Sie sind auf Seite 1von 20

Monk, J. C.

“Propulsion Systems”
The Engineering Handbook.
Ed. Richard C. Dorf
Boca Raton: CRC Press LLC, 2000

© 1998 by CRC PRESS LLC


176
Propulsion Systems
176.1 Performance Characteristics
176.2 Liquid Rocket Engine Cycles
Pressure-fed • Expander • Gas Generator • Staged Combustion
176.3 Major Components
Main Injector • Thrust Chamber • Turbomachinery
176.4 System Preliminary Design Process
176.5 Conclusion

Jan C. Monk
National Aeronautics and Space Administration

Rocket propulsion is an application of Newton's first, second, and third laws of motion. Newton's
first law of motion states that a particle not subjected to external forces remains at rest or moves
with constant velocity in a straight line. A rocket lifting off the launch pad goes from a state of rest
to a state of motion. Newton's second law of motion states that force equals mass times
acceleration. Force in the equation is the rocket thrust, where mass is the amount of rocket fuel
being burned and converted into gas, which expands and then escapes from the rocket. As the gas
exits the combustion chamber through a nozzle, it picks up speed. Newton's third law of motion
states for every action, there is an equal and opposite reaction. With rockets, the action is the
expelling of gas out of the engine; the reaction is the force or thrust of the rocket in the opposite
direction.

176.1 Performance Characteristics


In the process of producing thrust, rocket engines generate more power per unit weight than any
other engine. To enable a rocket to climb into low-earth orbit, it is necessary to achieve velocities
in excess of 28 000 km per hour. Escape velocity is a speed of about 40 250 km per hour. To
achieve these velocities, the rocket engine must burn a large amount of fuel and push the resulting
gas out of the engine as rapidly as possible. Containing and controlling this power is the basic
challenge in the development of these devices. For example, the power density produced by liquid
hydrogen (LH2) turbomachinery utilized by the space shuttle main engine (SSME) is
approximately 83 horsepower per pound of turbopump weight.
Rocket propulsion system design solutions are quite varied: thrust levels from ounces to millions
of pounds force; liquid and solid propellants; and liquid systems with pressures that are
maintained by turbopumps or pressurized tanks. Liquid system applications vary from small

© 1998 by CRC PRESS LLC


pressure-fed, storable monopropellant thrusters for keeping satellites stationary, to large
turbopump-fed, cryogenic bipropellant engines for boost propulsion. Combustion chamber
pressures vary from a few pounds per square inch (psi) to several thousand psi. Generally, liquid
propulsion systems consist of a propellant feed system, an injector, a combustion chamber, and a
nozzle. The propellant feed system includes ducting and valves for controlling flows and, in the
case of pump-fed systems, turbomachinery that draws propellants from lightweight propellant
tanks and increases the pressure to the level necessary to support the desired combustion chamber
pressure.
The ideal rocket propulsion equation is

M0
¢Videal = g0 Isp ln (176:1)
M1

where ¢Videal is the ideal delta velocity imparted on a vehicle, g0 is the gravitational constant, Isp
is the propulsion system's specific impulse, M0 is the initial mass of the vehicle, and M1 is the
final or burnout mass of the vehicle. This equation provides two important performance
parameters: specific impulse, which is a measure of propulsion system efficiency expressed in
seconds, and vehicle burnout mass, which includes all structures (tankage, thrust structure, etc.),
residual propellants, engine systems, feed systems, pressurization systems, auxiliary systems,
electronic systems, upper stages, payload supporting structures, and the payload itself.
One of the more important internal rocket engine parameters is characteristic exhaust velocity,
commonly referred to as C-star (C ¤ ) , which relates combustion chamber pressure, chamber throat
area, and propellant flow rate. Theoretical characteristic exhaust velocity C ¤ is computed as
follows:

Pns At g0
C¤ = (176:2)
w_ tc

where Pns is nozzle stagnation pressure in psi, At is throat area in square inches, and w_ tc is
chamber propellant mass flow rate in pounds-mass per second. A number of losses will reduce the
actual C ¤ realized. These losses are generally a function of injector design and are related to
mixture ratio maldistribution, mixing, etc. The actual C ¤ realized by a given design is,
¤
Cact = ´c¤ C ¤ (176:3)

where ´c¤ is C ¤ efficiency, typically between 0.80 and 0.99.


