the correct absolute

pressure reading, i.e. when an element of [V]i is equal to five volts the same element of

[p]i does not equal 3000 psia but 3750 psia. An adjustment factor must be used to correct

the pressure calculated in equation 5.1 to the actual pressure. Typically this is done by

Page 100: Thesis

85

taking two seconds of data, called zero data in Chapter 4, while the entire system is at

ambient pressure. At this point all pressure transducers should be reading approximately

14.7 psia. The voltage data taken during this initial two seconds is multiplied by the

slope of the calibration curve using equation 5.1 and then averaged. The adjustment

factor is calculated as follows:

avgzeroiptatmavgzeroiatmadji Vmpppp _____ −=−= , (5.2)

The constant adjustment factor is represented as pi_adj, the atmospheric pressure is

denoted by patm, and the average zero pressure and voltage data are represented as

pi_zero_avg and Vi_zero_avg respectively. The subscript ‘i’ is used in equation 5.2 because the

averaged zero values and adjustment factors are different for each transducer. The actual

pressure as measured by the transducers can now be calculated from:

[ ] [ ] adjiipti pVmp _+= , (5.3)

The data reduction code is also used to plot pressure and temperature data as a

function of time. Time is calculated by divided the sample number by the sample rate. A

typical plot of chamber pressure, pc2, versus time during a test in which autoignition

occurred is shown in Figure 5.1. The monoprop, biprop, and shutdown segments of the

test are clearly distinguished in the plot. A steady chamber pressure is achieved within

approximately 0.2 seconds of the start of both the monoprop and biprop sections of the

test.

Figure 5.1: Plot of pc2 from an autoignition test using PCI catalyst bed. Test conditions:

98% H2O2, φ = 1.84, CR = 3.0.

Page 101: Thesis

86

5.2.2 FFT and Filtering

A benefit of using MATLAB for data reduction is that it has built- in fast Fourier

transform, FFT, and filter design functions. The fast Fourier transform is an algorithm

used to speed up the computation of the discrete Fourier transform, DFT, of an array of

data. The DFT is used to convert data from the time domain to the frequency domain and

is used to determine which frequencies are most dominant in a set of noisy or oscillatory

data. Knowledge of the dominant frequency in chamber pressure data can help in

designing a filter to drown out the noise or to determine if an instability exists in the

chamber. A plot of a noisy chamber pressure signal during monoprop operation is shown

in Figure 5.2a).

A signal in the time domain is broken down into a sum of sine and cosine

functions with specific magnitudes and natural frequencies by the DFT. Resolution in the

frequency domain, or the resolution in the natural frequencies of the sine and cosine

functions, is dependent on the resolution in the time domain and the number of data

points used to compute the DFT. The time resolution for this study is equal to 0.001

seconds, or one over the scan rate. Most FFT algorithms require 2p points to compute the

FFT more efficiently, usually the number of points is set to 256, or p = 8. This would set

the frequency resolution at approximately 4 Hz based on the time resolution used in this

study. During data reduction the number of points used in the FFT was typically much

greater than 256, often nearer to 1024 points or p = 10, which dropped the frequency

resolution to less than 1 Hz. The ‘fft’ function in MATLAB performs a discrete Fourier

transform on an array of data and outputs the real and imaginary parts of the Fourier

coefficients corresponding to each frequency. The magnitude of the Fourier coefficients

is computed by taking the square root of the sum of the squares of the real and imaginary

parts. A plot of the magnitude of the Fourier coefficients versus frequency is called a

power spectrum. The power spectrum indicates which frequencies are most dominant in

a particular set of data. The power spectrum computed from the data in Figure 5.2a) is

shown in Figure 5.3a) and b). The high magnitude at a frequency of approximately 43

Hz indicates that this is the dominant frequency in the data.

Page 102: Thesis

87

A low-bypass filter is used to eliminate frequencies inherent in a set of data that

are greater than a critical value, called the cutoff frequency. Plotting the power spectrum

of a set of data is often helpful in determining the cutoff frequency. A function in

MATLAB, called ‘butter’, designs a low-bypass digital Butterworth filter of order n with

a specified cutoff frequency. The coefficients that are to be used in the numerator and

denominator of the nth order transfer function describing the filter are output as two

separate arrays by the ‘butter’ function. These arrays as well as the array of data to be

filtered are input to a second MATLAB function called ‘filter.’ This function outputs the

filtered form of the data which excludes all frequencies greater than the cutoff frequency.

The data plotted in Figure 5.2a) was filtered using a cutoff frequency of 15 Hz and the

result is shown in Figure 5.2b). This plot shows that nearly all the noise and oscillations

were removed from the signal.

a) b)

Figure 5.2: Plots of pc2 in a) unfiltered and b) filtered form. Test conditions: 87.5% H2O2, φ = 1.40, and CR = 3.0.

Page 103: Thesis

88

a) b)

Figure 5.3: a) Full and b) partial power spectrum of monoprop chamber pressure data shown boxed in Figure 5.2a).

5.2.3 Calculations using Measured Data

The temperature and pressure data collected during the steady state portions of

both the monoprop and biprop sections of each test are averaged using the data reduction

file. These average values of pressures and temperatures are used in subsequent

calculations. The oxidizer flow rate is calculated using equation 4.4 with the averaged

oxidizer cavitating venturi inlet pressure from each mode of test operation. The density

of the hydrogen peroxide is calculated using equation 4.8 based on the measured

concentration and the averaged temperature of the hydrogen peroxide in the tank. The

pressure drop across the catalyst bed, ?pcb, is calculated from the following equation:

outcbincbcb ppp __ −=∆ , (5.4)

The fuel flow rate is also calculated from equation 4.4 using the fuel cavitating

venturi inlet pressure during biprop mode. The density of JP-8 fuel varies with

temperature, but the fuel temperature was never measured during testing. Since the fuel

was stored at room temperature the density was calculated assuming an average

temperature of 72.5°F, the density varies by less than 1.0% with a ±5°F variance from

this average value. The mixture ratio, O/F, during biprop mode is calculated as follows:

Page 104: Thesis

89

f

ox

mm

FO&&

= , (5.5)

The equivalence ratio was calculated using equation 3.6. The pressure drop

across the fuel injector, ?pf, is calculated from:

1_ cinjff ppp −=∆ , (5.6)

The discharge coefficient of the fuel orifices cannot be directly calculated from

the pressure drop found in equation 5.6. This is due to the fact that pf_inj is measured at

the entrance to the injector manifold not at the entrance to the orifices. As a result, the

calculated pressure drop contains the pressure drop across the feed line and manifold as

well as the orifices. Calculating a discharge coefficient using the pressure drop found

from equation 5.6 would actua lly be the coefficient of the entire injector not just the

orifice. The static pressure behind the rearward step, pc1, is used in this calculation

because the velocity of the gas at that point is closest to what the velocity would be in the

gas port. The velocity at the end of the chamber is much lower than in the gas port and as

such using pc2 for this calculation would give erroneous results. Since the diameter at the

exit of the PCI catalyst bed is roughly equivalent to the diameter of the gas port the

catalyst bed exit pressure, pcb_out, could be used in place of pc1. The exit diameter of the

GK catalyst bed, on the other hand, is significantly larger than that of the gas port and, as

a result, the velocity is much lower there. This prevents the use of pcb_out in equation 5.6

when using the GK catalyst bed. The velocity of the fuel exiting the orifices was

calculated using equations 2.3 and 2.4 assuming a CD of 0.8.

Using the calculated mass flow rates the characteristic velocity can be calculated

from the following equation which is a modified version of equation 2.11:

tot

cttotc

m

gApC

&_2=∗ , (5.7)

Due to the small contraction ratios used during testing the Mach number at the

end of the chamber was too significant to ignore. As a result, the static pressure

measured at the end of the chamber, pc2, was adjusted to a total pressure, pc2_tot, and used

in equation 5.7. Table 5.1 below lists the Mach number and total-to-static pressure ratio

at the end of the chamber based on contraction ratio for both monoprop and biprop

Page 105: Thesis

90

conditions. The Mach number at the end of the chamber, Mc, was calculated iteratively

from the following equation:

( )121

2

21

11

21 −+

++

γ

γγ c

c

MM

CR , (5.8)

For the biprop case the specific heat ratio, ?, varies widely with mixture ratio and

chamber pressure. A value of 1.22 was used as an average based on the tested conditions.

In monoprop mode and average specific heat ratio was also used and set at 1.265, which

corresponds to that of decomposed 90% H2O2, since the Mach number and pressure ratio

vary by less than 0.1% over the range of hydrogen peroxide concentrations used during

testing. As Table 5.1 shows the Mach number varies by less than 0.1% between

monoprop and biprop specific heat ratios as well. The total pressure ratio is calculated

from the isentropic flow relation as follows:

12

2

_2

21

1−

+=γ

γγ

cc

totc Mp

p, (5.9)

For the monoprop case, as discussed in Chapter 2, the total mass flow rate from

equation 5.7 is equal to the oxidizer flow rate only. In biprop mode the total mass flow

rate is equal to the sum of the fuel and oxidizer flow rates. The efficiency of

decomposition in monoprop mode and the efficiency of combustion in biprop mode were

calculated using the following equation for C* efficiency, *Cη :

=th

C CC

*η , (5.10)

The theoretical C*, ∗thC , in monoprop mode is dependent on the concentration of

the hydrogen peroxide. Table 5.2 lists theoretical C* values along with other

decomposed gas properties based on concentration. For biprop mode the theoretical C* is

strongly dependent on the mixture ratio and hydrogen peroxide concentration and weakly

on the chamber pressure. The variation in theoretical C* with equivalence ratio for the

tested concentrations of H2O2 assuming a chamber pressure of 500 psia is shown in

Figure 5.4.

Page 106: Thesis

91

Table 5.1: Variation in chamber gas properties based on chamber contraction ratio. CR (--) 3.0 5.0 6.5

Conditions Monoprop Biprop Monoprop Biprop Monoprop Biprop ? (--) 1.265 1.220 1.265 1.220 1.265 1.220

Mc (--) 0.200 0.201 0.118 0.119 0.091 0.091

2

_2

c

totc

p

p 1.026 1.025 1.009 1.009 1.005 1.005

The trajectory of the liquid fuel jet is most important at the instant of fuel

injection prior to combustion. For this reason, the gas flow properties and momentum

ratio were calculated only for the monoprop chamber conditions. The decomposed gas

Mach number, velocity, and density were calculated assuming both compressible and

incompressible flow. In both cases the total pressure of the decomposed gas was

assumed to be equal to the monoprop total chamber pressure, pc2_tot_mono, and the total

temperature of the decomposed gas was calculated using the following equation:

thtoxmonoCtox TT _2

_*η= , (5.11)

Here the monoprop decomposition efficiency is denoted by ?C*_mono, and the

theoretical total temperature of the decomposed gas is represented by Ttox_th. The

theoretical decomposed gas temperature is shown in Table 5.2 as a function of

concentration. Equation 5.11 was due to unreliable decomposed gas temperature

measurement. For the case of compressible flow equations 2.7, 2.9, and 2.10 and the

method described in Chapter 2 were used to calculate gas Mach number, velocity, and

density. For the case of incompressible flow these values were calculated from modified

forms of equations 2.7, 2.9, and 2.10 as follows:

γox

toxu

oxtox

oxox MW

TRAp

mM

&= , (5.12)

ox

toxuoxoxoxox MW

TRMaMV

γ== , (5.13)

toxu

oxtoxox TR

MWp=ρ , (5.14)

Page 107: Thesis

92

The value of the specific heat ratio and molecular weight varies with the

concentration of hydrogen peroxide as shown in Table 5.2. The momentum ratio of the

liquid fuel to the decomposed hydrogen peroxide was calculated using equation 2.24 and

the shear layer characteristic time was calculated using equation 2.12.

Table 5.2: Decomposition properties of hydrogen peroxide as a function of concentration.

W (%) 85% 87.5% 90% 94% 98% Ttox_th (°F) 1173 1283 1393 1570 1746 C*th (ft/s) 2904 2996 3083 3217 3343

MWox (lbm/lbm-mol) 21.83 21.97 22.11 22.33 22.56 ? (--) 1.274 1.269 1.265 1.258 1.251

Mass Fraction O2 0.400 0.412 0.423 0.442 0.461 Mass Fraction H2O 0.600 0.588 0.577 0.558 0.539

O/Fs (--) 8.49 8.25 8.02 7.68 7.36

3500

3750

4000

4250

4500

4750

5000

5250

5500

0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0

Equivalence Ratio, φ (--)

Ch

arac

teri

stic

Vel

oci

ty, C

* (f

t/s)

98% H2O294% H2O290% H2O287.5% H2O285% H2O2

Figure 5.4: Variation in C* with φ for H2O2 combusting with JP-8.

Page 108: Thesis

93

5.3 Uncertainty Analysis

An analysis was performed to determine the relative uncertainty in all calculated

parameters and the route by which error propagated through each calculation. The

method of uncertainty analysis presented in Fox and McDonald61 was used in this

autoignition study. The uncertainty analysis is based around the relative uncertainty, uxi,

of a measured quantity, xi, and is defined as the variation or random error in the value, dx i,

divided by the measured value itself. The variation in any measurement is the sum of

fixed and random error. Fixed error is repeatable and can be eliminated through proper

calibration of a measuring device. In this analysis the fixed error is assumed to be

negligible. The goal of any uncertainty analysis is to estimate the random error, which is

nonrepeatable, in each calculated value and to determine how random error propagates

through to each calculated value. Typically the random error in any measurement is

estimated as plus or minus one-half of the smallest measuring increment of the device.

For instance, the accuracy of a 3000 psia pressure transducer is stated at ±7.5 psia or

0.25% of full scale. This correlates to a random error estimate of ±3.25 psia or 0.125%

of full scale. In most cases sound engineering judgment based on experience is used to

estimate random error. Generally, a conservative approach is taken in order to avoid

underestimation of random error.

The relative uncertainty of a calculated variable, R, is found from the following

equation assuming R is a function of x1, x2, … xn:61

21

22

22

2

2

11

1 ...

∂∂

++

∂∂

+

∂∂

±= xnn

nxxR u

xR

Rx

uxR

Rx

uxR

Rx

u , (5.15)

This guideline for computing relative uncertainty was applied to all the calculated

values used in this study. A summary of the resulting equations for uncertainty of each

calculated variable is shown in Appendix E. As an example, the uncertainty in the

characteristic velocity can be determined as follows:

( ) ( ) ( )[ ] 21

222_2* 2 totmDthtotpcC uuuu &++±= , (5.16)

Page 109: Thesis

94

The estimated random error for each measured test parameter is shown in Table

5.3. All pressure measurements were assumed to have an error equal to the accuracy of

the pressure transducer, which may be a slight overestimation. Also, the error in the total

pressure ratio used to correct the static chamber pressure measurement was assumed to be

negligible since it does not vary significantly with the specific heat ratio. The error in the

measurement of the throat diameter of each cavitating venturi was set equal to zero.

Instead this error was lumped into the estimation of the error in the discharge coefficient.

This was also done in estimating the error of the fuel injector orifice diameter and

discharge coefficient. All other diameter and length measurements were assumed to have

an error of ±0.005 inches, which is based on machining tolerances and wear and tear on

the parts. Since the temperature of the fuel was not measured it was assumed to have a

fairly large error of ±5.0°F. The oxidizer temperature was measured by a thermocouple

and the error was likely overestimated at ±2.0°F. The measurement of the refractive

index of hydrogen peroxide was accurate to within ±0.0005, neglecting human error, and

puts the error in the measurement of concentration at approximately ±0.1%.

The error in the theoretical decomposition temperature and monoprop C* were

estimated very conservatively. The values of the decomposition temperature and C*

were calculated using a thermochemistry code assuming an exact hydrogen peroxide

concentration, pressure, and liquid temperature. In reality these theoretical values vary

based on the operating conditions of each test and for this reason the error in

decomposition temperature was estimated at ±10.0 °F and the error in C* at ±10.0 ft/s. A

similar logic was applied to estimating the error in the biprop theoretical C*. The

difference here is that the theoretical C* for biprop mode varies significantly with

equivalence ratio especially over the range of values used during autoignition testing,

shown boxed in Figure 5.4. The slope of the theoretical C* curves is very steep in this

region of equivalence ratios and small deviations in equivalence ratio have a large effect

on C*. In addition, small errors in hydrogen peroxide concentration and chamber

pressure can also result in large changes in C*. For these reasons the error in theoretical

C* for the biprop case was estimated at ±20.0 ft/s.

Page 110: Thesis

95

Table 5.3: Estimated random error for all measured variables.

Measurement Description Estimated Random Error Pressure ± 7.5 psia

Chamber Diameter, Chamber Throat Diameter,

Oxidizer Port Diameter, Rearward Step Height

± 0.005 in.

Oxidizer Temperature ± 2.0 °F Hydrogen Peroxide

Concentration ± 0.1 %

Theoretical Decomposition Temperature

±10.0 °F

Theoretical Monoprop C* ± 10.0 ft/s Theoretical Biprop C* ± 20.0 ft/s All Cavitating Venturi Discharge Coefficients ± 0.007

All Cavitating Venturi Throat Diameters

± 0.0 in.

Fuel Injector Orifice Discharge Coefficient ± 0.05

Fuel Injector Orifice Diameter

± 0.0 in.

Fuel Temperature ± 5.0 °F

Page 111: Thesis

96

5.4 Data Summary

A total of 24 tests were conducted to investigate the relative affects of

equivalence ratio, decomposed gas temperature, and contraction ratio on the autoignition

of JP-8 in decomposed hydrogen peroxide. Hydrogen peroxide concentration was varied

from 85 to 98%, equivalence ratios ranged from 1.4 to 4.0, and contraction ratio was set

to values of 3.0, 5.0 and 6.5. Autoignition was classified into three regimes: strong, weak,

and no ignition based on visual observations from real time and slow-motion video and

recorded chamber pressure data. Tables of measured and calculated test data are

contained in Appendix F.

5.4.1 Strong Autoignition

Tests which resulted in strong autoignition were characterized by bright, stable

red-orange flames and calculated bipropellant C* efficiencies of over 90%. The delay

between the initiation of fuel injection and autoignition was typically very small. A plot

of chamber pressure, pc2, and fuel injector pressure, pfu_inj, from a strong autoignition test

running 90% H2O2 through the GK catalyst bed at an equivalence ratio of 1.59 and a

contraction ratio of 3.0 is shown in Figure 5.5a). As the figure shows the chamber

pressure rises sharply, indicating autoignition, approximately 0.1 seconds following the

rise in the fuel injector pressure. This most likely corresponds to the time required to fill

the fuel manifold as the pressure drop across the injector reaches approximately 100 psid

at this instant as well. The small hump in the fuel injector pressure prior to the initiation

of fuel flow is a result of mechanical vibration caused when the fuel main valve is opened.

There is a slight oscillation in the chamber pressure following autoignition before it

steadies out. The chamber pressure for a strong autoignition test typically increases by

approximately 100% from monoprop to biprop modes. For this particular test the

monoprop decomposition efficiency was approximately 97% while the biprop C*

efficiency was about 94%.

Page 112: Thesis

97

A frame-by-frame view of the startup of the biprop portion of this test is shown in

Figure 5.6. Clear decomposed hydrogen peroxide gas is being exhausted from the

engine in frames a) and b) of Figure 5.6. The decomposed gas contacts the water

cooling the deflector plate and vaporizes some of it to produce the white cloud of vapor at

the bottom left of frames a) and b). In frame c) the white cloud of vapor seems to grow a

little bit denser possibly indicating the presence of fuel vapor in the exhaust. In the

following frame, frame d), a bright red-orange flame appears indicating that autoignition

has occurred. Following this in frames e) and f) the flame continues to burn with the

same intensity indicating that it is quite stable. The sequence of frames pictured in

Figure 5.6 agree very well with the conclusions drawn from the pressure data plotted in

Figure 5.5a). In fact, the 0.1 second delay between the fuel injector pressure rise and the

chamber pressure rise make present itself as the increase in vapor density seen in frame c)

in Figure 5.6 prior to autoignition.

A similar result was seen in the pressure data and frame-by-frame representation

of the startup transient running the PCI catalyst bed at nearly identical conditions. This

test was run using 90% H2O2 at a contraction ratio of 3.0 and a nearly identical

equivalence ratio of 1.58. Monoprop decomposition efficiency was calculated at 94%

while the biprop C* efficiency was approximately 96%. The pressure data plotted in

Figure 5.5b) for this test is almost identical to that of the GK bed, shown in Figure 5.5a).

The main differences between the two are that chamber pressure rises slightly earlier

following the rise in fuel injector pressure than for the GK bed and there are no large

chamber pressure oscillations present following autoignition in the PCI case. The

chamber pressure in the PCI takes about 0.1 seconds to reach the steady-state value while

for the GK case it is only about 0.05 seconds.

The frame-by-frame representation of the startup of the biprop portion of this test

is shown in Figure 5.7. A similar white water vapor cloud is seen during the monoprop

phase in frames a) and b) of Figure 5.7 and it also seems to thicken in frame c) possibly

indicating fuel injection. The flame appears in frame d) indicating autoignition and

seems to have more predominantly orange color than frame d) in Figure 5.6. This could

be due to the difference in ambient light settings between the two tests. The flame

Page 113: Thesis

98

remains stable and its color and luminosity are nearly identical in frames e) and f). Again

the frame-by-frame view of this test agrees very well with the pressure data in Figure

5.5b). The orange flame seen in frame c) of Figure 5.7 may be a result of the long delay

in reaching the steady state chamber pressure during biprop mode.

a) b)

Figure 5.5: Plots of pfu_inj and pc2 at the point of injection for a strong autoignition test using: a) the GK catalyst bed using 90% H2O2 at φ = 1.59 and CR = 3.0 and b) the PCI

catalyst bed using 90% H2O2 at φ = 1.58 and CR = 3.0.

Page 114: Thesis

99

a) b)

c) d)

e) f) Figure 5.6: Frame-by-frame view of a strong autoignition test using GK catalyst bed.

Test conditions: 90% H2O2, φ = 1.59, CR = 3.0.

Page 115: Thesis

100

a) b)

c) d)

e) f) Figure 5.7: Frame-by-frame view of a strong autoignition test using PCI catalyst bed.

Test conditions: 90% H2O2, φ = 1.58, CR = 3.0.

Page 116: Thesis

101

5.4.2 Weak Autoignition

Tests classified as resulting in weak autoignition typically produced flames that

were very unstable and varied in color, from red-orange to green, and intensity. These

tests were characterized by large chamber pressure instabilities, in most cases, and

resulted in biprop C* efficiencies ranging from 60 to 90%. A typical plot of fuel injector

and chamber pressure at the start of the biprop portion of a weak autoignition test is

shown in Figure 5.8a). As the plot shows, the chamber pressure data for this test

contained significant pressure oscillations, the amplitude of which is nearly equivalent to

the average chamber pressure. The data was fed through a low-pass Butterworth filter

with a cut-off frequency of 15 Hz to eliminate these fluctuations. The filtered data is

plotted in Figure 5.8b). This test was run using the GK catalyst bed using 87.5% H2O2 at

a contraction ratio of 3.0 and an equivalence ratio of 1.62. The delay between the time in

which the fuel was turned on and the sudden rise in chamber pressure was approximately

equivalent if not shorter than that presented in Figure 5.5a) and b). However, the rise is

not as sharp and does not reach as high of a magnitude as in the strong ignition cases.

