CHAPTER 7
LIQUID PROPELLANTS

The classification of liquid propellants was first introduced in Section 6.1. In this chapter, we discuss properties, performance, hazards, and other characteristics of commonly used propellants that are stored as liquids (and a few as gases). These characteristics influence engine and vehicle design, test facilities, and propellant storage and handling. At the present time, we ordinarily use three liquid bipropellant combinations: (1) the cryogenic oxygen–hydrogen propellant system, used in upper stages and sometimes booster stages of space launch vehicles, giving the highest specific impulse nontoxic propellant combination and one that is best for high vehicle velocity missions; (2) the liquid oxygen–hydrocarbon propellant combination, used for booster stages (and a few second stages) of space launch vehicles —having a higher average density allows more compact booster stages with less inert mass when compared to the previous combination (historically, it was developed first and was originally used with ballistic missiles); (3) not a single bipropellant combination but several ambient temperature storable propellant combinations used in large rocket engines for first and second stages of ballistic missiles and in almost all bipropellant low‐thrust, auxiliary or reaction control rocket engines (this term is defined below); these allow for long‐term storage and almost instant readiness (starting without the delays and the precautions that come with cryogenic propellants). Each of these propellant systems is further described in this chapter. Presently, Russia and China favor nitrogen tetroxide as the oxidizer and unsymmetrical dimethylhydrazine, or UDMH, as the fuel for ballistic missiles and for auxiliary engines. The U.S. has used nitrogen tetroxide and a fuel mixture of 50% UDMH with 50% hydrazine in the Titan II and III missiles' large engines. For auxiliary engines in many satellites and upper stages, the United States uses a nitrogen‐tetroxide/monomethylhydrazine bipropellant. A subcategory of item (3) above is the storable monopropellant such as hydrogen peroxide or hydrazine. The International Space Station and many U.S. satellites use monopropellant hydrazine for low‐thrust auxiliary engines.

No truly new liquid propellant has been adopted for operational rocket flight vehicles in the past 30 years. Some new propellants (such as hydroxyl ammonium nitrate) were synthesized, manufactured, and ground tested in thrust chambers and flown in experimental vehicles in the past two decades, but they have not found their way into operational rocket engine applications. Between 1942 and 1975, a number of other propellants were successfully flown; these included ammonia (X‐15 Research test aircraft), ethyl alcohol (German V‐2 or U.S. Redstone missile), and aniline (WAC Corporal). They each had some disadvantages and are no longer used in operational flights today. Liquid fluorine and fluorine containing chemicals (such as chlorine pentafluoride and oxygen difluorine) give excellent performance and have been investigated and experimentally evaluated but, because of their extreme toxicity, they are no longer being considered.

A comparative listing of various performance quantities for a number of propellant combinations is given in Table 5–5 and in Ref. 7–1. Some important physical properties of selected common liquid propellants are shown in Table 7–1 (water is also listed for comparison). Specific gravities and vapor pressures are shown in Figs. 7–1 and 7–2. The specific gravity is defined to represent the ratio of the density of any given liquid to that of water at standard conditions (273 K and 1.0 atm) and thus carries no dimensions.

Table 7–1 Physical Properties of Liquid Propellants

Propellant Liquid Oxygen Nitrous oxide Nitrogen Tetroxide Nitric Acida (99% pure) Rocket Fuel RP‐1, RP‐2
Chemical formula O2 N2O N2O4 HNO3 Hydrocarbon CH1.97
Molecular mass 31.988  44.013  92.016  63.016  ∼175       
Melting or freezing point (K) 54.8    182.29   261.95   231.6    225      
Boiling point (K) 90.2    184.67   294.3    355.7    460–540
Heat of vaporization (kJ/kg) 213      374.3    (at 1 atm) 4132     480      246b     
Specific heat 0.4  0.209  0.374 0.042  0.48+
(kcal/kg‐K) (65 K)  (290 K)    (311 K)    (298 K) 
0.447 0.163 
(360 K)    (373 K)   
Specific gravityc 1.14  1.23b 1.38   1.549  0.58 
(90.4 K)  (293 K)    (273.15 K) (422 K) 
1.23  1.447 1.476  0.807
(77.6 K)  (322 K)    (313.15 K) (289 K) 
Viscosity 0.87  0.0146
(gas at 300 K)
0.47   1.45   0.75 
(centipoises) (53.7 K)  (293 K)    (273 K)    (289 K) 
0.19  0.33   0.21 
(90.4 K)  (315 K)    (366 K) 
Vapor pressure 0.0052 5.025  0.1014 0.0027 0.002
(MPa) (88.7 K) (293 K)    (293 K)    (273.15 K) (344 K) 
0.2013 0.605  0.023
(328 K)    (343 K)    (422 K) 
para–H2 CH4 CH3 NHNH2 N2 H4 (CH3)2 NNH2 H2O
2.016  16.04  46.072  32.045  60.099 18.02 
14.0    90.67 220.7    275.16   216   273.15  
20.27   111.7 360.8    387.46 335.5 373.15  
446       510b   808      1219b      543   22532      
2.34b   0.835b  0.700  0.736 0.704 1.008
(20.27 K) (298 K)   (293 K)  (298 K) (273.15 K)
0.735  0.758 0.715
(393 K)   (338 K)  (340 K)
0.071 0.424 0.8702 1.0037 0.7861 1.002
(20.4 K)  (111.5 K) (298 K)   (298 K)  (298 K) (373.15 K)
0.076 0.857 0.952 0.784 1.00
(14 K)    (311 K)   (350 K)  (244 K) (293.4 K) 
0.024 0.12 0.775 0.97 0.492 0.284
(14.3 K)  (111.6 K) (298 K)   (298 K)  (298 K) (373.15 K)
0.013 0.22 0.40 0.913 0.48 1.000
(20.4 K)  (90.5 K) (344 K)   (330 K)  (300 K) (277 K)   
0.2026     0.033 0.0066 0.0019 0.0223    0.00689
(23 K)    (100 K) (298 K)   (298 K)  (298 K) (312 K)   
0.87  0.101 0.638 0.016 0.1093    0.03447
(30 K)    (111.7 K) (428 K)   (340 K)  (339 K) (345 K)   

a Red fuming nitric acid (RFNA) has 5 to 20% dissolved NO2 with an average molecular mass of about 60, and a density and vapor pressure somewhat higher than those of pure nitric acid.

b At boiling point.

c Reference for specific gravity ratio: 103 kg/m3 or 62.42 lbm/ft3.

A plot with Specific gravity on the vertical axis, Temperature in Kelvin and Fahrenheit on the horizontal axis, and different curves plotted and marked with arrows.

Figure 7–1 Specific gravities of several liquid propellants as a function of temperature.

A plot with vertical pressure on the horizontal axis, Temperature in Kelvin and Fahrenheit on the vertical axis, and different curves plotted and marked with arrows.

Figure 7–2 Vapor pressures of several liquid propellants as a function of temperature.

Green propellants (Ref. 7–2), a recently minted term, represent those liquid propellants and their exhaust gases that are “environmentally friendly” and can be used without causing damage to people, equipment, or the surroundings. An excellent example is the liquid oxygen–liquid hydrogen propellant combination—they are not toxic, not corrosive, not hypergolic and will not decompose or explode. Some authors use a more restricted interpretation for the green propellant category, namely, one that can replace a toxic and/or potentially explosive propellant with a chemical substance that is harmless.

7.1 PROPELLANT PROPERTIES

It is important to distinguish between characteristics and properties of liquid propellants (i.e., fuel and oxidizer liquids in their unreacted condition) and those of the hot gas mixture resulting from their reaction in the combustion chamber. The chemical nature of liquid propellants and their mixture ratio determine the properties and characteristics of both storage and reaction products. Because none of the known practical propellants encompass all properties deemed desirable, the selection of propellant combinations is usually a compromise between various economic factors, such as those listed below.