Another useful parameter is thrust coefficient, which relates thrust F, chamber pressure, and
throat area as follows:

F = CF Pns At (176:4)

where CF is the thrust coefficient, Pns is nozzle stagnation pressure, and At is throat area. Once
again, additional parameters must be added to reflect actual values. This yields the following:

© 1998 by CRC PRESS LLC


F = ´C F CF Pns At ¡ Pa Ae (176:5)

where ´C F is thrust coefficient efficiency, typically between 0.90 and 0.97, Pa is local atmospheric
pressure in psi, and Ae is exit area in square inches. This equation yields thrust at any point
between sea level and vacuum conditions.
Specific impulse Isp is an overall efficiency term and is defined as

F
Isp = (176:6)
w_ t

where F is thrust level in pounds-force and w_ t is the total mass flow rate in pounds-mass per
second. Specific impulse can be computed for the engine or thrust chamber by utilizing either
engine thrust and flow rate or thrust chamber thrust and flow rate, as appropriate. Specific impulse
can also be computed if C ¤ and the thrust coefficient are known. This relationship is expressed as

C ¤ CF
Isp = (176:7)
g0

Again, one must maintain consistency between theoretical values and actual values.
Thrust and specific impulse are commonly calculated at either sea level or vacuum conditions for
reference or comparative purposes. Later discussions will refer to sea level thrust (Fsl ) , vacuum
thrust (Fvac ) , sea level specific impulse (Isp sl ) , and vacuum specific impulse (Isp vac ) .
Mixture ratio is the ratio between the oxidizer and fuel flow rates, and is expressed in equation
form as

w_ o
MR = (176:8)
w_ F

where w_ o is oxidizer flow rate in pounds per second and w_ F is fuel flow rate in pounds per second.
Mixture ratio can be computed for the engine or thrust chamber by utilizing either engine flow
rates or thrust chamber flow rates, as appropriate.
Expansion ratio " is a ratio of the thrust chamber nozzle exit area, Ae ; and the thrust chamber
throat area, At :

Ae
"= (176:9)
At

A more complete definition of these and other rocket engine equations, including solid propellant
systems, can be found in Rocket Propulsion Elements [Sutton, 1992].

176.2 Liquid Rocket Engine Cycles


A number of power cycles are available for liquid propellant systems. These include pressure-fed,

© 1998 by CRC PRESS LLC


expander, gas generator, and staged combustion cycles. Each cycle has advantages and
disadvantages; the one selected for a given application is determined after a series of system trade
studies. A description of a number of engine systems is given in Table 176.1. A brief description of
each of these power cycles follows.

Table 176.1Liquid Rocket Engine Characteristics

Pressure-fed
This system consists of a thrust chamber assembly, associated ducting and valves necessary for
control, pressurized tankage, and the pressurization system for the tankage. This system is widely
utilized for satellite attitude control, orbital transfer, and as auxiliary propulsion for most major
launch vehicles. Pressure-fed systems are perhaps the simplest of all propulsion systems, but are
performance limited because of the weight penalty associated with increasing chamber pressures.
As pressures increase, tank wall thickness and the mass of the gases needed to maintain tank
pressures increase. Tank pressures are set by chamber pressure plus pressure losses in the cooling
circuit (if any), injector, valves, and ducting. In most pressure-fed applications, combustion
chambers are passively cooled (i.e., film-cooled or radiative/ablative). The space shuttle utilizes
pressure-fed systems for the orbital maneuvering system and the reaction control system. A
schematic of a simple pressure-fed system is given in Fig. 176.1.

© 1998 by CRC PRESS LLC


Figure 176.1 Pressure-fed propulsion system schematic.

Expander
This is the simplest of the turbopump-fed systems primarily because the power source for the
turbines is the thrust chamber cooling circuit. Only the thrust chamber requires an ignition system.
Pump discharge pressures are set by chamber pressure plus pressure losses in the cooling circuit,
turbine, injector, valves, and ducting. The combustion chamber is regeneratively cooled. In some
applications, extensible radiation-cooled nozzle extensions are used to increase area ratio while
maintaining a short stowed length. Expander cycles are limited in the combustion chamber
pressure that can be attained because the energy available to drive the turbine(s) is obtained from
the combustion chamber cooling circuit. For applications that require operation at sea level, this
reduces the area ratio that can be achieved without side loads. Nozzle flow separation is discussed

© 1998 by CRC PRESS LLC


later. The RL10 engine is utilized by the Centaur upper stage for the Atlas-Centaur and
Titan-Centaur launch vehicles. A schematic of a simple expander system is given in Fig. 176.2 .

