You are on page 1of 10

1586_C199.

fm Page 1 Tuesday, May 11, 2004 3:13 PM

199
Spacecraft and
Mission Design
199.1 Spacecraft Environments
199.2 Fundamental Principles
199.3 Spacecraft/Mission Categories
Launch Vehicles Sounding Rockets Earth-Orbiting
Spacecraft Lunar Spacecraft Interplanetary Spacecraft Entry
Vehicles Landers Returning Spacecraft

199.4 Spacecraft Subsystems

Wallace T. Fowler
University of Texas, Austin

Structure Subsystem Power Subsystem Attitude Control


Subsystem Three-Axis Stabilization Spin Stabilization
Dual-Spin Spacecraft Gravity Gradient Stabilization
Magnetic Stabilization Solar Radiation Stabilization
Telecommunications Subsystem Propulsion Subsystem
Thermal Control Subsystem Spacecraft Mechanisms Launch
Vehicles Upper Stages Entry/Landing Subsystems
Subsystem Integration and Redundancy

199.5 Spacecraft/Mission Design Process

The design of a spacecraft and the design of its mission are closely related parts of the same iterative
process. The mission concept usually comes first, but new missions for existing spacecraft designs have
also been developed. The key to any successful design is the development of a good set of requirements.
In any case, the mission concept is often changed by design factors, and the spacecraft design is driven
by mission requirements. Effort spent in defining and refining requirements is crucial to good design.
The design process starts with either a mission concept or spacecraft and arrives at a complete design.
However, every design process is unique, and no design is optimal. Factors will always arise that do not
seem to fit. When this happens, the mission requirements or the design process should be modified to
accommodate or eliminate the unique factors.
The match between spacecraft and mission is determined by a number of factors. The two most
important preliminary design factors are the mass of the mission payload and the amount of velocity
change, DV, that must be provided to the payload to carry out the mission. These two factors dictate
much of the mission and spacecraft design, as they are the primary drivers for the mass of the booster
assembly. Other primary design factors are mission duration and destination to which the payload is
being sent. A 2-day mission on which the spacecraft stays near the Earth will require a very different
hardware/mission plan combination than will a multiyear mission to an outer planet. If a spacecraft is
manned, its design and mission planning are greatly complicated. Safety factors that are acceptable for
unmanned flight are unacceptable for manned flight. Life support and safety equipment, as well as
backups for such systems, dominate the design. One factor that should not be ignored during the design

2005 by CRC Press LLC

1586_book.fm Page 2 Tuesday, May 11, 2004 1:46 PM

199-2

The Engineering Handbook, Second Edition

of a spacecraft/mission combination is operations. A detailed operations plan is as important to a


successful space mission as is good hardware and a good mission plan. Inclusion of operations planning
in the initial spacecraft and mission design process, including planning for contingencies, has the potential
for lowering mission costs, increasing the chances of mission success, and lowering the potential for
problems caused by unforeseen operator/hardware interactions.

199.1 Spacecraft Environments


Every spacecraft must be able to survive in several environments. All those built on earth must be able
to survive the earths atmosphere or be protected from it. The Earths standard atmosphere is described
in Chapter 198, and atmosphere models for other planets are available in NASA literature. The launch
environment, characterized by vibration, noise, g-loads, aerodynamic loads, transition from air to vacuum, and so on, constitutes a major test for spacecraft. The space environment presents another set of
problems for the designer. Hard vacuum, radiation, and temperature extremes are common to all
missions. Spacecraft that fly through or beyond the Van Allen radiation belts experience more severe
radiation hazards than those spacecraft that stay inside the Van Allen belts. For manned spacecraft,
extended flight beyond the protection of the Van Allen belts implies special crew radiation shielding.
Orbital debris is also a hazard that must be considered.
Vehicles that reenter the Earths atmosphere or enter the atmosphere of another planet must contend
with entry aerodynamics and accompanying loads and heating. For landers, abrasive dust can be a
problem. Finally, if a spacecraft has a crew, its internal environment must support life while its structure
protects the habitable volume. Good discussions of the role of environment in spacecraft design can be
found in Fortescue and Stark [2003] and Griffin and French [1991].