Economic Factors

Availability in quantity and a low cost are very important considerations in propellant selection. In military applications, consideration has to be given to the logistics of production, supply, storage, along with other factors. The production process should require only ordinarily available chemical equipment and available raw materials. It is more expensive to use toxic or cryogenic propellants than storable, nontoxic ones, because the former require additional steps in their operation, more safety provisions, additional design features, longer check‐out procedures prior to launch, and often better trained personnel.

Performance of Propellants

The performance rocket engines may be compared on the basis of their specific impulse, exhaust velocity, characteristic velocity, and/or other engine parameters. These were introduced in Chapters 3, 5, and 6. The specific impulse and exhaust velocity are functions of pressure ratio, specific heat ratio, combustion temperature, mixture ratio, and molecular mass. Equilibrium values of performance parameters for various propellant combinations can be calculated with a high degree of accuracy and several are listed in Table 5–5. Very often, performance is also expressed in terms of flight performance parameters for specific rocket applications, as explained in Chapter 4. Here, average propellant density, total impulse, and engine mass ratio usually enter into the various flight relation descriptions.

For high performance, high chemical energy content per unit of propellant mixture is desirable because this yields high chamber temperatures. A low molecular mass for the combustion products gases is also desirable. This can be accomplished by using fuels rich in hydrogen obtained when a significant portion of the hydrogen gas injected or produced remains uncombined. In general, therefore, the best mixture ratio for many bipropellants is not stoichiometric (which results in complete oxidation and yields the highest flame temperature) but fuel‐rich, containing a large amounts of low‐molecular‐mass reaction products as shown in Chapter 5.

When relatively small metallic fuel particles (such as beryllium or aluminum) are suspended in the liquid fuel, it is theoretically possible to increase the specific impulse by between 9 and 18%. A particular chemical propellant combination having the highest ideal specific impulse known (approximately 480 sec at 1000 psia chamber pressure and expansion to sea‐level atmosphere, and 565 sec in a vacuum with a nozzle area ratio of 50) uses a toxic liquid fluorine oxidizer with hydrogen fuel plus suspended toxic solid particles of beryllium; as yet, acceptably safe and practical means for storing such seeded propellants and their practical rocket engine use remain undeveloped.

Gelled propellants are materials that have additives that make them thixotropic. They have the consistency of thick paint or jelly when at rest, but do liquefy and flow through pipes, valves, pumps and/or injectors when an adequate pressure or shear stress is applied. In spite of extensive research and development work and demonstrations of better safety and certain “green” qualities, to date they have not been adopted for any production rocket engine. Gelled propellants have been described in the Sixth, Seventh, and Eighth editions of this book.

Common Physical Hazards

Although several hazard categories are described below, they do not all apply to each propellant or to every bipropellant combination. Hazards can be different for each specific propellant and must be carefully understood before working with it. The consequences of unsafe operation or unsafe design are usually also unique to each propellant.

Corrosion

Several propellants, such as nitrogen tetroxide, nitric acid, nitric oxide and/or hydrogen peroxide, can only be handled in containers and pipelines made from special materials. When any propellant gets contaminated with corrosion products, its physical and chemical properties may sufficiently change to make it unsuitable for its intended operation. Corrosion caused by expelled gaseous reaction products is most critical in applications where the reaction products are likely to damage launch or ground test structures and parts of the vehicle, and/or affect communities and housing near a test facility or launch site.

Explosion Hazard

Over time some propellants (e.g., hydrogen peroxide or nitromethane) can become unstable in their storage tanks and may even detonate under certain conditions, depending on local impurities, temperatures, and shock magnitudes. When liquid oxidizers (e.g., liquid oxygen) and fuels get unintentionally mixed together, detonation may sometimes result. Unusual flight vehicle launch mishaps or transport accidents have caused such mixing and subsequent explosions to occur (see Refs. 7–3 and 7–4).

Fire Hazard

Many oxidizers will react with a large variety of organic compounds. Nitric acid, nitrogen tetroxide, fluorine, and/or hydrogen peroxide react spontaneously when in contact with many organic substances resulting in fires. Most of rocket fuels exposed to air are readily ignitable when heated. Also some household dusts, certain paints, or smoke particles can oxidize. Oxygen by itself will not usually start a fire with organic materials but it will greatly enhance an existing fire.

Accidental Spills

Unforeseen mishaps during engine operation or traffic accidents on highways or railroads while transporting hazardous materials, including many propellants, have on occasion caused spills which expose people to intense fires and/or potential health hazards. The U.S. Department of Transportation has strict rules for marking and containing hazardous materials during transport and also guidelines for emergency actions (see Ref. 7–5).

Health Hazards

Exposure to many commonly used propellants represents a health hazard. Toxic unburned chemicals or poisonous exhaust species affect the human body in a variety of ways and the resulting health disorders are propellant specific. Nitric acid causes severe skin burns and tissue disintegration. Skin contact with aniline or hydrazine may cause nausea and other adverse health effects. Hydrazine, monomethylhydrazine, unsymmetrical dimethylhydrazine, or hydrazine hydrate are known animal and suspected human carcinogens. Many propellant vapors cause eye irritation, even in small concentrations. Inadvertent ingestion of many propellants may result severe health degradation.

Inhalation of toxic exhaust gases or gaseous or vaporized liquid propellants is perhaps the most common health hazard. It can cause severe damage if the exposure is long or in concentrations that exceed established maximum threshold values. In the United States, the Occupational Safety and Health Administration (OSHA) has established limits or thresholds on the allowable exposure and concentration for most propellant chemicals. Several of these threshold limits are mentioned later in this chapter. References 7–3 and 7–6 give more information on toxic effects.

Toxic Propellants

These require special safety provisions, strict rules, and specific procedures for handling, transferring to other containers, road transport, inspection, and for working on rocket engines that contain them (such as the removal of residuals after testing or after return from a space mission). Instruments are available to detect toxic vapors or toxic contents in liquids or water. For personnel protection, face shields and gas masks (some with an oxygen supply), special gloves and boots, sealed communications equipment, medical supplies and periodic medical examinations need to be available. For accidental toxic spills or leaks there should be equipment for diluting with water or other suitable chemical, and for safe disposal of the contaminants from a test stand or launch platform. Chemicals for neutralization and detoxification should always be on hand. In case of mishaps instructions and procedures for notifying management, certain government offices and relevant others should be available and not overlooked. When compared to nontoxics, operations with toxic‐propellants require two to four times as many trained workers. Only subsets of the precautions listed above apply to any one operation. Toxic propellant safety is further discussed in Chapters 20 and 21.

Materials Compatibility

For several liquid propellants there are only a limited number of truly compatible materials, both metals and nonmetals, particularly for making gaskets or O‐rings. There have been many failures (causing fires, leakage, corrosion, or other malfunctions) when improper or incompatible hardware materials have been used in rocket engines. Depending on the specific component and loading conditions, structural materials have to withstand high stresses, stress corrosion, and in some applications high temperatures and/or abrasion. Several specific material limitations are mentioned in the next section. Certain storage materials may act to catalyze the self‐decomposition of hydrogen peroxide into water and oxygen making long‐term storage difficult and causing any closed containers to explode. Many structural materials, when exposed to cold (cryogenic) propellants become unacceptably brittle.

Desirable Physical Properties

Low Freezing Point

This permits operation of rockets in cold environments. The addition of small amounts of freezing point depressants has been found to help lower the freezing point in some liquid propellants that might otherwise solidify at environmental storage conditions.

High Specific Gravity

Denser propellants provide a larger propellant mass for a given vehicle tank volume. Alternatively, for a given mass, they permit smaller tank volumes and, consequently, lower structural vehicle mass and lower aerodynamic drag. The specific gravity of a propellant, therefore, has an important effect on the maximum flight velocity and range of any rocket‐powered vehicle or missile flying within the Earth's atmosphere as explained in Chapter 4. Specific gravities for various propellants are plotted in Fig. 7–1. Variations of ambient temperature in stored propellants cause changes of the liquid level in their storage tanks.