Figure 176.2 Expander engine system schematic.

Gas Generator
This is the most common engine cycle in use today. Turbine power is derived from a separate
combustor or gas generator (GG) which utilizes the same propellants as the main system. This hot
gas is routed through the turbopump turbines and is dumped overboard. Pump discharge pressures
are set by chamber pressure plus pressure losses in the cooling circuit, injector, valves, and
ducting. Because the gas generator is parallel to the main chamber, turbine pressure losses do not
impact pump discharge pressure in most designs. This highlights one of the disadvantages of this
cycle. The gas generator propellants are not used in the main chamber to produce thrust. Some
concepts use GG gases for cooling nozzle extensions, but the thrust added is minimal. Gas

© 1998 by CRC PRESS LLC


generators are operated at relatively low mixture ratios because turbine temperatures must be
maintained in the 1000 to 2000 degrees Rankine range. The main combustor mixture ratio is biased
higher to offset this parasitic flow. The combination of poor thrust efficiency of the GG gases and
main chamber mixture ratio bias results in a specific impulse penalty. The gas generator cycle was
utilized on the F-1 and J-2 engines of the Saturn V launch vehicle and is currently in use on the
Delta, Atlas, and Titan launch vehicles. A schematic of a simple gas generator system is given in
Fig. 176.3.

Figure 176.3 Gas generator engine system schematic.

© 1998 by CRC PRESS LLC


Staged Combustion
The staged combustion cycle provides the highest performance of conventional chemical rocket
engines. Turbine power is derived from a separate combustor or preburner which also utilizes the
same propellants as the main system. In bipropellant systems, the hot gas is routed through the
turbopump turbines to the main injector where it is mixed with the other propellant and is
combusted in the main chamber. Pump discharge pressures are set by chamber pressure plus
pressure losses in the cooling circuit, turbine, injector, valves, and ducting. Thrust chambers are
regeneratively cooled. Staged combustion cycle engines developed in the U.S. have utilized a
fuel-rich preburner. Several rocket engine systems developed in Russia have utilized an
oxidizer-rich preburner. In the former case, the fuel-rich hot gases are mixed with oxidizer in the
main chamber. In the latter, oxidizer-rich hot gases are mixed with fuel in the main chamber. The
staged combustion cycle utilizes all propellants in the main combustion chamber, which provides
maximum performance. A schematic of a simple staged combustion system is given in Fig. 176.4.

Figure 176.4 Staged combustion cycle engine system schematic.

© 1998 by CRC PRESS LLC


A variant of the staged combustion cycle is the full flow cycle, in which the oxidizer pump
turbine is driven with an oxidizer-rich preburner and the fuel pump is driven with a fuel-rich
preburner. This cycle offers some simplification in turbomachinery design because of the
simplified seal design between the pump end and the turbine end, and a significant reduction in
turbine temperatures because all the propellants can be utilized in the turbine drive circuits. This
concept is currently under study as a candidate engine system for the next generation launch
vehicle. A schematic of a simple full flow staged combustion system is given in Fig. 176.5 .

Figure 176.5 Full flow staged combustion cycle engine system schematic.

© 1998 by CRC PRESS LLC


176.3 Major Components
Main Injector
The purpose of the injector is to introduce propellants into the combustion chamber in a controlled
manner, to atomize the propellants, and to mix the propellants at the proper mixture ratio in a
homogenous manner. Mixture ratio variations across the injector face are one of the most common
problems that the designer will encounter, and these maldistributions lead to combustion efficiency
losses. In some cases, maldistributions are deliberately introduced. To enhance the durability of
combustion chambers, a film of fuel is injected at the outer circumference of the injector. In order
to produce a stable combustion process, baffle elements which are cooled with fuel are commonly
used. The most common injector concepts are coaxial, showerhead, and impinging. These are
illustrated in Fig. 176.6.

Figure 176.6 Injector concepts.