The monoprop decomposition efficiency for this test was about 93% and the biprop C*

efficiency was approximately 69%.

a) b)

Figure 5.8: a) Unfiltered and b) filtered plots of pfu_inj and pc2 at the point of injection for a weak autoignition test using the GK catalyst bed. Test conditions: 87.5% H2O2,

φ = 1.62, and CR = 3.0.

Page 117: Thesis

102

a) b)

c) d)

e) f)

Figure 5.9: Frame-by-frame view of a weak autoignition test using GK catalyst bed. Test conditions: 87.5% H2O2, φ = 1.62, CR = 3.0.

Page 118: Thesis

103

g) h)

i) j)

k) l) Figure 1.5 (cont.): Frame-by-frame view of a weak autoignition test using GK

catalyst bed. Test conditions: 87.5% H2O2, φ = 1.62, CR = 3.0.

Page 119: Thesis

104

A frame-by-frame view of this test shows how the instability affects the

appearance of the flame exiting the combustor. Frames a) through c) in Figure 5.5 show

the clear decomposed peroxide steam and increase in vapor density prior to autoignition

which are typical based on the tests discussed so far. In Frame d) and e) a faint

red-orange flame appears exiting the nozzle. This flame is much smaller in size and less

intense than those seen in the strong autoignition cases. In frames f) through i) an intense

green flame appears, which is similar in size to the strong ignition cases. The flame

reverts to a small red-orange flame in frames j) and k) and then back to a green flame

again in frame l). The unstable, oscillatory nature in the color and intensity of the flame

exiting the combustor agrees with the large oscillations in the measured chamber pressure

data. Each color flame last for approximately three frames or about 0.1 seconds,

assuming 30 frames per second. The peaks and valleys in the chamber pressure appear at

a much shorter interval of about 0.02 seconds based on Figure 5.8a). The green flame

indicates that it is a cooler flame, which may correspond to the valleys in the chamber

pressure data. The smaller, red-orange are hotter flames and could correspond to the

peaks in the pressure data. If this is the case it may mean that the camera did not

completely record the changes in the flame since many of them could have occurred

between frames.

5.4.3 No Autoignition

During tests in which no autoignition occurred a thick cloud of vapor was

typically seen issuing from the nozzle following fuel injection. In some cases, the

difference in appearance between the exhaust during monoprop mode and following fuel

injection was almost unnoticeable. In general, flames were not seen exiting the engine

during tests classified as no ignition. However, sparks of flame were sometimes seen at

shutdown. A plot of fuel injector inlet pressure and chamber pressure for a no

autoignition test is shown in Figure 5.10. This test was run using 94% H2O2 at a

contraction ratio of 3.0 and an equivalence ratio of 2.15. Following fuel injection the

chamber pressure rises by only about 10 psi or 6% of the monoprop chamber pressure.

Page 120: Thesis

105

This small increase in pressure is a result of the fuel being vaporized in the chamber. A

sharp increase in pressure does not occur following fuel injection like that of the strong

and weak autoignition cases, which suggests that the decomposed gas had only enough

energy to vaporize the fuel and not enough to ignite the resulting mixture. The monoprop

decomposition efficiency for this test was approximately 82% while the biprop C*

efficiency was about 47%. Biprop C* efficiencies for all no ignition cases were less than

65%.

A frame-by-frame view of this test is shown in Figure 5.11. Again, frames a)

through c) of Figure 5.11 are similar to the corresponding frames for the strong and weak

autoignition cases. However, following fuel injection a flame is not visible exiting the

engine as shown in frames d) through f). In fact, the difference in the plume following

fuel injection is so small that the only way to tell fuel is actually in the plume is by

difference in the density of the water vapor cloud surrounding the deflector plate between

frames b) and d). This result agrees with the conclusions drawn based on the small

increase in the measured chamber pressure.

Figure 5.10: Plot of pfu_inj and pc2 at the point of injection for a test using PCI catalyst bed.

Test conditions: 94% H2O2, φ = 2.15, and CR = 3.0.

Page 121: Thesis

106

a) b)

c) d)

e) f)

Figure 5.11: Frame-by-frame view of a test with no autoignition using PCI catalyst bed. Test conditions: 94% H2O2, φ = 2.15, CR = 3.0.

Page 122: Thesis

107

5.4.4 Trends in Autoignition Data

One of the goals of this study was to determine the effect of equivalence ratio and

decomposition temperature on autoignition of JP-8. A total of 20 tests were conducted at

a contraction ratio of 3.0 to determine autoignition boundary in terms of equivalence ratio

at a constant H2O2 concentration. Hydrogen peroxide concentrations of 85, 87.5, 90, 94,

and 98% were used to vary the decomposition gas temperature. The GK catalyst bed was

used for concentrations of 90% and below while the PCI catalyst bed was used for

concentrations 90% and above. The data resulting from these tests is shown in Error!

Reference source not found.. The uncertainties in equivalence ratio and hydrogen

peroxide concentration were less than 1.5% for all test conditions.

As Figure 5.12 shows, the strong autoignition boundary in terms of equivalence

ratio widens with increasing hydrogen peroxide concentration. It stretches from a value

of less than 1.37 at a concentration of 85% to 2.06 at a concentration of 98%. However,

the autoignition boundary is not necessarily clearly defined especially at concentrations

of 85, 87.5, and 90%. At a concentration of 85% a strong or weak autoignition condition

was not found even at an equivalence ratio as low as 1.37. A test was not conducted at a

lower equivalence ratio because the oxidizer flow rate, 2.33 lbm/s, needed to reach a φ of

1.37 was close to the highest recommended flow rate for the GK catalyst bed. At

concentrations of 87.5 and 90% weak autoignition was observed over a wide range of

equivalence ratios before reaching a no autoignition condition. For instance, at 90%

H2O2 strong autoignition was observed at the baseline equivalence ratio of 1.59 using the

GK catalyst bed. However, a no ignition condition was not found up to an equivalence

ratio of 2.19. Similarly, at a concentration of 87.5% using the GK catalyst bed strong

autoignition was observed at an equivalence ratio of 1.40, but a no autoignition condition

was not achieved until a φ of 1.86.

The PCI catalyst bed was tested at three test conditions that were nearly identical

to those run with GK catalyst bed. At the baseline conditions the PCI bed produced a

strong autoignition result in agreement with that of the GK bed. Running the PCI catalyst

bed at equivalence ratios of 2.01 and 2.23 resulted in no autoignition rather than weak

Page 123: Thesis

108

autoignition in both cases. In fact, the PCI catalyst bed did not produce weak

autoignition results at any of the tested conditions. At concentrations of 94 and 98%

strong autoignition was observed at equivalence ratios of 1.70 and 2.06 respectively.

1.0

1.2

1.4

1.6

1.8

2.0

2.2

2.4

2.6

2.8

3.0

84.0 86.0 88.0 90.0 92.0 94.0 96.0 98.0 100.0

H2O2 Concentration, W (%)

Equ

ival

ence

Rat

io, φ

(--)

Strong, GKWeak, GKNo Ign., GKStrong, PCIWeak, PCINo Ign., PCI

Figure 5.12: Autoignition of JP-8 as a function of φ and H2O2 concentration at CR = 3.0.

Another goal of this study was to determine the effect of gas velocity on the

autoignition of JP-8 in decomposed peroxide. Four additional tests were performed at a

contraction ratio of 5.0 and a constant concentration of 85% using the GK catalyst bed to

determine the autoignition boundary in terms of equivalence ratio. The data resulting

from these tests along with those done at a contraction ratio of 3.0 with 85% H2O2 are

shown in Figure 5.13. Strong autoignition was achieved at an equivalence ratio of 1.36,

a point at which no autoignition was achieved at a contraction ratio of 3.0. Following this

test the equivalence ratio was increased in an attempt to find a point a no autoignition

point. Weak autoignition was observed at equivalence ratios of 1.70 and 2.31, and a no

autoignition point was found at a φ of 2.73. As with the lower contraction ratio the GK

catalyst bed produced two weak autoignition points at a contraction ratio of 5.0.

Page 124: Thesis

109

1.0

1.2

1.4

1.6

1.8

2.0

2.2

2.4

2.6

2.8

3.0

2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0

Contraction Ratio, CR (--)

Eq

uiv

alen

ce R

atio

, φ (

--)

Strong, GKWeak, GKNo Ign., GK

Figure 5.13: Autoignition of JP-8 as a function of equivalence ratio and contraction ratio

at a constant H2O2 concentration of 85%.

5.4.5 Temperature Data

As described in Chapter 4, the temperature of the decomposed hydrogen

peroxide was measured using an exposed thermocouple. In addition, four grounded

thermocouples were located in the chamber to measure the temperature at the combustor

wall. During testing it was found that the thermocouples did not produce reliable

temperature readings. The response time of the exposed thermocouple was too slow to

achieve a steady state decomposition temperature reading in the one-half second of

monoprop lead in each autoignition test. Two plots of the decomposition temperature

measured during two separate tests are shown in Figure 5.14a) and b). The plot in

Figure 5.14a) shows that the measured decomposition temperature is still increasing at

the point at which autoignition occurs. Soon after autoignition there is a sudden increase

in the measured temperature from the previously measured value. This may indicate that

there is some heat soak-back to the thermocouple following autoignition. In Figure

5.14b) the measured decomposition temperature does not reach a steady value until the

end of the biprop section of the test. In both cases the monoprop chamber pressure is

Page 125: Thesis

110

very steady which indicates very steady decomposition. If the decomposition is steady

then the temperature of the gases should remain steady as well. Since this is not the case

it leads one to believe that the thermocouples are not responding quickly enough. This

casts some doubt as to the accuracy of the readings produced by the thermocouples at the

point of autoignition as well. For this reason, equation 5.11 was used to calculate the

decomposition temperature for each test. The only problem with this method is that the

uncertainty in the calculated temperature is quite large, up to 20.0%, due to the large error

in decomposition efficiency. The uncertainty in decomposition efficiency is discussed in

the following section.

The four grounded thermocouples in the combustion chamber behaved in a

similar manner. The response time of these thermocouples was slow as well and seldom

reached a steady state value at any time during the test. It was also unclear, due to the

flush mounted position in the wall, what the physical meaning was of each temperature

reading. It may have been reading the wall temperature, the temperature in the boundary

layer, or the free stream gas temperature. The thermocouple located behind the

rearward-facing step, Tc1, was somewhat useful in providing qualitative evidence of

autoignition. The temperature measured behind the step was lower in no autoignition and

weak autoignition cases than in strong autoignition. However, this temperature

measurement alone was not sufficient to classify the type of autoignition that occurred in

each test.

a) b)

Figure 5.14: Measured Td over the course of tests run at a) 90% H2O2 at φ = 1.59 and CR = 3.0 with GK bed and b) 98% H2O2 at φ = 1.84 and CR = 3.0 with PCI bed.

Page 126: Thesis

111

5.5 Catalyst Bed Performance

As described in the previous section there was a significant difference in the

performance of each catalyst bed during autoignition testing. This was especially evident

when the catalyst beds were run at nearly identical operating conditions. Remember,

from Chapter 4, that each catalyst bed is of a different internal design and should be

expected to behave differently. Each catalyst bed was subjected to similar operating

conditions in terms of concentration, mass flow rate, and monoprop chamber pressure

using 90% H2O2 and a contraction ratio of 3.0. The pressure drop and decomposition

efficiency of each catalyst bed as a function of mass flow rate at these conditions are

shown in Figure 5.15 and Figure 5.16. From Figure 5.15 it appears that the GK catalyst

bed produces higher average decomposition efficiency for each mass flow rate condition

than the PCI catalyst bed. However, the error bars shown in Figure 5.15 indicate that

there is substantial uncertainty in the calculated values of the decomposition efficiency.

In fact, the relative uncertainty in the decomposition efficiency ranges from 6.2 to 8.5%

at these conditions. The main reason for this large error is due to the fact that 3000 psia

transducers, with an estimated random error of 7.5 psia, are being used to measure the

monoprop chamber pressure which varies from 74 to 115 psia. Thus the relative

uncertainty in the pressure measurement itself is 6.5% to 8.8%.

Regardless of the error in the calculation, if the decomposition efficiency of the

PCI catalyst bed is lower than that of the GK bed then the decomposition temperature is

lower as well. Since autoignition is strongly dependent on the gas temperature this could

be one explanation for the inconsistent autoignition results. Using equation 5.11 the

2.5% difference in average decomposition efficiency at a flow rate of 1.80 lbm/s equates

to a 65°F difference in decomposition temperature. This is very significant seeing that

the theoretical decomposition temperature for 87.5 and 90% differs by only 110°F. At a

flow rate of approximately 1.60 lbm/s the decomposition efficiency differs by almost

10% which is a temperature difference of almost 200°F according to equation 5.11. This

indicates that it is not sufficient to plot these results in terms of concentration. Plotting in

terms of decomposition temperature may be more appropriate.

Page 127: Thesis

112

The pressure drop across the GK catalyst bed was consistently less than that of the

PCI bed under similar flow rate and chamber pressure conditions according to Figure

5.16. The uncertainty in the calculated pressure drop is less than 5% for each test

condition shown in Figure 5.16. The pressure drop across a catalyst bed does not

directly affect autoignition, but it does affect the range of operation of a cavitating venturi

and resistance of the catalyst bed to instability. Since the PCI bed has a higher pressure

drop it would seem to have an advantage in stability while the GK bed would seem to

have the advantage in operating range due to its lower pressure drop.

70.0%

75.0%

80.0%

85.0%

90.0%

95.0%

100.0%

105.0%

110.0%

1.40 1.60 1.80 2.00 2.20 2.40

H2O2 Mass Flow Rate, m H2O2 (lbm/s)

C*

Eff

icie

ncy

, ηC

* (%

)

GK, Monoprop

PCI, Monoprop

Figure 5.15: Comparison in decomposition efficiency produced by GK and PCI catalyst

beds using 90% H2O2 at approximately equivalent monoprop operating conditions.

Page 128: Thesis

113

0.0

100.0

200.0

300.0

400.0

500.0

600.0

1.40 1.60 1.80 2.00 2.20 2.40

H2O2 Mass Flow Rate, m H2O2 (lbm/s)

Cat

alys

t B

ed P

ress

ure

Dro

p, ∆

p cb

(p

sid

)

GK, Monoprop

PCI, Monoprop

Figure 5.16: Comparison in pressure drop across GK and PCI catalyst beds using 90%

H2O2 at approximately equivalent monoprop operating conditions.

During testing the GK catalyst bed consistently produced large pressure

oscillations in the chamber even in monoprop mode. In most cases, this instability could

be heard from the control room during the test. The effect was observed mainly during

tests classified as weak or no autoignition, but was sometimes present in strong

autoignition cases as well. A noise parameter, ?, was defined to quantify these pressure

oscillations. The standard deviation, s , in the chamber pressure, pc2, was calculated using

the same range of data used to calculate the average value of pc2 in each mode of

operation. Using the standard deviation the noise parameter is defined as:

2cpσ

ς = , (5.17)

The noise parameter is an indication of the relative magnitude of the pressure

oscillations in the chamber and can be expressed as a percentage. In addition, the

dominant frequency, fd, in the pressure data over this same averaging range was

determined using an FFT. The noise parameter is plotted against the dominant frequency

for the monoprop and biprop modes of each test in Figure 5.17 and Figure 5.18

respectively.

Page 129: Thesis

114

0.0

5.0

10.0

15.0

20.0

25.0

0.0 100.0 200.0 300.0 400.0 500.0

Dominant Frequency, f d (Hz)

Noi

se P

aram

eter

, ζ (

%)

GK, StrongGK, WeakGK, No Ign.PCI, StrongPCI, WeakPCI, No Ign.

Figure 5.17: Plot of the chamber pressure noise parameter against the dominant

frequency during the monoprop mode each autoignition test.

0.0

5.0

10.0

15.0

20.0

25.0

30.0

35.0

0.0 100.0 200.0 300.0 400.0 500.0Dominant Frequency, f d (Hz)

Noi

se P

aram

eter

, ζ (%

)

GK, StrongGK, WeakGK, No Ign.PCI, StrongPCI, WeakPCI, No Ign.

Figure 5.18: Plot of the chamber pressure noise parameter against the dominant

frequency during the biprop mode each autoignition test.

Page 130: Thesis

115

The type of autoignition that occurred for each test is also shown in Figure 5.17

and Figure 5.18. As the plots show the noise parameter calculated for eight out of the

fifteen tests using the GK catalyst bed exceeded 5% during monoprop mode. In some of

the tests the noise parameter was calculated as high as 20%. This indicates that the GK

catalyst bed is susceptible to instability. The instability may result from the low

monoprop chamber pressures used during testing. In biprop mode nine out of the fifteen

tests which used the GK bed had a noise parameter above 5% some reaching nearly 35%.

Many of these were classified as no or weak autoignition and some were not present at a

magnitude of greater than 5% during monoprop mode. This may suggest that the

addition of unburned fuel vapor to the chamber can enhance or trigger instability in the

engine. None of the tests run using the PCI bed had a noise parameter above 5%, in

monoprop or biprop mode. One must keep in mind that the standard deviation is only an

average deviation from the mean value, and the amplitude of the pressure oscillations

may be much greater than the noise parameter indicates in some cases. The dominant

frequency of each test with a noise parameter above 5% is between 27 and 47 Hz. It is

unclear as to why the large pressure oscillations exist within this particular range of

frequencies. However, all of the tests which were classified as weak autoignition have a

dominant frequency within this band during biprop mode even those with a noise

parameter of less than 5%.

All of this data suggests that the existence of a fluctuating pressure in the chamber

influenced the autoignition process. In more general terms this may mean that the

pressure of the decomposed gas very strongly influences autoignition. For an unstable

weak autoignition case, as discussed previously, the high end of the instability may

trigger an autoignition event. The heat produced by the combustion reaction may be

enough to sustain it as the pressure reaches the low point of the oscillation. This may be

why the color of the flame fluctuates during a weak autoignition test. When the pressure

is high the energy of the decomposed gas flow sustains the flame producing a red-orange

color. Whereas, at low pressure, the energy produced by the hot flame must sustain the

reaction and produces a cooler, green or light orange flame.

Page 131: Thesis

116

It is unclear why this instability was present for one catalyst bed and not the other,

but it may have to do with the internal design of each bed. However, it is interesting to

note the difference in the exit geometry of each catalyst bed. The bed with the instability

had an exit diameter 1.2 inches greater than that of the oxidizer port so it was contracted

down at an angle of 40° to match the port using an adapter piece, see Figure 4.5. That

means that the decomposed gas would exit the catalyst bed at a low velocity then be

accelerated through the contraction and then expanded again downstream of the

rearward-facing step. The exit diameter of the other catalyst bed was roughly the same

diameter as the oxidizer port and did not require a contraction. Therefore, the

decomposed gas was not accelerated after it exited the catalyst bed. No conclusions can

be drawn from this discussion, but the difference in geometry downstream of the catalyst

bed may have played a role in producing instability.

5.6 DMAZ Fuel

A total of six tests were run to compare the autoignition characteristics of DMAZ

fuel to those of JP-8. Tests were conducted using the GK catalyst bed at hydrogen

peroxide concentrations of 85 and 90% at a constant contraction ratio of 3.0. The same

transverse injector was used for these tests due to the similarity in density between JP-8

and DMAZ. As a result a similar, constant fuel flow rate was used as well. Strong

autoignition was achieved at equivalence ratios ranging from 2.02 to 2.60 using 90%

H2O2. The highest equivalence ratio at which strong autoignition was achieved using

JP-8 was only about 1.60 with weak autoignition up to an equivalence ratio of about 2.20

using the GK catalyst bed. A weak or no autoignition point was not found at 90% H2O2

using DMAZ. At 85% H2O2 the equivalence ratio was varied between 1.81 and 2.70.

Weak autoignition was achieved at each of these points, whereas no autoignition was

achieved at all using JP-8 down to an equivalence ratio of 1.37. The DMAZ autoignition

data points are compared those of JP-8 using the GK catalyst bed in Figure 5.19. The

data suggests that DMAZ is much easier to autoignite than JP-8.

Page 132: Thesis

117

Since the stoichiometric mixture ratio of DMAZ and hydrogen peroxide is about

half that of JP-8 lower mixture ratios were required to achieve similar equivalence ratios.

The lower mixture ratios required a low oxidizer flow rate and thus created a monoprop

chamber pressure of less than 50 psia for all of the DMAZ autoignition tests. Thus, the

calculated uncertainty in parameters based on chamber pressure was extremely high. In

addition the noise parameter calculated for the DMAZ tests was very high as well

particularly during tests conducted using 85% H2O2. During these tests the noise

parameter was calculated to be above 20% during both monoprop and biprop mode. The

dominant frequency was between 27 and 32 Hz as well. This is most likely the reason

why weak autoignition was observed for DMAZ during all of the tests conducted with

85% H2O2.

1.0

1.2

1.4

1.6

1.8

2.0

2.2

2.4

2.6

2.8

3.0

84.0 85.0 86.0 87.0 88.0 89.0 90.0 91.0 92.0

H2O2 Concentration, W (%)

Eq

uiv

alen

ce R

atio

, φ (-

-)

Strong, JP-8Weak, JP-8No Ign.,JP-8Strong,DMAZWeak, DMAZNo Ign., DMAZ

Figure 5.19: Comparison of DMAZ autoignition points to those of JP-8 using the GK catalyst bed at a CR = 3.0.

Page 133: Thesis

118

5.7 Data Analysis

5.7.1 Pressure Effects

As discussed in the previous section, it is more appropriate to plot autoignition

data in terms of decomposition temperature rather than concentration. This is due to the

fact that the efficiency of decomposition dictates the gas temperature. A plot of

equivalence ratio against the calculated decomposition temperature is shown in Figure

5.20 for JP-8 a contraction ratio of 3.0. Note the large error associated with the

decomposition temperature. This plot indicates, as did Figure 5.12, that at a constant

equivalence ratio the likelihood of strong autoignition increases as the decomposition

temperature increases. Again, this data indicates that equivalence ratio plays an

important role in determining the conditions for autoignition. However, because of the

fact that the oxidizer flow rate was varied to change the equivalence ratio for each test the

monoprop chamber pressure changed for each test as well.

As was discussed in Chapter 3, past data on the autoignition of kerosene

indicated that pressure strongly influences ignition delay. In addition, the unstable

chamber pressures encountered during all of the weak autoignition tests also suggest that

pressure is an important factor. The measured average monoprop chamber pressure for

tests conducted at a contraction ratio of 3.0 are shown in Figure 5.21 as a function of

calculated decomposition temperature. The figure shows that as the monoprop chamber

pressure falls at a particular temperature so does the likelihood of achieving strong

autoignition. There is one strong autoignition point that does not follow this trend. This

particular test was run using 94% H2O2 and achieved a monoprop efficiency of about

82%, which pushed the calculated decomposition temperature down. This may indicate

that there is another factor involved or that the measurement of chamber pressure and

thus the calculated decomposition temperature could be in error. The weak autoignition

points are all close enough to the strong autoignition points that small oscillations in

chamber pressure could push them into the strong autoignition regime.

Page 134: Thesis

119

Based on this discussion it is not conclusive that equivalence ratio is the factor

which most effected autoignition under the tested conditions. The monoprop chamber

pressure seems to be a very influential factor as well. There also has been mixed results

regarding these two factors from past studies as well. Nearly all autoignition correlations

include the effect of pressure in some way or another, but many neglect the effect of

equivalence ratio.55 However, several studies have shown that equivalence ratio as well

as pressure influence autoignition.8,58 In dump combustors, it has been shown that

equivalence ratio is influential as well.41-43 It is suspected that both factors exert a

combined influence on the autoignition process, but from the present data it is very

difficult to separate the two.