For a given bipropellant mixture ratio r, the average specific gravity of any propellant combination δav can be determined from the specific gravities of the fuel δf and of the oxidizer δo. This average specific gravity δav is defined below. Here r is the mixture ratio; it represents the oxidizer mass flow rate divided by the fuel mass flow rate (see Eq. 6–1):

Values of δav for various propellant combinations are listed in Table 5–5. The value of δav can be increased by adding high density materials to the propellants, either by solution or colloidal suspension. An identical equation can be written for the average density ρav in terms of the fuel and oxidizer densities:

Though the specific gravity is unitless, in the SI system it has the same numerical value as the density expressed in units of grams per cubic centimeter or kg/liter. In some performance comparisons the parameter density specific impulse Id is used. It is defined as the product of the average specific gravity δav and the specific impulse Is:

A propellant density increase will allow increases in mass flow and total propellant mass when all other system parameters remain the same. For example, Fig. 7–1 shows that lowering the temperature of liquid oxygen from –250 to –280 °F raises its specific gravity by approximately 8%. Therefore, the system's mass flow rate and the total mass will increase by approximately the same amount. These changes will enable increases in chamber pressure, total impulse, and thrust; changes of less than 1% in specific impulse may also be noticed. So there is some benefit to vehicle performance in operating with the liquid propellants at their lowest practical temperature. In ground‐based systems, cooling of liquid oxygen may be achieved with a (colder) liquid nitrogen heat exchanger just before launch. The percent increase in performance with propellants other than oxygen is usually much smaller.

Stability

Proper chemical stability means no decomposition of the liquid propellant during operation or storage, even at elevated temperatures. With many propellants, insignificant amounts of deterioration and/or decomposition during long‐term (over 15 years) storage and minimal reaction with the atmosphere have been attained. A desirable liquid propellant should also experience no chemical deterioration when in contact with tubing, pipes, tank walls, valve seats, and gasket materials even at relatively high ambient temperatures. No appreciable absorption of moisture and no adverse effects of small amounts of impurities are also desirable properties. There should also be no appreciable chemical deterioration when liquids flow through hot cooling jacket passages of a regeneratively cooled thrust chamber. When carbon‐containing coolants decompose and form (carbonaceous) deposits on hot inside surfaces of cooling passages, such deposits may harden, reducing the heat flow, increasing metal temperatures locally, and thus may cause the unit to weaken and eventually fail. Unavoidably, even in well‐insulated tanks, between 1 and 20% of a cryogenic propellant may evaporate daily when stored in the flight vehicle.

Heat Transfer Properties

High specific heat, high thermal conductivity, low freezing temperature, and high boiling or decomposition temperature are all desirable properties for propellants used for thrust chamber cooling (see Section 8.5).

Pumping Properties

Low vapor pressures permit not only easier handling of propellants, but also more effective pump designs in applications where propellants must be pumped. Low vapor pressures also reduce the potential for pump cavitation, as explained in Chapter 10. When the propellant viscosity is too high, then pumping and engine‐system calibrations become difficult. Propellants with high vapor pressures (such as LOX, liquid hydrogen, and other liquefied gases) require special design provisions, unusual handling techniques, and special low‐temperature materials.

Temperature Variation of Physical Properties

Temperature variations in the physical properties of any liquid propellant should be small and should be very similar for the fuel and oxidizer. For example, wide temperature variations in vapor pressure and density or unduly high changes in viscosity with temperature make it difficult to accurately calibrate rocket engine flow systems and/or to predict their performance over any reasonable range of operating temperatures. While stored in a flight vehicle, if one of the propellants experiences a larger temperature change than the other, this may cause a noticeable change in mixture ratio and in specific impulse, and possibly a significant increase in unusable or undesirable propellant residue.

Ignition, Combustion, and Flame Properties

When a propellant combination is spontaneously ignitable burning is initiated as soon as the oxidizer and the fuel come in contact with each other. Spontaneously or self‐ignitable propellant combinations are called hypergolic propellants. Although ignition systems are not necessarily objectionable, their elimination simplifies the propulsion system. In order to reduce potential explosion hazards during starting, all rocket propellants should be readily ignitable and exhibit acceptably short ignition time delays in order to eliminate potential explosion hazards. Starting and ignition problems are discussed further in Section 8.6.

Nonspontaneously ignitable propellants must be energized by external means for ignition to begin. Igniters are devices that accomplish a localized initial chamber pressurization and initial heating of the propellant mixture to a state where steady flow combustion can be self‐sustained. The amount of energy needed from the igniter to activate the propellants should be small so that low‐power, light‐weight ignition systems may be used. The energy required for satisfactory ignition usually diminishes with increasing propellant storage temperature. At low ambient temperatures ignition can become relatively slow (0.05 to 0.02 sec).

Certain propellant combinations burn very smoothly, without vibration (i.e., gas pressure oscillations). Other propellant combinations do not exhibit such combustion stability and, therefore, become less desirable. Combustion stability is treated in Chapter 9.

Smoke formation is objectionable in many applications because it may deposit on the surrounding equipment and parts. Smoke and brilliantly luminous exhaust flames are objectionable for certain military applications since they can be easily detected. In some applications, the condensed species from gaseous exhausts can cause surface contamination on spacecraft windows or optical lenses, and the presence of free electrons in a flame may cause undesirable interference or attenuation of communications radio signals. See Chapter 20 for information on exhaust plumes.

Property Variations and Specifications

Propellant properties and quality must not vary from delivered batch to batch because this can affect engine performance and combustion from changes in physical and/or chemical properties. The same propellant must have consistent composition and storage properties; rocket operating characteristics should not change even when propellants are manufactured at different times or made by different manufacturers. For these reasons propellants are purchased to conform to strict specifications which define ingredients, maximum allowable impurities, packaging methods and/or compatible materials, allowable tolerances on physical properties (such as density, boiling point, freezing point, viscosity, or vapor pressure), quality control requirements, container cleaning procedures, documents of inspections, laboratory analyses, and/or test results. A careful chemical analysis of composition and impurities is always necessary. Reference 7–7 describes some of these methods of analysis.

Additives

Altering and tailoring propellant properties may be achieved with additives. For example, a reactive ingredient is added to make a nonhypergolic fuel become hypergolic (readily ignitable). To desensitize concentrated hydrogen peroxide and reduce self‐decomposition, it is diluted with 3 to 15% of pure water. To increase the density or to alleviate certain combustion instabilities, a fine powder of a heavy solid material (such as aluminum) is suspended in the fuel. The use of additives to lower the freezing‐point temperature of nitrogen tetroxide is treated later.

7.2 LIQUID OXIDIZERS

The most energetic known oxidizer producing the highest specific impulse and having the highest density is liquid fluorine. It has been tested in several complete experimental rocket engines but abandoned because of its extreme hazards. Many different types of new storable and cryogenic liquid oxidizer propellants have been synthesized, tested in small thrust chambers, or proposed; these included mixtures of liquid oxygen and liquid fluorine, oxygen difluoride (OF2), chlorine trifluoride (ClF3), and chlorine pentafluoride (ClF5). None of these are used today because they are highly toxic and very corrosive. Several of the most commonly used oxidizers are listed below.

Liquid Oxygen (O2) (LOX)

Liquid oxygen is listed in Table 7–1. It is widely used as an oxidizer and burns with a bright white‐yellow flame with most hydrocarbon fuels. It has been used in combination with alcohols, jet fuels (kerosene‐type), gasoline, and hydrogen. As shown in Table 5–5, attainable performances are relatively high, and LOX is therefore a desirable and commonly used propellant in large rocket engines. The following missiles and space launch vehicles burn oxygen: (1) with jet fuel or RP‐1 (kerosene)—Atlas V and Soyuz (Russia); (2) with hydrogen—Ariane‐V (France), Delta IV, and Centaur upper stage; (3) with ethanol—the historic V‐2 (German) and the Redstone missile. Figs. 1–12, 1–13, and 6–1 show units that use LOX; Figs. 5–1 to 5–6 present theoretical performance information for LOX with one kerosene‐type fuel.