© 1998 by CRC PRESS LLC


The coaxial injector consists of a series of concentric tubes into which the oxidizer is introduced
through a center tube and the fuel introduced through the annular area formed by a second tube.
This type of injector is commonly used in oxygen-hydrogen engines. Impinging and showerhead
injectors, which are commonly used in oxygen-kerosene and storable propellant engines, consist of
a series of two sets of orifices. One injects the oxidizer, and the other the fuel. The number of
orifices, injection velocities, and injection angles are selected to provide consistent atomization and
mixing of propellants. Impinging injectors slant the orifices to impinge the two propellant streams
against each, enhancing mixing. The design utilized by the conventional impinging injector
consists of a series of concentric copper rings containing the injection orifices. The rings alternate
between oxidizer and fuel. The outer ring is generally a fuel ring and contains a set of smaller
orifices that control film coolant for the combustion chamber. Separate manifolding routes
propellants to each set of orifices.

Thrust Chamber
A number of design solutions have been utilized in thrust chambers, varying from passively cooled
ablatives to a number of regeneratively cooled concepts. In some applications, the thrust chamber
is composed of two separate components. The upper portionincluding the throat region and a
portion of the expansion regionis commonly called a combustion chamber. The lower
portionconsisting of the remainder of the expansion regionis called a nozzle. Regeneratively
cooled thrust chamber designs include brazed tube bundles, copper with milled channels, and steel
with milled channels. The bundled tube concept utilizes: steel tubes, pressed to vary the shape
necessary for formation of the overall thrust chamber shape; a structural shell in the combustion
chamber region with a number of straps spaced along the thrust chamber length for additional
strength; and necessary manifolding for inlet and discharge coolant flow. For higher pressure
applications (greater than approximately 1800 psia), the heat load produced by the combustion
process exceeds the capability of brazed tube designs. For these applications, a copper liner is
required in the high heat flux region. This configuration consists of a slotted copper liner, structural
jacket, and manifolding. Figure 176.7 illustrates these two thrust chamber concepts.

© 1998 by CRC PRESS LLC


Figure 176.7 Two thrust chamber concepts.

© 1998 by CRC PRESS LLC


MANNED MANEUVERING UNIT
Astronaut Bruce McCandless II is a few meters away from the cabin of the earth-orbiting Space
Shuttle Challenger in this 70 mm photograph taken on February 7, 1984. McCandless is one of the
two 41-B mission specialists who participated in this historical extravehicular activity (EVA). This
spacewalk represented the first use of a nitrogen-propelled, hand-controlled device called the

© 1998 by CRC PRESS LLC


Manned Maneuvering Unit (MMU), which allows for much greater mobility than that afforded
previous spacewalkers, who had to use restrictive tethers.
The MMU is a self-contained backpack with nitrogen gas propulsion that allows orbiter crews to
move outside the payload bay to other parts of the orbiter or to other spacecraft. The MMU latches
to the spacesuit (Extravehicular Mobility Unit, EMU) backpack and can be donned and doffed by
an astronaut unassisted.
MMU controls follow the layout familiar to spacecraft crews: the left-hand controller governs
fore−aft, right−left, and up−down translations, while the right-hand controller handles roll, pitch,
and yaw motions. The controllers may be used singly or in combination to give a full range of
movement within the operating logic of 729 command combinations, including attitude hold.
Thrust impulses are from 24 dry nitrogen gas thrusters each with 7.56 newtons thrust. Two
25-by-76 centimeter (9.8-by-30 inch) Kevlar filament-wrapped aluminum nitrogen tanks each hold
5.9 kilograms (13 pounds) of nitrogen when fully charged. Two 16.8 volt, 752 watt-hour silver
zinc batteries supply MMU electrical power, enough for one six-hour EVA. The nitrogen tanks
could be recharged in less than 20 minutes at the payload bay MMU service rack.
Built by Martin Marietta, Denver, CO, the MMU is 1.2 m (49.4 in.) high, 81 cm (32.5 in.) wide,
and 1.1 m (44.2 in.) deep with control arms extended. The MMU weighs 136 kg (300 lb) when
charged with nitrogen. With a spacesuited crewman and consumables added, on-orbit mass is
about 335 kg (740 lb). (Photo courtesy of National Aeronautics and Space Administration.)