1.0

1.2

1.4

1.6

1.8

2.0

2.2

2.4

2.6

400.0 600.0 800.0 1000.0 1200.0 1400.0 1600.0 1800.0 2000.0

Decomposition Temperature, T tox (oF)

Eq

uiv

alen

ce R

atio

, φ (-

-)

Strong, GKWeak, GKNo Ign., GKStrong, PCIWeak, PCINo Ign., PCI

Figure 5.20: Autoignition of JP-8 as a function of equivalence ratio and decomposition

temperature at a contraction ratio of 3.0.

Page 135: Thesis

120

60.0

70.0

80.0

90.0

100.0

110.0

120.0

130.0

400.0 600.0 800.0 1000.0 1200.0 1400.0 1600.0 1800.0

Decomposition Temperature, T tox (oF)

Mon

opro

p C

ham

ber

Pre

ssur

e, p

c_to

t_m

ono

(p

sia) Strong, GK

Weak, GKNo Ign., GKStrong, PCIWeak, PCINo Ign., PCI

Figure 5.21: Autoignition of JP-8 as a function of pc2_tot_mono and decomposition

temperature at a contraction ratio of 3.0.

5.7.2 Jet Trajectory

Chapter 2 discussed the effect of the fuel jet trajectory on the distribution and

atomization of the fuel in the gas port. The jet trajectory is dependent on chamber

pressure and, as a result, engine operating condition. Once autoignition occurs in the

chamber the pressure increases altering the trajectory, as described in Chapter 2. Prior

to autoignition, even when vaporized fuel is present in the chamber, the pressure is

roughly equal to the monoprop value. Therefore, when investigating the affects of jet

trajectory on autoignition it is beneficial to consider the trajectory based on monoprop

conditions. With this in mind, the fuel jet trajectory should be affected by changes in the

oxidizer mass flow rate along with chamber pressure. As the oxidizer flow rate is

lowered to increase equivalence ratio the oxidizer momentum will decrease. Assuming

the fuel momentum is constant, the fuel jet penetration should increase as a result. In

general, fuel distribution improves with increasing jet penetration and one would expect

Page 136: Thesis

121

the likelihood for strong autoignition to increase as well. However, this is not the case

from the results presented to this point.

At a contraction ratio of 3.0 the calculated momentum ratio, assuming monoprop

conditions, ranged from 4.2 to 8.1. Since the momentum ratio was calculated based on

the decomposition temperature its uncertainty reached 50% in some cases. Using the

calculated uncertainty values the momentum ratio ranged from 2.8 to 14.0. The

difference in trajectory resulting from this range of momentum ratios is shown in Figure

5.22. In the case of Q = 2.8 the fuel only penetrates about a quarter of the distance to the

centerline of the chamber prior to breakup. For Q = 14.0 the fuel penetrates slightly more

than halfway to the centerline of the chamber. Both of these trajectories achieved a no

autoignition result even though the penetration differed significantly. Since the rest of

the fuel trajectories at a contraction ratio of 3.0 are contained somewhere between those

shown in Figure 5.22 it suggests that the trajectory does not play a significant role in

autoignition at a constant contraction ratio. The Mach number of the decomposed gas

was approximately 0.45, with a maximum uncertainty of 15%, for all tests at this

contraction ratio.

When the contraction ratio was increased to 5.0 the calculated momentum ratio

ranged from approximately 7.1 to 18.1, or 5.4 to 25.7 based on uncertainty. These

trajectories are plotted in Figure 5.23. The Mach number of the decomposed gas during

these tests was approximately equal to 0.28 with a 15% uncertainty. The test run at a

minimum Q of 5.4, based on uncertainty, achieved strong autoignition. According to

Figure 5.23 the fuel jet penetrated only about one-third of the way to the centerline. For

the maximum Q case of 25.7 no autoignition was achieved and the fuel penetrated very

closely to the centerline prior to breakup. When the fuel penetrates this closely to the

centerline it is possible that fuel droplets from multiple jets could collide following

breakup forming larger drops that take longer to vaporize. Also, when the fuel is

centralized in the oxidizer port it may have difficultly making its may to the recirculation

zone behind the rearward-facing step. This may be reasons as to why autoignition did not

occur at this condition, but it is not known for certain. As with the low contraction ratio

there is no evidence to suggest that the fuel trajectory had an affect on autoignition.

Page 137: Thesis

122

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Axial Direction, x (in.)

Tra

nsv

erse

Dir

ecti

on

, y (i

n.)

Q = 2.8Q = 14.0Axial Fracture Point

Chamber WallsChamber Centerline

Figure 5.22: Trajectory variations in tests run at a contraction ratio of 3.0.

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Axial Direction, x (in.)

Tra

nsv

erse

Dir

ecti

on

, y (i

n.)

Q = 5.4Q = 25.7Axial Fracture Point

Chamber WallsChamber Centerline

Figure 5.23: Trajectory variations in tests run at a contraction ratio of 5.0.

Page 138: Thesis

123

5.7.3 Shear Layer Residence Time

The most important feature of a dump combustor is its flameholding capability

which is provided by a rearward-facing step, a described in Chapters 2 and 3. In

Chapter 3 it was suggested that autoignition could be correlated with the residence time

of the shear layer eddies created by the rearward-step. For this combustor the height of

the rearward step was kept fixed, which meant that the primary means of varying the

shear layer time was through the gas velocity. The gas velocity in the engine was altered

through the contraction ratio. Since many of the autoignition tests were performed at a

single contraction ratio the affect of the shear layer residence time was not explored in

depth. However, a shown in Figure 5.13, at a hydrogen peroxide concentration of 85%

the contraction ratio did affect autoignition. At a contraction ratio of 3.0 the shear layer

residence time was on the order of 0.075 milliseconds, ms. At this contraction ratio no

autoignition occurred at any of the tested conditions. When the contraction ratio was

increased to 5.0 the residence time increase to about 0.130 ms. Strong autoignition

occurred at an equivalence ratio of 1.36 at this contraction ratio and weak autoignition

was found up to 2.31 due to pressure oscillations. However, it is important to point out

that the monoprop chamber pressure also increased with contraction ratio and could have

also contributed to autoignition in addition to residence time. Since the shear layer

residence time was dependent on gas temperature the uncertainty approached 25% in

some cases.

5.7.4 Autoignition Correlation

The proposed correlation for autoignition based on shear layer residence time,

Chapter 3, was not attempted for a number of reasons. For one, the contraction ratio

was only varied at one hydrogen peroxide concentration, therefore the effect of residence

time was not observed at other temperatures. Also, even at a concentration of 85%,

where the contraction ratio was varied, a strong autoignition point was not achieved at a

Page 139: Thesis

124

contraction ratio of 3.0. As a result, there is no data indicating the conditions required for

autoignition at this contraction ratio and there was no way to compare the results between

the two. The pressure also changed with contraction ratio and therefore it may have

influenced autoignition as well as the residence time. In addition to the lack residence

time data, it is not clear from the results as to whether pressure or equivalence ratio had

the most influence on autoignition at a constant contraction ratio. Both parameters varied

when switching between test conditions and neither effect was ever isolated from the

other. Also, unstable chamber pressures in monoprop and biprop modes seemed to skew

the autoignition data. These instabilities created conditions for weak autoignition which

may have been classified as no autoignition under stable chamber pressure conditions.

Unreliable temperature measurements and low chamber pressures measured with high

range pressure transducers contributed to high uncertainty in many calculated parameters,

such residence time and decomposition temperature. This large uncertainty would cast

significant doubt into any autoignition correlation.

Page 140: Thesis

125

CHAPTER 6: SUMMARY AND CONCLUSIONS

A total of 24 tests were conducted to investigate the autoignition characteristics of

kerosene-based JP-8 fuel in decomposed hydrogen peroxide. Testing was performed

using staged-bipropellant rocket engine in a dump combustor configuration. The engine

used a catalyst bed to decompose the hydrogen peroxide and a transverse injector to

inject the JP-8 into the decomposed gas stream. A fuel jet trajectory analysis was

performed during injector design to model jet breakup and fuel distribution in the

oxidizer port and to prevent jet impingement. Downstream of the injection point a

rearward-facing step was used to provide flame stabilization at the entrance to the

combustion chamber. This design is commonly referred to as a dump combustor

configuration. Testing was structured to study the affects of gas temperature, gas

velocity, and equivalence ratio on autoignition.

Each test was classified into one of three groups based on visual observations and

measured chamber pressure data. Tests classified as strong autoignition produced a

stable, red-orange flame at the nozzle exit and bipropellant C* efficiencies of greater than

90%. The chamber pressure in these tests rose sharply within one-tenth of a second

following the initiation of fuel flow. The second classification, weak autoignition, was

typified by highly unstable flames that varied in color from red-orange to green. The C*

efficiencies for these tests ranged from 65 to 90%. The delay between fuel initiation and

chamber pressure rise was on the same order as the strong autoignition case, but the rise

was not as sharp. The third test classification was no autoignition. During these tests the

fuel was vaporized in the chamber but did not autoignite producing either a thick, white

vapor cloud or a clear plume at the nozzle exit. The biprop C* efficiencies for these tests

were less than 65% in most cases and the chamber pressure rise was minimal resulting

from the vaporization of the fuel.

Page 141: Thesis

126

It was determined that severe chamber pressure instabilities present during weak

autoignition tests caused the unstable flame structure. The instability caused the pressure

in the chamber to oscillate, and in some cases the average amplitude of the oscillation

was almost 30% of the chamber pressure. It is believed that the fuel and decomposed gas

mixture was initially autoignited at a high pressure point in the oscillation producing a

bright, red-orange flame. As the chamber pressure fell to a low point in the oscillation it

most likely altered the path of the combustion reaction causing a change in the color and

possibly the temperature of the flame. In some cases the pressure may have fell far

enough to quench the flame completely. This was seen during some tests when a flame

was visible at one instant and then a vapor cloud the next instant. As the pressure

continued to oscillate after the initial point of autoignition so too did the flame.

The gas temperature, gas velocity, and equivalence ratio were controlled during

testing by varying the H2O2 concentration, chamber contraction ratio, and oxidizer mass

flow rate respectively. Each test series was set up such that the concentration and

contraction ratio remained constant while the equivalence ratio was varied to determine

the boundary between strong autoignition and no autoignition for fuel rich conditions.

Once the boundary was determined at a particular concentration it was increased for the

next test series and the process was repeated again. Results showed that as the

concentration, or decomposition temperature, was increased the equivalence ratio at the

boundary between strong and no autoignition increased as well. At a contraction ratio of

3.0 and a concentration of 85% H2O2 no autoignition was achieved down to an

equivalence ratio of 1.37 while at a concentration of 98% strong autoignition was

achieved up to an equivalence ratio of 2.06. This trend agrees with past data from Mestre

and Ducourneau58 for kerosene in air as well as Walder8 for kerosene in decomposed

hydrogen peroxide. Both show that higher temperatures are required for autoignition as

equivalence ratio increases, or as the mixture becomes more fuel rich, for a specified

mixture residence time. Other studies done with kerosene fuel in air at oxidizer rich

equivalence ratios suggest that equivalence ratio plays a negligible role in autoignition.

However, due to the fact that the oxidizer flow rate was varied to change

equivalence ratio the monoprop chamber pressure was altered as well. There is nearly

Page 142: Thesis

127

universal agreement from past autoignition studies that the autoignition temperature

decreases with increasing pressure at a fixed residence time or contraction ratio in this

case. Data from Mestre and Ducourneau suggests that a pressure increase from 100 to

115 psia can alter the autoignition temperature of kerosene fuel in air by approximately

90°F at an equivalence ratio of 2.0. Data from Walder suggests a temperature difference

of about 20°F for kerosene fuel in decomposed hydrogen peroxide at the same pressures

and a stoichiometric equivalence ratio. It is believed that the variations in both the

pressure and equivalence ratio combined to influence the autoignition temperature at a

constant contraction ratio not just one or the other. However, the test conditions were

such that the variations could not be isolated from one another.

The effect of gas velocity on autoignition was investigated by varying the

contraction ratio of the engine. Increasing the contraction ratio decreases the Mach

number of the gases in the chamber as well as the decomposed gas in the oxidizer port.

As previously discussed, at contraction ratio of 3.0 no autoignition was achieved at all the

tested conditions using 85% H2O2. At this contraction ratio the Mach number in the

chamber is about 0.20 and 0.45 in the oxidizer port. When the contraction ratio was

increased to 5.0, however, strong autoignition was achieved at an equivalence ratio of

1.36 and weak autoignition was achieved up to an equivalence ratio of 2.31. At this

contraction ratio the Mach number in the chamber is about 0.12 and 0.20 in the oxidizer

port. Therefore, as the contraction ratio is increased, or the gas velocity decreased, the

temperature required to achieve autoignition at a particular equivalence ratio decreases.

This result also agrees with past data from Mestre and Ducourneau with regards to

residence time of a kerosene/air mixture. Walder made a similar conclusion and chose to

correlate the temperature decrease with the characteristic length of a rocket combustion

chamber instead of residence time. Intuitively this result makes sense because as the

available reaction time of the mixture increases the probability of autoignition should

increase as well at a certain temperature. Both studies suggest that the affect of residence

time, characteristic length, or gas velocity on the autoignition temperature is only

significant up to a point after which the effects are negligible. Since the fuel flow rate

was kept constant during contraction ratio variations as well the chamber pressure

Page 143: Thesis

128

increased with increasing contraction ratio. Based on the previous discussion on pressure,

it is believed that pressure affects also contributed to the decrease in autoignition

temperature at a larger contraction ratio.

Changes in contraction ratio also affected the trajectory of the fuel jet in the

oxidizer port. Fuel trajectory analysis was performed using momentum ratios calculated

from measured test data. All of the calculated trajectories at a contraction ratio of 3.0

penetrated halfway to the centerline of the duct or less, which includes both strong

autoignition tests and no autoignition tests. This may suggest that the fuel trajectory and

atomization did not play a critical role in autoignition. However, the variations in

momentum ratio were accompanied by changes in the equivalence ratio and flow rate of

hydrogen peroxide. Therefore, the affect of varying jet trajectory at a constant

equivalence ratio and monoprop chamber pressure was not determined.

As the contraction ratio was increased the shear layer residence time created by

the rearward-facing step was increased as well. Past studies have shown that the shear

layer time can be correlated with ignition delay to predict the stability of a flame. In this

study, the initial intention was to deve lop a correlation or model for autoignition relating

the shear layer residence time to an ignition delay parameter, which was similar in form

to that of past studies. The ignition delay parameter included the effects of temperature

and equivalence ratio, while velocity effects were included in the residence time

parameter. However, the correlation was not attempted due to inconclusive data

separating the effects of pressure, equivalence ratio, and contraction ratio. Most likely a

term would need to be added to the ignition delay portion of the correlation to include

pressure effects. In addition, large uncertainties were present in some of the calculated

data, including shear layer residence time. The large uncertainty originated from the

chamber pressure measurements, which were made using 3000 psia range transducers

with an accuracy of ±7.5 psia. During testing the monoprop chamber pressure was on the

order of 100 psia making the measurement uncertainty about 7.5%. The uncertainty

increased from this point through the rest of the calculations.

Despite this, it is believed that the rearward-facing step did provide enough

residence time to improve the autoignition limits based on past data. Many flight-rated

Page 144: Thesis

129

staged-bipropellant engines that used hydrogen peroxide and kerosene have had

contraction ratios of seven of higher and ran at stoichiometric equivalence ratios. A

study by Walder on the autoignition of kerosene in hydrogen peroxide used contraction

ratios of six and higher at stoichiometric conditions. In addition, data from Wu et al22

using a similar engine design running 85% H2O2 at an equivalence ratio somewhere

between 0.8 and 1.4 did not achieve autoignition at contraction ratio of approximately 5.0

even at a monoprop chamber pressure of 340 psia. Data from this study proved that

autoignition was possible at an equivalence ratio of 1.4 using 85% H2O2 at a monoprop

chamber pressure of approximately 100 psia. The data from this study also proves that

stable autoignition is possible at contraction ratios as low as 3.0 and equivalence ratios

less than 1.4 using 90% H2O2.

Future work should be aimed at developing the correlation for flame stability

based on shear layer residence time and ignition delay. This would require a study of the

affects of monoprop chamber pressure and equivalence ratio individually. To study

equivalence ratio effects at constant monoprop chamber pressure and contraction ratio the

flow rate of hydrogen peroxide would need to be held constant. As a result the fuel flow

rate would need to be varied to alter the equivalence ratio, which may require the use of

multiple fuel injectors to more accurately control the trajectory and pressure drop.

Unfortunately, to change out injectors the entire engine would need to be disassembled

with the current design. To study the effects of monoprop chamber pressure while

maintaining constant equivalence ratio and contraction ratio would require a change in

both the oxidizer and fuel flow rates. Most likely this would also require new injector

designs. In addition, a wider range of contraction ratios would need to be studied to

understand the effects of shear layer residence time on autoignition. The monoprop

chamber pressure would need to be held constant between contraction ratios to truly

determine this effect. Again this would require multiple fuel injectors to be made and a

catalyst bed would be required that performs adequately over a wide range of mass flow

rates.

Other suggestions for improvement to this autoignition study included finding an

improved method for measuring the temperature of the decomposed hydrogen peroxide.

Page 145: Thesis

130

The lack of an accurate temperature reading affects the uncertainty in the calculation of

the decomposed gas velocity and thus residence time and momentum ratio. Also, the use

of a more accurate pressure transducer in the chamber would also reduce the uncertainty

in many of the calculated parameters. It would also be beneficial to determine operating

conditions that reduce the probability of chamber pressure instability and to run

autoignition tests around these conditions. It would be preferable to eliminate pressure

oscillations from the autoignition tests. Use of a high-speed camera would create a better

picture of what occurs at the point of fuel injection and autoignition at the nozzle exit.

This could possibly allow the calculation of ignition delay following fuel injection. The

chamber length could also be varied to investigate its effect on residence time and

autoignition boundaries.

Despite its short comings this study generated some valuable data regarding the

autoignition of kerosene in a stream of decomposed hydrogen peroxide in a dump

combustor. In addition, a transverse injector design approach was developed which

proved to be both simple and reliable. Experimental data showed that the autoignition

limits were dependent on hydrogen peroxide concentration and chamber contraction ratio.

Also, a model for autoignition and flame stability was outlined based on the residence

time created by a rearward-facing step and ignition delay. Data showed that by adding a

flameholding capability to the engine autoignition could be achieved at lower contraction

ratios than in previous designs. This result may help in making staged-bipropellant

engines us ing hydrogen peroxide and kerosene lighter in weight and more reliable in

performance in the future.

Page 146: Thesis

131

LIST OF REFERENCES

1. Ventura, M., Mullens, P., “The Use of Hydrogen Peroxide for Propulsion and

Power,” AIAA Paper 99-2880, June 1999.

2. Andrews, D., “Advantages of Hydrogen Peroxide as a Rocket Oxidant,” Journal of The British Interplanetary Society, Vol. 43, 1990, pp. 319-328.

3. AFRPL-TR-67-144, “Hydrogen Peroxide Handbook,” Rocketdyne, Inc., July 1967.

4. TEP Version 1.5, SEA Software, Inc., Carson City, Nevada, 1999.

5. Humble, R. W., Henry, G. N., Larson, W. J., Space Propulsion Analysis and Design, McGraw-Hill Companies Inc., New York, 1995.

6. Hurlbert et al, “Nontoxic Orbital Maneuvering and Reaction Control Systems for Reusable Spacecraft,” Journal of Propulsion and Power, Vol. 14, No. 5, Sept.-Oct. 1998, pp. 676-687.

7. Chen et al, “Testing Research on Nontoxic H2O2/Kerosine Liquid Propellant Engines,” IAF-01-S309, 52nd International Astronautical Congress, Paris, 2001.

8. Walder, H., “An Investigation into the Thermal Ignition of Hydrogen Peroxide and Kerosine,” Report Number RPD 7, Royal Aircraft Establishment, May 1950.

9. Walder, H., “Further Investigations into the Thermal Ignition of Hydrogen Peroxide and Kerosine,” Technical Note Number RPD 43, Royal Aircraft Establishment, December 1950.

10. Walder, H. and Purchase, L. J., “The Influence of Injector Design on the Thermal Ignition of Hydrogen Peroxide and Kerosene,” Technical Note Number RPD 80, Royal Aircraft Establishment, April 1953.

Page 147: Thesis

132

11. Harlow, J., “Hydrogen Peroxide Engines: Early Work on Thermal Ignition at Westcott,” 2nd International Symposium on Hydrogen Peroxide, November 7-10, 1999.

12. Austin, B. L., Heister, S. D., “Characterization of Pintle Engine Performance for Nontoxic Hypergolic Bipropellants,” AIAA Paper 2002-4026, July 2002.

13. Long, M. R., Anderson, W. E., Humble, R. W., “Bi-Centrifugal Swirl Injector Development for Hydrogen Peroxide and Non-Toxic Hypergolic Miscible Fuels,” AIAA Paper 2002-4026, July 2002.

14. Andrews, D. “Rocket Engines for Satellite Launchers,” Proceeding from the International Symposium on Space Technology and Science, Tokyo, 1966, pp 111-120.

15. Healey, G. T., “Possibilities for Future Versions of Black Arrow-1,” Journal of Spacecraft, Vol. 10, No. 11, November 1968, pp. 394-401.

16. Gibbon et al, “Engergetic Green Propulsion for Small Spacecraft,” AIAA Paper 2001-3247, July 2001.

17. Coxhill, I., Richardson, G., Sweeting, M., “An Investigation of a Low Cost HTP/Kerosene 40N Thruster for Small Satellites,” AIAA Paper 2002-4155, July 2002.

18. Ventura, M., Wernimont, E., “History of the Reaction Motors Super Performance 90% H2O2/Kerosene LR-40 Rocket Engine,” AIAA Paper 2001-3838, July 2001.

19. Frazier, S. R., Moser, D. J., “Low Cost First Stage for Small Launch Vehicles,” AIAA Paper 95-3088, July 1995.

20. Phillips, E. H. “Beal Aerospace Developing New Launch Vehicle,” Aviation Week & Space Technology, Vol. 148, No. 14, April 6, 1998, pp. 74-75.

21. Phillips, E. H. “Beal Tests Stage 2 Liquid Fuel Engine,” Aviation Week & Space Technology, Vol. 152, No. 11, March 13, 2000, pp. 36.

22. Wu et al, “Development of a Pressure-Fed Rocket Engine Using Hydrogen Peroxide and JP-8,” AIAA Paper 99-2877, June 1999.

Page 148: Thesis

133

23. Anderson et al, “Upper Stage Flight Experiment 10K Engine Design and Test Results,” AIAA Paper 2000-3558, July 2000.

24. Fitzpatrick, S., Prater, D., Anderson, W., “A Design, Build, Test Course in Rocket Combustors,” AIAA Paper 2002-4186, July 2002.

25. Walder, H., Broughton, L. W., “Thermal Ignition Tests of Hydrogen Peroxide and Kerosine in a 2200lb Thrust Rocket Motor,” Technical Note Number RPD 70, Royal Aircraft Establishment, August 1952.