Although it usually does not burn spontaneously with organic materials at ambient pressures, combustion and/or explosions have occurred when confined mixtures of oxygen and organic matter suddenly come under pressure. Impact tests show that mixtures of LOX with many commercial oils or organic materials will detonate. Liquid oxygen also supports and accelerates the combustion of other materials. Handling and storage can be safe only when all contact materials are clean. Liquid oxygen is a noncorrosive and nontoxic liquid that will not cause the deterioration of clean container walls. Because of its low temperature, this propellant causes severe frostbite during any prolonged contact with human skin. Because LOX evaporates rapidly at ambient conditions, it cannot be stored readily for any extended length of time. When LOX is used in large quantities, it is often produced very close to its geographical point of application. Liquid oxygen can be obtained in several ways, but most commonly from the fractionated distillation of liquid nitrogen out of liquid air.

It is necessary to insulate all lines, tanks, valves, and parts that contain liquid oxygen in order to reduce evaporation losses. Rocket propulsion systems which have to remain filled with liquid oxygen for several hours and liquid oxygen storage systems need to be well insulated against heat absorption from the surroundings. External drainage provisions have to be made on all LOX tanks and lines to allow the water that condenses on the cold outside walls to drain from the rocket during launch preparations. Provisions for minimizing and removing ice formations are also needed.

Hydrogen Peroxide (H2O2)

This propellant is not only a powerful liquid oxidizer when burning with an organic fuel but also clean burning. It produces a nontoxic exhaust when used as a monopropellant. In rocket applications, hydrogen peroxide needs to be in a highly concentrated form of 70 to 98% (known as high‐test peroxide or HTP), the remainder being mostly water. Hydrogen peroxide was used in gas generators and rocket applications between 1938 and 1965 (X‐1 and X‐15 research aircraft). Since that time, 90% hydrogen peroxide has been the most common concentration for rocket engine use and gas generator applications.

As a monopropellant, it decomposes according to the following chemical reaction, forming superheated steam and hot gaseous oxygen:

images

This decomposition is brought about by the action of catalysts such as silver screens, various liquid permanganates, solid manganese dioxide, platinum, or iron oxide but, in fact, many materials can catalyze hydrogen peroxide. Theoretical specific impulses for 90% hydrogen peroxide can be around 154 sec, when used as a monopropellant with a solid catalyst bed. As bipropellant, H2O2 is hypergolic with hydrazine and will burn well with kerosene.

Concentrated peroxide causes severe burns when in contact with human skin and may ignite when in contact with wood, oils, and many other organic materials. In the past, rocket engines with hydrogen peroxide as oxidizer have been used for aircraft boosting (German Me 163, and U.S. F 104) and for satellite launching (Britain: Black Arrow). As a monopropellant it has been used for attitude control thrusters in the Soyuz space capsule and for steam generators driving the main propellant pumps in the Soyuz launch vehicle. It was not used in the United States for some time, partly because it was hard to predict its long‐term storage stability. In the past, concentrated hydrogen peroxide would self‐decompose during storage at about 1% per year. This progressively diluted this propellant until no longer useful. As oxygen gas would bubble out of the stored liquid, the decomposition rate would accelerate. Small amounts of impurities in the liquid or in wall materials would also accelerate decomposition and cause a rise in liquid temperature. Before reaching 448 K (175 °C), warm stored H2O2 or contaminated H2O2 must be diluted and discarded because an explosion may occur.

Recent progress in the manufacture of storage tanks and piping materials (and stricter cleaning methods) has much reduced the amount of impurities and lengthened storage life of HTP from about 3 to 4 years to 12 to 16 years, and there may further improvements. This has much renewed interest in this dense oxidizer and several different investigations and development programs have been underway to date (Ref. 7–8). To date, the authors have not become aware of any new production applications for 90% H2O2.

Nitric Acid (HNO3)

Several types of nitric acid mixtures were used as oxidizers between 1940 and 1965; these are not used extensively today in the U.S. The most common type, red fuming nitric acid (RFNA), consists of concentrated nitric acid (HNO3) that contains between 5 and 27% dissolved nitrogen dioxide (NO2). Compared to concentrated nitric acid (also called white fuming nitric acid), RFNA is more energetic, more stable in storage, and slightly less corrosive to many tank materials.

Only certain types of stainless steel, gold, and a few other materials are satisfactory as storage containers and/or tubing materials for concentrated nitric acid. Small additions of fluoride ion (less than 1% of hydrofluoric acid, or HF) inhibit nitric acid corrosion, forming a protective fluoride layer on walls and reducing the corrosion with many metals. This combination is called inhibited red fuming nitric acid (IRFNA). Even with an inhibitor, nitric acid reacts with many wall materials and forms dissolved nitrates and sometime insoluble nitrates. This changes the properties of this oxidizer and may cause blocking of valve and injector orifices. Nitric acid reacts with gasoline, various amines, hydrazine, dimethylhydrazine, and alcohols. It ignites spontaneously with hydrazine, furfuryl alcohol, aniline, and other amines. The specific gravity of nitric acid varies from 1.5 to 1.6, depending on the percentages of nitrogen dioxide, water, and impurities. This high density permits compact vehicle construction.

When RFNA is exposed to air, the evaporating red‐brown fumes are exceedingly annoying and poisonous. In case of accidental spilling, the acid must be quickly diluted with water or chemically deactivated. Lime and alkali metal hydroxides and carbonates are common neutralizing agents. However, any nitrates formed by the neutralization also act as oxidizing agents and must be handled accordingly. Vapors from nitric acid or red fuming nitric acid have an OSHA eight‐hour personnel exposure limit or a threshold work allowance of 2 ppm (parts per million, or about 5 mg/m3) and a short‐term exposure limit of 4 ppm. Droplets on the skin cause burns producing sores that do not heal readily.

Nitrogen Tetroxide (N2O4) (NTO)

The proper chemical name for N2O4 is ‘dinitrogen tetroxide’ as widely used by chemists and the chemical industry, but the rocket industry has always called it nitrogen tetroxide or NTO and this is the designation used in this book. It is a high‐density yellow‐brown liquid (specific gravity of 1.44). NTO is hypergolic with hydrazine, monomethylhydrazine (MMH), and unsymmetrical dimethylhydrazine (UDMH); it is the most common storable oxidizer employed today. Its liquid temperature range is narrow and thus easily accidentally frozen or vaporized. It is only mildly corrosive when pure but forms strong acids when moist or allowed to mix with water, readily absorbing moisture from the air. It can be stored indefinitely in sealed containers made of compatible materials. NTO is hypergolic with many fuels but can also cause spontaneous ignition with many common materials, such as paper, grease, leather, or wood. When it decomposes, the resulting NO2 fumes are reddish brown and extremely toxic. Because of its high vapor pressure, it must be kept sealed in relatively heavy tanks. The freezing point of N2O4 may be lowered (for example, by adding a small amount of nitric oxide or NO) but at the expense of a higher vapor pressure and slightly reduced performance. Such mixtures of NO and N2O4 are called mixed oxides of nitrogen (MON) and different grades are available between 2 and 30% by weight NO content. Several MON oxidizers have flown.

Nitrogen tetroxide is used as the oxidizer with UDMH in many Russian engines and in almost all their small thrusters. It was also used with MMH fuel in the Space Shuttle orbital maneuver system and multithruster reaction control system, and is still used in many U.S. spacecraft propulsion systems. For many applications, care has to be taken to avoid freezing the nitrogen tetroxide. The OSHA eight‐hour personnel exposure limit is 5 ppm NO2 or 9 mg/m3.