199.2 Fundamental Principles


The primary factor linking the spacecraft and the mission scenario is linear momentum change, and this
is the key driver in overall sizing of the spacecraft and its booster system. Except for missions dominated
by attitude-control requirements, the mission scenario that requires the lowest total linear momentum
change will be the most efficient in terms of propellant. Thus, in designing mission and spacecraft, it is
important to (1) explore many mission scenarios to ensure that the total momentum change required is
as low as possible; (2) use staging to discard unneeded mass and reduce the propellant required during
each mission phase; (3) employ engine/propellant combinations that produce large momentum changes
with relatively small amounts of propellant (i.e., propellants with high exhaust velocities relative to the
vehicle); (4) keep all spacecraft component masses as small as possible; and (5) use unmanned spacecraft
whenever possible to keep vehicle masses small. It is also important to use space-qualified hardware when
available (and appropriate) to reduce cost and risk.
The ability of a propulsion system to impart momentum changes to a spacecraft is the velocity of the
exhaust with respect to the spacecraft. The key equation is
c = Ispg

(199.1)

where c is exhaust velocity relative to the vehicle, Isp is specific impulse, and g is the acceleration of gravity
at the surface of the Earth. c is in m/sec (ft/sec), g is in m/sec2 (ft/sec2), and Isp is in units of seconds.

Note: The value used for g in Equation (199.1) does not change when the spacecraft is in a non-Earth
gravity field.
Propellant combinations (fuel and oxidizer) are often listed in propulsion references with a corresponding Isp value. These values imply an appropriate engine in which to combine the propellants
efficiently. In spacecraft design studies, trade-offs between propellant storability and Isp are common. Isp
2005 by CRC Press LLC

1586_book.fm Page 3 Tuesday, May 11, 2004 1:46 PM

199-3

Spacecraft and Mission Design

values for most large solid rockets are just under 300 seconds. The Isp value for the space shuttle main
engines (liquid oxygenliquid hydrogen) is more than 400 seconds.
A second basic equation, which relates the mass of the spacecraft assembly at the ith stage in the
mission (the total mass to be propelled at that point in the mission) to propellant mass that produces
the velocity change, is
DVi = c i ln(m oi / m fi )

(199.2)

where DVi is the velocity increment provided to the spacecraft assembly by the propellant in the ith stage,
ci is the exhaust velocity of the propulsion system in the ith stage relative to the spacecraft assembly, moi
is the mass of the spacecraft assembly prior to burning the propellant in the ith stage, and mfi is the mass
of the spacecraft assembly after the ith stage propellant is expended. The mass of the propellant burned
in the ith stage is the difference between moi and mfi.
Sequential staging is a very important concept. Mass that is no longer needed should be discarded as
soon as is practical in order to avoid using propellant to change the momentum of useless mass. The
following equation shows the total DV for a two-stage vehicle.
DV = c 1 ln(m o1 / m f 1 ) + c 2 ln(m o 2 / m f 2 )

(199.3)

When staging occurs, excess hardware (tanks, engines, pumps, structure) is dropped after a stage burns
out and before the next stage begins its burn. Thus, mf1 will be larger than mo2. This difference is one of
the most important factors in space vehicle design and performance. A good discussion of staging can
be found in Wertz and Larson [1998].
Equation (199.2) does not apply directly when the two different types of engines with different Isp
values are being used simultaneously. The space shuttle first stage (with solid rocket boosters and main
engines burning simultaneously) and the Delta booster (with strap-on solid rockets around a liquid
rocket core vehicle) are examples of this. A simple modification of Equation (199.2) applies, however.
This equation looks like Equation (199.2) but has a significant difference in the definition of the exhaust
velocity. For such vehicle mission stages,
DV = c *ln(m o / m f )

(199.4)

where c* is a weighted combination of the exhaust velocities of the two types of engines being used.
Specifically,
c * = f1*c1 + f 2*c 2

(199.5)

where c1 and c2 are the exhaust velocities for the two engine types being used, and f1 and f2 are the
respective mass flow rate fractions for the two engine types (note that f1 + f2 = 1). Equations (199.4) and
(199.5) allow characteristic DV values to be calculated for vehicle stages involving parallel burning of
dissimilar engine types.