26. Miller, K. J., Sisco, J. C., Austin Jr., B. L., Martin, T. N., Anderson, W. E., “Design and Ground Testing of a Hydrogen Peroxide/Kerosene Combustor for a RBCC Application,” AIAA Paper 2003-4477, July 2003.

27. McCormick, J., “Hydrogen Peroxide Rocket Manual,” FMC Corporation, 1965.

28. Ventura, M., Wernimont, E., “Advancements in High Concentration Hydrogen Peroxide Catalyst Beds,” AIAA Paper 2001-3250, July 2001.

29. Andrews, D., Sunley, H., “The Gamma Rocket Engines for Black Knight,” Journal of The British Interplanetary Society, Vol. 43, 1990, pp. 301-310.

30. Helms, W. J., Mok, J. S., Sisco, J. C., Anderson, W. E., “Decomposition and Vaporization Studies of Hydrogen Peroxide,” AIAA Paper 2002-4028, July 2002.

31. Lefebvre, Arthur H., Atomization and Sprays, Hemisphere, New York, 1989.

32. Bayvel, L., Orzechowski, Z., Liquid Atomization, Taylor & Francis, 1993.

33. Doumas, M., Laster, R., “Liquid-Film Properties for Centrifugal Spray Nozzles,” Chemical Engineering Progress, Vol. 49, No. 10, 1953, pp 518-526.

34. Lin, K. C., Kennedy, P.J., Jackson, T. A., “Penetration Heights of Liquid Jets in High-Speed Crossflows,” AIAA Paper 2002-0873, January 2002.

35. Wu, P. K., Kirkendall, K. A., Fuller, R. P., Nejad, A. S., “Breakup Processes of Liquid Jets in Subsonic Crossflows,” AIAA Paper 96-3024, July 1996.

Page 149: Thesis

134

36. Chen, T. H., Smith, C. R., Schommer, D. G., Nejad, A. S., “Multi-Zone Behavior of Transverse Liquid Jet in High-Speed Flow,” AIAA Paper 93-0453, January 1993.

37. Mazallon, J., Dai, Z., Faeth, G. M., “Aerodynamic Primary Breakup at the Surface of Nonturbulent Round Liquid Jet in Crossflow,” AIAA Paper 98-0716, January 1998.

38. Sallam, K. A., Aalburg, C., Faeth, G. M., “Primary Breakup of Round Nonturbulent Liquid Jets in Gaseous Crossflows,” 16th Annual Conference on Liquid Atomization and Spray Systems, Monterey, CA, May 2003.

39. Ingebo, R. D., “Capillary and Acceleration Wave Breakup of Liquid Jets in Axial-Flow Airstreams,” NASA TP-1791, 1981.

40. Hautman, D. J., Rosfjord, T. J., “Transverse Liquid Injection Studies,” AIAA Paper 90-1965, July 1990.

41. Prior, R. C., Fowler, D. K., Mellor, A. M., “Engineering Design Models for Ramjet Efficiency and Lean Blowoff,” Journal of Propulsion and Power, Vol. 11, No. 1, Jan-Feb 1995, pp 117-123.

42. Plee, S. L., Mellor, A. M., “Characteristic Time Correlation for Lean Blowoff of Bluff-Body-Stabilized Flames,” Combustion and Flame, Vol. 35, 1979, pp. 61-80.

43. Craig, R. R., Drewry, J. D., Stull, F. D., “Coaxial Dump Combustor Investigations,” AIAA 78-1107, July 1978.

44. NASA SP-8089, “Liquid Rocket Engine Injectors,” National Aeronautics and Space Administration, March 1976.

45. Pitz, R. W., Daily, J. W., “Combustion in a Turbulent Mixing Layer Formed at a Rearward-Facing Step,” AIAA Journal, Vol. 21, No. 11, November 1983, pp. 1565-1570.

46. Hsiao, C. C., Oppenheim, A. K., Ghoniem, A. F., Chorin, A. J., “Numberical Simulation of a Turbulent Flame Stabilized Behind a Rearward-Facing Step,” Twentieth Symposium (International) on Combustion, The Combustion Institute, 1984, pp 495-504.

Page 150: Thesis

135

47. CRC Report No. 530, “Handbook of Aviation Fuel Properties,” Coordinating Research Council, 1983.

48. Kihm, K. D., Lyn, G. M., Son, S. Y., “Atomization of Cross-Injecting Sprays into Convective Air Stream,” Atomization and Sprays, Vol. 5, 1995, pp. 417-433.

49. Edwards, T., Harrison III, W. E., Maurice, L. Q., “Properties and Usage of Air Force Fuel: JP-8,” AIAA Paper 2001-0498, January 2001.

50. Edwards, T., “Kerosene Fuels for Aerospace Propulsion – Composition and Properties,” AIAA Paper 2002-3874, July 2002.

51. Turns, S. R., An Introduction to Combustion: Concepts and Applications, 2nd Edition, McGraw-Hill Companies, Inc., Boston, 2000.

52. Bodner, G. M., Pardue, H. L., Chemistry: An Experimental Science, 2nd Edition, John Wiley & Sons, Inc., New York, 1995.

53. Maron, S. H., Lando, J. B., Fundamentals of Physical Chemistry, Macmillan Publishing Co. Inc., New York, 1974.

54. AFAPL-TR-75-70, “Summary of Ignition Properties of Jet Fuels and Other Aircraft Combustible Fluids,” U.S. Bureau of Mines, September 1975.

55. Spadaccini, L. J., TeVelde, J. A., “Autoignition Characteristics of Aircraft-Type Fuels,” NASA CR-159886, June 1980.

56. Freeman, G., Lefebvre, A. H., “Spontaneous Ignition Characteristics of Gaseous Hydrocarbon-Air Mixtures,” Combustion and Flame, Vol. 58, 1984, pp. 153-162.

57. Freeman, G., “The Spontaneous Ignition Characteristics of Gaseous Hydrocarbon Fuel-Air Mixtures at Atmospheric Pressure,” M.S. Thesis, Purdue University, 1984.

58. Mestre, A., Ducourneau, F., “Recent Studies on the Spontaneous Ignition of Rich Air-Kerosene Mixtures,” Combustion Institute European Symposium, Academic Press, London, 1973, pp 225-229.

59. Colket III, M. B., Spadaccini, L. J., “Scramjet Fuels Autoignition Study,” Journal of Propulsion and Power, Vol. 17, No. 2, March-April 2001, pp 315-323.

Page 151: Thesis

136

60. Austin Jr., B. L., “Characterization of Pintle Engine Performance for Non-Toxic Hypergolic Bipropellants,” M.S. Thesis, Purdue University, 2002.

61. Fox, R. W., McDonald, A. T., Introduction to Fluid Mechanics, 5th Edition, John Wiley and Sons, Inc., New York, 1998.

Page 152: Thesis

137

Appendix A: Part Drawings

Page 153: Thesis

138

Figure A.1: Engine assembly using GK catalyst bed, page one.

Page 154: Thesis

139

Figure A.2: Engine assembly using GK catalyst bed, page two.

Page 155: Thesis

140

Figure A.3: Extension piece for GK catalyst bed, allows temperature and pressure

measurement upstream of the fuel injector.

Page 156: Thesis

141

Figure A.4: Mounting plate for engine assembly using GK catalyst bed.

Page 157: Thesis

142

Figure A.5: Page one of transverse fuel injector drawing, indicates manifold dimensions.

Page 158: Thesis

143

Figure A.6: Page two of transverse fuel injector drawing, indicates orifice dimensions.

Page 159: Thesis

144

Figure A.7: Transverse fuel injector seat, the fuel feed line is attached to this piece.

Page 160: Thesis

145

Figure A.8: Drawing of fuel film cooling injector seat, fuel feed line was capped for

autoignition testing.

Page 161: Thesis

146

Figure A.9: Drawing of fuel film cooling injector, fuel was not flowed through this piece

during autoignition testing.

Page 162: Thesis

147

Figure A.10: Page one of combustion chamber part drawing.

Page 163: Thesis

148

Figure A.11: Page two of combustion chamber part drawing.

Page 164: Thesis

149

Figure A.12: Page three of combustion chamber part drawing.

Page 165: Thesis

150

Figure A.13: Page four of combustion chamber part drawing.

Page 166: Thesis

151

Figure A.14: Drawing of nozzle piece with a contraction ratio of 3.0.

Page 167: Thesis

152

Figure A.15: Drawing of nozzle piece with contraction ratio of 5.0.

Page 168: Thesis

153

Figure A.16: Drawing of nozzle piece with contraction ratio of 6.5.

Page 169: Thesis

154

Figure A.17: Assembly drawing of water cooled deflection plate, water cooling apparatus

not shown.

Page 170: Thesis

155

Figure A.18: Engine assembly using PCI catalyst bed.

Page 171: Thesis

156

Figure A.19: Mounting plate used for engine assembly using PCI catalyst bed, this piece

was attached to the top of the catalyst bed.

Page 172: Thesis

157

Figure A.20: Page one of drawing of transition piece between PCI catalyst bed and transverse fuel injector. This piece also allowed the measurement of pressure and

temperature at the exit of the catalyst bed.

Page 173: Thesis

158

Figure A.21: Page two of drawing of PCI transition piece, indicates dimension of

V-shaped groove for metal o-ring.

Page 174: Thesis

159

Appendix B: DMAZ Material Safety Data Sheet

Page 175: Thesis

160

MATERIAL SAFETY 3M DATA SHEET 3M Center (Experimental) St. Paul, Minnesota 55144-1000 1-800-364-3577 or (651) 737-6501 (24 hours) Copyright, 2001, Minnesota Mining and Manufacturing Company. All rights reserved. Copying and/or downloading of this information for the purpose of properly utilizing 3M products is allowed provided that: 1) the information is copied in full with no changes unless prior agreement is obtained from 3M, and 2) neither the copy nor the original is resold or otherwise distributed with the intention of earning a profit thereon. DIVISION: 3M SPECIALTY MATERIALS MATERIAL: L-15686 DEVELOPMENTAL MATERIAL ISSUED: February 14, 2001 SUPERSEDES: February 13, 2001 DOCUMENT: 09-0400-3 ---------------------------------------------------------------------------- 1. INGREDIENT C.A.S. NO. PERCENT ---------------------------------------------------------------------------- DIMETHYL-2-AZIDOETHYLAMINE.............. 86147-04-8 100 This material is not listed on the TSCA inventory and should be used for research and development purposes only under the direct supervision of a technically qualified individual. This material is not on EINECS and should be used in Europe only for research purposes in order to establish its properties. ---------------------------------------------------------------------------- 2. PHYSICAL DATA ---------------------------------------------------------------------------- BOILING POINT:................. 135 C VAPOR PRESSURE:................ 11 - 50 mmHg @ 20C VAPOR DENSITY:................. N/D EVAPORATION RATE:.............. N/D SOLUBILITY IN WATER:........... slight SPECIFIC GRAVITY:.............. 0.93 Water=1 PERCENT VOLATILE:.............. 100 % pH:............................ N/A VISCOSITY:..................... < 10 centipoise MELTING POINT:................. N/A APPEARANCE AND ODOR: Clear mobile liquid. Strong amine-type odor. ---------------------------------------------------------------------------- Abbreviations: N/D - Not Determined N/A - Not Applicable CA - Approximately

Page 176: Thesis

161

MSDS: L-15686 DEVELOPMENTAL MATERIAL February 14, 2001 PAGE 2 ---------------------------------------------------------------------------- 3. FIRE AND EXPLOSION HAZARD DATA ---------------------------------------------------------------------------- FLASH POINT:................... 30 C CC FLAMMABLE LIMITS - LEL:........ N/A FLAMMABLE LIMITS - UEL:........ N/A AUTOIGNITION TEMPERATURE:...... N/D EXTINGUISHING MEDIA: Water, Carbon dioxide, Dry chemical, Foam SPECIAL FIRE FIGHTING PROCEDURES: Wear full protective clothing, including helmet, self-contained, positive pressure or pressure demand breathing apparatus, bunker coat and pants, bands around arms, waist and legs, face mask, and protective covering for exposed areas of the head. UNUSUAL FIRE AND EXPLOSION HAZARDS: See Hazardous Decomposition section for products of combustion. ---------------------------------------------------------------------------- 4. REACTIVITY DATA ---------------------------------------------------------------------------- STABILITY: Stable INCOMPATIBILITY - MATERIALS/CONDITIONS TO AVOID: Strong Acids, Strong Oxidizing Agents May react with acrylic monomers or oxirane compounds. Reacts violently with fuming nitric acid or nitrogen tetroxide. HAZARDOUS POLYMERIZATION: Hazardous polymerization will not occur. HAZARDOUS DECOMPOSITION PRODUCTS: Carbon Monoxide and Carbon Dioxide, Toxic Vapors, Gases or Particulates. ---------------------------------------------------------------------------- 5. ENVIRONMENTAL INFORMATION ---------------------------------------------------------------------------- SPILL RESPONSE: Refer to other sections of this MSDS for information regarding physical and health hazards, respiratory protection, ventilation, and personal protective equipment. Ventilate area. Extinguish all ignition sources. Contain spill. Evacuate unprotected personnel from hazard area. Cover with absorbent material. Cover spill area with Light Water Brand or other ATC foam. (For further information on ATC foam usage, contact 3M Fire Protection Systems.) Collect using non-sparking tools. Clean up residue. Place in an approved metal container. Seal the container. ---------------------------------------------------------------------------- Abbreviations: N/D - Not Determined N/A - Not Applicable CA - Approximately

Page 177: Thesis

162

MSDS: L-15686 DEVELOPMENTAL MATERIAL February 14, 2001 PAGE 3 ---------------------------------------------------------------------------- 5. ENVIRONMENTAL INFORMATION (continued) ---------------------------------------------------------------------------- RECOMMENDED DISPOSAL: Incinerate in a permitted hazardous waste incinerator. ENVIRONMENTAL DATA: Not determined. REGULATORY INFORMATION: Volatile Organic Compounds: N/D. VOC Less H2O & Exempt Solvents: N/A. Since regulations vary, consult applicable regulations or authorities before disposal. In the event of an uncontrolled release of this material, the user should determine if the release qualifies as a reportable quantity. U.S. EPA Hazardous Waste Number = D001 (Ignitable) ---------------------------------------------------------------------------- 6. SUGGESTED FIRST AID ---------------------------------------------------------------------------- EYE CONTACT: Immediately flush eyes with large amounts of water for at least 15 minutes. Get immediate medical attention. SKIN CONTACT: Immediately flush skin with large amounts of water for at least 15 minutes in a chemical safety shower while removing contaminated clothing and shoes. Get immediate medical attention. Wash contaminated clothing before reuse. INHALATION: If signs/symptoms occur, remove person to fresh air. If signs/symptoms continue, call a physician. IF SWALLOWED: If swallowed, do NOT induce vomiting. Give victim two glasses of water. Call a physician immediately. Never give anything by mouth to an unconscious person. ---------------------------------------------------------------------------- 7. PRECAUTIONARY INFORMATION ---------------------------------------------------------------------------- EYE PROTECTION: Avoid eye contact. Wear vented goggles. ---------------------------------------------------------------------------- Abbreviations: N/D - Not Determined N/A - Not Applicable CA - Approximately

Page 178: Thesis

163

MSDS: L-15686 DEVELOPMENTAL MATERIAL February 14, 2001 PAGE 4 ---------------------------------------------------------------------------- 7. PRECAUTIONARY INFORMATION (continued) ---------------------------------------------------------------------------- SKIN PROTECTION: Avoid skin contact. Wear appropriate gloves when handling this material. Use one or more of the following personal protection items as necessary to prevent skin contact: apron. RECOMMENDED VENTILATION: Provide appropriate local exhaust ventilation at transfer points. Provide appropriate local exhaust ventilation on open containers. If exhaust ventilation is not adequate, use appropriate respiratory protection. RESPIRATORY PROTECTION: Avoid breathing of vapors, mists or spray. Select one of the following NIOSH approved respirators based on airborne concentration of contaminants and in accordance with OSHA regulations: Half-mask organic vapor respirator with dust/mist prefilter, full-face supplied air respirator. PREVENTION OF ACCIDENTAL INGESTION: Wash hands after handling and before eating. RECOMMENDED STORAGE: Keep container closed when not in use. Keep container in well- ventilated area. FIRE AND EXPLOSION AVOIDANCE: Keep container tightly closed. Flammable liquid and vapor. Keep away from heat, sparks, open flame, and other sources of ignition. Ground containers securely when transferring contents. Wear low static or properly grounded shoes. No smoking while handling this material. Vapors may ignite explosively. EXPOSURE LIMITS INGREDIENT VALUE UNIT TYPE AUTH SKIN* ---------------------------------------------------------------------------- DIMETHYL-2-AZIDOETHYLAMINE........... NONE NONE NONE NONE * SKIN NOTATION: Listed substances indicated with 'Y' under SKIN refer to the potential contribution to the overall exposure by the cutaneous route including mucous membrane and eye, either by airborne or, more particularly, by direct contact with the substance. Vehicles can alter skin absorption. SOURCE OF EXPOSURE LIMIT DATA: - NONE: None Established ---------------------------------------------------------------------------- Abbreviations: N/D - Not Determined N/A - Not Applicable CA - Approximately

Page 179: Thesis

164

MSDS: L-15686 DEVELOPMENTAL MATERIAL February 14, 2001 PAGE 5 ---------------------------------------------------------------------------- 8. HEALTH HAZARD DATA ---------------------------------------------------------------------------- EYE CONTACT: Chemical Related Eye Burns (chemical corrosivity): signs/symptoms can include cloudy appearance of the cornea, chemical burns, pain, tearing, ulcers, impaired vision or loss of vision. SKIN CONTACT: Skin Burns (chemical corrosivity): signs/symptoms can include redness, swelling, itching, pain, blistering, ulceration, sloughing, and scar formation. INHALATION: No information was found regarding effects from inhalation exposure. IF SWALLOWED: No information was found regarding effects from swallowing. Ingestion may cause: Irritation of Gastrointestinal Tissues: signs/symptoms can include pain, vomiting, abdominal tenderness, nausea, blood in vomitus, and blood in feces. ---------------------------------------------------------------------------- SECTION CHANGE DATES ---------------------------------------------------------------------------- PHYSICAL DATA SECTION CHANGED SINCE February 13, 2001 ISSUE ---------------------------------------------------------------------------- Abbreviations: N/D - Not Determined N/A - Not Applicable CA - Approximately ---------------------------------------------------------------------------- The information in this Material Safety Data Sheet (MSDS) is believed to be correct as of the date issued. 3M MAKES NO WARRANTIES, EXPRESSED OR IMPLIED, INCLUDING, BUT NOT LIMITED TO, ANY IMPLIED WARRANTY OF MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE OR COURSE OF PERFORMANCE OR USAGE OF TRADE. User is responsible for determining whether the 3M product is fit for a particular purpose and suitable for user's method of use or application. Given the variety of factors that can affect the use and application of a 3M product, some of which are uniquely within the user's knowledge and control, it is essential that the user evaluate the 3M product to determine whether it is fit for a particular purpose and suitable for user's method of use or application. 3M provides information in electronic form as a service to its customers. Due to the remote possibility that electronic transfer may have resulted in errors, omissions or alterations in this information, 3M makes no representations as to its completeness or accuracy. In addition, information obtained from a database may not be as current as the information in the MSDS available directly from 3M.

Page 180: Thesis

165

Appendix C: Test Cell P&ID

Page 181: Thesis

166

Figure C.1: Plumbing & Instrumentation Diagram of Test Cell A at APCL.

Page 182: Thesis

167

Appendix D: Data Reduction

The following are examples of the MATLAB scripts which were used to reduce

the data collected during each test. The first file called ‘DataReduction.m’ was used to

read the data file, convert voltage readings to pressure, calculate mass flow rate, mixture

ratio, and C*, as well as filter and plot the data. The second file called ‘Bipropavg.m’

was used to average the measured and calculated data over a set period in time during the

biprop portion of each test. This file also performed an FFT on the pressure data during

this period of time and calculated the standard deviation in the chamber pressure data.

The averaged results were written to a text file. A similar file was used to perform the

same functions for the monoprop section of each test and was called ‘Monoavg.m.’ The

mass flow rates, C*’s, efficiencies, and other parameters were recalculated for monoprop

and biprop mode in a spreadsheet using the mean pressures determined in the averaging

files. This was done so that all the calculations for each test were contained in one file

and to facilitate using the uncertainty analysis.