Nitrous Oxide (N2O)

This oxidizer is also known as dinitrogen monoxide and as “laughing gas” (a medical anesthetic), see Refs. 7–9 and 7–10. It is a less potent oxidizer than the other four listed above. At ambient temperature it is not flammable but supports combustion at elevated temperatures. Its handling and safety are discussed in Refs. 7–10 and 7–11. Compared to nitrogen tetroxide (N2O4) it is much less toxic, its eight‐hour exposure limit being about 20 times less severe. Various catalysts will cause the decomposition of N2O into O2 and N2, and many materials and some impurities are incompatible with N2O. Its cryogenic liquid temperature range is very narrow, see Table 7–1 for boiling and freezing points; liquid N2O is not as cold as LOX. Although somewhat difficult to control, this oxidizer can self‐pressurize itself, avoiding the need for separate helium pressurizing systems.

Nitrous oxide has flown in space for the last 14 years. One principal recent flight application has been that of liquid nitrous oxide with HTPB (hydroxyl‐terminated polybutadiene) as a solid fuel in a hybrid rocket propulsion system (Ref. 7–12) for a suborbital manned space plane, which is discussed in Chapter 16. In a bipropellant combination, it has also been tested as the oxidizer with organic liquid fuels and gaseous monopropellants using a catalyst.

Oxidizer Cleaning Process

The internal surfaces of all newly manufactured components for rocket engine liquid oxidizer systems (pipes, tanks, valves, pumps, seals, injectors, etc.) usually undergo a special cleaning process for removing all traces of organic materials or other impurities that can react with the oxidized and cause bubbles or overpressure. This cleaning includes removal of even minute amounts of grease, oils, machine cuttings, fluids, paints, and small deposits of carbon, plastics, and common dust. The cleaning process may differ for each particular oxidizer and consists of scraping‐off visible deposits, successive multiple flushings and rinsing (with appropriate liquids), followed by drying with clean hot air. After the components are “oxygen clean” they are usually sealed to avoid any contamination prior to assembly or operation.

7.3 LIQUID FUELS

As with oxidizers, many different chemicals have been proposed, investigated, and tested as fuels. Liquid fuels not listed in this chapter that have been used in experimental rocket engines, in older experimental designs, and in older production engines include aniline, furfuryl alcohol, xylidine, gasoline, hydrazine hydrate, borohydrides, methyl and/or ethyl alcohol, ammonia, and some mixtures of these.

Hydrocarbon Fuels

Petroleum derivatives encompass a large variety of different hydrocarbon chemicals that can be used as rocket fuels. The most common are types already in use with other engines applications, such as gasoline, kerosene, diesel oil, and turbojet fuel. Their physical properties and chemical composition vary widely with the type of crude oil from which they were refined, with the chemical process used in their production, and with the accuracy of control exercised in their manufacture. Typical values are listed in Table 7–2.

Table 7–2 Properties of Some Typical Hydrocarbon Fuels Made from Petroleum

Jet Fuel Kerosene Aviation Gasoline 100/130 RP‐1 RP‐2
Specific gravity at 289 K   0.78   0.81   0.73 0.80–0.815 0.80–0.815
Freezing point (K) 213 (max.) 230    213    225    225   
Viscosity at 289 K (centipoise)   1.4    1.6    0.5  0.75 (at 289 K) 0.75 (at 289 K)
Flash pointa (K) 269    331    244    333    333   
ASTM distillation (K)
 10% evaporated 347    337    458–483 458–483
 50% evaporated 444    363   
 90% evaporated 511    391   
Reid vapor pressure (psia) 2 to 3 Below 1   7   
Specific heat (kcal/kg‐K)   0.50   0.49   0.53   0.50   0.50
Average molecular mass (g/mol) 130    175     90   
Sulfur, total mg/kg 30 (max) 1 (max)

a Tag closed cup method.

In general, these petroleum fuels form yellow‐white, brilliantly radiating flames and give good performance. They are relatively easy to handle and abundant at relatively low cost. Specifically refined petroleum products particularly suitable as rocket propellants carry the designation of RP‐1 or ‐2 (rocket propellant number 1 or 2). These are basically a kerosene‐type fuel mixture of hydrocarbons with a somewhat narrow range of densities and vapor pressures. Under some conditions, hydrocarbon fuels may form carbon deposits on the inside of cooling passages, impeding heat transfer and raising wall temperatures. Ref. 7–13 indicates that such carbon formations depend on fuel temperature in the cooling jacket, the particular fuel, the rate of heat transfer, and the chamber wall material. Fortunately, RP‐1 is low in olefins and aromatics (the ones likely to cause such solid carbonaceous deposits inside fuel cooling passages). RP‐1 has been used with liquid oxygen in many early rocket engines (see Figs. 5–1 to 5–6 and Ref. 7–14). A very similar kerosene‐type fuel is used in Russia today. In about 2003, RP‐2 was substituted for RP‐1 in U.S. rocket applications; the principal difference is a reduced sulphur content (see Table 7–2) because this impurity was believed to have caused corrosion in cooling jackets of certain thrust chambers. See Ref. 7–14.

Methane (CH4) is a cryogenic liquid hydrocarbon and the main constituent of ordinary “natural gas.” Liquid methane is abundant and relatively low in cost. Compared to petroleum‐refined hydrocarbons, it has highly reproducible properties. A bipropellant consisting of LOX and methane gives a lower characteristic velocity images and/or specific impulse Is than LOX‐hydrogen but higher than LOX‐kerosene or LOX‐RP‐1, see Table 5–5. Several rocket propulsion organizations in different countries have been developing and ground testing experimental liquid propellant thrust chambers and/or small reaction control thrusters with LOX‐CH4. Since the density of liquid methane is approximately six times higher than liquid hydrogen, methane fuel tanks can be much smaller and less costly. Methane is being considered as a bipropellant fuel (with LOX) for future rocket engines in large, multistage space launch vehicles for both manned and unmanned missions to Mars. In anticipation of three Mars flights, a new larger oxygen‐methane engine is being developed in the United States; while these investigations are presently ongoing, there have not been any flights (as of 2015, personal communication with J. H. Morehart of The Aerospace Corporation).

Liquid Hydrogen

As shown in Table 5–5, hydrogen when burned with oxygen gives a high performance colorless flame (shock waves may however be visible in the rocket plume). Liquid hydrogen is an excellent regenerative coolant. Of all known fuels, liquid hydrogen is by far the lightest and the coldest, having a specific gravity of 0.07 and a boiling point of about 20 K. This extremely low fuel density requires bulky and large fuel tanks, which necessitate rather large vehicle volumes. The extremely low storage temperatures limit available materials for pumps, cooling jackets, tanks, and piping because many metals become brittle at such temperatures.

Because of their very low temperatures, liquid hydrogen tanks and supply lines have to be well insulated to minimize hydrogen evaporation or the condensation of moisture or air on the outside, with subsequent formations of liquid or solid air and ice. A vacuum jacket has often been used in addition to the insulating materials. All common liquids and gases solidify in liquid hydrogen. Such solid particles in turn plug orifices and valves. Therefore, care must be taken to purge all lines and tanks of air and moisture (flushing with helium and/or creating vacuums) before introducing any liquid hydrogen propellant. Mixtures of liquid hydrogen and solid oxygen or solidified air can actually explode.

Liquid hydrogen is manufactured from gaseous hydrogen by successive compression, cooling, and expansion processes. Hydrogen has two isomers, namely, orthohydrogen and parahydrogen, which differ in the orientation of their nuclear spin state. As hydrogen is liquefied and cooled to lower and lower temperatures, the relative equilibrium composition of ortho‐ and parahydrogen changes. The transformation from orthohydrogen to parahydrogen is exothermic and results in excessive boil‐off, unless complete conversion to parahydrogen is achieved during liquefaction. Parahydrogen is the only stable form of liquid hydrogen.

Hydrogen gas, when mixed with air, is highly flammable and explosive over a wide range of mixture ratios. To avoid this danger, escaping excess hydrogen gas (at tank vent lines) is often intentionally ignited or “flared” in air. Liquid hydrogen is used with liquid oxygen in the Delta IV, Ariane V and ‐HII launch vehicles, Centaur upper stage, and upper stage space engines developed in Japan, Russia, Europe, India, and China.