199.3 Spacecraft / M ission Categories


Launch Vehicles
Launch vehicles and launches are a common element/system for almost all spacecraft. The choice of launch
vehicle is determined by the mission, the mass to be boosted to orbit, the required DV, the compatibility
between launcher and payload, and the overall cost of the launch (launcher plus launch services). The
International Reference Guide to Space Launch Systems [Isakowitz, 1999] is an excellent source of informa 2005 by CRC Press LLC

1586_book.fm Page 4 Tuesday, May 11, 2004 1:46 PM

199-4

The Engineering Handbook, Second Edition

tion on launchers. A good rule of thumb to keep in mind is that for eastward launches into low Earth
orbit (LEO) from a latitude of about 30, between 9300 m/sec (30,500 ft/sec) and 9,450 m/sec (31,000 ft/
sec) total DV must be supplied by the booster. (This accounts for gravity and drag losses as well as increasing
the potential and kinetic energy of the spacecraft.) The time from launch to orbit insertion will depend
on the booster acceleration capabilities and constraints, but will generally be less than 8 min.

Sounding Rockets
Sounding rockets are usually small, and their missions are characterized by short mission duration and
short times out of the atmosphere of the Earth (or another planet). Their objectives are usually scientific,
their equipment is powered by batteries, and they usually possess a recovery system for the instrumentation package (or perhaps the entire vehicle). Sounding rocket payloads usually consist of structure,
instrumentation, a control system, a power system, communications, and a recovery system. The mass
and DV requirement for a sounding rocket depend primarily on the mass to be boosted, the propellant
combination to be used, and the altitude to which the payload is to be sent.

Earth-Orbiting Spacecraft
There are many varieties of Earth-orbiting spacecraft. Examples include the original Sputnik and Explorer
spacecraft, the space shuttle, and communications satellites such as Intelsat. There are resource mappers,
weather satellites, military satellites, and scientific satellites. There are many active satellites in orbit around
the Earth today. Useful satellite lifetimes range from a few hours to years. It is likely that the Lageos
satellite, a passive laser reflector, will be providing useful data for at least 100 years. Typical Earth-orbiting
spacecraft systems are the structure, power system, thermal management system, communications, sensors, computation/data storage, propulsion, and possibly a deorbit/entry system. Launch dates, orbits,
and missions of a wide variety of Earth-orbiting spacecraft have been summarized in compact form in
the TRW Space Log [Thompson, 1987]. Especially useful are the summary documents in this series.

Lunar Spacecraft
A special category of Earth-orbiting spacecraft meriting separate mention is lunar spacecraft. Such
spacecraft must climb out of the Earths gravity well, requiring a DV of more than 3100 m/sec (10,200
ft/sec) to reach the moon starting in LEO. Spacecraft that orbit the moon require about 1070 m/sec (3500
ft/sec) to enter an orbit about the moon and 2070 m/sec (6800 ft/sec) to land. Special thermal considerations are necessary for vehicles designed to rest on the lunar surface because of the slow rotation rate
of the moon (the lunar day and night are 14 Earth days each). Lunar spacecraft have the systems typical
of Earth-orbiting spacecraft, with the possible exception of the deorbit/entry system.

Interplanetary Spacecraft
Interplanetary spacecraft have much in common with lunar vehicles, but they also have many unique
characteristics. Spacecraft targeted to the outer planets have longer mission times and different power
supplies than those that remain in the inner solar system. The inner planets (Venus and Mercury) present
particular problems for the designer of landers. The atmosphere of Venus is hot, dense, and corrosive. The
surface of Mercury, except near the poles, sees wide swings in temperature. A point on the equator of
Mercury sees 88 Earth days of sunlight followed by 88 Earth days of darkness. Obviously, structure, materials,
and thermal systems are important for such spacecraft. Injection onto a minimum-energy transfer to Mars
requires 3600 m/sec (11,800 ft/sec). Orbiting Mars will require another 2100 m/sec (7000 ft/sec).

Entry Vehicles
Entry vehicles are spacecraft (or spacecraft segments) designed to use aerodynamic lift and drag to slow
or turn the spacecraft, leading to an orbital plane change, lowering of an orbit, or a descent into the
2005 by CRC Press LLC

1586_book.fm Page 5 Tuesday, May 11, 2004 1:46 PM

Spacecraft and Mission Design

199-5

planetary atmosphere. Aerothermodynamic considerations (both stagnation point temperature and integrated heat load) and deceleration loading dominate the design of such vehicles. Aerodynamic control
surfaces are the primary means of trajectory shaping, which in turn controls the heating and deceleration.
Survivability is a primary design consideration. An entry vehicle is a special type of hypervelocity vehicle,
usually a glider (see Chapter 198 for additional information).