Turbomachinery
The turbomachinery design process of liquid rocket engines is very similar to a normal
pump/turbine design, except for two critical areas. The first is the critical need to minimize weight.
This is perhaps the greatest difference. As stated earlier, the power density of the space shuttle
main engine turbopump is 83 horsepower per pound of turbopump weight. The second difference
is the dynamic and steady state environments that rocket engines require. Although a number of
turbojet engines operate at turbine temperatures significantly higher than most rocket engines, they
attain the steady state operating point in a matter of minutes, not in one to four seconds as do
rocket engine turbines. This produces severe thermal strains that tax the ability of materials to
sustain. Other environments that provide problems in some materials are oxygen and hydrogen.
Particle impact, fretting, and rubbing in an oxygen environment can lead to disastrous fires.
Susceptibility of materials to hydrogen embrittlement reduces the variety of materials available for
the designer or requires platings to protect materials. Another environment to which rocket engine
turbomachinery is susceptible is rotor dynamics, which is considerably more critical than in
conventional rotating machinery because of the reduced weight of rocket turbopumps. Structural
design considerations, including explanation of the processes utilized in the SSME, can be found in
Structural Design/Margin Assessment [Ryan, 1993].

176.4 System Preliminary Design Process


A number of theoretical thrust chamber performance computer models are readily available that
provide the basic performance parameters needed to support a conceptual design. The most
common is a Finite Area Combustor Theoretical Rocket Performance Program, commonly referred
to as the One-Dimensional Equilibrium (ODE) Program referenced in Computer Program for
Calculation of Complex Chemical Equilibrium Compositions and Applications, Supplement
I−Transport Properties [Gordon et al., 1984]. This model generates performance data for various

© 1998 by CRC PRESS LLC


propellant combinations as a function of mixture ratios, combustion chamber pressures, and nozzle
expansion ratios. A sample set of data for liquid oxygen (LO2 )/liquid hydrogen propellants is
given in Table 176.2.

Table 176.2 Theoretical Performance of Oxygen-Hydrogen Combustor

© 1998 by CRC PRESS LLC


The following process outlines a methodology of determining the initial set of overall system
requirements for a liquid rocket engine. A preliminary set of top-level engine requirements must be
established by the vehicle systems designer for a booster engine. For this exercise, they are as
follows:
Propellants: Liquid Oxygen/Liquid Hydrogen
Sea-Level Thrust: 400000 pounds force
Mixture Ratio: 6.0:1.0
In some instances, a minimum specific impulse value is specified. However, in most cases, the
design value should be selected as the result of vehicle-engine trade studies, along with engine
weight, recurring costs, and nonrecurring costs.
One of the first choices to be made by the engine designer is the value of combustion chamber
pressure. This choice is also the result of a series of trades. For this exercise, 2000 psia has been
chosen and is a good first approximation for the optimum value when recurring costs are one of the
more important parameters. This value will provide good performance, while simplifying
turbomachinery to two pump stages with moderate turbine temperatures. An initial assumption
needs to be made as to engine cycle. For a booster application, either a gas generator cycle or a
staged combustion cycle usually provides optimum performance. For this exercise, the staged
combustion cycle is selected. This simplifies the initial set of calculations in that the engine flow
rate and thrust chamber flow rate are approximately identical. Table 176.3 provides a typical ODE
output for a thrust chamber operating at a chamber pressure of 2000 psia and a mixture ratio of 6.0.
The first parameter to select is area ratio. From previous vehicle trade studies, the optimum nozzle
exit pressure (Pe ) for a first stage or booster vehicle is approximately 6.5 psia, the optimum for a
single-stage-to-orbit vehicle is approximately 4.0 psia, and the optimum for a parallel burn core
stage (booster and core stages ignite at sea level) is 2.5 psia. Standard atmospheric conditions for
sea level and various altitudes can be found in Terrestrial Environment (Climatic) Criteria
Guidelines for Use in Aerospace Vehicle Development, 1993 Revision [Johnson, 1993].

Table 176.3 Theoretical Performance of Oxygen-Hydrogen Combustor at a Chamber Pressure of


2000 psia and a Mixture Ratio of 6.0

© 1998 by CRC PRESS LLC


Another consideration is side loads on the nozzle. For an overexpanded nozzle (Pe < Pa ) ,
unsymmetrical flow separation results if significant dynamic loads are applied to the nozzle
(pressure times surface area forces). These side loads can ultimately destroy a nozzle and, at a
minimum, result in significant weight increases. Side loads are computed using both empirical
techniques and detailed nozzle fluid flow with computational fluid dynamics techniques.
Calculating the locations within the nozzle where flow separation will occur is very difficult and
side loads can usually be quantified accurately with test data.
As the application for this exercise is a booster, an exit pressure of approximately 6.5 psia is
desired. Using Table 176.3, an expansion ratio of 30 yields an exit pressure of 6.2 psia, which is
close enough for a first approximation of the engine characteristics. Using the initial vehicle thrust
and mixture ratio requirements, theoretical values of C ¤ and CF obtained from Table 176.3, and
assumed values of C ¤ efficiency of 0.99 (readily obtainable with oxygen/hydrogen coaxial tube
injectors) and CF efficiency of 0.97, various engine parameters can be computed using Eqs.
(176.3), (176.5), (176.6), (176.8), and (176.9) and are shown below in order of
computation:

Fvac 400 000 lb


Propellan Oxygen/hydrogen
ts
MR 6.0
Cycle Staged combustion
Pns 2000 psia
" 30:1
CF 1.8311
C* 7539.8 ft/s
Ai 109.22 in2
Ae 3276.7 in2
w_ t 932.27 lbm/s
w_ f 133.18 lbm/s
w_ o 799.09 lbm/s
Ispvac 429.1 s
Fsl 351 846 lbf
IspSl 377.41 s

176.5 Conclusion
The science of rocketry has enabled some of humankind's greatest achievements, ranging from
instantaneous global communications and accurate weather forecasting via geostationary satellites
to trips to the moon. NASA's space shuttle is one of the most complex flying machines ever built
and is the only partially reusable launch vehicle. While several countries have rockets capable of

© 1998 by CRC PRESS LLC


carrying a variety of payloads, the U.S. and Russia are the only countries with spacecraft that can
transport a crew to and from orbit. France and China have expendable launch vehicles now in use,
and Japan is developing another.
The U.S. is on the threshold of a next generation launch system. NASA aerospace engineers and
industry experts are exploring new concepts, including the first fully reusable launch vehicle.
Developing rockets for 21st century missions promotes enhanced technologies to meet new
challenges, including balancing design requirements between operability, performance, weight,
and cost. The next generation of spaceship will open new doors to the space frontier.

Defining Terms
Ablation: A passive cooling technique in which heat is carried away from a vital part by
absorption into a nonvital part, which may melt or vaporize and then fall away, taking the
heat with it.
Combustion chamber: A devicewhich includes a throat regionto mix, burn, and control
propellants.
Injector: A device to distribute and inject propellants into the combustion
chamber.
Nozzle: A device used to accelerate the combusted gases.
Propellant: Fuel [the chemical(s) the rocket burns] and an oxidizer (oxygen compounds) to ignite
the fuel.
Sea level: Standard atmospheric conditions at an altitude of zero feet.
Side loads: Unsymmetrical loads put on a nozzle because of internal flow separation of
overexpanded gases.
Regeneratively cooled: A cooling technique in which propellants, usually fuel, are utilized to
remove heat from the inner wall of a combustor in a heat exchange
process.
Thrust chamber assembly: An assembly consisting of the main injector, combustion chamber,
and nozzle. Depending upon fabrication techniques, the combustion chamber and nozzle can
be separate components or combined into a single component.
Vacuum: Conditions where atmospheric pressure can be considered to be 0.0
psia.

References
Gordon, S., McBride, B., and Zeleznik, F. 1984. Computer Program for Calculation of Complex
Chemical Equilibrium Compositions and Applications, Supplement ITransport
Properties. NASA Technical Memorandum 86885, National Aeronautics and Space
Administration, Office of Management, Scientific and Technical Information
Program.
Johnson, D. 1993. Terrestrial Environment (Climatic) Criteria Guidelines for Use in Aerospace
Vehicle Development, 1993 Revision. NASA Technical Memorandum 4511, National
Aeronautics and Space Administration, Office of Management, Scientific and Technical

© 1998 by CRC PRESS LLC


Information Program.
Ryan, R. 1993. Structural Design/Margin Assessment. NASA Technical Paper 3410, National
Aeronautics and Space Administration, Office of Management, Scientific and Technical
Information Program.
Sutton, G. P. 1992. Rocket Propulsion Elements, 6th ed. John Wiley & Sons, New York.

Further Information

• American Institute of Aeronautics and Astronautics (AIAA)


• American Society of Mechanical Engineers (ASME)
• National Space Society
• Marshall Space Flight CenterCentral Technical Library (Phone: 205-544-4524)

© 1998 by CRC PRESS LLC

Das könnte Ihnen auch gefallen