Hydrogen burning with oxygen forms a nontoxic essentially invisible exhaust. This propellant combination has been applied successfully to space launch vehicles because of its high specific impulse (payload capability usually increases greatly for relatively small increases in specific impulse) even if the low density of liquid hydrogen makes for a very bulky fuel tank, a large vehicle, and relatively higher drag. Studies have shown that, when burned with liquid oxygen, hydrocarbons (such as methane or RP‐1) give a small advantage in space launch vehicle first stages. Here, the higher average propellant density allows a smaller vehicle with less mass and lower drag, which compensates for the lower specific impulse of the hydrocarbon when compared to hydrogen. Therefore, some concepts exist for operating the same booster‐stage rocket engines initially with a hydrocarbon fuel and then switching during flight to hydrogen. Engines using LOX with these two fuels, namely, hydrocarbon and hydrogen, are called tripropellant engines. They have not yet been fully developed or flown. Some work on experimental tripropellant engines was done in Russia, but there is no current effort known to the authors.

Hydrazine (N2H4)

Reference 7–15 gives a good discussion of this propellant, which is widely used as a bipropellant fuel as well as a monopropellant. Hydrazine, monomethylhydrazine (MMH), and unsymmetrical dimethylhydrazine (UDMH) have similar physical and thermochemical properties. Hydrazine is a toxic, colorless liquid with a high freezing point (275.16 K or 35.6 °F). Hydrazine tanks, pipes, injectors, catalysts, and valves are usually electrically heated to prevent freezing in cool ground weather or in outer space. Hydrazine has a short ignition delay and is spontaneously ignitable with nitric acid, nitrogen tetroxide, and concentrated hydrogen peroxide.

Pure anhydrous hydrazine is a stable liquid; it has been safely heated to near 416 K. It has been stored in sealed tanks for over 15 years. Space probes with monopropellant hydrazine have been operating for about 40 years while traveling beyond the boundary of the solar system. With impurities or at higher temperatures it decomposes releasing energy. Under pressure shock (blast wave or adiabatic compression) hydrazine vapor or hydrazine mist can decompose at temperatures as low as 367 K. Under some conditions this decomposition can be a violent detonation, and this has caused problems in cooling passages of experimental injectors and thrust chambers. Harmful effects to personnel may result from ingestion, inhalation of its toxic vapors, or prolonged contact with skin. The American Conference of Government Industrial Hygienists (ACGIH) recommended eight‐hour personnel exposure limit is 0.01 ppm or 0.013mg/m3. Hydrazine is a known animal carcinogen and a suspected carcinogen for people.

Hydrazine vapors may form explosive mixtures with air. If hydrazine is spilled on a porous surface or a cloth, spontaneous ignition with air may occur. It reacts with many materials, and care must be exercised to avoid contact with storage materials that cause decomposition (see Refs. 7–15 and 7–16). Tanks, pipes, injectors, catalysts, or valves must be cleaned and free of all traces of impurities. Compatible materials include certain stainless steels (303, 304, 321, or 347), nickel, and the 1100 and 3003 series of aluminum. But iron, copper, and its alloys (such as brass or bronze), monel, magnesium, zinc, and some types of aluminum alloy must be avoided.

Hydrazine cannot be safely used in cooling jackets of bipropellant thrust chambers because it will explode when a certain detonation temperature is exceeded. Such explosions occur when heat transfer is unusually high or immediately after thrust termination when the cooling fluid is stagnant; heat soaking back from hot inner walls can overheat any trapped hydrazine and this may cause a violent decomposition. Presently, bipropellant chambers with hydrazine are radiation cooled (using molybdenum or rhenium metals) and/or are ablatively cooled since internal thrust chamber insulation layers can erode and oxidize. Hydrazine injectors have to be designed without inside stagnant pockets or manifolds where liquid hydrazine can remain just after shutdown. Helium purges may be used to solve this situation. Radiation cooling and ablation cooling are discussed in Chapter 8. Reference 7–2 describes recent work on green‐propellant hydrazine replacements.

Unsymmetrical Dimethylhydrazine [(CH3)2NNH2]

This derivative of hydrazine, abbreviated as UDMH, is often used instead of or in mixtures with hydrazine because it forms a thermally more stable liquid. Furthermore, it has a lower freezing point (215.9 K) and a lower boiling point (335.5 K) than hydrazine itself. When UDMH is burned with an oxidizer it gives only slightly lower values of Is than pure hydrazine. UDMH has been used when mixed with 30 to 50% hydrazine or with 25% hydrazine hydrate. UDMH is used in many Russian and Chinese small thrusters and some main rocket engines. The ACGIH (Ref. 7–6) recommended eight‐hour personnel exposure limit for vapor is 0.01 ppm, and UDMH is a known animal carcinogen.

Monomethylhydrazine (CH3NHNH2)

Monomethylhydrazine, abbreviated as MMH, is being used extensively in U.S. spacecraft rocket engines, particularly in small attitude control engines, usually with N2O4 as the oxidizer. It has a better shock resistance to blast waves and a better liquid temperature range than pure hydrazine. Like hydrazine, its vapors are easily ignited in air; flammability limits are from 2.5 to 98% by volume at atmospheric sea‐level pressure and ambient temperature. Any materials compatible with hydrazine are also compatible with MMH. The specific impulse with storable oxidizers usually is 1 or 2% lower with MMH than with N2H4.

Both MMH and UDMH are soluble in many hydrocarbons, though pure hydrazine is not. Monomethylhydrazine, when added in quantities of 3 to 15% by volume to hydrazine, has a substantial quenching effect on the explosive decomposition of hydrazine. MMH decomposes at 491 K, whereas pure hydrazine can explode at 369 K when subjected to certain pressure shocks. MMH is a known animal carcinogen and the ACGIH recommended personnel 8‐hour exposure limit is 0.01 ppm. Of all hydrazines, MMH vapor is the most toxic when inhaled.

7.4 LIQUID MONOPROPELLANTS

The simplicity associated with monopropellant feed and control systems makes this kind of propellant very attractive for certain applications. Hydrazine is used extensively as monopropellant in small attitude and trajectory control rockets for the control of satellites and other spacecraft and also in hot gas generators (this is discussed in a preceding section). Other monopropellants (ethylene oxide or nitromethane) were tried experimentally, but are no longer used today. Concentrated hydrogen peroxide (usually 90%) was used for small thrusters between 1945 and 1965 and for monopropellant gas generator in the United States, United Kingdom or Britain, and Germany and is still being used in Russia and elsewhere (see Section 7.2. and Example 5–1). Presently, green monopropellants based on hydroxylammonium nitrate or ammonium dinitrate aqueous solutions are under development as alternatives to hydrazine.

Decomposition of monopropellants can be achieved thermally (electrically or flame heated) or by a catalytic material. A monopropellant must be chemically and thermally sufficiently stable to ensure proper liquid storage properties, and yet it must be easily decomposed and reactive to quickly give complete decomposition.

Hydrazine as a Monopropellant

Hydrazine stores well and is used as monopropellant when decomposed by a suitable solid catalyst; such catalyst often needs to be preheated for fast startups and/or for extending the useful catalyst life. Iridium on a porous alumina base is an effective catalyst at room temperature. At elevated temperature (about 450 K) many other materials decompose hydrazine, including iron, nickel, and cobalt. See Ref. 7–15. Different catalysts and different configurations produce different decomposition products, resulting in gases of varying composition and temperature. Monopropellant hydrazine units are used in gas generators and/or attitude control rockets, and as propellant in electrothermal and arcjet thrusters. A typical hydrazine monopropellant thrust chamber, its injection patterns, and its decomposition reaction are described in Fig. 8–14. Typical operating parameters are shown in Fig. 7–3 and Table 7–1.

Five plots with ammonia decomposition on the horizontal axes. In the plot at the bottom, three different curves are plotted and marked N2, H2, and NH3. N2 and NH3 are dashed line curves. One solid line curve is plotted in each of the four plots above.

Figure 7–3 Operating parameters for decomposed hydrazine at the exit of a catalytic reactor as a function of the ammonia dissociation fraction.