Landers
Landers are a special class of spacecraft. Some landers are also entry vehicles and some are not, depending
on whether the body on which the landing is to take place has an atmosphere. Landings on targets with
atmospheres can involve parachutes, cushions, shock absorbers, and so on. Landings on airless bodies
usually involve either rocket braking or braking via penetration of the surface of the target body.

Returning Spacecraft
Spacecraft that return to the planet of origin are a very special case. Vehicles that fly to Earth orbit and
then deorbit and land on Earth, although complex, are much simpler than vehicles that fly from the
Earth to a target, land operate, launch, return to Earth, enter the atmosphere, and are recovered. For
such vehicles, mass is by far the most important driver for all design decisions, and repeated staging is
the primary feature of the overall design. Often the redesign of the staging process can greatly simplify
an otherwise complex vehicle and mission (the lunar orbit rendezvous used in the Apollo program is an
excellent example).

199.4 Spacecraft Subsystems


Once the overall mission and spacecraft requirements have been defined, the process of defining requirements for various spacecraft subsystems can begin. The integration of the various subsystems into a
harmoniously functioning spacecraft can be very difficult. One way to reduce the difficulties in spacecraft
integration is to spend sufficient time and effort early in carefully and completely defining the interfaces
between the subsystems of the spacecraft. Time spent in defining interfaces will repay itself many times
as the design progresses. The following is a brief discussion of the major design considerations of the
principal spacecraft subsystems. Chetty [1991], Fortescue and Stark [2003], and Griffin and French [1991]
provide excellent discussions of various spacecraft subsystems.

Structure Subsystem
The structure of the spacecraft holds the components together, provides load paths for launch forces,
maneuver loads, and so on, and is usually an integral part of the thermal control system and/or the
electrical system for the spacecraft. Design constraints that the structural subsystem must usually meet
include stiffness requirements, placing principal inertia axes in preferred directions, sustaining mission
loads, and serving as a ground for all spacecraft electrical equipment. Spacecraft structures can be metallic,
composite, or ceramic. Structural mass is generally 5 to 20% of the overall spacecraft launch mass.

Power Subsystem
The type of power system chosen for a spacecraft is strongly dependent on the mission of the spacecraft
and the power demands of the equipment to be carried. Short-duration missions often use batteries.
Longer-duration missions that go no farther from the sun that Mars can use photoelectric power (solar
cells). Long-duration, low-power missions on which the spacecraft go very far from the sun (e.g., Voyager)
use radioisotope thermoelectric generators (RTGs). Other potential power sources are solar thermionic
systems, chemical dynamic systems, and nuclear power systems. Limitations on power often place constraints on spacecraft designs and mission operations.
2005 by CRC Press LLC

1586_book.fm Page 6 Tuesday, May 11, 2004 1:46 PM

199-6

The Engineering Handbook, Second Edition

Attitude Control Subsystem


Most spacecraft require attitude stabilization to point antennas, properly orient solar arrays, point sensors,
and orient thrusters. Attitude control consists of sensing the current attitude of the spacecraft and applying
torques to reorient the spacecraft to a desired attitude.

Three-Axis Stabilization
The most complex and precise spacecraft attitude control is three-axis stabilization accomplished using
momentum wheels or reaction wheels. Attitude thrusters are also used for three-axis stabilization, either
as the primary attitude control mechanism or to desaturate the momentum wheels. Three-axis stabilization systems can be fast and accurate. For example, the attitude control system on the Hubble Space
Telescope, which uses reaction wheels, can point the optical axis of the telescope to within 0.007 arc
seconds of the desired direction.

Spin Stabilization
Spin stabilization is much simpler and less precise than three-axis stabilization. Spin-stabilized spacecraft
must be rigid and spinning about the principal axis with the largest mass moment of inertia. The first
U.S. orbiting satellite, Explorer 1 (launched 31 January 1958), provided an early demonstration of the
need for spinning spacecraft to be rigid. Explorer 1 was a long, slender cylinder spinning about its long
axis. It possessed flexible whip antennas that dissipated energy and caused the spacecraft to tumble within
a day of orbit insertion.

Dual-Spin Spacecraft
Some spacecraft have a spinning section and a despun section. Such spacecraft must have systems to
maintain the attitude of the despun section and to control the direction of the spin vector. Reactions
between the spinning and despun sections are major design and operational considerations.

