You are on page 1of 17

Analytical Study Of Paraffins C12H26 And ItsAutoCumbustion Process Using Decomposed Hydrogen Peroxide

NAME OF RESEARCHER ADDRESS: EMAIL ID Abstract This study recognizes the automatic burning sequence of paraffin with the use of hydrogen peroxide. The rocket combustors uses the 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 autoburn 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 autocombustion. 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 autoburning characteristics of JP-8 in decomposed hydrogen peroxide. The goals of the investigation were to develop a greater understanding of the autocombustion process and, if possible, develop auto-combustion 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 autocombustion boundary increased as well. At a contraction ratio of 3.0, no autocombustion was achieved down to

an equivalence ratio of 1.37 using 85% hydrogen peroxide, but at 98% hydrogen peroxide autocombustion occurred up to an equivalence ratio of 2.06. When the

contraction ratio was increased to 5.0 autocombustion 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 auto-combustion model.

Introduction: Hydrogen peroxide and kerosene rocket engines have a long history of use in propulsion systems dating back prior to World War II.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, H2 O2 , 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 below outlines performance parameters of

common rocket systems operating at similar conditions. Spacecraft reaction control, RCS, and orbital maneuvering systems, OMS, have typically used hydrazine and NTO/MMH rocket systems since the 1960s. This was due to the storability of the propellants, hydrazines 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.

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.There is also increasedinterest to develop low cost, reusable satellite launch vehicles to replace the expendable vehicles currently used in industry.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.

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 Optimum O/F Ratio Characteristic Velocity, C* (ft/s) Chamber Temperature, Tc (F) Specific Impulse, Isp (sec) Density Specific Impulse, Density Isp (sec) 90% H2 O2 JP-8 7.8 5300 4600 267 344 NTO MMH 2.2 5710 5650 288 346 LOx LH2 3.5 7940 4450 386 101 LOx RP-1 2.6 5890 6160 300 308

Hydrogen peroxide and kerosene rocket systems have the potential to replace their toxic and cryogenic predecessors. Table above shows that both NTO/MMH and

H2O2/kerosene systems have comparable density specific impulse, density I , and are sp 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 H2 O2/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. 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 H O2 /kerosene system have similar simplicity as 2 well.Hydrogen peroxide and kerosene are not hypergolic by themselves, and there is research being done make these propellants ignite on contact.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, autocombustion is dependent on a number of different factors such as fuel injector design, hydrogen peroxide concentration, decomposed gas velocity, chamber pressure, and mixture ratio. A better understanding of the autocombustion 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 autocombustion under varying engine operating conditions, and creating a model to aid in the prediction of autocombustion. Results of this research

may make the staged-bipropellant H O2 /kerosene rocket a lighter, more reliable, and 2 higher performing engine in the future. 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 H O2 by 2

weight of solution. Propellant-grade H2 O2 , 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 H2 O2 is less than 0.1% per year over normal atmospheric temperature and pressure ranges.Decomposition is significantly accelerated as the temperature of the 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 b e l o w outlines the variation in physical and decomposition properties of hydrogen peroxide with concentration. Table : Properties of liquid and decomposed hydrogen peroxide with concentration. Concentration Molecular Weight Specific Gravity Boiling Point (F) Vapor Pressure (psia) Heat Capacity (Btu/lbm- R) Temperature (F) Molecular Weight Specific Heat Ratio Mass Fraction O2 Mass Fraction H2O 70 % H2O2 80 % H2O2 90 % H2O2 98 % H2O2 Liquid Properties (@ STP) 26.86 28.89 31.29 33.42 1.283 1.333 1.387 1.432 257 -287 299 0.137 -0.065 0.045 0.738 -0.663 0.633 Decomposed Gas Properties 504 952 1393 1746 21.04 21.56 22.11 22.56 1.315 1.287 1.265 1.251 0.341 0.376 0.423 0.461 0.659 0.624 0.577 0.539

Hydrogen peroxide has a low vapor pressure, as Table above 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. It is advantageous to use a propellant with a low vapor pressure in rocket system for several reasons. 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 As a

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

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 above 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. It is considered to be a non-toxic propellant as well. Toxic propellants are poisonous to humans through inhalation or However, hydrogen peroxide solutions and vapors are

contact with the body tissue. 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. The most important feature of hydrogen peroxide as propellant is its reactivity. The hot gases produced when H2 O2 is decomposed contain a significant amount of energy, see Table above .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% conc entration has been used in RCS systems in the past, such as the Mercury space capsule, and new systems using H2O2 are currently in development. These gases can also be expanded through a row of turbine blades imparting its energy to generate turbine rotation. This is important