DataReduction.m

%============================== %FILENAME: DataReduction.m %BY: B.J. Austin %Modified By: Jim Sisco %============================== clear all; close all; %Setting Plot Titles PlotTitle1='Autoignition Testing: Test #90305002, 01/21/03 Unfiltered'; PlotTitle2='Autoignition Testing: Test #90305002, 01/21/03 15Hz Filter';

Page 183: Thesis

168

%Loading Data File NameIn = input('Input File Name: ','s'); datafile = load(NameIn); %------ %Inputs %------ % Scan rate used by LabVIEW code ScanRate = 1000; %[samples/sec] % Assumed atmospheric pressure Patm = 14.7; %[psia] % Set H2O2 concentration Percent_HP = 90.5; %[%] % Assumed fuel temperature T_Fu = 72.5; %[F] % Set gravitational constant g = 32.174; %[lbm-ft/lbf-s] %Cavitating venturi characteristics % Oxidizer Cd_CV_ox = 0.989; %C_D for 0.111" cavitating venturi D_CV_ox = 0.111; %in % Fuel Cd_CV_fu = 0.844; %Cd for the 0.058" Dt cavitating venturi D_CV_fu = 0.058; %in % Chamber throat diameter Throat_Diam = 1.478; %in % Monoprop theoretical C*, based on concentration cstarthmono = 3086; %ft/s %--------------------------- %Calculating Time Parameters %--------------------------- % Total number of scans scans = length(datafile(:,1)); % Time resolution based on scan rate dt = 1/ScanRate; %[sec] % Time array from number of scans and time resolution Time = (1:scans)*dt; %[sec]

Page 184: Thesis

169

%--------------------------------------------- %Assigning Variables from Columns of Test File %--------------------------------------------- % Load Cell F_LoadCell_raw = datafile(:,1); %Thrust (not used for autoignition testing) % Pressure Transducers P_OxUllage_raw = datafile(:,2); %Primary Oxidizer Tank Ullage Pressure P_FuUllage_raw = datafile(:,3); %Primary Fuel Tank Ullage Pressure P_Fu2Ullage_raw = datafile(:,5); %Secondary Fuel Tank Ullage Pressure P_OxCav_raw = datafile(:,8); %Oxidizer Cavitating Venturi Inlet Pressure P_Catbed_raw = datafile(:,9); %Catalyst Bed Inlet Pressure P_FuCav_raw = datafile(:,10); %Fuel Cavitating Venturi Inlet Pressure P_FuInj_raw = datafile(:,11); %Fuel Injector Inlet Pressure P_CatbedOut_raw = datafile(:,12); %Catalyst Bed Outlet Pressure P_Chamber1_raw = datafile(:,13); %Chamber Pressure 1 P_Chamber2_raw = datafile(:,14); %Chamber Pressure 2 % Thermocouples T_OxLiquid = datafile(:,6); %Oxidizer Tank Liquid Temperature T_CatbedOut = datafile(:,15); %Catalyst Bed Outlet Temperature T_Chamber1 = datafile(:,16); %Chamber Temperature 1 T_Chamber2 = datafile(:,17); %Chamber Temperature 2 T_Chamber3 = datafile(:,18); %Chamber Temperature 3 T_Chamber4 = datafile(:,19); %Chamber Temperature 4 T_ChamberSkin1 = datafile(:,20); %Chamber Skin Temperature %Clear raw data variables to empty memory storage clear datafile; %---------------------------------------------- %Converting Transducer Voltage Data to Pressure %---------------------------------------------- % Setting slope of the calibration curve for each transducer Slope_P_OxUllage = 750; %Primary Oxidizer Ullage Pressure Slope_P_FuUllage = 750; %Fuel Ullage Pressure Slope_P_Fu2Ullage = 750; %Secondary Fuel Ullage Pressure Slope_P_OxCav = 750; %Oxidizer Cavitating Venturi Pressure Slope_P_Catbed = 750; %Catalyst Bed Inlet Pressure Slope_P_FuCav = 750; %Fuel Cavitating Venturi Pressure Slope_P_FuInj = 750; %Fuel Injector Inlet Pressure Slope_P_CatbedOut = 750; %Catalyst Bed Outlet Pressure Slope_P_Chamber1 = 750; %Chamber Pressure 1 Slope_P_Chamber2 = 750; %Chamber Pressure 2

Page 185: Thesis

170

%Creating Filter Design % Low-bypass digital Butterworth filter of 2nd order with 15Hz cutoff frequency % Cutoff frequency expressed as fraction of the Nyquist frequency (= ScanRate/2) [b,a] = butter(2,15*2/ScanRate); %Setting zeroing range in terms of scans (500 scans = 0.5 seconds) % Zero data is collected within the first 2000 scans (2 seconds) m = 500; n = 1500; %Calculating zero offset, absolute pressure, and filtered pressure P_OxUllage = P_OxUllage_raw * Slope_P_OxUllage; P_OxUllage_zero = mean(P_OxUllage(m:n)); P_OxUllage = P_OxUllage - P_OxUllage_zero + Patm; P_OxUllage_filt = filter(b,a,P_OxUllage); P_FuUllage = P_FuUllage_raw * Slope_P_FuUllage; P_FuUllage_zero = mean(P_FuUllage(m:n)); P_FuUllage = P_FuUllage - P_FuUllage_zero + Patm; P_FuUllage_filt = filter(b,a,P_FuUllage); P_Fu2Ullage = P_Fu2Ullage_raw * Slope_P_Fu2Ullage; P_Fu2Ullage_zero = mean(P_Fu2Ullage(m:n)); P_Fu2Ullage = P_Fu2Ullage - P_Fu2Ullage_zero + Patm; P_Fu2Ullage_filt = filter(b,a,P_Fu2Ullage); P_OxCav = P_OxCav_raw * Slope_P_OxCav; P_OxCav_zero = mean(P_OxCav(m:n)); P_OxCav = P_OxCav - P_OxCav_zero + Patm; P_OxCav_filt = filter(b,a,P_OxCav); P_Catbed = P_Catbed_raw * Slope_P_Catbed; P_Catbed_zero = mean(P_Catbed(m:n)); P_Catbed = P_Catbed - P_Catbed_zero + Patm; P_Catbed_filt = filter(b,a,P_Catbed); P_FuCav = P_FuCav_raw * Slope_P_FuCav; P_FuCav_zero = mean(P_FuCav(m:n)); P_FuCav = P_FuCav - P_FuCav_zero + Patm; P_FuCav_filt = filter(b,a,P_FuCav); P_FuInj = P_FuInj_raw * Slope_P_FuInj; P_FuInj_zero = mean(P_FuInj(m:n)); P_FuInj = P_FuInj - P_FuInj_zero + Patm; P_FuInj_filt = filter(b,a,P_FuInj);

Page 186: Thesis

171

P_CatbedOut = P_CatbedOut_raw * Slope_P_CatbedOut; P_CatbedOut_zero = mean(P_CatbedOut(m:n)); P_CatbedOut = P_CatbedOut - P_CatbedOut_zero + Patm; P_CatbedOut_filt = filter(b,a,P_CatbedOut); P_Chamber1 = P_Chamber1_raw * Slope_P_Chamber1; P_Chamber1_zero = mean(P_Chamber1(m:n)); P_Chamber1 = P_Chamber1 - P_Chamber1_zero + Patm; P_Chamber1_filt = filter(b,a,P_Chamber1); P_Chamber2 = P_Chamber2_raw * Slope_P_Chamber2; P_Chamber2_zero = mean(P_Chamber2(m:n)); P_Chamber2 = P_Chamber2 - P_Chamber2_zero + Patm; P_Chamber2_filt = filter(b,a,P_Chamber2); %Calculating Differential Pressures % Catalyst bed pressure drop DP_Catbed = P_Catbed - P_CatbedOut; DP_Catbed_filt = filter(b,a,DP_Catbed); % Fuel injector pressure drop DP_FuInj = P_FuInj - P_Chamber_avg; DP_FuInj_filt = filter(b,a,DP_FuInj); %--------------------------- %Calculating Mass Flow Rates %--------------------------- %Calculating densities % Average H2O2 temperature T_Ox = mean(T_OxLiquid(1:scans)); %[F] % Calculate density of H2O2 rho_Ox = 66.166 + ; %[lbm/ft^3] % Calculate fuel temperature using user-defined function rho_Fu = jp8density(T_Fu); %[lbm/ft^3] %Oxidizer flow rate A_CV_ox = pi*(D_CV_ox/2)^2; %[in^2] Mdot_Ox_CV = Cd_CV_ox*A_CV_ox*sqrt(2*g*rho_Ox*P_OxCav./144);%[lbm/s] % Set oxidizer flow rate to zero when it is not flowing for i = 1:scans, if P_Catbed(i) < 50 Mdot_Ox_CV(i) = 0.0; end end

Page 187: Thesis

172

%Fuel flow rate A_CV_fu = pi*(D_CV_fu/2)^2; %[in^2] Mdot_Fu_CV = Cd_CV_fu*A_CV_fu*sqrt(2*g*rho_Fu*P_FuCav./144);%[lbm/s] % Set fuel flow rate to zero when it is not flowing for i = 1:scans, if P_FuInj(i) < 200 Mdot_Fu_CV(i) = 0.0; end end %------------------------ %Performance Calculations %------------------------ %Calculate mixture ratio Mix_Rat = Mdot_Ox_CV./Mdot_Fu_CV; %Set dummy array for theoretical C* for biprop and monoprop cstarth = Mix_Rat; %Read data file containing theoretical C* for biprop mode as function of % mixture ratio for H2O2/JP8 at 500 psia A = dlmread('cstarpc500.txt','\t',6,0); %Create arrays of theoretical data of1 = A(:,1); cstar1 = A(:,3); %Calculate Theoretical C* %Set theoretical C* based on mixture ratio % If Mix_Ratio > 0, interpolate from data file % If Mix_Ratio = 0, set to monoprop theoretical C* for i = 1:scans, if Mdot_Fu_CV(i) > 0.0; Mix_Rat(i) = Mix_Rat(i); cstarth(i) = interp1(of1,cstar1,Mix_Rat(i)); %[ft/s] else Mix_Rat(i) = 0; cstarth(i) = cstarthmono; %[ft/s] end end

Page 188: Thesis

173

%Calculate C* and Efficiency %Calculation for C* cstar = P_Chamber2.*((Throat_Diam)^2*pi/4)*g./(Mdot_Ox_CV+Mdot_Fu_CV);%[ft/s] %Create dummy array for efficiency cstareff = (cstar./cstarthmono)*100; %[%] %Create final array for C* and efficiency % This if statement sets C* and efficiency to zero % when the oxidizer flow rate is zero for i = 1:scans if Mdot_Ox_CV(i) > 0 cstar(i) = cstar(i); %[ft/s] cstareff(i) = (cstar(i)/cstarth(i))*100; %[%] else cstar(i) = 0; cstareff(i) = 0; end end %------------- %Plotting Data %------------- figure(1) plot (Time, P_OxUllage_raw, Time, P_Fu2Ullage_raw, Time, P_OxCav_raw, Time, P_Catbed_raw, Time, P_FuCav_raw, Time, P_FuInj_raw, Time, P_CatbedOut_raw, Time, P_Chamber1_raw, Time, P_Chamber2_raw) xlabel ('Time [sec]') ylabel ('Volts [VDC]') title (['Raw Data, ' PlotTitle1]) legend ('POxUlg','PFu2Ulg', 'POxCav','PCatbed','PFuCav','PFuInj','PCatbedOut','Pc1','Pc2',3) grid on figure(2) plot (Time, P_OxUllage, Time, P_Fu2Ullage, Time, P_OxCav, Time, P_Catbed, Time, P_FuCav, Time, P_FuInj, Time, P_CatbedOut, Time, P_Chamber1, Time, P_Chamber2) xlabel ('Time [sec]') ylabel ('Pressure [psia]') title (['System Pressures, ' PlotTitle1]) legend ('POxUlg','PFu2Ulg','POxCav','PCatbed','PFuCav','PFuInj','PCatbedOut','Pc1','Pc2',3) grid on

Page 189: Thesis

174

figure(3) plot (Time, P_OxUllage_filt, Time, P_Fu2Ullage_filt, Time, P_OxCav_filt, Time, P_Catbed_filt, Time, P_FuCav_filt, Time, P_FuInj_filt, Time, P_CatbedOut_filt, Time, P_Chamber1_filt, Time, P_Chamber2_filt) xlabel ('Time [sec]') ylabel ('Pressure [psia]') title (['System Pressures, ' PlotTitle2]) legend ('POxUlg','PFu2Ulg','POxCav','Pcb','PFuCav','PFuInj','PCatbedOut','Pc1','Pc2',3) grid on figure(4) plot ( Time, Mdot_Ox_CV, Time, Mdot_Fu_CV) xlabel ('Time [sec]') ylabel ('Mass Flow [lbm/sec]') title (['Mass Flow Rate, ' PlotTitle1]) legend('Oxidizer','Fuel'); grid on; figure(5 plot ( Time, Mix_Rat) xlabel ('Time [sec]') ylabel ('Mixture Ratio') title (['Mixture Ratio, ' PlotTitle1]) grid on figure(6) plot(Time, cstar) xlabel ('Time [sec]') ylabel ('Cstar [ft/sec]') title (['Characteristic Velocity, ' PlotTitle2]) grid on figure(7) plot(Time, cstareff) xlabel ('Time [sec]') ylabel ('Cstar Efficiency [%]') title (['Cstar Efficiency, ' PlotTitle1]) grid on figure(8) plot (Time, T_OxLiquid) xlabel ('Time [sec]') ylabel ('Temperature [deg F]') title (['Oxidizer Liquid Temperature, ' PlotTitle2]) grid on

Page 190: Thesis

175

figure(9) plot(Time,P_Chamber_avg) xlabel ('Time [sec]') ylabel ('Pressure [psia]') title(['Average Chamber Pressure ' PlotTitle1]) grid on figure(10) plot(Time, DP_Catbed) xlabel ('Time [sec]') ylabel ('Pressure [psid]') title(['CatBed Pressure Drop ' PlotTitle1]) grid on figure(11) plot(Time, DP_Catbed_filt) xlabel ('Time [sec]') ylabel ('Pressure [psid]') title(['CatBed Pressure Drop ' PlotTitle2]) grid on figure(12) plot(Time, DP_FuInj) xlabel ('Time [sec]') ylabel ('Pressure [psid]') title(['Fuel Injector Pressure Drop ' PlotTitle1]) grid on figure(13) plot(Time, DP_FuInj_filt) xlabel ('Time [sec]') ylabel ('Pressure [psid]') title(['Fuel Injector Pressure Drop ' PlotTitle2]) legend('Transverse') grid on figure(14) plot(Time,P_OxCav ) xlabel ('Time [sec]') ylabel ('Pressure [psi]') title(['Oxidizer Venturi Inlet Pressure ' PlotTitle1]) grid on

Page 191: Thesis

176

figure(15) plot(Time,P_Catbed) xlabel ('Time [sec]') ylabel ('Pressure [psi]') title(['Catalyst Bed Inlet Pressure ' PlotTitle1]) grid on figure(16) plot(Time,P_FuCav ) xlabel ('Time [sec]') ylabel ('Pressure [psi]') title(['Fuel Venturi Inlet Pressure ' PlotTitle1]) grid on figure(17) plot(Time,P_FuInj) xlabel ('Time [sec]') ylabel ('Pressure [psi]') title(['Fuel Injector Inlet Pressure ' PlotTitle1]) grid on figure(18) plot(Time,P_CatbedOut ) xlabel ('Time [sec]') ylabel ('Pressure [psi]') title(['CatBed Outlet Pressure ' PlotTitle1]) grid on figure(19) plot(Time,P_Chamber1,Time,P_Chamber2) xlabel ('Time [sec]') ylabel ('Pressure [psi]') title(['Chamber Pressure' PlotTitle1]) legend('Pc1','Pc2'); grid on figure(20) plot (Time, T_CatbedOut) xlabel ('Time [sec]') ylabel ('Temperature [deg F]') title (['Catbed Outlet Temperature, ' PlotTitle2]) grid on

Page 192: Thesis

177

figure(21) plot (Time,T_Chamber1,Time,T_Chamber2,Time,T_Chamber3,Time,T_Chamber4 ) xlabel ('Time [sec]') ylabel ('Temperature [deg F]') title (['Chamber NFFC Temperatures, ' PlotTitle2]) legend('Tc1','Tc2','Tc3','Tc4',2) grid on figure(22) plot (Time,T_ChamberSkin1) xlabel ('Time [sec]') ylabel ('Temperature [deg F]') title (['Chamber Skin Temperatures, ' PlotTitle2]) grid on

Page 193: Thesis

178

Bipropavg.m

%========================== %FILENAME: Bipropavg.m %BY: Jim Sisco %========================== %Create output file fid = fopen('ign90305002out.txt','w'); %Set times which bound the averaging range t1 = 21.8; t2 = 22.5; %Convert the times to scans t1 = round(t1*1000); t2 = round(t2*1000); %----------------------- %Calculating Mean Values %----------------------- %Pressure data P_OxUllage2 = mean(P_OxUllage(t1:t2)); %Primary Oxidizer Tank Ullage Pressure P_OxCav2 = mean(P_OxCav(t1:t2)); %Oxidizer Cavitating Venturi Inlet Pressure P_Catbed2 = mean(P_Catbed(t1:t2)); %Catalyst Bed Inlet Pressure P_Fu2Ullage2 = mean(P_Fu2Ullage(t1:t2)); %Secondary Fuel Tank Ullage Pressure P_FuCav2 = mean(P_FuCav(t1:t2)); %Fuel Cavitating Venturi Inlet Pressure P_FuInj2 = mean(P_FuInj(t1:t2)); %Fuel Injector Inlet Pressure P_CatbedOut2 = mean(P_CatbedOut(t1:t2));%Catalyst Bed Outlet Pressure P_Chamber12 = mean(P_Chamber1(t1:t2)); %Chamber 1 Pressure P_Chamber22 = mean(P_Chamber2(t1:t2)); %Chmaber 2 Pressure %Temperature Data T_CatbedOut2 = mean(T_CatbedOut(t1:t2)); %Catalyst Bed Outlet Temperature %Calculated Data Mdot_Ox_CV2 = mean(Mdot_Ox_CV(t1:t2)); %Oxidizer Flow Rate Mdot_Fu_CV2 = mean(Mdot_Fu_CV(t1:t2)); %Fuel Flow Rate Mixture_Ratio = Mdot_Ox_CV2/Mdot_Fu_CV2; %Mixture Ratio cstar2 = mean(cstar(t1:t2)); %C* cstareff2 = mean(cstareff(t1:t2)); %C* Efficiency %Calculate differential pressures from average data dp_catbed = P_Catbed2 - P_CatbedOut2; %Catalyst Bed Pressure Drop dp_fuel_injector = P_FuInj2 - P_Chamber12; %Fuel Injector Pressure Drop %--------------

Page 194: Thesis

179

%Performing FFT %-------------- %Determine n, such that FFT uses 2^n points % (if 2^n is less than t1-t2 zero values are used to fill vector) p = ceil(log10(t2-t1)/log10(2)); N = 2^p; %Determine average value of important pressure parameters over time period P_Catbed_avg = mean(P_Catbed(t1:t2)); P_FuInj_avg = mean(P_FuInj(t1:t2)); P_CatbedOut_avg = mean(P_CatbedOut(t1:t2)); P_Chamber1_avg = mean(P_Chamber1(t1:t2)); P_Chamber2_avg = mean(P_Chamber2(t1:t2)); %Subtract out average value such that only noise remains P_Catbed_noise = P_Catbed - P_Catbed_avg; P_FuInj_noise = P_FuInj - P_FuInj_avg; P_CatbedOut_noise = P_CatbedOut - P_CatbedOut_avg; P_Chamber1_noise = P_Chamber1 - P_Chamber1_avg; P_Chamber2_noise = P_Chamber2 - P_Chamber2_avg; %Perform FFT on pressure noise [P_Catbed_FFT] = fft(P_Catbed_noise(t1:t2), N); [P_FuInj_FFT] = fft(P_FuInj_noise(t1:t2), N); [P_CatbedOut_FFT] = fft(P_CatbedOut_noise(t1:t2), N); [P_Chamber1_FFT] = fft(P_Chamber1_noise(t1:t2), N); [P_Chamber2_FFT] = fft(P_Chamber2_noise(t1:t2), N); %Calculate magnitude of FFT amplitude Mag_P_Catbed_FFT = sqrt(P_Catbed_FFT.*conj(P_Catbed_FFT)); Mag_P_FuInj_FFT = sqrt(P_FuInj_FFT.*conj(P_FuInj_FFT)); Mag_P_CatbedOut_FFT = sqrt(P_CatbedOut_FFT.*conj(P_CatbedOut_FFT)); Mag_P_Chamber1_FFT = sqrt(P_Chamber1_FFT.*conj(P_Chamber1_FFT)); Mag_P_Chamber2_FFT = sqrt(P_Chamber2_FFT.*conj(P_Chamber2_FFT)); %Calculate power spectrum Spectrum_P_Catbed = Mag_P_Catbed_FFT.^2./N; Spectrum_P_FuInj = Mag_P_FuInj_FFT.^2./N; Spectrum_P_CatbedOut = Mag_P_CatbedOut_FFT.^2./N; Spectrum_P_Chamber1 = Mag_P_Chamber1_FFT.^2./N; Spectrum_P_Chamber2 = Mag_P_Chamber2_FFT.^2./N;

Page 195: Thesis

180

%Calculate step in time and frequency del_t = 1/ScanRate; del_freq = 1/(N*del_t); %Calculate frequency vector (0-->Nyquist Frequency) freq = del_freq.*(0:N/2); %Calculate Max & Min of pressure noise as well as peak-to-peak amplitude Max_P_Catbed_noise = max(P_Catbed_noise(t1:t2)); Min_P_Catbed_noise = min(P_Catbed_noise(t1:t2)); Max_P_FuInj_noise = max(P_FuInj_noise(t1:t2)); Min_P_FuInj_noise = min(P_FuInj_noise(t1:t2)); Max_P_CatbedOut_noise = max(P_CatbedOut_noise(t1:t2)); Min_P_CatbedOut_noise = min(P_CatbedOut_noise(t1:t2)); Max_P_Chamber1_noise = max(P_Chamber1_noise(t1:t2)); Min_P_Chamber1_noise = min(P_Chamber1_noise(t1:t2)); Max_P_Chamber2_noise = max(P_Chamber2_noise(t1:t2)); Min_P_Chamber2_noise = min(P_Chamber2_noise(t1:t2)); PTP_P_Catbed_noise = Max_P_Catbed_noise - Min_P_Catbed_noise; PTP_P_FuInj_noise = Max_P_FuInj_noise - Min_P_FuInj_noise; PTP_P_CatbedOut_noise = Max_P_CatbedOut_noise - Min_P_CatbedOut_noise; PTP_P_Chamber1_noise = Max_P_Chamber1_noise - Min_P_Chamber1_noise; PTP_P_Chamber2_noise = Max_P_Chamber2_noise - Min_P_Chamber2_noise; %Find magnitude and frequency of highest peak in power spectrum [Max_Catbed, I_Catbed] = max(Spectrum_P_Catbed(1:(N/2)+1)); [Max_FuInj, I_FuInj] = max(Spectrum_P_FuInj(1:(N/2)+1)); [Max_CatbedOut, I_CatbedOut] = max(Spectrum_P_CatbedOut(1:(N/2)+1)); [Max_Chamber1, I_Chamber1] = max(Spectrum_P_Chamber1(1:(N/2)+1)); [Max_Chamber2, I_Chamber2] = max(Spectrum_P_Chamber2(1:(N/2)+1)); dom_freq_Catbed = freq(I_Catbed); dom_freq_FuInj = freq(I_FuInj); dom_freq_CatbedOut = freq(I_CatbedOut); dom_freq_Chamber1 = freq(I_Chamber1); dom_freq_Chamber2 = freq(I_Chamber2); %Calculate standard deviation of chamber pressure data std_Chamber2 = std(P_Chamber2(t1:t2));

Page 196: Thesis

181

%-------------------- %Plot Power Spectrums %-------------------- figure(80) plot(freq, Spectrum_P_Catbed(1:(N/2)+1), freq, Spectrum_P_FuInj(1:(N/2)+1)); legend('Catbed','FuInj'); grid; xlabel('Frequency (Hz)'); ylabel('Power (psi^2/Hz ???)'); title (['Power Spectrum, 'PlotTitle1]) figure(81) plot(freq, Spectrum_P_CatbedOut(1:(N/2)+1), freq, Spectrum_P_Chamber1(1:(N/2)+1), freq, Spectrum_P_Chamber2(1:(N/2)+1)); legend('CatbedOut','Chamber1','Chamber2'); grid; xlabel('Frequency (Hz)'); ylabel('Power (psi^2/Hz ???)'); title (['Power Spectrum, 'PlotTitle1]) %------------------ %Write Data to File %------------------ fprintf(fid,'Ignition Testing - Test 90-30-50-02 (PCI Catbed)\n\n'); fprintf(fid,'Oxidizer Ullage Pressure (psi) = \t\t\t\t\t\t\t%8.2f\n',P_OxUllage2); fprintf(fid,'Oxidizer Cavitating Venturi Inlet Pressure (psi) = \t\t\t%8.2f\n',P_OxCav2); fprintf(fid,'Catalyst Bed Inlet Pressure (psi) = \t\t\t\t\t\t%8.2f\n',P_Catbed2); fprintf(fid,'Catalyst Bed Outlet Pressure (psi) = \t\t\t\t\t\t%8.2f\n',P_CatbedOut2); fprintf(fid,'Catalyst Bed Pressure Drop (psi) = \t\t\t\t\t\t\t%8.2f\n',dp_catbed); fprintf(fid,'Oxidizer Flow Rate (lbm/s) = \t\t\t\t\t\t\t\t%8.3f\n\n',Mdot_Ox_CV2); fprintf(fid,'Fuel Ullage Pressure (psi) = \t\t\t\t\t\t\t\t%8.2f\n',P_Fu2Ullage2); fprintf(fid,'Fuel Cavitating Venturi Inlet Pressure (psi) = \t\t\t\t%8.2f\n',P_FuCav2); fprintf(fid,'Fuel Injector Supply Pressure (psi) = \t\t\t\t\t\t%8.2f\n',P_FuInj2); fprintf(fid,'Fuel Injector Pressure Drop (psi) = \t\t\t\t\t\t%8.2f\n',dp_fuel_injector); fprintf(fid,'Fuel Flow Rate (lbm/s) = \t\t\t\t\t\t\t\t\t%8.3f\n\n',Mdot_Fu_CV2); fprintf(fid,'Mixture Ratio = \t\t\t\t\t\t\t\t\t\t\t%8.2f\n',Mixture_Ratio); fprintf(fid,'Chamber Pressure 1 (psi) = \t\t\t\t\t\t\t\t\t%8.2f\n',P_Chamber12); fprintf(fid,'Chamber Pressure 2 (psi) = \t\t\t\t\t\t\t\t\t%8.2f\n',P_Chamber22); fprintf(fid,'Characteristic Velocity (ft/s) = \t\t\t\t\t\t\t%8.2f\n',cstar2); fprintf(fid,'C* Efficiency (percent) = \t\t\t\t\t\t\t\t\t%8.2f\n\n',cstareff2); fprintf(fid,'Catbed Outlet Temperature (F) = \t\t\t\t\t\t\t%8.2f\n\n\n',T_CatbedOut2); fprintf(fid,'Biprop Pressure Oscillations:\n\n'); fprintf(fid,'Catalyst Bed Inlet Pressure Frequency (Hz) =\t\t\t\t%8.2f\n',dom_freq_Catbed);