Gravity Gradient Stabilization


Spacecraft can use the gradient of the gravitational field of a planet to provide restoring torques and
maintain the spacecraft in an orientation with the spacecrafts long axis pointing along the local vertical.
Spacecraft designed to employ gravity gradient stabilization usually feature mass concentrations at the
ends of a long boom that lies along the local vertical. Gravity gradient stabilization works best with nearcircular orbits and would not work on even moderately elliptic orbits. A pointing accuracy of about 1
with some wobble is achievable with gravity gradient stabilization of Earth satellites.

Magnetic Stabilization
Spacecraft attitudes can be stabilized or changed using torques developed by running currents through
coils (torque rods) mounted in the spacecraft if the spacecraft is orbiting a body with a significant
magnetic field. The primary difficulty with this type of attitude stabilization system is keeping track of
the relative orientations of the spacecraft axes and the magnetic field lines as the spacecraft moves along
in its orbit. Magnetic stabilization is slow and coarse.

Solar Radiation Stabilization


The pressure of solar radiation falling on a spacecraft can be used to create torques. Spacecraft that
employ solar radiation stabilization are large and lightweight.

2005 by CRC Press LLC

1586_book.fm Page 7 Tuesday, May 11, 2004 1:46 PM

Spacecraft and Mission Design

199-7

Telecommunications Subsystem
The primary function of the telecommunications subsystem is to receive, process, and transmit electronic
information. Some satellites (communications, intelligence, and weather satellites) are designed to meet
one or more telecommunications mission objectives. The telecommunications subsystem often demands
a significant amount of electrical power, and this demand often drives the design of the power system.
The spacecraft computer/sequencer can be considered part of the telecommunications system or a
separate system. Chetty [1991], Fortescue and Stark [2003], and Griffin and French [1991] provide
excellent discussions of spacecraft telecommunications.
The telecommunications system often drives the attitude and pointing requirements for the spacecraft.
Furthermore, the capabilities of the spacecraft telecommunications antenna(s), coupled with limited
power available for signal transmission, often require the use of large and expensive receiving antennas
on Earth. This can greatly increase operational costs. Finally, deployable spacecraft antennas carry additional risks. The very successful Galileo mission was severely impacted by the failure of the onboard highgain antenna to deploy properly.

Propulsion Subsystem
The spacecraft propulsion subsystem will be considered separately from the launch vehicle and upper
stages. The propulsion subsystem is responsible for orbit maintenance, small orbit changes, attitude
control system desaturation, and so on. Spacecraft onboard propulsion systems use chemically reacting
fuels and oxidizers, monopropellants and catalysts, cold gas, and ion thrusters. Chetty [1991], Fortescue
and Stark [1992], and Griffin and French [1991] provide excellent discussions of propulsion systems.

Thermal Control Subsystem


Spacecraft thermal control is of utmost importance. Maintaining appropriate thermal environments for
the various spacecraft is essential to proper component function and longevity. The primary problem
with thermal management of spacecraft is that all excess heat must ultimately be radiated by the spacecraft.
For many unmanned spacecraft, passive thermal control involving coatings, insulators, and radiators is
ideal. For other spacecraft, active thermal control is necessary. Active systems can employ heaters, radiators, heat pipes, thermal louvers, and flash evaporators. Significant spacecraft mass reductions can
sometimes be made if the excess heat from one component is conducted to another component that
would otherwise require active heating.

Spacecraft Mechanisms
Some spacecraft are mechanically passive and have no mechanisms in their designs. However, even these
spacecraft require mechanisms to deploy them from their boosters. Mechanically active spacecraft can
feature a wide variety of mechanisms. There are deployment systems for booms and antennas, docking
mechanisms, robot arms, spin/despin systems, sensor pan and tilt mechanisms, gyroscopes, reaction
wheels (also considered as part of the attitude control system), landing legs, and airlocks. The types of
mechanisms depend on the mission of the spacecraft.

Launch Vehicles
The launch vehicle, or booster, is an integral part of most spacecraft systems. The International Reference
Guide to Space Launch Systems [Isakowitz, 1999] covers the capabilities of most currently available launch
vehicles in detail. The rotation of the Earth has a major effect on launch vehicle performance. The
eastward motion of the launch site must be considered when computing the payload that a launch vehicle
can place in orbit. The launch azimuth and the latitude-dependent eastward velocity of the launch site
can be combined to determine how much velocity assist is provided by Earths rotation. For westward

2005 by CRC Press LLC

1586_book.fm Page 8 Tuesday, May 11, 2004 1:46 PM

199-8

The Engineering Handbook, Second Edition

launches (to retrograde orbits), the rotation of the Earth extracts a penalty, requiring additional velocity
to be provided by the booster to make up for the eastward velocity of the launch site.