Adapted with permission from Ref. 7–15

The catalytic decomposition of hydrazine can be described ideally as a two‐step process; such a simplified scheme ignores other steps and intermediate products. First, hydrazine (N2H4) decomposes into gaseous ammonia (NH3) and nitrogen (N2); this reaction is highly exothermic (e.g., it releases heat). Second, hot ammonia decomposes further into nitrogen and hydrogen gases, but this reaction is endothermic and absorbs heat. These simplified reaction steps may be written as

Here, x is the degree of ammonia dissociation; it is a function of catalyst type, size, and geometry; chamber pressure; and dwell time within the catalyst bed. Figure 7–3 shows several calculated rocket engine parameters for hydrazine monopropellant as a function of x. The shown values are for an ideal thruster at 1000 psia chamber pressure with a nozzle area ratio of 50 expanding at high altitudes. The best specific impulse is attained when little ammonia is allowed to dissociate.

For more than 25 years now a number of investigations have been conducted to find a substitute for hydrazine monopropellant because it is toxic, its freezing point is relatively high (275 K or 2 °C), and it requires heating of all parts it comes in contact with. To date, none of the propellants that have been evaluated so far are fully satisfactory. One possible candidate is a monopropellant (identified as LMP‐1035) made from ammonium dinitramine (ADN, a solid) dissolved in small amounts of water; it was first flown in the European satellite PRISMA, which was launched by the Russian DNEPR launch vehicle on June 15, 2010. This relatively new experimental monopropellant has a higher density than hydrazine, is not toxic, and has an acceptable liquid temperature range but a slightly lower performance. However, LMP‐1035 will decompose only when its catalyst has been preheated to 533 K or 500 °F. This higher decomposition temperature requires extra thermal insulation and this heating may shorten the catalyst life. Another substitute candidate is based on hydroxyl ammonium nitrate (HAN) and this monopropellant has also been under investigation. Both HAN‐ and ADN‐based monopropellants are scheduled for tests in U.S. satellites, but to date neither has been fully proven. Investigations continue on other potential hydrazine monopropellant replacements.

7.5 GASEOUS PROPELLANTS

Propellants that store as a gas at ambient temperatures or cold gas propellants have been used successfully in reaction control systems (RCSs) for more than 70 years. The phrase “cold gas” distinguishes them from “warm gas,” those expanded after being heated. The applicable engine system components are relatively simple, consisting of one or more high‐pressure gas tanks, multiple simple nozzles (often aluminum or plastic) each with an electrical control valve, a pressure regulator, and provisions for filling and venting the gas. Tank sizes are smaller when the storage pressures are high. Because these pressures can typically be between 300 and 1000 MPa (about 300 to 10,000 psi), strong (often thick‐walled and massive) gas tanks are needed.

Typical cold gas propellants and some relevant properties and characteristics are listed in Table 7–3. Nitrogen, argon, dry air, and helium have been employed for spacecraft RCS. With high‐pressure hydrogen or helium as the cold gas, the specific impulse is much higher, but because their densities are lower they require much larger gas storage volumes and/or more massive high‐pressure tanks; in most applications any extra inert mass outweighs the advantages of better performance. In a few applications the gas (and sometimes also its storage tank) may be heated electrically or chemically. This improves the specific impulse and allows for smaller tanks, but it also introduces complexity (see Chapter 17).

Table 7–3 Properties of Gaseous Propellants Used for Auxiliary Propulsion

Propellant Molecular Mass Densitya (lbm/ft3) Specific Heat Ratio k Theoretical Specific Impulseb (sec)
Hydrogen  2.0  1.77 1.40 284
Helium  4.0  3.54 1.67 179
Methane 16.0 14.1  1.30 114
Nitrogen 28.0 24.7  1.40  76
Air 28.9 25.5  1.40  74
Argon 39.9 35.3  1.67  57
Krypton 83.8 74.1  1.63  50

a At 5000 psia and 20 °C.

b In vacuum with nozzle area ratio of 50:1 and initial temperature of 20 °C.

Selection of propellant gas, storage tanks, and RCS design depend on many factors, such as volume and mass of the storage tanks, maximum thrust and total impulse, gas density, required maneuvers, duty cycle, and flight duration. Cold gas systems have been used for producing total impulses of up to 22,200 N‐sec or 5000 lbf‐sec; higher values usually require liquid mono‐ or bipropellants.

During short operations (most of the gas utilized in a few minutes while the main engine is running), gas expansions will be close to adiabatic (no heat absorption by gas) which are often analyzed as isentropic expansions. The gas temperature in high‐pressure storage tanks, the (unregulated) pressures, and the specific impulse will drop as the gas is being utilized. For intermittent low‐duty‐cycle operations (months or years in space) heat from the spacecraft may transfer to the gas and tank temperatures stay essentially constant; then expansions will be nearly isothermal.

Advantages and disadvantages of cold gas thrusters and systems are further described in Section 8.3 in the discussion of low thrust.

7.6 SAFETY AND ENVIRONMENTAL CONCERNS

To minimize any hazards and potential damages inherent in reactive propellant materials, it is necessary to be very conscientious of all likely risks and hazards (see Refs. 7–5, 7–17, and 7–18). These relate to toxicity, explosiveness, fires and/or spill dangers, and others mentioned in Section 7.1. Before an operator, assembler, maintenance mechanic, supervisor, or engineer is allowed to transfer or use any particular propellant, he or she should receive safety training in that particular propellant, its characteristics, its safe handling or transfer, potential damage to equipment or the environment, and in countermeasures for limiting the consequences in case of accidents. Staff must also be aware of potential hazards to personnel health, first aid remedies in case of contact exposure of the skin, ingestion, or inhaling, and of how to use safety equipment. Examples of safety equipment are protective clothing, face shields, detectors for toxic vapors, remote controls, warning signals, and/or emergency water deluges. Personnel working with or close to highly toxic materials usually must undergo periodic health monitoring. Also, rocket engines need to be designed for safety to minimize the occurrence of leaks, accidental spills, unexpected fires, and/or other potentially unsafe conditions. Most organizations have one or more safety specialists who review the safety of test plans, manufacturing operations, designs, procedures, and/or safety equipment. With proper training, equipment, precautions, and design safety features, all propellants can be handled safely.

When safety violation occur or if an operation, design, procedure, or practice is found to be (or appears to be) unsafe, then a thorough investigation of the particular item or issue should be undertaken, the cause of the lack of safety should be investigated and identified, and an appropriate remedial action should be selected and initiated as soon as possible.

The discharge of toxic exhaust gases to the environment and their wind dispersion may cause exposure to operating personnel as well as the general public in nearby areas, and result in damage to plants and animals. This is discussed in Section 20.2. The dumping or spilling of toxic liquids contaminates subterranean aquifers and surface waters, and their vapors pollute the air. Today the type and amount of gaseous and liquid discharges are regulated and monitored by government authorities. These discharge quantities must be controlled or penalties will be assessed against violators. Obtaining a permit to discharge can be a lengthy and involved procedure.

One way to enhance safety against accidents (explosions, fires, spills, bullet impacts, etc.) is to use gelled propellants. They have additives to make them thixotropic materials, with the consistency of very thick paint when at rest; they readily flow through valves or injectors when under a pressure gradient or shear stress. Different gelling agents have been extensively investigated with several common propellants and the reader should consult the literature for details (e.g., Ref. 7–19 or Section 7.5 of the Eighth edition of this book). As far as the authors know, there is as yet no production engine using gelled propellants.