Page 197: Thesis

182

fprintf(fid,'Catalyst Bed Inlet Pressure Peak-to-Peak (psid) =\t\t\t%8.2f\n\n',PTP_P_Catbed_noise); fprintf(fid,'Fuel Injector Inlet Pressure Frequency (Hz) =\t\t\t\t%8.2f\n',dom_freq_FuInj); fprintf(fid,'Fuel Injector Inlet Pressure Peak-to-Peak (psid) =\t\t\t%8.2f\n\n',PTP_P_FuInj_noise); fprintf(fid,'Catalyst Bed Outlet Pressure Frequency (Hz) =\t\t\t\t%8.2f\n',dom_freq_CatbedOut); fprintf(fid,'Catalyst Bed Outlet Pressure Peak-to-Peak (psid) =\t\t\t%8.2f\n\n',PTP_P_CatbedOut_noise); fprintf(fid,'Chamber 1 Pressure Frequency (Hz) =\t\t\t\t\t\t\t%8.2f\n',dom_freq_Chamber1); fprintf(fid,'Chamber 1 Pressure Peak-to-Peak (psid) =\t\t\t\t\t%8.2f\n\n',PTP_P_Chamber1_noise); fprintf(fid,'Chamber 2 Pressure Frequency (Hz) =\t\t\t\t\t\t\t%8.2f\n',dom_freq_Chamber2); fprintf(fid,'Chamber 2 Pressure Peak-to-Peak (psid) =\t\t\t\t\t%8.2f\n\n',PTP_P_Chamber2_noise); fclose(fid);

Page 198: Thesis

183

Appendix E: Uncertainty Analysis

The following equations for uncertainty in each calculated variable were derived

by the method outlined in Fox and McDonald.61

Propellant Density

The density of hydrogen peroxide, ?ox, is calculated from its temperature, Tox, and

concentration, W, using equation 4.8, which has the following form:

oxoxoxox FWTETDTCWBWA +++++= 22ρ , (E.1)

Therefore, the uncertainty in the hydrogen peroxide density can be calculated

from the following equation:

21

22

∂∂

+

∂∂

= Toxox

ox

ox

oxW

ox

oxox u

TT

uW

Wu

ρρ

ρρρ , (E.2)

Taking the derivative of the hydrogen peroxide density with respect to

concentration and temperature gives the following:

oxox FTCWB

W++=

∂∂

, (E.3)

FWETDT ox

ox

ox ++=∂∂

, (E.4)

For JP-8, the density, ?f, is a linear function of temperature, Tf, and can be written

as follows:

ff QTP +=ρ , (E.5)

Page 199: Thesis

184

The derivative of this equation with respect to fuel temperature is equal to the

constant P. Then, the uncertainty in the density of JP-8 can be written as:

21

2

= Tf

f

ff u

TPu

ρρ , (E.6)

Data regarding the variation in the density of DMAZ with its temperature is not

currently known. As a result, the uncertainty in the DMAZ density was assumed to be

equal to zero.

Mass Flow Rate and Equivalence Ratio

Using equation 4.4 the uncertainty in the mass flow rate of hydrogen peroxide,

JP-8, or DMAZ can be determined from the following equation:

21

2222

∂∂

+

∂∂

+

∂∂

+

∂∂

= pdtt

tCD

D

Dm u

pm

mp

um

mu

dm

md

uCm

mC

u&

&&

&&

&&

&& ρρρ

, (E.7)

This equation can be simplified substantially, after substituting the appropriate

derivatives, to express the uncertainty in flow rate as:

( ) ( )2

122

22

222

+

++= p

dtCDm

uuuuu ρ

& , (E.8)

Starting from equations 3.6 and 5.5 the uncertainty in equivalence ratio can be

expressed as:

21

222

∂+

∂∂

+

∂∂

= FsOs

sfm

f

foxm

ox

ox uFO

FOu

m

mu

mm

φφ

φφ

φφ && &

&

&&

, (E.9)

The equation for uncertainty in equivalence ratio can be simplified a great deal as

well to give:

( ) ( ) ( )[ ] 21

222FsOfmoxm uuuu ++−= &&φ , (E.10)

The uncertainty in the total mass flow rate, which is used to calculate C*, can be

determined from the following equation:

Page 200: Thesis

185

21

22

∂∂

+

∂∂

= fmf

tot

tot

foxm

ox

tot

tot

oxtotm u

mm

m

mu

mm

mm

u &&& &&

&

&

&&

&&

, (E.11)

Since the derivatives of the total mass flow rate with respect to the fuel and

oxidizer flow rate are equal to one this equation can be simplified to:

21

22

+

= fm

tot

foxm

tot

oxtotm u

m

mu

mm

u &&& &

&

&&

, (E.12)

Performance Parameters

The uncertainty in the characteristic velocity, equation 5.7, in both monoprop and

biprop modes can be calculated from the following equation:

21

222

*

∂∂

+

∂∂

+

∂∂

=∗

∗ totmtot

totDth

th

thpc

c

cC u

mC

Cm

uDC

CD

upC

Cp

u &&&

, (E.13)

Upon simplification the equation reduces to:

( ) ( ) ( )[ ] 21

222* 2 totmDthpcC uuuu &−++= , (E.14)

The C* or decomposition efficiency, equation 5.10, can be found from the

following:

21

2

**

*

2

**

**

∂∂

+

∂∂

=∗

thCth

C

C

thC

C

CC u

CC

uC

Cu

ηη

ηηη , (E.15)

This equation reduces to:

( ) ( )[ ] 21

2*

2** thCCC uuu −+=η , (E.16)

Page 201: Thesis

186

Momentum Ratio and Residence Time

There are a number of uncertainty parameters which must be determined to

calculate the uncertainty in the momentum ratio. One of these parameters is the

uncertainty in the fuel velocity exiting the injector, calculated from equations 2.3 and

2.4. The uncertainty in fuel velocity can be found from the following equation:

21

22

22

∂+

∂+

∂+

=

doo

f

f

oCD

D

f

f

D

ff

f

f

ffm

f

f

f

f

Vf

ud

V

Vd

uC

V

VC

uV

Vu

m

V

V

m

uρρ

ρ&&

&

, (E.17)

Upon simplification this equation becomes:

( ) ( ) ( ) ( )[ ] 21

2222 2 doCDffmVf uuuuu −+−+−+= ρ& , (E.18)

To calculate the uncertainty in the properties of the decomposed gases the

uncertainty in the temperature of the gas must be determined first. The uncertainty in the

decomposition temperature, equation 5.11, is calculated from the following:

21

2

__

_2

_**

_*

∂∂

+

∂∂

= thTtoxthtox

tox

tox

thtoxmonoC

C

tox

tox

monoCTtox u

TT

T

Tu

TT

u ηη

η, (E.19)

Simplifying this equation gives:

( ) ( )[ ] 21

2_

2_*2 thTtoxmonoCTtox uuu += η , (E.20)

Using the uncertainty in temperature the uncertainty in the density of the

decomposed gas, equation 5.14, is determined from the following equation assuming an

incompressible fluid:

21

2

22

∂+

∂+

=

Ttoxtox

oxg

oxg

tox

MWoxgoxg

oxg

oxg

oxgptox

tox

oxg

oxg

tox

oxg

uT

T

uMW

MWu

pp

ρ

ρ

ρ

ρ

ρρ , (E.21)

This equation can be simplified to:

Page 202: Thesis

187

( ) ( ) ( )[ ] 21

222TtoxMWoxgptoxoxg uuuu −++=ρ , (E.22)

With the uncertainty in the density known the uncertainty in the velocity of the

decomposed gas, equation 5.13 which assumes incompressible flow, can be found from:

21

222

∂∂

+

∂∂

+

∂∂

= Doxox

ox

ox

oxoxg

oxg

ox

ox

oxgoxm

ox

ox

ox

oxVox u

DV

VD

uV

Vu

mV

Vm

u ρρ

ρ&&

&, (E.23)

After simplification this equation becomes:

( ) ( ) ( )[ ] 21

222 2 DoxoxgoxmVox uuuu −+−+= ρ& , (E.24)

The uncertainty in momentum ratio, equation 2.24, can now be determined from

the following:

21

22

22

∂∂

+

∂∂

+

∂∂

+

∂∂

=

Voxox

oxoxg

oxg

oxg

Vff

ff

f

f

Q

uVQ

QV

uQ

Q

uVQ

Q

Vu

QQ

u

ρ

ρ

ρ

ρ

ρ

ρ

, (E.25)

This expression simplifies to:

( ) ( ) ( ) ( )[ ] 21

2222 22 VoxoxgVffQ uuuuu −+−++= ρρ , (E.26)

The uncertainty in the shear layer residence time, equation 2.12, is determined

from the following equation:

21

22

∂+

∂∂

= hsl

slVox

ox

sl

sl

oxsl u

hh

uV

Vu

ττ

τττ , (E.27)

After simplification this equation becomes:

( ) ( )[ ] 21

22hVoxsl uuu +−=τ , (E.28)

Finally, the uncertainty in Mach number, equation 5.12, neglecting uncertainty in

the molecular weight and specific heat ratio of the decomposed gas is determined from:

Page 203: Thesis

188

21

22

22

∂∂

+

∂∂

+

∂∂

+

∂∂

=

Ttoxtox

ox

ox

toxDox

ox

ox

ox

ox

ptoxtox

ox

ox

toxoxm

ox

ox

ox

ox

Mox

uTM

MT

uDM

MD

upM

Mp

umM

Mm

u&&

&

, (E.29)

Following simplification this equation becomes:

( ) ( ) ( )2

12

222

22

+−+−+= Ttox

DoxptoxoxmMoxu

uuuu & , (E.30)

Other Uncertainty Parameters

The uncertainty in the pressure drop across the catalyst bed, equation 5.4, can be

calculated from the following equation:

21

2

__

_

2

__

_

∂∆∂

∆+

∂∆∂

∆=∆ outpcb

outcb

cb

cb

outcbinpcb

incb

cb

cb

incbpcb u

pp

p

pu

pp

p

pu , (E.31)

The uncertainty in the contraction ratio is calculated from:

21

22

∂∂

+

∂∂

= Dthth

thDc

c

cCR u

DCR

CRD

uDCR

CRD

u , (E.32)

This equation simplifies to the following:

( ) ( )[ ] 21

22 22 DthDcCR uuu −+= , (E.33)

Page 204: Thesis

189

Appendix F: Test Data

Page 205: Thesis

190

Table F.1: Measured and calculated test data during bipropellant operation.

Test Number Injector W Fuel CR Dth L* p_ox_cv dth_ox_cv CD_ox_cv T_ox rho_ox mdot_ox p_fu_cv dth_f_cv CD_f_cv rho_f mdot_f(xx-xx-xx-xx) (Catbed) (%) (--) (in.) (in.) (psia) (in.) (--) (F) (lbm/ft^3) (lbm/s) (psia) (in.) (--) (lbm/ft^3) (lbm/s)90-30-17-01 B (GK) 90.6 DMAZ 3.0 1.478 24.1 959.9 0.078 0.839 79.6 86.72 0.774 1682.1 0.058 0.844 57.98 0.46690-30-18-01 B (GK) 89.5 DMAZ 3.0 1.478 24.1 1160.7 0.078 0.839 73.2 86.57 0.850 1697.9 0.058 0.844 57.98 0.46890-30-22-01 B (GK) 89.5 DMAZ 3.0 1.478 24.1 1610.8 0.078 0.839 72.2 86.60 1.001 1704.5 0.058 0.844 57.98 0.46985-30-17-01 B (GK) 85.0 DMAZ 3.0 1.478 24.1 1026.6 0.078 0.839 73.9 85.00 0.792 1692.2 0.058 0.844 57.98 0.46785-30-21-01 B (GK) 84.7 DMAZ 3.0 1.478 24.1 1530.7 0.078 0.839 72.0 84.97 0.967 1686.6 0.058 0.844 57.98 0.46685-30-25-01 B (GK) 84.7 DMAZ 3.0 1.478 24.1 1640.9 0.078 0.839 73.3 84.93 1.001 1221.8 0.058 0.844 57.98 0.39785-30-50-01 B (GK) 85.4 JP-8 3.0 1.478 24.1 1374.7 0.111 0.989 82.8 84.82 2.186 1651.0 0.058 0.844 50.68 0.43185-30-64-01 B (GK) 85.4 JP-8 3.0 1.478 24.1 1446.4 0.111 0.989 79.6 84.94 2.243 1201.1 0.058 0.844 50.68 0.36885-30-64-03 B (GK) 85.4 JP-8 3.0 1.478 24.1 1558.6 0.111 0.989 73.9 85.14 2.331 1251.1 0.058 0.844 50.68 0.37687-30-44-01 B (GK) 87.7 JP-8 3.0 1.478 24.1 1073.5 0.111 0.989 72.3 85.98 1.944 1697.1 0.058 0.844 50.68 0.43790-30-40-02 B (PCI) 89.4 JP-8 3.0 1.478 24.1 1633.3 0.094 0.994 77.5 86.38 1.733 1666.8 0.058 0.844 50.68 0.43390-30-35-02 B (PCI) 90.5 JP-8 3.0 1.478 24.1 1318.1 0.094 0.994 78.8 86.71 1.560 1663.0 0.058 0.844 50.68 0.43394-30-35-01 B (PCI) 94.1 JP-8 3.0 1.478 24.1 1258.8 0.094 0.994 68.0 88.38 1.539 1652.4 0.058 0.844 50.68 0.43294-30-40-01 B (PCI) 94.1 JP-8 3.0 1.478 24.1 1666.2 0.094 0.994 80.9 87.90 1.765 1668.2 0.058 0.844 50.68 0.43498-30-30-01 B (PCI) 98.1 JP-8 3.0 1.478 24.1 938.4 0.094 0.994 56.8 90.24 1.342 1666.9 0.058 0.844 50.68 0.43385-50-31-01 B (GK) 85.2 JP-8 5.0 1.145 40.4 1033.2 0.094 0.994 72.3 85.13 1.368 1715.7 0.058 0.844 50.68 0.44087-30-60-01 B (GK) 87.3 JP-8 3.0 1.478 24.1 1562.3 0.111 0.989 72.8 85.83 2.344 1399.4 0.058 0.844 50.68 0.39790-30-50-01 B (GK) 90.0 JP-8 3.0 1.478 24.1 1342.4 0.111 0.989 75.1 86.67 2.183 1666.2 0.058 0.844 50.68 0.43390-30-50-02 B (PCI) 90.5 JP-8 3.0 1.478 24.1 1342.1 0.111 0.989 67.8 87.11 2.188 1657.5 0.058 0.844 50.68 0.43294-30-45-01 B (PCI) 93.5 JP-8 3.0 1.478 24.1 1068.6 0.111 0.989 81.9 87.65 1.959 1662.3 0.058 0.844 50.68 0.43398-30-35-01 B (PCI) 98.1 JP-8 3.0 1.478 24.1 1240.1 0.094 0.994 76.3 89.51 1.537 1648.4 0.058 0.844 50.68 0.43198-30-40-01 B (PCI) 98.1 JP-8 3.0 1.478 24.1 1585.2 0.094 0.994 80.7 89.35 1.736 1675.0 0.058 0.844 50.68 0.43485-50-64-01 B (GK) 86.0 JP-8 5.0 1.145 40.4 1511.0 0.111 0.989 79.5 85.14 2.296 1201.4 0.058 0.844 50.68 0.36887-30-52-01 B (GK) 87.4 JP-8 3.0 1.478 24.1 1425.9 0.111 0.989 72.6 85.87 2.239 1716.7 0.058 0.844 50.68 0.44087-30-49-01 B (GK) 87.7 JP-8 3.0 1.478 24.1 1265.2 0.111 0.989 74.3 85.91 2.110 1685.4 0.058 0.844 50.68 0.43690-30-35-01 B (GK) 89.5 JP-8 3.0 1.478 24.1 710.7 0.111 0.989 74.9 86.51 1.587 1668.6 0.058 0.844 50.68 0.43490-30-40-01 B (GK) 89.8 JP-8 3.0 1.478 24.1 933.4 0.111 0.989 64.0 87.00 1.824 1676.7 0.058 0.844 50.68 0.43590-30-44-01 B (GK) 89.8 JP-8 3.0 1.478 24.1 1071.4 0.111 0.989 68.4 86.85 1.952 1681.5 0.058 0.844 50.68 0.43585-50-37-01 B (GK) 84.8 JP-8 5.0 1.145 40.4 1442.7 0.094 0.994 77.2 84.82 1.614 1705.0 0.058 0.844 50.68 0.43885-50-50-01 B (GK) 86.0 JP-8 5.0 1.145 40.4 1334.2 0.111 0.989 79.5 85.14 2.157 1649.3 0.058 0.844 50.68 0.431

Page 206: Thesis

191

Table F.2: Measured and calculated bipropellant test data (cont.)

IgnitionStrong,

Test Number mdot_tot O/F φ p_cb_in p_cb_out ∆p_cb p_c1 p_c2 p_c2/ P_c2_tot C* C*_th η_C* Weak, p_fu_inj ∆p_f CD_inj(xx-xx-xx-xx) (lbm/s) (--) (--) (psia) (psia) (psid) (psia) (psia) p_c2_tot (psia) (ft/s) (ft/s) (%) No Ign. (psia) (psid) (--)90-30-17-01 1.239 1.66 2.60 -- 115.7 -- 93.4 98.9 0.9753 101.4 4522.6 4636.2 97.6% Strong 211.7 118.3 0.6990-30-18-01 1.318 1.82 2.38 -- 123.5 -- 102.6 107.6 0.9756 110.3 4622.7 4704.0 98.3% Strong 217.9 115.3 0.7090-30-22-01 1.470 2.14 2.02 -- 138.1 -- 116.8 119.7 0.9756 122.7 4610.7 4817.6 95.7% Strong 236.5 119.7 0.6985-30-17-01 1.259 1.70 2.70 -- 107.5 -- 98.8 97.5 0.9753 99.9 4384.2 4454.2 98.4% Weak 206.3 107.5 0.7285-30-21-01 1.433 2.07 2.20 -- 120.4 -- 110.4 107.5 0.9756 110.2 4246.2 4691.5 90.5% Weak 219.5 109.1 0.7285-30-25-01 1.398 2.52 1.81 -- 113.3 -- 108.0 104.2 0.9756 106.8 4222.4 4846.5 87.1% Weak 185.2 77.2 0.7385-30-50-01 2.617 5.07 1.68 470.1 145.1 325.0 114.6 130.6 0.9757 133.8 2825.0 4982.8 56.7% No Ign. 225.2 110.6 0.7185-30-64-01 2.611 6.10 1.39 483.4 138.0 345.4 114.1 130.6 0.9757 133.9 2832.6 5102.0 55.5% No Ign. 205.2 91.1 0.6685-30-64-03 2.707 6.21 1.37 537.2 131.3 405.9 110.2 116.7 0.9757 119.6 2441.3 5116.8 47.7% No Ign. 198.3 88.0 0.6987-30-44-01 2.382 4.45 1.86 476.9 118.6 358.3 109.6 105.2 0.9757 107.9 2501.7 5002.4 50.0% No Ign. 235.9 126.3 0.6790-30-40-02 2.166 4.00 2.01 521.2 88.8 432.4 86.7 93.1 0.9757 95.4 2432.6 5003.9 48.6% No Ign. 185.1 98.4 0.7590-30-35-02 1.992 3.60 2.23 462.8 76.1 386.7 75.4 78.8 0.9757 80.7 2238.8 4898.4 45.7% No Ign. 179.9 104.4 0.7394-30-35-01 1.970 3.57 2.15 527.0 82.2 444.8 83.1 83.1 0.9757 85.2 2387.9 5049.3 47.3% No Ign. 181.3 98.2 0.7594-30-40-01 2.199 4.07 1.89 626.2 98.3 528.0 98.0 102.4 0.9757 105.0 2637.7 5168.3 51.0% No Ign. 196.1 98.1 0.7598-30-30-01 1.776 3.10 2.38 637.8 77.6 560.2 78.1 83.0 0.9757 85.1 2646.7 5027.4 52.6% No Ign. 175.5 97.3 0.7685-50-31-01 1.808 3.11 2.73 379.8 156.5 223.3 157.3 151.5 0.9915 152.8 2801.8 4502.1 62.2% No Ign. 281.9 124.6 0.6887-30-60-01 2.741 5.90 1.40 653.9 250.6 403.4 232.4 236.5 0.9757 242.4 4886.2 5175.5 94.4% Strong 320.6 88.2 0.7390-30-50-01 2.616 5.04 1.59 512.2 243.3 268.9 222.8 224.9 0.9757 230.5 4866.1 5159.2 94.3% Strong 323.9 101.1 0.7490-30-50-02 2.621 5.06 1.58 636.8 226.1 410.7 227.7 229.7 0.9757 235.4 4963.1 5183.1 95.8% Strong 330.6 102.9 0.7394-30-45-01 2.392 4.53 1.70 865.9 194.6 671.3 195.5 200.3 0.9757 205.3 4741.5 5236.8 90.5% Strong 301.3 105.8 0.7298-30-35-01 1.968 3.57 2.06 709.8 168.7 541.1 161.7 170.2 0.9757 174.4 4896.9 5191.7 94.3% Strong 265.0 103.3 0.7398-30-40-01 2.171 4.00 1.84 743.9 185.3 558.6 182.6 190.6 0.9757 195.3 4971.5 5282.5 94.1% Strong 286.1 103.5 0.7385-50-64-01 2.664 6.24 1.36 678.8 397.0 281.8 380.6 392.1 0.9915 395.5 4921.9 5109.6 96.3% Strong 457.0 76.4 0.7287-30-52-01 2.679 5.09 1.62 597.7 182.7 415.0 164.0 166.2 0.9757 170.3 3512.3 5098.8 68.9% Weak 284.8 120.8 0.6987-30-49-01 2.546 4.84 1.70 571.9 170.3 401.6 150.6 153.3 0.9757 157.1 3408.4 5066.0 67.3% Weak 282.4 131.8 0.6590-30-35-01 2.021 3.66 2.19 329.5 107.2 222.3 91.7 97.2 0.9757 99.6 2722.4 4898.4 55.6% Weak 188.2 96.5 0.7690-30-40-01 2.259 4.20 1.91 402.7 141.0 261.6 128.1 136.2 0.9757 139.6 3414.0 5026.1 67.9% Weak 244.3 116.2 0.6990-30-44-01 2.388 4.48 1.79 578.3 215.0 363.3 198.4 200.7 0.9757 205.7 4760.1 5092.7 93.5% Weak 304.3 105.9 0.7385-50-37-01 2.052 3.68 2.31 496.9 269.4 227.5 273.6 264.8 0.9915 267.1 4314.0 4729.1 91.2% Weak 371.1 97.5 0.7685-50-50-01 2.588 5.00 1.70 629.0 379.6 249.4 362.3 371.0 0.9915 374.1 4791.5 4982.8 96.2% Weak 464.0 101.7 0.74

Assume γ = 1.22Pc Correction Biprop Performance

C*_th @ 500 psiFuel Injector Performance

Page 207: Thesis

192

Table F.3: Measured and calculated monopropellant test data.

Test Number p_ox_cv mdot_ox G p_cb_in p_cb_out ∆p_cb p_c1 p_c2 p_c2/ P_c2_tot C* C*_th η_C* Ttox_th Ttox(xx-xx-xx-xx) (psia) (lbm/s) (lbm/s/in^2) (psia) (psia) (psid) (psia) (psia) p_c2_tot (psia) (ft/s) (ft/s) (%) (F) (F)90-30-17-01 956.4 0.772 0.12 -- 42.1 -- 32.2 36.0 0.9751 36.9 2638.7 3083 85.6% 1393 102090-30-18-01 1080.0 0.820 0.12 -- 49.1 -- 40.7 43.9 0.9751 45.1 3036.8 3083 98.5% 1393 135290-30-22-01 1608.8 1.001 0.15 -- 59.3 -- 50.0 52.8 0.9751 54.2 2990.5 3083 97.0% 1393 131185-30-17-01 1023.1 0.791 0.12 -- 41.3 -- 34.5 36.9 0.9750 37.9 2646.6 2904 91.1% 1173 97485-30-21-01 1528.4 0.966 0.15 -- 46.9 -- 38.1 38.2 0.9750 39.2 2240.3 2904 77.1% 1173 69885-30-25-01 1639.2 1.000 0.15 -- 48.7 -- 40.2 40.5 0.9750 41.6 2295.8 2904 79.1% 1173 73385-30-50-01 1375.0 2.186 0.33 451.7 122.3 329.4 95.0 106.8 0.9750 109.5 2768.1 2904 95.3% 1173 106685-30-64-01 1501.0 2.285 0.35 515.9 129.0 386.9 103.8 114.1 0.9750 117.0 2828.2 2904 97.4% 1173 111385-30-64-03 1562.6 2.334 0.35 508.0 106.4 401.6 84.0 91.3 0.9751 93.7 2216.4 2904 76.3% 1173 68387-30-44-01 1072.3 1.943 0.29 461.6 96.9 364.7 85.7 83.8 0.9750 85.9 2443.4 2996 81.6% 1283 85390-30-40-02 1631.6 1.732 0.55 519.0 81.3 437.7 80.9 87.4 0.9751 89.6 2858.4 3083 92.7% 1393 119790-30-35-02 1317.3 1.559 0.50 462.9 69.7 393.3 70.5 74.1 0.9751 76.0 2693.5 3083 87.4% 1393 106394-30-35-01 1262.1 1.541 0.49 519.7 68.8 450.9 70.6 71.7 0.9752 73.6 2637.9 3217 82.0% 1570 105694-30-40-01 1626.5 1.744 0.56 616.4 84.3 532.1 84.9 90.4 0.9752 92.7 2936.4 3217 91.3% 1570 130898-30-30-01 936.6 1.341 0.43 623.6 66.4 557.3 67.1 72.7 0.9753 74.5 3070.3 3343 91.8% 1746 147385-50-31-01 1031.8 1.367 0.21 326.6 102.8 223.8 98.0 94.0 0.9912 94.8 2299.1 2904 79.2% 1173 73587-30-60-01 1561.9 2.343 0.35 612.6 132.2 480.5 109.2 118.3 0.9750 121.4 2861.3 2996 95.5% 1283 117090-30-50-01 1337.2 2.179 0.33 461.9 127.6 334.4 106.6 114.9 0.9751 117.8 2987.5 3083 96.9% 1393 130890-30-50-02 1345.0 2.191 0.70 622.0 104.1 517.9 104.2 112.2 0.9751 115.0 2900.7 3083 94.1% 1393 123394-30-45-01 1070.8 1.961 0.62 867.6 85.7 781.9 90.3 93.9 0.9752 96.2 2711.5 3217 84.3% 1570 111598-30-35-01 1238.5 1.536 0.49 691.5 78.3 613.2 78.5 84.7 0.9753 86.8 3122.6 3343 93.4% 1746 152398-30-40-01 1584.4 1.736 0.55 721.6 85.9 635.7 88.0 95.1 0.9753 97.5 3104.0 3343 92.8% 1746 150585-50-64-01 1510.5 2.295 0.35 549.3 196.0 353.3 184.2 188.7 0.9912 190.4 2749.3 2904 94.7% 1173 105187-30-52-01 1433.9 2.246 0.34 569.2 124.3 444.9 102.4 110.7 0.9750 113.5 2792.8 2996 93.2% 1283 111587-30-49-01 1265.4 2.110 0.32 538.1 111.6 426.5 91.0 96.6 0.9750 99.1 2595.0 2996 86.6% 1283 96390-30-35-01 709.3 1.585 0.24 325.6 92.3 233.3 77.1 83.0 0.9751 85.1 2965.9 3083 96.2% 1393 128990-30-40-01 929.1 1.820 0.28 374.7 103.7 271.0 85.6 94.2 0.9751 96.6 2934.0 3083 95.2% 1393 126290-30-44-01 1072.6 1.953 0.30 546.4 112.8 433.6 92.6 102.8 0.9751 105.5 2982.8 3083 96.8% 1393 130485-50-37-01 1449.9 1.618 0.24 412.0 127.3 284.7 124.7 115.8 0.9912 116.8 2393.6 2904 82.4% 1173 79785-50-50-01 1375.7 2.190 0.33 498.4 179.4 319.0 166.5 169.9 0.9912 171.4 2594.3 2904 89.3% 1173 936

Page 208: Thesis

193

Table F.4: Calculated decomposed gas and fue l flow conditions.

Fuel Orifice

Test Number γ MW_ox D_ox M_ox V_ox rho_ox M Vox rho_ox CD V_f Q tsl(xx-xx-xx-xx) (--) (lbm/lbm-mol) (in) (--) (ft/s) (lbm/ft^3) (--) (ft/s) (lbm/ft^3) (--) (ft/s) (--) (ms)90-30-17-01 1.265 22.11 1.71 0.461 946.4 0.051 0.465 941.1 0.052 0.800 118.4 17.7 0.07590-30-18-01 1.265 22.11 1.71 0.443 1006.3 0.051 0.447 1001.1 0.052 0.800 118.9 15.8 0.07190-30-22-01 1.265 22.11 1.71 0.445 998.8 0.063 0.448 993.6 0.063 0.800 119.2 13.1 0.07185-30-17-01 1.274 21.83 1.71 0.454 925.7 0.054 0.458 920.8 0.054 0.800 118.7 17.7 0.07785-30-21-01 1.274 21.83 1.71 0.482 883.0 0.069 0.486 877.7 0.069 0.800 118.5 15.2 0.08185-30-25-01 1.274 21.83 1.71 0.477 887.7 0.071 0.482 882.5 0.071 0.800 100.9 10.6 0.08085-30-50-01 1.274 21.83 1.71 0.447 941.6 0.146 0.451 936.7 0.147 0.800 125.4 6.2 0.07685-30-64-01 1.274 21.83 1.71 0.445 949.8 0.151 0.448 945.0 0.152 0.800 107.0 4.2 0.07585-30-64-03 1.274 21.83 1.71 0.484 881.1 0.167 0.489 875.8 0.168 0.800 109.2 4.7 0.08187-30-44-01 1.269 21.97 1.71 0.470 912.4 0.134 0.474 907.2 0.135 0.800 127.2 7.3 0.07890-30-40-02 1.265 22.11 1.71 0.450 978.2 0.111 0.454 972.9 0.112 0.800 126.0 7.6 0.07390-30-35-02 1.265 22.11 1.71 0.458 954.0 0.103 0.462 948.7 0.103 0.800 125.9 8.6 0.07594-30-35-01 1.258 22.33 1.71 0.466 959.5 0.101 0.469 953.9 0.102 0.800 125.5 8.6 0.07494-30-40-01 1.258 22.33 1.71 0.452 1005.5 0.109 0.455 1000.0 0.110 0.800 126.1 7.3 0.07198-30-30-01 1.251 22.56 1.71 0.451 1040.5 0.081 0.454 1034.8 0.082 0.800 126.0 9.2 0.06885-50-31-01 1.274 21.83 1.71 0.286 532.8 0.161 0.287 531.7 0.162 0.800 127.9 18.1 0.13487-30-60-01 1.269 21.97 1.71 0.447 967.2 0.152 0.451 962.2 0.153 0.800 115.5 4.7 0.07490-30-50-01 1.265 22.11 1.71 0.445 998.4 0.137 0.449 993.1 0.138 0.800 126.0 5.9 0.07190-30-50-02 1.265 22.11 1.71 0.449 984.7 0.140 0.452 979.4 0.141 0.800 125.7 5.9 0.07294-30-45-01 1.258 22.33 1.71 0.462 970.2 0.127 0.466 964.7 0.128 0.800 125.9 6.7 0.07398-30-35-01 1.251 22.56 1.71 0.449 1049.9 0.092 0.452 1044.1 0.093 0.800 125.3 7.8 0.06898-30-40-01 1.251 22.56 1.71 0.450 1046.6 0.104 0.453 1040.8 0.105 0.800 126.3 7.1 0.06885-50-64-01 1.274 21.83 1.71 0.269 563.4 0.256 0.270 562.4 0.257 0.800 107.0 7.1 0.12687-30-52-01 1.269 21.97 1.71 0.450 957.3 0.148 0.454 952.2 0.148 0.800 127.9 6.1 0.07487-30-49-01 1.269 21.97 1.71 0.460 930.6 0.143 0.464 925.4 0.143 0.800 126.7 6.6 0.07690-30-35-01 1.265 22.11 1.71 0.446 994.9 0.100 0.449 989.7 0.101 0.800 126.1 8.1 0.07290-30-40-01 1.265 22.11 1.71 0.447 989.9 0.116 0.451 984.6 0.116 0.800 126.4 7.1 0.07290-30-44-01 1.265 22.11 1.71 0.445 997.6 0.123 0.449 992.4 0.124 0.800 126.6 6.6 0.07185-50-37-01 1.274 21.83 1.71 0.282 538.2 0.189 0.283 537.1 0.190 0.800 127.5 15.0 0.13285-50-50-01 1.274 21.83 1.71 0.274 551.6 0.250 0.275 550.5 0.250 0.800 125.4 10.5 0.129

Incompressible Calculations Compressible Calculations

Page 209: Thesis

194

Table F.5: FFT results for bipropellant portion of each test including maximum range of pressure oscillations.

Test Number Frequency Peak-to-Peak Frequency Peak-to-Peak Frequency Peak-to-Peak Frequency Peak-to-Peak Frequency Peak-to-Peak Standard Dev. ζ(xx-xx-xx-xx) (Hz) (psid) (Hz) (psid) (Hz) (psid) (Hz) (psid) (Hz) (psid) (psid) (%)90-30-17-01 2.0 200.9 2.0 24.2 2.0 22.7 21.5 14.4 2.0 13.5 2.2 1.1%90-30-18-01 2.0 0.8 2.0 26.8 31.3 17.1 31.3 13.2 31.3 12.2 2.7 1.3%90-30-22-01 29.3 3016.1 29.3 118.0 29.3 140.7 29.3 121.1 29.3 117.9 38.7 16.2%85-30-17-01 2.0 17.3 25.4 113.8 25.4 154.5 25.4 137.0 25.4 134.5 41.6 21.3%85-30-21-01 2.0 20.4 29.3 164.2 29.3 186.2 29.3 175.7 29.3 165.2 46.7 21.7%85-30-25-01 2.0 28.2 31.3 184.3 31.3 165.4 31.3 165.4 31.3 146.1 42.4 20.3%85-30-50-01 33.2 1321.5 33.2 455.3 33.2 297.8 33.2 331.5 33.2 356.3 79.4 30.4%85-30-64-01 34.2 1734.8 34.2 492.8 34.2 307.5 34.2 330.0 34.2 438.0 77.2 29.6%85-30-64-03 39.1 1892.4 39.1 437.6 39.1 280.1 39.1 255.2 39.1 303.2 56.9 24.4%87-30-44-01 37.1 1887.0 37.1 438.8 37.1 247.8 37.1 243.3 37.1 193.7 59.2 28.1%90-30-40-02 123.1 38.3 1.0 6.8 1.0 4.5 350.6 5.3 350.6 9.8 1.9 1.0%90-30-35-02 147.5 54.8 1.0 10.5 1.0 3.0 1.0 3.8 406.3 7.5 1.2 0.8%94-30-35-01 82.0 82.5 1.0 15.0 1.0 4.5 82.0 6.8 348.6 7.5 1.5 0.9%94-30-40-01 146.5 67.5 1.0 14.3 1.0 7.5 1.0 8.3 1.0 9.8 1.7 0.8%98-30-30-01 93.8 64.5 1.0 11.3 1.0 5.3 1.0 8.3 1.0 7.5 1.1 0.6%85-50-31-01 27.3 2139.7 27.3 768.0 27.3 313.1 27.3 324.3 27.3 292.5 92.4 30.5%87-30-60-01 365.2 239.0 230.5 18.0 230.5 71.5 230.5 22.0 230.5 25.5 6.1 1.3%90-30-50-01 294.4 453.8 0.5 29.3 221.2 48.8 221.2 61.5 221.2 94.5 14.1 3.1%90-30-50-02 1.0 27.8 1.0 105.0 1.0 18.8 1.0 22.5 1.0 26.3 3.8 0.8%94-30-45-01 245.2 97.5 1.0 9.8 1.0 11.3 1.0 21.8 245.1 26.3 0.2 0.1%98-30-35-01 160.2 78.8 1.9 9.8 1.0 10.5 1.0 13.5 489.3 25.5 5.1 1.5%98-30-40-01 211.9 81.8 1.0 12.8 1.0 12.8 2.9 18.0 398.4 57.8 5.6 1.5%85-50-64-01 308.6 270.8 2.0 44.3 2.0 77.3 2.0 51.0 2.0 62.3 9.5 1.2%87-30-52-01 35.2 1023.7 35.2 260.5 35.2 250.4 35.2 242.4 35.2 286.8 78.3 23.6%87-30-49-01 35.2 949.7 35.2 352.0 35.2 299.4 35.2 297.4 35.2 271.9 75.7 24.7%90-30-35-01 233.4 108.8 30.3 7.5 31.3 6.0 31.3 3.0 31.3 4.5 0.7 0.3%90-30-40-01 30.8 936.0 30.8 273.0 30.8 310.5 30.8 633.0 30.8 311.3 89.6 32.9%90-30-44-01 335.9 165.4 1.0 37.4 257.8 59.2 257.8 62.1 36.1 37.8 5.9 1.5%85-50-37-01 43.0 612.5 43.0 290.3 43.0 262.2 43.0 250.6 43.0 246.1 74.6 14.1%85-50-50-01 31.3 369.8 31.3 209.3 31.3 225.8 31.3 181.5 31.3 214.5 29.3 3.9%

p_c2p_c1p_fu_inj p_cb_outp_cb_in

Page 210: Thesis

195

Table F.6: FFT result for monopropellant section of each test including maximum range of pressure oscillations.

Test Number Frequency Peak-to-Peak Frequency Peak-to-Peak Frequency Peak-to-Peak Frequency Peak-to-Peak Frequency Peak-to-Peak Standard Dev. ζ(xx-xx-xx-xx) (Hz) (psid) (Hz) (psid) (Hz) (psid) (Hz) (psid) (Hz) (psid) (psid) (%)90-30-17-01 3.9 0.9 3.9 1.7 3.9 2.1 3.9 1.7 3.9 1.8 0.4 0.6%90-30-18-01 2.0 9.5 25.4 5.0 23.4 7.8 25.4 6.5 23.4 7.3 1.7 2.0%90-30-22-01 2.0 1.1 27.3 28.2 27.3 50.4 27.3 41.6 27.3 43.0 6.4 6.1%85-30-17-01 3.9 0.9 27.3 9.8 27.3 17.1 27.3 13.1 27.3 14.7 2.3 3.2%85-30-21-01 3.9 3.6 31.3 37.8 31.3 82.7 31.3 65.4 31.3 63.6 19.1 24.9%85-30-25-01 3.9 5.5 31.3 37.4 31.3 81.8 31.3 62.0 31.3 60.5 18.4 22.7%85-30-50-01 3.9 192.0 3.9 14.3 3.9 20.3 3.9 18.0 3.9 21.0 3.7 1.7%85-30-64-01 332.0 210.8 2.0 6.8 2.0 6.8 2.0 7.5 2.0 9.8 2.0 0.9%85-30-64-03 43.0 1694.9 43.0 67.5 43.0 152.3 43.0 109.6 43.0 133.2 41.8 22.9%87-30-44-01 39.1 1581.2 39.1 117.0 39.1 187.4 39.1 131.4 39.1 126.7 37.1 22.2%90-30-40-02 85.9 63.8 2.0 5.3 2.0 6.0 85.9 6.8 85.9 9.0 1.8 1.0%90-30-35-02 148.4 46.5 2.0 3.0 2.0 4.5 2.0 3.8 2.0 6.0 1.1 0.7%94-30-35-01 2.0 81.8 2.0 10.5 2.0 10.5 2.0 11.3 2.0 13.5 2.5 1.7%94-30-40-01 148.4 75.8 2.0 6.8 2.0 6.0 2.0 6.0 2.0 8.3 1.5 0.9%98-30-30-01 3.9 63.8 3.9 7.5 3.9 7.5 3.9 8.3 3.9 9.0 1.8 1.3%85-50-31-01 37.1 2203.4 37.1 93.2 37.1 174.0 37.1 151.0 37.1 136.4 39.6 21.1%87-30-60-01 43.0 413.1 43.0 39.9 43.0 64.2 43.0 51.8 43.0 56.9 13.7 5.8%90-30-50-01 367.2 320.3 2.0 22.5 2.0 5.3 2.0 5.3 2.0 5.3 1.2 0.5%90-30-50-02 191.4 24.0 2.0 3.0 2.0 3.8 2.0 5.3 418.0 7.5 1.4 0.6%94-30-45-01 255.9 92.3 2.0 6.0 2.0 7.5 2.0 8.3 2.0 8.3 0.3 0.1%98-30-35-01 3.9 86.3 3.9 6.8 3.9 7.5 3.9 7.5 3.9 8.3 1.5 0.9%98-30-40-01 2.0 82.5 2.0 9.8 2.0 9.8 2.0 9.8 2.0 11.3 2.0 1.1%85-50-64-01 345.7 248.3 2.0 14.3 2.0 13.5 2.0 13.5 2.0 15.8 3.9 1.0%87-30-52-01 39.1 659.6 39.1 74.9 39.1 134.4 39.1 105.9 39.1 116.6 31.9 14.4%87-30-49-01 37.1 900.6 37.1 85.4 37.1 164.8 37.1 138.2 37.1 131.7 39.6 20.5%90-30-35-01 238.3 87.8 2.0 17.3 2.0 4.5 2.0 5.3 2.0 6.0 1.1 0.7%90-30-40-01 343.8 226.5 2.0 8.3 2.0 10.5 2.0 10.5 2.0 10.5 2.4 1.3%90-30-44-01 37.1 115.3 37.1 7.4 37.1 11.1 37.1 10.2 37.1 12.0 2.7 1.3%85-50-37-01 46.9 2881.4 46.9 90.8 46.9 184.1 46.9 160.2 46.9 138.9 42.2 18.2%85-50-50-01 37.1 498.0 37.1 77.3 37.1 102.0 37.1 96.0 37.1 100.5 18.7 5.5%

p_cb_in p_fu_inj p_cb_out p_c1 p_c2

Page 211: Thesis

196

Table F.7: Calculated uncertainty in densities and mass flow rates.