Upper Stages
Launch vehicles are often combined with upper stages to form a more capable launch vehicle. Commonly
used upper stages include the IUS, the Centaur, and several types of PAM (payload assist modules).
Upper stages are also carried aloft in the payload bay of the space shuttle to boost spacecraft to higher
orbits. Designers should always consider the option of using an available upper stage if the mission
requires more DV than can be provided by the basic launch vehicle.

Entry/Landing Subsystems
Entry subsystems can be considered either as a separate set of devices or as parts of other subsystems.
The heat shield can be reusable or ablative, and could be considered to be part of the thermal control
subsystem. The retrorockets, if any, might be considered to be part of the propulsion system. Parachutes,
landing legs, cushions, wings, aerosurfaces, a streamlined shape, and landing gear, when needed, belong
solely to the category of entry and landing subsystems.

Subsystem Integration and Redundancy


The interplay among spacecraft subsystems is great, and their integration into an efficient, smoothly
operating, and reliable spacecraft is an immense challenge. Redundant system elements can provide added
reliability if the risk of failure must be lowered. However, redundancy increases initial costs and usually
increases overall mass. It is often good to avoid physically redundant elements, choosing physically
different but functionally redundant elements instead. System elements that are to be functionally duplicated should be identified as early in the design process as possible.

199.5 Spacecraft/Mission Design Process


The spacecraft/mission design process can be outlined as follows:
1. Beginning with a set of mission goals, develop a set of mission and spacecraft requirements that
must be met in order to achieve the goals. An excellent discussion of requirements and how to
develop them is found in Wertz and Larson [1998]. Mission goals may change during the design
process, and when this happens, the requirements must be reevaluated.
2. Before beginning a methodical approach to the design, it is almost always helpful to bring the
design team together and conduct a brainstorming session. For those who are unfamiliar with
brainstorming, Adams [1986] is recommended. The brainstorming session should be expected to
identify several unorthodox candidate scenarios or components for the mission and spacecraft.
3. Develop a conceptual model for the spacecraft and the mission, identifying major systems and
subsystems, and most important, the interfaces between systems, both within the spacecraft itself
and between the spacecraft and its support elements on the ground and in space. A preliminary
mission scenario and rough timeline should also be developed at this time. These conceptual
elements should be developed from the mission and spacecraft requirements. This step will allow
the definition of major systems and subsystems and will identify major interactions between
mission events and spacecraft hardware elements. Preliminary indications of pointing requirements, required sensor fields of view, DV requirements, allowable spacecraft mass, and required
and desired component lifetimes should come from these considerations. Brown [1998], Fortescue
and Stark [2003], Griffin and French [1991], and Isakowitz [1999] give information appropriate
to this topic.

2005 by CRC Press LLC

1586_book.fm Page 9 Tuesday, May 11, 2004 1:46 PM

Spacecraft and Mission Design

199-9

4. The role of heritage in spacecraft design is important. Systems and technologies that have worked
well on past spacecraft are often good candidates for new spacecraft with similar missions. Examine
the recent past for similar missions, spacecraft with similar requirements, and so on. Although
sufficiently detailed systems overviews of recent spacecraft are often difficult to obtain, the effort
necessary to obtain them is almost always worth the effort. However, the design team should be
careful to avoid early adoption of a candidate system from earlier spacecraft in order to avoid
being locked into a system that only marginally meets or does not meet design requirements.
5. Once the first four steps have been taken (some more than once), the design team should develop
preliminary spacecraft design candidates (spacecraft/scenario/timeline combinations), concentrating on systems and subsystems and their performance, masses, power requirements, system interactions, thermal input/output, and costs. Several candidate designs should be considered at this
stage and their predicted relative characteristics compared. These considerations allow the design
team to learn more about the requirements and how best to meet them. It is likely that the design
team will recycle through some or all of the design process at this point for one or more of the
candidate designs. The possibility of developing a new composite design with the best features of
several candidate designs should not be ignored.
6. As candidate designs are refined, the design team should develop criteria for use in choosing
among candidate designs. The information on the candidate designs should be organized and
presented at a preliminary design review. The feedback from those attending the review is valuable
in the identification of strengths and weaknesses of candidate designs.
7. At some point, determined by budget, timeline, technical considerations, and the characteristics
of the candidate designs, one design concept must be chosen for development. Once the choice
is made, the design team should carry out a complete analysis of the chosen design. This analysis
should include a detailed mission scenario and timeline, launch vehicle choice, DV requirements
for every maneuver, hardware choices for every system and subsystem, system interaction analyses,
manufacturability considerations, component and overall cost analyses, and failure mode effects
analyses. Any remaining system/subsystem hardware choices should be made, and the design
should be presented at a critical design review. The feedback from the critical design review will
often result in major improvements in the design. Questions arising at the critical design review
will often precipitate another pass through the design process.
8. The effects of real or imagined environmental impacts on the public of mission and/or spacecraft
features and resulting political pressures cannot be overemphasized. The Galileo mission to Jupiter
provides good examples. First, there were public protests at the Kennedy Space Center in Florida
because Galileo carried a nuclear power device (an RTG). Most protestors had no idea that RTGs
had been launched safely for over 2 decades and that RTGs carried by Apollo were still in operation
on the Moon. Also, since Galileo flew by Venus and then used an Earth fly-by to provide a gravity
assist to help get it to Jupiter, there were those who predicted that this nuclear device might drift
off course and hit the Earth. Mission designers must learn to anticipate such protests and provide
pertinent information to the public in a straightforward and clear format.