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 H2 O2 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. Purpose of the study: o To investigate the auto-combustion characteristics of paraffin-based JP-8 fuel in decomposed hydrogen peroxide. o To perform testing using staged-bipropellant rocket engine in a dump combustor configuration. o 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. o Downstream of the injection point a rearward- facing step was used to provide flame stabilization at the entrance to the combustion chamber. o Testing was structured to study the affects of gas temperature, gas velocity, and equivalence ratio on autocombustion. o To perform the test at three stsges. 1)Strong Autocombustion 2)Week Autocombustion 3)No Autocombustion

o To analyze the test cases by collecting the data from Mestre and Ducourneau and Walder. o To check that the variations in both the pressure and equivalence ratio combined to influence the autocombustion temperature at a constant contraction ratio not just one or the other. o However, the test conditions were such that the variations could not be isolated from one another. o To investigated the effect of gas velocity on autocombustion. o Changes in contraction ratio will also affected the trajectory of the fuel jet in the oxidizer port or not. o As the contraction ratio was increased the shear layer residence time created by

the rearward- facing step was increased as well or not . o Mainly to proves that the stable autocombustion is possible at contraction ratios as low as 3.0 and equivalence ratios less than 1.4 using 90% H2 O2 . METHODOLOGY:
The test conducted in various phases to determine the actual performance using hydrogen

peroxide in monopropellant and bipropellant system. 1. Test Facility verview 2. Cavitating Venturi Flow Control 3. Data Acquisition and Control 4. Instrumentatio 5. Test Article and etup. 6. Hydrogen Peroxide Dilution. 7. Pressure Budget 8. Firing Sequence 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.

Conclusion: A total of 24 tests were conducted to investigate the autocombustion 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 rearwardfacing 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 autocombustion. Each test was classified into one of three groups based on visual observations and measured chamber pressure data.Tests classified as strong autocombustion 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 autocombustion, 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 autocombustion case, but the rise was not as sharp. The third test classification was no autocombustion. 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. It was determined that severe chamber pressure instabilities present during weak autocombustion 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 autocombustion so too did the flame. The gas temperature, gas velocity, and equivalence ratio were controlled during testing by varying the H2 O2 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 autocombustion and no autocombustion 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 autocombustion increased as well. At a contraction ratio of 3.0 and a concentration of 85% H2O2 to an equivalence ratio of 1.37 while no autocombustion was achieved down at a concentration of 98% strong

autocombustion was achieved up to an equivalence ratio of 2.06. This trend agrees with past data from Mestre and Ducourneau for kerosene in air as well as Walder for kerosene in decomposed hydrogen peroxide. Both show that higher temperatures are required for autocombustion 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 equivale nce ratio plays a negligible role in autocombustion. 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 nearlyuniversal

agreement from past autocombustion studies that the autocombustion 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 autocombustion temperature of kerosene fuel in air by approximately 90F at an equivalence ratio of 2.0. Data from Walder suggests a temperature difference of about 20F 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 autocombustion 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 autocombustion 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 autocombustion was achieved at all the tested conditions using 85% H O2 . At this contraction ratio the Mach number in the 2 chamber is about 0.20 and 0.45 in the oxidizer port.When the contraction ratio was increased to 5.0, however, strong autocombustion was achieved at an equivalence ratio of 1.36 and weak autocombustion 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 autocombustion 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 autocombustion should increase as well at a certain temperature. Both studies suggest that the affect of residence time, characteristic length, or gas velocity on the autocombustion 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 increased with increasing contraction ratio. Based on the previous discussion on pressure, it is believed that pressure affects also contributed to the decrease in

autocombustion 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 autocombustion tests and no autocombustion tests. This may suggest that the fuel

trajectory and atomization did not play a critical role in autocombustion.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 autocombustion 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%. increased from this point through the rest of the calculations.

The uncertainty

Despite this, it is believed that the rearward-facing step did provide enough residence time to improve the autocombustion limits based on past data. Many flightrated 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 autocombustion of kerosene in hydrogen peroxide used contraction ratios of six and higher at stoichiometric conditions. In addition, data from Wu et al using a similar engine design running 85% H O2 at an equivalence ratio 2 somewhere between 0.8 and 1.4 did not achieve autocombustion at contraction ratio of approximately 5.0 even at a monoprop chamber pressure of 340 psia. Data from this study proved that autocombustion 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 autocombustion is possible at contraction ratios as low as 3.0 and equivalence ratios less than 1.4 using 90% H2O2. BIBLIOGRAPHY 1. Ventura, M., Mullens, P., The Use of Hydrogen Peroxide for Propulsion and Power, AIAA Paper 99-2880, June 1999. Andrews, D., Advantages of Hydrogen Peroxide as a Rocket Oxidant, Journal The British Interplanetary Society, Vol. 43, 1990, pp. 319-328. 3. AFRPL-TR-67-144, Hydrogen Peroxide Handbook, Rocketdyne, Inc., July 1967. 4. 5. TEP Version 1.5, SEA Software, Inc., Carson City, Nevada, 1999. Humble, R. W., Henry, G. N., Larson, W. J., Space Propulsion Analysis and Design, McGraw-Hill Companies Inc., New York, 1995. 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. Chen et al, Testing Research on Nontoxic H2 O2 /Kerosine Liquid Propellant Engines, IAF-01-S309, 52nd International Astronautical Congress, Paris, 2001.

2. of

6.

7.

8.

Walder, H., An Investigation into the Thermal Ignition of Hydrogen Peroxide and Kerosine, Report Number RPD 7, Royal Aircraft Establishment, May 1950. Walder, H., Further Investigations into the Thermal Ignition of Hydrogen Peroxide and Kerosine, Technical Note Number RPD 43, Royal Aircraft Establishment, December 1950.

9.

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. 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% H2 O2 /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. 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. 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 980716, 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. 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. 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.

You might also like