Test Number uT_ox uW urho_ox drho_ox uT_f urho_f drho_f uCD_ox_cv udt_ox_cv up_ox_cv umdot_ox dmdot_ox uCD_f_cv udt_f_cv up_fu_cv umdot_f dmdot_f(xx-xx-xx-xx) (--) (--) (%) (lbm/ft^3) (--) (%) (lbm/ft^3) (--) (--) (--) (%) (lbm/s) (--) (--) (--) (%) (lbm/s)90-30-17-01 2.51E-02 1.10E-03 0.1% 0.081 -- 0.0% -- 8.34E-03 0.00E+00 7.81E-03 0.9% 0.007 8.29E-03 0.00E+00 4.46E-03 0.9% 0.00490-30-18-01 2.73E-02 1.12E-03 0.1% 0.080 -- 0.0% -- 8.34E-03 0.00E+00 6.46E-03 0.9% 0.008 8.29E-03 0.00E+00 4.42E-03 0.9% 0.00490-30-22-01 2.77E-02 1.12E-03 0.1% 0.080 -- 0.0% -- 8.34E-03 0.00E+00 4.66E-03 0.9% 0.009 8.29E-03 0.00E+00 4.40E-03 0.9% 0.00485-30-17-01 2.71E-02 1.18E-03 0.1% 0.079 -- 0.0% -- 8.34E-03 0.00E+00 7.31E-03 0.9% 0.007 8.29E-03 0.00E+00 4.43E-03 0.9% 0.00485-30-21-01 2.78E-02 1.18E-03 0.1% 0.078 -- 0.0% -- 8.34E-03 0.00E+00 4.90E-03 0.9% 0.008 8.29E-03 0.00E+00 4.45E-03 0.9% 0.00485-30-25-01 2.73E-02 1.18E-03 0.1% 0.079 -- 0.0% -- 8.34E-03 0.00E+00 4.57E-03 0.9% 0.009 8.29E-03 0.00E+00 6.14E-03 0.9% 0.00485-30-50-01 2.42E-02 1.17E-03 0.1% 0.079 6.90E-02 0.3% 0.139 7.08E-03 0.00E+00 5.46E-03 0.8% 0.017 8.29E-03 0.00E+00 4.54E-03 0.9% 0.00485-30-64-01 2.51E-02 1.17E-03 0.1% 0.079 6.90E-02 0.3% 0.139 7.08E-03 0.00E+00 5.19E-03 0.8% 0.017 8.29E-03 0.00E+00 6.24E-03 0.9% 0.00385-30-64-03 2.71E-02 1.17E-03 0.1% 0.079 6.90E-02 0.3% 0.139 7.08E-03 0.00E+00 4.81E-03 0.7% 0.017 8.29E-03 0.00E+00 5.99E-03 0.9% 0.00387-30-44-01 2.77E-02 1.14E-03 0.1% 0.080 6.90E-02 0.3% 0.139 7.08E-03 0.00E+00 6.99E-03 0.8% 0.015 8.29E-03 0.00E+00 4.42E-03 0.9% 0.00490-30-40-02 2.58E-02 1.12E-03 0.1% 0.080 6.90E-02 0.3% 0.139 7.04E-03 0.00E+00 4.59E-03 0.7% 0.013 8.29E-03 0.00E+00 4.50E-03 0.9% 0.00490-30-35-02 2.54E-02 1.10E-03 0.1% 0.081 6.90E-02 0.3% 0.139 7.04E-03 0.00E+00 5.69E-03 0.8% 0.012 8.29E-03 0.00E+00 4.51E-03 0.9% 0.00494-30-35-01 2.94E-02 1.06E-03 0.1% 0.082 6.90E-02 0.3% 0.139 7.04E-03 0.00E+00 5.96E-03 0.8% 0.012 8.29E-03 0.00E+00 4.54E-03 0.9% 0.00494-30-40-01 2.47E-02 1.06E-03 0.1% 0.082 6.90E-02 0.3% 0.139 7.04E-03 0.00E+00 4.50E-03 0.7% 0.013 8.29E-03 0.00E+00 4.50E-03 0.9% 0.00498-30-30-01 3.52E-02 1.02E-03 0.1% 0.083 6.90E-02 0.3% 0.139 7.04E-03 0.00E+00 7.99E-03 0.8% 0.011 8.29E-03 0.00E+00 4.50E-03 0.9% 0.00485-50-31-01 2.77E-02 1.17E-03 0.1% 0.079 6.90E-02 0.3% 0.139 7.04E-03 0.00E+00 7.26E-03 0.8% 0.011 8.29E-03 0.00E+00 4.37E-03 0.9% 0.00487-30-60-01 2.75E-02 1.15E-03 0.1% 0.079 6.90E-02 0.3% 0.139 7.08E-03 0.00E+00 4.80E-03 0.7% 0.018 8.29E-03 0.00E+00 5.36E-03 0.9% 0.00490-30-50-01 2.66E-02 1.11E-03 0.1% 0.080 6.90E-02 0.3% 0.139 7.08E-03 0.00E+00 5.59E-03 0.8% 0.017 8.29E-03 0.00E+00 4.50E-03 0.9% 0.00490-30-50-02 2.95E-02 1.10E-03 0.1% 0.080 6.90E-02 0.3% 0.139 7.08E-03 0.00E+00 5.59E-03 0.8% 0.017 8.29E-03 0.00E+00 4.53E-03 0.9% 0.00494-30-45-01 2.44E-02 1.07E-03 0.1% 0.082 6.90E-02 0.3% 0.139 7.08E-03 0.00E+00 7.02E-03 0.8% 0.016 8.29E-03 0.00E+00 4.51E-03 0.9% 0.00498-30-35-01 2.62E-02 1.02E-03 0.1% 0.083 6.90E-02 0.3% 0.139 7.04E-03 0.00E+00 6.05E-03 0.8% 0.012 8.29E-03 0.00E+00 4.55E-03 0.9% 0.00498-30-40-01 2.48E-02 1.02E-03 0.1% 0.083 6.90E-02 0.3% 0.139 7.04E-03 0.00E+00 4.73E-03 0.7% 0.013 8.29E-03 0.00E+00 4.48E-03 0.9% 0.00485-50-64-01 2.51E-02 1.16E-03 0.1% 0.079 6.90E-02 0.3% 0.139 7.08E-03 0.00E+00 4.96E-03 0.8% 0.017 8.29E-03 0.00E+00 6.24E-03 0.9% 0.00387-30-52-01 2.75E-02 1.14E-03 0.1% 0.079 6.90E-02 0.3% 0.139 7.08E-03 0.00E+00 5.26E-03 0.8% 0.017 8.29E-03 0.00E+00 4.37E-03 0.9% 0.00487-30-49-01 2.69E-02 1.14E-03 0.1% 0.080 6.90E-02 0.3% 0.139 7.08E-03 0.00E+00 5.93E-03 0.8% 0.016 8.29E-03 0.00E+00 4.45E-03 0.9% 0.00490-30-35-01 2.67E-02 1.12E-03 0.1% 0.080 6.90E-02 0.3% 0.139 7.08E-03 0.00E+00 1.06E-02 0.9% 0.014 8.29E-03 0.00E+00 4.49E-03 0.9% 0.00490-30-40-01 3.13E-02 1.11E-03 0.1% 0.080 6.90E-02 0.3% 0.139 7.08E-03 0.00E+00 8.04E-03 0.8% 0.015 8.29E-03 0.00E+00 4.47E-03 0.9% 0.00490-30-44-01 2.92E-02 1.11E-03 0.1% 0.080 6.90E-02 0.3% 0.139 7.08E-03 0.00E+00 7.00E-03 0.8% 0.015 8.29E-03 0.00E+00 4.46E-03 0.9% 0.00485-50-37-01 2.590E-02 1.179E-03 0.1% 0.079 6.90E-02 0.3% 0.139 7.04E-03 0.00E+00 5.20E-03 0.8% 0.012 8.29E-03 0.00E+00 4.40E-03 0.9% 0.00485-50-50-01 2.514E-02 1.163E-03 0.1% 0.079 6.90E-02 0.3% 0.139 7.08E-03 0.00E+00 5.62E-03 0.8% 0.016 8.29E-03 0.00E+00 4.55E-03 0.9% 0.004

Page 212: Thesis

197

Table F.8: Calculated uncertainty in equivalence ratio and bipropellant performance parameters.

Test Number uφ dφ umdot_tot up_c2_tot uD_th uC* dC* dC*_th uC*_th uη_C* dη_C*(xx-xx-xx-xx) (%) (--) (--) (--) (--) (%) (ft/s) (ft/s) (--) (%) (%)90-30-17-01 1.3% 0.033 6.601E-03 7.393E-02 3.383E-03 7.5% 337.1 20.0 4.314E-03 7.47% 7.28%90-30-18-01 1.2% 0.030 6.532E-03 6.972E-02 3.383E-03 7.0% 325.2 20.0 4.252E-03 7.05% 6.93%90-30-22-01 1.2% 0.025 6.511E-03 6.266E-02 3.383E-03 6.3% 292.1 20.0 4.151E-03 6.35% 6.08%85-30-17-01 1.3% 0.034 6.561E-03 7.696E-02 3.383E-03 7.8% 339.9 20.0 4.490E-03 7.77% 7.64%85-30-21-01 1.2% 0.027 6.505E-03 6.979E-02 3.383E-03 7.0% 299.0 20.0 4.263E-03 7.05% 6.38%85-30-25-01 1.2% 0.022 6.692E-03 7.196E-02 3.383E-03 7.3% 306.5 20.0 4.127E-03 7.27% 6.33%85-30-50-01 1.2% 0.019 6.507E-03 5.744E-02 3.383E-03 5.8% 164.4 20.0 4.014E-03 5.83% 3.31%85-30-64-01 1.2% 0.016 6.610E-03 5.741E-02 3.383E-03 5.8% 164.8 20.0 3.920E-03 5.83% 3.24%85-30-64-03 1.2% 0.016 6.569E-03 6.426E-02 3.383E-03 6.5% 158.6 20.0 3.909E-03 6.51% 3.10%87-30-44-01 1.2% 0.022 6.649E-03 7.127E-02 3.383E-03 7.2% 179.9 20.0 3.998E-03 7.20% 3.60%90-30-40-02 1.1% 0.023 6.187E-03 8.059E-02 3.383E-03 8.1% 197.3 20.0 3.997E-03 8.12% 3.95%90-30-35-02 1.2% 0.026 6.249E-03 9.520E-02 3.383E-03 9.6% 214.1 20.0 4.083E-03 9.57% 4.38%94-30-35-01 1.2% 0.025 6.279E-03 9.026E-02 3.383E-03 9.1% 216.7 20.0 3.961E-03 9.08% 4.30%94-30-40-01 1.1% 0.022 6.190E-03 7.321E-02 3.383E-03 7.4% 194.6 20.0 3.870E-03 7.39% 3.77%98-30-30-01 1.2% 0.028 6.488E-03 9.035E-02 3.383E-03 9.1% 240.4 20.0 3.978E-03 9.09% 4.79%85-50-31-01 1.2% 0.032 6.366E-03 4.950E-02 4.367E-03 5.1% 142.0 20.0 4.442E-03 5.09% 3.17%87-30-60-01 1.2% 0.016 6.529E-03 3.171E-02 3.383E-03 3.3% 161.6 20.0 3.864E-03 3.33% 3.14%90-30-50-01 1.2% 0.018 6.522E-03 3.335E-02 3.383E-03 3.5% 168.6 20.0 3.877E-03 3.49% 3.29%90-30-50-02 1.2% 0.018 6.526E-03 3.265E-02 3.383E-03 3.4% 168.6 20.0 3.859E-03 3.42% 3.27%94-30-45-01 1.2% 0.020 6.670E-03 3.745E-02 3.383E-03 3.9% 183.2 20.0 3.819E-03 3.88% 3.52%98-30-35-01 1.2% 0.024 6.293E-03 4.407E-02 3.383E-03 4.5% 220.5 20.0 3.852E-03 4.52% 4.26%98-30-40-01 1.1% 0.021 6.203E-03 3.935E-02 3.383E-03 4.0% 200.9 20.0 3.786E-03 4.06% 3.82%85-50-64-01 1.2% 0.016 6.594E-03 1.913E-02 4.367E-03 2.2% 108.5 20.0 3.914E-03 2.24% 2.16%87-30-52-01 1.2% 0.019 6.482E-03 4.513E-02 3.383E-03 4.6% 161.9 20.0 3.922E-03 4.63% 3.19%87-30-49-01 1.2% 0.020 6.543E-03 4.894E-02 3.383E-03 5.0% 169.9 20.0 3.948E-03 5.00% 3.36%90-30-35-01 1.2% 0.027 7.190E-03 7.720E-02 3.383E-03 7.8% 211.9 20.0 4.083E-03 7.79% 4.33%90-30-40-01 1.2% 0.023 6.792E-03 5.507E-02 3.383E-03 5.6% 190.8 20.0 3.979E-03 5.60% 3.81%90-30-44-01 1.2% 0.021 6.659E-03 3.737E-02 3.383E-03 3.9% 183.5 20.0 3.927E-03 3.88% 3.62%85-50-37-01 1.1% 0.027 6.199E-03 2.832E-02 4.367E-03 3.0% 130.6 20.0 4.229E-03 3.06% 2.79%85-50-50-01 1.2% 0.020 6.522E-03 2.022E-02 4.367E-03 2.3% 110.1 20.0 4.014E-03 2.33% 2.24%

Page 213: Thesis

198

Table F.9: Calculated uncertainty in monopropellant performance parameters.

Test Number up_cb_in up_cb_out u∆p_cb d∆p_cb up_c2_tot uC* dC* dC*_th uC*_th uη_C* dη_C* uTtox_th uTtox dTtox(xx-xx-xx-xx) (--) (--) (%) (psid) (--) (%) (ft/s) (ft/s) (--) (%) (%) (--) (%) (°F)90-30-17-01 -- 1.784E-01 -- -- 20.34% 20.4% 537.5 5.0 1.622E-03 20.4% 17.4% 7.179E-03 40.7% 415.890-30-18-01 -- 1.526E-01 -- -- 16.64% 16.7% 506.6 5.0 1.622E-03 16.7% 16.4% 7.179E-03 33.4% 451.090-30-22-01 -- 1.265E-01 -- -- 13.85% 13.9% 415.4 5.0 1.622E-03 13.9% 13.5% 7.179E-03 27.8% 364.285-30-17-01 -- 1.816E-01 -- -- 19.80% 19.8% 524.9 5.0 1.722E-03 19.8% 18.1% 8.525E-03 39.7% 386.685-30-21-01 -- 1.601E-01 -- -- 19.14% 19.2% 429.6 5.0 1.722E-03 19.2% 14.8% 8.525E-03 38.4% 267.885-30-25-01 -- 1.540E-01 -- -- 18.04% 18.1% 415.0 5.0 1.722E-03 18.1% 14.3% 8.525E-03 36.2% 265.185-30-50-01 1.661E-02 6.134E-02 3.22% 10.6 6.85% 6.9% 191.6 5.0 1.722E-03 6.9% 6.6% 8.525E-03 13.9% 147.985-30-64-01 1.454E-02 5.814E-02 2.74% 10.6 6.41% 6.5% 183.6 5.0 1.722E-03 6.5% 6.3% 8.525E-03 13.0% 144.885-30-64-03 1.476E-02 7.052E-02 2.64% 10.6 8.01% 8.1% 178.9 5.0 1.722E-03 8.1% 6.2% 8.525E-03 16.2% 110.587-30-44-01 1.625E-02 7.742E-02 2.91% 10.6 8.73% 8.8% 214.7 5.0 1.669E-03 8.8% 7.2% 7.794E-03 17.6% 150.290-30-40-02 1.445E-02 9.222E-02 2.42% 10.6 8.37% 8.4% 241.0 5.0 1.622E-03 8.4% 7.8% 7.179E-03 16.9% 202.190-30-35-02 1.620E-02 1.077E-01 2.70% 10.6 9.87% 9.9% 267.2 5.0 1.622E-03 9.9% 8.7% 7.179E-03 19.9% 211.194-30-35-01 1.443E-02 1.091E-01 2.35% 10.6 10.20% 10.2% 270.3 5.0 1.554E-03 10.2% 8.4% 6.369E-03 20.5% 216.594-30-40-01 1.217E-02 8.900E-02 1.99% 10.6 8.09% 8.2% 239.4 5.0 1.554E-03 8.2% 7.4% 6.369E-03 16.3% 213.598-30-30-01 1.203E-02 1.130E-01 1.90% 10.6 10.06% 10.1% 310.7 5.0 1.496E-03 10.1% 9.3% 5.727E-03 20.2% 298.285-50-31-01 2.297E-02 7.299E-02 4.74% 10.6 7.91% 8.0% 183.9 5.0 1.722E-03 8.0% 6.3% 8.525E-03 16.0% 117.887-30-60-01 1.224E-02 5.675E-02 2.21% 10.6 6.18% 6.3% 179.2 5.0 1.669E-03 6.3% 6.0% 7.794E-03 12.6% 146.990-30-50-01 1.624E-02 5.879E-02 3.17% 10.6 6.37% 6.4% 192.6 5.0 1.622E-03 6.4% 6.2% 7.179E-03 12.9% 169.090-30-50-02 1.206E-02 7.206E-02 2.05% 10.6 6.52% 6.6% 191.4 5.0 1.622E-03 6.6% 6.2% 7.179E-03 13.2% 163.194-30-45-01 8.644E-03 8.752E-02 1.36% 10.6 7.79% 7.9% 213.2 5.0 1.554E-03 7.9% 6.6% 6.369E-03 15.7% 175.698-30-35-01 1.085E-02 9.575E-02 1.73% 10.6 8.64% 8.7% 271.7 5.0 1.496E-03 8.7% 8.1% 5.727E-03 17.4% 265.298-30-40-01 1.039E-02 8.727E-02 1.67% 10.6 7.69% 7.8% 240.8 5.0 1.496E-03 7.8% 7.2% 5.727E-03 15.5% 233.785-50-64-01 1.365E-02 3.827E-02 3.00% 10.6 3.94% 4.1% 112.9 5.0 1.722E-03 4.1% 3.9% 8.525E-03 8.3% 86.887-30-52-01 1.318E-02 6.033E-02 2.38% 10.6 6.61% 6.7% 186.7 5.0 1.669E-03 6.7% 6.2% 7.794E-03 13.4% 149.387-30-49-01 1.394E-02 6.723E-02 2.49% 10.6 7.57% 7.6% 198.1 5.0 1.669E-03 7.6% 6.6% 7.794E-03 15.3% 147.290-30-35-01 2.304E-02 8.124E-02 4.55% 10.6 8.81% 8.9% 263.4 5.0 1.622E-03 8.9% 8.5% 7.179E-03 17.8% 229.290-30-40-01 2.002E-02 7.232E-02 3.91% 10.6 7.76% 7.8% 229.8 5.0 1.622E-03 7.8% 7.5% 7.179E-03 15.7% 197.990-30-44-01 1.373E-02 6.648E-02 2.45% 10.6 7.11% 7.2% 214.4 5.0 1.622E-03 7.2% 7.0% 7.179E-03 14.4% 187.785-50-37-01 1.820E-02 5.892E-02 3.73% 10.6 6.42% 6.5% 156.1 5.0 1.722E-03 6.5% 5.4% 8.525E-03 13.1% 104.285-50-50-01 1.505E-02 4.181E-02 3.32% 10.6 4.38% 4.5% 117.4 5.0 1.722E-03 4.5% 4.0% 8.525E-03 9.1% 85.2

Page 214: Thesis

199

Table F.10: Calculated uncertainties in fuel and oxidizer flow parameter as well as momentum ratio and residence time.

Test Number uCD ud_o uV_f dV_f urho_ox_g drho_ox_g uD_ox uV_ox dV_ox uM_ox dM_ox uQ dQ utau_sl dtau_sl uCR dCR(xx-xx-xx-xx) (--) (--) (%) (--) (%) (lbm/ft^3) (--) (%) (ft/s) (%) (--) (%) (--) (%) (sec) (%) (--)90-30-17-01 6.25E-02 0.0 6.31% 7.47 45.5% 0.023 2.929E-03 45.6% 431.1 28.8% 0.133 102.6% 18.13 45.6% 0.034 0.8% 0.02390-30-18-01 6.25E-02 0.0 6.31% 7.50 37.3% 0.019 2.929E-03 37.3% 375.4 23.6% 0.105 84.4% 13.33 37.3% 0.026 0.8% 0.02390-30-22-01 6.25E-02 0.0 6.31% 7.52 31.0% 0.020 2.929E-03 31.1% 310.3 19.6% 0.087 70.6% 9.24 31.1% 0.022 0.8% 0.02385-30-17-01 6.25E-02 0.0 6.31% 7.49 44.3% 0.024 2.929E-03 44.4% 410.6 28.1% 0.127 100.0% 17.74 44.4% 0.034 0.8% 0.02385-30-21-01 6.25E-02 0.0 6.31% 7.48 42.9% 0.030 2.929E-03 42.9% 378.7 27.1% 0.131 96.7% 14.67 42.9% 0.035 0.8% 0.02385-30-25-01 6.25E-02 0.0 6.31% 6.37 40.4% 0.029 2.929E-03 40.4% 358.9 25.6% 0.122 91.3% 9.64 40.4% 0.032 0.8% 0.02385-30-50-01 6.25E-02 0.0 6.32% 7.92 15.5% 0.023 2.929E-03 15.5% 146.0 9.8% 0.044 36.9% 2.27 15.5% 0.012 0.8% 0.02385-30-64-01 6.25E-02 0.0 6.32% 6.76 14.5% 0.022 2.929E-03 14.5% 138.1 9.2% 0.041 34.9% 1.48 14.6% 0.011 0.8% 0.02385-30-64-03 6.25E-02 0.0 6.32% 6.90 18.0% 0.030 2.929E-03 18.1% 159.2 11.4% 0.055 42.3% 1.98 18.1% 0.015 0.8% 0.02387-30-44-01 6.25E-02 0.0 6.32% 8.03 19.6% 0.026 2.929E-03 19.7% 179.4 12.4% 0.058 45.7% 3.36 19.7% 0.015 0.8% 0.02390-30-40-02 6.25E-02 0.0 6.32% 7.96 18.8% 0.021 2.929E-03 18.9% 184.5 11.9% 0.054 44.0% 3.32 18.9% 0.014 0.8% 0.02390-30-35-02 6.25E-02 0.0 6.32% 7.95 22.2% 0.023 2.929E-03 22.2% 211.7 14.0% 0.064 51.2% 4.39 22.2% 0.017 0.8% 0.02394-30-35-01 6.25E-02 0.0 6.32% 7.93 22.9% 0.023 2.929E-03 22.9% 219.9 14.5% 0.067 52.8% 4.53 22.9% 0.017 0.8% 0.02394-30-40-01 6.25E-02 0.0 6.32% 7.96 18.2% 0.020 2.929E-03 18.2% 183.4 11.5% 0.052 42.7% 3.12 18.3% 0.013 0.8% 0.02398-30-30-01 6.25E-02 0.0 6.32% 7.96 22.6% 0.018 2.929E-03 22.6% 235.5 14.3% 0.064 52.2% 4.78 22.7% 0.015 0.8% 0.02385-50-31-01 6.25E-02 0.0 6.32% 8.08 17.9% 0.029 2.929E-03 17.9% 95.3 11.3% 0.032 41.9% 7.58 17.9% 0.024 1.0% 0.04887-30-60-01 6.25E-02 0.0 6.32% 7.30 14.0% 0.021 2.929E-03 14.0% 135.6 8.9% 0.040 33.8% 1.60 14.1% 0.010 0.8% 0.02390-30-50-01 6.25E-02 0.0 6.32% 7.96 14.4% 0.020 2.929E-03 14.4% 144.1 9.1% 0.041 34.6% 2.04 14.5% 0.010 0.8% 0.02390-30-50-02 6.25E-02 0.0 6.32% 7.94 14.7% 0.021 2.929E-03 14.8% 145.5 9.3% 0.042 35.4% 2.09 14.8% 0.011 0.8% 0.02394-30-45-01 6.25E-02 0.0 6.32% 7.95 17.6% 0.022 2.929E-03 17.6% 170.7 11.1% 0.051 41.3% 2.77 17.6% 0.013 0.8% 0.02398-30-35-01 6.25E-02 0.0 6.32% 7.92 19.4% 0.018 2.929E-03 19.5% 204.3 12.3% 0.055 45.3% 3.55 19.5% 0.013 0.8% 0.02398-30-40-01 6.25E-02 0.0 6.32% 7.98 17.3% 0.018 2.929E-03 17.4% 181.6 11.0% 0.049 40.8% 2.89 17.4% 0.012 0.8% 0.02385-50-64-01 6.25E-02 0.0 6.32% 6.76 9.2% 0.023 2.929E-03 9.2% 51.8 5.8% 0.016 24.1% 1.72 9.3% 0.012 1.0% 0.04887-30-52-01 6.25E-02 0.0 6.32% 8.08 14.9% 0.022 2.929E-03 15.0% 143.3 9.5% 0.043 35.8% 2.19 15.0% 0.011 0.8% 0.02387-30-49-01 6.25E-02 0.0 6.32% 8.00 17.1% 0.024 2.929E-03 17.1% 159.1 10.8% 0.050 40.2% 2.65 17.1% 0.013 0.8% 0.02390-30-35-01 6.25E-02 0.0 6.32% 7.96 19.8% 0.020 2.929E-03 19.9% 197.7 12.6% 0.056 46.2% 3.75 19.9% 0.014 0.8% 0.02390-30-40-01 6.25E-02 0.0 6.32% 7.98 17.5% 0.020 2.929E-03 17.5% 173.5 11.1% 0.050 41.2% 2.94 17.6% 0.013 0.8% 0.02390-30-44-01 6.25E-02 0.0 6.32% 8.00 16.1% 0.020 2.929E-03 16.1% 160.5 10.2% 0.045 38.1% 2.52 16.1% 0.012 0.8% 0.02385-50-37-01 6.25E-02 0.0 6.32% 8.05 14.6% 0.028 2.929E-03 14.6% 78.6 9.2% 0.026 35.0% 5.26 14.6% 0.019 1.0% 0.04885-50-50-01 6.25E-02 0.0 6.32% 7.92 10.1% 0.025 2.929E-03 10.1% 55.9 6.4% 0.017 25.9% 2.72 10.2% 0.013 1.0% 0.048

Fuel Orifice Using Incompressible Flow Equations