Defining Terms
Mission scenario A sequence of events and times that, in their entirety, meet the objectives of the
mission.
Specific impulse The amount of thrust force obtained from a fuel/oxidizer/engine combination when
a unit weight of fuel is burned in 1 second. The units of specific impulse are seconds because
the force and weight units cancel.
Subsystem A part of an overall system that can be isolated to some degree. Subsystems interact and
cannot actually be isolated, but it is convenient to consider subsystems in sequence rather than
all at once.

2005 by CRC Press LLC

1586_book.fm Page 10 Tuesday, May 11, 2004 1:46 PM

199-10

The Engineering Handbook, Second Edition

References
Adams, J. L. 1986. The Care and Feeding of Ideas: A Guide to Encouraging Creativity, 3rd ed. AddisonWesley, Reading, MA.
American Institute of Aeronautics and Astronautics. 1998. AIAA Aerospace Design Engineers Guide, 4th
ed. American Institute of Aeronautics and Astronautics, Washington, DC.
Brown, C. D. 1998. Spacecraft Mission Design, 2nd ed. AIAA Education Series, American Institute of
Aeronautics and Astronautics, Washington, DC.
Chetty, P. R. K. 1991. Satellite Technology and Its Applications. TAB Professional and Reference Books,
Blue Ridge Summit, PA.
Fortescue, P. and Stark, J. 2003. Spacecraft Systems Engineering, 3rd ed. John Wiley & Sons, New York.
Griffin, M. D. and French, J. R. 1991. Space Vehicle Design. AIAA Education Series, American Institute
of Aeronautics and Astronautics, Washington, DC.
Isakowitz, S. J., Ed. 1999. International Reference Guide to Space Launch Systems, 3rd ed. AIAA Space
Transportation Technical Committee, AIAA, Washington, DC.
Sellers, J. J. 1994. Understanding Space An Introduction to Astronautics. McGraw-Hill, New York.
Thompson, T. D., Ed. 1987. TRW Space Log, 19571987. TRW Space & Technology Group, Redondo
Beach, CA.
Wertz, J. R. and Larson, W. J., Eds. 1998. Space Mission Analysis and Design, 3rd ed. Space Technology
Library, Kluwer Academic, Boston.

Further Information
An excellent and inexpensive source of information on the orbital mechanics necessary for mission
planning is Bate, R. R., Mueller, D. D., and White, J. E. 1971. Fundamentals of Astrodynamics. Dover, New
York.
Details on the locations of the planets, their moons, the orbit of the Earth about the sun, values of
astrodynamical constants, and many other items pertinent to mission planning are found in The Astronomical Almanac, from the Nautical Almanac Office, Naval Observatory, published yearly by the U.S.
Government Printing Office, Washington, DC; and Seidelmann, P. K., Ed. 1992. Explanatory Supplement
to the Astronomical Almanac. University Science Books, Mill Valley, CA.
An excellent overview of the interplay between spacecraft, launch vehicle, trajectory, and operations
is given in Yenne, B., Ed. 1988. Interplanetary Spacecraft. Exeter, New York.

2005 by CRC Press LLC

You might also like