SYMBOLS

Id density specific impulse, sec
Is specific impulse, sec
k ratio of specific heats for cold‐gas propellants
mf/m0 ratio of final to initial mass
r mixture ratio (mass flow rate of oxidizer to mass flow rate of fuel)

Greek Letters

δav average specific gravity of mixture
δf specific gravity of fuel
δo specific gravity of oxidizer
ρav, ρf, ρo densities average, fuel, and oxidizer, kg/m3 (lbm/ft3)

PROBLEMS

  1. Plot the variation of the density specific impulse (product of average specific gravity and specific impulse) against mixture ratio and explain the meaning of the curve, using the theoretical shifting specific impulse values in Fig. 5–1 and the specific gravities from Fig. 7–1 or Table 7–1 for the liquid oxygen/RP‐1 propellant combination.Answer: Check point at images.
  2. Prepare a list comparing the relative merits of liquid oxygen and of nitrogen tetroxide as rocket engine oxidizers.
  3. Derive Eq. 7–1 for the average specific gravity.
  4. A rocket engine uses liquid oxygen and RP‐1 as propellants at a design mass mixture ratio of 2.40. The pumps used in the feed system are basically constant‐volume flow devices. The RP‐1 hydrocarbon fuel has a nominal temperature of 298 K which can vary by about ±25 °C. The liquid oxygen is nominally at its boiling point (90 K), but, after the tank is pressurized, this temperature may increase by up to 30 K during long time storage. What are the extreme mixture ratios that result under unfavorable temperature conditions? If this engine has a nominal mass flow rate of 100 kg/sec and duration of 100 sec, what is the maximum residual propellant mass when the other propellant is fully consumed? Use the curve slopes of Fig. 7–1 to estimate changes in density. Assume that the specific impulse is constant for relatively small changes in mixture ratio, that small vapor pressure changes have no influence on the pump flow, that there is no evaporation of the oxygen in the tank, and that the engine has no automatic control for mixture ratio. Assume also zero residual propellant at the design condition.
  5. The vehicle stage propelled by the rocket engine in Problem 7–4 has a design mass ratio mf/m0 of 0.50 (see Eq. 4–6). For the specific impulse use a value halfway between the shifting and the frozen equilibrium curves of Fig. 5–1. How much will the worst combined changes in propellant temperatures affect the mass ratio and the ideal gravity‐free vacuum velocity?
    1. What should be the approximate percent ullage volume for nitrogen tetroxide tank when the vehicle is exposed to ambient temperatures between about 50 °F and about 150 °F?
    2. What is maximum tank pressure at 150 °F.
    3. What factors should be considered in part (b)?Answer: (a) 15 to 17%; the variation is due to the nonuniform temperature distribution in the tank; (b) 6 to 7 atm; (c) vapor pressure, nitrogen monoxide content in the oxidizer, chemical reactions with wall materials, or impurities that result in largely insoluble gas products.
  6. An insulated, long, vertical, vented liquid oxygen tank has been sitting on the sea‐level launch stand for a period of time. The surface of the liquid is at atmospheric pressure and is 10.2 m above the closed outlet at the bottom of the tank. What will be the temperature, pressure, and density of the oxygen at the tank outlet if (a) liquid oxygen is allowed to circulate within the tank and (b) if it is assumed that there is no heat transfer throughout the tank wall to the liquid oxygen?

REFERENCES

  1. 7–1. S. F. Sarner, Propellant Chemistry, Reinhold, New York, 1966.
  2. 7–2. C. H. McLean et al., “Green Propellant Infusion Mission Program Overview, Status, and Flight Operations,” AIAA Paper 2015‐3751, Orlando, FL, 2015; R. L. Sackheim and R. K. Masse, “Green Propulsion Advancement—Challenging the Propulsion Maturity of Monopropellant Hydrazine,” AIAA Paper 2013‐3988, July 2013; K. Anflo and R. Moellerberg, “Flight Demonstration of a New Thruster and Green Propellant Technology on the PRISMA Satellite,” Acta Astronautica, Vol. 65, 1238–1249, Nov. 2009.
  3. 7–3. Chemical Rocket Propellant Hazards, Vol. 1, General Safety Engineering Design Criteria, Chemical Propulsion Information Agency (CPIA) Publication 194, Oct. 1971; http://handle.dtic.mil/100.2/AD0889763
  4. 7–4. L. C. Sutherland, “Scaling Law for Estimating Liquid Propellant Explosive Yields, Journal of Spacecraft and Rockets, Mar.–Apr. 1978, pp. 124–125; doi: 10.2514/3.28002.
  5. 7–5. NIOSH Pocket Guide to Chemical Hazards, DHHS (NIOSH) Publication No. 2005‐149, Department of Health and Human Services, Washington DC, 454 pages, Sept. 2005; http://www.cdc.gov/niosh/docs/2005‐149/pdfs/2005‐149.pdf
  6. 7–6. 2014 Threshold Limit Values for Chemical Substances and Physical Agents and Biological Exposure Indices, American Conference of Government Industrial Hygienists, Cincinnati, OH, 2013, ACGIH Publication #0113, ISBN: 978‐1‐607260‐59‐2 (revised periodically).
  7. 7–7. H. E. Malone, The Analysis of Rocket Propellants, Academic Press, New York, 1976.
  8. 7–8. M. C. Ventura, “Long Term Storability of Hydrogen Peroxide,” AIAA Paper 2005‐4551, Jul. 2005; A. Pasini et al., “Testing and Characterization of a Hydrogen Peroxide Monopropellant Thruster,” Journal of Propulsion and Power, Vol. 24, No. 3, May–June 2008.
  9. 7–9. “Nitrous Oxide Summary Properties,” M4 Liquid Propellant Manual, Chemical Propulsion Information Analysis Center, Columbia, MD, November 2009.
  10. 7–10. F. Mackline, C. Grainger, M. Veno, and S. Benson, “New Applications for Hybrid Propulsion,” AIAA Paper 2003‐5202, 20‐23 July 2003.
  11. 7–11. R. Haudy, “Nitrous Oxide/Hydrocarbon Fuel Advanced Chemical Propulsion: DARPA Contract Overview,” Qualis Corporation, Huntsville AL, 2001.
  12. 7–12. Yen‐Sen Chen, A. Lau, T. H. Ch, and S.S. Wu, “N2O‐HTPB Hybrid Rocket Combustion Modeling with Mixing Enhancement Designs,” AIAA Paper 2013‐3645, 14‐17 July, 2013.
  13. 7–13. K. Liang, B. Yang, and Z. Zhang, “Investigation of Heat Transfer and Coking Characteristics of Hydrocarbon Fuels,” Journal of Propulsion and Power, Vol. 14, No. 5, Sept.–Oct. 1998; doi: 10.2514/2.5342.
  14. 7–14. Detail Specification—Propellant Rocket Grade Kerosene, Department of Defense, Washington DC, MIL‐DTL‐25576D, 20 May 2005.
  15. 7–15. E. W. Schmidt, Hydrazine and Its Derivatives, Preparation, Properties, Applications, 2nd ed., John Wiley & Sons, New York, 2001.
  16. 7–16. O. M. Morgan and D. S. Meinhardt, “Monopropellant Selection Criteria—Hydrazine and other Options,” AIAA Paper 99‐2595, June 1999; doi: 10.2514/6.1999-2595.
  17. 7–17. J. A. Hannum, Hazards of Chemical Rockets and Propellants, Vol. I, Safety, Health and the Environment, AD‐A160951, and Vol. III, Liquid Propellants, Chemical Propulsion Information Analysis Center, AD‐A158115, CPIA Publication 394, 1984.
  18. 7–18. Emergency Response Guidebook, U.S. Dept. of Transportation, Pipeline and Hazardous Materials Safety Administration, Emergency Response Guidebook 2012, 392 pages. http://phmsa.dot.gov/staticfiles/PHMSA/DownloadableFiles/Files/Hazmat/ERG2012.pdf
  19. 7–19. K. Madlener, et al., “Characterization of Various Properties of Gel Fuels with Regard to Propulsion Application,” AIAA Paper 2008‐4870, Jul. 2008; J. von Kampen, F. Alberio, and H. K. Ciezki, “Spray and Combustion Characteristics of Aluminized Gelled Fuels with an Impinging Jet Injector,” Aerospace Science and Technology, Vol. 11, No. 1, Jan. 2007, pp. 77–83.
..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset